[ { "title": "1207.1525v1.Double_Exchange_Ferromagnetism_in_the_Peierls_Insulator_State.pdf", "content": "arXiv:1207.1525v1 [cond-mat.str-el] 6 Jul 2012Double Exchange Ferromagnetism in the Peierls Insulator St ate\nS. Nishimoto1and Y. Ohta2\n1Institut f¨ ur Theoretische Festk¨ orperphysik, IFW Dresde n, 01171 Dresden, Germany and\n2Department of Physics, Chiba University, Chiba 263-8522, J apan\n(Dated: November 8, 2018)\nWe studythe effects of opening of theband gap on the double exc hange ferromagnetism. Applying\nthe density-matrix renormalization group method and an ana lytical expansion from the dimer limit\nto the one-dimensional double exchange model, we demonstra te for a relevant region of the exchange\ncoupling that, in the weak dimerization regime, the Peierls gap opens in the fully spin-polarized con-\nduction band without affecting its ferromagnetism, whereas in the strong dimerization regime, the\nferromagnetism is destroyed and the Mott gap opens instead, leading the system to the antiferro-\nmagnetic quasi-long-range order. An insulator version of t he double exchange ferromagnetism is\nthus established.\nPACS numbers: 71.10.Fd, 75.10.-b, 71.30.+h\nMagnetism and electronic transport properties of ma-\nterials are closely related to each other, e.g., insulat-\ning transition-metal oxides are typically antiferromag-\nnetic and ferromagnetismusually goeshand in hand with\nmetallicity [1], and elucidation of the mechanisms of this\nrelationship is one of the major issues in the field of\nphysics of strong electron correlations. A well-known ex-\nampleis the doubleexchangeferromagnetismthat occurs\nin mixed systems of localized spins and itinerant elec-\ntrons interacting via the Hund’s rule coupling, where the\ncoherent motion of the itinerant electrons aligns the lo-\ncalized spins ferromagnetically to gain in kinetic energy\nof the itinerant electrons [2–4].\nA question then arises as to what happens in this fer-\nromagnetism if the coherent motion of electrons ceases\ndue, e.g., to the opening of the band gap. This issue\ncomes into a real question in ferromagnetic hollandite\nK2Cr8O16[5], where the double exchange mechanism in-\nduces the three-dimensional (3D) full spin polarization\nin the system [6], and then the metal-insulator transition\nfollows in its fully spin-polarized quasi-one-dimensional\n(1D) conduction band by the Peierls mechanism [7, 8],\nwithout affecting its 3D ferromagnetism. Thus, the un-\ncommon ferromagnetic insulating (FI) state is realized in\nthis material [7].\nA naive answer to the above frequently asked question\nmay then be that the ferromagnetism can survive if the\nbandgapissmallenoughincomparisonwiththewidthof\nthe conduction band and therefore the motion of conduc-\ntion electrons, though not coherent, is not significantly\nsuppressed. However, to the best of our knowledge, no\nquantitative theoretical studies have been made on this\nissue whether this is actually the case.\nIn this paper, we address this issue from the theoret-\nical standpoint. We apply the numerical density-matrix\nrenormalization group (DMRG) technique [9] and ana-\nlytical expansion from the dimer limit to the 1D dou-\nble exchange model that simulates the quasi-1D chain of\nK2Cr8O16and study the effects of opening of the bandgap on the double exchange ferromagnetism. We calcu-\nlate the total-spin quantum number and charge gap of\nthe system and extract the ground-state phase diagram\nof the model.\nWe will thereby demonstrate for a relevant region of\nthe exchange coupling that, in the weak dimerization\nregime, the Peierls gap opens in the fully spin-polarized\nconduction band without affecting its ferromagnetism,\nwhereas in the strong dimerization regime, the ferromag-\nnetism is destroyed and the Mott gap opens instead, due\nto the effective “on-dimer” Coulomb interaction, which\nleads the system to the antiferromagnetic quasi-long-\nrange order. The metallicity itself is not therefore a\nnecessary condition for the realization of the double ex-\nchangeferromagnetismandthusan“insulatorversion”of\nthe double exchange ferromagnetism is established. This\nis a route to realization of insulating ferromagnets, dif-\nferent from those discussed in doped LaMnO 3where the\norbital ordering plays an essential role [10–12].\nOur model (see Fig. 1) contains the terms of Peierls\ndimerization and Hund’s rule coupling and is defined by\nthe Hamiltonian\nH=−L/summationdisplay\ni=1,σti,i+1(c†\ni,σci+1,σ+H.c.)−JHL/summationdisplay\ni=1si·Si\nsi=1\n2/summationdisplay\nσ,σ′c†\ni,στσ,σ′ci,σ′, (1)\nwherec†\ni,σis the creation operator of an electron with\nspinσ(=↑,↓) at site i,siis the spin operator of a\nconduction electron at site i,τis the vector of Pauli\nmatrices, and Siis the quantum spin operator (of spin\n1/2) of a localized electron at site i. The hopping pa-\nrameter between the nearest-neighbor sites is defined as\nti,i+1=/bracketleftbig\n1−(−1)iδ/2/bracketrightbig\ntwith the dimerization strength\nδof 0≤δ <2; we in particular define t±= (1±δ/2)t.\nJHis the strength of the Hund’s rule coupling. Lis the\nnumber of sites in the system, where the site contains a\nconduction orbital and a localized spin. We confine our-\nselvestothecaseatquarterfillingofconductionelectrons2\n-1 0 1-202(a)\n(b)\n(c)\n/c50/c100strong-dimerization limit\nweak-dimerization limit\nk//c112/c101( )/k tt-t+\nJH\nFIG. 1: (Color online) Schematic representations of the dou -\nble exchange model in (a) the weak dimerization regime and\n(b) strong dimerization regime. In (c), we illustrate the no n-\ninteracting band structure ε(k) of our model with the lattice\ndimerization δin the unfolded Brillouin zone.\nn=N/L= 1/2, where Nis the number of conduction\nelectrons in the system. We introduce δas a controllable\ninput parameter rather than the order parameter of the\nspontaneous lattice dimerization since our purpose is to\nstudytheeffectsof δontheelectronicstatesofthemodel.\nWe use the DMRG method to investigate the ground-\nstate properties of the system Eq. (1), where the open-\nend boundary conditions are applied. We study the\nmodel with several lengths of L= 8 to 20 with keeping\nm= 1200to3200density-matrixeigenstatesintherenor-\nmalization procedure; in this way, the largest truncation\nerror, or the discarded weight, is of the order of 10−11.\nNote that we have to keep relatively-numerous density-\nmatrix eigenstates to extract the true ground state from\na number of nearly degenerate magnetic states, in par-\nticular, in the vicinity of the phase boundaries. The ex-\ntrapolation to the thermodynamic limit L→ ∞is made\nin the results presented unless otherwise indicated.\nFirst, let us present the total-spin quantum number\nSof the system, which is calculated directly from the\nground-state expectation value of S2defined as ∝angbracketleftS2∝angbracketright=\nS(S+1) =/summationtext\nij/parenleftbig\n∝angbracketleftsi·sj∝angbracketright+2∝angbracketleftsi·Sj∝angbracketright+∝angbracketleftSi·Sj∝angbracketright/parenrightbig\n. Theresults\nare given in Fig. 2 as a ground-state phase diagram. We\nfind that there are three phases, S= 0, 0< S < S max,\nandS=Smax, depending on the values of JHandδ,\nwhereSmaxdenotes the full spin polarization. As ex-\npected, and in agreement with the previous calculations\natδ= 0 [13–16], the S=Smaxphase appears when\nJHis large and S= 0 phase appears when JHis small.\nAnd in between, the phase with the intermediate spinFI ( = )S Smax\nPI ( =0)SFM\nPMFI (0< < ) S Smax\n/c100JH/t\n0 1 20102030\nFIG. 2: (Color online) Calculated ground-state phase dia-\ngram of the 1D double exchange model at quarter filling with\nthe lattice dimerization δ. We find the FI (ferromagnetic in-\nsulating), PI (paramagnetic insulating), FM (ferromagnet ic\nmetallic), and PM (paramagnetic metallic) phases. The FI-\nPI phase boundary at δ→2 is determined by the analytical\nexpansion from the dimer limit and is shown by the open di-\namond and dashed line. The FM and PM phases at δ= 0 or\nJH= 0 are also indicated by the thick lines.\npolarization also appears as in Ref. 15 at δ= 0 (see also\nFig. 3(a) below). These phases are retained even when\nthe dimerization δis introduced. We note that the two\ncritical values of JHthat separate between the S=Smax\nandS= 0 phases increase with increasing δand that the\nregion with the intermediate spin polarization becomes\nnarrower and vanishes at δ→2 (see below).\nNext, let us calculate the charge gap ∆ defined as ∆ =\nlimL→∞∆(L) with ∆( L) =1\n2/bracketleftbig\nEN+2\n0(L) +EN−2\n0(L)−\n2EN\n0(L)/bracketrightbig\n, whereEN\n0(L) is the ground-state energy of the\nsystem of size LwithNelectrons. The gap ∆ is defined\nwith the prefactor 1 /2, so that the single-particle band\ngap is identical to ∆ in the present case where pairing\ninteractions are absent. The results are shown in Figs. 3\n(b)-(d). We find that the charge gap opens in the entire\nparameter space except at the lines δ= 0 and JH= 0.\nThe model at JH= 0 is trivial, where the conduction\nelectrons, decoupled completely from the localized spins,\nbehave just as the noninteracting electrons, resulting in\nthe PM phase.\nThe model Eq. (1) at δ= 0 on the other hand is highly\nnontrivial and much work has been done in recent years\n[13–16]: basically,thereappearstheferromagneticmetal-\nlic(FM) phaseforthelarge JHregion,whichchangesinto\nthe paramagnetic metallic (PM) phase when JHbecomes\nsmall. In addition, it has been claimed [15, 17, 18] that\nthere appears the region of phase separation in particu-3\nlar near half filling and the “spiral” phase with a long-\nwavelengthantiferromagneticcorrelationsat the FM-PM\nphase boundary. Although we have not detected any in-\ndications of the phase separation at least at quarter fill-\ning in our accurate DMRG calculations with very large\nmvalues, our results obtained are consistent with the\nresults of the previous work [13–16]: we find that either\nthe FM phase (when JHis large) or PM phase (when JH\nis small) is realized, and in between there is the partially\nspin-polarized metallic phase (see Figs. 2 and 3(a)) that\nmay correspond to the spiral phase predicted in Ref. 15.\n0 0.5 1012\n0 1 200.10.20.30.40.5\n0 1 200.20.40.6\nL=16\nL=12\nL=80 1 200.20.40.610 15 2000.20.40.60.81\nL=16\nL=12\nL=8\n0 1 200.20.40.6\nL=16\nL=12\nL=800.05 0.100.5\n/c100=0.1/c100=1.900.05 0.100.511.5\n/c100=0.1\n/c100/c100\nJH/t/c68/t\n/c68/t/c68/t\nS S/max\nS q( )S q( )(a)\n(b)\n(c)\n(d)\n(e) (f)JH=20/t\nJH=1 /t\n/c100/c61/c49/c46/c53JH/t/c100/c61/c48/c46/c49\nJ tH/ =2, /c100/c61/c48 J tH/ =2, /c100/c61/c48/c46/c53\nq//c112 q//c112/c68 /c61/c50/c100/t\ndimer limit1/L1/L\nFIG. 3: (Color online) (a) Calculated normalized total-spi n\nquantum number S/Smaxas a function of JH/t. Also shown\nare the calculated results for the charge gap ∆: (b) δdepen-\ndenceintheFIphase, (c) δdependenceinthePIphase(where\nthe dotted line indicates the result of the strong-dimeriza tion\nexpansion), and (d) JH/tdependence in the PI phase. Inset\nof (b) and (c) shows examples of the finite-size scaling of the\ngap. In (e) and (f), the calculated spin structure factors S(q)\nin the PI phase without and with the lattice dimerization,\nrespectively, are shown.\nLet us then introduce the lattice dimerization δ >0.\nThe results are the following: In the S=Smaxregion,\nwe find that the charge gap of ∆ /t= 2δopens as shownin Fig. 3(b). This can simply be understood because in\nthis region we have the noninteracting band of spinless\nfermions at half filling, and therefore the lattice dimer-\nizationδopens the band gap of the size ∆ /t= 2δ. The\nFI phase is thus realized. Since the FM phase due to\nthe double exchange mechanism is continuous to this FI\nphase, we may naturally refer to it as the insulator ver-\nsion of the double exchange ferromagnetism.\nIn theS= 0 region, we find that the gap actually\nopens as ∆ ∝δin the weak dimerization limit as shown\nin Fig. 3(c). The size of the gap increases as the value\nofJHincreases as shown in Fig. 3(d). This phase with\n∆>0 may then be denoted as the paramagnetic insu-\nlating (PI) phase. The Fourier transform of the spin-\nspin correlation function for the conduction electrons\nS(q) =1\nL/summationtext\ni,jeiq(Rj−Ri)∝angbracketleftsi·sj∝angbracketright(as well as that for the\nlocalizedelectrons, seeFig.2ofRef. 15) maycharacterize\nthis phase. The calculated results are shown in Figs. 3\n(e) and (f), wherewefind that the antiferromagneticspin\ncorrelation of the wavevector of q=π/2 is enhanced and\nthat the lattice dimerization inducing the localization of\nconduction electrons further enhances this correlation.\nThis results may therefore be interpreted as an enhance-\nment by the lattice dimerization of the “island” state\npredicted in Ref. 15, where the high-spin ( S= 3/2) clus-\nters, formed by a conduction electron coupled ferromag-\nnetically with the two neighboring localized spins, are\narrangedantiferromagneticallywith the quasi-long-range\norder [15, 19]. In higher spatial dimensions, this phase\nmay well fall into the true long-range antiferromagnetic\norder, resultingin the antiferromagneticinsulating(AFI)\nphase. Thissituationresemblesthatofthe“dimer-Mott”\nphase[20, 21] in the dimerized Hubbard model atquarter\nfilling although in the latter the spins are of S= 1/2.\nIn the 0< S < S maxregion, the charge gap also opens\natδ >0, where its size increases rapidly with increasing\nJH/t. We thus have the FI phase here as well, which may\nbe the insulating spiral phase continuous to the metallic\none predicted in Ref. 15.\nNow, let us discuss the strong dimerization limit,\nwhere we start with the highly correlated clusters Cl\n(l= 1,···,L/2) coupled weakly to each other through\nthe hopping parameter t−(see Fig. 1(b)). Each of the\nclusters consists of the two conduction orbitals and two\nlocalized spins. In the ground state, the cluster contains\noneconduction electron(andtwolocalizedspins) and the\ninternal three spins are fully polarized. The lowest en-\nergy of the single cluster is e(1) =−JH/4−t+, where\nthe conduction electron is in the bonding state of the\ntwo conduction orbitals. The eigenstates of the cluster\nwith the 0 and 2 conduction electrons are also derived\nexactly and the lowest energies are found to be e(0) = 0\nande(2) =−/radicalBig\nJ2\nH+16t2\n+/slashbig\n2, respectively. In the strong-\ndimerization limit of the PI phase, we can therefore map\nour system Eq. (1) onto an effective single-band Hub-4\nbard model defined in terms of the bonding orbital on\neach dimer. The Hamiltonian may be written as\nHeff=teffL/2/summationdisplay\ni=1,σ(b†\ni,σbi+1,σ+H.c.)+UeffL/2/summationdisplay\ni=1nb\ni,↑nb\ni,↓,(2)\nwith the creation operator of an electron on the bond-\ning orbital b†\ni= (c†\n2i−1,σ+c†\n2i,σ)/√\n2 andnb\ni,σ=b†\ni,σbi,σ.\nWe obtain the effective hopping integral teff=t−/2\nand effective “on-dimer”” Coulomb interaction Ueff=\ne(2)+e(0)−2e(1). In this mapping, the localized spins\ncontribute only to Ueffand their degrees of freedom are\nnot explicitly involved in the operator b†\ni,σ. An analytical\nexpression for the charge gap may thus be derived as\n∆ =Ueff−2t−\n=−/radicalBig\nJ2\nH+4t2(2+δ)2/2+2tδ+JH/2,(3)\nwhichisvaliduptothefirstorderof2 −δ. Theresultthus\nobtained is shown in Fig. 3(c) as a dotted line, where we\nfind that the agreement with the DMRG result is very\ngood. In the FI phase, on the other hand, the charge\ngap is always ∆ /t= 2δin its entire region, independent\nofJH. The two gaps are therefore discontinuous at the\nphase boundary.\nNext, we derive the effective exchange interaction Jeff\nbetween conduction electrons on the neighboring clus-\nters. We first take a direct product of the isolated clus-\nters (/producttextL/2\nl=1Cl) to be the unperturbed ground state of the\nsystem. Then, taking into account all the processes up\nto the second order of the hopping t−, we obtain the\nexpression for Jeffas\nJeff=t2\n−/bracketleftBig(J2\nH+8t2\n+)\n(JH+4t+)(J2\nH+16t2\n+)+1\n8(JH+2t+)\n+(/radicalBig\nJ2\nH+16t2\n++4t+)2(JH+4t+)\n16JHt+(J2\nH+16t2\n+)−3\n16t+/bracketrightBig\n,(4)\nwhere we should note that, if only the lowest intermedi-\nate state is taken into account in the second-order pro-\ncess, we obtain Jeff= 4t2\n−/Ueff, which is always positive\n(or antiferromagnetic). We thus find that, depending on\nthe value of Jeff(either positive or negative), the ground\nstate realized is either the PI phase or the FI phase. The\ncritical value of JHatδ→2 is found to be 29 .004t,which\ndetermines the FI-PI phase boundary at δ→2. The re-\ngion of the intermediate spin state does not appear here.\nThe result for the phase boundary obtained from Eq. (4)\nis shown as a dashed line in Fig. 2, where we find again\nthat the agreement with our DMRG result is very good,\nreinforcing the validity of our phase diagram of Fig. 2.\nIn summary, we studied the effects of opening of the\nband gap on the double exchange ferromagnetism. We\napplied the DMRG technique and analytical expansion\nfrom the dimer limit to the 1D double exchange modelat quarter filling with lattice dimerization and obtained\nthe ground-state phase diagram. We found three phases:\nthe FI phase with the Peierls gap and full spin polariza-\ntion, the PI phase with the Mott gap and dominant an-\ntiferromagnetic spin correlations, and the FI phase with\nthe partial spin polarization. The results for JH/t/greaterorsimilar15\ndemonstrated that, in the weak dimerization regime, the\nPeierls gap opens in the fully spin-polarized conduction\nband without affecting its ferromagnetism. Therefore,\nthe metallicity itself is not a necessary condition for\nthe realization of the double exchange ferromagnetism.\nThe concept of the insulator version of the double ex-\nchange ferromagnetism was thus established. In the\nstrong dimerization regime, on the other hand, the ferro-\nmagnetism is destroyed at JH/t/lessorsimilar29 and the Mott gap\ndue to the effective on-dimer Coulomb interaction opens\nthere with the antiferromagnetic quasi-long-range order\nin the system.\nThe uncommon FI state realized in K 2Cr8O16[7] then\nmeans that this material is in the weak dimerization\nregime with the Peierls gap and full spin polarization.\nA recent experiment suggests [22] that, by applying high\npressures of /greaterorsimilar2 GPa, the FM phase is suppressed very\nrapidly, while the metal-insulator transition remains al-\nmost unchanged, leading to the transition from the PM\nphase to the PI or AFI phase by lowering temperature.\nIt may then be quite interesting to point out that, if the\napplied pressure decreases the value of JH/t, this might\ncorrespond to the phase change in the quasi-1D chains\nfrom the FI to PI phase as in our phase diagram given\nin Fig. 2, where the intermediate spin state, or the spiral\nstate of Ref. 15, may also be predicted to appear under\nhigh pressure. We hope that ourworkpresentedhere will\nstimulate further searches for new phenomena and ma-\nterials with intriguing magnetic and transport properties\nderived from the interplay between the double exchange\nand Peierls/Mott mechanisms.\nEnlightening discussions with R. Eder, D. I. Khomskii,\nT. Konishi, K. Nakano, T. Toriyama, Y. Ueda, and T.\nYamauchi are gratefully acknowledged. This work was\nsupported in part by a Kakenhi Grant No. 22540363 of\nJapan. A part of computations was carried out at the\nResearchCenter for Computational Science, OkazakiRe-\nsearch Facilities, Japan.\n[1] D.I.KhomskiiandG.A.Sawatzky, SolidStateCommun.\n102, 87 (1997).\n[2] C. Zener, Phys. Rev. 81, 440 (1951); 82, 403 (1951).\n[3] P. W. Anderson and H. Hasegawa, Phys. Rev. 100, 675\n(1955).\n[4] P.-G. de Gennes, Phys. Rev. 118, 141 (1960).\n[5] K. Hasegawa, M. Isobe, T. Yamauchi, H. Ueda, J. Ya-\nmaura, H. Gotou, T. Yagi, H. Sato, and Y. Ueda, Phys.\nRev. Lett. 103, 146403 (2009).5\n[6] M. Sakamaki, T. Konishi, and Y. Ohta, Phys. Rev. B 80,\n024416 (2009); 82, 099903(E) (2010).\n[7] T. Toriyama, A. Nakao, Y. Yamaki, H. Nakao, Y. Mu-\nrakami, K. Hasegawa, M. Isobe, Y. Ueda, A. V. Ushakov,\nD. I. Khomskii, S. V. Streltsov, T. Konishi, and Y. Ohta,\nPhys. Rev. Lett. 107, 266402 (2011).\n[8] A. Nakao, Y. Yamaki, H. Nakao, Y. Murakami, K.\nHasegawa, M. Isobe, and Y. Ueda, J. Phys. Soc. Jpn.\n81, 054710 (2012).\n[9] S. R. White, Phys. Rev. Lett. 69, 2863 (1992); Phys.\nRev. B48, 10345 (1993).\n[10] H. Y. Hwang, S.-W. Cheong, P. G. Radaelli, M. Marezio,\nand B. Batlogg, Phys. Rev. Lett. 75, 914 (1995).\n[11] T. Mizokawa, D. I. Khomskii, and G. A. Sawatzky, Phys.\nRev. B61, R3776 (2000).\n[12] J. Geck, P. Wochner, S. Kiele, R. Klingeler, A.\nRevcolevschi, M. v. Zimmermann, B. B¨ uchner, and P\nReutler, New J. Phys. 6, 152 (2004).\n[13] J. Kienert and W. Nolting, Phys. Rev. B 73, 224405\n(2006).[14] S. Henning and W. Nolting, Phys. Rev. B 79, 064411\n(2009).\n[15] D. J. Garcia, K. Hallberg, B. Alascio, and M. Avignon,\nPhys. Rev. Lett. 93, 177204 (2004).\n[16] M. Gul´ acsi, Adv. Phys. 53, 769 (2004).\n[17] S. Yunoki,J. Hu, A.L.Malvezzi, A.Moreo, N.Furukawa,\nand E. Dagotto, Phys. Rev. Lett. 80, 845 (1998).\n[18] E. Dagotto, S. Yunoki, A. L. Malvezzi, A. Moreo, J. Hu,\nS. Capponi, D. Poilblanc, and N. Furukawa, Phys. Rev.\nB58, 6414 (1998)\n[19] See, e.g., P. Fazekas, Lecture Notes on Electron Correla-\ntion and Magnetism (World Scientific, Singapore, 1999),\npp. 383-384.\n[20] K. Penc and F. Mila, Phys. Rev. B 50, 11429 (1994).\n[21] H. Seo, J. Merino, H. Yoshioka, and M. Ogata, J. Phys.\nSoc. Jpn. 75, 051009 (2006).\n[22] T. Yamauchi, K. Hasegawa, H. Ueda, M. Isobe, and Y.\nUeda, unpublished." }, { "title": "2008.06757v1.The_inverse_proximity_effect_in_strong_ferromagnet_superconductor_structures.pdf", "content": " 1 \n The inverse proximity effect in strong \nferromagnet -superconductor structures \nV. O. Yagovtsev1, N. G. Pugach1,2 and M. Eschrig3 \n1 National Research University Higher School of Economics, Moscow, 101000 , Russia \n2 Skobeltsyn Institute of Nuclear Physics, Lomonosov Moscow State University, Leninskie Gory, 1 (2), Moscow GSP -2, 119991 , Russia \n3 Institut für Physik, Universität Greifswald, D-17489 Greifswald , Germany \n \nE-mail: vyagovtsev @hse.ru \n \nAbstract \nThe magnetization in a superconductor induced due to the inverse proximity effect is investigated in hybrid \nbilayers containing a superconductor and a ferromagnetic insulator or a strongly spin -polarized ferromagnetic \nmetal. The study is performe d within a quasiclassical Green function framework, wherein Usadel equations are \nsolved with boundary conditions appropriate for strongly spin -polarized ferromagnetic materials. A comparison \nwith recent experimental data is presented . The singlet to triple t conversion of the superconducting correlations as \na result of the proximity effect with a ferromagnet is studied. \nKeywords: superconductivity, superconductor -ferromagnet, superconductor -ferromagnetic insulator, inverse \nproximity effect, induced magnetiza tion \n \n1. Introduction \nIn classical electronics, charge currents are used to transfer \ninformation, while in spintronics, spin -polarized currents are \nused for this purpose [1]. The heat, generated in the process of \nusing spin -polarized currents , can be an undesirable spurious \neffect. The use of superconductors in magnetic nanostructures \ncan greatly reduce this heating, increasing the energy \nefficiency of spintronic devices. This idea underlies \nsuperconducting spintronics. \nIn the last 30 years, superconducting spintronics has b een \nactively studied by many experimental and theoretical groups \n[2–6]. The main focus of the research was the theoretical and \nexperimental description of hybrid nanostructures. In \nparticular, the proximity effect in superconductor -ferromagnet \n(SF) structures was inves tigated. Its main features are spatial \noscillations of the amplitude of the superconducting \ncorrelation penetrating into the ferromagnet [4,5,7 –9] and the \nappearance of triplet superconducting correlations with \ninhomogeneous magnetization of the layers [6,10] . It leads to \na long -range proximity effect [5,6] . This was discovered in \nexperimental studies and it is now known that the long -range proximity effect leads to change in the critical temperature of \nsuperconductors in bilayer and multilayer SF structures, \noscillations of the Josephson current in the presence of a \nferromagnet, and the long -range Josephson effect [2–5]. \nDevices of superconducting spintron ics are considered as \npromising to create sensitive sensors and element base for a \nquantum computer [5,11] . \nThe inverse proximity effect in a superconductor in contact \nwith a ferromagnet was first described in the work [12]. \nTriplet superconducting correlations are the origin of the non -\nzero induced magnetization in the superconductor due to \ninverse proximity effect. Triplet Cooper pairs are symmetric \nin spin space [13–16]. In theoretical work [17] the induced \nmagnetization in such structures was estimated in the presence \nof long -range triplet supe rconducting correlations . In \nexperimental studies [18,19] evidence of the presence of \ninduced magnetization in the proximity of a superconductor to \na ferromagnet ic metal was demonstrated. \nThe origin of the magnetization is a spin -splitting of the \nlocal density of states in the superconductor near the SF \ninterface. This spin -splitting is directly related to the spin -\npolarization of the Cooper pair . As a result, they will create a \n 2 \n noticeable magnetic field, which has the feature that it depends \non the concentration of Cooper pairs. In the absence of \nsuperconductivity, this magnetic field will be restricted to a \nmuch shorter length scale, related to the Thomas -Fermi length \n[20]. \nAnother way to explain the appearance of magnetization is \nthe spin mixing angle [21]. It quantifies how large the relative \nscattering phase shift between the electrons of the Cooper pair \nis after they are reflected from an S–F interface. This \ndifference between phase of the spin -up and spin -down \nelectrons of the Cooper pair results in an imbalance in the \nDOS for electrons with different spin , and it leads to the \nappearance of an induced magnetization. \nHowever, most of these t heoretic al and experimental \nstudies were focused on structures with ferromagnetic metals \n(FM). Later phenomen a associated with the inverse proximity \neffect in structures with ferromagnetic insulators (FI) were \nexperimentally demonstrated [21,22] . In addition, in \nexperiments [23–25] phenomena related to the proximity \neffect of such structures were demon strated. They include \nsplitting in the density of states of the superconductor due to \nthe effective exchange field created by the proximity with a \nferromagnetic insulator [23] and the possibility of \nmanipulating spin transfer by adding ferromagnetic insulators \nto the s uperconducting layer [25]. \nThe present work is devoted to the theoretical descr iption \nof induced magnetization in these structures. In Section 3, a \ncomparison is made with experimental studies [25–27], where \nthe value of the induced magnetization is obtained. In the \ntheoretical model of induced magnetization in S -FI and S -FI-\nS structures [28,29] the Hamiltonian consists of a general \nHamiltonians of the S layers and a tunnelling term with spin -\ndependent matrix elements. The latter represent the band \nsplitting in the ferromagnetic barrier which leads to different \ntunnel barrier heights for spin -up and spin -down electrons, \nand consequently, to differences between tunnelling \namplitudes for electrons with different spin projections . \nElectrons with spin -up and spin -down will penetrate the \ninsulating barrier of the S-FI bilayer to different depths. In \nboth cases this will lead to a spin imbalance in the S layer near \nthe S-FI interface and produce spin magnetization in the \nsuperconductor. And again, absence of Cooper pairs will \nresult in absence of this mechanism for induced \nmagnetization. \nOur work also describes induced magnetization in \nstructures with strongly ferromagnetic metals (Section 2.3). \nThe work is based on the quasiclassical app roximation using \nUsadel equations with boundary conditions appropriate for the \ncase of strong spin polarization of the ferromagnet [21,30] . \nThese equations are applicable in the dirty limit of a \nsuperconductor, in which the mean free path of an electron is \nmuch l ess than the coherence length of a Cooper pair . \nFigure 1. The schematic view of the simulated \nsuperconducting -ferromagnet structure. \nThe structure we describe is shown schematically in Fig. 1. \nLayer S is the superconductor layer, and F is a ferromagnetic \ninsulator or ferromagnetic metal. Here L is the thickness of the \nsuperconduct ing layer . \nSection 2.1 provides a model describing the S -F bilayer \nusing linearized U sadel equations. The solution to these \nequations serves as the basis for calculating the magnetization \ninduced in the superconductor. Section 2.2 shows boundary \nconditions and results for the induced magnetization for the S-\nFI case, Section 2.3 .1 shows them for the S -FM case at the \ntemper ature close to its critical value and Section 2.3 .2 \npresents the equations for S-FM case in the limit of weak \nproximity effect . \nSection 3.1 presents the determination of the angle of spin \nmixing from comparison with the theoretical model [31], and \nSection 3.2 presents the determination of the angle using \nrelated experimental data [25–27]. Section 3.3 is devoted to \nthe results of the numerical calculations and shows how the \nmagnetization depends on various parameters of the model. \nSection 4 discusses the work’s main results. \n \n 3 \n 2. Model \n2.1. Equations for superconductor in contact with \nferromagnet \nTo estimate the value of the induced magnetization in the \nsupercon ductor we use the quasiclassical approximation in the \nform of Usadel equations, assuming that the mean free path is \nthe smallest characteristic length in all layers of our structure. \nIn the ferromagnet, uniform magnetization is assumed over \nthe entire volume of the material . It is assumed that the \ndiffus ive limit is realized in the structure, as real \nsuperconducting nanostructures created by sputtering usually \nsatisfy the conditions of the diffusive, or “dirty” limit . The \ncoherence length in the superconductor is 𝜉=\n√ℏ𝐷𝑠2𝜋𝑘B𝑇cb ⁄ , where ℏ is the reduced Planck constant, 𝑘B \nis the Boltzmann constant (we use units where ℏ=𝑘B=\n1), 𝐷𝑠 is the electron diffusion coefficient of the \nsuperconductor , and Tcb is the criti cal temperature of the bulk \nsuperconductor. \nThe linearized U sadel transport equations within the \nsuperconductor, applicable in the dirty limit [32] with \ntemperature in the vicinity of the S film’s critical \nsuperconducting temperature Tс, are: \n(𝐷𝑠𝛁2−2|𝜔𝑛|)𝑓𝑠=−2𝜋Δ,\n(𝐷𝑠𝛁2−2|𝜔𝑛|)𝐟𝑡=0. (1) \nHere the Matsubara frequencies are 𝜔𝑛=𝜋𝑇(2𝑛+1),𝑛 is \nan integer, T is the temperature, 𝑓𝑠 and 𝐟𝑡 are the singlet and \ntriplet components of the anomalous Green function, \nrespectively, and 𝐟𝑡 is the vector (𝑓𝑡𝑥,𝑓𝑡𝑦,𝑓𝑡𝑧), Δ is the scalar \nsuperconducting order parameter which we assu me to be real -\nvalued, and in the bulk equal to the superconducting energy \ngap. As a result of the influence of the ferromagnetic layer, the \nsinglet superconducting correlations near the interface with \nthe insulator is suppressed and converted to the triplet ones, its \nvalue depends on the distance to the S -F interface . The F layer \nmagnetization is considered to be aligned along the z axis. \nThe boundary condition at the boundary of the \nsuperconductor with the (outer) environment has the form \n𝜕𝑓𝑠\n∂𝑥=𝜕𝑓𝑡𝑧\n∂𝑥=0, \nreflecting zero current flow through this interface. \nIn both cases of FI and FM, after finding the Green's \nfunctions by solving the Usadel equation with boundary \nconditions, it is necessary to take into account the suppression \nof the superconducting correlations. The suppression of the \nsinglet component at the interface in the limit of small \ntransparency of the S -F interface goes like the square of the \nspin-mixing angle, and can thus be neglected because we are \ninterested only in effects of linear order in the spin -mixing \nangle [33]. To show this, we calculate the singlet and triplet \npair amplitudes and make sure that the order parameter has almost constant value. In order to do this, the self -consistent \nequation [15] \n∆(x)ln𝑇c\n𝑇=𝜋𝑇∑(∆(x)\n𝜔𝑛−𝑓𝑠(𝑥)\n𝜋).\n𝑛>0 (2) \nIs also solved approximately in the appropriate case . The \ninduced magnetization i s calculated by the formula [15]: \n𝛿𝑀(𝑥)=2𝜇𝐵𝑁0π𝑘B𝑇∑𝐠(𝑥,𝜔𝑛)\n𝑛>0, (3) \nwhere 𝜇𝐵 is the Bohr magneton, 𝑁0 is the density of states \nat the Fermi level in the normal state . \nEquations (1) are supplemented by the boundary condition \nfor the boundaries parallel to the YOZ plane. In our work, we \nconsider boundary conditions appropriate for a ferromagnetic \ninsulator and for a ferromagnetic metal. \n2.2. Bilayer with f erromagnetic insulator \nIn this case the bound ary condition [21,30] at the boundary \nof the S and FI layers is: \n𝐴𝜎𝑆g0𝜕(𝑓𝑡𝑧+𝑓𝑠)\n∂𝑥=−2𝑁𝐺𝑄𝑃(φ), (4) \nwhere A is the contact area, 𝜎𝑆 is the conductivity of the \nsuperconductor in the non -superconducting state, g0 is the \nsinglet component of the Green function g, N is the number of \nconduction channels at the boundary (in the general case, it is \ndetermined by the polarization of the ferromagnet, the \ncoefficient of particle transmission through the boundary, and \nthe conductivity of the boundary ), 𝐺𝑄=𝑒2\n𝜋⁄ is the quantum \nof conductivity. P(φ) is the function defined in [21], which \ndepends on the value of φ, the angle of spin mixing. \nThe spin mixing angle describes how s trongly the magnetic \nexchange field affects the phase diff erence for electrons with \nspin-up and spin -down at the superconductor -ferromagnet \ninterface. Since the ferromagnetic insulator may be considered \n[28,29] as a spin-dependent potential barrier , spin-up and spin -\ndown electrons penet rate into FI at different depths. Th e \nsimplest approach leads to the assumption that the stronger the \nmagnetic exchange field in the ferromagnetic material, the \ngreater the difference in penetration depth , and t herefore, the \nphase difference created by the boundary will be larger. \nHowever , it was shown [34] that the dependence of the created \nphase difference on the exchange splitting is nonlinear at large \nsplitting . The angle of spin mixing φ can be estimated using \nthe model given in the work [31]. \nThe vector component of the normal Green function g is \nexpressed through the combination of the components of the \nanomalous Green function, \n𝑓𝑓̃+g2=σ0. \nHere σ0 is a unit matrix, g=g0+(𝐠∙𝛔),𝑓=[𝑓𝑠+\n(𝒇𝑡∙𝛔)]𝑖σy, where g is the spin vector part of the normal \nGreen function g, 𝛔 is the vector of spin Pauli matrices. \nA linear approximation was used, which is justifie d by a \nsmall change in the normal (diagonal part of the Nambu - \n 4 \n Gor’kov matrix) Green function and the smallness of the \nanomalous (off-diagonal) Green function . To obtain an \nanalytic solution of the Usadel equation with two boundary \nconditions, the limit of temperature close to the critical \ntemperature was taken. \nAfter expanding the normal Green function g into a Taylor \nseries over the anomalous Green function components and \ndiscarding the terms of third and higher order using equation \n(3), one obtain s the following expression for the induced \nmagnetization [17]: \n𝛿𝑴(𝑥)=−4𝜇𝐵𝑁0π𝑘𝐵𝑇∑Im(𝑓𝑠(𝑥,𝜔𝑛)𝑓̃𝑡𝑧(𝑥,𝜔𝑛)).\n𝑛>0(5) \nNote that neither a pure singlet nor a pure triplet pair can \ncreate such a spin -splitting, only the simultaneous presence of \nboth, see discussion around Eq. (87) in review [5]. \nWe assume that i t is possible to neglect the Meissner effect \nfor this structure, which acts only at lengths of the order of the \nLondon penetration depth λ, [35] i.e. it is assumed that the \nthickness of the S layer is 𝐿~𝜉𝑆≪λ. \n \n2.3. Ferromagn etic metal case \n2.3.1 Limit of temperature close to the critical \ntemperature \nNow we turn to the case of a ferromagnetic metal (FM) as \nmagnetic element in the bilayer . We consider the case of a \nstrongly spin -polarized ferromagnet where the period of space \noscillations of the superconducting correlations and its decay \nlength are much smaller than the mean free path lF and the \nlayer thi ckness L. It means that the exchange magnetic energy \nH is large enough to satisfy the inequality ξH <𝐿. For the same reason as for S -FI case, \nwe assume that 𝐿 is much less than the London penetration \ndepth λ. We also consider limit of temperature close to the \ncritical temperature. \nThe Usadel equation in the superconductor has the s imilar \nform as (1), \n(𝐷𝑠\n𝜋g𝑑2\n𝑑𝑥2+2𝑖𝜔)(𝑓𝑠±𝑓𝑡𝑧)=−2Δg, but the boundary condition at the boundary of the S and F M \nlayers is different from the S -FI case. F or the ferromagnetic \nmetal the boundary condition [36] at the interface ( 𝑥=𝐿) may \nbe written in the simple closed form \nℏ𝐴𝜎𝑆\n𝑒2𝜋𝜔𝑛\n√𝜔𝑛2+∆2𝜕(𝑓𝑠±𝑓𝑡𝑧)\n∂𝑥=(𝑡∓𝑖φ)(𝑓𝑠±𝑓𝑡𝑧), (6) \nwhere 𝑡=𝑡↑+𝑡↓ is the probability for the particles to be \ntransmit ted through the S-FM interface . Here, 𝑡↑=\n∑|𝑡↑↑𝑛|2\n𝑛↑ ,𝑡↓=∑|𝑡↓↓𝑛|2\n𝑛↓ are the transmission probability \nmatrix elements for each of the 𝑛↑ or 𝑛↓ transmission channels \nwith spin up ( ↑) or down ( ↓), respectivel y [36]. We assume no \nspin-flip scattering on the SF interface, therefore the \ntransmission matrix has a diagonal form. There are different \nnumbers of transmission channels for spin -up and spin -down \nin the ferromagnet, therefore 𝑛↑≠𝑛↓. \nThe imaginary part 𝑖φ appears due to the spin splitting in \nthe ferromagnet. For a structure without Josephson current, \nwhen ∆ may be assumed to be real, (𝑓𝑠−𝑓𝑡𝑧)=(𝑓𝑠+𝑓𝑡𝑧)∗. \nEarlier calculation s of the proximity effect between a \nsuperconductor an d a strong ferromagnet [37,38] were \nperformed for only a very thin F layer L ~ξH. In this case the \nboundary conditions at the SF interface had approximately the \nsame form [38] with a complex coefficient between the \nfunctions (𝑓𝑠±𝑓𝑡𝑧) and their derivati ves, but these \ncoefficients oscillated with thickness dF. \nThe inverse proximity effect leads to the singl et to triplet \ntransform of the superconducting correlations near the SF \ninterface. If L~ξ, the order parameter and the anomalous Green \nfunction 𝑓𝑠+𝑓𝑡𝑧=𝜋∆(𝑥)\n|𝜔𝑛|+𝐷𝑠𝑘2/2 may be determined by the \nansatz [38] \n∆(𝑥)=∆0cos[𝑘(𝑥−𝐿)]. (7) \nThis ansatz satisfies the Usadel equation (1) with the \nboundary condition at the free interface. For this case, for the \ndependence of all the quantities of interest, we will \nconcentrate only on thicknesses of the order of the coherence \nlength, since the ansatz does not work well at larger \nthicknesses . \nThe value of k is found from the boundary condition (6) that \nleads t o the equation \n𝑘∙tan(𝑘𝐿)=𝑒2\n2𝜋2ℏ𝐴𝜎𝑆(𝑡−𝑖φ). (8) \nNote that here k depends on the values 𝐿, 𝑡 and φ, and does \nnot depend on 𝜔𝑛. Using the obtained value of k we may \ncalculate properties of the proximity structure (consisting of a \nsuperconductor – strongly spin -polarized ferromagnet bilayer) \nas functions of the parameters of the SF boundary , 𝑡 and φ. \nAs it was mentioned above, the inverse proximity effect \nbetween a ferromagnet and a superconductor leads to an \ninduced magnetization at the S side [14]. Such induced \nmagnetization was detected in experiments using the nuclear \nmagnetic resonance [19], and the polar Kerr effect [39]. The \nvector of this induced magnetization ma y be found by the \n 5 \n formula (3). Using this formula, we calculated the \nmagnetization at the S -FM interface depending on the \nthickness of the superconductor L. \n \n2.3.2 The weak proximity effect limit \nHere w e assume that the transparency of the SF interface is \nsmall, and therefore the Green functions weakly differ from \ntheir bulk values g=−𝑖𝜔𝑛\n√𝜔𝑛2+∆2, 𝑓=∆\n√𝜔𝑛2+∆2 and 𝑓𝑡𝑧,𝑓𝑠 are \ndefined b y the boundary condition (6). \nThe corrections to the bulk values of functions g and 𝑓 \nappear only in the second order on 𝑓𝑡𝑧 and may be neglected. \nThe solution of the Usadel equation ( 1) satisfying boundary \ncondition at the free interface has the form 𝑓𝑡𝑧=\n𝐶cosh 𝑘𝑆(𝑥−𝐿), where 𝑘𝑆=2√𝜔𝑛2+∆2ℏ𝐷𝑠 ⁄ . The \nconstant C may be found from th e boundary condition at the \nSF interface ( 6), \n𝐶=𝑖𝑓φ𝑠𝑖𝑔𝑛(𝜔𝑛)(𝜋𝐴𝜎𝑆\n𝑒22|𝜔𝑛|\n𝐷𝑠𝑘𝑆sinh 𝑘𝑆𝐿−𝑡cosh 𝑘𝑆𝐿)−1\n. \nAs it is expected, the amplitude 𝐶 of the triplet component \nis proportional to the phase shift φ, which the anomalous \nfunction experiences at the interface with the ferromagnet. \nThis value depends on the exchange field 𝐻 and is equal to \nzero in the absence of 𝐻. The expression for the induced \nmagnetization ( 5) may be written as \n𝛿𝑀(𝑥)=−4𝜇𝐵𝑁0π𝑘𝐵𝑇∙ \n∙∑∆2\n𝜔𝑛√𝜔𝑛2+∆2\n𝜔𝑛>0φcosh [𝑘𝑆(𝑥−𝐿)]\n𝜋𝐴𝜎𝑆\n𝑒22𝜔𝑛\n𝐷𝑠𝑘𝑆sinh 𝑘𝑆𝐿−𝑡cosh 𝑘𝑆𝐿. \nFor a thick superconductor for 𝐿»𝜉 we may write this \nequation in the form: \n𝛿𝑀(𝑥)=−4𝜇𝐵𝑁0π𝑘𝐵𝑇∑∆2\n𝜔𝑛√𝜔𝑛2+∆2\n𝜔𝑛>0φexp(−𝑘𝑆𝑥)\n𝜋𝐴𝜎𝑆\n𝑒22𝜔𝑛\n𝐷𝑠𝑘𝑆−𝑡. \nThe induced magnetization decreases exponentially from \nthe SF boundary at a distance of the order of the \nsuperconducting coherence length 𝜉 at low temperature. \n3. Results \n3.1. Comparison with experiment \nThe spin mixing angle can be estimated by comparing the \nmagnetization obtained for different values of the angle with \nthe magnetization which was taken from experimental studies \n[25–27]. In these works, the conductivity of the structures \nunder study was measured, and the densities of states (DOS) \nwere obtained. The values of the exchange field in the \nsuperconductor were obtained from DOS. \nTable 1 shows values of the spin -mixing angle that were \nobtained from experimental studies as a result of fitting the \ndata. In these works, Al was taken as a s uperconductor and EuS was taken as a ferromagnetic insulator. 𝜎𝑆 was taken as \n3.8∙107 𝑂ℎ𝑚−1∙𝑚−1. If it were not mentioned in the article, \n𝐷𝑠 was taken as 8.68∙10−3𝑠−1∙𝑚2, because we can assume \nthat its value in all articles is approximately the sa me due to \nalmost identical production conditions. \n \n№ L, nm 𝐷𝑠, s−1∙m2 𝐵, T φ \n22 10 2∙10−3 1.75 0.033 \n23 10-15 8.68∙10−3 2.25-4.5 0.14-0.27 \n24 20 8.68∙10−3 0.25 0.8 \nTable 1. Comparison with the experimental works. \nOne can see that the spin mixing angle does not reach its \nmaximum value π/2. This is consistent with the theoretical \ndependence of the magnetization on the spin mixing angle . \nThe model calculation fit with the experimental data yields \nsome variation of the spin mixing angle for different \nsuperconductor thicknesses , although the spin-mixing angle \ndepends only on the properties of the interface between \nmaterials. Probably this is a consequence of the constant order \nparameter approximation, which nevertheless allows \nestimation of φ. \n3.2. Estimation of the angle of spin mixing \nTo determine the spin mixing angle , we used a comparison \nwith the boundary condition [31] written in a linearized form. \nAfter carrying out the corresponding transformations, we \nobtain the following relationship between the \nphenomenological parameters: \nsin(φ)=−𝐴g0\n2𝑁GQ𝑅𝐵𝐿GΦ\nGT, \nHere, 𝑅𝐵 is the barrier resistance, which is a \nphenomenological parameter, and GΦ and GT are the \nphenomenological parameters given in the work [31]. \nThe following estimat es were obtained for the spin mixing \nangle in radians: φ=0.82 at 𝑁𝐴⁄=2.48∙1019 μm−2, 𝑇=\n0.8∙𝑇𝑐, 𝑇𝑐=1.2 К, 𝐿=5𝜉, 𝜉= 93 nm. \nParameters appropriate for aluminium are taken for the \nsuperconductor. Aluminium has 𝑇𝑐=1.2 К, and in the \ncurrently used preparation technology 𝜎=3.8∙107 Ohm−1∙\nm−1, 𝐷=8.68∙10−3 s−1∙m2. Parameters appropriate for \ncobalt are taken for the ferromagnet metal in our work. \n3.3. Numerical results and discussion \nThe magnetization profile was obtained over the thickness \nof the superconducting layer as a function o f the spin mixing \nangle . The structure was modelled with the following \nparameters: 𝑁𝐴⁄=2.48∙1019 μm−2, 𝑇=0.8∙𝑇c, 𝑇c is 1.2 \nK, 𝜎=3.8∙107 Ohm−1∙m−1, 𝐷=8.68∙10−3 s−1∙m2, \n 6 \n 𝐿=4𝜉, 𝜉= 93 nm, 𝑡=0.7. All dependencies are presented \nfor given parameter values, unless otherwise indicated. \nThe calculation of pairing amplitudes was carried out: the \nvalues of Green functions with a constant order parameter \nwere substituted into the equation (2). The result is presented \nin Fig. 2. It allo ws estimation of the suppression of the \nsuperconducting correlations at the boundary. It is seen that \nthe superconducting correlations, and consequently, the order \nparameter of the superconductor is weakly suppressed by the \nmagnetization of the neighbourin g insulator layer at small spin \nmixing angles. \n \n \nFigure 2. Pairing amplitudes of the superconductor for S -FI \ncase as a function of the coordinate for various values of the \nangle of spin mixing, φ. \nThe obtained dependences of the components of the \nanomalou s Green function are presented in \nFig. 3. \nThe relations between the various components of the \nanomalous Green function are: \n𝑓𝑆=𝑓↑↓−𝑓↓↑\n2,𝑓𝑡𝑧=𝑓↑↓+𝑓↓↑\n2. \nThe dependences of the components 𝑓𝑆 and 𝑓𝑡𝑧 of the \nanomalous Green function on the x coordinate, shown in Fig. \n3, demonstrate that the singlet component of the Green \nfunction is suppressed when approaching the F boundary . The \ntriplet component, on the contrary, decreases when moving \naway from the S -FI boundary. This corresponds to physical \nexpectations, according to which, singlet Cooper pairs are \ntransformed into triplet pairs at the S -FI border. \nThe functions shown in fig. 3 were substituted into formula \n(5). As a result, the induced magnetization values were \nobtained. \n \nFigure 3. Real and imaginary components fs and ftz of the \nanomalous Green functions in the superconductor as \nfunctions of the coordinates , of (a) S -FI interface, (b) S-FM \ninterface for the temperature close to the critical one, (c ) S-FM \ninterface for the weak proximity effect. \n \n 7 \n \nFigure 4. Magnetization versus the x coordinate for various \nvalues of the angle of spin mixing, φ in the S layer of (a) S-FI \nbilayer , (b) S-FM bilayer for the temperature close to the \ncritical limit, (c) S-FM bilayer for the weak proximity effect \nlimit. As it can be seen from Fig. 4(a), a local maximum of \nmagnetization appears on the curve at the extremal value of \nthe coordinate x. It becomes clear that such a behaviour of \nmagnetization should have been e xpected. The reason for this \nbehaviour is the decrease of the singlet component and the \ngrowth of the triplet component when approaching the \nsuperconductor -ferromagnetic interface. \nWe can see from Fig. 4 (b, c) the absence of a local extreme \npoint of the i nduced magnetization profile at a contact with a \nferromagnetic metal. We may assume that this is a \nconsequence of the difference in boundary conditions, \ndescribing a nonzero transparency of the interface for \nelectrons in the metallic case. The S -FM interfa ce provides not \nonly an inverse, but also a direct proximity effect with \ndrainage of Cooper pairs from S into the ferromagnet. The S -\nFI interface produces only singlet -to-triplet conversion of \nCooper pairs. This is the reason for different behaviour of the \ninduced magnetization in the superconductor in the vicinity of \nthese interfaces. \n \nFigure 5. Magnetization in the S layer near the S -FM interface \nversus the superconductor thickness L for various values of \nthe angle, 𝜑, (a) for the temperature close t o the critical limit, \n(b) for the weak proximity effect limit, here 𝑇=0.8∙𝑇𝑐, 𝑡=\n0.7 for (a), 𝑡=0.1 for (b). \n \n 8 \n Fig. 5 (a) shows that, as one would expect, the \nmagnetization at the S -FM boundary for small L is minuscule, \nbecause superconductivity is suppressed by the magnetization \nof a ferromagnet. The higher the angle, the less magnetization \noccurs. It is clear that growth of the ferromagnetic insulator’s \nmagnetization increases the suppression of superconductivity \nin the S -layer. \nFor the S -FM case w ithin weak proximity effect, see Fig.5 \n(b). In this case, for small thickness of superconductor the \nmagnetization is bigger than for much larger thickness. \nHowever, this might be due to the fact that in this case the \nsuppression of the order parameter in t he superconductor layer \nwas not taken into account, unlike in case (a) . \nFrom Fig. 6(a), we see that there is a maximum of the \naverage magnetization on the spin mixing angle φ due to the \nfollowing factors. For a small magnetization of the \nferromagnet, the growth of the spin -mixing angle leads to the \nappearance of more triplet pairs, and the magnetization in a \nsuperconductor increases. However, at some point, the \nmagnetization of the ferromagnet reaches such a value that its \ngrowth already leads to a decrea se in the number of Cooper \npairs due to the suppression of superconductivity, and, as a \nconsequence, to a decrease in the number of singlet pairs from \nwhich triplet pairs are obtained. In addition, it is seen that with \nincrease in temperature, the magnetiz ation decreases. This is \ndue to the fact that the superconductivity is suppressed with \nincreasing temperature and, as a result, fewer singlet and \ntriplet Cooper pairs are generated and contribute to the \nmagnetization. There is no maximum of the average ind uced \nmagnetization function IM(φ) at Fig.6 (b,c) for the S -FM \nbilayer. This is also a consequence of the transparency of the \ninterface. \n4. Conclusion \nWe have presented results for the induced magnetization \nand superconducting correlations profile in the \nsuperconducting layer of a bilayer superconductor -\nferromagnet structure. We have shown that the induced \nmagnetization monotonically decreases with increasing \ntemperature as a consequence of the inverse proximity effect \nwhich vanishes at the critical temperat ure. The d ependence of \nthe induced magnetization on the coordinate and on the spin \nmixing angle, has a maximum for S -FI, and is monotonous for \nS-FM structures. The reason for this behaviour is a decrease \nin the singlet component and a growth of the triplet component \nwhen approaching the superconductor -ferromagnetic \ninterface. As the magnetization is proportional to the product \nof both, a local maximum in the profile appears. \n \nFigure 6. Average magnetization at (a) the S -FI bilayer, (b) the \nS-FM interface for the temperature close to the critical limit, \n(c) the S -FM interface for the weak proximity effect limit, \nversus the angle of spin mixing, φ, for various values of the \ntemperature, T. Here 𝜉= 93 nm. The transparency 𝑡=0.7 \n(b), 𝑡=0.1 (c). \n \n 9 \n The reason for the local maximum in the averaged \nmagnetization as function of the spin mixing angle is the \ncompetition between singlet -to triplet conversion, provided by \nspin-mixing. A stronger spin mixing leads to suppression of \nsuperconductivity and, the refore, to a decrease of the induced \nmagnetization. \nThe magnitude of the induced magnetization is also \ndifferent for S -FI and S -FM structures. In the S -FI case it is \nusually large r as a consequence of the finite transparency of \nthe S -FM interface for elec trons. \nSuperconductor -ferromagnet structures as discussed in this \npaper may have various applications, e.g. as sensitive \nbolometers and thermometers [40,41] . \nAcknowled gements \nPugach N. and Yagovtsev V. thank Prof. Ekomasov E. G. \nfor valuable discussions and the program Mirror Labs of \nNational Research University Higher School of Economics for \nsupport of interrussian cooperation. \nThe calculations of inverse proximity effect in S -FI bilayers \nand estimations of the spin -mixing angle were funded by the \nRussian Ministry of Education and Science, Megagrant \nproject N 2019 -220-07-6383. The study of S -FM structures \nand comparison with expe riment on inverse proximity effect \nwere funded by RFBR according to the research project № 19 -\n02-00316. \n \nReferences \n1. Žutić, I.; Fabian, J.; Sarma, S. Das. Spintronics: \nFundamentals and Applications. Reviews o f Modern \nPhysics . 2004, pp 323 –410. \nhttps://doi.org/10.1103/RevModPhys.76.323. \n2. Linder, J.; Robinson, J. W. A. Superconducting \nSpintronics. Nat. Phys. 2015 , 11 (4), 307 –315. \nhttps://doi.org/10.1038/nphys3242. \n3. Robinson, M. G.; Blamire, J. W. A. The I nterface \nbetween Superconductivity and Magnetism: \nUnderstanding and Device Prospects. J. Phys. \nCondens. Matter 2014 , 26. \n4. Buzdin, A. I. Proximity Effects in Superconductor -\nFerromagnet Heterostructures. Rev. Mod. Phys. \n2005 , 77 (3), 935 –976. \nhttps://doi. org/10.1103/RevModPhys.77.935. \n5. Eschrig, M. Spin -Polarized Supercurrents for \nSpintronics: A Review of Current Progress. Reports \nProg. Phys. 2015 , 78 (10), 104501. \nhttps://doi.org/10.1088/0034 -4885/78/10/104501. \n6. Bergeret, F. S.; Volkov, A. F.; Efetov , K. B. Odd \nTriplet Superconductivity and Related Phenomena in \nSuperconductor -Ferromagnet Structures. Rev. Mod. \nPhys. 2005 , 77, 1321 –1373. \n7. Heim, D. M.; Pugach, N. G.; Kupriyanov, M. Y.; \nGoldobin, E.; Koelle, D.; Kleiner, R. Ferromagnetic Planar Josephson Junction with Transparent \nInterfaces: A φ Junction Proposal. J. Phys. Condens. \nMatter 2013 , 25 (21), 215701. \nhttps://doi.org/10.1088/0953 -8984/25/21/215701. \n8. Vedyayev, A. V.; Ryzhanova, N. V.; Pugach, N. G. \nCritical Current Oscillations in S/F Heterostructures \nin the Presence of s –d Scattering. J. Magn. Magn. \nMater. 2006 , 305 (1), 53 –56. \nhttps://doi.org/10.1016/j.jmmm.2005.11.026. \n9. Vedyayev, A.; Lacroix, C.; Pugach, N.; Ryzhanova, \nN. Spin -Valve Magnetic Sandwich in a Josephson \nJunction. Europhys. Lett. 2005 , 71 (4), 679 –685. \nhttps://doi.org/10.1209/epl/i2005 -10118 -y. \n10. Pugach, N. G.; Safonchik, M.; Champel, T.; \nZhitomirsky, M. E.; Lähderanta, E.; Eschrig, M.; \nLacroix, C. Superconducting Spin Valves Controlled \nby Spiral Re -Orientation in B20 -Family Magnets. \nAppl. Phys. Lett. 2017 , 111 (16), 162601. \nhttps://doi.org/10.1063/1.5000315. \n11. Klenov, N.; Kornev, V.; Vedyayev, A.; Ryzhano va, \nN.; Pugach, N.; Rumyantseva, T. Examination of \nLogic Operations with Silent Phase Qubit. J. Phys. \nConf. Ser. 2008 , 97, 012037. \nhttps://doi.org/10.1088/1742 -6596/97/1/012037. \n12. Tokuyasu, T.; Sauls, J. A.; Rainer, D. Proximity \nEffect of a Ferromagneti c Insulator in Contact with a \nSuperconductor. Phys. Rev. B 1988 , 38 (13), 8823 –\n8833. https://doi.org/10.1103/PhysRevB.38.8823. \n13. Fazio, R.; Lucheroni, C. Local Density of States in \nSuperconductor -Ferromagnetic Hybrid Systems. \nEurophys. Lett. 1999 , 45 (6), 707 –713. \nhttps://doi.org/10.1209/epl/i1999 -00225 -3. \n14. Bergeret, F. S.; Volkov, A. F.; Efetov, K. B. Induced \nFerromagnetism Due to Superconductivity in \nSuperconductor -Ferromagnet Structures. Phys. Rev. \nB 2004 , 69 (17), 174504. \nhttps://doi.org/10.1103/ PhysRevB.69.174504. \n15. Champel, T.; Eschrig, M. Effect of an \nInhomogeneous Exchange Field on the Proximity \nEffect in Disordered Superconductor -Ferromagnet \nHybrid Structures. Phys. Rev. B - Condens. Matter \nMater. Phys. 2005 , 72 (5), 1 –11. \nhttps://doi.org/10.1103/PhysRevB.72.054523. \n16. Linder, J.; Yokoyama, T.; Sudbø, A. Theory of \nSuperconducting and Magnetic Proximity Effect in \nS/F Structures with Inhomogeneous Magnetization \nTextures and Spin -Active Interfaces. Phys. Rev. B \n2009, 79 (5), 054523. \nhttps://doi.org/10.1103/PhysRevB.79.054523. \n17. Pugach, N. G.; Buzdin, A. I. Magnetic Moment \nManipulation by Triplet Josephson Current. Appl. \nPhys. Lett. 2012 , 101 (24). \nhttps://doi.org/10.1063/1.4769900. \n18. Teichert, A.; Zhernenkov, K. N.; Bottyán, L.; Rühm, \nA.; Steitz, R.; Nikitenko, Y. V.; Khaydukov, Y. N.; \nNagy, B.; Aksenov, V. L. Magnetic Proximity \nEffects in V/Fe Superconductor/Ferromagnet Single \nBilayer Revealed by Waveguide -Enhanced Polarized \n 10 \n Neutron Reflectometry. J. Supercon d. Nov. Magn. \n2010 , 24 (1–2), 961 –968. \nhttps://doi.org/10.1007/s10948 -010-1041 -0. \n19. Salikhov, R. I.; Garif’Yanov, N. N.; Garifullin, I. A.; \nTagirov, L. R.; Westerholt, K.; Zabel, H. Spin \nScreening Effect in Superconductor/Ferromagnet \nThin Film Heterostr uctures Studied Using Nuclear \nMagnetic Resonance. Phys. Rev. B - Condens. \nMatter Mater. Phys. 2009 , 80 (21), 1 –8. \nhttps://doi.org/10.1103/PhysRevB.80.214523. \n20. Kiwi, M. Origin of the Magnetic Proximity Effect. In \nMaterials Research Society Symposium - \nProceedings ; 2002. https://doi.org/10.1557/proc -746-\nq5.2. \n21. Ouassou, J. A.; Pal, A.; Blamire, M.; Eschrig, M.; \nLinder, J. Triplet Cooper Pairs Induced in Diffusive \nS-Wave Superconductors Interfaced with Strongly \nSpin-Polarized Magnetic Insulators or Half -Metallic \nFerromagnets. Sci. Rep. 2017 , 7 (1), 24 –32. \nhttps://doi.org/10.1038/s41598 -017-01330 -1. \n22. Giazotto, F.; Solinas, P.; Braggio, A.; Bergeret, F. S. \nFerromagnetic -Insulator -Based Superconducting \nJunctions as Sensitive Electron Thermometers. Phys. \nRev. Appl. 2015 , 4 (4), 1 –12. \nhttps://doi.org/10.1103/PhysRevApplied.4.044016. \n23. Pal, A.; Blamire, M. G. Large Interfacial Exchange \nFields in a Thick Superconducting Film Coupled to a \nSpin-Filter Tunnel Barrier. Physical Review B - \nCondensed Matter and Materials Physics . 2015. \nhttps://doi.org/10.1103/PhysRevB.92.180510. \n24. Münzenberg, M.; Eich, M.; Assaf, B. A.; Moodera, J. \nS.; Li, B.; Epstein -Martin, M.; Roschewsky, N.; \nHeiman, D. Superconducting Spin Switch with \nInfinite Magnetoresistance Induced by an Internal \nExchange Field. Phys. Rev. Lett. 2013 , 110 (9), 1 –5. \nhttps://doi.org/10.1103/physrevlett.110.097001. \n25. Wolf, M. J.; Sürgers, C.; Fischer, G.; Beckmann, D. \nSpin-Polarized Quasiparticle Transport in Exchange -\nSplit Superconducting Aluminum on Europium \nSulfide. Phys. Rev. B - Condens. Matter Mater. Phys. \n2014 , 90 (14), 1 –5. \nhttps://doi.org/10.1103/PhysRevB.90.1 44509. \n26. Hao, X.; Moodera, J. S.; Meservey, R. Thin -Film \nSuperconductor in an Exchange Field. Phys. Rev. \nLett. 1991 , 67 (10), 1342 –1345. \nhttps://doi.org/10.1103/PhysRevLett.67.1342. \n27. Kolenda, S.; Sürgers, C.; Fischer, G.; Beckmann, D. \nThermoelectric Effects in Superconductor -\nFerromagnet Tunnel Junctions on Europium Sulfide. \nPhys. Rev. B 2017 , 95 (22), 1 –5. \nhttps://doi.org/10.1103/PhysRevB.95.224505. \n28. Bergeret, F. S.; Verso, A.; Volkov, A. F. Electronic \nTransport through Ferromagnetic and \nSupercon ducting Junctions with Spin -Filter \nTunneling Barriers. Phys. Rev. B - Condens. Matter \nMater. Phys. 2012 . \nhttps://doi.org/10.1103/PhysRevB.86.214516. \n29. Bergeret, F. S.; Verso, A.; Volkov, A. F. Spin -Polarized Josephson and Quasiparticle Currents in \nSuper conducting Spin -Filter Tunnel Junctions. Phys. \nRev. B - Condens. Matter Mater. Phys. 2012 . \nhttps://doi.org/10.1103/PhysRevB.86.060506. \n30. Eschrig, M.; Cottet, A.; Belzig, W.; Linder, J. \nGeneral Boundary Conditions for Quasiclassical \nTheory of Superconduc tivity in the Diffusive Limit: \nApplication to Strongly Spin -Polarized Systems. \nNew J. Phys. 2015 , 17 (8). \nhttps://doi.org/10.1088/1367 -2630/17/8/083037. \n31. Linder, J.; Sudbø, A.; Yokoyama, T.; Grein, R.; \nEschrig, M. Signature of Odd -Frequency Pairing \nCorrelations Induced by a Magnetic Interface. Phys. \nRev. B - Condens. Matter Mater. Phys. 2010 , 81 \n(21), 1 –13. \nhttps://doi.org/10.1103/PhysRevB.81.214504. \n32. Champel, T.; Eschrig, M. Switching \nSuperconductivity in Superconductor/Ferromagnet \nBilayers by Mult iple-Domain Structures. Phys. Rev. \nB - Condens. Matter Mater. Phys. 2005 , 71 (22), \n220506. \nhttps://doi.org/10.1103/PhysRevB.71.220506. \n33. Eschrig, M.; Löfwander, T. Triplet Supercurrents in \nClean and Disordered Half -Metallic Ferromagnets. \nNat. Phys. 2008 , 4 (2), 138 –143. \nhttps://doi.org/10.1038/nphys831. \n34. Eschrig, M. Theory of Andreev Bound States in S -F-\nS Junctions and S -F Proximity Devices. Philos. \nTrans. R. Soc. A Math. Phys. Eng. Sci. 2018 , 376 \n(2125), 20150149. \nhttps://doi.org/10.1098/rsta.2015.0149. \n35. Mironov, S.; Mel’nikov, A. S.; Buzdin, A. \nElectromagnetic Proximity Effect in Planar \nSuperconductor -Ferromagnet Structures. Appl. Phys. \nLett. 2018 , 113 (2). \nhttps://doi.org/10.1063/1.5037074. \n36. Eschrig, M. Scattering Problem in Nonequilibrium \nQuasiclassical Theory of Metals and \nSuperconductors: General Boundary Conditions and \nApplications. Phys. Rev. B - Condens. Matter Mater. \nPhys. 2009 . \nhttps://doi.org/10.1103/PhysRevB.80.134511. \n37. Vodo pyanov, 1B. P.; Tagirov, L. R. Andreev \nConductance of a Ferromagnet -Superconductor Point \nContact. J. Exp. Theor. Phys. Lett. 2003 , 77 (3), 126 –\n131. https://doi.org/10.1134/1.1567773. \n38. Vodopyanov, B. P.; Tagirov, L. R. Oscillations of \nSuperconducting Tr ansition Temperature in Strong \nFerromagnet -Superconductor Bilayers. J. Exp. \nTheor. Phys. Lett. 2003 , 78 (9), 555 –559. \nhttps://doi.org/10.1134/1.1641483. \n39. Xia, J.; Shelukhin, V.; Karpovski, M.; Kapitulnik, \nA.; Palevski, A. Inverse Proximity Effect in \nSuperconductor -Ferromagnet Bilayer Structures. \nPhys. Rev. Lett. 2009 , 102 (8), 087004. \nhttps://doi.org/10.1103/PhysRevLett.102.087004. \n40. Heikkilä, T. T.; Ojajärvi, R.; Maasilta, I. J.; \nStrambini, E.; Giazotto, F.; Bergeret, F. S. \n 11 \n Thermoelectric Radiation Detector Based on \nSuperconductor -Ferromagnet Systems. Phys. Rev. \nAppl. 2018 , 10 (3), 1 –10. \nhttps://doi.org/10.1103/PhysRevApplied.10.034053. \n41. Giazotto, F.; Robinson, J. W. A.; Moodera, J. S.; \nBergeret, F. S. Proposal for a Phase -Coherent \nThermoelectric Transistor. Appl. Phys. Lett. 2014 , \n105 (6), 1 –5. https://doi.org/10.1063/1.4893443. \n " }, { "title": "2206.08196v1.Tunable_Ferromagnetism_in_LaCoO3_Epitaxial_Thin_Films.pdf", "content": " \n1 \n \nTunable Ferromagnetism in LaCoO 3 Epitaxial Thin Films \n \nDongwon Shin#, Sangmoon Yoon#, Sehwan Song, Sungkyun Park, Ho Nyung Lee , and Woo \nSeok Choi* \n \nD. Shin, W. S. Choi \nDepartment of Physics, Sungkyunkwan University, Suwon 16419, Republic of Korea \nE-mail: choiws@skku.edu \n \nS. Yoon, H. N. Lee \nMaterials Science and Technology Division, Oak Ridge National Laboratory, Oak Ridge, \nTennessee 37831, United States \n \nS. Yoon \nDepartment of Physics, Gachon University, Seongnam, 13120, Republic of Korea \n \nS. Song, S. Park \nDepartment of Physics, Busan National University, Busan 46241, Republic of Korea \n \nKeywords: ferromagnetic insulator, p erovskite LaCoO 3, crystallographic orientation , \nferromagnetic ordering, pulsed laser epitaxy \n \nFerroma gnetic insulators play a crucial role in the develop ment of low-dissipation quantum \nmagnetic device s for spintronics. Epitaxial LaCoO 3 thin film is a prominent ferromagnetic \ninsulator , in which the robust ferromagnet ic ordering emerges owing to epitaxial strain. \nWhe reas it is evident that strong spin-lattice coupling induces ferromagnetism, the reported \nferromagnetic properties of epitaxially strained LaCoO 3 thin films were highly consistent. For \nexample, even under largely modulated degree of st rain, the reported Curie temperatures of \nepitaxially strained LaCoO 3 thin films lie within 80 –85 K , without much deviation . In this \nstudy , substantial enhancement (~18%) in the Curie temperature of epitaxial LaCoO 3 thin \nfilms is demonstrated via crystallog raphic orientation dependence . By changing the \ncrystallographic orientation of the films from (111) to (110), the crystal -field energy was \nreduced and the charge transfer between the Co and O orbitals was enhanced. These \nmodifications led to a considerable enhancement of the ferromagnetic properties (including \n2 \n \nthe Curie temperature and magnetization), despite the identical nominal degree of epitaxial \nstrain . The findings of this study provide insights into facile tunability of ferromagnetic \nproperties via structural symmetry control in LaCoO 3. \n \n1. Introduction \nFerromagnetic insulator s (FM-Is) are essential component s in dissipation -less spintronic \ndevices with highly suppressed leakage currents.[1, 2] They generat e pure spin currents by \nfilter ing charges and in troducing spins into adjacent nonmagnetic layer s via the magnetic \nproximity effect.[3-8] However , FM -Is are scarce in nature compared to more conventional FM \nmetals because the general superexchange mechanism in insulators promotes \nantiferromagnetic (AFM) ordering.[9-12] Iron yttrium garnet (YIG), α-Fe2O3, spinel CoFe 2O4, \nEuS, EuO, and strained EuTiO 3 thin films are some of examples of FM -I materials.[13-19] \nEpitaxially -strained LaCoO 3 (LCO) , e.g., epitaxially deposited on SrTiO 3 (STO) , \n(LaAlO 3)0.3(Sr2AlTaO 8)0.7 (LSAT), or YAlO 3 substrates, is an emerging FM-I with a Curie \ntemperature above the liquid N 2 temperature (TC = ~80 K), with a robust ferromagnetic \norigin.[20, 21] \n \nTable 1. TC obtained in LCO thin films: A summary of reported TC of LCO thin films, \ndepending on the applied magnetic field ( H), type of substrates, crystallographic orientation , \nand film thickness. Note that a robust TC with few exceptions is found for the LCO thin films. \nIn contrast a substantial increase of TC (~ 18%) is observed in the current study. \nTC (K) H [Oe] substrate orientation thickness [nm] reference \n43.2 200 LAO (100) 100 \nPhys. Rev. B 77, 014434 (2008) 69.2 200 SLAO (100) 100 \n71.6 200 SLGO (100) 100 \n76.0 200 LSAT (100) 100 \n77.6 200 STO (100) 100 \n80/87 200 LSAT (100) 8/50 Phys. Rev. B 79, 024424 (2009) \n86 2000 STO (100) 100 \nEur. Phys. J. B 76, 215 –219 (2010) 75 2000 LAO (100) 100 \n85 2000 LSAT (100) 100 \n87 2000 PMN -PT (100) 100 \n84 2000 SLAO (100) 100 \n80 2000 STO/LSAT (100) 20-90 J. Appl. Phys. 105, 07E503 (2009) \n76-83 50 LAO (100) 30-95 J. Appl. Phys. 109, 07D717 (2011) \n75.9/81.2 50 STO/LSAT (100) 15 Phys. Rev. B 91, 144418 (2015) \n80 1000 STO/LSAT (100) 30 Nano Lett. 12, 4966 -4970 (2012) \n94 1000 LSAT (110) 60 Phys. Rev. Lett. 111, 027206 (2013) \n24 1000 LSAT (111) 60 \nPhys. Rev. B 92, 195115 (2015) \n92/90 500 STO/ LSAT (111)/ (100) 35/38 \n3 \n \n85 500 STO/LSAT (100) 30 AIP Advances 8, 056317 (2018) \n85 500 STO (100) 12 PNAS 115, 2873 –2877 (2018) \n78 500 STO (100) 70 Curr. Appl. Phys. 28, 87-92 (2021) \n50 500 STO (111) 80 Phys. Status Solidi B , 2100424 (2021) \n76.1-90.2 1000 STO (100), (110) and (111) 30 This work \n \nLong-range FM ordering in epitaxial LCO thin film s is facilitated by strain -induced lattice \ndistortion s. Although bulk LCO is nonmagneti c with a rhombohedral crystal structure, the \ntensile strain in the thin film is known to induce monoclinic and tetragonal phases, resulting in \na spin-ordered state.[20, 22, 23] The large and facile modification of the crystalline symmetry \noriginates from the ferroelastic nature of LCO , which favors the formation of twin-domain \nstructure s under mechanical pressure or epitaxial strain .[24-31] Such twin -domain structures \nhave been observed as dark stripe patterns in high -angle annular dark -field scanning \ntransmission electron microscopy (HAADF STEM) imaging.[20, 23, 32, 33] Detailed \nmicrostructure analyses using HAADF STEM on (100) -oriented LCO thin films enable \nclassif ication of basic lattice units into compressed - (c-unit), tensile -stretched - (t-unit) , and \nbulk-units (b-unit).[23] Among the se, c-units with coexisting low (LS) and high spin- (HS) \nstates of the Co spin configuration were found to be responsible for the FM phase. It should \nbe noted that the FM phase , characterized by TC, in the (100) -oriented LCO thin film was \nmostly unchang ed. As summarized in Table 1 , the reported values of TC of (100) -oriented \nLCO thin films are mostly in the range 80 –85 K,[20, 34-45] despite the large difference in the \ndegree of epitaxial strain. Note that TC might decrease further below 80 K, if the crystalline \nquality is not ideal .[43] Such unsusceptible behavior in TC suggests that the magnetically active \nc-units are highly resilient against orthogonal stress, likely stemming from the ferroelastic \nnature of LCO.[23] \n \nModifications in the crystallographic orientation lead to disparate lattice symmetry, offering \nextra tunability in the functional properties of epitaxial thin films. Even though an identical \nfilm/substrate system is used, implying the same degree of nominal epitaxial strain , the \nstructural symmetry of the thi n film can be modulated by the crystallographic orientation of \nthe substrate , leading to stress that is not orthogonal to the lattice . For instance, the FM TC of \nepitaxial SrRuO 3 thin film s can be increased from 147 to 157 K, a s the crystallographic \norient ation changes from (100) to (111) , because of the decreas ed spin dimensionality within \nthe trigonal symmetry .[46] In epitaxial La0.7Sr0.3MnO 3 thin films, (110) and (111) -oriented thin \nfilms exhibit larger spin moment s than (100)-oriented thin film s.[47] As the crystallographic \norientation of LaAlO 3 thin films on KTaO3 substrate s changed from (100) to (110) or (111), \n4 \n \nemergent superconductivity w as observed . The superconducting critical temperatures could \nbe further modulated between the (110) and (111) -oriented thin films , with the values of 1 and \n2 K, respectively.[48-50] Because the crystallographic orientation determines the \ncrystallographic symmetry of the lattice unit, it is expected to modify the robust FM -I phase \nof LCO epitaxial thin films as well. Th is strategy would provide fundamental insight into the \nstructure -property relationship of LCO, which would not have been possible using only (100) -\noriented thin films with different strain state s, but the same type of orthogonal stress. Indeed, \nstudies on LCO thin films with crystallographic orientations other than (100) -orientation are \nhighly limit ed and report inconsistent TC values , as shown in Table 1, and a systematic \ncomparison between (100) -, (110) -, and (111) -oriented LCO epitaxial thin films has not been \nconducted thus far. \n \nIn this study, we demonstrate a large modulation of the FM phase in LCO epitaxial thin films, \nincluding the substantial tuning of TC, by tailoring the crystalline symmetry via \ncrystallographic orientation control. We fabricated epitaxial LCO thin films on (100)-, (110) - \nand (111) -oriented STO substrates using pulsed laser epitaxy (PLE) . The structural quality \nand crystallographic symmetry of the thin films have been determined by X -ray diffraction \n(XRD) measurements and HAADF STEM imaging. The FM phases for the (100) -, (110) - and \n(111) -oriented LCO thin films have TCs of 81.5, 90.2, 76.1 K, respectively . The large \nmodulation of the FM TC, i.e., ~18% enhancement in the (110) -oriented thin film compared to \nthe (111) -oriented one, captures the effectiveness of the crystallographic symmetry control on \nthe lattice units. The physical origin of the TC enhancement was investigated by characteriz ing \nthe microscopic lattice unit s, electronic structure measurements , and theoretical calculations . \nIn particular, the reduced crystal -field splitting energy and stronger Co -O charge transfer in \nthe (110) -oriented thin film resulted in considerable enhance ment of the FM properties. \n \n2. Results and Discussion \nThe global lattice structures of the LCO thin films (30 nm) were characterized using XRD to \nreveal the systematic crystallographic orientation dependence. Figure 1 a shows the XRD \nθ−2θ scans for the LCO thin films . Fine thickness fringes are commonly observed along with \nthe corresponding LCO peaks (Figure 1a; Figure S1 a and b , Supporting Information) , \nindicating consistent crystalline phases with atomically smooth surfaces and interface s. In \naddition, atomic force microscopy images show atomically smooth surfaces of the epitaxial \nLCO thin films with low surface roughness of < 1 u.c. ( Figure S2, Supporting Information ). \n5 \n \nFigure 1b shows the XRD reciprocal space maps around the (103), (120), and (212) Bragg \nreflections for the LCO thin films epitaxially grown on the (100)-, (110) -, and (111) -oriented \nSTO substrates, respectively. Despite the difference in the crystallographic orien tations, all \nfilms were fully strained to the substrates. Although the nominal lattice mismatch between the \nthin film and the substrate is identical , it is important to note that the degree of strain still \ndiffers owing to symmetry modification . The differ ence in the degree of strain is shown in \nFigures 1c , in terms of the pseudocubic unit cell volume ( Vu.c.) and out -of-plane stress ( εoop \n(%) = 100 × (dfilm − dbulk) / dbulk, where dfilm and dbulk are the out -of-plane distance s between \nthe atomic planes of the thin film and the bulk, respectively) for distinctive crystallographic \norientation s. The s ame trends are shown for the 10- and 20 -nm-thick LCO thin films (Figure \nS1c and d, Supporting Information) . Both Vu.c. and εoop were the largest for the (110) -oriented \nthin film, indicating the largest lattice distortion, and the smallest for the (111) -oriented film. \n \nFigure 1. Crystal structures of epitaxial LCO thin films with various crystallographic \norientations: a) XRD θ-2θ scans of epitaxial LCO thin films (#) grown on STO substrates \nwith different orientations . b) XRD reciprocal space maps (RSM) of the (100) -, (110) -, and \n(111) -oriented LCO thin films, shown around the (103), (120), and (212) Bragg reflections of \nthe STO substrates, respectively. Crystallographic orientation -dependent c) unit cell volume \n(Vu.c.) and out-of-plane strain ( εoop). Z-contrast HAADF STEM images for LCO t hin films \ngrown on d) (100), e) (110), and f) (111) STO substrates. Dark stripe patterns are observed in \nall crystallographic orientations of the LCO thin films. Two distinct regions are marked with \nwhite horizontal lines. Region I near the interface is a “uniform region” without any dark \n0.20 0.210.740.760.78(212)\nLCO\nSTO\n0.17 0.18 0.190.540.56\nqx (Å-1)(120)\nLCO\nSTO58.058.1(c)\nVu.c. (Å3)\n(100) (110) (111)-0.5-0.4-0.3\neoop (%)\nOrientation0.24 0.260.760.780.80\nLCO(b)\nqz (Å-1)(103)\nSTO\n10 20 30 40 50 60 70 80##\n#(300)(220)\n(100)\n#\n#(200)\n(111)(a)\nIntensity (arb. units)\n2q (degrees)#(110)\n(e) (d) (f)\nLCO (100)\nSTO (100)III\n2 nm[100][001]\n[010]\nSTO (111)[01][111]\n[11]\n2 nmLCO (111)\nSTO (110)[00][110]\n[10]LCO (110)\n2 nm \n6 \n \nstripe patterns. Region II away from the interface is an “ordered region” with dark stripe \npatterns. \n \nFigure 1d–f show s high-resolution HAADF STEM images of epitaxial LCO thin films grown \non d) (100) -, e) (110) -, and f) (111) -oriented STO substrates, respectively . The STEM images \ncommonly exhibit a coherent alignment of the atomic columns without any defects and/or \nmisfit dislocations at the interfaces. Interestingly, distinctive dark stripe patterns (with ~3 u.c. \nperiodicity) were observed , above a certain thickness of the interfacial layer. We carefully \nexclude the presence of oxygen vacancy ordering at the dark stripe pa tterns by electron \nenergy loss spectroscopy (EELS) measurement s (Figure S3, Supporting Information). While \nthe dark stripe patterns have been frequently recognized for the (100) -oriented LCO thin \nfilms,[20, 23, 32, 33] we report similar patterns in (110) - and (111) -oriented thin films with \nsystematic changes for the first time. The LCO thin films can be categorized into two regions \nconsidering the dark stripe pattern , i.e., Region I (near the interface) and Region II (above the \ninterface) .[20] Region I is free of the dark strip e patterns, suggesting a conventional fully \nstrained state . The thickness of Region I is the thickest for the (110) -oriented thin film, and \nthe thinnest for the (111) -oriented thin film , consistent with the fact that the (110) -oriented \nthin film has the largest deviation in Vu.c. from the bulk value (55.604 Å3). On the other hand, \nRegion II presents twin -domain structures due to the ferroelastic nature of the LCO thin film, \nas mentioned earlier. For more detailed structural analyses, we measured the distance of La \ncolumns in Region II along the two orthogonal axes perpendicular to the HAADF STEM \nimage (Table S1, Supporting Information). The lattice structure of magnetically active c-units , \nwhich was robust in the (100) -oriented LCO thin films,[21] is distorted differently depending \non the substrate orientations, suggesting that the robust structure can be tuned by symmetry \nconstraints. The volume of c-unit is the largest in the (110) -orien ted LCO thin film, while it is \nthe smallest on the (111) -oriented film, which is consistent with the XRD results. \n \nThe FM ordering in the LCO thin films is strongly coupled to the crystallographic symmetry . \nFigure 2 a show s the temperature -dependent magnetization, M(T), of the LCO thin films \nunder field -cooled (FC , 1000 Oe ) cooling along the in -plane direction. A clear FM transition \nwith a sudden upturn of M(T) is observed at 81.5, 90.2, and 76. 1 K for the (100)-, (110) -, and \n(111)-oriented LCO thin films, respectively. The inset of Fig ure 2a shows the linear \nextrapolation of M(T) for the TC determination . The m agnetic -field-dependent magnetization, \nM(H), measured at 2 K , demonstrates a clear FM hysteresis loop, as shown in Figure 2b. The \n7 \n \nvalues of saturation magnetization ( Ms) and remanent magnetization ( Mr) are summarized in \nFigure 2c and 2d, respectively, exhibiting consistent orientation -dependent trend s with TC \n(Figure 2e) . Indeed, a substant ial enhancement in ferromagnetic properties, i.e., TC (~18%), \nMs (~36%), and Mr (~49%), is observed, by comparing the (110) -oriented thin film to the \n(111) -oriented one. The same trend was observed from the Curie -Weiss (C -W) law fitting of \nthe M(T) curves, in terms of the effective ratio of the HS-state (Figure S4, Supporting \nInforma tion).[29] Within the prevailing spin configuration model of LCO epitaxial thin \nfilms,[23, 29, 33, 51] in which the LS - and HS -states coexist, the effective ratios of HS -state are \n68.4, 69.5, and 57.2%, respectively, for the (100) -, (110) -, and (111) -orien ted LCO thin films. \nX-ray photoelectron spectroscopy results consistently support that the (110) -oriented LCO \nthin film exhibits the largest HS state ratio (Figure S5, Supporting Information) . \n \nFigure 2. Crystallographic orientation dependent magnetic properties : a) M(T) curves of \nLCO thin films on (100), (110), and (111) STO substrates along the in -plane direction. M(T) \ncurves were obtained during field-cooled cooling at 100 0 Oe of magnetic field. The i nset of \nFigure 2a show s TC’s defined by linear extrapolation . b) M(H) curves of LC O thin films on \n(100), (110), and (111) STO substrates along the in -plane direction at 2 K. M(H) curves of \nLCO thin films show FM hysteresis loops. Crystallographic orientation -dependent FM \nproperties i ncluding c) saturation magnetization ( Ms), d) remanent magnetization ( Mr), and e) \nCurie temperature ( TC) are summarized for the LCO thin films. \n \nThe electronic structures of LCO thin films provide a microscopic mechanism f or FM \nmodulation with distinctive crystallographic symmetries. In particular , we performed \nspectroscopic ellipsometry to capture minute changes in the electronic band structures of the \nLCO thin films . The real part of the optical conductivity , σ1(ω), of the (100) -, (110) -, and \n(111) -oriented LCO thin film s is shown in Figure 3a, which close ly resembl es σ1(ω) of bulk \n50 100 150 200 250 3000.00.10.20.30.4(a)m0H // ab @ 1000 Oe\nFCM (mB Co-1)\nT (K) (100)\n (110)\n (111)\n-4 -2 0 2 4-1.0-0.50.00.51.0(b)\nm0H // ab \n@ 2 KM (mB Co-1)\nm0H (T) (100)\n (110)\n (111) 0.60.8(c)Ms (mB Co-1)\n0.30.4(d)Mr (mB Co-1)\n(100) (110) (111)75808590 (e)TC (K)\nOrientation40 60 80 1000.00.10.20.3\n76.1 KM (mB/Co)\nT (K)90.2 K81.5 K \n8 \n \nLCO . [20, 52] Below 4 eV , σ1(ω) consists of two Lorentz oscillators centered at ~1.5, and ~ 3.0 \neV, labelled as the β (d-d transition within Co 3 d orbitals, brown dotted lines) and γ peaks \n(charge transfer transition, O 2p → Co 3 d, green dotted lines ), respectively, as shown in \nFigure 3b.[52] The quantitative values of the peak positions ( ωj) and spectral weight s (SWj) of \nthe β and γ peaks are summarized in Fig ure 3c and 3d. The (110) -oriented LCO thin film \nexhibit ed a lower ωj for both the β and γ transition s compared to those of the (100 )- and \n(111) -oriented thin films. In addition , the (110) -oriented thin film show ed a larger SW j value \nfor both the β and γ transitions compared to those of the (100) - and (111) -oriented thin films. \n \nFigure 3. Optical properties and electronic structures : a) σ1(ω) of LCO thin films on \ndifferently oriented STO substrates. For comparison, we displayed the reported σ1(ω) of LCO \nthin film (black dotted line) and bulk (grey dotted line). b) σ1(ω) can be deconvoluted by the \nLorentzian oscillators corresponding to β (Co d-d transition) and γ (O 2p to Co 3d charge \ntransfer transition) peaks. Each c) ωj and d) SWj of the individual Lorentzian oscillators is \nshown as a function of the crystallographic orientation of the LCO thin films. \nLCO thin film: Reproduced with permission.[20] Copyright 2012 , ACS Publications . Bulk \nLCO: Reproduced with permission.[52] Copyright 2014, Nature portfolio . \n \nThe strongly intertwined electronic structure and FM properties of the LCO epitaxial thin \nfilms m odulated by crystallographic orientation are illustrated in Figure 4 . When we consider \nan HS configuration within LCO, which is necessary for FM ordering, the net spin moment \ncan be evaluated from the energy states of the electron orbitals within the atomic picture. In \nparticular, th is tendency can be qualitatively captured by taking into account the energy cost \n(Δ) of the spin-state transition from a n LS to HS configuration . A larger spin moment, and \nhence stronger FM behavior , is expected for a smaller Δ. Δ can be defined as, Δ = ΔCF – Δex – \nW/2, where ΔCF and Δex are the crystal -field splitting and Hund energ ies, respectively, and W \n01000(b)\n g (O 2 p ® Co 3 d)\nb (Co d-d) (100)\n01000 (110)\ns1(w) (W-1 cm-1)\n0 1 2 3 401000 (111)\nw (eV)0 2 4 6010002000\ns1(w) (W-1 cm-1)\nw (eV) (100)\n (110)\n (111)\n LaCoO3 thin film\n Bulk LaCoO3(a)\n1.41.52.62.7(c)\nwj (eV) g\n b\n(100) (110) (111)451516(d)\nSWj (eV2)\nOrientation \n9 \n \nis the bandwidth.[29] Thus, the competition between ΔCF, Δex, and W determines Δ. Both the \nbond length (rCo-O, average distance between Co and O ions) and bond angle ( θ, Co-O-Co \nangle ) influence Δ and the energy scales,[51, 53, 54] in which ΔCF ~ rCo-O–5 and W ~ rCo-O–3.5cos(π \n− θ).[29, 55] Apparently, ΔCF has a stronger dependence on rCo-O, and hence, is a dominant factor \nfor determin ing Δ. For the CoO 6 octahedra, ΔCF is determined by the energy difference \nbetween the Co t2g and eg states , as depicted in Figure 4a. Because ΔCF is directly related to ωβ \nand partially related to ωγ, the observation of the lowest ωβ and ωγ in the (110) -oriented LCO \nthin film (Figure 3c) indicates that the film has the lowest ΔCF, leading to a stronger FM spin \nordering (small Δ, Figure 2).[55, 56 ] We can directly infer this insight from the lattice structure, \nas the (110) -oriented LCO thin film has the largest Vu.c. among the films studied (Figures 1c \nand 1d; Table S1, Supporting Information). We further note that the charge transfer transition \nfrom the O 2 p to Co 3 d orbital s, depicted by SWγ, is the largest for the (110) -oriented LCO \nthin film. We expect that (110) -oriented LCO thin film to have θ value close r to 180° than the \nother orientations, as it has the largest Vu.c., that would lead to a stronger charge transfer \nbetween Co and O orbitals.[34, 35] It has been reported that a stronger charge transfer transition \nleads to enhanced magnetic ordering in LCO thin films.[34, 57] In general, the facile spin and \ncharge transfer between Co 3 d and O 2 p orbitals introduces a strong FM superexchange \ninteraction.[34, 58] \n \n024(100)pDOS (state/eV f.u.) HS Co3+ t2g LS Co3+ t2g\n HS Co3+ eg LS Co3+ eg\n O 2p0246\nwg(110)\nwb\n-3 -2 -1 0 1 2 3024(111)\nEnergy (eV)-0.5 2.0 2.50.02.55.0total DOS\n(state/eV f.u.)\nEnergy (eV) (100)\n (110)\n (111)LCO ( 100)LCO ( 110)\nLCO ( 111)(b) ΔCF ~ rCo-O-5\nVu.c.= 58.10 Å3\nεoop= −0.26 %\nVu.c.= 58.03 Å3\nεoop= −0.38 %\nVu.c.= 57.96 Å3\nεoop= −0.49 %(a)\nΔCF\nΔCFeg\nt2g\neg\nt2g\negσ\negπ\na1gΔCF\n[010][001]\n[100]\n[00][110]\n[10]\n[01][111]\n[11]\nCo O\n \n10 \n \nFigure 4. Origin of the FM modulation in the LCO thin films: a) Schematic diagram of \nthe crystallographic orientation -dependent spin -state configuration. Crystallographic \nsymmetry modulates the orbital energy level for the Co3+ state to determine ΔCF. b) The pDOS \nand total DOS (inset of bottom panel) of LCO thin films under different symmetry obtained \nby DFT calculation . Optical transitions β (brown arrow) and γ (green arrow) indicate the Co \nd-d transition and Co -O charge transfer transition, respectively . \n \nThe total (total DOS) and partial density of states (pDOS) calculated by ab initio density \nfunctional th eory (DFT) calculations consistently support the modulation in ΔCF and charge \ntransfer between the Co and O orbitals , as shown in Figure 4b. We used the experimental \nlattice parameters obtained from the HAADF STEM analys es (Table S1 , Supporting \nInformation ) for the calculation s. Previously, it was found that the rock salt-type HS/LS -FM \norder within the c-unit is responsible for the magnetic properties of epitaxially strained \nLCO.[23, 54] Therefore, w e used the rock salt-type HS/LS -FM order, which is the most stable \nstate regardless the crystallographic symmetry (Table S2, Supporting Information). The \nmodulated optical transitions β (brown arrow) and γ (green arrow) from the DFT calculation s \nare systematically reproduced , as shown in the inset of the bottom most panel of Figure 4b. As \nthe crystallographic orientation changes from (111) to (100) to (110), the separation between \nthe Co -O hybridized states and the Co 3 d states increases on the order of 0.1 eV, which is \nconsistent with the results of σ1(ω) shown in Fig ure 3. \n \nRecently, diffusion Monte Carlo calculations show ed that the magnetic ground state of the \nLCO thin film highly depends on the La -La distance , with the HS/LS -FM order becoming the \nmost stable state below a specific compressive strain.[59] In terms of the exchange mechanism, \nthis order mainly results from correlation superexchange , which refer s to the magnetic \ninteraction whereby two electrons of O2− mediate the exchange coupling between neighboring \nHS and LS Co3+. More specific ally, the double occurrence of AFM superexchange coupling \nalong the HS Co3+ dx2−y2 – O2− pσ – LS Co3+ dx2−y2 – O2− pσ – HS Co3+ dx2−y2 pathway \nfacilitate s the FM ordering ( Figure S6, Supporting Information). In this mechanism, the net \nspin moment of LS Co3+ dx2−y2 (which would be empty in the ideal LS Co3+ configuration) \nmay be a descriptor for the superexchange coupling strength because it reflect s the spin -\npolarized charge transfer from neighboring O2− pσ orbitals. \n \n11 \n \nTo quantitatively assess the spin -polarized charge transfer, we analyzed the eigen -occupations \nof the d orbital density matrix of the HS and LS Co3+ stabilized in the c-units of (100) -, \n(110) -, and (111) -oriented LCO thin films (Table 2). The eigen -occupations are fictitiou s \nauxiliary physical quantities, which often provide important insight s into the orbital states of \nthe transition metal ions.[60-62] The HS Co3+ was confirmed to have an HS d6 orbital \nconfiguration with all majority -spin orbitals and the minority -spin dxy orbital fully occupied , \nbut the other minority -spin orbitals almost empty. The result for LS Co3+ shows that the three \nlower -energy d orbitals, i.e., dxy, dyz, and dzx orbital s, are almost fully occupied, which is \nconsistent with LS Co3+. In contrast , the two higher -energy d orbitals ( the eg orbitals), i.e., \ndx2−y2 and d3z2−r2 orbitals , are expected to be empty in the ideal LS Co3+ orbital configuration, \nbut as mentioned above, they are partially occupied because of the spin -polarized charge \ntransfer from O2−. Interestingly, the net spin moment of LS Co3+ dx2−y2, estimated by the \ndifference between the eigen -occupations of its majority - and minority -spins, was the largest \nin the LCO with the (110) orient ation (Table 2). This DFT calculation result directly implies \nthat the spin -polarized charge transfer, which determines the superexchange coupling \nstrength , occurs more actively in the (110) -oriented LCO thin films than in the (100) - or \n(111) - oriented films. In addition, local magnetic moments of HS Co3+ within HS/LS -FM \norder are summarized in Table S3. Upon comparison, the local magnetic moment was the \nlargest in the c-unit stabilized the (110) -orientation, followed by that in the (100) - and (111) -\norientations, which also matches with experimental result s. \n \nTable 2. Orbital occupations of the HS and LS states for LCO thin films: A summary for \noccupations of t2g (dxy, dyz and dzx) and eg orbital ( dx2−y2 and d3z2−r2) for HS and LS states. \n HS LS \n majority -spin minority -spin majority – \nminority spin majority -spin minority -spin majority – \nminority spin \n(100) eg 3z2-r2 1.0211 0.2333 0.7878 0.4811 0.3258 0.1553 \nx2-y2 1.0635 0.4318 0.6317 0.4886 0.3359 0.1527 \nt2g xy 0.9726 0.9721 0.0005 0.9737 0.9741 -0.0004 \nyz 0.9943 0.2213 0.7730 0.9801 0.9459 0.0342 \nzx 0.9932 0.2175 0.7757 0.9797 0.9437 0.0360 \n(110) eg 3z2-r2 1.0215 0.2304 0.7911 0.4872 0.3322 0.1550 \nx2-y2 1.0601 0.4279 0.6322 0.4806 0.3244 0.1562 \nt2g xy 0.9733 0.9730 0.0003 0.9745 0.9747 -0.0002 \nyz 0.9932 0.2154 0.7778 0.9791 0.9466 0.0325 \nzx 0.9920 0.2167 0.7753 0.9776 0.9435 0.0341 \n(111) eg 3z2-r2 1.0250 0.2318 0.7932 0.4957 0.3376 0.1581 \nx2-y2 1.0725 0.4331 0.6394 0.4972 0.3502 0.1470 \nt2g xy 0.9732 0.9727 0.0005 0.9750 0.9751 -0.0001 \n12 \n \nyz 0.9957 0.2333 0.7624 0.9814 0.9417 0.0397 \nzx 0.9957 0.2332 0.7625 0.9814 0.9417 0.0397 \n \n3. Conclusion \nWe studied the unexpectedly large modulation of ferromagnetic properties , including a \nsizable increase in TC, in crystallographic orientation dependent -LCO thin films. By \nsystematically controlling the crystallographic symmetry, we obtained LCO thin films with \nvarious structural parameters. The c lear FM phases strongly coupled to the crystallographic \nsymmetry demonstrate d that the (110) -oriented LCO thin film exhibit ed highly enhanced FM \nproperties. Optical spectroscopy and DFT calculation s revealed that LCO thin films with \nsmall ΔCF and large spin -polarized charge transfer between the Co 3 d and O 2 p orbitals \nfacilitate FM ordering in the (110) -oriented LCO thin film. T his study implies that the \nsymmetry constraint is an efficient tuning parameter for the exchange coupling strength in \nferromagnetic insulator LCO thin films , in which ferromagnetism is highly robust against \nconventiona l strain due to its ferroelastic nature. \n \n4. Experimental Section \nEpitaxial thin film growth : High -quality LCO epitaxial thin films were fabricated using PLE on \nconventionally HF -treated STO substrates with various orientation s [(100), (110), and (111) ]. \nAn excimer KrF laser (λ = 248 nm, IPEX864; Lightmachinery) with a fluence of 0.9 J cm−2 and \na repetition rate of 2 Hz was used. The thin films were synthesized at 500 ºC under an oxygen \npartial pressure of 100 mTorr . Each set of thin films with distinctive orientations (with the same \nnumber of laser shots) was simultaneously fabricated. \nAtomic and crystal structure characterization : The crystalline structure and lattice parameters \nof the LCO thin films were characteri zed using high-resolution X RD (PANalytical X’Pert \nPro). HAADF STEM measurements were performed on a Nion UltraSTEM200 microscope \noperated at 200 kV. The microscope was equipped with a cold field emission gun and a \ncorrector of third - and fifth -order aberra tions for sub -angstrom resolution. The collection \ninner half -angle for the HAADF STEM w as 65 mrad. Cross -sectional TEM specimens were \nprepared via ion milling after conventional mechanical polishing. \nSurface characterization: Atomic force microscopy measurements were performed using a \ncommercial system (Park Systems, NX10) to examine the surface topography and roughness. \nMagnetization measurements : A magnetic property measurement system (MPMS3; Quantum \nDesign) was used to characterize the i n-plane magnetic properties of the thin films . The \n13 \n \ntemperature -dependent magnetization , M(T), was measured from 300 to 2 K under a magnetic \nfield of 1000 Oe. The magnetic field -dependent magnetization , M(H), was measured at 2 K. \nChemical composition characterization : X-ray photoelectron spectroscopy (XPS; AXIS \nSUPRA, KRATOS Analytical) with Al Kα radiation was used to study the chemical state and \ncomposition of the LCO thin films. All the X -ray photoelectron spectra were calibrated using \nthe C -C bonding peak (284.5 eV). The Co 2 p and valence state s of the LCO thin films were \ndeconvoluted using Gaussian -Lorentz curves. \nOptical spectroscopy : The optical conductivit y of the LCO thin films w as measured using \nspectroscopic ellipsometry (J. A. Woollam Co., Inc.). A wavelength range from mid -infrared \nto UV (0.6 –6.2 eV) with incident angle s of 60 ° and 65 ° was used. We employed a three -layer \nmodel analysis (surface roughness (50% of material and 50% of void s), LCO, and STO layers) \nto obtain dielectric function s and optical conductivities of the thin films. \nTheoretic al calculation s: Ab initio DFT calculations were performed using the Vienna ab \ninitio simulation package (VASP) code.[63] The Perdew –Burke –Ernzerhof plus Hubbard \ncorrection (PBE + U + J) was used for the exchange -correlation functional ,[64] in which the \ndouble -counting interactions were corrected using the full localized limit (FLL).[65] The \nvalues used for the on -site direct Coulomb paramete r (U) and the anisotropic Coulomb \nparameter ( J) were 4.5 and 1.0 eV, respectively.[54] A plane wave basis set at a cutoff energy \nof 600 eV was used to expand the electronic wave functions, and the valence electrons were \ndescribed using projector -augmented wave potentials. All atoms were relaxed by the \nconjugate gradient algorithms until none of the remaining Hellmann –Feynman forces acting \non any atoms exceeded 0.02eV Å−1. \n \nSupporting Information \nSupporting Information is available from the Wiley Online Li brary or from the author . \n \nAcknowledgements \nDongwon Shin and Sangmoon Yoon contributed equally to this work. The authors thank M. \nF. Chisholm for technical assistance on TEM work . This work was supported by the Basic \nScience Research Programs through the National Research Foundation of Korea ( NRF -\n2021R1A2C2011340 and NRF -2020K1A3A7A09077715 ). The microstructural analysis work \n14 \n \nat ORNL was supported by U.S. DOE, Basic Energy Sciences, Materials Sciences and \nEngine ering Division. \n \nReceived: ((will be filled in by the editorial staff)) \nRevised: ((will be filled in by the editorial staff)) \nPublished online: ((will be filled in by the editorial staff)) \n \nReferences \n[1] K. Uchida, J. Xiao, H. Adachi, J. Ohe, S. Takahashi , J. Ieda, T. Ota, Y. Kajiwara, H. \nUmezawa, H. Kawai, G. E. Bauer, S. Maekawa, E. Saitoh, Nat. Mater. 2010 , 9, 894. \n[2] A. Hirohata, K. Yamada, Y. Nakatani, I. -L. Prejbeanu, B. Diény, P. Pirro, B. \nHillebrands, J. Magn. Magn. Mater. 2020 , 509. \n[3] H. Haugen , D. Huertas -Hernando, A. Brataas, Phys. Rev. B 2008 , 77. \n[4] H. X. Yang, A. Hallal, D. Terrade, X. Waintal, S. Roche, M. Chshiev, Phys. Rev. Lett. \n2013 , 110, 046603. \n[5] Z. Wang, C. Tang, R. Sachs, Y. Barlas, J. Shi, Phys. Rev. Lett. 2015 , 114, 016603. \n[6] P. Wei, S. Lee, F. Lemaitre, L. Pinel, D. Cutaia, W. Cha, F. Katmis, Y. Zhu, D. \nHeiman, J. Hone, J. S. Moodera, C. T. Chen, Nat. Mater. 2016 , 15, 711. \n[7] C. Tang, C. -Z. Chang, G. Zhao, Y. Liu, Z. Jiang, C. -X. Liu, M. R. McCartney, D. J. \nSmith, T. Chen , J. S. Moodera, J. Shi, Sci. Adv. 2017 , 3. \n[8] S. Sakai, S. V. Erohin, Z. I. Popov, S. Haku, T. Watanabe, Y. Yamada, S. Entani, S. \nLi, P. V. Avramov, H. Naramoto, K. Ando, P. B. Sorokin, Y. Yamauchi, Adv. Funct. \nMater. 2018 , 28, 1800462. \n[9] M. A. Ruderman, C. Kittel, Phys. Rev. 1954 , 96, 99. \n[10] T. Kasuya, Prog. Theor. Phys. 1956 , 16. \n[11] K. Yosida, Phys. Rev. 1957 , 106, 893. \n[12] C. Sohn, E. Skoropata, Y. Choi, X. Gao, A. Rastogi, A. Huon, M. A. McGuire, L. \nNuckols, Y. Zhang, J. W. Freeland, D. Haskel, H. N. Lee, Adv. Mater, 2019 , 31, \ne1805389. \n[13] T. Lin, C. Tang, J. Shi, Appl. Phys. Lett. 2013 , 103. \n[14] A. H. Hill, F. Jiao, P. G. Bruce, A. Harrison, W. Kockelmann, C. Ritter, Chem. Mater. \n2008 , 20, 4891. \n15 \n \n[15] E.-J. Guo, A. Herklotz, A. Kehlberger, J. Cramer, G. Jakob, M. Kläui, Appl. Phys. \nLett. 2016, 108. \n[16] T. R. McGuire, B. E. Argyle, M. W. Shafer, J. S. Smart, Appl. Phys. Lett. 1962 , 1, 17. \n[17] S. V. Houten, Phys. Lett. A 1962 , 2. \n[18] J. H. Lee, L. Fang, E. Vlahos, X. Ke, Y. W. Jung, L. F. Kourkoutis, J. W. Kim, P. J. \nRyan, T. Heeg, M. Roec kerath, V. Goian, M. Bernhagen, R. Uecker, P. C. Hammel, \nK. M. Rabe, S. Kamba, J. Schubert, J. W. Freeland, D. A. Muller, C. J. Fennie, P. \nSchiffer, V. Gopalan, E. Johnston -Halperin, D. G. Schlom, Nature 2010 , 466, 954. \n[19] D. Shin, I. Kim, S. Song , Y.-S. Seo, J. Hwang , S. Park, M. Choi , W. S. Choi , J. Am. \nCeram. Soc. 2021, 104, 4606 . \n[20] W. S. Choi, J. H. Kwon, H. Jeen, J. E. Hamann -Borrero, A. Radi, S. Macke, R. \nSutarto, F. He, G. A. Sawatzky, V. Hinkov, M. Kim, H. N. Lee, Nano. Lett. 2012 , 12, \n4966. \n[21] N. Zhang, Y. Zhu, D. Li, D. Pan, Y. Tang, M. Han, J. Ma, B. Wu, Z. Zhang, X. Ma, \nACS Appl. Mater. Interfaces 2018 , 10, 38230. \n[22] S. Xu, Y. Moritomo, K. Mori, T. Kamiyama, T. Saitoh, A. Nakamura, J. Phys. Soc. \nJapan 2001 , 70, 3296{3299. \n[23] S. Yoon, X. Gao, J. M. Ok, Z. Liao, M. G. Han, Y. Zhu, P. Ganesh, M. F. Chisholm, \nW. S. Choi, H. N. Lee, Nano. Lett. 2021 , 21, 4006. \n[24] T. Vogt, J. A. Hriljac, N. C. Hyatt, P. Woodward, Phys. Rev. B 2003 , 67. \n[25] D. P. Kozlenko, N. O. Golosova, Z. Jirák, L. S. Dubrovinsky, B. N. Savenko, M. G. \nTucker, Y. Le Godec, V. P. Glazkov, Phys. Rev. B 2007 , 75. \n[26] M. Capone, C. J. Ridley, N. P. Funnell, M. Guthrie, C. L. Bull, Phys. Status Solidi A \n2019 , 216. \n[27] K. Kleveland, N. Orlovskaya, T. Grande, A. M. M. Moe, M. -A. Einarsrud, J. Am. \nCeram. Soc. 2001 , 84, 2029. \n[28] S. Faaland, T. Grande, M. -A. Einar srud, P. E. Vullum, R. Holmestad, J. Am. Ceram. \nSoc. 2005 , 88, 726. \n[29] E.-J. Guo, R. Desautels, D. Keavney, M. A. Roldan, B. J. Kirby, D. Lee, Z. L. T. \nCharlton, A. Herklotz, T. Z. Ward, M. R. Fitzsimmons, H. N. Lee, Sci. Adv. 2019 , 5. \n[30] P. E. Vullum, R. Holmestad, H. L. Lein, J. Mastin, M. A. Einarsrud, T. Grande, Adv. \nMater, 2007 , 19, 4399. \n[31] P. E. Vullum, H. L. Lein, M. A. Einarsrud, T. Grande, R. Holmestad, Philos. Mag. \n2008 , 88, 1187. \n16 \n \n[32] J.-H. Kwon, W. S. Choi, Y. -K. Kwon, R. Jung, J. -M. Zuo, H. N. Lee, M. Kim, Chem. \nMater. 2014 , 26, 2496. \n[33] E. J. Guo, R. Desautels, D. Lee, M. A. Roldan, Z. Liao, T. Charlton, H. Ambaye, J. \nMolaison, R. Boehler, D. Keavney, A. Herklotz, T. Z. Ward, H. N. Lee, M. R. \nFitzsimmons, Phys. Rev. Lett. 2019 , 122, 187202. \n[34] D. Fuchs, E. Arac, C. Pinta, S. Schuppler, R. Schneider, H. v. Löhneysen, Phys. Rev. \nB 2008 , 77, 014434 . \n[35] D. Fuchs, L. Dieterle, E. Arac, R. Eder, P. Adelmann, V. Eyert, T. Kopp, R. \nSchn eider, D. Gerthsen, H. v. Löhneysen, Phys. Rev. B 2009 , 79, 024424 . \n[36] A. D. Rata, A. Herklotz, L. Schultz, K. Dörr, Eur. Phys. J. B 2010 , 76, 215. \n[37] V. V. Mehta, M. Liberati, F. J. Wong, R. V. Chopdekar, E. Arenholz, Y. Suzuki, J. \nAppl. Phys. 2009 , 105. \n[38] V. Mehta, Y. Suzuki, J. Appl. Phys. 2011 , 109. \n[39] V. V. Mehta, N. Biskup, C. Jenkins, E. Arenholz, M. Varela, Y. Suzuki, Phys. Rev. B \n2015 , 91. \n[40] J. Fujioka, Y. Yamasaki, H. Nakao, R. Kumai, Y. Murakami, M. Nakamura, M. \nKawasaki, Y. Tokura, Phys. Rev. Lett. 2013 , 111, 027206. \n[41] J. Fujioka, Y. Yamasaki, A. Doi, H. Nakao, R. Kumai, Y. Murakami, M. Nakamura, \nM. Kawasaki, T. Arima, Y. Tokura, Phys. Rev. B 2015 , 92. \n[42] Y. Li, S. J. Peng, D. J. Wang, K. M. Wu, S. H. Wang, AIP Adv. 2018 , 8. \n[43] D. Meng, H. Guo, Z. Cui, C. Ma, J. Zhao, J. Lu, H. Xu, Z. Wang, X. Hu, Z. Fu, R. \nPeng, J. Guo, X. Zhai, G. J. Brown, R. Knize, Y. Lu, PNAS 2018 , 115, 2873. \n[44] H. Liu, J. Fan, H. Zheng, A. Rahman, F. Qian, C. Wang, C. Ma, Y. Zhu, H. Yang, \nCurr. Appl. Phy s. 2021 , 28, 87. \n[45] H. Liu, J. Fan, J. Zhao, F. Qian, Y. Zhu, C. Ma, A. Rahman, L. Zhang, H. Yang, Phys. \nStatus Solidi B 2021 , DOI: 10.1002/pssb.202100424. \n[46] H. I. Seo, S. Woo, J. Kim, S. G. Jeong, T. Park, W. S. Choi, Phys. Rev. B 2021 , 103. \n[47] R. V. Chopdekar, E. Arenholz, Y. Suzuki, Phys. Rev. B 2009 , 79. \n[48] C. Liu, X. Yan, D. Jin, Y. Ma, H. -W. Hsiao, Y. Lin, T. M. Bretz -Sullivan, X. Zhou, J. \nPearson, B. Fisher, J. S. Jiang, W. Han, J. -M. Zuo, J. Wen, D. D. Fong, J. Sun, H. \nZhou, A. Bhattacharya, Science 2021 , 371, 716. \n[49] Z. Chen, Y. Liu, H. Zhang, Z. Liu, H. Tian, Y. Sun, M. Z. Y. Zhou, J. Sun, Y. Xie, \nScience 2021 , 372, 721. \n17 \n \n[50] Z. Chen, Z. Liu, Y. Sun, X. Chen, Y. Liu, H. Zhang, H. Li, M. Zhang, S. Hong, T. \nRen, C. Zhang, H. Tian, Y. Zhou, J. Sun, Y. Xie, Phys. Rev. Lett. 2021 , 126, 026802. \n[51] Y. Yokoyama, Y. Yamasaki, M. Taguchi, Y. Hirata, K. Takubo, J. Miyawaki, Y. \nHarada, D. Asakura, J. Fujiok a, M. Nakamura, H. Daimon, M. Kawasaki, Y. Tokura, \nH. Wadati, Phys. Rev. Lett. 2018 , 120, 206402. \n[52] D. W. Jeong, W. S. Choi, S. Okamoto, J. Y. Kim, K. W. Kim, S. J. Moon, D. Y. Cho, \nH. N. Lee, T. W. Noh, Sci. Rep. 2014 , 4, 6124. \n[53] C. Pinta, D. Fuchs, M. Merz, M. Wissinger, E. Arac, H. v. Löhneysen, A. Samartsev, \nP. Nagel, S. Schuppler, Phys. Rev. B 2008 , 78, 177402 . \n[54] G. E. Sterbinsky, R. Nanguneri, J. X. Ma, J. Shi, E. Karapetrova, J. C. Woicik, H. \nPark, J. W. Kim, P. J. Ryan, Phys. Rev. Lett. 2018, 120, 197201. \n[55] G. E. Sterbinsky, P. J. Rya n, J. W. Kim , E. Karapetrova, J. X. Ma, J. Shi, J. C. Woicik, \nPhys. Rev. B 2012, 85, 020403(R) . \n[56] V. Chaturvedi , J. Wlater, A. Paul, A. Gr utter, B. Kirby, J. S. Jeong, H. Zhou, Z. \nZhang, B. Yu, M. Greven, K. A. Mkhoyan, T. Birol, C. Leighton, Phys. Rev. Mater . \n2020, 4, 034403 . \n[57] E.-J. Guo, R. D. Desautels, D. Keavney, A. Herklotz, T. Z. Ward, M. R. Fitzsimmons, \nH. N. Lee, Phys. Rev. Mate r. 2019 , 3. \n[58] D. I. Khomskii, G. A. Sawatzky, Solid State Communications 1997 , 102, 87. \n[59] K. Saritas, J. T. Krogel, S. Okamoto, H. N. Lee, F. A. Reboredo, Phys. Rev. Mater. \n2019 , 3. \n[60] T. A. Mellan, F. Corà, R. Grau -Crespo, S. Ismail -Beigi, Phys. R ev. B 2015 , 92. \n[61] S. Yoon, K. Jin, S. Lee, K. T. Nam, M. Kim, Y. K. Kwon, Phys Chem Chem Phys \n2021 , 23, 859. \n[62] S. Yoon, S. Lee, S. Pang, M. Kim, Y. -K. Kwon, Phys. Rev. B 2021 , 104. \n[63] G. Kresse, J. Furthmüller, Phys. Rev. B 1996 , 54, 11169. \n[64] J. P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 1997 , 77, 3865. \n[65] A. I. Liechtenstein, V. V. Anisimov, J. Zaanen, Phys Rev B Condens Matter 1995 , 52, \nR5467. \n \n18 \n \n \n \nSupporting Information \n \n \nTunable Ferromagnetism in LaCoO 3 Epitaxial Thin Films \n \nDongwon Shin1,#, Sangmoon Yoon2,3,#, Sehwan Song4, Sungkyun Park4, Ho Nyung Lee2, and \nWoo Seok Choi1,* \n \n1Department of Physics, Sungkyunkwan University, Suwon 16419, Republic of Korea \n2Materials Science and Technology Division, Oak Ridge National Laboratory, Oak Ridge, \nTennessee 37831, United States \n3Department of Physics, Gachon University, Seongnam, 13120, Republic of Korea \n4Department of Physics, Pusan National University, Busan 46241, Republic of Korea \n \n*Corresponding author: choiws@skku.edu \n \n19 \n \nThickness dependent crystal structures of epitaxial LCO thin films \nHigh -quality LCO thin films were epitaxially grown on STO substrat es using PLE as shown \nin Fig ure 1a (30 nm), Fig ure S1a (10 nm) and Fig ure S1b (20 nm). Figure s S1c and 1d show \nthat 10 - and 20 -nm-thick LCO thin films also exhibit the same orientation -dependent trend in \nVu.c. and εoop, as the 30 nm -thick films, as shown in Figures 1c and 1d. \n \nAbsence of oxygen vacancy ordering \nFigure S3a shows HAADF STEM image for the film -interior region that EELS linescan is \nperformed. The EELS spectra of twin -wall and bulk -like domains are compared in Figure \nS3band c. First, there is no noticeable difference between two regions at Co -L2, 3 edge. It is \nwidely known that HS and LS Co3+ is hard to be distinguished in Co -L2, 3 edge because of \nsubtle changes. This results also exclude the presence of oxygen vacancy ordering in pristine \ntwin-wall domains, because the red -shift of Co -L3 edge is not observed in the tw in-wall \ndomains. Second, the prepeak (denoted by the black arrow) emerges at the O -K edge spectra \nof twin -wall domain. It would be the fingerprint that electronic structure, presumably spin \nstates, is different between twin -wall domain and bulk -like domain s. However, it is not the \ndirect experimental evidence that ferromagnetism forms in dark -striped regions. Our scenario \nis supported by theoretical results, and also explains the trend of unconventional magnetic \nproperties. Atomic -scale electron magnetic ci rcular dichroism at low temperature under \ndevelopment may provide direct experimental evidence to end this issue. \n \nEstimation of spin moment and HS/LS ratio \nTo characterize the FM properties of the LCO thin films, we analyzed the M(T) curves using \nthe Curie -Weiss (C -W) law. According to the C -W law, the C-W temperature, spin moment, \nand ratio of the HS-state were calculated. The inverse of the susceptibility H/M as a function \nof T above TC can be fitted by the C-W law H/M = (T – θ)/C, where C is a const ant and θ is \nthe C -W temperature (Fig ure S4a). All the LCO thin films exhibit ed a robust FM phase with θ \nsimilar to the TC values obtained in Fig ure 2e (Fig ure S4b). From the C values (slope of the \nC-W law fitting), we used μ = (3kBC/NA)1/2 ≈ 2.84 C1/2, where kB is Boltzmann’s constant and \nNA is Avogadro’s number, to determine the spin moment.[1] Furthermore, considering a \nsimple model of μ = g√S(S+1), where g ~ 2 is the gyromagn etic ratio of a free electron and S \nis the spin value, we determine the ratio of the HS -state, as summarized in Fig ure S4c. \n \n20 \n \nDetermining Co valence state and the ratio of HS state \nWe performed the X -ray photoelectron spectroscopy (XPS) to examine the binding energy of \nCo, which leads to the estimation of valence state as shown in Fig ure S5a. Distinct peaks of \nCo 2 p3/2 and 2 p1/2 are observed without any satellite peaks, indicating the Co3+ state within the \nLCO thin films. The absence of Co2+ indicates the absence of oxygen vacancies in the thin \nfilms. In addition, the amounts of LS and HS states were estimated. In Fig ure S5b, the peaks \nnear ~0.235, 0.935, 2.335 , and 4.735 eV correspond to the contributions of Co3+ HS, LS, O \n2p, and Co3+ 3d, respectively.[2, 3] From the integration of the HS and LS peaks, we obtained \nthe HS (LS) ratio of (100) -, (110) - and (111) -oriented LCO thin films to be 63.07% (36.93%), \n65.08% (34.92%) , and 61.16% (38.84%), respectively. \n \n \n \n \n \n \nFigure S1. Crystal structures of the epitaxial LCO thin films (10 and 20 nm) grown on \nSTO substrates. XRD θ-2θ scans near the (002), (110), and (111) Bragg planes of a) 10 nm \nand b) 20 nm LCO thin films (#) on (100), (110), and (111) STO substrates planes (*), \nrespectively . Crystallographic orientation -dependent unit cell volume ( Vu.c.), and out -of-plane \nstrain (εoop = 100 × (dbulk − dfilm) / dbulk) of the c) 10- and d) 20-nm-thick LCO thin films, \nrespectively. \n \n30 35 40 45 50(a)Intensity (arb. units)\n2q (degrees)10 nm\n (100)\n (110)\n (111)\n30 35 40 45 50(b)Intensity (arb. units)\n2q (degrees)20 nm\n (100)\n (110)\n (111)57.357.6(c)Vu.c. (Å3)10 nm\n-1.8-1.5-1.2\neoop (%)\n57.858.0(d)Vu.c. (Å3)20 nm\n(100) (110) (111)-0.9-0.6\neoop (%)\nOrientation \n21 \n \n \nFigure S2. Surface roughness of STO substrates and LCO thin films. The rms roughness \nare 0.118, 0.157, and 0.100 nm for the (100) -, (110) -, and (111) -oriented STO substrates, \nrespectively. In addition, the rms roughness are 0.141, 0.226, and 0.148 nm for ( 100)-, (110) -, \nand (111) -oriented LCO thin films, respectively. \n \n \n \n \nFigure S3. STEM -EELS (electron energy loss spectroscopy) results. a) HAADF STEM \nimage for the film -interior region that EELS linescan is performed. EELS b) O K-edge and c) \nCo L-edge spectra are respectively shown for the twin -wall and bulk -like domains. \n1 μmSTO (100) LCO (100)\nSTO (110) LCO (110)\nSTO (111) LCO (111)(a)\n(b)\n(c)\n \n22 \n \nc\n c\n c cctcc\nabc \nc-unit (100) (110) (111) \na [Å] 3.86 3.85 3.85 \nb [Å] 3.905 3.905 3.85 \nc [Å] 3.68 3.7 3.62 \nV [Å3] 55.47 55.63 53.66 \n \nTable S1. Structural prop erties of the c-units with different crystallographic symmetry \nin the FM state. Lattice parameters of a, b, c, and V of the c-units, shown for LCO thin films \nwith different crystallographic orientations . The schematics show distorted c-units by the \ncrystal lographic symmetry modulation. \n \n \n \nFigure S4. FM properties of LCO thin films from the Curie -Weiss law fitting. a) The \ninverse of the susceptibility H/M as a function of T above TC. The solid lines indicate Curie -\nWeiss law fitting H/M = (T − θ) / C, where C is constant, θ is the C -W temperature. The \ntemperature -dependent magnetic susceptibility provides spin moment and ratio of the HS state \nestimated by the C -W law. The summary of t he orientation -dependent FM properties; b) θ and \nc) HS state ratio. \n40 60 80 100010203040 (100)\n (110)\n (111)H/M (T (mB Co-1)-1)\nT (K)(a)\nq707580(b)\nq (K)\n(100) (110) (111)0.60.7(c)HS (%)\nOrientation \n23 \n \n \nFigure S5. Characterization of Co valence state and spin states. a) XPS spectra of Co 2 p \nlevels. b) XPS spectra near the valence -band regio n. c) Summary of the LS and HS ratio in the \nLCO thin films with distinct crystallographic orientations. \n \n \n \n \n \n \n \n \nAFM (HS) FM (HS/LS) FM (HS) NM (LS) \nEnergy \n[meV/f.u.] (100) 0 -55 274 297 \n(110) 0 -50 -25 277 \n(111) 0 -129 227 185 \n \nTable S2. Total energy calculations of the c-units with different symmetry in four different \nmagnetic configurations. The magnetic configurations are schematically demonstrated in the \nfirst raw of the table, where HS (orange sphere) and LS (blue sphere) of Co atoms and spins \n(black arrow) are shown. The energy of each magnetic configuration is represented relative to \nthat of the HS AFM configuration, which is set to zero energy. \n820 810 800 790 780 770\nBinding energy (eV)Co 2 p\n(111)(110)Intensity (arb. units)Co3+ 2p1/2Co3+ 2p3/2(100) (a) (c) (b) Co3+ 3d HS O 2p \n Co3+ 3d LS Co 3+ 3d\n Peak sum(100)Intensity (arb. units)(110)\n8 4 0\nBinding energy (eV)(111)(100) (110) (111)34363840606264\nLSSpin state (%)\nOrientationHS\n \n24 \n \n \nFigure S6. Proposed FM superexchange mechanism for the epitaxial LCO thin film. \nSchematic representation of FM HS Co3+ - Co3+ superexchange interaction promoted via empty \nLS Co3+ and filled O 2 p orbitals. HS Co3+ dx2−y2 (blue and yellow lobes), O2− pσ (purple lobes) \nand spin (black arrow) are shown. \n \n \n \n \n \n \n (100) (110) (111) \nMHS Co3+ \n(μB) 2.986 2.993 2.976 \n \nTable S3. The local magnetic moment of Co3+ within the c-units. \n \n \n \n \n \n \nReference s \n[1] M. Herlitschke, B. Klobes, I. Sergueev, P. Hering, J. Perßon , R. P. Hermann, Physical \nReview B 2016 , 93. \n[2] T. Saitoh, T. Mizokawa, A. Fujimori, M. Abbate, Y. Takeda, M. Takano, Physical \nReview B 1997 , 55. \n[3] A. G. Thomas, W. R. Flavell, P. M. Dunwoody, C. E. J. Mitchell, S. Warren, S. C. Grice, \nP. G. D. Marr, D. E. Jewitt, N. Khan, V. R. Dhanak, D. Teehan, E. A. Seddon, K. Asai, \nY. Koboyashi, N. Yamada, Journal of Physics: Condensed Matters 2000 , 3, 9259. \n \n \ndx2-y2pσ\nxy\nx\ny\nx\ny\nxy\nxy\npσHS Co3+ HS Co3+ LS Co3+ O O\ndx2-y2 dx2-y2" }, { "title": "1305.3158v1.Interface_induced_room_temperature_ferromagnetism_in_hydrogenated_epitaxial_graphene.pdf", "content": "1\nInterface induced room-temperature \nferromagnetism in hydrogenated epitaxial graphene \n \nA.J.M. Giesbers1,*, K. Uhlířová2,M. Konečný1, 4, E. C. Peters3, M. Burghard3, J. Aarts2, and \nC.F.J. Flipse1,† \n \n1Molecular Materials and Nanosystems, Eindhoven University of Technology, 5600 MB \nEindhoven, The Netherlands. \n2Magnetic and Superconducting Materials, Leiden Institute of Physics, 2333 CA Leiden, The \nNetherlands. \n3Max-Planck Institute for Solid State Research, Heisenbergstrasse 1, D-70569 Stuttgart \nGermany. \n4CEITEC BUT, Technická 10, 616 69 Brno, Czech Republic \n \nContact information \n*A.J.M.Giesbers@tue.nl, †C.F.J.Flipse@tue.nl \n 2\nDue to the predominantly surf ace character of graphene, it is highly suitable for \nfunctionalization with external atoms and/or molecules leading to a plethora of new and \ninteresting phenomena. Here we show ferromagnetic properties of hydrogen-\nfunctionalized epitaxial graphene on SiC. Ferromagnetism in such a material is not \ndirectly evident as it is inherently compos ed of only non-magnetic constituents. Our \nresults nevertheless show strong ferromagnet ism, which cannot be explained by simple \nmagnetic impurities. The ferromagnetism is unique to hydrogenated epitaxial graphene \non SiC, where interactions with the interfacial buffer layer play a crucial role. We argue \nthat the origin of the observed ferromagn etism is governed by electron correlation \neffects of the narrow Si-dangling-bond (S i-DB) states in the buffer layer exchange-\ncoupled to localized states in the hydroge nated graphene layer. This forms a quasi-\nthree-dimensional ferromagnet with a Curie temperature higher than 300 K. \n \nOwing to its capability of ballistic transport over micrometer distances1, as well as its very \nlong spin relaxation time and spin relaxation length2, 3, graphene represents a close-to-ideal \nmaterial for spintronic applications4. In this context, considerable effort has recently been \ndirected to rendering graphene ferromagnet ic via chemical modi fication. Thus far, \nferromagnetic order in graphene has been attained through covalent functionalization, \ninvolving the linkage of radical species like the spin-bearing carbon atom of an organic molecule or hydrogen atoms to the graphene layer\n5-17. Along these lines, functionalization of \nepitaxial graphene by aryl radicals has been reported to yield disordered magnetism, \ncomprising a mixture of ferromagnetic, supe rparamagnetic and antiferromagnetic regions18. \nWith the aid of combined atomic and magnetic force microscopy, it could be proven that these randomly dispersed regions are constituted by the attached moieties. This lack of a \nperiodic functionalization pattern of the gra phene sheet prevents the achievement of long \nrange ferromagnetic order, thus limiting the use of such samples in spintronic devices. \nFurthermore, room temperature ferromagnetism has been detected in partially hydrogenated \nepitaxial graphene grown on silicon carbide (SiC), and attributed to hydrogen monomers bonded to the graphene sheet\n12. Despite these accomplishments, however, both the \nmechanism underlying the ferromagnetic ordering, and the role played by the SiC substrate \nused for the epitaxial graphene growth, has not yet been clarified. Here, we experimentally \ndemonstrate that spin ordering within hydrogena ted epitaxial graphene critically depends on \nthe presence of the underlying buffer layer. In a ddition, it is shown that the created magnetic 3\nareas are distributed over the entire graphene sheet, thus enabling to effectively tune the \noverall magnetization through the density of attached hydrogen atoms. \n \nTo explore the ferromagnetism in epitaxial graphene, we use samples grown on insulating \n6H-SiC substrates following the procedure described in ref. 19 (all samples originate from \nthe same wafer). The atomic force micrograph (A FM) of the sample surface (Fig. 1a) reveals \nterrace steps originating from a slight miscut of the SiC substrate. The terraces are typically \n3-5 um wide and approximately 10 nm high and are overgrown with a continuous carpet of \ngraphene20, 21. The inset in Fig. 1a shows a schematic cross-section of the layer sequence at \nthe surface with a graphene layer on top of an interfacial carbon layer (buffer layer) partly \nbonded to the SiC substrate. On the terrace edges an unintentional region of bilayer graphene has formed under the current gowth conditions\n19, discernible as brighter areas in the \ncorresponding AFM phase image (Fig. 1b). The presence of a small bilayer area is confirmed \nby Raman microscopy and low energy electron microscopy (LEEM) investigations (see \nsupplementary information Fig. S1) and has the same coverage in all samples. After growth, \nthe graphene samples are hydrogenated by an atomic hydrogen source in an ultrahigh vacuum \nchamber, for different exposure times. Succ essful hydrogenation is testified by an \nenhancement of the sp3-defect associated Raman D-peak, whose intensity increases with \ntreatment time (Fig. 1c and d), as discussed in more detail in the supplementary information. Increasing hydrogen exposure also leads to a rising C-H signal in x-ray photo-absorption \nspectra (XPS) (see supplementary information). The inset in Fig. 1d illustrates the hydrogen \nbonded on the top graphene layer. \n The magnetic properties of the hydrogenated graphene samples are determined using a \ncommercial SQUID with a sensitivity of 5·10\n-8 emu. All measurements are performed at \nroom temperature unless stated otherwise. Figure 2a shows the magnetization of an epitaxial \ngraphene sample hydrogenated for 3 minutes. The linear background is related to the bulk \nSiC diamagnetism and can be subtracted by a li near fit to the high field part of the curve \nwhere all other forms of magnetism are assumed to be saturated. The resulting diamagnetic \nsusceptibility χ = 0M/Hm , with m = (1.97 ± 0.05)·10-5 kg the sample mass and 0 = 4·10-7 \nTm/A the vacuum permeability, is χSiC = -(4.1 ± 0.1)·10-9 m3/kg, within its error in good \nagreement with literature ( χSiC = -4.01·10-9 m3/kg). Consistent values for χSiC, within the error \nrange, were found for all samples used in this work. The data obtained after subtraction of the 4\nSiC diamagnetic background are shown in Fig. 2b for three different temperatures. The \ncurves show a clear ferromagnetic response fr om the hydrogenated epitaxial graphene. The \nhysteresis loop shows a saturation magnetization of Ms = ± 27·10-7 emu, a remanent \nmagnetization of Mr = ± 7·10-7 emu and a coercive field of Hc = ± 91 Oe at 300 K. Upon \ndecreasing the temperature a small increase in the high field magnetization occurs. A similar \ntrend is observed for the coercive field and th e remanent magnetization (inset Fig. 2b). The \nmeasured saturation magnetization at room temp erature corresponds to a value of about 0.9 B \nper unit cell. \n \nFigure 2c compares the ferromagnetic signal for the 3 min hydrogenated sample under in-\nplane magnetic field, along ( θ = 0 deg) and perpendicular ( θ = 90 deg) to the terraces, as well \nas for out-of-plane (OofP) orientation (inset Fig. 2c). A notable anisotropy can be discerned, \nwith easier magnetization along the terrace steps (black curve), as compared to perpendicular \nalignment (red curve) and the out-of-plane dir ection (blue curve). This difference manifests \nitself in a lower saturation magnetization and in the case of the out-of-plane signal in a more stretched hysteresis loop. The preferred ma gnetization along the terrace edges might result \nfrom the predominant formation of double site hydrogen sites aligned along the zigzag \ndirection of graphene\n22.The double H-sites show elongated sh aped charge structures of 3 nm \nor more with 6-fold symmetry coinciding with the 6 fold symmetry of the graphene \nhoneycomb lattice. In atomic resolution STM it was shown that the armchair edge of the graphene layer coincides with the SiC terrace structure\n23. Combined, these results lead to \nanisotropy between the terrace edge and perpendicular to the edge direction which could \nexplain the observed anisotropy in the magn etization. The out of plane magnetization \ncontribution is probably due to a non collinear spin orientation in the buffer layer, similar as \nfor the √3x√3R30 6H -SiC(0001) structure of SiC24, which will be discussed later. \n \nTo tune the ferromagnetic signal we can use the hydrogen coverage, as is shown in Fig. 2d for hydrogenation times between 0 and 120 minutes. While the pristine graphene (0 min, \nblack curve) displays no magnetic signal, a short hydrogen exposure (0.5 min, red curve) \nresults in a clear ferromagnetic signal. From the corresponding hysteresis loop, a coercive \nfield of H\nc = ± 65 Oe and remanent magnetization of Mr = ± 2.4·10-7 emu is extracted. At \nhigh fields ( H = 3000 Oe), the magnetization reaches a saturation value of 14·10-7 emu. This \nsaturation magnetization, Ms, increases up to a treatment time of 3 minutes (27·10-7 emu), 5\nwhich is followed by a decrease for longer treatments, finally resulting in Ms (120 min) = \n13·10-7 emu. The same trend is observed in the coer cive field and the remanent magnetization \nfor the different samples. \n \nIn order to determine the origin of the ferromagnetic behavior, we have investigated the \nmagnetization properties of severa l control samples (Fig. 2c). Firstly, a sample prepared in \nthe same manner as the 3 minute sample, except that the hydrogen bottle is kept closed, is \nfound to exhibit no ferromagnetic signal (red curve). Secondly, the same procedure is applied \nto an untreated bare SiC sample, which likewise does not lead to ferromagnetic signatures \n(not shown). Thirdly, to test the influence of the underlying substrat e, a quasi-freestanding \nmonolayer of graphene24 (QFMG) is used as a third control sample (Fig. 3f shows a \nschematic). It is obtained by growing only a buffer layer24-26 on the SiC, followed by \nhydrogen intercalation to passivate the SiC subs trate and turn the buffer layer into a QFMG. \nOwing to the reduced substrate interaction, QFMG is of superior quality compared to \nepitaxial graphene24. Pristine QFMG samples not subjected to hydrogenation (exemplified by \ngreen curve) do not display ferromagnetism as expected for pure graphene (in total two \nsamples were studied). Remarkably, also afte r 3 minutes of hydrogenation, no ferromagnetic \nsignal at room temperature emerges for such samples (hQFMG, schematic in Fig. 3f) (blue \ncurve representative for one out of two samples). The above findings highlight that the hydrogenated graphene is not ferromagnetic at room temperature and the buffer layer is \ncrucial to render the epitaxial graphene ferromagnetic. Finally, in a fourth control experiment \nusing two buffer layer samples (one shown, cyan curve) and three hydrogenated buffer layer \nsamples (one shown, magenta curve) no ferromagne tic signal is detected (schematics of the \nsamples are shown in Fig. 3f). This absence consolidates that hydrogenated epitaxial \ngraphene requires both the hydrogenated graphene and the underlying buffer layer to become \nferromagnetic. At low temperatures, the linear background magnetization, observable for \nboth the buffer layer samples and the hQFMG, leads to a smaller χ\nSiC compared to the pure \nSiC substrates. This difference hints toward an unsaturated low temperature paramagnetic contribution in these samples, akin to fluorinated graphene laminates\n27.The presence of \nlocalized paramagnetic-like states in the buffer layer was recently also suggested from spin \ntransport experiments in epitaxial graphene28. In our preliminary high magnetic field \nmagnetization measurements the paramagnetism of the bufferlayer is indeed confirmed, \nsaturating at H/T 25 kOe/K (see supplementary information Fig S4). 6\n \nFurther insight into the ferromagnetic properties of the hydrogenated graphene is gained by \ndetecting the remanent magnetization with th e aid of Magnetic Force Microscopy (MFM) \n(see Fig. 3). By placing the sample briefly on either the south pole (-B) or north pole (+B) of \na permanent magnet prior to MFM measurements, we can magnetize the sample in \nrespectively a negative or positive out-of-plane remanent magnetization state as we observed in the SQUID measurements of Fig. 2c. Figure 3a and b show the magnetic signal of the \nsame area for the two magnetization directions with their respective cross-sections in panel c. \nThe highlighted dirt particle is an artifact due to crosstalk with the topography\n29 and serves as \na position marker on the sample. The topology of the sample is similar to Figure 1a and b. \nThe clear difference in MFM contrast between the single (1L) and bilayer (2L) areas, indicate \ntheir different magnetization. This might be due to different hydrogen coverages30, 31, in \naccord with the lower overall D-peak intensity on the bilayer regions observed in Raman \nimages. Other possible contributions are the different electronic structure of the bilayer \ngraphene, as well as different interactions among the hydrogen sites, or in the specific case of \nthe bilayer graphene between hydrogen sites and the buffer layer due to the increased \ndistance between the buffer layer and the hydrogenated layer. In the SQUID measurements \nthe bilayer areas will reduce the overall saturation magnetization, however since the bilayer \ncoverage is similar for all samples the results above are not affected. The switching of the out-of-plane remanent magnetization direction is clearly visible in the MFM cross-sections in \nFig. 3c. Specifically, after positive B-field magn etization, the MFM signal is positive and the \nsignal from the single layer is slightly larger th an that from the bilayer. After negative B-field \nmagnetization the MFM signal has reversed the sign and the response from the single layer is again highest. These changes show that the color inversion between panel a and b is due to a \ncomplete flip of the magnetization direction, while the signal from the single layer is always \nhigher than that from the bilayer. That the flip is not symmetric around zero indicates a \nconstant background phase shift, and might be attributed to electrostatic interactions \nsimultaneously probed by the metallic tip. Electric field microscopy (EFM) confirmed this magnetic-field independent electrostatic background\n32 (see supplementary figure S3). \n \nThe MFM measurements corroborate the ferromagnetism of the hydrogenated epitaxial \ngraphene sample and show that the signal originates from the whole surface. Together with \nthe observed variation of the ferromagnetic strength with hydrogen coverage, the magnetic anisotropy and control sample magnetic measurements these results form a conclusive set of 7\nobservations which rule out any possible magnetic contaminations as the origin of the \nobserved magnetic behavior. \n \nThe observed ferromagnetism in our hydrogenated epitaxial graphene is best interpreted in \nterms of an exchange coupled interaction between localized electron states of the buffer-layer \nand either spin-polarized localized states or th e mid-gap states of the hydrogenated graphene \nlayer22. The overall paramagnetic behavior of the buffer-layer indicates the presence of \nlocalized magnetic moments, which are the localized defect states forming an insulating \nbehavior as has been shown by STM and STS experiments33, 34.These states are attributed to \nSi-dangling bond (DB) states which probably be have as Hubbard Coulomb repulsion driven \nNeel like states, described by a non-collinear spin density wave, similar as was shown by \nAnisimov35 for a smaller unit cell reconstructed surface, the √3x√3R30 6H -SiC(0001). The \nSiC buffer-layer has a 6 times larger unit cell, 6 √3x6√3R30 SiC(0001) surface structure with \na band gap of 1 eV, formed by localized Si-DB states33, 34. Upon hydrogen adsorption on top \nof the graphene layer, carbon-hydrogen bonds are created, forming a mid-gap state22.This \nlocalized mid-gap state can be spin-split in filled and unfilled localized states close to the \nFermi-level due to the Coulomb interaction of the Si-DB states of the buffer layer, forming a \nquasi 3-dimensional ferromagnetic state with a Curie temperature of 300 K or higher. A \nsecond option to explain the FM behavior at 300 K is that the hydrogenated graphene layer is intrinsically ferromagnetic, but with a much lower Curie temperature due to the two-\ndimensionality, which would become quasi-three dimensional if the paramagnetic buffer-\nlayer will exchange couple to it. \n To conclude, hydrogenated epitaxial graphene shows a ferromagnetic behavior with a Curie \ntemperature higher than 300 K and a magnetic moment of 0.9\nB per effective carbon hexagon \narea. We have shown that both the hydrogen coverage and the buffer-layer with the Si-\ndangling bonds, play a crucial role for the hi gh temperature ferromagnetic properties. To \nexplain the ferromagnetism in our graphene system at room temperature, we tentatively \npropose an exchange coupled interaction between the Coulomb induced localized Si-DB \nstates of the buffer-layer and the localized mid-gap state or the two-dimensional \nferromagnetic hydrogenated graphene layer. The buffer-layer stabilizes the ferromagnetic \nbehavior at room temperature and this quasi three-dimensional system can explain the relatively high Curie temperature, higher than 300 K. The high Curie temperature in 8\ncombination with a small coercive field (100 Oe ) and high spin relaxation time in graphene \nmakes hydrogenated epitaxial graphene a favor able material for spintronic applications. \n \n \n \n \n \n \n 9\nReferences \n1. Ki, D-K., Jeong, D., Choi, J-H., Lee, H-J., Park, K-S., Inelastic scattering in a monolayer graphene sheet: A weak-localization study, Phys. Rev. B 78, 125409 (2008). \n2. Tombros, N., Jozsa, C., Popinciuc, M., Jonkman, H. T., van Wees, B. J., Electronic spin \ntransport and spin precession in single graphene layers at room temperature, Nature 448, 571 \n(2007). 3. Maassen, T. et al., Long spin relaxation times in wafer scale epitaxial graphene on \nSiC(0001), Nano Lett. 12, 1498 (2012). \n4. Candini, A., Klyatskaya, S., Ruben, M., Wernsdorfer, W., Affronte, M., Graphene \nSpintronic Devices with Molecular Nanomagnets, Nano Lett. 11, 2634 (2011). \n5. Chen, L., Yu, D., Liu, F., Magnetism in nanopatterned graphite film, Appl. Phys. Lett. 93, \n223106 (2008). \n6. Son, Y-W., Cohen, M. L., Louie, S. G., Half-metallic graphene nanoribbons, Nature 444, \n347 (2006). \n7. Hong, J. et al., Room-temperature Magn etic Ordering in Func tionalized Graphene, Sci. \nRep. 2, 624 (2012). \n8. Brey, L., Fertig, H. A. , Das Sarma, S., Diluted graphene antiferromagnet, Phys. Rev. Lett. \n99, 116802 (2007). \n9. Lehtinen, P. O. et al., Magnetic Properties and Diffusion of Adatoms on a Graphene Sheet, \nPhys. Rev. Lett. 91, 017202 (2003). \n10. Wu, M., Liu, E-Z, Jiang, J. Z., Magnetic behavior of graphene absorbed with N, O, and F \natoms: A first-principles study, Appl. Phys. Lett. 93, 082504 (2008). \n11. Boukhvalov, D. W., Katsnelson, M. I., Lichtenstein, A. I., Hydrogen on graphene: \nElectronic structure, total Energy, structural distortions and magnetism from first-principles \ncalculations, Phys. Rev. B 77, 035427 (2008). \n12. Xie, L. et al., Room temperature ferromagnetism in partially hydrogenated epitaxial \ngraphene, Appl. Phys. Lett. 98, 193113 (2011). \n13. Boukhvalov D. W. and Katsnelson, M. I ., Chemical functionalization of graphene, J. \nPhys.: Condens. Matter 21, 344205 (2009). \n14. Kuila, T. et al., Chemical functiona lization of graphene and its applications, Progress in \nMaterials Science 57, 1061–1105 (2012). \n15. Wei, W. and Qu, X., Extraordinary Physical Properties of Functionalized Graphene, \nSmall 8, 2138 (2012). 10\n16. Papagno, M. et al., Large Band Gap Opening between Graphene Dirac Cones Induced by \nNa Adsorption onto an Ir Superlattice, ACS Nano 6, 199 (2012). \n17. Suk, J. W., Piner, R. D., An, J., Ruoff, R. S., Mechanical Properties of Monolayer \nGraphene Oxide, ACS Nano 4, 6557 (2010). \n18. Hong, J. et al., Effect of Nitrophenyl F unctionalization on the Magnetic Properties of \nEpitaxial Graphene, Small 7, 1175–1180 (2011). \n19. Emtsev, K. V. et al., Towards wafer-size graphene layers by atmospheric pressure \ngraphitization of silicon carbide, Nature Mat. 8, 203 (2009). \n20. Norimatsu, W., Kusunoki, M., Fromat ion process of graphene on SiC (0001), Physica E \n42, 694 (2010). \n21. Poon, S. W., Chen, W., Wee, A. T. S., and Tok E. S., Growth dynamics and kinetics of monolayer and multilayer graphene on a 6H-SiC(0001) substrate, Phys. Chem. Chem. Phys. \n12, 13522 (2010). \n22. Scheffler, M. et al., Probing Local Hydrogen Impurities in Quasi-Free-Standing \nGraphene, ACS Nano 6, 10590 (2012). \n23. Hu, T. W., Ma, D. Y., Ma, F., Xu, K. W., Preferred armchair edges of epitaxial graphene \non 6H-SiC(0001) by thermal decomposition, Appl. Phys. Lett. 101, 241903 (2012). \n24. Speck, F. et al., The quasi-free-standing na ture of graphene on H-saturated SiC(0001), \nAppl. Phys. Lett. 99, 122106 (2011). \n25. Ferralis, N., Maboudian, R., and Carraro, C., Ev idence of Structural Strain in Epitaxial \nGraphene Layers on 6H-SiC(0001), Phys. Rev. Lett. 101, 156801 (2008). \n26. Emtsev, K. V., Speck, F., Seyller, Th., Ley, L., Riley, J. D., Interaction, growth, and \nordering of epitaxial graphene on SiC{0001} surfaces: A comparative photoelectron spectroscopy study, Phys. Rev. B 77, 155303 (2008). \n27. Nair, R. R. et al., Spin-half paramagnetism in graphene induced by point defects, Nature \nPhys. 8, 199 (2012). \n28. Maassen, T. et al., Localized States Infl uence Spin Transport in Epitaxial Graphene, \nPhys. Rev. Lett. 110, 067209 (2013). \n29. Schendel, P. J. A. van, Hug, H. J., Stiefel, B., Martin, S., and Güntherodt, H-J., A method \nfor the calibration of magnetic force microscopy tips, J. Appl. Phys. 88, 435 (2000). \n30. Luo, Z. et al., Thickness-dependent reversible hydrogenation of graphene layers, ACS \nNano 3, 1781-1788 (2009). \n31. Ryu, S. et al., Reversible Basal Plane Hydrogenation of Graphene, Nano Lett. 8, 4597–\n4602 (2008). 11\n32. Martínez-Martín, D., Jaafar, M., Pérez, R., Gómez-Herrero, J., and Asenjo, A., Upper \nBound for the Magnetic Force Gradient in Graphite, Phys. Rev. Lett. 105, 257203 (2010). \n33. Nie, A., Feenstra, R. M., Tunneling spectros copy of graphene and related reconstructions \non SiC(0001), J.Vac.Sci.Technol. A27, 1052 (2009). \n34. Cervenka, J., Ruit van der, K., Flipse, C. F. J., Effect of local doping on the electronic \nproperties of epitaxial graphene on SiC, Phys.Status Solidi A 207, 595 (2010); Cervenka, J., \nRuit van der, K., Flipse, C. F. J., Giant inelastic tunneling in epitaxial graphene mediated by \nlocalized states, Phys. Rev. B 81, 205403 (2010). \n35. Anisimov, V. I. et al., SiC(0001): A surface Mott-Hubbard insulator, Phys Rev. B 61, \n1752 (2000). \n 12\nMethods \nFor the magnetization measurements we used a Magnetic Properties Measurement System \nfrom Quantum Design with Reciprocating samp le magnetometer. The best resolution was \nobtained with an amplitude of 3 cm and a frequency of 2Hz with a system resolution of 5·10-8 \nemu. The samples were placed into standard ø 6 mm PE straws which exactly fixated the 4x4 \nmm sample without further adhesives. \nThe magnetic signal in the MFM measurements is acquired by a magnetic Co-Cr coated tip \nwith a force constant of k = 2.8 N/m and a quality factor Q = 226 in constant lift mode of 50 \nnm and an oscillation amplitude of 30 nm is used to avoid short range and van der Waals \ninteractions. The MFM signal, or phase-shift \nθ, is directly related to the magnetic force F \nbetween the tip and substrate: \n Δߠ ൎ െொ\nௗி\nௗ௭ , \nleading to a force in the order of nano-Newtons. \n 13\nFigures: \n \nFigure 1| Epitaxial graphene on 6H-SiC. (a) Atomic force micrograph of the typical SiC \nterrace structure on top of which the graphene is grown. The inset shows a schematic of the \nsurface structure with from top to bottom a graphene layer, a buffer layer and the SiC \nsubstrate. The gray spheres represent carbon, the blue spheres silicon, and the yellow ovals the silicon dangling bonds. (b) Phase image showing the single layer graphene areas on top of \nthe terraces and the narrow bilayer regions at the terrace edges. (c) Raman spectra of \nhydrogenated graphene with a treatment time t = 0 min. (black), 30 min. (red) and 120 min. \n(blue). Clearly visible is the upcoming D-peak intensity with increasing treatment time. (d) \nD-peak intensity as a function of treatment time (the line is a guide to the eye). The inset shows the schematic bonding of hydrogen (orange spheres) to the graphene layer. \n \n4 m 4 m\n1500 2000 2500 3000 04 0 8 0 1 2 02D G\n Intensity (A.U.)\nRaman shift (cm-1) 120 min\n 30 min\n 0 minDD-Peak Intensity (A.U.)\nt (min)\n0 180 nm 90\na b\ncd\n14\n \nFigure 2| Magnetization of hydrogenated epitaxial graphene. (a) Room temperature \nmagnetization as a function of the applied magn etic field for hydrogena ted epitaxial graphene \ntreated for 3 minutes. The red line shows the di amagnetic contribution of the SiC substrate. \n(b) Temperature dependence of the magnetiza tion after subtraction of the diamagnetic \nbackground showing a clear ferromagnetic hysteresis loop. The inset shows a zoom of the \ncoercive field and remanent magnetization (3 min. treatment). (c) Direction dependence of \nthe magnetization (3 min. treatment). (d) Ferromagnetic signal for different hydrogen \ntreatment times, t = 0, 0.5, 3, 10, 30, 90, and 120 minutes. (e) Magnetization for different \ncontrol samples. (f) Schematic representati ons of the various control samples. \n -1000 0 1000-30-1501530\n90o0oOofP\n-1000 0 1000-30-1501530\n-1000 0 1000-30-150153002 0 0 0 4 0 0 0-30-1501530\n-150 0 150-10010\n0 2000 4000-300-200-1000100\n \n 10 min 0 min\n 0.5 min\n 3 min \n 30 min\n 90 min\n 120 min\n M (10-7 emu)\nH (Oe)\nM (10-7 emu)\nH (Oe) =0o\n = 90o\n OofPM (10-7 emu)\n hQFMG\n BL\n hBL 3 min\n SiC+G\n QFMG\n \nH (Oe)M (10-7 emu)\n 4K\n 77K\n 300K\n \nH (Oe)M (10-7 emu)\n \nH (Oe) M (10-7 emu)\nH (Oe) 3 min\n SiC diam.\nT = 300 Kabc\nde\nQFMG hQFMG\nBL hBLf15\n \nFigure 3| Magnetic Force Micrograp hs of hydrogenated graphene. (a) Magnetic force \nmicrograph after applying a positive magnetic field to the sample showing high and low remanent magnetization for single and bilayer respectively (scale: +3.3 ± 0.2 deg). (b) \nInversion of the remanent magnetization after applying a negative magnetic field to the sample (scale: -14.4± 0.2 deg). (c) Cross-s ection of the positive (a) and negative (b) \nmagnetization. We repeated the switching be tween positive and negative magnetization \nseveral times yielding the same result. \n ac\nb3.23.33.4\n0123-14.5-14.4-14.3MFM signal (deg)\n \nx (m)-B+B2L 1L 2L 1L 2L\n2L 1L 2L 1L 2L\n1 m2L 1L 2L 1L 2L\n+B\n‐B 1 m2L 1L 2L 1L 2L16\nSupplementary information \n \nSamples \nThe epitaxial graphene samples used in th is research are grown on insulating 6H-SiC \nsubstrates from II-VI Inc. following the procedur e described in ref. [S1]. All samples in this \narticle originate from the same wafer from II-VI inc. (EI1037-07-EV) that was diced in 4x4 \nmm pieces by laser cutting. After growth the sample surface consists of 1-5 m wide terraces \nwith a step height of about 10 nm as was shown by AFM in the main text. The terraces are \ncovered with single layer graphene with a narrow strip of bilayer graphene at the edges. In \nthe main text this was visualized by the AFM phase image, figure S1 shows a typical Raman map of the surface confirming the alternating structure by plotting the 2D-peak area and 2D-\npeak position, respectively Fig. S1a and b. The dark red areas indicate single layer graphene \nand the bright yellow areas indicate bilayer graphene as can be seen from the complete \nspectra at the red and blue dot displayed in Fig. S1c with matching colors. Figure S1d shows \na low energy electron microscopy (LEEM) image of a similar sample again showing the alternating single and bilayer area. The respective intensity versus energy spectra taken on \nthe numbered areas are shown in Fig. S1e. The number of dips in the spectra corresponds to \nthe number of layers. \n \nHydrogenation of graphene \nAfter growth the samples are covered with hydr ogen by an atomic hydrogen source in an \nultra high vacuum chamber. The hydrogenation is done at a pressure of 2·10\n-6 mbar and \ndifferent coverages are achieved by varying the exposure time. After hydrogenation the \nRaman signature of the sample has drastically changed. Figure 1c (main text) shows the \nRaman spectrum before and after hydrogenation with a treatment time of 30 minutes and 120 \nminutes. The main difference between the hydrogenated spectra and the non-hydrogenated \nspectra is the appearance of an additional peak at 1375 cm-1 [S2], the so-called D-peak. The \nD-peak is caused by breathing modes of sp2 rings and is activated by a defect [S3]. The \nchemisorption of hydrogen on grap hene is expected to create sp3 bonds. These sp3 bonds act \nas defects and allow sp2 breathing modes, which show up as a D-peak in the Raman \nspectrum. The intensity of the D-peak is dire ctly related to the amount of defects present \n[S3], in our case the amount of hydrogen. Figure 1d (main text) shows the evolution of the D-\npeak intensity as a function of treatment ti me, which clearly shows the increased hydrogen 17\ncoverage with treatment time. Recently it was phenomenologically shown [S4] that the \nintensity ratio of the D and D’ peak (at 1620 cm-1) for sp3 defects was ~10 or larger. In our \nSiC-graphene samples the SiC background [S1] impedes a proper analysis of the D’ peak, \nhowever we repeated the same hydrogen treatments on graphene flakes on SiO 2 which indeed \nshowed a I(D)/I(D’) ratio of 11, corroborating that we indeed have hydrogenated our \nsamples. X-ray photo-absorption spectroscopy \nIn addition, the presence of hydrogen is confirmed by X-ray photo-absorption spectroscopy (XPS) measurements (see Fig. S2a-c) [S5-S7]. Figure 2a shows the XPS spectrum for an \nuntreated sample (black line). In accordance to ref. [S5] the spectra can be fitted with four \nmain peaks, one originating from the SiC substrate (red), two from the two differently \nbonded carbon atoms in the bufferlayer (S1: magenta, S2: cyan) and one from the graphene \nlayer (green). After the sample is treated with hydrogen the main left peak starts to shift to \nhigher binding energies. This shift can be accounted for by adding a new peak for the binding \nenergy of carbon-hydrogen bonds at 284.74 ± 0.10 eV (brown peak in Fig. S2b and c) [S6]. \nThe intensity of this C-H peak increases w ith increasing treatment time, whereas the C-C \npeak intensity decreases as expect ed for higher hydrogen coverage. \n \nElectric Force Microscopy \nTo rule out any electrostatic contributions to the MFM signal in the main text these interactions between a metallic AFM tip and the hydrogenated graphene sample, are \nmeasured by electric force microscopy (EFM). Figure S3a shows the AFM topography of the sample measured simultaneously with the EF M signal. Panel b and c show the EFM signal \nafter the sample was place on respectively the south pole (-B) and the north pole (+B) of a \npermanent magnet. Neither of the EFM measurements shows a response to the magnetic field \ndirection. This indeed confirms that the magnetic field dependent MFM results are related to \nthe remanent magnetization of the hydrogena ted epitaxial graphene. The electrostatic \ninteractions with the metallic tip only lead to a constant offset in the MFM measurements as \nwas indeed shown in the main text. \n Paramagnetism in the buffer layer \nTo confirm if the smaller diama gnetic background (smaller than χ\nSiC), measured for the \nbuffer layer, is related to a pa ramagnetic signal in this layer we perform a high magnetic field 18\nmagnetization measurement at low temperatures (see Fig. S4). After subtraction of the \ndiamagnetic sample substrate background ( χSiC= -(4.1 ± 0.1)·10-9 m3/kg) a clear paramagnetic \nsignal becomes visible. As a function of H/T both the 2.5 K and 4 K curve coincide and can \nbe fitted with a single Brillouin function \nܯൌߤܬ݃ܰ ቈ2ܬ1\n2ܬ݄݊ݐܿቆሺ2ܬ1ሻݖ\n2ܬቇെ1\n2ܬ݄݊ݐܿ൬ݖ\n2ܬ൰ \nwith z = gJBH/kBT, N the number of spins, g the g-factor, J the angular momentum quantum \nnumber, B the Bohr magneton, H the magnetic field, kB the Boltzman constant and T the \ntemperature. The best fit was obtained for J = 1/2 with N and g as fitting parameters leading \nto N = (3.7 ± 0.3)·1014 spins and g = 2.7 ± 0.2. An enhancement of the g-factor of the buffer \nlayer was already shown by Maassen et al. [S8] but has also been observed in graphene [S9]. \nThe extracted number of spins N on our 4 mm × 4 mm sample is equivalent to 1.2 B per \nhexagon area and can originate from the Silicon dangling bonds [S10] but also from carbon \nvacancies [S11] and other defect structures present in the unscreened buffer layer. \n \nReferences supplementary information \nS1. Emtsev, K. V. et al., Nature Mat. 8, 203 (2009). \nS2. Elias, D. C. et al., Science 323, 610-613 (2009). \nS3. Ferrari, A. C., and Robertson, J., Phys. Rev. B 61, 14095 (2000). \nS4.Eckmann, A. et al., Nano Lett. 12, 3925 (2012). \nS5. Emtsev, K. V., Speck, F., Seyller, Th., Ley, L., Riley, J. D., Phys. Rev. B 77, 155303 \n(2008). \nS6. Haberer, D. et al., Adv. Mater. 23, 4497 (2011). \nS7. Luo, Z. et al., ASC Nano 3, 1781 (2009). \nS8. Maassen, T. et al., Localized States Influence Spin Transport in Epitaxial Graphene, Phys. Rev. Lett. 110, 067209 (2013). \nS9. Kurganova, E. V. et al., Spin splitting in graphene studied by means of tilted magnetic-\nfield experiments, Phys. Rev. B 84, 121407(R) (2011). \nS10. Anisimov, V. I. et al., SiC(0001): A surface Mott-Hubbard insulator, Phys Rev. B 61, \n1752 (2000). \nS11. Yazyef, O. V., Emergence of magnetism in graphene materials and nanostructures, Rep. \nProg. Phys. 73, 056501 (2010). \n 19\n \n 20\nSupplementary figures \n \n \nFigure S1| Number of graphene layers. Raman map image of the 2D-peak Area (a) and \nPosition (b) showing the alternating single (red) and bilayer (yellow) areas. (c) Full Raman \nspectra at the red and blue locations indicated in (a) and (b). In blue the typical spectra for a \nsingle layer graphene and in red for the bila yer graphene. (d) LEEM image showing a clear \ncontrast difference between the single layer graphene in the middle of a terraces and bilayer \ngraphene on the terrace edges. (d) LEEM intens ity as a function of energy on the positions as \nindicated in (d). The number of dips corres ponds to the number of layers, corroborating the \nalternating single and bilayer areas with incidental multilayer patches. \n \n21\n\nFigure S2| X-ray photoemission spectra for hydrogenated graphene. (a) spectra for an \nuntreated epitaxial graphene samples with the S1 (magenta) and S2 (cyan) peaks originating from the bufferlayer, the SiC (red) peak from the substrate and the CC (green) peak from the \nsp\n2 graphene carbon atoms. (b) spectrum with a hydrogen treatment of 30 min showing the \nupcomming CH (brown) peak of sp3 carbon-hydrogen bonds. (c) Same spectra after 120 min \nhydrogen treatment. 288 286 284 282\nCHSiC CC\nS1\ncb\n120 min30 minI (A.U.)0 mina\nS2 I (A.U.) I (A.U.)\nB.E. (eV)22\n \nFigure S3| Electric Force Microscopy of hydrogenated graphene. (a) Atomic force \nmicroscope height image taken with a Pt/Ir conductive tip. (b) Electrical force image after the \nsample was placed on the south pole of a permanent magnet. (c) Electrical force image after \nthe sample was placed on the north pole of a permanent magnet. (d) Cross section of the \nimages in (b) and (c) after respective negativ e (-B) and positive (+B) magnetization showing \nno magnetic field response. \n \n+B1 mSL DL\n-BSL DL\n1 m\nSL DL\n1 m\n024-6.5-6.0-5.5-5.0-BEFM signal (deg) \nx (m)+Ba c b d23\n \nFigure S4| Paramagnetism in bufferlayer on SiC. Magnetization of the buffer layer for 2.5 \nK and 4 K. The curves overlap as a function of H/T after subtraction of the diamagnetic \nbackground and show a clear paramagnetic signal saturating at high magnetic fields. The red \nline is a fit to the Brillouin function for J = 1/2. \n 01 0 2 00153045\n 2.5 K\n 4 K \n Brillouin fitM (10-7 emu)\nH/T (kOe/K)" }, { "title": "1401.0878v1.Hybrid_paramagnetic_ferromagnetic_quantum_computer_design_based_on_electron_spin_arrays_and_a_ferromagnetic_nanostripe.pdf", "content": "arXiv:1401.0878v1 [quant-ph] 5 Jan 2014Hybrid paramagnetic-ferromagnetic\nquantumcomputerdesignbased on\nelectron spin arrays anda ferromagneticnanostripe\nJérômeTRIBOLLET∗\nInstitutdeChimiedeStrasbourg,StrasbourgUniversity,U MR 7177(CNRS-UDS),\n4 rueBlaisePascal,CS 90032,F-67081StrasbourgCedex, Fra nce\nE-mail: tribollet@unistra.fr\nAbstract\nDesigning an assembly of quantum nano-objects which can int eract between themselves\nand be manipulated by external fields, while staying isolate d from their noisy environment is\nthe key for the development of future quantum technologies, such as quantum computers and\nsensors. Electron spins placed in a magnetic field gradient, interacting by dipolar magnetic\ncouplings and manipulated by microwave pulses represent a p ossible architecture for a quan-\ntum computer. Here, a general design for the practical imple mentation of such nanodevice is\npresented and illustrated on the example of electron spins i n silicon carbide placed nearby a\npermalloyferromagnetic nanostripe. Firstly,theconfined spinwaveresonance spectrum ofthe\nnanostripe and the properties of its magnetic field gradient are calculated. Then, one shows\nhow to avoid microwave driven electron spin decoherence. Fi nally, one shows that decoher-\nence due to ferromagnetic fluctuations is negligible below r oom temperature for spins placed\nfar enough from the nanostripe.\n∗To whomcorrespondenceshouldbeaddressed\n1Introduction\nDuring last decades, the development of nanotechnologies h as allowed the experimental demon-\nstration of entanglement between the quantum states of diff erent pairs of nanosystems,1–4a key\npreliminarysteptowardsquantuminformationprocessing. However,upscalingthoseexperiments\nto at least ten coupled quantum systems, called quantum bits or qubits, on a single chip remains\nextremely challenging. This difficulty is linked to the need to couplequantum bitsovernanoscale\ndistancestoperformconditionallogicaloperations,exce ptwhensomealternativestrategyforlong\ndistance entanglement is available.5,6In the context of quantum bits encoded on electron spins in\nsolids, the easiest solution to couple them is to use the dipo lar magnetic coupling,3,4,7which has\nhowevertwodrawbacks. Firstly,itbecomestooweakbeyondt ennanometersforefficientquantum\nstates entanglement. Secondly, it is a permanent coupling, which requires methods to effectively\nswitch on/off those couplings. The up scaling problem is thu s directly related in this context to\nthe difficulty to precisely positiontheelectron spins on th e chip and also to manipulatethem with\nnanoscale resolution. State of art nanolithography and ion implantation methods used to build\nsilicon8or diamond3,4,9based small scale quantum devices with few coupled electron spins are\nstill far from reaching the ten qubits scale. Building solid state arrays of dipolar coupled electron\nspin qubits thus requieres an alternative nanotechnology a nd also a method and a nanodevice al-\nlowing to switch on/offthosedipolar couplings. Transmiss ionElectron Microscope(TEM) based\nnanofabrication methods with nanometer scale precision we re recently demonstrated with carbon\nnanotubes10andgraphenematerials,11andcouldprovidetherequiredalternativenanotechnology .\nAlso, few years ago, it was shown that selectiveon/off switc hing of the dipolar coupling is possi-\nble using sequences of microwave pulses applied to dipolar c oupled electron spins submitted to a\nmagnetic field gradient.7The problem was thus shifted to the production of strong magn etic field\ngradient. Inthefieldofquantumcomputingwithnuclearspin s,itwasproposedtousedysprosium\nnanoferromagnet toproduce ahugemagneticfield gradient.12However,in thecontextofelectron\nspin qubits, the nanoferromagnet also produces unwanted ad ditional decoherence of the electron\nspins13dueeithertoincoherentthermalexcitationortocoherentm icrowaveexcitationofthespin\n2waves14confined in the nanoferromagnet.15,16Here I present a quantum register design contain-\ning a permalloy ferromagnetic nanostripe and nearby arrays of electron spins in silicon carbide\n(SiC). The spin qubits are silicon vacancies17,18,20,21,23that could be created in SiC using TEM\nnanofabricationmethods.10,11,18,22,23Usingmagnetostatictheory,12theLandauLifschitzequation\nof motion of magnetization in the nanoferromagnet,16a one dimensional Schrodinger equation\nsolved numerically by a transfer matrix method,24and the density matrix theory of spin qubit de-\ncoherence,13Ishowhowtodesignsimultaneouslythemagneticfieldgradie ntandtheconfinedspin\nwaves spectrum of the nanoferromagnet in order to avoid addi tional decoherence. The designed\nquantum register could be operated with microwave pulses38and read out with high sensitivity\nusing ensemble optical measurements of the paramagnetic re sonance of silicon vacancy electron\nspinsinSiC.17,18\nDesignandfabricationofananodevicewithelectronspinar rays\nnearbya ferromagneticnanostripe\nThegeneral designproposedhereforquantuminformationpr ocessingcouldbeapplied,inprinci-\nple,toallkindofelectronspinqubitslikephosphorousdon orsinsilicon,8,34nitrogendonorsorNV\ncenters in diamonds,3,4,9transition metals ions or shallow donors in Zinc Oxide,13,25–28N@C60\nmolecules,29paramagnetic defects in graphene,33paramagnetic dangling bonds arrays on hydro-\ngenpassivatedsiliconsurface,31orsupramolecularassembliesofparamagneticmoleculessu chas\nneutral radical molecules30or copper phtalocyanine molecules,32as long as regular dense arrays\nof them could be produced at precise positions nearby a ferro magnetic nanostripe. Without loss\nof generality and to be more quantitative, I focus here on a na nodevice made with silicon carbide\nand permalloy materials and build with top-down methods. Th e successive steps of two possible\nnanofabrication processes allowing to build such a nanodev ice are described on figure 1a and 1b.\n3The resulting nanodevice is shown on figure 1c. Each of the thr ee SiC polytypes, 4H, 6H, and\n/MT97/MT47\n/MT49/MT47 /MT50/MT47 /MT51/MT47 /MT52/MT47 /MT53/MT47 /MT54/MT47\n/MT99/MT47\n/MT120/MT121\n/MT122\n/MT87\n/MT84/MT100/MT87/MT47/MT50\n/MT79/MT76/MT47/MT50/MT66/MT48/MT122\n/MT66/MT49/MT120/MT40/MT116/MT41/MT80/MT121\n/MT54/MT72/MT32/MT45/MT32/MT83/MT105/MT67\n/MT32/MT119/MT97/MT102/MT101/MT114/MT54/MT72/MT45/MT83/MT105/MT67/MT98/MT47\n/MT49/MT47 /MT50/MT47 /MT51/MT47 /MT53/MT47 /MT54/MT47 /MT52/MT47\nFigure 1: a/ First process: 6H-SiC wafer / step in wafer produ ced by focused ion beam (FIB)\nmilling/ epitaxial growth of isotopically purified (withou t nuclear spins) 6H-SiC on patterned\nwafer/ nanotrench created by focused ion beam (FIB) milling method35nearby the edge of the\nstep: depth W, widthT, lengthL/ ion assistedtrench filling methods36used to fill thenanotrench\nwith Permalloy / TEM production10,18,22,23,32of identical and parallel linear arrays of silicon va-\ncancies (white dots) aligned along the x axis. b/ Second proc ess: it combines electron beam\nlithography and permalloy plating:37first steps as in 1a and then: 1b4: resist deposition and pla-\nnarization,1b5: nanotrenchproducedinresistbyebeamlit hography,furtherfilledwithPermalloy\n(red), 1b6: removal of the resist, and then TEM production of silicon vacancies. c/ Design of the\nproposed SiC-Py quantum register corresponding to nanofab rication method a/. In the following\ncalculations: L=100µm,T=100nm,W=800nm.\n3C, has some paramagnetic defects related to silicon vacanc ies which could potentially encode a\ngood electron spin quantum bit.17,18,20,21,23Some silicon vacancies defects in 6H-SiC has a spin\nS=3/2 with an isotropic g factor g=2.0032 and a small zero field splitting.18Their spin state\ncan also be prepared by optical alignment18and it can be readout by the highly sensitive ODMR\nmethod(opticallydetectedmagneticresonance).19Itseemsthushighlyrelevanttoencodeanelec-\n4tronspinqubitonthiskindofsiliconvacancydefectin6H-S iC(forexampleatXbandorathigher\nmicrowave frequency, using:/vextendsingle/vextendsingleS=3\n2,MS=−3\n2/angbracketrightbig\n≡ |0/angbracketrightand/vextendsingle/vextendsingleS=3\n2,MS=−1\n2/angbracketrightbig\n≡ |1/angbracketright). In the\nfollowing calculations, one simplifies the spin qubit model by using a simple pseudo spin s=1/2\nwithag factorg=2, withoutanylossofgenerality.\nDipolarmagneticfieldgradientoutsidethenanoferromagne tand\neffectiveIsingcouplingbetweenelectronspinqubits\nThepracticalimplementationofsingleandtwoqubitsquant umgatesonchainsofdipolarcoupled\nelectron spin qubits, as those shown on figure 1, requires app ropriate microwave pulse sequences\nand a strong magnetic field gradient along the direction of th e spin chains of the nanodevice to\ntransform their dipolar coupling into an effectiveIsing li ke coupling.7As here I also consider the\npossibility to perform ensemble measurements of the spin st ates of the spin qubits, for example\nusing the highly sensitiveODMR method,19the magnetic field gradient produced by the permal-\nloy ferromagnetic stripe shown on figure 1 has not only to be st rong enough along x, but also\nto be enough homogeneous along y and z. Strong enough along x m eans here that the energy\ndifference between one qubit j (at position xj) and the next one j+1 (at position xj+1) along the\nspinchain has to bemuch larger thantheenergy associated to thedipolarcouplingbetween them,\nwhich depends on their relative distance linter. In the nanodevice described on figure 1, linteris\nthus also the periodicity of each spin chain. The condition f or an effective Ising coupling trans-\nlates into/vextendsingle/vextendsinglegµB/parenleftbig\nB0+Bdip,z/parenleftbig\nxj+1/parenrightbig/parenrightbig\n−gµB/parenleftbig\nB0+Bdip,z/parenleftbig\nxj/parenrightbig/parenrightbig/vextendsingle/vextendsingle>>µ0µ2\nBg2\n4πl3\ninter, or approximately,\n/vextendsingle/vextendsingle/vextendsinglegµBdBdip,z(x)\ndxlinter/vextendsingle/vextendsingle/vextendsingle>>µ0µ2\nBg2\n4πl3\ninter, whereBdip,z(x)is the value of the z component of the dipolar\nmagnetic field produced by the ferromagnetic nanostripe at p osition x, assuming that z≈0 for\neach spinqubit,and thatthey coordinateof each spinqubitd oes not matterforthevalueof Bdip,z\napplied to the qubit. The irrelevance of the y coordinate is d ue to the approximate invariance by\ntranslation of the gradient produced by the long ferromagne tic nanostripe (infinite stripe model).\nThis approximation is valid as long as the spin chains are cre ated far from the edges of the ferro-\n5magnetic stripe along the y direction. Assuming a permalloy ferromagnetic stripe with a length\nL=100µmalongtheyaxis,awidth T=100nmalongthexaxis,andadepth W=800nmalong\nthe z axis, further assuming that the ferromagnetic stripe i s fully magnetized along the z axis by\nthe staticapplied magneticfield B0, one can calculate using magnetostatictheory12the properties\nofthedipolarmagneticfield producedby thisnanoferromagn et,as shownonfigure 2.\n−1000 −500 0 500 1000−500005000\nz (nm)Bz (z) (G)\n \n−800 −400 0 400 800−4000−200002000\nx (nm)C(x) (G.nm)\n−1000 −600 −200 200 600 1000−1000−5000100\nx (nm)Bz(x) (G)\n−800 −400 0 400 800−2−1012\nx (nm)dBz(x)/dx (G/nm)z= 0 nm z= 0 nmfull: x = 0 nm\ndash: x= 230 nm\nc/ d/b/ a/\nFigure 2: a/ Bdip,z(z,x)for two different x position: x=0 (continuous black) and x=xoptim=\n230nm(dash red). b/ C(x) =/integraltext+100\n−100dz/parenleftbig\nBdip,z(x,z)−Bdip,z(x,0)/parenrightbig\n. One defines the position\nxoptimbyC(xoptim) =0. For the permalloy ferromagnetic nanostripe considered h ere (see figure\n1), one finds xoptim=230nm. This position xoptimis the position where the homogeneity of\nBdip,z(x,z)is optimalalong thezdirection. c/ Bdip,z(z=0,x). d/ Gradient along x of thedipolar\nmagnetic field produced by the ferromagnetic nanostripe, as suming z=0 and y=0:dBdip,z(x)\ndx. For\ntheferromagneticPermalloynanostripe,thesaturationma gnetizationis Msat,Py,withµ0Msat,Py=\nBsat,Py=11300G.\nFigure2showsthataslongasthespinqubitsencodedontosil iconvacanciesinSiCarecreated\nfar from the edges of the ferromagnetic stripe along the y dir ection and close enough to the z=0\nplane, then the relevant z component of the dipolar magnetic field produced by the ferromagnetic\nnanostripe can be considered as only dependent on x, and thus its gradient can also be approxi-\nmatelyconsideredasbeingonedimensionalalongx. Onefinds hereBdip,z(xoptim=230nm,0)≈\n−676GanddBdip,z(xoptim=230nm,0)\ndx≈1.44G/nm. From those results, one can thus estimate\n6whether thecondition givenaboveto obtain an effectiveIsi ng coupling between the spin qubits is\nsatisfied. Assumingthat most of the spin qubits are created i n SiC by TEM at x positionscloseto\nxoptim=230nmand z=0, and further assuming a distance linter=5nmbetween successive spin\nqubits along the x direction, one finds that/vextendsingle/vextendsingle/vextendsinglegµBdBdip,z(x)\ndxlinter/vextendsingle/vextendsingle/vextendsingleis more than hundred times larger\nthan thedipolarcouplingenergyµ0µ2\nBg2\n4πl3\ninter. One could thus,at first sight,concludethat theproposed\nnanodevice described on figure 1, satisfies the requirements for the practical implementation of\nquantum information processing with dipolar coupled elect ron spin qubits manipulated by appro-\npriatesequencesofmicrowavepulses.7However,inordertoavoidmicrowavedrivendecoherence,\nitisthusabsolutelynecessary todesigntheferromagnetic nanostripesuchthatno spectraloverlap\nbetween the precession frequencies of the spin qubits and th e precession frequencies of the spin\nwavesconfined insidetheferromagneticnanostripe14–16occurs.\nDesigning the field sweep confined spin wave resonance spec-\ntrumtoavoidcoherentlydrivenelectronspinqubitdecoher ence\nOneassumeshere thatthepermalloyferromagneticnanostri peis almostfullymagnetizedat equi-\nlibrium along z at sufficiently low temperature and sufficien tly high magnetic field (see figure 1).\nThe excitation of a ferromagnetic spin wave implies that the magnetization inside the ferromag-\nnetic nanostripe is no more uniform. In other words that mean s that the dipolar magnetic field\nacting on the nearby electron spin qubits outside the nanofe rromagnet will become fluctuating,\nleadingto spinqubitdecoherence.13,38As itcan beseen on figure 2a(in thecase x=0)and on fig-\nure 3a, all the spinscontained insidethe ferromagneticnan ostripeproduce a very inhomogeneous\ndipolar magnetic field inside the ferromagnetic nanostripe itself. This leads to the confinement of\nthe ferromagnetic spin waves15along the z axis, mainly on the edges of the ferromagnetic nan os-\ntripe. Here I have developed a method to calculate the confine d spin wave resonance spectrum of\na long ferromagnetic nanostripe fully magnetized. First, t he Landau Lifschitz equation of motion\nofmagnetizationinthenanoferromagnet16islinearized. Then,theonedimensionaleigenenergies\n70 200 400 600 800−80−60−40−20020406080\nz* (nm)1D eigenenergie (micro−eV)\n0 200 400 600 800−5−4−3−2−101x 105\nz* (nm)1D wavefunction amplitude (a.u.)a/ b/\nFigure3: Theonedimensionaleigenenergiesandeigenfunct ionsofthespinwavesconfinedinthe\ninhomogeneouseffectivepotentialexistinginsidethefer romagneticnanostripealongthedirection\nz of the static magnetic field applied: a/ one dimensional eig enenergies of the spin waves along\nz axis (horizontal lines) represented on top of the inhomoge neous effective confining potential\ninside the nanostripe; b/ one dimensional eigenfunctions o f the spin waves confined along z (the\nout of equilibrium magnetization component δmx(z)has been plotted here). Here the permalloy\nferromagnetic nanostripe has the dimensions: length L=100µmalong the y axis, width T=\n100nmalong thex axis, and depth W=800nmalong thez axis. It was furtherassumed that the\nferromagneticstripeisfullymagnetizedalongthezaxisby thestaticmagneticfieldapplied B0,and\nthe magnetization of Permalloy at saturation was taken equa l toµ0Msat,Py=Bsat,Py=11300G.\nNote also that for numerical calculations, the new variable z∗was used, defined by z∗=z+450\n(in nm).\nand eigenfunctions of the spin waves confined in the inhomoge neous effective potential existing\ninside the ferromagnetic nanostripe along z are calculated , as it is shown respectively on figure\n3a and figure 3b, using an effective Schrodinger equation sol ved numerically by a transfer matrix\nmethod.24Then, the three dimensional eigenenergies of the confined sp in waves are determined,\nand finally the resonance fields at which the microwave magnet ic field of fixed frequency can ex-\nciteaconfinedspinwavearedetermined,simplybywritingth eenergyconservationrule. Thisrule\nsays that the energy of the microwave photon absorbed should be equal to the three dimensional\neigenenergie of the confined spin wave excited. It is importa nt to note that the combinationof the\ntheoretical and of the experimental field sweep microwaveab sorption spectrum, providesan indi-\n8rect method to determine the dipolar magnetic field produced by the ferromagnetic nanostripe on\nthespinqubits,becauseitisthesamemagnetizationdistri butionthatproducesthedipolarmagnetic\nfield insideand outsidethenanostripe. The full field sweep s pectrum expected forthe nanodevice\ndescribed on figure 1 is plottedon figure 4a. It includes thefie ld sweep spectrum expected for the\n0.80.911.11.21.31.4\nx 10400.20.40.60.811.21.41.61.82\nBo (Gauss)mw abs (a.u.)\n1.25 1.26 1.27 1.28 1.29\nx 10400.20.40.60.811.21.41.61.82\nBo (Gauss)mw abs (a.u.)a/ b/\nB\nCBA\nC\nFigure 4: a/ Full field sweep microwave absorption spectrum o f the nanodevice designed on\nfigure1,includingthespinwavesresonancespectrum(bluer esonancelines)andtheelectron spin\nqubits paramagnetic resonance spectrum (red resonance lin es), as it could be measured at a fixed\nmicrowave frequency ν=34GHzand with a varying magnetic field (see text for more details).\nb/ Zoom in the high magnetic field range of the field sweep micro wave absorption spectrum of\nthe proposed nanodevice, showing that no spectral overlap b etween the ferromagnetic spin waves\nand the electron spin qubits occur with this design (only red resonance lines are present in this\nfield range). Note that the linewidth and the strength of the r esonances (oscillator strength) were\narbitrarily chosen here, that means, without taking into ac count the relative symmetry of the spin\nwavemodeandofthemicrowavephotonmodeinsidethemicrowa vecavity,whichdeterminesthe\noscillatorstrength,inordertoseeanyspinwaveresonance satisfyingtheenergyconservationrule\nalone.\nferromagnetic nanostripe (blue resonance lines with reson ance magneticfields occurring between\n0.80and1.40Tesla). Italsoincludesthefieldsweepspectru mexpectedfor16electronspinqubits\ncreated at positions x close to xoptim=230nm, with x values comprised between x=199nm\nandx=274nm(ensemble C of 16 red resonance lines with resonance magneti c fields occurring\naround 1.28 Tesla). In order to see more clearly that those el ectron spin qubits resonance lines\n9(C) are shifted towards high magnetic field values by the dipo lar magnetic field Bdip,z(x,0)pro-\nducedbytheferromagneticnanostripeandpreviouslycalcu lated,Ialsoaddedonfigure4athefield\nsweep spectrumexpected foran electron spinqubitcreated a t infinitedistancefrom theferromag-\nnetic stripe along the x axis (single A red resonance line wit h resonance magnetic field occurring\naround 1.215 Tesla) and also the field sweep spectrum expecte d for an electron spin qubit created\nat distance x=500nmfrom the ferromagnetic stripe (single B red resonance line w ith reso-\nnancemagneticfieldoccurringaround1.250Tesla). Themost importantresultshownonfigure4a\nand highlighted by figure 4b, is the possibility to design the electron spin based quantum register\nproposed here such that no spectral overlap occur between th e 16 resonances lines of the shifted\nelectron spin qubits and the resonance lines of the two confin ed spin waves of the ferromagnetic\nnanostripeoccurring at thehighest magneticfield values,t he so called edgespin wavemodes, oc-\ncurring here at Q band, at 1.227 Tesla and at 1.362 Tesla. This provides a magnetic field interval\nof1350 G free of anyspin waveresonance. Further assumingan inhomogeneouslinewidthbelow\n1 G for each silicon vacancy electron spin qubit in nuclear sp in free isotopicallypurified SiC, one\ncouldhopetobuildidenticalparallelarrays of675electro nspinqubitsalongthexaxis. However,\nstate of art pulse EPR spectrometers operating at Q band have a resonator bandwidth of around\n1 GHz, corresponding here to a magnetic field interval of arou nd 350 G, in which only around\n175 electron spin qubits could be manipulated using selecti ve microwave pulses.38It seems thus\npossible in principle to build a model quantum register of th is kind having tens of electrons spins\nqubitswith paramagneticresonance lines not overlappingt heconfined spin waveresonance lines,\nthusavoidinganycoherentlydrivenspindecoherence.\nInvestigation of the spin qubit decoherence process due to i nco-\nherentmagneticfluctuationsofthenearbynanoferromagnet\nTillnow,Ihavenotconsideredthefactthattheincoherentt hermalexcitationofferromagneticspin\nwaves can produce some time dependent fluctuating dipolar ma gnetic field outside the ferromag-\n10netic nanostripe, at the positions where the electron spin q ubits are created. Those fluctuations\nalways exist in a ferromagnetic nanostripe at equilibrium a t a non zero temperature T. As it is\nalso known from the standard density matrix theory of electr on spin decoherence,13any fluctuat-\ning magnetic field lead to electron spin decoherence. Thus he re, one expects that the fluctuating\ndipolarmagneticfieldactingonelectronspinqubitswillpr oduceanewspindecoherenceprocess,\nwhich must be added to the other electron spin decoherence pr ocesses, which are intrinsic to the\nsiliconvacanciesinthesiliconcarbidematrixalone. Thea nalysisofthisnewdecoherenceprocess\nrequires the knowledge of the quantum correlation function s of this time dependent fluctuating\ndipolar magnetic field outside the ferromagnetic nanostrip e.13This is a complicated problem not\naddressed to date. Here I present the results of a theoretica l study of this problem based on the\nformalismsofthesecondquantizationofspinwavesandofth edensitymatrix(detailswillbepub-\nlished elsewhere). The theory developed shows that this new electron spin decoherence process\ndepends mainly, on the saturation magnetization of the nano ferromagnet, on the temperature, on\nthespectraldetuningbetweentheparamagneticresonanceo ftheelectronspinqubitandtheferro-\nmagnetic resonance of the nanostripe, and also on the distan cex of the qubit to the ferromagnetic\nnanostripe, as long as the qubit has a position assumed far fr om the edges of the ferromagnetic\nnanostripealongthey andz axis(y«L,z«W and x>T/2). Result softhecalculation oftheelectron\nspin coherence time T2and of the electron spin longitudinalrelaxation time T1of an electron spin\nqubitofthenanodeviceshownonfigure1andoperatedatmicro waveQbandarepresentedonfig-\nure 5. At helium temperature, at microwaveQ band, and with an applied magneticfield ofaround\n1.28 Tesla, one finds: T1(x=230nm,2K) =3.4sandT2(x=230nm,2K) =6.4s. In those\nexperimental conditions, this new decoherence process for the electron spin qubitsshould thus be\nnegligible compared to other decoherence processes intrin sic to natural SiC. However, it could\nalso become the dominant decoherence process for electron s pin qubits embedded in a nuclear\nspinfreeisotopicallypurifiedSiCmatrix,assumingintrin sicspincoherencetimeinSiCsimilarto\nthosefoundinSi,intherangeoffewseconds.34Thissuggestthatquantuminformationprocessing\nwith tens of electrons spins qubits encoded onto silicon vac ancies in SiC, coherently manipulated\n1120025030035040045050055010−210−1100101102103\n x (nm)T1 or T2 (s)\n05010015020025030010−1100101\nT (K)T1 or T2 (s)\n \n30 K\n300 K3 Kb/ a/\nx = 230 nm\nFigure 5: Effects of the thermal fluctuations of the dipolar m agnetic field produced by the ferro-\nmagnetic nanostripe on the electron spin coherence time T2 a nd on the electron spin longitudinal\nrelaxationtimeT1ofanelectronspinqubitofthenanodevic eshownonfigure1. a/T2(thinlines)\nand T1 (thick lines), as a function of the qubit position x for three different temperatures T= 3 K,\n30 K, and 300 K. b/ T2 (squarre) and T1 (cross), as a function of the temperature of the nanode-\nvice, assumingan electron spin qubit position x=xoptim=230nm, where its resonant magnetic\nfield at Q band (34 GHz) is around 1.28 Tesla, and also assuming y=z=0. Permalloy nanostripe\ndimensions considered: length L=100µmalong the y axis, width T=100nmalong the x\naxis, and depth W=800nmalong the z axis, as shown on figure 1. The estimated T2 and T1\nare valid as long as the shifted spin qubit resonance line do n ot overlap any confined spin wave\nresonanceline,asitisrequieredtousethenanodeviceshow nonfigure1. Itwasalsoassumedthat\ntheferromagneticresonance linewidthofthepermalloynan ostripeisequal to3 µeVatQ band.\nby microwave pulse and submitted to the strong dipolar magne tic field gradient of the permalloy\nferromagneticnanostripeshouldbepossible,atleastatcr yogenictemperatures. Also,atroomtem-\nperature one finds T1(x=230nm,300K) =25msandT2(x=230nm,300K) =47ms. Thus,\nthosetheoreticalestimates,combinedwiththerecentobse rvationofroomtemperaturespincoher-\nenceofsomesiliconvacanciesinSiCover80 µs,18suggestthataSiC-Permalloyquantumregister\nsuch as the one designed here should keep its quantum coheren ce properties at room temperature\nat a sufficiently high level to perform fundamental studies o f quantum entanglement over tens of\ndipolarcoupledelectrons spinsqubitsofsiliconvacancie sat roomtemperature.\n12Discussion\nSiC is a very attractive material for the practical implemen tation with a top-down approach of\nthe general quantum register design proposed here, firstly d ue to the long intrinsic electron spin\ncoherence time of its paramagnetic defects18similar to the one in silicon,34secondly due to the\npossibilitytocreatesiliconvacanciesarraysinSiCusing advancedTEMmethods,andthirdlydue\nto the possibilities for optical initialization and optica l readout of the spin state of some silicon\nvacancies in SiC.17,18Note that a dysprosium ferromagnetic nanostripe would prod uce a three\ntimes bigger magnetic field gradient than in the case of perma lloy. However, one expects with\ndysprosium an electron spin decoherence rate at least one or der of magnitude larger than in the\ncase of permalloy at cryogenic temperature due to its three t imes larger saturation magnetization\nand its muchlarger ferromagneticresonance linewidth. In a ddition,Dysprosiumisno moreferro-\nmagnetic at room temperature, contrary to permalloy, which suppresses any possibility for room\ntemperature quantum information processing. Thus the hybr id paramagnetic-ferromagnetic nan-\nodevicepresentedhere,basedontheSiCmaterialwidelyuse dforhightemperatureelectronicsand\non the permalloy material widely used for magnetic elements in computers , seems to be an ex-\ncellent choice for quantum information processing with dip olar coupled electron spins selectively\nand coherentlymanipulatedbyresonant microwavepulses.\nReferences\n(1) R. Blatt et al.,Nature 453,1008 (2008).\n(2) M.Neeley et al.,Nature 467,570 (2010).\n(3) F. Doldeet al., Naturephysics 9, 139(2013).\n(4) H.Bernien et al.,Naturephysics 6,249 (2010).\n(5) H.Bernien et al.,Nature 497, 86(2013).\n(6) G.Burkard et al.,Phys.Rev. B 74, 041307R(2006).\n13(7) R. deSousaet al.,Phys.Rev. A 70,52304(2004).\n(8) R. Vrijenet al., Phys.Rev.A 62, 012306(2000).\n(9) D.M.Toyliet al.,Nano Lett. 10, 3168(2010).\n(10) J.A.Rodriguez-Manzo et al.,Nano Lett. 9, 2285(2009).\n(11) A.W.Robertson et al.,Nano Lett. 13, 1468(2013).\n(12) J.R. Goldmanet al., Appl.Phys.A 71,11 (2000).\n(13) J.Tribollet,Eur. Phys.J.B 72,531(2009).\n(14) C. Kittel,Phys.Rev. 110,1295 (1958).\n(15) J.Jorzick et al.,Phys. Rev.Lett. 88, 47204(2002).\n(16) I. Leeet al.,Nature 466, 845(2010).\n(17) A.L.Falk et al.,NatureCommunications 4, 1819(2013).\n(18) D.Riedel etal., Phys.Rev.Lett. 108,226402(2012).\n(19) H.Kraus et al., NaturePhysics 2826,1(2013).\n(20) H.Itoh et al.,IEEE TransactionsonNuclearScience 37, 1732(1997).\n(21) T.Wimbaueretal., Phys.Rev.B 56,7384(1997).\n(22) J.W.Steeds et al., DiamondRelat. Mater. 11,1923(2002).\n(23) J.Lefevreet al., J. ofAppl.Phys. 105,23520(2009).\n(24) C. Jirauschek,IEEE J. ofQuantumElectronics 45, 1059(2009).\n(25) J.Tribolletet al., EurophysicsLetters 84, 20009(2008).\n(26) S.T.Ochsenbein etal., NatureNanotechnology 6,112(2011).\n14(27) R.E. Georgeet al., Phys.Rev.Lett. 110, 27601(2013).\n(28) K.Benzid etal., EurophysicsLetters 104,47005(2013).\n(29) W.Harneit, Phys.Rev.A 65, 32322(2002).\n(30) A.Tamuliset al.,Chem. Phys.Lett. 436,144(2007).\n(31) S.R. Schofield etal., NatureCommunications 4,1649, 1(2013).\n(32) M.Warner et al.,Nature 503,504 (2013).\n(33) R.R. Nairet al., Naturephysics 8, 199(2012).\n(34) A.M.Tyryshkinet al., Naturematerials 11, 143(2012).\n(35) R. Menzel et al.,Appl.Surf. Science 136, 1(1998).\n(36) A.Cheryl, ElectronicDesign 47,38 (1999).\n(37) X.Yang et al., J.Vac. Sci. Technol.B 21, 3017(2003).\n(38) A. Schweiger and G. Jeschke, Principles of pulse electron paramagneticresonance (Oxford\nUniversityPress, Oxford, UK ;New York,2001).\nCompetingfinancialinterests\nTheauthordeclare thathehas nocompetingfinancial interes ts.\n15" }, { "title": "0704.3975v3.Spectroscopy_and_critical_temperature_of_diffusive_superconducting_ferromagnetic_hybrid_structures_with_spin_active_interfaces.pdf", "content": "arXiv:0704.3975v3 [cond-mat.supr-con] 12 Nov 2007Spectroscopy and critical temperature of diffusive superco nducting/ferromagnetic\nhybrid structures with spin-active interfaces\nAudrey Cottet\nLaboratoire de Physique Th´ eorique et Hautes ´Energies, Universit´ es Paris 6 et 7,\nCNRS, UMR 7589, 4 place Jussieu, F-75252 Paris Cedex 05, Fran ce and\nLaboratoire de Physique des Solides, Universit´ e Paris-Su d,\nCNRS, UMR 8502, F-91405 Orsay Cedex, France\n(Dated: July 3, 2021)\nThedescriptionoftheproximityeffectinsuperconducting/ ferromagnetic heterostructuresrequires\nto use spin-dependent boundary conditions. Such boundary c onditions must take into account the\nspin dependence of the phase shifts acquired by electrons up on scattering on the boundaries of fer-\nromagnets. The present article shows that this property can strongly affect the critical temperature\nand the energy dependence of the density of states of diffusiv e heterostructures. These effects should\nallow a better caracterisation of diffusive superconductor /ferromagnet interfaces.\nPACS numbers: 73.23.-b, 74.20.-z, 74.50.+r\nI. INTRODUCTION\nWhen a ferromagnetic metal ( F) with uniform mag-\nnetization is connected to a BCS superconductor ( S),\nthe singlet electronic correlations characteristic of the\nSphase can propagate into Fbecause electrons and\nholes with opposite spins and excitation energiesare cou-\npled coherently by Andreev reflections occurring at the\nS/Finterface. Remarkably, the ferromagnetic exchange\nfield induces an energy shift between the coupled elec-\ntrons and holes, which leads to spatial oscillations of the\nsuperconducting order parameter in F1,2. This effect\nhas been observed experimentally through oscillations\nof the density of states (DOS) in Fwith the thickness\nofF3, or oscillations of the critical current I0through\nS/F/Sstructures4,5,6,7, with the thickness of For the\ntemperature. The oscillations of I0have allowed to ob-\ntainπ-junctions8, i.e. Josephson junctions with I0<0,\nwhich could be useful in the field of superconducting\ncircuits9,10. A reentrant behavior of the superconducting\ncritical temperature of S/Fbilayers with the thickness\nofFhas also been observed11. At last, some F/S/F\ntrilayers have shown a lower critical temperature for an\nantiparallelalignment ofthe magnetizationsin the two F\nlayers as compared with the parallel alignment12, which\nshould offer the possibility of realizing a superconducting\nspin-switch13,14.\nFor a theoretical understanding of the behavior of\nS/Fhybrid circuits, a proper description of the inter-\nfaces between the different materials is crucial. For a\nlong time, the only boundary conditions available in\nthe diffusive case were spin-independent boundary con-\nditions derived for S/normal metal interfaces15. Re-\ncently, spin-dependentboundaryconditionshavebeenin-\ntroduced for describing hybrid diffusive circuits combin-\ning BCS superconductors, normal metals and ferromag-\nnetic insulators16. These boundary conditions take into\naccount the spin-polarization of the electronic transmis-\nsion probabilities through the interface considered, but\nalso the spin-dependence of the phase shifts acquired byFIG. 1: a. Diffusive F/S/F trilayer consisting of a BCS\nsuperconductor Swith thickness dSplaced between two fer-\nromagnetic electrodes F1andF2with thickness dF. In this\npicture, the directions of the magnetic polarizations in F1and\nF2are parallel [antiparallel], which corresponds to the con-\nfiguration C=P[AP]. b.S/Fbilayer consisting of a BCS\nsuperconductor Swith thickness dS/2 contacted to a ferro-\nmagnetic electrode Fwith thickness dF.\nelectronsupon transmissionorreflectionby the interface.\nThe first property generates widely known magnetoresis-\ntance effects17. The second property is less commonly\ntakeninto account. However, the Spin-Dependence ofIn-\nterfacial Phase Shifts (SDIPS) can modify the behavior\nof many different types of mesoscopic circuits with ferro-\nmagneticelements, likethoseincludingadiffusivenormal\nmetal island18, a resonant system19,20, a Coulomb block-\nade system20,21,22, or a Luttinger liquid23. It has also\nbeen shown that the SDIPS has physical consequences\ninS/Fhybrid systems16,24,25,26. One can note that, in\nsome references, the SDIPS is called ”spin-mixing angle”\nor ”spin-rotation angle” (see e.g. Refs. 24,26). In the\ndiffusive S/Fcase, the spin-dependent boundary condi-\ntions of Ref. 16 have been applied to different circuit\ngeometries27,28,29,30,31but the only comparison to exper-\nimental data has been performed in Ref. 29. The authors2\nof this Ref. have generalized the boundary conditions of\nRef. 16 to the case of metallic S/Finterfaces with a su-\nperconducting proximity effect in F. They have showed\nthat the SDIPS can induce a shift in the oscillations of\nthe critical current of a S/F/SJosephson junction or of\nthe DOS of a S/Fbilayer with the thickness of F. Signa-\ntures of this effect have been identified in the Nb/PdNi\nhybrid structures of Refs. 3,5. Nevertheless, the problem\nof characterizing the SDIPS of diffusive S/Finterfaces\nhas raised little attention so far, in spite of the numerous\nexperiments performed.\nA good characterization of the properties of diffusive\nS/Finterfaces would be necessary for a better control\nof the superconducting proximity effect in diffusive het-\nerostructures. The present article presents other conse-\nquences of the SDIPS than that studied in Ref. 29, which\ncould be useful in this context. In particular, the SDIPS\ncan generate an effective magnetic field in a diffusive Sin\ncontact with a diffusive F, like found for a ballistic Sin\ncontact with a ferromagnetic insulator24. This effective\nfield can be detected, in particular, through the DOS of\nthe diffusive Flayer, with a visibility which depends on\nthe thickness of F. A strong modification of the varia-\ntions of the critical temperature of diffusive S/Fstruc-\ntures with the thickness of Fis also found. These effects\nshould allow to characterize the SDIPS of diffusive S/F\ninterfaces through DOS and critical temperature mea-\nsurements, by using the heterostructures currently fabri-\ncated. The calculations reported in this paper are also\nappropriate to the case of a diffusive Slayer contacted\nto a ferromagnetic insulator ( FI).\nThis paper is organized as follows: Section II presents\nthe initial set of equations used to describe the het-\nerostructures considered. The case of F/S/Ftrilayers\nis mainly addressed, but the case of S/F(orS/FI) bi-\nlayers follows straightforwardly. Section III specializes\nto the case of a weak proximity effect in Fand a super-\nconducting layer with a relatively low thickness dS≤ξS,\nwithξSthe superconducting coherence length in S. The\nspatial evolution of the electronic correlations in the S\nandFlayers is studied in Section III.A. The energy-\ndependent DOS of S/Fheterostructures is calculated in\nSection III.B. Section III.C considers briefly the limit of\nS/FIbilayers. Section III.D discusses SDIPS-induced\neffective field effects in other types of systems. Section\nIII.E compares the present work to other DOS calcu-\nlations for data interpretation in S/Fheterostructures.\nCritical temperatures of S/Fcircuits are calculated and\ndiscussed in Section III.F. Conclusions are presented in\nSection IV. Throughout the paper, I consider conven-\ntional BCS superconductors with a s-wave symmetry.\nII. INITIAL DESCRIPTION OF THE PROBLEM\nThis article mainly considers a diffusive F/S/F tri-\nlayer consisting of a BCS superconductor Sfor−dS/2<\nx < d S/2, and ferromagnetic electrodes F1forx∈{−dS/2−dF,−dS/2}andF2forx∈ {dS/2,dS/2+dF}\n(see Figure 1.a). The magnetic polarization of the two\nferromagnets can be parallel (configuration C=P) or\nanti-parallel (configuration C=AP), but the modu-\nlus|Eex|of the ferromagnetic exchange field is assumed\nto be the same in F1andF2. Throughout the struc-\nture, the normal quasiparticle excitations and the su-\nperconducting condensate of pairs can be characterized\nwith Usadel normal and anomalous Green’s functions\nGn,σ= sgn(ωn)cos(θn,σ) andFn,σ= sin(θn,σ), with\nθn,σ(x) the superconducting pairing angle, which de-\npends on the spin direction σ∈ {↑,↓}, the Matsubara\nfrequency ωn(T) = (2n+1)πkBT, and the coordinate x\n(see e.g. Ref. 32). The Usadel equation describing the\nspatial evolution of θn,σwrites\n/planckover2pi1DS\n2∂2θn,σ\n∂x2=|ωn|sin(θn,σ)−∆(x)cos(θn,σ) (1)\ninSand\n/planckover2pi1DF\n2∂2θn,σ\n∂x2= (|ωn|+iEexσsgn(ωn))sin(θn,σ) (2)\ninF1andF2, withDFthe diffusion constant of the fer-\nromagnets and DSthe diffusion constant of S. The self-\nconsistentsuperconductinggap∆( x) occurringin(1) can\nbe expressed as\n∆(x) =πkBTλ\n2/summationdisplay\nσ∈{↑,↓}\nωn(T)∈{−ΩD,ΩD}sin(θn,σ) (3)\nwith Ω Dthe Debye frequency of S,λ−1=\n2πkBTBCS\nc/summationtext\nωn(TBCS\nc)∈{0,ΩD}ω−1\nnthe BCS coupling\nconstant and TBCS\ncthe bulk transition temperature of\nS. I assume ∆ = 0 in F1andF2. The above equations\nmustbe supplemented with boundaryconditionsdescrib-\ning the interfaces between the different materials. First,\none can use\n∂θn,σ\n∂x/vextendsingle/vextendsingle/vextendsingle/vextendsingle\nx=±(dS/2+dF)= 0 (4)\nfor the external sides of the structure. Secondly, the\nboundary conditions at the S/Finterfaces can be cal-\nculated by assuming that the interface potential locally\ndominates the Hamiltonian, i.e. at a short distance it\ncauses only ordinary scattering (with no particle-hole\nmixing) (see e.g. Ref. 33). This ordinary scattering\ncan be described with transmission and reflection am-\nplitudes tS(F)\nn,σandrS(F)\nn,σfor electrons coming from the\nS(F) side of the interface in channel nwith a spin\ndirection σ. The phases of tS(F)\nn,σandrS(F)\nn,σcan be\nspin-dependent due to the exchange field EexinF1(2)\nand a possible spin-dependence of the barrier poten-\ntial between SandF1(2). Boundary conditions tak-\ning into account this so-called Spin-Dependence of In-\nterfacial Phase Shifts (SDIPS) have been derived for3\n|tS\nn,↑|2,|tS\nn,↓|2≪1 and a weakly polarized F16,29. When\nthere is no SDIPS, the boundary conditions involve the\ntunnel conductance GT=GQ/summationtext\nnTnand the magne-\ntoconductance GMR=GQ/summationtext\nn(|tS\nn,↑|2− |tS\nn,↓|2), with\n↑(↓) the majority(minority) spin direction in the Felec-\ntrode considered, GQ=e2/h, andTn=|tS\nn,↑|2+|tS\nn,↓|2.\nIn the case of a finite SDIPS, one must also use the\nconductances GF(S)\nφ= 2GQ/summationtext\nn(ρF(S)\nn−4[τS(F)\nn/Tn]),\nGF(S)\nξ=−GQ/summationtext\nnτS(F)\nnandGF(S)\nχ=GQ/summationtext\nnTn(ρF(S)\nn+\nτS(F)\nn)/4, withρm\nn= Im[rm\nn,↑rm∗\nn,↓] andτm\nn= Im[tm\nn,↑tm∗\nn,↓]\nform∈ {S,F}. In the following, I will focus on the ef-\nfects ofGF\nφandGS\nφ, and I will assume GMR,GF(S)\nξand\nGF(S)\nχto be negligible, like found with a simple barrier\nmodel in the limit Tn≪1 andEex≪EF29. In this\ncase, one finds that the boundary conditions for the S/F\ninterface located at x=xj= (−1)jds/2, withj∈ {1,2},\nwrite\nξF∂θF\nn,σ\n∂x/vextendsingle/vextendsingle/vextendsingle/vextendsingle/vextendsingle\nxj= (−1)jγTsin[θF\nn,σ−θS\nn,σ]+iγF\nφǫC,j\nn,σsin[θF\nn,σ]\n(5)\nand\nξF∂θF\nn,σ\n∂x/vextendsingle/vextendsingle/vextendsingle/vextendsingle/vextendsingle\nxj−ξS\nγ∂θS\nn,σ\n∂x/vextendsingle/vextendsingle/vextendsingle/vextendsingle/vextendsingle\nxj=/summationdisplay\nm∈{F,S}iγm\nφǫC,j\nn,σsin[θm\nn,σ]\n(6)\nwhere the indices SandFindicate whether θn,σ\nand its derivative are taken at the SorFside\nof the interface. These equations involve the re-\nduced conductances γT=GTξF/AσFandγF(S)\nφ=\nGF(S)\nφξF/AσF, the barrier asymmetry coefficient γ=\nξSσF/ξFσS, the superconducting coherence lengthscale\nξS= (/planckover2pi1DS/2∆BCS)1/2, the magnetic coherence length-\nscaleξF= (/planckover2pi1DF/|Eex|)1/2, the gap ∆ BCSfor a bulk S,\nthe normal state conductivity σF(S)of theF(S) mate-\nrial, and the junction area A. The coefficient ǫC\nn,jtakes\ninto account the direction of the ferromagnetic polar-\nization of electrode Fjin configuration C ∈ {P,AP}.\nOne can use the convention ǫP,j\nn,σ= (−1)jσsgn(ωn) and\nǫAP,j\nn,σ=σsgn(ωn), in which the factor sgn( ωn) arising\nfrom the definition of θn,σand the terms ( −1)jandσ\narising from the boundary conditions have been included\nforcompactnessoftheexpressions. Notethatinthepres-\nence of a finite SDIPS i.e. γF(S)\nφ/ne}ationslash= 0, the right hand side\nof equation (6) is not zero contrarily to what found in\nthe spin-degenerate case15. In the general case, γF\nφand\nγS\nφare different (see e.g. Appendix A). This implies that,\nwith the present approximations, each interface is char-\nacterized by three parameters: γT,γF\nφandγS\nφ. For the\nsake of simplicity, symmetric F/S/Ftrilayers are con-\nsidered, so that γT,γF\nφandγS\nφare the same for the two\nS/Finterfaces.\nBefore working out the above system of equations, it is\ninteresting to note that the angle θn,σcalculated in theparallel configuration C=Pforx >0 also corresponds\nto the angle θn,σexpected for a S/Fbilayer consisting of\na superconductor Sfor 0< x < d S/2, and a ferromag-\nnetic electrode Fforx∈ {dS/2,dS/2+dF}(Figure 1.b).\nIn practice, using a F/S/Fgeometry can allow one to\nobtain more information on spin effects, as shown below.\nIII. CASE OF A THIN SUPERCONDUCTOR\nAND A WEAK PROXIMITY EFFECT IN F\nA. Spatial variations of the pairing angle\nI will assume that the amplitude of the superconduct-\ning correlations in F1(2)is weak, i.e. |θn,σ| ≪1 for\nx∈ {−dS/2−dF,−dS/2}andx∈ {dS/2,dS/2 +dF}\n(hypothesis 1) so that one can develop the Usadel equa-\ntion (2) at first order in θn,σ. This leads to\n∂2θn,σ\n∂x2−/parenleftBigg\nkC,j\nn,σ\nξF/parenrightBigg2\nθn,σ= 0 (7)\nin the ferromagnet Fj, withj∈ {1,2},kC,j\nn,σ= (2[iηC,j\nn,σ+\n|ωn/Eex|)])1/2andηC,j\nn,σ= (−1)jǫC,j\nn,σ. Combining Eqs.\n(4) and (7), one finds in Fj\nθn,σ(x) =θF\nn,σ(xj)cosh/parenleftbigg/bracketleftbig\nx−(−1)j/parenleftbig\ndF+dS\n2/parenrightbig/bracketrightbigkC,j\nn,σ\nξF/parenrightbigg\ncosh/parenleftBig\ndFkC,j\nn,σ\nξF/parenrightBig\n(8)\nThis result together with boundary condition (5) leads\nto\nθF\nn,σ(xj) =γTsin(θS\nn,σ(xj))\nγTcos(θSn,σ(xj))+iγF\nφηC,j\nn,σ+BC,j\nn,σ(9)\nwithBC,j\nn,σ=kC,j\nn,σtanh[dFkC,j\nn,σ/ξF]. Thisallowstorewrite\nthe boundary condition (6) in closed form with respect\ntoθS\nn,σ, i.e.\nξS∂θS\nn,σ\n∂x/vextendsingle/vextendsingle/vextendsingle/vextendsingle/vextendsingle\nxj= (−1)j+1LC\nj,n,σsin[θS\nn,σ(xj)] (10)\nfor theS/Finterface located at x=xj, with\nLC\nj,n,σ\nγ=γT(BC,j\nn,σ+iγF\nφηC,j\nn,σ)\nγTcos(θSn,σ)+BC,j\nn,σ+iγF\nφηC,j\nn,σ+iγS\nφηC,j\nn,σ(11)\nInthe following, Iwill assume |Eex| ≫∆BCSlikeinmost\nexperiments, so that kC,j\nn,σ= 1+iηC,j\nn,σ. I will also assume\ndS/ξS≤1, so that one can use, for the Cconfiguration\nand−dS/2< x < d S/2,\nθn,σ(x) =/tildewideθC\nn,σ−αC\nn,σ(x/ξS)−βC\nn,σ(x/ξS)2(12)4\nwith/vextendsingle/vextendsingle/vextendsingleθn,σ(x)−/tildewideθC\nn,σ/vextendsingle/vextendsingle/vextendsingle≪1 (hypothesis 2 ). Note that al-\nthough experiments are often performed in the limit of\nthick superconducting layers dS> ξS, assuming dS≤ξS\nis not unrealistic since one can obtain diffusive supercon-\nducting layerswith athickness dS∼ξS(seee.g. Ref. 34).\nFurthermore, using relatively low values of dSis more\nfavorable for obtaining efficient superconducting spin-\nswitches35. Hypothesis 2 allows one to develop sin( θn,σ)\nandcos(θn,σ)atfirstorderwithrespectto θn,σ−/tildewideθC\nn,σinS.\nAccordingly, I will neglect the space-dependence of ∆( x)\nand assign to it the value ∆Cin configuration C. Note\nthat I do not make any assumption on the value of the\nangle/tildewideθC\nn,σ, which is not necessarily close to the bulk BCS\nvalue. The coefficient LC\nj,n,σis transformed into its conju-\ngate when the magnetic polarizationofelectrode Fjis re-\nversed. Therefore, I will note LP\n1,n,σ=LP(AP)\n2,n,σ=Lσ\nnand\nLAP\n1,n,σ= (Lσ\nn)∗. The aboveassumptionsleadto αP\nn,σ= 0,\nβP\nn,σ=4Lσ\nnsin(/tildewideθC\nn,σ)\n4δS+cos(/tildewideθCn,σ)δ2\nSLσn(13)\nαAP\nn,σ=iIm[Lσ\nn]4sin(/tildewideθC\nn,σ)−cos(/tildewideθC\nn,σ)δ2\nSβAP\n4+2Re[ Lσn]cos(/tildewideθCn,σ)δS(14)\nand\nβAP\nn,σ=/parenleftBig\n8Re[Lσ\nn]+4|Lσ\nn|2cos(/tildewideθC\nn,σ)δS/parenrightBig\nsin(/tildewideθC\nn,σ)\n8δS+6Re[Lσn]cos(/tildewideθCn,σ)δ2\nS+|Lσn|2cos2(/tildewideθCn,σ)δ3\nS\n(15)\nwithδS=dS/ξS. On the other hand, from (1), one finds\nβC\nn,σ=∆Ccos(/tildewideθC\nn,σ)−|ωn|sin(/tildewideθC\nn,σ)\n2∆BCS(16)\nThe comparison between equations (13), (15) and (16)\nallowsoneto find /tildewideθC\nn,σasa function of ∆C. Then, one has\nto calculate ∆Cby using the self-consistency relation (3).\nI will study below the DOS and the critical temperature\nfollowingfromthese Eqs., in alimit whichleads tosimple\nanalytical expressions.\nB. Low-temperature density of states of S/F\nheterostructures\nThe DOS of the ferromagnets F1andF2of Figure 1.a\ncan be probed at x=±(dF+dS/2) by performing tun-\nnelling spectroscopy through an insulating layer. So far,\nthis quantity has been less measured3,36,37,38than criti-\ncal temperatures or supercurrents. However, this way of\nprobing the superconducting proximity effect is very in-\nteresting because it allows one to obtain spectroscopic\ninformation. It has been shown that the zero-energy\nDOS of a Flayer in contact with a superconductor os-\ncillates with the thickness of F. For certain thicknesses,this zero-energy DOS can even become higher than its\nnormal state value N039,40,41,42, as shown experimen-\ntally in Ref. 3. Remarkably, the SDIPS can shift these\noscillations29. Although the energy dependence of the\nDOS of diffusive S/Fstructures has raised some theoret-\nical and experimental interest, the effect of the SDIPS on\nthis energy dependence has not been investigated so far.\nFor calculating analytically the low-temperature\nDOS of the structure of Fig. 1.a., one can as-\nsumeγ2\nTcos(θS\nn,σ(xj))/|(γTcos(θS\nn,σ(xj)) +iγF\nφηP,2\nn,σ+\nBP,2\nn,σ)(γT+iγS\nφηP,2\nn,σ)| ≪1 (hypothesis 3 ), which leads\ntoLσ\nn=γ/parenleftBig\nγT+iγS\nφσsgn(ωn)/parenrightBig\n. This hypothesis is e.g.\nvalid for dF≥ξFand any value of γS(F)\nφifγTis rel-\natively small (see e.g. Fig. 2). I will also assume\nthat the lowest order terms in δSprevail in the numera-\ntors and denominators of expressions (13) and (15), i.e.\nβP\nn,σ∼ Lσ\nnsin(/tildewideθC\nn,σ)δ−1\nSandβAP\nn,σ∼Re[Lσ\nn]sin(/tildewideθC\nn,σ)δ−1\nS\n(hypothesis 4 ). Taking into account hypothesis 3 and\nγ∼1, hypothesis 4 is valid provided γTand/vextendsingle/vextendsingle/vextendsingleγS\nφ/vextendsingle/vextendsingle/vextendsingleare\nrelatively small compared to 1. Importantly, hypotheses\n3 and 4 are less restrictive regarding the value of γF\nφ. Ac-\ncordingly, I will often use γT,/vextendsingle/vextendsingle/vextendsingleγS\nφ/vextendsingle/vextendsingle/vextendsingle≪/vextendsingle/vextendsingle/vextendsingleγF\nφ/vextendsingle/vextendsingle/vextendsinglein the Figs. of\nthis paper. Hypotheses 3 and 4 lead to\n/tildewideθC\nn,σ= arctan/parenleftbigg∆C\n|ωn|+ΩCn,σ/parenrightbigg\n(17)\nwith\nΩP\nn,σ= 2∆BCSγ(γT+iγS\nφσsgn(ωn))δ−1\nS(18)\nand\nΩAP\nn,σ= 2∆BCSγγTδ−1\nS (19)\nEquations (17) and (18) show that an effective magnetic\nfieldHeffappears in the Slayer in the Pconfiguration,\ndue toγS\nφ/ne}ationslash= 0. From Eqs. (17) and (18), Heffcan be\nexpressed as\ngµBHeff=/planckover2pi1vS\nF\nds2GS\nφℓS\ne\n3σSA= 2ES\nTHGS\nφ\nGS(20)\nHere,vS\nFandℓS\nedenote the Fermi velocity and mean\nfree path in S, andGS=σSA/dSandES\nTH=/planckover2pi1DS/d2\nS\ndenote the normal state conductance and the Thouless\nenergy of the Slayer of Fig. 1.a. From Equations (17)\nand (19), the effective field effect disappears in the AP\nconfiguration because the two contacts are assumed to\nbe symmetric, and therefore, their contributions to Heff\ncompensate each other in the APcase. Note that, in\nprinciple, the γF\nφterm can induce an effective field ana-\nlogue to Heffin theFlayer, but this effect is not rele-\nvant in the regime studied in this paper (see Appendix\nB). The effects of Heffon the DOS of the structure will\nbe investigated in next paragraphs. In order to calculate5\nFIG. 2: Zero energy density of states NF(ε= 0) at x=\ndF+dS/2 as a function of dF, for the F/S/Fstructure of\nFigure 1.a in the Pconfiguration, with γ= 1,γT= 0.12,\nanddS/ξS= 1. The different curves correspond to different\nSDIPS parameters: γF\nφ=γS\nφ= 0 (black dotted curve), {γF\nφ=\n−2.1,γS\nφ= 0}(black full curve), and {γF\nφ=−2.1,γS\nφ= 0.12}\n(red full curve). The low-temperature self-consistent gap ∆P\nfound for these three different cases is indicated at the bott om\nright of the Figure. Using a strong value for γF\nφallows one\nto change significantly the phase of the oscillations of NF\nσ(0)\nwithdF(effect similar to Ref. 29). One can check that the\nhypotheses 1 to 4 are valid for the parameters used in this\nFigure. Note that for the different cases considered here, th e\ncritical temperature TP\ncof the structure (see Sec. III.F) is\nsuch that 0 .615TBCS\nc< TP\nc<0.635TBCS\nc. The four red\npoints correspond to the red full curves shown in Fig. 3.\n∆C, one has to combine the self-consistency relation (3)\nwith Eq. (17), which gives, at low temperatures,\nRe[log(ΩC\nn,σ+/radicalBig/parenleftbig\nΩCn,σ/parenrightbig2+(∆C)2\n∆BCS)] = 0 (21)\nThis equation can be solved numerically. The resulting\n∆Cis independent from the values of nandσused in\nEq. (21). Then, the value of the pairing angle θn,σin\nthe ferromagnets can be found by using Eqs. (8), (9),\n(12) and (17). Note that for γ≪1, the above Eqs.\nare in agreement with formula (5) of Ref. 29, obtained\nwith rigid boundary conditions, i.e. θn,σequal to its bulk\nvalue at the Sside. The energy dependence of θn,σcan\nbe found by performing the analytic continuation ωn=\n−iε+ Γ and sgn( ωn) = 1 in the above equations. The\nrate Γ = 0 .05 is used to account for inelastic processes43.\nAt last, the density of states Nσ(x,ε) at position xfor\nthe spin direction σ∈ {↑,↓}can be calculated by using\nNσ(x,ε) = (N0/2)Re[cos[ θn,σ(x)]], where N0/2 is the\nnormal density of states per spin direction.\nIn the following, I will mainly focus on NF(ε) =/summationtext\nσ∈{↑,↓}Nσ(x=dF,ε). Figure 2 shows the variations\nof the zero energy density of states NF(ε= 0) as a func-\ntion ofdF, for interface parameters γT=γS\nφ= 0.12 and\nγF\nφ=−2.1. Importantly, the value γT= 0.12 seems real-FIG.3: Densityofstates NF(ε)atx=dF+dS/2asafunction\nof the energy ε, for the F/S/Fstructure of Figure 1.a with\nγ= 1,γT= 0.12,dS/ξS= 1,γF\nφ=−2.1 andγS\nφ= 0.12.\nThe different panels correspond to different values of dF/ξF.\nThe red full curves correspond to the Pconfiguration and the\nblue dashed curves to the APconfiguration. For finite values\nofγS\nφ, some characteristic ”double-gap” structures appear in\nNF\nσ(ε) in the Pcase, due to a SDIPS-induced effective field\nappearing in S. The visibility of this effect strongly depends\non the thickness dFof theFlayers. For the different cases\nconsidered here, one can check that hypotheses 1 to 4 are\nvalid and that the critical temperature TC\ncof the structure\n(see Sec. III.F) is such that 0 .62TBCS\nc< TC\nc<0.67TBCS\nc.\nNote that for the above parameters and γS\nφ= 0, the curves\nobtained in the PandAPcase would be identical and very\nclose to the APcurves shown here.\nistic, at least for the weakly polarized Nb/Pd 0.9Ni0.1bi-\nlayers used in Ref. 3, for which one finds γT∼0.15 (see\nRef. 44). In addition, a simple barrier model suggests\nthat the situation/vextendsingle/vextendsingle/vextendsingleGF\nφ/vextendsingle/vextendsingle/vextendsingle≫GT, withGF\nφ<0 can happen\n(see Appendix A). The value γF\nφ=−2.1 used in Fig. 2\nthus seems possible .One can see that γF\nφcan change6\nsignificantly the phase of the oscillations of NF(ε= 0)\nwithdF(this effect has already been studied in Ref. 29\nin the case of rigid boundary conditions but I recall it\nhere for the sake of completeness). In Fig. 2, using for\nγF\nφa strong negative value allows to get NF(0)< N0for\nthe lowest values of dF, like often found in experiments.\nNote that from Fig. 2, γS\nφcan also shift the oscillations\nofNF(ε= 0) with dF. Here, this effect is much weaker\nthan that of γF\nφ, but one has to keep in mind that the\nlimitγS\nφ≪γF\nφis considered.\nEquation (20) shows that a measurement of Heff\nshould allow to determine the conductance GS\nφof a diffu-\nsiveS/Finterface. In this context, studying the energy\ndependence of NF(ε) is very interesting, because it can\nallow to see clear signatures of Heff, as shown below.\nFigure 3 shows the energy dependence of NF(ε) in the\nPandAPconfigurations, for a finite value of γS\nφand\ndifferent values of dF. ForC=P,NF(ε) shows some\n”double-gap” structures which disappear if the device is\nswitched to the APconfiguration. These double struc-\ntures are an indirect manifestation of the effective mag-\nnetic field Heffoccurring in Sin thePconfiguration,\ndue toγS\nφ/ne}ationslash= 0. Although Heffis localized in the S\nlayer, the double-gap structure that this field produces\nin the DOS of Sis transmitted to the DOS of Fdue to\nAndreev reflections occurring at the S/Finterfaces, as\nshown by Eq. (9). Interestingly, Rowell and McMillan\nhave already observed that an internal property of a S\nlayer can be seen through the superconducting proximity\neffect occurringin a nearbynormal layer. More precisely,\nthese authors have found that the DOS of an Ag layer\ncan reveal the phonon spectrum of an adjacent supercon-\nducting Pb layer47. Remarkably, the visibility of Heffin\nNF(ε) is modulated by quantum interferences occurring\ninF. Indeed, Heffis more visible for certain values of\ndF(e.g.dF/ξF= 1.0 or 1.2 in Fig. 3) than others (e.g.\ndF/ξF= 2.1 in Fig. 3), due to the dF-dependence of\nEq.(9).\nIt is useful to note that the SDIPS-induced effective\nfieldHeffshould also occur in the S/Fbilayer of Figure\n1.b. In this case, the Thouless energy and normal state\nconductance of the Slayer correspond to /tildewideES\nTH= 4ES\nTH\nand/tildewideGS= 2GSrespectively, sothat one finds gµBHeff=\n/tildewideES\nTHGS\nφ//tildewideGS. Double gap structures strikingly similar to\nthose shown in Figure 3 were indeed measured very re-\ncently by P. SanGiorgio et al., at the ferromagnetic side\nof diffusive Nb/Ni bilayers, in the absence of any exter-\nnal magnetic field48. Remarkably, the visibility of the\nobserved double structures varies with dF, as predicted\nabove. Note that in S/Fbilayers, the field Heffshould\nalso be observable directly at the Sside by measuring\nNS(ε) =/summationtext\nσ∈{↑,↓}Nσ(x= 0,ε). However, for param-\neters comparable to those of Figure 3, this should not\nenhance the resolution on Heff(see figure 4, left). For\ncertain values of dF,Heffis even more visible in NF(ε)\nthan inNS(ε)(see Figure 3).\nBefore concluding this section, I would like to empha-FIG. 4: Energy dependentdensityofstates NS(ε)measurable\nat the S side ( x= 0) of the S/Fstructure of Figure 1.b. The\nleft panel corresponds to the case of a metallic Fcontact\nwith parameters corresponding to that of Figs. 2 and 3. In\nthis case, the DOS measured at the Sside does not allow to\nresolve the SDIPS-induced effective field Heffbetter than a\nDOS measurement at the Fside of the structure, as can be\nseen from a comparison with Fig. 3. For certain values of dF,\nHeffis even more visible in NF(ε) thanNS(ε). The right\npanel corresponds to the case in which Fis not a metal but\nan insulating ferromagnet, i.e. γT= 0 . In this case, one\ncan use the reduced SDIPS parameter λS\nφ=GS\nφξS/AσS. For\nλS\nφ/ne}ationslash= 0,NS(ε) shows strong signatures of the effective field\nHeffinduced by the FIlayer inS. One can check that the\nhypotheses 1 to 4 are valid for the parameters used in this\nFigure.\nsize that from Eqs. (17) and (19), in the APconfig-\nuration, the SDIPS-induced effective field Heffdisap-\npears for the F/S/Fstructure considered in this paper\nbecause the two contacts are assumed to be symmetric\nand have thus opposite contributions to Heffin theAP\ncase. In the case of a dissymmetric structure, this should\nnot be true anymore, but the SDIPS-induced effective\nfield should nevertheless vary from the Pto theAPcase.\nThis is one practical advantage of working with F/S/F\ntrilayers instead of S/Fbilayers.\nC. Low-temperature density of states of S/FI\nbilayers\nTwenty years ago, internal Zeeman fields were ob-\nservedin superconductingAllayerscontactedtodifferent\ntypes of ferromagnetic insulators ( FI) (see Refs. 49,50,\n51,52). Using a ballistic S/FIbilayer model, Ref. 24\nsuggested that the observed internal fields could be in-\nduced by the SDIPS24. However, the inadequacy of this\nballistic approach for modeling the actual experiments\nwas pointed out in Ref. 52. Most of the experiments\non Al/FIinterfaces were interpreted by their authors\nin terms of a diffusive approach with no SDIPS, and an\ninternal Zeeman field added arbitrarily in the Al layer\n(see Refs. 51,52,53). The calculations of Section III.B.\nprovide a microscopic justification for the use of such an\ninternalfield in the diffusive model. Indeed, using γT= 0\nin the above calculations allows one to address the case7\nof diffusive S/FIbilayers. One finds that the SDIPS-\ninduced effective field Heffof Eq. (20) can occur in a\nthin diffusive Slayer contacted to a FIlayer. This effec-\ntive field effect can be seen e.g. in the density of states\nNS(ε) of theSlayer atx= 0 (see Figure 4, right). Re-\nmarkably, it wasfound experimentally52thatHeffscales\nwithd−1\ns, in agreement with Eq. (20)54.\nD. SDIPS-induced effective fields in other types of\nsystem\nInterestingly, the SDIPS can induce effective field ef-\nfects in othertypes ofsystems. First, the caseof S/N/FI\ntrilayers with a thickness dNof normal metal Nhas\nbeen studied theoretically27,28. In this case, a con-\nductance GN\nφsimilar to GS\nφcan be introduced to take\ninto account the SDIPS for electrons reflected by the\nFIlayer. The Nlayer is subject to an effective field\ngµBH′\neff=ETHGN\nφ/GNwithGNtheconductanceof N.\nThe expression of H′\neffis analogue to that of Heff(see\nEq. 20), up to a factor 2 which accounts for the symme-\ntry of the F/S/Fstructure with respect to x= 0 in the\nPconfiguration. Secondly, an effective field H′′\neffdefined\nbygµBH′′\neff= (/planckover2pi1vW\nF/2L)(ϕ↑\nL+ϕ↑\nR−ϕ↓\nL−ϕ↓\nR) has been\npredicted for a resonantsingle-channel ballistic wire with\nlengthLplaced between two ferromagnetic contacts19,\nwithϕσ\nL(R)the reflection phase of electrons with spin σ\nincident from the wire onto the left(right) contact, and\nvW\nFis the Fermi velocity in the wire. The expression of\nH′′\neffalsoshowsstrongsimilaritieswith that of Heff(see\nEq. 20, middle term). In practice, signatures of the field\nH′′\neffcould be identified in a carbon nanotube contacted\nwith two ferromagnetic contacts22,56. The fields Heff,\nH′\neffandH′′\neffhave the same physical origin: the en-\nergies of the states localized in the central conductor of\nthe structure depend on spin due to the spin-dependent\nphase shifts acquired by electrons at the boundaries of\nthis conductor. In all cases, the DOS of the central con-\nductor reveals the existence of the SDIPS-induced effec-\ntivefieldonlyifitalreadypresentsastrongenergydepen-\ndence near the Fermi energy in the absence of a SDIPS.\nIn theF/S/Fcase, this energy dependence is provided\nby the existence of the superconducting gap in S57. In\ntheS/N/FI case, it is provided by the existence of a su-\nperconducting minigap in N. At last, in the case of the\nballistic wire, it is provided by the existence of resonant\nstates in the wire.\nE. Comparison between the present work and\nother models for data interpretation in S/F\nheterostructures\nForcharacterizingthe propertiesof S/Finterfaces, one\nhas to interpret the experimental data showing the oscil-\nlationsof the density ofstates NF(ε) inFwith the thick-nessdFofF(or the oscillations of the critical current I0\nof aS/F/SJosephson with dF). However, if one uses a\nsimple description with spin-degenerate boundary condi-\ntions, the amplitude and the phase of these signals are\nnot independent, which makes the agreement with ex-\nperimental results impossible in most cases. The SDIPS\nconcept can solve this problem since it produces a shift\nof the signals oscillations with respect to the GS(F)\nφ= 0\ncase. However, in many cases, the observed shifts were\nattributed to the existence of a magnetically dead layer\n(MDL) at the Fside of the S/Finterface (see e.g. Refs.\n3,58,59). In other cases, the discrepancy between the\ntheory and the data was solved by taking into account\nspin-scattering processes in the Flayer (see e.g. Ref. 6).\nIn order to have a better insight into superconducting\nproximity effect experiments, one must stress the impor-\ntance of estimating experimentally the MDL thickness\nand the spin-scattering rate. In principle, spin-scattering\nrates can be estimated experimentally, as was done for\ninstance for the CuNi alloy60which is frequently used\nin proximity effect measurements (see e.g. 6,58). An\nexperimental determination of the MDL thickness has\nalso been performed in a few structures used to measure\nTcorI061,62,63,64,65, but, so far, this parameter has not\nbeen used for a real quantitative analysis of the data. In\nsomesituations, amodel combiningtheSDIPSwithspin-\nscattering and/or a MDL may be necessary. In any case,\nit is important to point out that descriptions based on\nspin-degenerateboundaryconditionsare, inprinciple, in-\ncomplete since they do not account for the effective field\neffect described in section III.B.\nBefore concluding this section, it is interesting to note\nthat the effective field effect produced by GS\nφinSor the\nphaseshift ofthe spatialoscillationsofthe DOS provided\nbyGF\nφinFwill remain qualitatively similar when the\nsigns ofGS\nφandGF\nφare changed (not shown). The sign\nofGF\nφintheexperimentsofRefs.3,5couldbedetermined\nfrom a quantitative comparison between the theory and\nthe data29. Below, we present a study of the critical tem-\nperatureof S/Fstructureswhichcangiveinformationon\nthe signs of GS\nφandGF\nφthrough qualitative signatures.\nF. Critical temperature of S/Fheterostructures\nThe critical temperature of S/Fhybrid structures\nhas already been the topic of many theoretical (see\ne.g. Refs. 13,14,35,66,67) and experimental (see e.g.\nRefs. 11,12) studies, but the effects of the SDIPS on this\nquantity have raised little attention so far. I show below\nthat the SDIPS can significantly modify the critical tem-\nperatures of S/Fdiffusive structures. Calculating the\ncritical temperature TC\ncof the structure of Figure 1.a in\nconfiguration Crequires to consider the limit in which\nsuperconducting correlations are weak in Sas well as in\nF(hypotheses 1 and 2 are then automatically satisfied).\nEquations (13), (15) and (16), then lead to8\nFIG. 5: Critical temperature TP\ncfor theF/S/Fstructure of\nFigure 1.a in the Pconfiguration (top panels) and difference\n∆Tc=TAP\nc−TP\ncbetween the critical temperatures in the P\nandAPconfigurations (bottom panels), as a function of the\nthickness dFof theFlayers. The left panels show the effect of\na finiteγS\nφand the right panels the effect of a finite γF\nφ. The\nfour panels show the case γS\nφ=γF\nφ= 0 with full red lines,\nfor comparison. The other parameters used in the Figure\nareγ= 1,γT= 0.12 anddS/ξS= 1. The SDIPS modifies\nTC\ncand ∆Tcin a quantitative or qualitative way, depending\non the case considered. The type of effects produced by the\nSDIPS on TC\ncand ∆Tcstrongly depend on the signs of γF\nφ\nandγS\nφ.\n/tildewideθC\nn,σ=∆C\n|ωn|+2∆BCSbCn,σδ−1\nS(22)\nwith\nbP\nn,σ=4L0\nn,σ\n4+δSL0n,σ(23)\nbAP\nn,σ=8Re[L0\nn,σ]+4/vextendsingle/vextendsingleL0\nn,σ/vextendsingle/vextendsingle2δS\n8+6Re[ L0n,σ]δS+/vextendsingle/vextendsingleL0n,σ/vextendsingle/vextendsingle2δ2\nS(24)\nand\nL0\nn,σ\nγ=γT(BP,2\nn,σ+iγF\nφσsgn(ωn))\nγT+BP,2\nn,σ+iγF\nφσsgn(ωn)+iγS\nφσsgn(ωn) (25)\nThese Eqs. together with (3) lead to\nlog/parenleftbiggTBCS\nc\nTCc/parenrightbigg\n= Re/bracketleftBigg\nΨ/parenleftBigg\n1\n2+bC\nn,σ\nδSexp(Γ)TBCS\nc\nTCc/parenrightBigg/bracketrightBigg\n−Ψ/parenleftbigg1\n2/parenrightbigg\n(26)where Γ denotes Euler’s constant. The resulting TC\ncis\nindependent from the values of nandσused in Eq. (26).\nNotethatinthecase γS\nφ=γF\nφ= 0andδS→0, thisequa-\ntion is in agreement with Eqs. (18) and (19) of Ref. 35.\nPerforminga numerical resolution of Eq. (26) together\nwith (23) and (24), one obtains the results of Fig. 5,\nwhich shows the critical temperature TP\ncof the structure\nin thePconfiguration (top panels) and the difference\n∆Tc=TAP\nc−TP\nc(bottom panels) as a function of dF,\nfor different interface parameters68. Here, for simplic-\nity, I consider cases where TC\ncdoes not show a strongly\nreentrant behavior, i.e. a cancellation of TC\ncin a certain\ninterval of dF. Such a behavior can happen for instance\nfor larger values of γT(see e.g. Ref. 35) or γ(S)F\nφ, but\nthis corresponds to a minority69of the experimental ob-\nservations made so far. The four panels of Fig. 5 show\nthe case γS\nφ=γF\nφ= 0 with full red lines, for comparison\nwith the other cases, where the SDIPS is finite. I first\ncomment the results obtained for the TP\nc(dF) curves (top\npanels of Fig. 5). In some cases, the SDIPS can modify\nquantitatively the TP\nc(dF) curves, for instance by am-\nplifying the dip expected in TP\nc(dF) in the absence of a\nSDIPS (see e.g. dotted curve in the upper left panel, cor-\nresponding to γS\nφ>0). In some other cases the SDIPS\ncan modify qualitatively the TP\nc(dF) curves, for instance\nby transforming the minimum expected in TP\nc(dF) into a\nmaximum (see e.g. curve with circles in the upper right\npanel, corresponding to γF\nφ<0). I now comment the\nresults obtained for the ∆ Tc(dF) curves (bottom pan-\nels of Fig. 5). In the range of parameters studied in\nthis work, one always finds ∆ Tc>0 because the effects\nof the two Flayers on Sare partially compensated in\ntheAPcase. In the absence of a SDIPS and for a low\nγT, the ∆Tc(dF) curve presents a maximum at a finite\nvalue ofdF(see red full lines in bottom panels). In some\ncases, the SDIPS can modify quantitatively the ∆ Tc(dF)\ncurves, for instance by increasing the value of this max-\nimum (see dotted curve in the bottom left panel, cor-\nresponding to γS\nφ>0). In other cases, the SDIPS can\nmodify qualitatively the ∆ Tc(dF) curves, for instance by\nturning the maximum expected in ∆ Tc(dF) into a mini-\nmum (see dashed curve in the bottom left panel, corre-\nsponding to γS\nφ<0), or by increasing the complexity of\nthe variations of ∆ TcwithdF(see curve with circles in\nthe bottom right panel, corresponding to γF\nφ<0), or by\ntransforming ∆ Tc(dF) into a monotonically decreasing\ncurve with a reduced amplitude (see dot-dashed curve in\nbottom right panel, corresponding to γF\nφ>0). Remark-\nably, the type of effects produced by the SDIPS on the\nTC\nc(dF) and ∆Tc(dF) curves depend on the sign of γF\nφ\nandγS\nφ. Critical temperature measurements can thus be\nan interesting way to determine the values of γF\nφandγS\nφ.\nFor simplicity, I have shown here results for γF\nφ/ne}ationslash= 0 and\nγS\nφ/ne}ationslash= 0 separately. Nevertheless, the behavior predicted\nforγF\nφ/ne}ationslash= 0 and γS\nφ/ne}ationslash= 0 simultaneously remain highly in-\nformative on the values of γF\nφandγS\nφ. Note that if γF\nφ9\nandγS\nφare increased compared to the values used in Fig.\n5, theTC\nc(dF) curve gets some cancellation points like\nin Ref. 35, but the behaviors of TC\ncand ∆Tcremain, in\nmany cases, qualitatively dependent on the signs of γF\nφ\nandγS\nφ(not shown).\nBeforeconcluding, Inotethat withthe approximations\nused in the previous section (III.B), the electronic corre-\nlations inside Swere affected by the presence of the F\nelectrodes through the parameter γS\nφonly. Consequently,\nthe self-consistent gap ∆CoftheSlayerwas independent\nfromdF, and it was furthermore identical for the Pand\nAPconfigurations for γS\nφ= 0. In the present section, I\nhave not neglected the dependence of TP\ncandTAP\ncondF\nandγF\nφbecause I have considered parameters for which\nhypothesis 3 is not acceptable anymore. In particular, I\nhave used lower values for dF70.\nIV. CONCLUSION\nThis article shows that the Spin-Dependence of In-\nterfacial Phase Shifts (SDIPS) can have a large variety\nof signatures in diffusive superconducting/ferromagnetic\n(S/F) heterostructures. Ref. 29 had already predicted\nthat the SDIPS produces a phase shifting of the oscilla-\ntions of the superconducting correlations with the thick-\nness ofFlayers. This article shows that this is not the\nonlyconsequenceoftheSDIPSin S/Fcircuits. Inpartic-\nular, the SDIPS can produce an effective magnetic field\nin a diffusive Slayer contacted to a diffusive Flayer.\nThis effective field can be seen e.g. through the DOS\nof the diffusive Flayer, with a visibility which oscillates\nwith the thickness of F. The SDIPS can also modify\nsignificantly the variations of the critical temperature of\naS/Fbilayer or a F/S/Ftrilayer with the thickness of\nF, either in a quantitative or in a qualitative way, de-\npending on the regime of parameters considered and in\nparticular the sign of the conductances GS\nφandGF\nφused\nto account for the SDIPS of the S/Finterfaces. In the\ncase of a F/S/Fspin valve, this last result also holds for\nthe thickness-dependence of the difference between the\ncritical temperatures in the parallel and antiparallel lead\nconfigurations. These effects should help to determine\nthe parameters GS\nφandGF\nφof diffusive S/Finterfaces.\nThe calculations shown in this paper are also appropri-\nate to the case of thin diffusive Slayers contacted to\nferromagnetic insulators.\nI thank P. SanGiorgio for showing me his experimental\ndata prior to publication, which stimulated part III.B of\nthis work. I acknowledge discussions with T. Kontos and\nW. Belzig. This work was supported by grants from the\nSwiss National Science Foundation and R´ egion Ile-de-\nFrance.V. APPENDIX A: PARAMETERS OF A S/F\nINTERFACE FROM A DIRAC BARRIER MODEL\nFIG. 6: Conductances GT(full line),˛˛GS\nφ˛˛(dashed lines), and˛˛GF\nφ˛˛(dash-dotted lines) reduced by the number of channels\nn, as a function of the spin-averaged barrier strength Zof a\nS/Finterface modeled with a Dirac barrier (see text). The\ncurveswere obtained with typical Fermi energies EF\nf= 0.8 eV\nandES\nf= 12 eV in FandSrespectively, a spin-polarization\nP= 0.06 of the density of states in F, and a spin asymmetry\nα= 0.06 for the barrier. Note that for the set of parameters\nused this Figure, one finds GF\nφ<0 andGS\nφ<0. However, for\nother parameters (e.g. by using α <0), one can reverse the\nsigns ofGF\nφandGS\nφ, or obtain opposite signs for GF\nφandGS\nφ\n(not shown).\nThe exact values of the conductances GT,GF\nφandGS\nφ\nof aS/Finterface depend on the details of this inter-\nface and on the microscopic structure of the contacted\nmaterials. Nevertheless, it is already interesting to study\na simplified Dirac barrier model which shows that the\nparameters regime assumed in this paper is, in princi-\nple, possible. Neglecting the transverse part of the elec-\ntronsmotion, one finds rS(F)\nn,σ= (kσ\nS(F)−kσ\nF(S)−iZσ)/Dσ\nandtS(F)\nn,σ= 2(kσ\nSkσ\nF)1/2/DσwithDσ=kσ\nS+kσ\nF+iZσ.\nHere,kσ\nS(F)is the Fermi electronic wavevector in S(F)\nandZσis the strength of the Dirac barrier for elec-\ntrons with spin σ. I use /planckover2pi1kσ\nS= (2meES\nf)1/2, with\nmethe free electron mass and ES\nfthe Fermi level in\nS. ForF, I use an s-band Stoner model, in which\n/planckover2pi1kσ\nF= (2meEF\nf[1±2P/(1 +P2)])1/2, withPthe spin-\npolarization of the density of states in FandEF\nfthe\nFermi level in F. I assume that the barrier can have a\nspin dependence α= (Z↓−Z↑)/(Z↑+Z↓) due to the\nferromagnetic contact material used to form the inter-\nface. The results given by this approach are shown in\nFig. 6. The conductances GT,/vextendsingle/vextendsingle/vextendsingleGF\nφ/vextendsingle/vextendsingle/vextendsingleand/vextendsingle/vextendsingle/vextendsingleGS\nφ/vextendsingle/vextendsingle/vextendsingle, reduced\nby the number of channels n, are shown as a function of\nthe spin-averagedbarrierstrength Z= (Z↑+Z↓)/2. The\nthree types of conductances go to zero when Zgoes to\ninfinity with αconstant71. One can see that in the limit\nTn≪1 (i.e. here GTh/ne2≪1) andP≪1 in which the10\nboundary conditions (5,6) have been derived,/vextendsingle/vextendsingle/vextendsingleGF\nφ/vextendsingle/vextendsingle/vextendsinglecan\nbe significantly stronger than GT, as assumed in Figs. 2\nand 3. Note that for the set of parameters used in Fig.\n6, one finds GF\nφ<0 andGS\nφ<0. However, for other pa-\nrameters (e.g. by using α <0), one can reverse the signs\nofGF\nφandGS\nφ, or obtain opposite signs for GF\nφandGS\nφ\n(not shown). This suggests that there is no fundamental\nconstraint on the signs of GF\nφandGS\nφin the general case.\nVI. APPENDIX B: EFFECTIVE FIELD EFFECT\nIN AFLAYER\nThis appendix reconsiders the case of the S/Fbilayer\nof Fig.1.b, in the limit dF≪ξFwhereθn,σ(x) can be\napproximatedwithaquadraticforminthe Flayer. From\nEqs. (2) and (5), one finds, for x∈[dS/2,dS/2+dF],\nθn,σ(x) =/tildewideθF\nn,σ+bF\nn,σ/bracketleftBigg\ndS−2x\ndF+/parenleftbiggdS−2x\n2dF/parenrightbigg2/bracketrightBigg\n(27)\nwith\n/tildewideθF\nn,σ= arctan\nγTsin(/tildewideθS\nn,σ)\nγTcos(/tildewideθSn,σ)+2ξF\ndFΩn\nEF\nTH+iγF\nφηP,2\nn,σ\n\n(28)\nbF\nn,σ=Ωnsin(/tildewideθF\nn,σ)\nEF\nTH(29)\nΩn=|ωn|+iEexηP,2\nn,σand/tildewideθS\nn,σ=θS\nn,σ(x=dS/2). Here,\nGF=σFA/dFandEF\nTH=/planckover2pi1DF/d2\nFdenote the con-\nductance and Thouless energy of the Flayer (note that\ndF≪ξF⇐⇒Eex≪EF\nTH). From completeness, I also\ngive the analogous equations for x∈[0,dS/2]. Assuming\nthatθn,σ(x) can be approximated with a quadratic form\nin theSlayer, one finds, from Eqs. (1) and (6),\nθn,σ(x) =/tildewideθS\nn,σ+bS\nn,σ/bracketleftBigg\n2(2x−dS)\ndS+/parenleftbigg2x−dS\ndS/parenrightbigg2/bracketrightBigg\n(30)\nwith\n/tildewideθS\nn,σ= arctan\nγγTsin(/tildewideθF\nn,σ)+4ξS\ndS∆\neES\nTH\nγγTcos(/tildewideθFn,σ)+4ξS\ndS|ωn|\neES\nTH+iγγS\nφηP,2\nn,σ\n\n(31)and\nbS\nn,σ=|ωn|sin(/tildewideθS\nn,σ)−∆cos(/tildewideθS\nn,σ)\n/tildewideES\nTH(32)\nThe notations used in the above equations are the same\nas in Section III. In particular, the Thouless energy and\nthe normal state conductance of the Slayer with thick-\nnessdS/2 are denoted /tildewideGS= 2σSA/dSand/tildewideES\nTH=\n4/planckover2pi1DS/d2\nS, and one has ηP,2\nn,σ=σsgn(ωn). In the limit\n/tildewideθF\nn,σ≪1, Eq. (31) is in agreement with Eqs. (17)\nand (18). The analytic continuation of Eq. (31) shows\nthat the Slayer is subject to the effective field Heff=\n/tildewideES\nTHGS\nφ//tildewideGS, in agreement with Eq. (20). The analytic\ncontinuationofEq. (28) showsthat the Flayerissubject\nto an analogue effective field HF\neff, defined by\ngµBHF\neff=EF\nTHGF\nφ\nGF(33)\nIt is interesting to compareEqs. (28) and (9). First, note\nthat in the regime/vextendsingle/vextendsingle/vextendsingle/tildewideθF\nn,σ/vextendsingle/vextendsingle/vextendsingle≪1, Eq. (28) can be recovered\nfromthe low dF-limit ofEq. (9). The interestofEq. (28)\nis that it is valid beyond the regime of a weak supercon-\nducting proximityeffect studied in part III. In particular,\nEq. (28) indicates that, in principle, the field EF\nTHcan\noccur in a thin Flayer for any value of the tunneling\nconductance GTof theS/Finterface. Interestingly, in\nthe case of a thick FlayerdF≥ξF/2, from Eq. (9), the\ncontribution of GF\nφto the pairing angle θF\nn,σ(x) cannot\nbe put anymore under the form of an effective exchange\nfield, due to the non-linearity of the BC,j\nn,σterm. Thus,\nstrictly speaking, the concept of a SDIPS-induced effec-\ntive field is valid only in the limit of thin metallic layers.\nHowever, it might be possible, in principle, to observe\nreminiscences of the double-gap structure appearing in\nNF(ε) in the regime of intermediate layer thicknesses\ndF∼ξFat least, due to the continuity of the equations.\nThen, one can wonder why this effect does not appear\nin section III. This is due to the regime of parameters\nchosen: section III assumes Eex≫∆BCSlike in most\nexperiments, and it furthermore focuses on the typical\nenergy range probed in superconducting proximity effect\nmeasurements, i.e. |ε|/lessorsimilar2∆BCS. In such a regime, one\ncan neglect the term |ωn|compared to iEexσsgn(ωn) in\nEqs. (9) and (28), so that HF\neffdoes not emerge in the\nmodel. Accordingly, signatures of the SDIPS-induced ef-\nfectivefield HF\neffarenotlikelytobeobservedinpractice.\n1A. I. Buzdin, L. N. Bulaevskii, and S. V. Panjukov, JETP\nLett.35, 178 (1982).\n2A. A. Golubov, M. Yu. Kupriyanov, and E. Il’ichev, Rev.Mod. Phys. 76, 411 (2004).\n3T. Kontos, M. Aprili, J. Lesueur, and X. Grison, Phys.\nRev. Lett. 86, 304 (2001).11\n4V. V. Ryazanov, V. A. Oboznov, A. Yu. Rusanov, A. V.\nVeretennikov, A. A. Golubov, and J. Aarts, Phys. Rev.\nLett.86, 2427 (2001).\n5T. Kontos, M. Aprili, J. Lesueur, F. Genˆ et, B. Stephanidis,\nand R. Boursier, Phys. Rev. Lett. 89, 137007 (2002).\n6H. Sellier, C. Baraduc, F. Lefloch, and R. Calemczuk,\nPhys. Rev. B 68, 054531 (2003).\n7Y. Blum, A. Tsukernik, M. Karpovski and A. Palevski,\nPhys. Rev. Lett. 89, 187004 (2002).\n8W. Guichard, M. Aprili, O. Bourgeois, T. Kontos, J.\nLesueur, and P. Gandit, Phys. Rev. Lett. 90, 167001\n(2003).\n9L. B. Ioffe, V. B. Geshkenbein, M. V. Feigel’man, A. L.\nFauch` ere, G. Blatter, Nature 398, 679 (1999).\n10T. Yamashita, K. Tanikawa, S. Takahashi and S. Maekawa,\nPhys. Rev. Lett. 95, 097001 (2005).\n11J. S. Jiang, D. Davidovic, D. H. Reich and C. L. Chien,\nPhys. Rev. Lett. 74, 314 (1995); L. V. Mercaldo, C. At-\ntanasio, C. Coccorese, L. Maritato, S. L. Prischepa and M.\nSalvato, Phys. Rev. B 53, 14040 (1996).\n12G. Deutscher and F. Meunier, Phys. Rev. Lett. 22,\n395(1969); J. Y. Gu, C.-Y. You, J. S. Jiang, J. Pear-\nson, Ya. B. Bazaliy, and S. D. Bader, Phys. Rev. Lett.\n89, 267001 (2002); A. Potenza and C. H. Marrows, Phys.\nRev. B71, 180503(R) (2005); Ion C. Moraru, W. P. Pratt,\nJr., and N. O. Birge, Phys. Rev. Lett. 96, 037004 (2006);\nI. C. Moraru, W. P. Pratt, Jr., and N. O. Birge, Phys. Rev.\nB74, 220507(R) (2006).\n13P. G. de Gennes, Phys. Lett. 23, 10 (1966).\n14L. R. Tagirov, Phys. Rev. Lett. 83, 2058 (1999); A. I.\nBuzdin, A. V. Vedyayev and N. V. Ryzhanova, Euro-\nphysics Lett. 48, 686 (1999).\n15M. Yu. Kuprianov and V. F. Lukichev, JETP 67, 1163\n(1988).\n16D. Huertas-Hernando, Yu. V. Nazarov, and W. Belzig,\ncond-mat/0204116; D. Huertas-Hernando, Phd Thesis,\nDelft University of Technology, The Netherlands (2002).\n17M. N. Baibich, J. M. Broto, A. Fert, F. Nguyen Van Dau,\nand F. Petroff, P. Etienne, G. Creuzet, A. Friederich, and\nJ. Chazelas, Phys. Rev. Lett. 61, 2472 (1988); G. Binasch,\nP. Gr¨ unberg, F. Saurenbach, and W. Zinn, Phys. Rev. B\n39, 4828 (1989).\n18A. Brataas, Y. V. Nazarov, and G. E. W. Bauer, Phys.\nRev. Lett. 84, 2481 (2000); D. Huertas Hernando, Yu. V.\nNazarov, A. Brataas, and G. E. W. Bauer, Phys. Rev. B\n62, 5700 (2000); A. Brataas, Y. V. Nazarov and G. E. W.\nBauer, Eur. Phys. J. B 22, 99 (2001).\n19A. Cottet, T. Kontos, W. Belzig, C. Sch¨ onenberger and C.\nBruder, Europhys. Lett. 74, 320 (2006).\n20A. Cottet, T. Kontos, S. Sahoo, H. T. Man, M.-S. Choi, W.\nBelzig, C. Bruder, A. F. Morpurgo and C. Sch¨ onenberger,\nSemicond. Sci. Technol. 21, S78 (2006).\n21W. Wetzels, G. E. W. Bauer, and M. Grifoni, Phys. Rev.\nB72, 020407(R) (2005); Phys. Rev. B 74, 224406 (2006).\n22A.CottetandM.-S.Choi, Phys.Rev.B. 74, 235316 (2006).\n23L. Balents and R. Egger, Phys. Rev. B 64, 035310 (2001);\nS. Koller, L. Mayrhofer and M. Grifoni, New J. Phys. 9,\n348 (2007).\n24T. Tokuyasu, J. A. Sauls and D. Rainer, Phys. Rev. B 38,\n8823 (1988).\n25A. Millis, D. Rainer, and J. A. Sauls, Phys. Rev. B 38,\n4504 (1988).\n26M. Fogelstr¨ om, ibid.62, 11812 (2000); J. C. Cuevas and\nM. Fogelstr¨ om, ibid. 64, 104502 (2001); M. Eschrig, J.Kopu, J. C. Cuevas, and G. Sch¨ on, Phys. Rev. Lett. 90,\n137003 (2003); J. Kopu, M. Eschrig, J. C. Cuevas, and\nM. Fogelstr¨ om, Phys. Rev. B 69, 094501 (2004); E. Zhao,\nT. L¨ ofwander, and J. A. Sauls, ibid. 70, 134510 (2004), M.\nEschrig, T. L¨ ofwander, T. Champel, J. C. Cuevas, J. Kopu\net Gerd Sch¨ on, J. Low Temp. Phys. 147, 457 (2007).\n27D. Huertas Hernando, Yu. V. Nazarov, and W. Belzig,\nPhys. Rev. Lett. 88, 047003 (2002).\n28D. Huertas-Hernando and Yu. V. Nazarov, Eur. Phys. J.\nB44, 373 (2005).\n29A. Cottet and W. Belzig, Phys. Rev. B. 72, 180503(R)\n(2005).\n30V.BraudeandYu.V.Nazarov, Phys.Rev.Lett.98, 077003\n(2007).\n31J. P. Morten, D. Huertas-Hernando, A. Brataas and W.\nBelzig, cond-mat/0612197.\n32N. B. Kopnin, Theory of Nonequilibrium Superconductiv-\nity, Clarendon Press, Oxford (2001); W. Belzig, F. K. Wil-\nhelm, C. Bruder, G. Sch¨ on and A. D. Zaikin, Superlatt.\nand Microstr. 25, 1251 (1999).\n33Yu. V. Nazarov, Superlatt. and Microstruc. 25, 1221\n(1999).\n34V. Fominov, N. M. Chtchelkatchev and A. A. Golubov,\nJETP Letters 74, 96 (2001).\n35I. Baladi´ e and A. Buzdin, Phys. Rev. B 67, 014523 (2003).\n36T. Kontos, M. Aprili, J. Lesueur, X. Grison, and L. Du-\nmoulin, Phys. Rev. Lett. 93, 137001 (2004).\n37L. Cr´ etinon, A. K. Gupta, H. Sellier, F. Lefloch, M. Faur´ e,\nA. Buzdin, and H. Courtois, Phys. Rev. B 72, 024511\n(2005).\n38S. Reymond, P. SanGiorgio, M. R. Beasley, J. Kim,\nT. Kim, and K. Char, Phys. Rev. B 73, 054505 (2006).\n39A. Buzdin, Phys. Rev. B 62, 11377 (2000); I. Baladi´ e and\nA. Buzdin, Phys. Rev. B 64, 224514 (2001).\n40M. Zareyan, W. Belzig, and Yu. V. Nazarov, Phys. Rev.\nLett.86, 308 (2001).\n41F. S. Bergeret, A. F. Volkov, and K. B. Efetov, Phys. Rev.\nB65, 134505 (2002).\n42T. Yokoyama, Y. Tanaka, and A. A. Golubov, Phys. Rev.\nB72, 052512 (2005).\n43W. Belzig , PhD thesis, Karlsruhe university (1999).\n44In practice, the conductance GTof aS/Finterface can be\nvery high (even for Tn≪1), and it can thus be very dif-\nficult to determine it from a normal state transport mea-\nsurement. In this case, an indirect determination of GT\nmust be realized. In Ref. 36, Kontos et al. have studied\nNb/Pd bilayers with a thickness dPd= 5 nm of Pd and a\ninterface area A= 100µm×100µm. They havefoundfrom\nminigap measurements46that the conductance GTof their\nNb/Pd interfaces was such that γB=AσN/GTdNb= 5.3,\ni.e.AσN/GT= 26.5 nm. In Ref. 3, Kontos et al. have stud-\nied Nb/Pd 0.9Ni0.1bilayers with the same interface area as\nfor the previous Nb/Pd bilayers. The two types of bilay-\ners were realized with the same sources of Nb and Pd and\nsimilar fabrication processes. Considering that only 10%\nof Ni were added to the Pd material in the S/Fcase, the\nconductivity of Pd 0.9Ni0.1was probably of the same order\nas that of Pd, and the conductance of the Nb/Pd interface\nwas probably of the same order as that of Nb/Pd 0.9Ni0.1.\nThis leads to AσF/GT∼26.5 nm. Then, using the value\nξF∼4 nm found in Refs. 3,45 for dPd0.9Ni0.1= 8 nm, one\nfindsγT=GTξF/AσF∼0.15.\n45T. Kontos, Ph.D. thesis, Universit´ e Paris-Sud, Orsay,\nFrance, 2002.12\n46The width ofthe minigap appearing inthe densityofstates\nof aS/Nbilayer versus energy gives a direct access to the\nparameter γTof this interface, as explained in A. Golubov\nand M. Yu. Kuprianov, Sov. Phys. JETP 69, 805 (1989).\n47J. M. Rowell and W. L. McMillan, Phys. Rev. Lett. 16,\n453 (1966).\n48Paul SanGiorgio, private communication.\n49P. M. Tedrow, J. E. Tkaczyk, and A. Kumar, Phys. Rev.\nLett.56, 1746 (1986).\n50R. Meservey and P. M. Tedrow, Phys. Rep. 238, 173\n(1994).\n51J. S. Moodera, X. Hao, G. A. Gibson, and R. Meservey,\nPhys. Rev. Lett. 61, 637 (1988), X. Hao, J. S. Moodera,\nand R. Meservey, Phys. Rev. B 42, 8235 (1990).\n52X. Hao, J. S. Moodera, and R. Meservey, Phys. Rev. Lett.\n67, 1342 (1991).\n53J. A. X. Alexander, T. P. Orlando, D. Rainer, and P. M.\nTedrow, Phys. Rev. B 31, 5811 (1985).\n54The fact that Heffscales with d−1\nsis also in agreement\nwith Eq. (1) of Ref. 13, proposed by De Gennes from\nCooper’s argument55.\n55L. N. Cooper, Phys. Rev. Lett. 6, 689 (1961); P. G. de\nGennes, Rev. Mod. Phys. 36, 225 (1964).\n56S. Sahoo, T. Kontos, J. Furer, C. Hoffmann, M. Graber, A.\nCottet and C. Sch¨ onenberger, Nature Phys. 1, 99 (2005).\n57Without superconductivity, Heffwould not be not resolv-\nable anymore in the S/Fstructures considered in this pa-\nper, since it was assumed implicitly that the normal-state\nDOS of these structures is constant near the Fermi level.\n58V. A. Oboznov, V. V. Bol’ginov, A. K. Feofanov, V. V.\nRyazanov and A. I. Buzdin, Phys. Rev. Lett. 96, 197003\n(2006).\n59M. Weides, M. Kemmler, E. Goldobin, D. Koelle, and R.\nKleiner, H. Kohlstedt, A. Buzdin, Appl. Phys. Lett. 89,\n122511 (2006).\n60W. P. Pratt, Jr., private communication; see also S.-Y.\nHsu, P. Holody, R. Loloee, J. M. Rittner, W. P. Pratt, Jr.,\nand P. A. Schroeder, Phys. Rev. B 54, 9027 (1996).61Th. M¨ uhge, N. N. Garif’yanov, Yu. V. Goryunov, G. G.\nKhaliullin, L. R. Tagirov, K. Westerholt, I. A. Garifullin,\nand H. Zabel, Phys. Rev. Lett. 77, 1857 (1996).\n62J. Aarts and J. M. E. Geers, E. Br¨ uck, A. A. Golubov, R.\nCoehoorn, Phys. Rev. B 56, 2779 (1997).\n63J. W. A. Robinson, S. Piano, G. Burnell, C. Bell, M. G.\nBlamire, Phys. Rev. Lett. 97, 177003 (2006).\n64S. Piano , J. W. A. Robinson, G. Burnell, M. G. Blamire,\nEur. Phys. J. B 58, 123 (2007).\n65C. Bell, R. Loloee, G. Burnell, and M. G. Blamire, Phys.\nRev. B71, 180501(R) (2005).\n66E. A. Demler, G. B. Arnold and M. R. Beasley, Phys.\nRev. B55, 15174 (1997); M. G. Khusainov and Yu. N.\nProshin, Phys. Rev. B 56, R14283 (1997); L. R. Tagirov,\nPhysica C 307, 145 (1998); Yu. N. Proshin and M. G.\nKhusainov JETP 86, 930 (1998); Ya. V. Fominov, N. M.\nChtchelkatchev, and A. A. Golubov, Phys. Rev. B 66,\n014507 (2002); A. Bagrets, C. Lacroix and A. Vedyayev,\nPhys. Rev. B 68, 054532 (2003).\n67C.-Y. You, Ya.B. Bazaliy, J. Y.Gu, S.-J. Oh, L.M. Litvak,\nand S. D. Bader, Phys. Rev. B 70, 014505 (2004).\n68With the parameters of Fig. 5, the approximations bP\nn,σ=\nL0\nn,σandbAP\nn,σ= Re[L0\nn,σ] used in Ref. 35 would be suffi-\ncient for calculating the TP\ncandTAP\nccurves, but not the\ndifference ∆ Tc. Indeed, for the different cases studied in\nthis Figure, the relative difference between the results ob-\ntained with these approximations and those given by Eqs.\n(23) and (24) is smaller than ∼0.8% forTP\ncandTAP\nc, but\nit can reach ∼18% for ∆ Tc. This is why I have used the\ngeneral equations (23) and (24) for plotting Figure 5.\n69L. R. Tagirov, I. A. Garifullin, N. N. Garifyanov, S. Ya.\nKhlebnikov, D. A. Tikhonov, K. Westerholt, and H. Zabel,\nJ. Magn. Magn. Mater. 240, 577 (2002).\n70Note that both regimes dF≤ξF(see e.g. Ref 37) and\ndF≥ξF(see e.g. Ref. 67) seem to be accessible in practice.\n71ForZ→+∞,GF(S)\nφvanishes because arg( rF(S)\nn,σ)→πand\narg(tF(S)\nn,σ)→ −π/2 forσ∈ {↑,↓}." }, { "title": "2401.04219v2.Ferromagnetic_Materials_for_Josephson_π_Junctions.pdf", "content": "Ferromagnetic Materials for Josephson πJunctions\nNorman O. Birge1,a)and Nathan Satchell2,b)\n1)Department of Physics and Astronomy, Michigan State University, East Lansing,\nMichigan 48824, USA\n2)Department of Physics, Texas State University, San Marcos, Texas, 78666,\nUSA\n(Dated: 12 March 2024)\nThe past two decades have seen an explosion of work on Josephson junctions contain-\ning ferromagnetic materials. Such junctions are under consideration for applications\nin digital superconducting logic and memory. In the presence of the exchange field,\nspin-singlet Cooper pairs from conventional superconductors undergo rapid oscilla-\ntions in phase as they propagate through a ferromagnetic material. As a result,\nthe ground-state phase difference across a ferromagnetic Josephson junction oscil-\nlates between 0 and πas a function of the thickness of the ferromagnetic material.\nπ-junctions have been proposed as circuit elements in superconducting digital logic\nand in certain qubit designs for quantum computing. If a junction contains two or\nmore ferromagnetic layers whose relative magnetization directions can be controlled\nby a small applied magnetic field, then the junction can serve as the foundation for a\nmemory cell. Success in all of those applications requires careful choices of ferromag-\nnetic materials. Often, materials that optimize magnetic properties do not optimize\nsupercurrent propagation, and vice versa. In this review we discuss the significant\nprogress that has been made in identifying and testing a wide range of ferromagnetic\nmaterials in Josephson junctions over the past two decades. The review concentrates\non ferromagnetic metals, partly because eventual industrial applications of ferromag-\nnetic Josephson junctions will most likely start with metallic ferromagnets (either in\nall metal junctions or junctions containing also an insulating layer). We will briefly\nmention work on non-metallic barriers, including ferromagnetic insulators, and some\nof the exciting work on spin-triplet supercurrent in junctions containing noncollinear\nmagnetic inhomogeneity.\na)Electronic mail: birge@msu.edu\nb)Electronic mail: satchell@txstate.edu\n1arXiv:2401.04219v2 [cond-mat.supr-con] 8 Mar 2024I. INTRODUCTION\nIt has been over 20 years since Ryazanov and co-workers first demonstrated a mag-\nnetic Josephson junction whose ground-state phase difference could be either zero or π\ndepending on the sample temperature.1,2That breakthrough was followed soon after by\nKontos et al. , who demonstrated the 0 −πtransition as a function of the thickness of\nthe ferromagnetic material in a series of magnetic junctions.3The evidence for the 0 −π\ntransition in those pioneering works was indirect in the sense that only the amplitude of\nthe critical current was measured; direct measurements of the junction phase were carried\nout soon thereafter.4–7The physical mechanism underlying these phenomena had been un-\nderstood already 20 years earlier;8–10nevertheless, the laboratory demonstrations of these\nphenomena sparked widespread interest in magnetic Josephson junctions and in π-junctions\nin general.11–13There have been many proposals to use π-junctions as circuit elements in\nsuperconducting digital logic14,15and in certain qubit designs for quantum computing,16,17\nand several of these ideas are being tested in the laboratory (see Section III of this review).\nAt about the same time as the first π-junction demonstrations, Bergeret, Volkov and\nEfetov predicted theoretically that magnetic Josephson junctions made from conventional\nspin-singlet superconductors could carry long-range spin-triplet supercurrent under specific\nconditions that include the presence of non-collinear magnetic inhomogeneities inside the\njunctions.18–21Unlike the demonstration of magnetic πjunctions, the Bergeret prediction\ncame as a total surprise to the superconductivity community. The idea that supercurrents\ncould be spin-polarized led to much excitement and the coining of the name “superconduct-\ning spintronics” for this new area of condensed matter physics.22–26\nPut together, the two breakthroughs mentioned above catalyzed a huge surge in interest\nin superconducting/ferromagnetic (S/F) hybrid systems. This review covers only a small\nsubset of the work that has been done since 2001, focusing mainly on S/F/S π-junctions\nand their variants, including S/I/F/S, S/I/s/F/S and S/FI/S junctions (where I = insulator\nand FI = ferromagnetic insulator). Our emphasis is on the wide range of magnetic materials\nthat have been used in such devices. The centerpiece of the review are the Tables in Section\nIV, which we hope provide a complete list of magnetic materials that have been used inside\nJosephson junctions, usually with the aim of making π-junctions. To our knowledge, the last\ntime such a compilation has been made was in 2006 by Kupriyanov, Golubov, and Siegel.27\n2That review covered both experimental and theoretical developments; ours focuses almost\nentirely on experimental results.\nThere has been a lot of work on other experimental signatures of S/F physics besides π-\nstate Josephson junctions, such as the critical temperature and critical field of S/F bilayers\nand multilayers. We do not discuss those results here, even though some of that work28–36\npredated the experimental demonstrations of πJosephson junctions discussed above.\nThe structure of this review is as follows: Section II discusses general considerations,\nincluding the physical parameters that describe Josephson junctions, ballistic vs diffusive\ntransport regimes, the different types of junctions, and how the junction response to a\nmagnetic field is modified when the junction contain ferromagnetic materials. Section III\ndiscusses applications of π-junctions in quantum and classical circuits. Section IV discusses\nmetallic S/F/S π-junctions and contains the Tables of magnetic materials mentioned above.\nIt includes a short subsection about “spin-valve” junctions whose critical current or phase\nstate are controllable by changing the magnetic configuration inside the junction. Section V\nmentions briefly the many different types of π-junctions that have been created in the past\nfew years in exotic materials other than metals. Section VI discusses spin-triplet Joseph-\nson junctions, including a very brief mention of so-called “ ϕandϕ0-junctions.” Finally,\nSection VII contains a discussion of some open challenges and directions for the future.\nII. GENERAL CONSIDERATIONS\nA. S/F physics, length scales, and theoretical predictions for IcRN\nThe physics of S/F systems has been reviewed many times;11–13,21,23we encourage readers\nnew to the topic to read the excellent pedagogical article by Demler et al.37Superconducting\nand ferromagnetic materials in contact or in close proximity to each other interact in two\ndistinct ways. Effects due to the magnetic field produced by F are generally referred to as\n“orbital effects”; they are nearly always present in experiments and should not be forgotten,\nbut they are not the main focus of this review. Here we focus on proximity effects, whereby\nelectron pair correlations are induced in a material in contact with a superconductor.\nWe start by defining several important length scales. The superconducting coherence\nlength is defined as ξbal\nS=ℏvF/∆ or ξdif\nS= (ℏDS/∆)1/2in the ballistic and diffusive limits,\n3respectively, where ∆ is the superconducting gap, and vFandDSare the Fermi velocity and\ndiffusion constant in S. The diffusion constant, D, is related to the electron mean free path,\nl, via D=1\n3vFl. In an S/N bilayer (N = normal metal), pair correlations penetrate N over\na distance called the “normal-metal coherence length”, which has the following forms in the\nballistic or diffusive limits: ξbal\nN=ℏvF/(2πkBT) orξdif\nN= (ℏDN/(2πkBT))1/2, where vFand\nDNare now in the normal metal, and Tis the temperature. That length can be quite long –\nup to several hundred nm in noble metals at dilution refrigerator temperatures. In contrast,\nin ferromagnetic materials the pair correlations oscillate in sign due to the exchange split-\nting between the majority and minority spin bands in F. (Those oscillations are sometimes\nreferred to as “FFLO oscillations” after Fulde-Ferrell38and Larkin-Ovchinnikov,39but we\nemphasize that the original FFLO prediction was for a bulk ferromagnetic superconductor\nin the ballistic limit, rather than a proximity system.) The period of the oscillations is 2 πξF,\nwhere ξFis sometimes called the “exchange length” or the “ferromagnetic coherence length”.\nIn the ballistic or diffusive limits with Eex>> k BT, the results are ξbal\nF=ℏvF/(2Eex) or\nξdif\nF= (ℏDF/Eex)1/2, respectively, where 2 Eexis the exchange splitting between the ma-\njority and minority spin bands. In addition to the oscillation, the pair correlations decay\nalgebraically in the ballistic limit or exponentially in the diffusive limit. Due to the large\nexchange energies in strong F materials, Eex≈1eV,ξFis less than 1 nm in those materials.\nThat is one reason why experimental progress in S/F systems lagged behind the theoretical\ndevelopments for so long. One of the keys to the experimental breakthroughs of 2001 was\nthe use of dilute ferromagnetic alloys: CuNi alloy in the case of the Ryazanov group2and\nPdNi alloy in the case of the Aprili group.3,40Diluting the Ni decreased Eexsubstantially\nthereby increasing ξFto values of several nm. In that limit one must include the contribution\nof the temperature to the exchange length ξF; in fact, Ryazanov’s first demonstration of a\nπ-junction was carried out by varying the temperature of a Josephson junction with fixed\nferromagnetic layer thickness.2\nWhen discussing the critical current, Ic, of Josephson junctions, we will quote results\nfor the IcRNproduct, where RNis the normal-state resistance of the junction, usually\ndetermined from the slope of the I–Vrelation for applied currents I >> I c.IcRNis a useful\nquantity because it is independent of junction area and because it can be compared with\nthe standard Ambegaokar-Baratoff result for short S/I/S junctions: IcRN=π∆/(2e).41\nCalculations of Icfor ferromagnetic Josephson junctions have been carried out in several\n4different regimes. In addition to the length scales defined above, one must also consider the\nelectron mean free path, lF, and the thickness dFof the F layer. The first calculation of\nIcRNwas performed for the purely ballistic or “clean” limit, defined by dF<< ξ S<< l F, lS\n– i.e. neglecting impurity scattering in both F and S.9The decay and oscillations of IcRN\nare given by the numerical maximum with respect to φof the ballistic limit supercurrent\nIS(φ),9\nIS(φ)RN=π∆α2\n2eZ∞\nαdy\ny3\"\nsin \nφ−y\n2!\ntanh \n∆ cosφ−y\n2\n2kBT!\n+ sin \nφ+y\n2!\ntanh \n∆ cosφ+y\n2\n2kBT!#\n,(1)\nwhere φis the phase difference across the junction, and α≡dF/ξbal\nF. In the limit of large α\nand for Tnear Tc, Eqn. 1 can be simplified to,9\nIcRN=π∆2\n4eT\f\f\f\f\fsin\u0000\ndF/ξbal\nF\u0001\n\u0000\ndF/ξbal\nF\u0001\f\f\f\f\f. (2)\nThe simplified form of the ballistic limit shows 0 −πoscillations at dF/ξbal\nF=nπ, but misses\nthe first transition, which occurs at dF/ξbal\nF≈0.33π.\nThe opposite limit is the diffusive or “dirty” limit, defined by lF, lS<< d F, ξS, ξF– i.e.\nthe limit where impurity scattering is very strong in both F and S. In that limit the Usadel\nequation is valid, and the original calculation produces the result:10,13\nIcRN=π∆2\n4eTc2x\f\f\f\fcos(x) sinh( x) + sin( x) cosh( x)\ncosh(2 x)−cos(2 x)\f\f\f\f, (3)\nwhere x=dF/ξdif\nF. In the limit dF>> ξdif\nFandTnear Tc, that result simplifies to:10,13\nIcRN=√\n2∆2\n4eTc \ndF\nξdif\nF!\nexp \n−dF\nξdif\nF!\f\f\f\f\fsin \ndF\nξdif\nF+π\n4!\f\f\f\f\f, (4)\nwhich predicts 0 −πoscillations at dF/ξdif\nF= (n+ 3/4)π.\nThere are several ways in which real Josephson junctions deviate from the strict assump-\ntions used to calculate the expressions above. We know of only one paper that incorporates\nrealistic models of the Fermi surfaces of the S and F materials into a calculation of Ic.42\nIn addition, strong F materials have different mean free paths for majority and minority\nelectrons, which are not captured in the usual theoretical formulations. In the presence of\n5spin-flip or spin-orbit scattering in F, the length scales governing the decay and oscillations\nare no longer equal,43,44hence one defines independent length scales ξF1andξF2for the two\nphenomena. In addition, it has been shown that the position of the first 0 −πtransition\nvaries depending on the boundary conditions at the S/F interface, including the presence of\nan insulating barrier or a normal-metal spacer layer.44–46\nIt is common to address some of the effects discussed above by utilizing a phenomeno-\nlogical expression for the diffusive regime:\nIcRN=IcRN(0) exp\u0012−dF\nξF1\u0013\f\f\f\fsin\u0012dF\nξF2+ϕ\u0013\f\f\f\f, (5)\nwhere ϕis not necessarily equal to π/4 and IcRN(0) is a fictitious zero thickness fit parameter.\nSpin-dependent scattering generally causes ξF1to decrease and ξF2to increase relative to\nξdif\nF, resulting in the relation ξF1< ξ F2, which is observed in junctions containing weakly-\nferromagnetic CuNi alloys, as seen in Table III.\nSeveral authors have pointed out that the strong elemental F materials, Ni, Fe, and Co, do\nnot satisfy the conditions of validity of the Usadel equation. In such materials the exchange\nenergy is very large, hence the length ξbal\nFis very short. Bergeret et al.47were the first to\ndiscuss the intermediate limit, defined by ξbal\nF<< l F<< ξbal\nS. This limit clearly requires\nthat Eex>>∆, which is true for the strong F materials. The expression for IcRNin the\nintermediate limit involves a sum over Matsubara frequencies, which we do not reproduce\nhere. (See Eqns. (19) and (20) in Ref. 47.) The dependence of IcRNondFhas a similar\nform to Eqn. (5), except now one finds that the oscillation period is governed by the ballistic\nexpression for the exchange length, ξF2=ξbal\nF, while the decay is governed by the mean free\npath, ξF1≈lF. Thus, in the intermediate limit one expects to observe ξF1> ξF2, which is\nindeed the case for many of the materials listed in Table I.\nWhen it comes to fitting experimental data, one requires data over a wide range of\nthicknesses to obtain accurate estimates of all the parameters in Eqn. (5). If only a single\n0−πtransition is observed in the measured thickness range, then the parameters ξF2and\nϕhave large uncertainties and are highly correlated. To remove that correlation, we prefer\nto use the following fitting form:\nIcRN=IcRN(0) exp\u0012−dF\nξF1\u0013\f\f\f\fsin\u0012dF−d0−π\nξF2\u0013\f\f\f\f, (6)\nwhere d0−πis the thickness of the first 0 −πtransition, which is often determined quite\n6accurately in the experiments. Expression (6) is the one we used when analyzing data from\nother groups, to extract the parameters shown in the Tables in Section IV.\nB. Josephson junction geometry: planar vs “sandwich-style”\nMicrofabricated Josephson junctions generally take one of two geometries. In planar or\nlateral junctions, the supercurrent flows in the plane of the substrate, while in “sandwich-\nstyle” junctions, the current flows perpendicular to the plane. Most S/I/S junctions are\nof the sandwich type due to the necessity of keeping the insulating barrier extremely thin.\nIn contrast, most S/N/S junctions studied nowadays are planar so that the supercurrent\nflows over a long distance in the N metal. Exceptions include junctions containing very\nhigh-resistivity metals such as silicides,48or junctions studied long ago on very thick alloy\nfilms.49Most S/F/S Josephson junctions have the sandwich structure due to the very short\nlength scale over which the critical current oscillates and decays in ferromagnetic materials.\nThe notable exceptions are junctions based on half-metallic CrO 250or on single-crystal\nferromagnetic nanowires,51where the supercurrent may propagate over very long distances.\nAll of the Josephson junctions analyzed in the Tables in Section IV have the sandwich\ngeometry.\nC. S/I/F/S vs S/F/S π-junctions\nJosephson junctions are characterized by several parameters. In addition to ∆ and IcRN,\nthese include the Josephson energy, EJ= (ℏ/2e)Icand the Stewart-McCumber parameter,\nβc= (2e/ℏ)IcR2\nNC, where Cis the junction capacitance. The value of βcdetermines whether\nthe junction dynamics are underdamped ( βc>1) or overdamped ( βc<1). Without an ex-\nplicit insulating tunnel barrier, all S/F/S Josephson junctions are overdamped. They can\nbe made underdamped by inserting an insulating barrier to make an S/I/F/S junction,3,52\nor by using a ferromagnetic insulator (FI) as the barrier material: S/FI/S.53–55The differ-\nence between overdamped and underdamped junction dynamics is crucial for all π-junction\napplications – both quantum and classical.\nA related issue relevant for applications is whether the π-junction is being used as an\nactive or passive device. In the former situation, the phase drop across the junction is an\n7important dynamical variable, whereas in the latter situation the π-junction acts as a passive\nphase shifter with the junction phase hardly deviating from πduring circuit operation. For\npassive applications, the π-junction is invariably paired up with conventional S/I/S junctions\nwhose critical currents are much smaller than that of the π-junction. That configuration is\ncommon in applications that require underdamped junction dynamics or large IcRNproduct\n– both of which are difficult to achieve with S/F/S junctions containing a metallic barrier.\nD. Dependence of critical current on applied magnetic field\nA common way to assess the quality of a Josephson junction is to measure the critical\ncurrent vs magnetic field applied in a direction perpendicular to the current flow, Ic(H).\nThe local supercurrent in a junction depends on the local vector potential; in the case of a\nuniform field, the vector potential varies linearly along the direction perpendicular to both\nthe current and applied field.56For the case of a small rectangular junction of width w, with\napplied magnetic field, H, parallel to one of the symmetry axes, the resulting interference\npattern follows the “sync” function familiar from the single-slit diffraction pattern in optics:\nIc(Φ) = Ic0\f\f\f\fsin(πΦ/Φ0)\n(πΦ/Φ0)\f\f\f\f(7)\nwhere Φ 0=h/2e= 2.07×10−15Tm2is the flux quantum for electron pairs, and Φ =\nµ0Hw(2λeff+dbarrier ) is the magnetic flux through the junction, where λeffis the effective\nLondon penetration depth of the S electrodes and dbarrier is the total thickness of the barrier.\nA small junction has wsmaller than the Josephson penetration depth and thus uniform\ncurrent distribution.56Because of the similarity with optics, a plot of such Ic(H) data is\ncolloquially referred to as the “Fraunhofer pattern.” The period of the Fraunhofer pattern\nin the applied field His inversely proportional to the width wof the junction in the direction\ntransverse to the applied field. For the case of circular junctions or elliptical junctions with\nthe field applied along a symmetry axis, the interference pattern follows an Airy function,56\nbut we still refer to the data as the Fraunhofer pattern out of habit.\nThe introduction of magnetic materials inside the junction leads to complications due to\nthe field produced by the magnetic material; the Fraunhofer pattern is modified and may\nbe severely distorted. In principle, if one knew the exact configuration of the magnetization\nMas a function of H, one could calculate the vector potential everywhere and deduce the\n8Ic(H) pattern.57,58The converse is not true, however, except in the simplest cases discussed\nbelow. Consider first the worst situation, which is an S/F/S Josephson junction of relatively\nlarge lateral size (more than a few µm) containing a thick layer of a strong F material such\nas Ni, Fe, Co, or Gd. Due to the complicated multi-domain magnetic configuration of the F\nmaterial, the Ic(H) pattern has a random-looking set of peaks and valleys with no discernible\nperiod and no clear central peak.59–61In such a situation it is neither possible to calculate the\ndomain structure from the Ic(H) data nor to deduce the maximum critical current density\nJcinside the junction.\nFortunately, there exist many situations that produce well-behaved Fraunhofer patterns\nfrom which one can extract key junction parameters – in particular the maximum value of\nJc. An illustrative example is the early work of Ryazanov on a Josephson junction contain-\ning a weakly-ferromagnetic CuNi alloy.62That material has a mild perpendicular magnetic\nanisotropy (PMA),63which is advantageous because the vector potential corresponding to\nan out-of-plane magnetization points in the plane and does not affect the Fraunhofer pat-\ntern. In addition, if both the magnetization and domain size are small, then the vector\npotential effectively undergoes a random walk throughout the sample and never gets very\nlarge in magnitude. For those reasons, Ryazanov’s CuNi virgin-state junction produced an\nexcellent Fraunhofer pattern centered at zero field. After the junction was magnetized in a\nlarge in-plane field, the Fraunhofer pattern was shifted in field with only minor distortion,\nindicating that the CuNi alloy had acquired a nearly-uniform in-plane component to its\nmagnetization. After demagnetizing the sample, the Fraunhofer pattern reverted back to\nits original shape centered at the origin.\nEven in junctions containing strong F materials, the Fraunhofer pattern is well-behaved\nif the magnetization of the material is uniform and parallel to the direction of the applied\nfieldH. If it stays uniform over the full measured field range, then Eqn.(7) is still valid but\nwith a total magnetic flux that includes the contribution from the F material:\nΦ =µ0Hw(2λeff+dF) +µ0M(1–Ndemag)wdF (8)\nwhere Mis the magnetization of the F material and Ndemag is the demagnetizing factor\ndetermined by the shape of the F layer. Most workers using Eqn.(8) ignore the demagnetizing\nfactor, which is small for thin oblong-shaped F layers. Note that Eqn.(8) neglects any\nmagnetic flux that returns inside the junction – a quantity that is difficult to calculate\n9FIG. 1. IcRNvsµ0Hfor Josephson junctions containing a PdFe alloy with 3% Fe with thicknesses\nlabeled in the figure. Junctions have elliptical shape and lateral dimensions of ≈1.3×0.55µm2.\nThe data acquired before the field at which the PdFe magnetization vector reverses direction (solid\nmarkers), and the corresponding fits (lines) show good agreement for both the positive (red, dashed)\nand negative (blue) field sweep directions. The hollow circles are the corresponding data points\nafter the PdFe magnetization switches. These Fraunhofer patterns display magnetic hysteresis\nand are increasingly shifted with larger dF. Adapted with permission from IEEE Trans. Appl.\nSupercond. 27, 1800205 (2017).64\naccurately. Setting the total flux Φ = 0 shows that the central peak of the Fraunhofer\npattern is shifted in field by an amount:\nHshift=−Md F\n2λeff+dF(9)\nwhere we have neglected Ndemag. Note that the denominator should also include the thick-\nnesses of any other non-superconducting layers inside the junction – e.g. the Cu spacers\nused in most of the junctions studied by the authors.\nWe illustrate Eqn. 9 with data from junctions containing a PdFe alloy with about 3% Fe,\nshown in Figs. 1 and 2. We have chosen this data set because this weakly-ferromagnetic alloy\nhas a large value of ξF(several nm, see Table III), hence thickness fluctuations are negligible\nand there is very little scatter in the data of IcRNvsdF. Nevertheless, the magnetization is\nlarge enough to shift the Fraunhofer patterns considerably. The PdFe thickness where IcRN\nhas its peak value in the πstate is about 24 nm. At that thickness, the Fraunhofer pattern\n101101000\n5 1 01 52 02 53 03 5051015200 \n IcRN (µV)a)b\n)π \n µ0Hshift (mT)d\nPdFe (nm)FIG. 2. Characterization of Josephson junctions containing a weakly ferromagnetic PdFe alloy. a)\nThe maximal IcRNproduct of the junctions versus PdFe thickness. The minimum indicates the\nPdFe thickness at which the junctions transition between the 0 and π-phase states. The solid line\nis a fit of the data to Eqn. 6. b) The field shift of the Ic(B) data (the “Fraunhofer pattern”) vs\nPdFe thickness. The field shift increases with dPdFedue to the increasing total magnetic moment\ninside the junctions and is fit to Eqn. 9. Reproduced with permission from IEEE Trans. Appl.\nSupercond. 27, 1800205 (2017).64Copyright 2017, IEEE.\nis shifted by about 14 mT, and the value of Icmeasured at zero field is significantly smaller\nthan the peak value extrapolated from the fit, Figure 1 (c). Hence this material would not\nbe ideal for π-junction applications unless the junctions were considerably narrower in their\nlateral dimension, with correspondingly wider Fraunhofer patterns.\nIII. APPLICATIONS OF π-JUNCTIONS\nSeveral applications of Josephson junctions involve superconducting loops threaded by\nhalf a quantum of magnetic flux, which has two degenerate ground states with supercurrent\ncirculating in opposite directions. The fundamental motivation for using a π-junction in\nplace of the flux bias is to improve the circuit performance in some way, e.g. by reducing\nnoise, reducing the size of the circuit (e.g. by eliminating the need for flux bias lines or a\nlarge geometrical inductance in the loop), or increasing the operating margins.\nEven before Ryazanov’s 2001 experimental breakthrough with S/F/S junctions, there\n11already existed experimental realizations of π-junctions without magnetic materials. The\nmost famous example involved the demonstration of the d-wave character of the high-T c\nsuperconductors, where it was shown that a phase shift of πis created by a Pb/YBCO/Pd\ndouble-junction based on two orthogonal crystal faces of a YBa 2Cu3O7single crystal,65or\nby three YBCO grain-boundary junctions in series.66π-junctions can also be realized in 4-\nterminal S/N/S junctions driven out of equilibrium by a control current flowing orthogonal\nto the supercurrent direction67–69or in 3-terminal junctions with the control current injected\ninto the junction.70–72While the nonequilibrium π-junctions demonstrate fascinating physics,\nwe would argue that they are not likely to be suitable for practical applications.\nA. π-junctions in superconducting qubits\nThe results from d-wave superconductors65,66inspired Ioffe and co-workers to propose\nusing π-junctions in superconducting phase or flux qubits for quantum computing.16,17\nA conventional flux qubit consists of a superconducting loop containing three Josephson\njunctions.73,74Under an applied flux bias of exactly Φ 0/2, states with clockwise and coun-\nterclockwise supercurrent flow are degenerate, and the symmetric and antisymmetric linear\ncombinations of those states can form the two states of a qubit. Replacing the Φ 0/2 flux\nbias with a π-junction is beneficial in several ways. In systems with many qubits the Φ 0/2\nflux bias can be provided either by a global magnetic field or by individual flux lines for\neach qubit. The former suffers from field inhomogeneity and fluctuations in loop area among\nthe qubits while the latter suffers from crosstalk between neighboring qubits and the cor-\nresponding need to calibrate every qubit flux precisely. Both suffer from fluctuations in dc\nsupply current. Replacing the flux bias with π-junctions alleviates those issues. In addition,\nremoving the large static flux bias means that the qubits can be made much smaller, thereby\nreducing sensitivity to magnetic field fluctuations and improving prospects for large-scale\nintegration. We note that individual flux bias lines are still needed for qubit operations, but\nthey supply only small, pulsed currents; there is no need for a large static dc current.\nThe idea to use π-junctions in various types of qubits has been developed further by\nseveral authors,75–81and is being pursued to this day.82–85While in the early proposals, the\nπ-junction was considered to be an active device,75–77later works emphasized the advantages\nof using the π-junction as a passive phase-shifter,78,80,81,85where the only constraint on the\n12critical current of the π-junction is that it be much larger than the critical currents of the\nconventional S/I/S junctions in the loop.\nA crucial issue confronting all types of qubits is decoherence. Kato et al. published the\nfirst calculation of the relaxation and dephasing rates of a flux qubit containing a ferro-\nmagnetic π-junction as a passive phase shifter.78Those authors state explicitly that only\nunderdamped π-junctions with very large critical current density have sufficiently long de-\nphasing times to be suitable for qubit applications. Nevertheless, the first demonstration\nof a ferromagnetic π-junction in a qubit, carried out by Feofanov et al. in 2010,80used\nan overdamped S/F/S π-junction in series with a standard underdamped S/I/S junction to\nform a phase qubit. The choice of using a phase qubit rather than a flux qubit was for device\nsimplicity. Feofanov et al. compared the properties of their π-junction qubit with a reference\nqubit without the π-junction. While there was no noticeable decrease in the decoherence\ntime of the π-junction qubit relative to the reference qubit, both qubits had coherence times\nof only a few nanoseconds. Calculation of the expected decoherence time using the theo-\nretical framework of Kato et al.78gave a result in the nanosecond range, but that theory\nwas not intended to describe the phase qubit of Feofanov et al. Later, Shcherbakova et al.81\nfabricated flux qubits with and without a π-junction and showed explicitly the shift of the\nmagnetic-field dependence of the qubit energy splitting by exactly half of a flux quantum,\nbut they did not measure the decoherence times of their qubits.\nVery recently, Kim et al. fabricated and measured flux qubits with and without an S/F/S\nπ-junction.85Theπ-junction was used as a passive phase shifter with a critical current much\nlarger than those of the conventional S/I/S junctions in the qubit loop. The coherence time\nof the standard qubit was more than 1,000 times larger than that of the phase qubit studied\nearlier by Feofanov et al.80, making this a much more stringent test of how the π-junction\naffects qubit performance. Kim et al. found that the energy relaxation time ( T1) of the\nflux qubit decreased from about 15 µsto 1.5µsupon insertion of the π-junction, while the\nvalue of T2decreased from about 11 µsto 1.5µs. The authors attribute the decreases to\ndissipation in the S/F/S junction due to quasiparticle excitations. While these results are\nsomewhat disappointing, they are not surprising given the insistence by Kato et al. that\nonly underdamped π-junctions are suitable for use in quantum circuits, even when the π-\njunction is used as a passive phase shifter.78Kim et al. suggest using a π-junction based on\na ferromagnetic insulator (FI) rather than a metallic ferromagnet, and we certainly agree.\n13Unfortunately, the only FI that has been studied extensively in Josephson junctions is GdN,\nwhere a number of complex behaviors have been observed. We delay our discussion of GdN\nJosephson junctions to Section V of this review. Another possibility would be to use an\nS/I/F/S junction, if the critical current density can be made large enough and the damping\nsmall enough to satisfy the requirements laid out by Kato et al.\nNowadays it is possible to fabricate so-called “fluxonium” qubits with coherence times as\nlarge as a millisecond.86Like traditional flux qubits, fluxonium qubits also have an optimal\nworking point when flux biased at Φ 0/2.87It would be very informative to repeat the ex-\nperiment of Kim et al. on a fluxonium qubit, but with an underdamped S/FI/S or S/I/F/S\nπ-junction in place of Kim’s S/F/S π-junction.\nB. π-junctions in superconducting digital logic\nA more promising application of π-junctions, in our opinion, is in superconducting digital\nelectronics.88,89There have been numerous proposals for using πphase shifters in supercon-\nducting digital circuits, starting with the pioneering papers by Terzioglu and Beasley in\n199814and Ustinov and Kaplunenko15in 2003, and continuing to the present.90–108The\ninitial motivation for introducing π-junctions into single-flux-quantum (SFQ) digital logic\ncircuits was to eliminate the need for the large geometrical loop inductances needed to store\nsingle flux quanta, thereby reducing the circuit size substantially.15Subsequently it was ar-\ngued that incorporating π-junctions could also alleviate the need for current bias lines and\nincrease circuit operating margins.90More recently, entire new families of superconducting\ndigital logic incorporating π-junctions are being explored, as we will see below.\nThe distinction between active and passive roles of the π-junction is crucial in this context\nof classical digital circuits, just as it was in qubits. The speed at which a Josephson junction\nswitches into the voltage state is limited by the time τswitch =ℏ/(eIcRN). Metallic S/F/S\nπ-junctions without an insulating barrier have very low values of IcRN, hence they are\nnot suitable as switching elements in high-speed circuits. The easiest way around that\nproblem is to use S/F/S junctions only as passive phase shifters embedded in circuits that\ncontain conventional S/I/S junctions.80The critical current of the S/F/S phase shifter must\nbe considerably larger than that of the S/I/S junctions, so that the S/F/S junction never\nswitches into the voltage state and its phase drop stays close to π. An alternative approach\n14is to develop π-junctions with a large IcRNproduct. That is not trivial; simply adding an\ninsulating barrier increases the value of RNby a large factor but simultaneously decreases\nIc– usually by a larger factor.52An ingenious solution is to insert a thin superconducting\nlayer between the insulating barrier and the F layer, to get S/I/s/F/S.109–112Depending on\nthe thickness of the “s” layer, the S/I/s/F/S behaves either as a single Josephson junction\nwith large IcRNproduct, or as a series combination of a standard S/I/S junction and an\nS/F/S junction.\nThe first laboratory demonstration of a single-flux-quantum (SFQ) digital logic circuit\ncontaining intrinsic πphase shifts was reported by Ortlepp et al.90using high-T c/low-T chy-\nbrid Josephson junctions.113Theπphase shifts were introduced into superconducting loops\nby using orthogonal facets of the YBCO a-b plane, as was done in the original measure-\nment of the d-wave symmetry in YBCO.65Ortlepp et al.90demonstrated proper functioning\nof a toggle flip-flop circuit, and they used circuit simulations to compare two identically-\nfunctioning circuits with and without πphase shifters. The circuit with the phase shifters\ncontained fewer current bias lines, took up considerably less area, and had wider operating\nmargins – i.e. better stability against variations in circuit parameters. Shortly thereafter,\nBalashov et al.114demonstrated an all-niobium toggle flip-flop with a πphase shifter based\non the “trapped flux” method.115Theπphase shifter replaces a large geometrical induc-\ntance, hence diminishes the circuit size substantially. Using the same type of trapped-flux\nπ-shifter, Wetzstein et al. compared the bit error rates of toggle flip-flops with and without\naπphase shifter, and concluded that the former had improved robustness against circuit\nnoise.93\nThe first demonstrations of all-Nb SFQ circuits containing ferromagnetic π-junctions were\nreported by Feofanov et al.80and by Khabipov et al.92in 2010. Those works represent a\nmajor milestone because all modern large-scale SFQ circuits are fabricated using only Nb-\nbased superconductors, and because they represent the first example of a scalable process\nfor incorporating π-junctions into SFQ circuits. In comparison to trapped-flux π-shifters,\nS/F/S junctions are more compact and do not require a delicate initialization procedure.\nBoth groups used the π-junction as a passive phase shifter in a circuit containing conventional\nS/I/S switching junctions. Both groups identified circuit size as the main advantage of using\nπ-junctions.\nThere has been a surge of interest in π-junctions in the past few years in Japan. The\n15Yoshikawa group at Yokohama University has shown through simulations that including\nπ-junctions in certain SFQ circuits reduces the junction count significantly, while the new\ncircuits have wider operating margins and lower static power dissipation that their conven-\ntional counterparts.95,98They and others propose similar advantages using π-junctions in\nultra-low-power adiabatic quantum flux parametron circuits.97,106,108The Fujimaki group at\nNogoya University recently invented a new family of superconducting logic called “half-flux-\nquantum” logic, where information propagates in the form of voltage pulses with area equal\nto Φ 0/2 rather than Φ 0as in standard SFQ circuits.94,96,100–104The basic unit of half-flux\nquantum logic was originally a 0 −πSQUID, where the 0-junction and π-junction in the\nSQUID must have similar critical currents and switching properties. The inventors soon\nrealized, however, that it is more practical to use a 0 −0−πSQUID instead, where the\nπ-junction is used as a passive phase shifter and never switches into the voltage state. Half-\nflux-quantum circuits offer several advantages compared to standard SFQ circuits, including\nsmaller bias currents, smaller circuit size, and lower power dissipation.\nAnother new logic family, called “phase logic”, was introduced very recently by the group\nat Moscow State University.105,107A major selling point of phase logic is the absence of induc-\ntors, which take up much of the space in standard SFQ circuits and cause crosstalk between\nneighboring cells, thereby being a major hindrance to scaling those circuits to smaller sizes.\nIn the original proposal, phase logic circuits were based on bistable Josephson junctions\nwith a dominant second harmonic in their current phase relation: Is(ϕ) =Icsin(2ϕ). The\nexistence of such junctions has been demonstrated in special situations,116,117and proposed\nto exist in others,118–121but we emphasize that they are not available as “off the shelf” com-\nponents ready for use in electronic circuits. Hence it was encouraging when Maksimovskaya\net al.107suggested a way to construct phase logic circuits using standard π-junctions. Phase\nlogic appears to be a promising avenue for scaling superconducting circuits to higher densi-\nties, and is being pursued also by circuit designers in the US.122–125\nWe note that the 0 −0−πSQUIDs that form the basic unit of the Nagoya group’s\nhalf-flux-quantum logic are equivalent to a single junction with a negative second harmonic\ncurrent-phase relation: Is(ϕ) =−Icsin(2ϕ).105The ground state of such a SQUID is doubly-\ndegenerate, with a phase shift of ±π/2 across the SQUID. In contrast, the basic unit of the\nMoscow group’s original phase logic proposal is a junction with a positive second harmonic\ncurrent-phase relation, which puts its ground state phase at either 0 or π. Those two\n16situations have very different consequences for circuit design.\nC. Controllable junctions: superconducting memory\nJosephson junctions with controllable properties are suitable for applications in super-\nconducting memory. The controllable property can be either the magnitude of Icor the\nphase state of the junction, but in the former case the junction must be active, i.e. it must\nswitch into the voltage state when the applied current surpasses Ic, which means it must\nhave a large IcRNproduct for high-speed operation, as discussed earlier.\nThere are several proposals for controlling the magnitude of Icin Josephson junctions\ncontaining a single magnetic layer. The Ryazanov group developed controllable S/F/S junc-\ntions containing a weakly-ferromagnetic PdFe alloy.126In a collaboration with the Hypres\nCorporation in the US, that idea was extended to controllable S/I/s/F/S junctions with a\nlarge IcRNproduct.109,110Control of Icwas achieved by the orbital effect – i.e. by shifting\nthe Fraunhofer pattern of the junction via partial magnetization of the PdFe. A disadvan-\ntage of that approach is that it does not scale well as the junction size is reduced. The\njunction size can be reduced in one in-plane dimension, but must remain large in the other\ndimension to achieve adequate Fraunhofer shift.127A recent proposal to control Icvia the\nstray field produced by the magnetic texture of a Co disk also relies on the orbital effect,\nbut the constraints on junction size are less severe because of the large magnetization of the\nCo.128\nSeveral other schemes for controlling Icof Josephson junctions have been proposed, and\nwe do not try to review all of them here. We mention one based on placing Abrikosov flux\nvortices in close vicinity to a Josephson junction,129one based on rotating the magnetization\nin a combination S/F/S and S/F/N/S device,130and another based on a multiterminal S/F\ndevice.131,132\nThe possibilities become richer when there are two independent F layers inside the junc-\ntion – one with fixed magnetization and the other with a magnetization that can be reori-\nented by application of a small external field. Either the magnitude of Icor the phase state\nof the junction can be controlled in such junctions, as predicted in a number of theoretical\nworks.133–141The first experimental verification of Icmodulation in such “spin-valve” junc-\ntions was performed by Bell et al. in 2004, using Co and Ni .80Fe.20(NiFe) as the fixed and\n17free F layer materials, respectively.142A similar experiment was revisited a decade later in a\nmuch more thorough study by Baek et al. ,143and also by Qadar et al. .144Those groups used\nNiFe-based alloys as the free layer because of their excellent magnetic switching properties,\neven though they are not so great for supercurrent transmission (see Table II). Baek et al.\nused Ni as the fixed layer material since Ni has the best supercurrent transport of any strong\nF material (see Table I in Section IV).\nControlling the phase state of the junction rather than the amplitude of Ichas two\nadvantages. First, there is no need for a large IcRNproduct, which is difficult to achieve in\nall-metal junctions. Second, the junction phase is intrinsically a digital rather than analog\nquantity; it can take only one of the two values zero or π, as long as the F-layer is far from\nthe thickness corresponding to the 0 −πtransition. The first demonstration of phase control\nin spin-valve junctions was carried out by one of the authors and his students,145,146using\nNi and NiFe for the fixed and free F-layer materials. Similar junctions formed the basis for\nNorthrop Grumman’s “Josephson Magnetic Random Access Memory” or JMRAM, which\nwas demonstrated on a 2 ×2 memory array in 2017.147A later demonstration on an 8 ×8\nmemory array revealed a difficulty with this approach, namely the inconsistent behavior of\nthe magnetic layers in different memory cells. The Ni fixed layers are multi-domain at the\nµm-scale lateral dimensions of the junctions, and require a very large initialization field to\nmagnetize completely.143And the switching fields of the NiFe free layers varied significantly\nfrom cell to cell. A great deal of materials research would be needed to turn this idea into\na successful memory technology.\nFinally, we mention proposals that involve intrinsic bistability of Icor of the junction\nphase state without requiring any change in the magnetic configuration.148,149\nIV. METALLIC π-JUNCTIONS\nThe Tables in this section are intended for the reader to make quick comparisons between\nmaterials and works. We attempt to include as much of the available literature as possible,\neven where fit parameters could not be extracted from the data; hence there are gaps in the\nTables. We present four Tables on single F layer junctions: I – pure elements, II – NiFe and\nNiFe-based alloys, III – weak ferromagnetic alloys, and IV – strong and other ferromagnetic\nalloys. Within each Table, materials are grouped together, and where there are multiple\n18works on the same material, the works are listed chronologically.\nThe methodology for putting together the Tables in this section is as follows. Fit param-\neters are either extracted from the original publications, or datasets are extracted from the\nplotted graphs, fit to Eqn. 6 and the best fit parameters are reported in the Table. We fit\nthe expression only to the nominal thickness of the ferromagnetic layer, ignoring magnetic\ndead layers and any role of additional buffer layers. The reported uncertainties in the orig-\ninal datasets are not taken into consideration for the fitting, so we quote all fit parameters\nto 2 significant figures without consideration of their uncertainties. For more rigorous data\nanalysis, the reader should refer to the published works, particularly for cases where Eqn. 6\ndoes not provide the best fit to the experimental data. We attempt to note in the main\ntext and figure captions as much as possible where subtleties exist in the experimental data\nwhich are not accounted for in the fitting here.\nThe parameters we report in the Tables include the fit parameters from Eqn. 6, corre-\nsponding to the decay and oscillation lengths ξF1andξF2, the thickness at which the first\n0-πtransition occurs, d0−π, and the fictitious zero thickness fit parameter, IcRN(0), which\nis necessary for the reader to reproduce the fits but doesn’t have a physical meaning. (Junc-\ntions without the ferromagnet will not have the extracted dF= 0 critical current as there\nare fewer interfaces in such devices.) We also extract the critical current at the peak of the\nπ-state, which is a useful parameter for device consideration and for comparing materials.\nWhen the range of thicknesses studied is not broad enough to cover multiple 0- πoscilla-\ntions (multiple minima in IcRN), the ξF2fitting parameter ambiguous, in which case that\nparameter will be missing from the Tables.\nThere are two common conventions for reporting the magnitude of critical Josephson\ncurrent. The first is the product of the critical current and normal state resistance, IcRN,\nquoted here in µV. The second convention is to report the critical current density, Jc,\nquoted here in kAcm−2. The main difference between IcRNandJcis that the junction\narea need not be known to report IcRN, which may be advantageous for a study where\nmultiple junction sizes are fabricated (intentionally or unintentionally). Where the original\npublications contain all the information necessary to perform the conversion, we report\nboth quantities at the peak of the π-state in the Tables. Otherwise, the reader can consider\nthe approximation that for metallic S/F/S junctions the product of area and normal state\nresistance, ARN≈10−14Ωm2, which arises primarily from the boundary resistances between\n19the S electrodes and the F layer.150Hence, for S/F/S junctions, one can estimate Jcby\ndividing IcRNbyARN(in the Tables, IcRNvalues quoted in µV can be multiplied by 10\nto get Jcin kAcm−2assuming ARN≈10−14Ωm2). For junctions containing an insulating\nlayer, such as the S/I/F/S geometry, the story is very different: ARNis typically several\norders of magnitude larger, and varies widely depending on the oxidation conditions used\nto produce the insulating barrier.\nA. Pure elements\nIn Table I, we highlight works studying Josephson junctions with pure elements.\nPure elemental Ni was among the first ferromagnetic barrier layers studied.151Ni has the\nlowest bulk magnetization of the transition metal ferromagnets (550 emu/cm3). Experimen-\ntally, it is found that Ni has favorable supercurrent carrying properties - the decay length is\nrelatively long for a strong ferromagnet and the IcRNis high. These favourable properties\nhave resulted in a number of reports on Ni Josephson junctions.143,147,151–157,160,162,163It is\nworthwhile noting that while Table I is obtained by fitting to Eqn. 6 which describes dif-\nfusive and intermediate limit transport, Ni junctions have been reported in the literature\nin both the diffusive154and ballistic155limits. While Ni may have favourable supercurrent\ncarrying properties, the magnetic properties of thin Ni layers are far from ideal. Ni layers are\nmagnetically ‘hard’, requiring a large initialization field to set the magnetization direction\nand a large switching field. It is well known that Ni films break up into small magnetic\ndomains, which sets a limit on single domain junction size, and can contribute stray fields\ninto the surrounding superconducting layers.\nFor determining the position of the 0 −πtransition for Ni from Table I, we guide the\nreader towards Refs. 155 and 147, which have the smallest increments in thickness around\nthe transitions. The data of Baek et al. covers both the first 0 −πandπ−0 transitions, which\noccur at 0.90 and 3.4 nm respectively.155The first π−0 transition thickness is confirmed by\nDayton et al. who observe it at 3.3 nm, again taking small increments in thickness around\nthe transition.147Most works on Ni do not study the very thin layers of Ni in the range\nof Baek et al. ’s observed 0 −πtransition, and we note that in some cases it may still be\npossible to model the data in those works assuming d0−π= 0.9 nm.\nCritical current oscillations have also been reported in pure elements of Co153,159,160,162\n20BarrierξF1\n(nm)ξF2\n(nm)d0−π\n(nm)IcRN(0)\n(µV)IcRN(π)\n(µV)Jc(π)\n(kAcm−2)dead layer\n(nm)Ref.Comments\nCu/Ni/Cu 3.8 0.98 2.6 2.8 1.0 10 151\nCu(Au)/Ni/Cu(Au) 1.7 0.45 152 ⃝\nNi 4.5 1.2 4.0 450 110 1.3 153\nAl2O3/Cu/Ni 0.66 0.53 3.0 3.7 3.4×10−32.3 154\nCu/Ni/Cu 2.4 0.79 0.90 330 ≥100 0.7 155\nCu/Ni/Cu 3.3⋆200 147 ⃝,†\nNi >30 >300 156 †\nNi & Cu/Ni 157 †\nCu/Co/Cu 158 †\nCo 2.1 0.29 1.0 130 65 0.8 153\nRh/Co/Rh 0.81 0.21 0.64 260 82 1600 0.0 159\nCo/Ru/Co 1.2 2.9 61 ‡\nCu/Co/Ru/Co/Cu 2.3 0.90 61 ‡\nFe 5.4 0.25 0.81 170 140 1.1 160\nAl/Gd/Al 1.2 20 59 ‡, 1 K\nHo 4.3 2300 2.0 161 ‡\nTABLE I. Extracted parameters for Nb based Josephson junctions with pure elements as the\nferromagnet in the barriers. The fitting methodology to Eqn 6 is given in the main text and best\nfit values are quoted to 2 significant figures without consideration to their uncertainty so should\nbe treated as approximations. The full barrier layers, including normal metals and insulators,\nare listed. Unless otherwise commented, the measurements were performed at 4.2 K. Comments:\n⃝The range of thickness values did not cover the thickness where the first 0- πtransition was\npredicted. †The range of thickness values studied was not sufficient for a meaningful fit to theory.\n‡No 0- πoscillations were observed or claimed, and the data are fitted here with only the exponential\ndecay component of Eqn 6. ⋆This thickness corresponds to the first π-0 transition.\nand Fe,160,164although they are less studied than Ni. Potential advantages of Co and Fe are\nthat thin films tend to be magnetically softer than Ni, with larger magnetic domains. An\nunexplored potential is that thin layers of Co can have perpendicular magnetic anisotropy\nwhen deposited on Pt or Pd normal metal spacer layers.\n21Finally we mention that the rare-earth metals Gd59and Ho161have been studied in\nJosephson junctions, including the thickness series to extract the decay length, however\nno critical current oscillations attributed to the 0- πtransition were observed. In the case\nof Gd, this may be due to the thickness increments (step size) between samples in that\nearly work being too large to observe the oscillation. Ho has been extensively studied as\na spin-triplet generator due to its intrinsic magnetic inhomogeneity,165which may make it\nunsuitable as a π-junction. Nb/Gd multilayers have been extensively studied for the obser-\nvation of critical temperature oscillations,33which can guide future junction experiments.\nIt is additionally noteworthy that Nb can seed rare-earth ferromagnet growth and that the\ninterface is considered to be sharp with minimal interdiffusion.166–168\nB. NiFe alloys\nIn Table II, we highlight works studying Josephson junctions with NiFe and NiFe-based\nalloys.\nNi.80Fe.20(hereafter NiFe) is a strong ferromagnetic alloy which is optimized for magnetic\nswitching at small magnetic fields, making it an excellent choice in spintronics research as\na soft magnetic layer. In many papers, NiFe is referred to as permalloy or Py. Josephson\nstudies of NiFe were well motivated by the spintronics community, and it has been used\nas the free layer of choice in superconducting spin valves.142,145,147Additional functionality\noffered by NiFe is the ability to set the direction of in-plane magnetic anisotropy by applying\na magnetic field during growth.\nFrom studying Table II, the reader will notice two possible positions of d0−πfor NiFe,\neither between 1 .3−1.8 nm or at 2.4 nm. We guide the reader towards the work of Dayton et\nal.,147which takes the smallest increments in thickness for NiFe around the 0 −πtransition\nand concludes that d0−π= 1.66 nm, falling in the middle of the 1 .3−1.8 nm range reported\nby other works.147\nAlthough the switching properties of NiFe at room temperature are excellent, the switch-\ning degrades at low temperatures relevant to Josephson junctions. To address this limita-\ntion, NiFe-based alloys with even lower coercivity have been developed174,175and tested in\nJosephson junctions, such as Ni .73Fe.21Mo.06(supermalloy), which we include also in Table\nII. Unfortunately, the decay length inside NiFe-based alloys is experimentally found to be\n22BarrierξF1\n(nm)ξF2\n(nm)d0−π\n(nm)IcRN(0)\n(µV)IcRN(π)\n(µV)Jc(π)\n(kAcm−2)dead layer\n(nm)Ref.Comments\nCu/NiFe/Cu 1.0 0.76 2.4 360 12 170 0.7 169 #\nNiFe 1.6 0.33 1.3 240 78 0.5 153\nNiFe 1.2 0.61 2.4 17 1.2 144\nCu/NiFe/Cu 1.5 0.58 1.8 69 12 160 0.0 170\nCu/NiFe/Cu 1.66 25 147 †\nAl/AlO x/Nb/NiFe 112 †\nNiFe 171 †\nCu/NiFe/Cu 0.71 0.74 1.5 250 10 100 172\nCu/Ni .73Fe.21Mo.06/Cu 0.48 0.96 2.3 150 0.20 1.8 0.2 173\nCu/Ni .65Fe.15Co.20/Cu 1.1 0.48 1.2 30 5 67 0.0 170\nCu/Ni/NiFe/Ni/Cu 0.64 0.78 0.92 800 50 530 172\nTABLE II. Extracted parameters for Nb based Josephson junctions with NiFe and NiFe-based\nalloys as the ferromagnet in the barriers. The fitting methodology to Eqn 6 is given in the main\ntext and best fit values are quoted to 2 significant figures without consideration to their uncertainty\nso should be treated as approximations. The full barrier layers, including normal metals and in\none case an additional ferromagnetic layer, are listed. Unless otherwise stated in the Table, the\ncomposition was Ni .80Fe.20and the measurements were performed at 4.2 K. Comments: # The\nbottom Nb electrode and barrier layers were epitaxial. †The range of thickness values studied was\nnot sufficient for a meaningful fit to theory.\neven shorter than NiFe, which reduces considerably the potential critical Josephson current\nof aπ-junction based on these materials. This is likely due to the short electron mean free\npath in these alloys.\nOther NiFe-based alloys which have been explored for use in Josephson junctions include\nNi.70Fe.17Nb.13, which was used by Baek et al. as the free layer in a spin-valve device.143\nIn later works, the same group used a more dilute non-magnetic concentration of the alloy\nas a normal metal spacer layer in junctions containing ferromagnetic Ni.155,163Qader et al.\ncharacterized Cu 1-x(Ni.80Fe.20)xas a possible barrier layer, however single layer Josephson\njunctions with this alloy have not been reported.176\nThe work of Bell et al. is noteworthy because the bottom Nb electrode and barrier layers\n23were grown epitaxially.169This work may, therefore, not be directly comparable to the other\nworks in Table II. As far as we are aware, this is the only work listed in any of the Tables to\nstudy an epitaxial barrier. Epitaxy plays a role in improving the mean free path of electrons\nthrough metals, which may improve the supercurrent carrying properties of junctions with\nepitaxial barriers. Epitaxy has not yet been systematically studied for π-junctions and may\nprovide fertile ground for future exploration.\nC. Weak ferromagnetic alloys\nIn Table III, we highlight works studying Josephson junctions with weak ferromagnetic\nalloys.\nIn the context of ferromagnetic Josephson junctions, weak ferromagnet refers to materials\nwith a low Curie temperature, and therefore low exchange energy. The structural and\nmagnetic properties of weak ferromagnetic alloys was reviewed in detail by Kupriyanov,\nGolubov, and Siegel.27\nFrom Table III, the reader will notice that although CuNi alloys have been widely studied,\nthey are far from an ideal material as a π-junction. The CuNi alloys are firmly in the diffusive\nlimit where ξF1< ξ F2, which in comparison to other materials presented in the Tables is\nsomewhat of a rare case. Superconducting junctions with weak ferromagnets are a system\nwhere the theoretical Usadel equations are valid, and therefore it is possible in these systems\nto directly test predictions of that theoretical framework. Oboznov et al. apply the Usadel\nequations to model their data, considering that their junctions showed strong spin dependent\nscattering, attributed to magnetic inhomogeneities in the barrier caused by the formation\nof Ni-rich clusters.178Further work studying Nb/CuNi bilayers using scanning tunneling\nspectroscopy support this idea, finding significant spatial variation in the thickness and\ncomposition of Ni in the CuNi layer.184In comparison to CuNi, the Pd-based alloys appear\nmuch more promising as potential π-junctions due to the larger critical Josephson currents\nobserved in the π-state.\nOther works on weak ferromagnetic alloys that are not included in Table III but which\nwe wish to highlight include an experiment where CuNi junctions were fabricated where the\nbarrier layer has a thickness step, such that junction has both 0 and πparts.185Further work\non CuNi includes observation of a positive second harmonic in the current-phase relation in\n24BarrierξF1\n(nm)ξF2\n(nm)d0−π\n(nm)IcRN(0)\n(µV)IcRN(π)\n(µV)Jc(π)\n(kAcm−2)dead layer\n(nm)Ref.Comments\nCu.48Ni.52/Cu 2 ▷ ◁\nCu.52Ni.48 177 ▷ ◁\nCu.47Ni.53/Cu 1.3 3.7 11 2200 0.15 1 4.3 178\nAl2O3/Cu .40Ni.60 0.78 1.35 5.2 400 5.0×10−33.1 52 2.11 K\nAlO x/Cu .40Ni.60 1.2 5.8 1.5×10−2179\nAlO x/Nb/Cu .40Ni.60 1.2 7.2 6.5×10−2179\nCu.47Ni.53 1.3 4.3 7.4 3200 2.0 20 180\nNbN/Cu .40Ni.60/NbN 1.8 2.0 5.6 160 2.2 8.8 181\nCu.48Ni.52/Al 1.3 3.2 15⋆>4.2 >60 182 ⃝\nAl/Al 2O3/Pd .88Ni.12 2.9 2.3 6.5 450 18 3 1.5 K\nPd.88Ni.12 7.7 4.4 7.4∗100 60 ⃝\nNbN/Pd .89Ni.11/NbN 4.4 3.6 8.5 540 ≥18 ≥71 183\nPd.89Ni.11/Al/AlO x 22 2.2×10−2103 †\nAl/AlO x/Nb/Pd .99Fe.01 109 †\nPd.97Fe.03 16.2 7.2 16 100 21 190 2.8 64\nPtxNi1-x 157 ▷ ◁\nTABLE III. Extracted parameters for Nb based Josephson junctions with weak ferromagnetic alloy\nbarriers. In Refs. 181 and 183, the superconductor was NbN, as indicated in the Table. The fitting\nmethodology to Eqn 6 is given in the main text and best fit values are quoted to 2 significant\nfigures without consideration to their uncertainty so should be treated as approximations. The full\nbarrier layers, including normal metals and insulators, are listed. Unless otherwise commented,\nmeasurements were performed at 4.2 K. Comments: ▷ ◁The main focus was to study the temper-\nature and/or compositional variation of the Josephson effect instead of the thickness dependence.\n†The range of thickness values studied was not sufficient for a meaningful fit to theory. ∗This\nvalue was extrapolated from measurements on junctions outside of this thickness. ⃝The range of\nthickness values did not cover the thickness where the first 0- πtransition was predicted. ⋆This\nthickness corresponds to the first π-0 transition.\njunctions with a CuNi thickness close to the 0 −πtransition.116\n25BarrierξF1\n(nm)ξF2\n(nm)d0−π\n(nm)IcRN(0)\n(µV)IcRN(π)\n(µV)Jc(π)\n(kAcm−2)dead layer\n(nm)Ref.Comments\nAl/Al 2O3/Ni3Al 4.6 0.45 ≥1.5⋄5 - 8 186 ⃝\nAl/AlO x/Cu/Fe .75Co.25 0.22 0.79 1.9∗>1000 <0.10 0.6 187 ⃝\nPt/Co .68B.32/Pt 0.28 0.20 0.30 56 7 35 0.0 188 1.8 K\nCu/Co .56Fe.24B.20/Cu 0.10 0.80 19 189\n[Nb/Cu] ML/Co .56Fe.24B.20/Cu 0.13 0.80 9.6 189\nCo.40Fe.40B.20 0.93 440 190‡, 1.6 K\nCo.47Gd.53 0.16 0.21 1.2 1.9 191\nNbN/GdN/NbN 0.40 80000 192‡, 0.3 K\nNbSe 2/Cr2Ge2Te6/NbSe 2 2.5 2.4 8.4 940 11 4.0×10−3193 1.6 K\nTABLE IV. Extracted parameters for Nb based Josephson junctions with strong and other fer-\nromagnetic alloy barriers. In Ref. 192, the superconductor was NbN and in Ref. 193, NbSe 2, as\nindicated in the Table. The fitting methodology to Eqn 6 is given in the main text and best fit\nvalues are quoted to 2 significant figures without consideration to their uncertainty so should be\ntreated as approximations. The full barrier layers, including normal metals and insulators, are\nlisted. Unless otherwise commented, measurements were performed at 4.2 K. Comments: ⃝The\nrange of thickness values did not cover the thickness where the first 0- πtransition was predicted. ‡\nNo 0- πoscillations were observed or claimed, and the data are fitted here with only the exponential\ndecay component of Eqn 6. ⋄This lower limit value is taken at the first observed oscillation, which\nis unlikely the first π-state due to the thickness range studied. ∗This value was extrapolated from\nmeasurements on junctions outside of this thickness.\nD. Strong and other ferromagnetic alloys\nIn Table IV, we highlight works studying Josephson junctions with strong and other\nferromagnetic alloys.\nIt is noteworthy that while most alloy barriers we have included in the Tables show\nsupercurrent transport in the diffusive or intermediate limits, the Ni 3Al barrier was found\nto be in the ballistic limit.186The long ξF1decay length and short ξF2oscillation lengths\nmeant that multiple 0- πoscillations in this material were accessible in the thickness range\n10-18 nm. The thickness range studied was unlikely to cover the first 0- πoscillation, so\nwe do not report a d0−πparameter for this material, and the reported Jc(π) corresponds to\n26the peak of the first oscillation in the measured thickness range, so a higher Jc(π) may be\npossible in thinner barriers. However, a potential downside of Ni 3Al is that the magnetic\ndead layer in this material was very large compared to others.\nThe amorphous Co-based alloys, CoFeB and CoB, are popular in spintronics and mag-\nnetic memory research due to their lack of crystalline anisotropy and weaker pinning of\nmagnetic domain walls due to the reduced density of grain boundaries.188,194For application\nin MRAM, they have been found to have large values of TMR when placed in magnetic tun-\nnel junctions.195In addition, thin layers can have perpendicular magnetic anisotropy when\nplaced on spacer layers such as Pt or Pd.188,194CoGd is an amorphous ferrimagnet, where\nthe composition can be tuned to have zero net magnetization while retaining spin-polarized\ntransport.196In Josephson junction studies, the general conclusion from Table IV is that\nthe amorphous alloys have very short ξF1decay length, most likely related to the short\nelectron mean free path in amorphous metals. This limitation may make them unsuitable\nasπ-junctions. Komori et al. showed that Jccan be enhanced in CoFeB by thermally\nannealing the junctions - which has been shown to crystallize the layer.190,197\nE. Advantages of weak vs strong ferromagnets (length scales)\nAn important consideration for constructing a π-junction is to consider the ratio of\nd0−π/ξF1. Since the supercurrent decays exponentially with ξF1, if the 0- πtransition oc-\ncurs at too great of a thickness with respect to ξF1, the resulting π-junctions will have small\ncritical Josephson currents.\nThe materials that we can describe as strong ferromagnets in Tables I, II, and IV have\nd0−π/ξF1ratios which can be ≤1. While this is great for optimizing the critical Josephson\ncurrent, one of the difficulties of fabricating π-junctions with strong ferromagnetic barriers\nis that the length scales ξF1andξF2are both short – particularly in the amorphous alloys\nin Table IV, where ξF1andξF2are<1 nm. As a result, very precise control over film\nthickness during deposition is needed to create a thickness series of samples with a small\nenough increments (step size) to observe the 0- πoscillations. For integrating junctions into\nan industrial process, the layer thickness margin of error for reproducibility of the π-junction\nwill be small, although sub-nm thick magnetic layers are routinely used in MRAM, so while\ndifficult, this level of control on an industrial scale is not impossible.\n27Weak ferromagnets offer the advantage that ξF1andξF2can be much longer, up to several\nnm in the Pd-based weak alloys (Table III), which similarly offer the d0−π/ξF1ratio≤1.\nLonger length scales in the Josephson junctions offer the advantage that the parameters of\naπ-junction, such as Jc, should in principle be more robust against small sample-to-sample\nvariations in the thickness of the barrier. In the weak ferromagnets, CuNi is somewhat of an\noutlier having a particularly large ratio of d0−π/ξF1, causing the critical Josephson current\nin the πstate to be small.\nF. Magnetic dead layers and normal metal spacer layers\nFerromagnetic materials are often modelled as slabs with uniform magnetization, however\nin reality they are rarely so simple. At the interface of a thin magnetic film, it is often ob-\nserved that the moments of the first few atomic layers may not align, reducing the observed\nmagnetization from the expected one. These are commonly referred to as magnetic dead\nlayers. Furthermore, the presence of such magnetic dead layers can degrade key properties\nsuch as the coercive field of the layer. Nb/ferromagnet interfaces are well known for having\nmagnetic dead layers160and much effort has been dedicated to studying the related inter-\nfacial roughness at Nb/ferromagnet interfaces.198,199Where possible, we have included the\nthickness of magnetic dead layers in the Tables.\nThe most reliable method to determine the presence of dead layers is to measure several\nsamples of varying ferromagnet thickness. The magnetometer returns a measurement of\nthe total magnetic moment of the sample which, so long as due care has been taken in the\nhandling of the sample and the data corrections, can be used to report the area normalized\nmagnetic moment.200If the moment/area is linear with ferromagnet thickness, then the total\ndead layer thickness ( di) can be determined according to moment/area = M(dF−di).\nAdditional normal metal layers between the Nb and ferromagnet layer are often intro-\nduced in the junctions, and we have included such layers as part of the “barrier” in the\nTables. These additional layers are commonly referred to in the literature as either buffers,\nseeds, spacers, or interlayers. Here we will use the term spacer layers to describe any addi-\ntional normal metal layers. The motivation for adding the spacers is often to provide better\nlattice matching for the ferromagnet to be deposited.\nConsidering the example of Co, the fcc spacer layers Cu, Pt and Rh have been studied.\n28The predicted mechanism is that the fcc spacer layers promote fcc growth of Co, which\notherwise may have a mixed fcc/hcp phase. The two desirable benefits that Co grown on\nthe fcc spacer layers gain compared to Co grown directly on Nb is that the magnetic dead\nlayer is reduced or removed completely, and that the improved Co crystal structure can\nimprove supercurrent carrying properties.61,159,201\nAs a counter example, the inclusion of Cu spacer layers in junctions containing Ni and\nCuNi alloys may be detrimental. Cu and Ni alloy very readily which can cause the Ni to\ninterdiffuse out of the intended layer and form interfacial alloy layers. Bolginov et al.180argue\nthat early works on CuNi, Refs. 4 and 178, contain large magnetic dead layers because of\ninterdiffusion of the Ni out of the CuNi layer into an adjacent Cu layer (which was added as\na necessary step in the fabrication process). By improving the fabrication process to exclude\nthe extra Cu layer, the resulting junctions without interdiffusion are shown in Table III to\nhave an order of magnitude improvement in Jc(π).180Kapran et al. describe a similar effect\ncomparing single layer Ni and bilayer Cu/Ni junctions, where the former junctions have an\norder of magnitude higher Jcfor the same thickness of Ni.157In addition to this Ni specific\nissue, a known downside of adding normal metal spacer layers is that every interface in the\nbarrier may reduce the critical Josephson current. In particular, normal metals with known\nlarge interface resistances should be avoided.202\nA further motivation for adding spacer layers is that they can provide a smoother surface\nfor growing the very thin ferromagnetic layers on compared to Nb. Thin film Nb has a\ncolumnar growth of grains which can result in a surface roughness comparable to the thick-\nness of the ferromagnetic barrier layer in a π-junction. Certain metals including Al, Au,\nand Cu have been shown to act as planarization layers which significantly reduce the surface\nroughness. We highlight two methods making use of this property as they are employed\nin works featured in the Tables. The first is to replace the bottom Nb electrode with a\nsuperlattice of Nb and a thin normal metal, where the superlattice will have lower surface\nroughness.203–205Several works in the Tables use this method, however they are not distin-\nguished in the Tables on the assumption that the use of a thick superlattice is unlikely to\naffect the reported junction parameters. The second approach is to place a much thinner\nsuperlattice on the surface of the single layer Nb electrode in place of the normal metal\nspacer layer, see Ref. 189 in Table IV.\n29BarrierIcRN(max)\n(µV)Jc(max)\n(kAcm−2)∆Ic(P vs AP)\n(%)Phase\ndetectionRef.Comment\nCo/Cu/NiFe 2 20 50 142\nNiFe/Cu/Co 3.5 25 206\nCu/Ni .70Fe.17Nb.13/Cu/Ni/Cu 4 50 500 143\nCu/Cu .70(NiFe) .30/Cu/NiFe/Cu 1 2.5 380 144\nNi/Cu/Ni 66 860 210 207\nCu/NiFe/Cu/Ni/Cu 11 110 93 yes 145\nNiFe/Cu/Ni yes 147 §\nCu/NiFe/Cu/Ni/Cu 9 90 800 208\nCu/NiFe/Cu/Ni/Cu 6.5 65 340 yes 146\nPt/Co/Pt/Co .68B.32/Pt 0.5 5 60 209 1.8 K\nTABLE V. Summary of literature on Nb based spin-valve Josephson junctions. The full barrier\nlayers, including normal metal spacer layers are listed. For single junction measurements, IcRN\n(max) corresponds to the largest critical Josephson current reported and ∆ Ic(P vs AP) is (the\ndifference in Icbetween P and AP)/(the lesser of Icin P or AP) expressed as a percentage,\nwhere P is the parallel and AP the antiparallel magnetic alignment. Unless otherwise commented,\nmeasurements were performed at 4.2 K. Comments: §The magnetic junctions were passive and\ntherefore the switching parameters were not directly measured.\nG. Spin-valve junctions\nIn Table V we highlight works studying controllable spin-valve junctions.\nFor spin-valve junctions we extract the parameters that may be useful for potential ap-\nplications, including the maximum critical Josephson currents, IcRNandJc, achieved in the\nwork, and what was the largest observed difference in critical current between the parallel\nand antiparallel magnetic alignments. The first quantity is an important consideration where\nthe junction is passive and does not switch (as the critical current must be larger than the\nactive junctions in the circuit) and the second is useful for memory schemes where readout\nis achieved by measuring directly the state of the spin-valve junction (superconducting or\nnormal) or any other application where large differences between the magnetic alignments\nis desirable.\n30It is possible that for the entries in Table V the two magnetic alignments correspond also\nto a change in phase difference between 0 and π. This can be inferred based on the 0 −π\ntransition of single layer junctions reported in the other Tables. Direct detection of the\nphase was achieved in spin-valve junctions by placing the junctions into a phase sensitive\nSQUID circuit.145–147\nThe highest reported critical current in Table V was for a junction where both magnetic\nlayers were pure Ni, which is not surprising from studying the presented Tables for single\nlayer junctions. While using NiFe as the free layer is magnetically beneficial (NiFe has much\nimproved switching properties compared to Ni) the resulting junctions have a much reduced\ncritical current.\nV. NON-METALLIC AND EXOTIC π-JUNCTIONS\nFerromagnetic insulators (FIs) have been predicted to show 0- πtransitions.54,77,210\nProgress in this area has been more difficult due to the limited availability of FI mate-\nrials, where GdN and Eu chalcogenides have been the most studied for potential integration\ninto superconducting devices. Of the available studies, the only FI material to show a\nJosephson current in S/FI/S junctions is GdN.55GdN Josephson junctions have shown\na range of interesting phenomena, suggesting that the physics inside these junctions is\ncomplex.55,83,117,192,211–215Complex behaviors of GdN junctions includes observation of a\npure second harmonic current-phase relation.117We also note an experiment on hybrid\njunctions containing InAs barriers coated in the FI EuS, where switching between 0 and π\ncoupling is attributed to the magnetic domain structure of the EuS.216\nCaruso et al. studied the thickness and temperature dependence of supercurrent in GdN\njunctions.192The thickness dependence at 0.3 K resulted in an exponential decay, where\nthe fit parameters are shown in Table IV. By studying the temperature dependence of the\njunctions, the authors characterize an incomplete 0- πtransition with temperature accom-\npanied by spin-triplet physics.192Follow up work reports on the possibility to quantitatively\ndescribe and eventually control spin-triplet transport through external magnetic fields which\nis correlated to an “extended” 0 −πtransition in temperature in these GdN FI barriers.215\nThe van der Waals family of materials have only recently been explored in the context of\nπ-junctions. In this case, both the superconductor and magnetic layer must be compatible\n31with the van der Waals fabrication method, so typically NbSe 2, ans-wave superconductor,\nis used. Reports on junctions and SQUIDs with a barrier of Cr 2Ge2Te6(which is described\nas either a ferromagnetic semiconductor or ferromagnetic insulator) showed a coexistence of\n0 and πphase in the junction region.217,218Another work studied the thickness dependence\nof S/F/S Josephson junctions with the Cr 2Ge2Te6barrier showing a pronounced minimum\nin supercurrent at a thickness of 8.4 nm, attributed to the 0 −πtransition.193Fitting Eqn 6\nto the thickness dependence yields the fit parameters shown on the final line of Table IV. In\na planar junction, a long ranged supercurrent was shown across a Fe 3GeTe 2van der Waals\nbarrier.219\nThe 0 −πtransition can occur in Josephson junctions without magnetic barriers if the\ngeometry of the junction allows for Zeeman splitting in a large enough magnetic field. The\nidea is that in a S/N/S junction, the normal metal in a large enough magnetic field will\nresemble a weak ferromagnet as the up and down spin bands are displaced by the Zeeman\nenergy.220,221An important factor here is that the applied magnetic field must be parallel to\nthe direction of supercurrent propagation, such that it does not contribute to the Fraunhofer\npattern discussed previously. Although an early attempt to observe a so-called Zeeman π-\njunction with the normal metal Ag by one of the authors was unsuccessful,72later works\nwith non-metallic barriers were successful. Zeeman π-junctions have been realized with\nbarriers of the Dirac semimetal Bi 1-xSnx,222the 2D material graphene,223the topological\ncrystalline insulator SnTe,224the two-dimensional electron gas system InSb,225,226and in a\nquantum dot.227In some of the successful implementations, the large g-factors and spin-orbit\ncoupling in these materials contributed along with the Zeeman effect to the observation of\n0-πtransitions. In the InSb junctions, it was possible to apply a gate voltage which could\ninduce the 0- πtransition at a set applied magnetic field.225\nElectron transport across a quantum dot is heavily influenced by the strong Coulomb\ninteraction, including the possibility of observing single electron tunneling.228A Josephson\njunction can be formed across a quantum dot when pairs can coherently tunnel. In a\nJosephson quantum dot junction, 0 −πtransitions are possible by tuning the occupancy\nlevels. Experimental realizations of π-junctions in quantum dots were achieved in an InAs\nnanowire229and in carbon nanotubes.230–232\nSeveral examples in this section explore planar junctions, as opposed to the sandwich\ngeometry of all the works presented in the Tables. Planar junctions open additional possi-\n32bilities for experimental control, such as adding voltage gates and manipulating strain by\nuse of piezoelectric substrate. Another possibility in planar junctions is to control the 0 −π\ntransition using only the geometry of the junction by introducing curvature.233,234\nVI. SPIN-TRIPLET JUNCTIONS\nThe data in the Tables show clearly that the length scales governing the decay and period\nof the 0- πoscillations, ξF1andξF2, are both very short – typically less than 1 nm in the\nstrong F materials and up to several nm in the weak F materials. In the late 1990’s, however,\nthere were some experimental hints of a long-range proximity effect in S/F systems based on\nmeasurements of the electrical resistance of long F wires connected to an S electrode.235–237\nThose results foreshadowed the 2001 theoretical breakthrough by Bergeret et al. mentioned\nin the Introduction. Further experimental support for the theory appeared in 2006,238,239\nbut the experimental breakthroughs that finally convinced the skeptics in the community\ndid not appear until 2010.50,165,240,241Those works all showed that the supercurrent in an\nS/F/S junction decayed over a distance much longer than ξFif the F layer contained the\nappropriate kind of noncollinear magnetization needed to convert spin-singlet Cooper pairs\nfrom the S electrodes into spin-triplet pairs in the central region of the junction. In the case\nof the “half-metal” CrO 2, a supercurrent can propagate hundreds of nanometers through the\nF material.50,242Since that time, there have been several reviews of spin-triplet proximity\neffects in S/F systems;21–24,243we do not review that large body of work here.\nA. Controllable π-junctions\nVery few experimental studies of spin-triplet supercurrents address the issue of the\nground-state phase difference across the junction. The theoretical papers, however, make\nspecific predictions about this issue.20,244–248Here we discuss only the sample geometry sug-\ngested by Houzet and Buzdin in 2007,244which was realized in many experiments. In that\ntheoretical work, the sequence of layers in the junction is S/F′/F/F′′/S, where the directions\nof the magnetizations of the three F layers can be controlled independently. (In experimen-\ntal devices, normal spacer layers are placed between adjacent F layers to prevent exchange\ncoupling between them.) In particular, the magnetizations Mof any two adjacent layers\n33must be non-collinear; let us choose adjacent layers to have orthogonal magnetizations to\nmaximize generation of spin-triplet supercurrents. If we define the direction of M′as the\nz-direction and the direction of Mas the x-direction, and if we keep all three magnetizations\ncoplanar, then the direction of M′′may be either plus or minus z. According to theory, the\nformer case produces a π-junction whereas the latter produces a 0-junction.20,244–246The\nonly experimental work that tests this prediction, to our knowledge, is the one by Glick\net al. in 2018.249To achieve the required orthogonal magnetizations, those authors used a\nPd/Co multilayer with perpendicular anisotropy as the central F layer. (Actually, two such\nmultilayers were placed back-to-back separated by a Ru spacer layer to achieve a synthetic\nantiferromagnet. That does not change the 0 −πphysics, according to theory.245,246) Those\nauthors used Ni as the fixed F′′layer and NiFe as the free F′layer. Seven out of eight\njunctions showed robust πphase switching when the direction of the NiFe magnetization\nwas reversed by a small applied field.\nFrom a practical perspective, spin-triplet junctions appear less attractive than the simpler\nspin-valve junctions discussed in Section III C. They do have one advantage, however: the\nconstraints on the thicknesses of the F layers needed for switching between the 0 and π\nstates of the junctions are less stringest in the case of spin-triplet junctions than they are\nfor spin-valve junctions.250\nB. ϕand ϕ0-junctions\nThe spin-singlet physics discussed in the majority of this article would appear to allow\nonly junctions with ground-state phase differences of 0 or π. That is not true, however.\nA spatially-extended junction containing only a single F layer with varying thickness, such\nthat some parts of the junction prefer the 0-state while others prefer the π-state, can have\ndegenerate ground states at phases ±ϕ.251–255Such junctions are called “ ϕ-junctions,” where\nthe phase ϕcan take any value between 0 and π. Such junctions still obey the conventional\ntime-reversal symmetry relation, Is(−φ) =−Is(φ), where φis the phase difference between\nthe superconducting condensates in the two S electrodes.\nSpin-triplet junctions containing three F-layers have an additional possibility, if the three\nmagnetizations are non-coplanar. In that case, time-reversal symmetry is broken, and cannot\nbe restored by a simple rotation of the entire spin system.256Such junctions may have a\n34ground-state phase difference of ϕ0, which can take any value between 0 and 2 π, and is\nnondegenerate on that interval. Such junctions are now called ϕ0-junctions, to distinguish\nthem from the ϕ-junctions discussed in the previous paragraph. The existence of ϕ0-junctions\nwas predicted in a number of theoretical works starting in 2007,256–266and two reviews\nhave appeared recently.267,268Because of the broken symmetries, ϕ0-junctions do not obey\nIs(−φ) =−Is(φ), hence this phenomenon is also referred to as the “anomalous Josephson\neffect”, or AJE. Since 2016, the AJE has been experimentally observed in a wide variety of\nexotic systems with strong spin-orbit interaction218,269–275and in planar out-of-equilibrium\ndevices,276but so far it has not been observed in metallic sandwich junctions.\nVII. OPEN CHALLENGES AND OUTLOOK\nWe hope that the presented Tables act as both a record of current literature and inspi-\nration for the future direction of the field. While magnetic π-junctions are somewhat of\na mature field, we feel from the presented Tables that a Goldilocks junction material that\npasses a large supercurrent in the π-state while having well controlled magnetic properties\nis still to be discovered. Such a material is particularly needed to realize practical spin-valve\njunctions. In this section, we outline some open challenges and provide an outlook on the\nfield.\nA. Spread in junction parameters and influence of fabrication method\nFrom examination of the Tables in this work, the reader may find themselves overwhelmed\nwith seemingly contradictory information on the same material - a thickness of ferromagnet\nthat provides a π-state from one work may not directly translate to the π-state in another\nwork. For Ni and NiFe, the two most studied barriers, the authors have directed the reader\ntowards what they consider to be the most reliable value for the thickness d0−π.\nIn some cases, the explanation for spread in the Tables may be as simple as differences\nin the way the ultra-thin layer thicknesses were calibrated, or that the range of thicknesses\nstudied unintentionally missed a 0 −πtransition. In other cases, the explanation may be\nmore complex. The decay length inside the ferromagnet, ξF1, is shown in the Tables to vary\nconsiderably between works. This parameter is sensitive to the exact layer morphology,\n35which itself is influenced by the presence or absence of normal metal spacer layers. We also\nmust consider what role, if any, does the junction fabrication process have on the measured\nparameters in the Tables. Bolginov et al. provide a detailed account of improvements made\nto their fabrication process, which resulted in a significant difference to the parameters,\nparticularly the thickness of d0−π, in junctions measured with CuNi barriers.180\nTwo common fabrication methods are based around either the ion beam etching or focused\nion beam techniques, which are described in detail elsewhere.204,277Here, we highlight a\nsignificant difference in the two methods which may affect junction parameters. The ion\nbeam etching process typically requires that the top Nb electrode be deposited separately\nto the bottom Nb electrode and barrier layers, post-fabrication and etching. To protect the\nbarrier layers from oxidation during fabrication, a normal metal capping layer must be added\nabove the barrier layers. In focused ion beam fabrication, both bottom and top electrodes as\nwell as the barrier layers can be deposited in the same vacuum cycle, meaning no additional\ncapping layers are required by the processing. We speculate that this difference may account\nfor the generally higher critical Josephson currents observed in focused ion beam junctions,\nalthough there has been no systematic study comparing the two techniques to date.\nBased on the discussion in this section, the authors highly recommend that while the\nTables in this work can be used to help choose materials to implement as π-junctions in\nan application, that researchers don’t assume that parameters provided in the Tables will\nexactly reproduce in their own process. It is therefore necessary to conduct a full systematic\nstudy of a material using the deposition and fabrication methods needed for the ultimate\napplication to determine the exact parameters for junctions produced by that process.\nB. Magnetic initialization\nIn the as-grown state, the direction of the F layer’s magnetization in a junction may be\nuncontrolled and could have broken up into magnetic domains. In a single F layer junction,\nthe physics driving the 0 −πoscillations should not depend on the direction of magnetiza-\ntion. However, for in-plane magnetization, the Fraunhofer pattern of an individual junction\ncan be strongly influenced by the direction and domain state of the F layer, as discussed in\nSection II D. Since determining the critical Josephson current in the junction relies on inter-\npretation of the Fraunhofer response, it may be preferable to initialize the magnetization to\n36a known direction and to minimize domains. Another case where setting the magnetization\ndirection of the ferromagnet in the junction is crucial are spin-valve junctions, where it is\nessential to know and control the relative orientations of the magnetic layers. For junctions\nwith perpendicular magnetic anisotropy, the magnetization is perpendicular to the applied\nfield so the Fraunhofer pattern should be less influenced and therefore magnetic initialization\nmay not be necessary.\nTo aid control of the in-plane magnetization direction in a junction, it is usual to create\na shape anisotropy by defining the junction area as an ellipse, where it is favourable for the\nmagnetization to point in the long axis. Even with a shape anisotropy, for a single domain F\nlayer, there are now two directions that the magnetization could point. When the magnetic\ndomains are small the situation is worse, as the magnetic layer may break up into domains\npointing in many different directions, and stray magnetic fields may emerge from the domain\nwalls.\nThe standard procedure, to set the direction of magnetization and minimize the number\nof domains, is to initialize the magnetic layer in a large magnetic field. The magnetic field\nmust be applied while the circuit is far below the Curie temperature – which might be below\nroom temperature in the case of very thin magnetic layers – but above the superconducting\ncritical temperature of Nb to avoid flux trapping. For Ni, where the small domain size is\nparticularly problematic, it is typical to apply a global initialization field of ≥350 mT to\nsaturate the Ni magnetization.155A superconducting digital logic circuit is unlikely to be\nimplemented in a cryogenic environment which includes the ability to apply a large global\nfield, hence it would be favourable to have a magnetic layer which forms a single magnetic\ndomain and does not require an initialization field procedure.\nIn some materials, such as NiFe, magnetic anisotropy can be set by a growth field during\ndeposition, which may be enough to set the direction of magnetization without the need for\nadditional initialization fields. Another mechanism which can set the magnetization is the\nexchange bias effect between a ferromagnetic and antiferromagnetic material.278However, it\nhas been found that supercurrent is severely suppressed in Josephson junctions containing\nmetallic antiferromagnetic barriers.279–281\n37C. Materials compatibility with Nb or CMOS fab (MRAM example)\nMRAM is an example where integrating the magnetic junctions into the traditional\nCMOS wafer fabrication has been extremely difficult. Generally, the magnetic layers of\nthe MRAM are incompatible with the CMOS process. One of the reasons for this is that\nelevated temperatures used for the CMOS process would cause the thin magnetic layers to\ninterdiffuse. The solution is that the standard CMOS wafer is manufactured in one front-\nend-of-the-line facility, then shipped to a separate facility for backend-of-the-line processing,\nwhere the MRAM is built. This approach has several technological disadvantages, including\nseverely limiting circuit design possibilities.\nFor superconducting electronics, although a multi-facility approach is possible without\ncompromising the overall transport properties of individual junctions,92,109,110,282,283in our\nopinion it would be best to avoid the multi-facility approach so that circuit design is not\nlimited. In 2019, MIT Lincoln Laboratory announced a new fabrication process for su-\nperconductor electronics called PSE2, which integrates Ni π-junctions - an important step\ntowards using π-junctions in logic circuits.156But we are not aware of any published circuit\ndemonstrations using that process. As we have previously motivated, pure Ni may not be\nthe ideal choice of material due to the high initialization field required and small magnetic\ndomains.\nThe Tables presented here might motivate the reader to revisit pure Fe or to pursue\nother Fe containing alloys (NiFe, PdFe, etc.) as potential π-junction materials. However,\nwe would urge caution on exploring further materials that include Fe when considering\nincorporation into an integrated fabrication process. Fe may be incompatible with scaled\nfabrication processes due to the potential detrimental effect Fe contamination would have\non the rest of the process.\nOur conclusion from a materials compatibility perspective is that Ni based alloys are well\nmotivated for further work, particularly those that are in the limit Kapran et al. describe as\n“strong-but-clean”, meaning strong ferromagnetism with ballistic transport properties.157\nWe also mention that it may be possible to explore alloys and compounds for magnetic\nbarriers where the component materials are otherwise non-magnetic, such as PdMn,27some\nHeuslers,241,284andL10-ordered MnAl285or MnGa.286Such an approach may negate any\nconcerns about ferromagnetic particle contamination in a Nb process.\n38D. Materials and read/write operations for spin-valves\nWe identify some outstanding challenges specific to spin-valve junctions. Firstly, to\nachieve a spin-valve junction with high critical current, it seems necessary to replace the\nNiFe free layer. Secondly, we must remove the requirement for global magnetic field for\nread/write operations and move towards on-chip operations.\nThe materials for spin-valve junctions have several requirements and important consid-\nerations: the fixed layer must be robust enough against the switching of the free layer, both\nlayers should be single domain so that the supercurrent and phase difference across the\njunction is uniform, both layers must be sufficiently thin to allow a significant supercurrent\nto propagate through the junction, magnetic dead layers (which can dominate in very thin\nlayers) must be minimized, and the 0 −πoscillations of each component layer must be well\ncharacterized. It is not obvious from the Tables presented in this work that there is an\nimmediate replacement for either NiFe as the free layer or Ni as the fixed layer.\nOn the prospect of on-chip writing, we note that there are adventurous proposals for\nswitching magnetic layers in the superconducting state of S-F systems,287–290although there\nis not yet experimental realization. We therefore suggest that established mechanisms used\nin the field of spintronics and technologies for MRAM can also be applied in the normal\nstate of a superconducting spin-valve. Spin transfer torque switching, where a current pulse\nis passed through the junction and induced magnetization dynamics causes the free layer\nto switch, was demonstrated by Baek et al. by applying a current density to their junction\nwhich greatly exceeded Jc.207Spin orbit torque switching, where a current pulse is passed\nthrough a layer with large spin Hall angle adjacent to the free layer causing it to switch, has\nshown promising results in room-temperature magnetic tunnel junctions, and promises faster\nswitching speeds than spin transfer torque.291Raytheon BBN Technologies have proposed a\nlow temperature superconductor-ferromagnet hybrid memory (not based on the Josephson\neffect) where magnetic switching is achieved by spin Hall effect.292\nE. Altermagnets\nSolids with an intrinsic magnetic phase are traditionally classed as having either fer-\nromagnetic or antiferromagnetic ordering. Recently, there has been great interest in the\n39fundamental study and potential applications of an emerging third class of magnetic solid,\nwhere the magnetic phase is known as altermagnetism.293Altermagnets have properties that\nblur the lines between the traditional ferromagnetic and antiferromagnetic orderings. On the\none hand, these materials have electronic properties consistent with ferromagnetism, they\nshow an anomalous Hall effect and have spin-polarized conduction bands. On the other\nhand, altermagnetic materials have antiparallel magnetic crystal order and zero net magne-\ntization, consistent with antiferromagnetism. Altermagnets are expected to be abundant in\nnature.293\nThe presence of a spin-polarized conduction band suggests that much of the rich physics\npresent in ferromagnet-superconductor hybrid systems are also present when the ferromagnet\nis replaced by an altermagnet, along with additional unique physics in the new altermagnet\nsystem. Notably in the altermagnet, the finite pair momentum appears in the absence of a\nnet magnetization.294Zero net magnetic moment junctions offers the significant advantages\nover ferromagnet systems that the junctions produce either zero or negligibly small stray\nmagnetic fields, and they don’t require any magnetic initialization.\nOf particular interest to the topic of this review are predictions of 0- πoscillations in\nJosephson junctions containing altermagnetic barriers.294–297To date, there are no experi-\nmental studies of altermagnet-superconductor hybrid systems. The most studied candidate\naltermagnet is RuO 2, a metal with a bulk room-temperature resistivity of ρ≈35µΩ-cm.298\nF. Neuromorphic and novel analogue computing\nMagnetic Josephson junctions have been proposed as artificial synapses for neuromorphic\ncomputing schemes.299–301Neuromorphic computing is seen as a potential solution to the\nproblem of poor energy efficiency when running neural networks on traditional computing\narchitectures. The Tutorial of Schneider et al. covers the topic of neuromorphic systems\nbased on magnetic Josephson junctions in detail.302Here we focus on the materials and\nrequirements for this application.\nThe requirements in this potential application is for the junction to provide a continuously\ntunable (analogue) response, provide a degree of plasticity/memory, and be scalable and\nenergy efficient enough for the architecture required to interconnect many such devices\nneeded for neuromorphic application.\n40So far, synapse behavior in magnetic Josephson junctions has been achieved in magnetic\nnanoclusters. The magnetic clusters do not spontaneously order, but can be aligned by\nsubsequent current pulses. The magnetic order can be tuned continuously from the dis-\nordered to ordered magnetic states, and the degree of order affects the critical current of\nthe junction.302–305Similar to the π-junctions presented in the Tables, there will be a wide\nvariation of performance in metrics such as the IcRNof barriers with different magnetic\nnanoclusters. We are aware of two magnetic nanocluster barriers which have been stud-\nied for this application, Mn clusters embedded in Si303and Fe clusters embedded in Ge.304\nCurrent limitations are that the IcRNproducts in these barriers are quite a low and the\nMn clusters in Si required an annealing step. A potential alternative to nanoclusters are\nultra-thin magnetic layers, sometimes called dusting layers, which can also be close to the\nsuperparamagnetic transition.\nVIII. CONCLUSIONS\nReaders new to this field may be overwhelmed by the sheer number of materials listed in\nthe Tables in Section IV, as well as the wide variability in parameters extracted by different\nresearch groups. To avoid that parting impression, we finish the review by expressing a few\nof our own opinions.\nFirst, if we had to choose a material from the Tables to make highly-reproducible π-\njunctions that can tolerate small variations in thickness, we would choose a PdNi alloy with\na Ni concentration in the range of 10 - 15%. Pham et al. report values of JcandIcRN\nin the π-state as large as 70 kA/cm2and 18 µV, respectively, in Pd 89Ni11junctions with\nNbN superconducting electrodes.183Because of its mild perpendicular anisotropy, junctions\ncontaining PdNi may not need magnetic initialization. The only drawback of this material\nis the cost of Pd; but we remind readers that only very small amounts of the material are\nneeded.\nSecond, if one wants the highest possible critical current density in a π-junction, then Ni\nis the obvious choice. But Ni junctions require magnetic initialization in a rather large mag-\nnetic field to establish a reproducible magnetized initial state. Without that initialization,\nthe Fraunhofer pattern will vary randomly from sample to sample, yielding non-reproducible\nvalues of Jc.\n41Third, if one wants a magnetically-soft free layer for a spin-valve junction, then NiFe is\nthe best choice so far. Our own attempts to dilute NiFe to lower its magnetization (and\nhence the magnetic switching energy) have been largely unsuccessful; in the case of doping\nwith Mo or Cr, the resulting material has a short mean free path, leading to a steep decay\nofJcwith thickness.\nAnd finally, we lament the heavy use of CuNi alloy by groups new to the field. We\nacknowledge the many “firsts” achieved by the Ryazanov group and their collaborators using\njunctions containing CuNi, but the high rate of spin-flip scattering in the material causes\nJcto decay very steeply with thickness, leading to very small values of JcandIcRNin the\nπ-state. The main advantages of CuNi are: i) ξFis long due to the small magnetization, so\nthickness fluctuations are tolerated; ii) the perpendicular anisotropy avoids distortions of the\nFraunhofer pattern; and iii) the material is less expensive than PdNi. We would recommend\nusing that material only when junction size is not an issue, so that the small value of Jcis\ntolerated. But we much prefer PdNi alloy.\nIn summary, we believe that π-junctions have an important role to play in the devel-\nopment of superconducting digital electronics. We hope that this review will help workers\nnavigate the vast literature as they explore new materials for π-junctions.\nACKNOWLEDGMENTS\nN.O.B. wishes to thank all the students who have worked with him on superconduct-\ning/ferromagnetic hybrid systems over the past two decades. Special thanks also go to B.\nBi, D. Edmunds, R. Loloee, and W.P. Pratt, Jr., without whom none of this work would\nhave been possible, and to the JMRAM team at Northrop Grumman Corporation, especially\nA.Y. Herr, D.L. Miller, O. Naaman, N. Rizzo, T.F. Ambrose, and M.G. Loving. N.S. wishes\nto thank G. Burnell and everyone who has worked with him on this topic, and acknowledges\nsupport from new faculty startup funding made available by Texas State University. We ac-\nknowledge helpful suggestions from F.S. Bergeret, A.I. Buzdin, A.A. Golubov, S. Jacobsen,\nM. Kupriyanov, O. Mukhanov, Z. Radovic, V.V. Ryazanov, I.I. Soloviev, and A.F. Volkov.\n42AUTHOR DECLARATIONS\nConflict of Interest\nThe authors have no conflicts to disclose.\nAuthor Contributions\nN.O. Birge : writing - original draft (equal) N. Satchell : writing - original draft (equal)\nDATA AVAILABILITY\nData sharing is not applicable to this article as no new data were created or analyzed in\nthis study.\nREFERENCES\n1A. Veretennikov, “Supercurrents through the superconduc-\ntor–ferromagnet–superconductor (SFS) junctions,” Physica B Condens. Matter.\n284-288 , 495–496 (2000).\n2V. V. Ryazanov, V. A. Oboznov, A. Y. Rusanov, A. V. Veretennikov, A. A. Golubov,\nand J. Aarts, “Coupling of Two Superconductors through a Ferromagnet: Evidence for\naπJunction,” Phys. Rev. Lett. 86, 2427–2430 (2001).\n3T. Kontos, M. Aprili, J. Lesueur, F. Genˆ et, B. Stephanidis, and R. Boursier, “Josephson\nJunction through a Thin Ferromagnetic Layer: Negative Coupling,” Phys. Rev. Lett. 89,\n137007 (2002).\n4V. V. Ryazanov, V. A. Oboznov, A. V. Veretennikov, and A. Y. Rusanov, “Intrinsi-\ncally frustrated superconducting array of superconductor-ferromagnet-superconductor π\njunctions,” Phys. Rev. B 65, 020501 (2001).\n5W. Guichard, M. Aprili, O. Bourgeois, T. Kontos, J. Lesueur, and P. Gandit, “Phase\nSensitive Experiments in Ferromagnetic-Based Josephson Junctions,” Phys. Rev. Lett.\n90, 167001 (2003).\n436A. Bauer, J. Bentner, M. Aprili, M. L. Della Rocca, M. Reinwald, W. Wegscheider, and\nC. Strunk, “Spontaneous supercurrent induced by ferromagnetic πjunctions,” Phys. Rev.\nLett. 92, 217001 (2004).\n7S. M. Frolov, D. J. Van Harlingen, V. A. Oboznov, V. V. Bolginov, and\nV. V. Ryazanov, “Measurement of the current-phase relation of superconduc-\ntor/ferromagnet/superconductor πJosephson junctions,” Phys. Rev. B 70, 144505 (2004).\n8L. N. Bulaevskii, V. V. Kuzii, and A. A. Sobyanin, “Superconducting system with weak\ncoupling with a current in the ground state,” Pis’ma Zh. Eksp. Teor. Fiz. 25, 314 – 318\n(1977).\n9A. I. Buzdin, L. N. Bulaevskii, and S. V. Panyukov, “Critical-current oscillations as a\nfunction of the exchange field and thickness of the ferromagnetic metal (F) in an SFS\nJosephson junction,” JETP Lett 35, 178–180 (1982).\n10A. Buzdin and M. Kupriyanov, “Josephson junction with a ferromagnetic layer,” JETP\nLetters 53, 321–326 (1991).\n11A. A. Golubov, M. Y. Kupriyanov, and E. Il’ichev, “The current-phase relation in Joseph-\nson junctions,” Rev. Mod. Phys. 76, 411–469 (2004).\n12I. F. Lyuksyutov and V. L. Pokrovsky, “Ferromagnet–superconductor hybrids,” Adv.\nPhys. 54, 67–136 (2005).\n13A. I. Buzdin, “Proximity effects in superconductor-ferromagnet heterostructures,” Rev.\nMod. Phys. 77, 935–976 (2005).\n14E. Terzioglu and M. Beasley, “Complementary Josephson junction devices and circuits:\na possible new approach to superconducting electronics,” IEEE Trans. Appl. Supercond.\n8, 48–53 (1998).\n15A. V. Ustinov and V. K. Kaplunenko, “Rapid single-flux quantum logic using π-shifters,”\nJ. Appl. Phys. 94, 5405 (2003).\n16L. B. Ioffe, V. B. Geshkenbein, M. V. Feigel’man, A. L. Fauch` ere, and G. Blatter, “En-\nvironmentally decoupled sds-wave Josephson junctions for quantum computing,” Nature\n398, 679–681 (1999).\n17G. Blatter, V. B. Geshkenbein, and L. B. Ioffe, “Design aspects of superconducting-phase\nquantum bits,” Phys. Rev. B 63, 174511 (2001).\n18F. S. Bergeret, A. F. Volkov, and K. B. Efetov, “Long-range proximity effects in\nsuperconductor-ferromagnet structures.” Phys Rev Lett 86, 4096–4099 (2001).\n4419A. Kadigrobov, R. I. Shekhter, and M. Jonson, “Quantum spin fluctuations as a source\nof long-range proximity effects in diffusive ferromagnet-superconductor structures,” Eu-\nrophys. Lett. 54, 394–400 (2001).\n20A. F. Volkov, F. S. Bergeret, and K. B. Efetov, “Odd triplet superconductivity in\nsuperconductor-ferromagnet multilayered structures,” Phys. Rev. Lett. 90, 117006 (2003).\n21F. S. Bergeret, A. F. Volkov, and K. B. Efetov, “Odd triplet superconductivity and related\nphenomena in superconductor-ferromagnet structures,” Rev. Mod. Phys. 77, 1321–1373\n(2005).\n22M. Eschrig, “Spin-polarized supercurrents for spintronics,” Phys. Today 64, 43–49 (2011).\n23M. Eschrig, “Spin-polarized supercurrents for spintronics: a review of current progress,”\nRep. Prog. Phys 78, 104501 (2015).\n24J. Linder and J. W. A. Robinson, “Superconducting spintronics,” Nat. Phys. 11, 307–315\n(2015).\n25A. S. Mel’nikov, S. V. Mironov, A. V. Samokhvalov, and A. I. Buzdin, “Superconducting\nspintronics: state of the art and prospects,” Phys.-Usp. 65, 1248–1289 (2022).\n26R. Cai, I. ˇZuti´ c, and W. Han, “Superconductor/ferromagnet heterostructures: A platform\nfor superconducting spintronics and quantum computation,” Adv. Quantum Technol. 6,\n2200080 (2023).\n27M. Y. Kupriyanov, A. A. Golubov, and M. Siegel, “Josephson junctions with ferromag-\nnetic materials,” in Micro- and Nanoelectronics 2005 , Vol. 6260, edited by K. A. Valiev\nand A. A. Orlikovsky, International Society for Optics and Photonics (SPIE, 2006) p.\n62600S.\n28H. K. Wong and J. B. Ketterson, “Superconducting properties of Fe/V/Fe sandwiches,”\nJ Low Temp Phys 63, 139–150 (1986).\n29H. K. Wong, B. Y. Jin, H. Q. Yang, J. B. Ketterson, and J. E. Hilliard, “Superconducting\nproperties of V/Fe superlattices,” J Low Temp Phys 63, 307–315 (1986).\n30C. Uher, J. Cohn, and I. Schuller, “Upper critical field in anisotropic superconductors.”\nPhys Rev, B Condens Matter 34, 4906–4908 (1986).\n31Z. Radovic, L. Dobrosavljevic-Grujic, A. Buzdin, and J. Clem, “Upper critical fields of\nsuperconductor-ferromagnet multilayers.” Phys Rev, B Condens Matter 38, 2388–2393\n(1988).\n4532Z. Radovic, M. Ledvij, L. Dobrosavljevic-Grujic, A. Buzdin, and J. Clem, “Transition\ntemperatures of superconductor-ferromagnet superlattices.” Phys Rev, B Condens Matter\n44, 759–764 (1991).\n33J. Jiang, D. Davidovic, D. Reich, and C. Chien, “Oscillatory Superconducting Transition\nTemperature in Nb/Gd Multilayers,” Phys. Rev. Lett. 74, 314–317 (1995).\n34T. M¨ uhge, N. Garif’yanov, Y. Goryunov, G. Khaliullin, L. Tagirov, K. Westerholt, I. Gar-\nifullin, and H. Zabel, “Possible origin for oscillatory superconducting transition tempera-\nture in superconductor/ferromagnet multilayers,” Phys. Rev. Lett. 77, 1857–1860 (1996).\n35T. M¨ uhge, K. Westerholt, H. Zabel, N. N. Garif’yanov, Y. V. Goryunov, I. A. Garifullin,\nand G. G. Khaliullin, “Magnetism and superconductivity of Fe/Nb/Fe trilayers,” Phys.\nRev. B 55, 8945–8954 (1997).\n36C. Chien and D. H. Reich, “Proximity effects in superconducting/magnetic multilayers,”\nJ. Magn. Magn. Mater. 200, 83–94 (1999).\n37E. A. Demler, G. B. Arnold, and M. R. Beasley, “Superconducting proximity effects in\nmagnetic metals,” Phys. Rev. B 55, 15174–15182 (1997).\n38P. Fulde and R. A. Ferrell, “Superconductivity in a strong spin-exchange field,” Phys.\nRev.135, A550–A563 (1964).\n39A. Larkin and Y. Ovchinnikov, “Nonuniform state of superconductors,” Sov. Phys. JETP\n20(1965).\n40T. Kontos, M. Aprili, J. Lesueur, and X. Grison, “Inhomogeneous superconductivity\ninduced in a ferromagnet by proximity effect,” Phys. Rev. Lett. 86, 304–307 (2001).\n41V. Ambegaokar and A. Baratoff, “Tunneling Between Superconductors,” Phys. Rev. Lett.\n10, 486–489 (1963).\n42H. Ness, I. A. Sadovskyy, A. E. Antipov, M. van Schilfgaarde, and R. M. Lutchyn,\n“Supercurrent decay in ballistic magnetic Josephson junctions,” npj Comput. Mater. 8,\n23 (2022).\n43M. Faur´ e, A. I. Buzdin, A. A. Golubov, and M. Y. Kupriyanov, “Properties of supercon-\nductor/ferromagnet structures with spin-dependent scattering,” Phys. Rev. B 73, 064505\n(2006).\n44N. G. Pugach, M. Y. Kupriyanov, E. Goldobin, R. Kleiner, and D. Koelle,\n“Superconductor-insulator-ferromagnet-superconductor Josephson junction: From the\ndirty to the clean limit,” Phys. Rev. B 84, 144513 (2011).\n4645A. Buzdin, “ π-junction realization due to tunneling through a thin ferromagnetic layer,”\nJetp Lett. 78, 583–586 (2003).\n46D. M. Heim, N. G. Pugach, M. Y. Kupriyanov, E. Goldobin, D. Koelle, R. Kleiner,\nN. Ruppelt, M. Weides, and H. Kohlstedt, “The effect of normal and insulating layers\non 0-πtransitions in Josephson junctions with a ferromagnetic barrier,” New J. Phys. 17,\n113022 (2015).\n47F. S. Bergeret, A. F. Volkov, and K. B. Efetov, “Josephson current in superconductor-\nferromagnet structures with a nonhomogeneous magnetization,” Phys. Rev. B 64, 134506\n(2001).\n48Y. Chong, P. D. Dresselhaus, and S. P. Benz, “Electrical properties of Nb–MoSi2–Nb\nJosephson junctions,” Appl. Phys. Lett 86, 232505 (2005).\n49J. Clarke, “Supercurrents in lead-copper-lead sandwiches,” Proc. R. Soc. Lond. A 308,\n447–471 (1969).\n50M. S. Anwar, F. Czeschka, M. Hesselberth, M. Porcu, and J. Aarts, “Long-range super-\ncurrents through half-metallic ferromagnetic CrO 2,” Phys. Rev. B 82, 100501 (2010).\n51J. Wang, M. Singh, M. Tian, N. Kumar, B. Liu, C. Shi, J. K. Jain, N. Samarth, T. E.\nMallouk, and M. H. W. Chan, “Interplay between superconductivity and ferromagnetism\nin crystalline nanowires,” Nat. Phys. 6, 389–394 (2010).\n52M. Weides, M. Kemmler, E. Goldobin, D. Koelle, R. Kleiner, H. Kohlstedt, and A. I.\nBuzdin, “High quality ferromagnetic 0 and πJosephson tunnel junctions,” Appl. Phys.\nLett. 89, 122511 (2006).\n53Y. Tanaka and S. Kashiwaya, “Theory of Josephson effect in superconductor-\nferromagnetic-insulator-superconductor junction,” Physica C: Supercond. 274, 357–363\n(1997).\n54S. Kawabata, Y. Asano, Y. Tanaka, A. A. Golubov, and S. Kashiwaya, “Josephson π\nstate in a ferromagnetic insulator,” Phys. Rev. Lett. 104, 117002 (2010).\n55K. Senapati, M. G. Blamire, and Z. H. Barber, “Spin-filter Josephson junctions,” Nat.\nMater. 10, 849–852 (2011).\n56A. Barone and G. Patern` o, Physics and Applications of the Josephson Effect (John Wiley\n& Sons, New York, 1982).\n57M. G. Blamire, C. B. Smiet, N. Banerjee, and J. W. A. Robinson, “Field modulation\nof the critical current in magnetic Josephson junctions,” Supercond. Sci. Technol. 26,\n47055017 (2013).\n58I. A. Golovchanskiy, V. V. Bol’ginov, V. S. Stolyarov, N. N. Abramov, A. Ben Hamida,\nO. V. Emelyanova, B. S. Stolyarov, M. Y. Kupriyanov, A. A. Golubov, and V. V.\nRyazanov, “Micromagnetic modeling of critical current oscillations in magnetic Josephson\njunctions,” Phys. Rev. B 94, 214514 (2016).\n59O. Bourgeois, P. Gandit, J. Lesueur, A. Sulpice, X. Grison, and J. Chaussy, “Josephson\neffect through a ferromagnetic layer,” Eur. Phys. J. B 21, 75–80 (2001).\n60T. S. Khaire, W. P. Pratt, and N. O. Birge, “Critical current behavior in Josephson\njunctions with the weak ferromagnet PdNi,” Phys. Rev. B 79, 094523 (2009).\n61M. A. Khasawneh, W. P. Pratt, and N. O. Birge, “Josephson junctions with a synthetic\nantiferromagnetic interlayer,” Phys. Rev. B 80, 020506 (2009).\n62V. V. Ryazanov, “Josephson superconductor - ferromagnet - superconductor - π-contact\nas an element of a quantum bit (experiment),” Phys.-Uspekhi 42, 825 – 827 (1999).\n63I. S. Veshchunov, V. A. Oboznov, A. N. Rossolenko, A. S. Prokokiev, L. Y. Vinnikov,\nA. Y. Rusanov, and D. V. Matveev, “Observation of the magnetic domain structure in\nCu0.47Ni0.53thin films at low temperature,” Pis’ma v. ZhETF 88, 873–876 (2008).\n64J. A. Glick, R. Loloee, W. P. Pratt, and N. O. Birge, “Critical Current Oscillations of\nJosephson Junctions Containing PdFe Nanomagnets,” IEEE Trans. Appl. Supercond. 27,\n1800205 (2017).\n65D. A. Wollman, D. J. Van Harlingen, W. C. Lee, D. M. Ginsberg, and A. J. Leggett,\n“Experimental determination of the superconducting pairing state in YBCO from the\nphase coherence of YBCO-Pb dc SQUIDs,” Phys. Rev. Lett. 71, 2134–2137 (1993).\n66J. R. Kirtley, C. C. Tsuei, J. Z. Sun, C. C. Chi, L. S. Yu-Jahnes, A. Gupta, M. Rupp,\nand M. B. Ketchen, “Symmetry of the order parameter in the high-T csuperconductor\nYBa 2Cu3O7−δ,” Nature 373, 225–228 (1995).\n67S.-K. Yip, “Energy-resolved supercurrent between two superconductors,” Phys. Rev. B\n58, 5803–5807 (1998).\n68F. K. Wilhelm, G. Sch¨ on, and A. D. Zaikin, “Mesoscopic Superconducting–Normal\nMetal–Superconducting Transistor,” Phys. Rev. Lett. 81, 1682–1685 (1998).\n69J. J. A. Baselmans, B. J. van Wees, and T. M. Klapwijk, “Nonequilibrium supercurrent\ntransport in controllable superconductor–normal-metal–superconductor junctions,” Phys.\nRev. B 63, 094504 (2001).\n4870B. J. van Wees, K. Lenssen, and C. Harmans, “Transmission formalism for supercurrent\nflow in multiprobe superconductor-semiconductor-superconductor devices,” Phys. Rev. B\n44, 470–473 (1991).\n71J. Huang, F. Pierre, T. T. Heikkil¨ a, F. K. Wilhelm, and N. O. Birge, “Observation of\na controllable πjunction in a 3-terminal Josephson device,” Phys. Rev. B 66, 020507\n(2002).\n72M. S. Crosser, J. Huang, F. Pierre, P. Virtanen, T. T. Heikkil¨ a, F. K. Wilhelm, and\nN. O. Birge, “Nonequilibrium transport in mesoscopic multi-terminal SNS Josephson\njunctions,” Phys. Rev. B 77, 014528 (2008).\n73J. E. Mooij, T. P. Orlando, L. Levitov, L. Tian, van der Wal C H, and S. Lloyd, “Joseph-\nson persistent-current qubit.” Science 285, 1036–1039 (1999).\n74T. P. Orlando, J. E. Mooij, L. Tian, C. H. van der Wal, L. S. Levitov, S. Lloyd, and\nJ. J. Mazo, “Superconducting persistent-current qubit,” Phys. Rev. B 60, 15398–15413\n(1999).\n75T. Yamashita, K. Tanikawa, S. Takahashi, and S. Maekawa, “Superconducting πqubit\nwith a ferromagnetic Josephson junction,” Phys. Rev. Lett. 95, 097001 (2005).\n76T. Yamashita, S. Takahashi, and S. Maekawa, “Superconducting πqubit with three\nJosephson junctions,” Appl. Phys. Lett 88, 132501 (2006).\n77S. Kawabata, S. Kashiwaya, Y. Asano, Y. Tanaka, and A. A. Golubov, “Macroscopic\nquantum dynamics of πjunctions with ferromagnetic insulators,” Phys. Rev. B 74, 180502\n(2006).\n78T. Kato, A. A. Golubov, and Y. Nakamura, “Decoherence in a superconducting flux\nqubit with a π-junction,” Phys. Rev. B 76, 172502 (2007).\n79S. Kawabata, A. A. Golubov, Ariando, C. J. M. Verwijs, H. Hilgenkamp, and J. R.\nKirtley, “Macroscopic quantum tunneling and quasiparticle-tunneling blockade effect in\ns-wave/ d-wave hybrid junctions,” Phys. Rev. B 76, 064505 (2007).\n80A. K. Feofanov, V. A. Oboznov, V. V. Bol’ginov, J. Lisenfeld, S. Poletto, V. V. Ryazanov,\nA. N. Rossolenko, M. Khabipov, D. Balashov, A. B. Zorin, P. N. Dmitriev, V. P. Koshelets,\nand A. V. Ustinov, “Implementation of superconductor/ferromagnet/ superconductor π-\nshifters in superconducting digital and quantum circuits,” Nat. Phys. 6, 593–597 (2010).\n81A. V. Shcherbakova, K. G. Fedorov, K. V. Shulga, V. V. Ryazanov, V. V. Bolginov,\nV. A. Oboznov, S. V. Egorov, V. O. Shkolnikov, M. J. Wolf, D. Beckmann, and A. V.\n49Ustinov, “Fabrication and measurements of hybrid Nb/Al Josephson junctions and flux\nqubits with π-shifters,” Supercond. Sci. Technol. 28, 025009 (2015).\n82M. Mori and S. Maekawa, “Half-integer Shapiro-steps in superconducting qubit with a\nπ-Josephson junction,” Appl. Phys. Express 14, 103001 (2021).\n83H. G. Ahmad, V. Brosco, A. Miano, L. Di Palma, M. Arzeo, D. Montemurro, P. Lucig-\nnano, G. P. Pepe, F. Tafuri, R. Fazio, and D. Massarotti, “Hybrid ferromagnetic transmon\nqubit: Circuit design, feasibility, and detection protocols for magnetic fluctuations,” Phys.\nRev. B 105, 214522 (2022).\n84D. Massarotti, H. G. Ahmad, R. Satariano, R. Ferraiuolo, L. Di Palma, P. Mastrovito,\nG. Serpico, A. Levochkina, R. Caruso, A. Miano, M. Arzeo, G. Ausanio, C. Granata,\nP. Lucignano, D. Montemurro, L. Parlato, A. Vettoliere, R. Fazio, O. Mukhanov, G. P.\nPepe, and F. Tafuri, “A feasible path for the use of ferromagnetic Josephson junctions\nin quantum circuits: The ferro-transmon,” Low Temp. Phys. 49, 794–802 (2023).\n85S. Kim, L. V. Abdurakhimov, D. Pham, W. Qiu, H. Terai, S. Ashhab, S. Saito, T. Ya-\nmashita, and K. Semba, “Superconducting flux qubit with ferromagnetic Josephson π\njunction operating at zero magnetic field,” arXiv:2401.14597 (2024).\n86A. Somoroff, Q. Ficheux, R. A. Mencia, H. Xiong, R. Kuzmin, and V. E. Manucharyan,\n“Millisecond coherence in a superconducting qubit,” Phys. Rev. Lett. 130, 267001 (2023).\n87I. M. Pop, K. Geerlings, G. Catelani, R. J. Schoelkopf, L. I. Glazman, and M. H.\nDevoret, “Coherent suppression of electromagnetic dissipation due to superconducting\nquasiparticles.” Nature 508, 369–372 (2014).\n88K. K. Likharev, “Superconductor digital electronics,” Physica C: Supercond. 482, 6–18\n(2012).\n89I. I. Soloviev, N. V. Klenov, S. V. Bakurskiy, M. Y. Kupriyanov, A. L. Gudkov, and\nA. S. Sidorenko, “Beyond Moore’s technologies: operation principles of a superconductor\nalternative,” Beilstein J. Nanotechnol 8, 2689–2710 (2017).\n90T. Ortlepp, Ariando, O. Mielke, C. J. M. Verwijs, K. F. K. Foo, H. Rogalla, F. H.\nUhlmann, and H. Hilgenkamp, “Flip-flopping fractional flux quanta,” Science 312, 1495–\n1497 (2006).\n91O. Mielke, T. Ortlepp, B. Dimov, and F. H. Uhlmann, “Phase engineering techniques in\nsuperconducting quantum electronics,” J. Phys.: Conf. Ser. 97, 012196 (2008).\n5092M. I. Khabipov, D. V. Balashov, F. Maibaum, A. B. Zorin, V. A. Oboznov, V. V.\nBol’ginov, A. N. Rossolenko, and V. V. Ryazanov, “A single flux quantum circuit with\na ferromagnet-based Josephson π-junction,” Supercond. Sci. Technol. 23, 045032 (2010).\n93O. Wetzstein, T. Ortlepp, R. Stolz, J. Kunert, H.-G. Meyer, and H. Toepfer, “Com-\nparison of RSFQ logic cells with and without phase shifting elements by means of BER\nmeasurements,” IEEE Trans. Appl. Supercond. 21, 814–817 (2011).\n94T. Kamiya, M. Tanaka, K. Sano, and A. Fujimaki, “Energy/Space-Efficient Rapid Single-\nFlux-Quantum Circuits by Using π-Shifted Josephson Junctions,” IEICE Trans. Electron.\nE101.C , 385–390 (2018).\n95Y. Yamanashi, S. Nakaishi, A. Sugiyama, N. Takeuchi, and N. Yoshikawa, “Design\nmethodology of single-flux-quantum flip-flops composed of both 0- and π-shifted Joseph-\nson junctions,” Supercond. Sci. Technol. 31, 105003 (2018).\n96D. Hasegawa, Y. Takeshita, K. Sano, M. Tanaka, A. Fujimaki, and T. Yamashita, “Mag-\nnetic Josephson junctions on Nb four-layer structure for half flux quantum circuits,” in\n2019 IEEE International Superconductive Electronics Conference (ISEC) (IEEE, 2019)\npp. 1–3.\n97K. Arai, N. Takeuchi, T. Yamashita, and N. Yoshikawa, “Adiabatic quantum-flux-\nparametron with πJosephson junctions,” J. Appl. Phys. 125, 093901 (2019).\n98Y. Yamanashi, S. Nakaishi, and N. Yoshikawa, “Simulation of the Margins in Single\nFlux Quantum Circuits Containing π-Shifted Josephson Junctions,” IEEE Trans. Appl.\nSupercond. 29, 1301805 (2019).\n99T. Yamashita, “Magnetic Josephson Junctions: New Phenomena and Physics with Di-\nluted Alloy, Conventional Ferromagnet, and Multilayer Barriers,” IEICE Trans. Electron.\nE104.C , 422–428 (2021).\n100Y. Takeshita, F. Li, D. Hasegawa, K. Sano, M. Tanaka, T. Yamashita, and A. Fujimaki,\n“High-Speed Memory Driven by SFQ Pulses Based on 0- πSQUID,” IEEE Trans. Appl.\nSupercond. 31, 1100906 (2021).\n101F. Li, Y. Takeshita, D. Hasegawa, M. Tanaka, T. Yamashita, and A. Fujimaki, “Low-\npower high-speed half-flux-quantum circuits driven by low bias voltages,” Supercond. Sci.\nTechnol. 34, 025013 (2021).\n102F. Li, D. Pham, Y. Takeshita, M. Higashi, T. Yamashita, M. Tanaka, and A. Fuji-\nmaki, “Energy efficient half-flux-quantum circuit aiming at milli-kelvin stage operation,”\n51Supercond. Sci. Technol. 36, 105006 (2023).\n103F. Li, Y. Takeshita, M. Tanaka, and A. Fujimaki, “Superconductor digital circuits with\nπjunctions alone,” Appl. Phys. Lett 122, 162601 (2023).\n104S. Tanemura, Y. Takeshita, F. Li, T. Nakayama, M. Tanaka, and A. Fujimaki, “Optimiza-\ntion of Half-Flux-Quantum Circuits Composed of π-Shift and Conventional Josephson\nJunctions,” IEEE Trans. Appl. Supercond. 33, 1701305 (2023).\n105I. I. Soloviev, V. I. Ruzhickiy, S. V. Bakurskiy, N. V. Klenov, M. Y. Kupriyanov, A. A.\nGolubov, O. V. Skryabina, and V. S. Stolyarov, “Superconducting Circuits without\nInductors Based on Bistable Josephson Junctions,” Phys. Rev. Appl. 16, 014052 (2021).\n106I. I. Soloviev, G. S. Khismatullin, N. V. Klenov, and A. E. Schegolev, “ πjunctions\nin adiabatic superconductor logic cells,” J. Commun. Technol. Electron. 67, 1479–1491\n(2022).\n107A. A. Maksimovskaya, V. I. Ruzhickiy, N. V. Klenov, S. V. Bakurskiy, M. Y. Kupriyanov,\nand I. I. Soloviev, “Phase logic based on πJosephson junctions,” Jetp Lett. 115, 735–741\n(2022).\n108G. S. Khismatullin, N. V. Klenov, and I. I. Soloviev, “Optimization of adiabatic super-\nconducting logic cells by using πJosephson junctions,” JETP Lett. 118, 220–229 (2023).\n109T. I. Larkin, V. V. Bol’ginov, V. S. Stolyarov, V. V. Ryazanov, I. V. Vernik, S. K. Tolpygo,\nand O. A. Mukhanov, “Ferromagnetic Josephson switching device with high characteristic\nvoltage,” Appl. Phys. Lett. 100, 222601 (2012).\n110I. V. Vernik, V. V. Bol’ginov, S. V. Bakurskiy, A. A. Golubov, M. Y. Kupriyanov, V. V.\nRyazanov, and O. A. Mukhanov, “Magnetic Josephson junctions with superconducting\ninterlayer for cryogenic memory,” IEEE Trans. Appl. Supercond. 23, 1701208 (2013).\n111S. V. Bakurskiy, N. V. Klenov, I. I. Soloviev, V. V. Bol’ginov, V. V. Ryazanov, I. V.\nVernik, O. A. Mukhanov, M. Y. Kupriyanov, and A. A. Golubov, “Theoretical model of\nsuperconducting spintronic SIsFS devices,” Appl. Phys. Lett 102, 192603 (2013).\n112L. Parlato, R. Caruso, A. Vettoliere, R. Satariano, H. G. Ahmad, A. Miano, D. Monte-\nmurro, D. Salvoni, G. Ausanio, F. Tafuri, G. P. Pepe, D. Massarotti, and C. Granata,\n“Characterization of scalable Josephson memory element containing a strong ferromag-\nnet,” J. Appl. Phys. 127, 193901 (2020).\n113H. Hilgenkamp, Ariando, H.-J. H. Smilde, D. H. A. Blank, G. Rijnders, H. Rogalla,\nJ. R. Kirtley, and C. C. Tsuei, “Ordering and manipulation of the magnetic moments in\n52large-scale superconducting π-loop arrays,” Nature 422, 50–53 (2003).\n114D. Balashov, B. Dimov, M. Khabipov, T. Ortlepp, D. Hagedorn, A. Zorin, F.-I. Buchholz,\nF. Uhlmann, and J. Niemeyer, “Passive phase shifter for superconducting Josephson\ncircuits,” IEEE Trans. Appl. Supercond. 17, 142–145 (2007).\n115J. B. Majer, J. R. Butcher, and J. E. Mooij, “Simple phase bias for superconducting\ncircuits,” Appl. Phys. Lett 80, 3638–3640 (2002).\n116M. J. A. Stoutimore, A. N. Rossolenko, V. V. Bolginov, V. A. Oboznov, A. Y. Rusanov,\nD. S. Baranov, N. Pugach, S. M. Frolov, V. V. Ryazanov, and D. J. Van Harlingen,\n“Second-Harmonic Current-Phase Relation in Josephson Junctions with Ferromagnetic\nBarriers,” Phys. Rev. Lett. 121, 177702 (2018).\n117A. Pal, Z. Barber, J. Robinson, and M. Blamire, “Pure second harmonic current-phase\nrelation in spin-filter Josephson junctions,” Nat. Commun. 5, 3340 (2014).\n118L. Trifunovic, “Long-range superharmonic Josephson current.” Phys Rev Lett 107, 047001\n(2011).\n119L. Trifunovic, Z. Popovi´ c, and Z. Radovi´ c, “Josephson effect and spin-triplet pairing\ncorrelations in SFFS junctions,” Phys. Rev. B 84, 064511 (2011).\n120A. S. Mel’nikov, A. V. Samokhvalov, S. M. Kuznetsova, and A. I. Buzdin, “Interfer-\nence phenomena and long-range proximity effect in clean superconductor-ferromagnet\nsystems.” Phys Rev Lett 109, 237006 (2012).\n121C. Richard, M. Houzet, and J. S. Meyer, “Superharmonic long-range triplet current in a\ndiffusive Josephson junction.” Phys Rev Lett 110, 217004 (2013).\n122I. Salameh, E. G. Friedman, and S. Kvatinsky, “Superconductive logic using 2 ϕJosephson\njunctions with half flux quantum pulses,” IEEE Trans. Circuits Syst. II 69, 2533–2537\n(2022).\n123T. Jabbari, M. Bocko, and E. G. Friedman, “All-JJ logic based on bistable JJs,” IEEE\nTrans. Appl. Supercond. 33, 1–7 (2023).\n124S. Razmkhah and M. Pedram, “High-density superconductive logic circuits utilizing 0 and\nπJosephson junctions,” Eng. Res. Express 6, 015307 (2024).\n125H. Cong, S. Razmkhah, M. A. Karamuftuoglu, and M. Pedram, “Superconductor logic\nimplementation with all-JJ inductor-free cell library,” arXiv:2310.13857 (2023).\n126V. V. Bol’ginov, V. S. Stolyarov, D. S. Sobanin, A. L. Karpovich, and V. V. Ryazanov,\n“Magnetic switches based on Nb-PdFe-Nb Josephson junctions with a magnetically soft\n53ferromagnetic interlayer,” JETP Lett. 95, 366–371 (2012).\n127L. N. Karelina, R. A. Hovhannisyan, I. A. Golovchanskiy, V. I. Chichkov, A. Ben Hamida,\nV. S. Stolyarov, L. S. Uspenskaya, S. A. Erkenov, V. V. Bolginov, and V. V. Ryazanov,\n“Scalable memory elements based on rectangular SIsFS junctions,” J. Appl. Phys. 130,\n173901 (2021).\n128R. Fermin, N. M. A. Scheinowitz, J. Aarts, and K. Lahabi, “Mesoscopic superconducting\nmemory based on bistable magnetic textures,” Phys. Rev. Research 4, 033136 (2022).\n129T. Golod, A. Iovan, and V. M. Krasnov, “Single Abrikosov vortices as quantized infor-\nmation bits.” Nat. Commun. 6, 8628 (2015).\n130I. I. Soloviev, N. V. Klenov, S. V. Bakurskiy, V. V. Bol’ginov, V. V. Ryazanov, M. Y.\nKupriyanov, and A. A. Golubov, “Josephson magnetic rotary valve,” Appl. Phys. Lett.\n105, 242601 (2014).\n131I. P. Nevirkovets, S. E. Shafraniuk, and O. A. Mukhanov, “Multiterminal\nSuperconducting-Ferromagnetic Device with Magnetically Tunable Supercurrent for\nMemory Application,” IEEE Trans. Appl. Supercond. 28, 1800904 (2018).\n132I. P. Nevirkovets and O. A. Mukhanov, “Memory Cell for High-Density Arrays Based on\na Multiterminal Superconducting-Ferromagnetic Device,” Phys. Rev. Appl. 10, 034013\n(2018).\n133F. S. Bergeret, A. F. Volkov, and K. B. Efetov, “Enhancement of the Josephson Current\nby an Exchange Field in Superconductor-Ferromagnet Structures,” Phys. Rev. Lett. 86,\n3140–3143 (2001).\n134V. N. Krivoruchko and E. A. Koshina, “From inversion to enhancement of the dc Joseph-\nson current in S/F−I−F/S tunnel structures,” Phys. Rev. B 64, 172511 (2001).\n135A. A. Golubov, M. Y. Kupriyanov, and Y. V. Fominov, “Critical current in SFIFS\njunctions,” Jetp Lett. 75, 190–194 (2002).\n136Y. S. Barash, I. V. Bobkova, and T. Kopp, “Josephson current in S-FIF-S junctions:\nNonmonotonic dependence on misorientation angle,” Phys. Rev. B 66, 140503 (2002).\n137N. M. Chtchelkatchev, W. Belzig, and C. Bruder, “Josephson effect in SFXSF junctions,”\nJETP Lett. 75, 772–776 (2002).\n138A. Zaitsev, “Josephson effect in superconducting junctions with different types of magnetic\nbarrier between the superconductors,” Physica B Condens. Matter. 329-333 , 1498–1499\n(2003).\n54139Y. M. Blanter and F. W. J. Hekking, “Supercurrent in long SFFS junctions with antipar-\nallel domain configuration,” Phys. Rev. B 69, 024525 (2004).\n140Z. Pajovi´ c, M. Boˇ zovi´ c, Z. Radovi´ c, J. Cayssol, and A. Buzdin, “Josephson coupling\nthrough ferromagnetic heterojunctions with noncollinear magnetizations,” Phys. Rev. B\n74, 184509 (2006).\n141B. Crouzy, S. Tollis, and D. A. Ivanov, “Josephson current in a superconductor-\nferromagnet junction with two noncollinear magnetic domains,” Phys. Rev. B 75, 054503\n(2007).\n142C. Bell, G. Burnell, C. W. Leung, E. J. Tarte, D.-J. Kang, and M. G. Blamire, “Con-\ntrollable Josephson current through a pseudospin-valve structure,” Appl. Phys. Lett. 84,\n1153–1155 (2004).\n143B. Baek, W. H. Rippard, S. P. Benz, S. E. Russek, and P. D. Dresselhaus, “Hybrid\nsuperconducting-magnetic memory device using competing order parameters,” Nat. Com-\nmun. 5, 3888 (2014).\n144M. Abd El Qader, R. K. Singh, S. N. Galvin, L. Yu, J. M. Rowell, and N. Newman,\n“Switching at small magnetic fields in Josephson junctions fabricated with ferromagnetic\nbarrier layers,” Appl. Phys. Lett. 104, 022602 (2014).\n145E. C. Gingrich, B. M. Niedzielski, J. A. Glick, Y. Wang, D. L. Miller, R. Loloee, W. P.\nPratt, and N. O. Birge, “Controllable 0- πJosephson junctions containing a ferromagnetic\nspin valve.” Nat. Phys. 12, 564–567 (2016).\n146A. E. Madden, J. C. Willard, R. Loloee, and N. O. Birge, “Phase controllable Josephson\njunctions for cryogenic memory,” Supercond. Sci. Technol. 32, 015001 (2018).\n147I. M. Dayton, T. Sage, E. C. Gingrich, M. G. Loving, T. F. Ambrose, N. P. Siwak, S. Kee-\nbaugh, C. Kirby, D. L. Miller, A. Y. Herr, Q. P. Herr, and O. Naaman, “Experimental\ndemonstration of a Josephson magnetic memory cell with a programmable π-junction,”\nIEEE Magn. Lett. 9, 3301905 (2018).\n148S. V. Bakurskiy, N. V. Klenov, I. I. Soloviev, M. Y. Kupriyanov, and A. A. Golubov,\n“Superconducting phase domains for memory applications,” Appl. Phys. Lett 108, 042602\n(2016).\n149S. V. Bakurskiy, N. V. Klenov, I. I. Soloviev, N. G. Pugach, M. Y. Kupriyanov, and A. A.\nGolubov, “Protected 0- πstates in SIsFS junctions for Josephson memory and logic,” Appl.\nPhys. Lett 113, 082602 (2018).\n55150C. Fierz, S. F. Lee, J. Bass, W. P. Pratt, and P. A. Schroeder, “Superconduc-\ntor/ferromagnet boundary resistances,” J. Phys.: Condens. Matter 2, 9701–9706 (1990).\n151Y. Blum, A. Tsukernik, M. Karpovski, and A. Palevski, “Oscillations of the Supercon-\nducting Critical Current in Nb-Cu-Ni-Cu-Nb Junctions,” Phys. Rev. Lett. 89, 187004\n(2002).\n152V. Shelukhin, A. Tsukernik, M. Karpovski, Y. Blum, K. B. Efetov, A. F. Volkov, T. Cham-\npel, M. Eschrig, T. L¨ ofwander, G. Sch¨ on, and A. Palevski, “Observation of periodic π-\nphase shifts in ferromagnet-superconductor multilayers,” Phys. Rev. B 73, 174506 (2006).\n153J. W. A. Robinson, S. Piano, G. Burnell, C. Bell, and M. G. Blamire, “Critical Current\nOscillations in Strong Ferromagnetic πJunctions,” Phys. Rev. Lett. 97, 177003 (2006).\n154A. A. Bannykh, J. Pfeiffer, V. S. Stolyarov, I. E. Batov, V. V. Ryazanov, and M. Weides,\n“Josephson tunnel junctions with a strong ferromagnetic interlayer,” Phys. Rev. B 79,\n054501 (2009).\n155B. Baek, M. L. Schneider, M. R. Pufall, and W. H. Rippard, “Phase Offsets in the\nCritical-Current Oscillations of Josephson Junctions Based on Ni and Ni-(Ni 81Fe19)xNby\nBarriers,” Phys. Rev. Applied 7, 064013 (2017).\n156S. K. Tolpygo, V. Bolkhovsky, R. Rastogi, S. Zarr, A. L. Day, E. Golden, T. J. Weir,\nA. Wynn, and L. M. Johnson, “Planarized Fabrication Process With Two Layers of SIS\nJosephson Junctions and Integration of SIS and SFS π-Junctions,” IEEE Trans. Appl.\nSupercond. 29, 1101208 (2019).\n157O. M. Kapran, T. Golod, A. Iovan, A. S. Sidorenko, A. A. Golubov, and V. M.\nKrasnov, “Crossover between short- and long-range proximity effects in superconduc-\ntor/ferromagnet/superconductor junctions with Ni-based ferromagnets,” Phys. Rev. B\n103, 094509 (2021).\n158C. S¨ urgers, T. Hoss, C. Sch¨ onenberger, and C. Strunk, “Fabrication and superconducting\nproperties of nanostructured SFS contacts,” J. Magn. Magn. Mater. 240, 598–600 (2002),\n4th International Symposium on Metallic Multilayers.\n159J. W. A. Robinson, Z. H. Barber, and M. G. Blamire, “Strong ferromagnetic Josephson\ndevices with optimized magnetism,” Appl. Phys. Lett. 95, 192509 (2009).\n160J. W. A. Robinson, S. Piano, G. Burnell, C. Bell, and M. G. Blamire, “Zero to π\ntransition in superconductor-ferromagnet-superconductor junctions,” Phys. Rev. B 76,\n094522 (2007).\n56161J. D. S. Witt, J. W. A. Robinson, and M. G. Blamire, “Josephson junctions incorporating\na conical magnetic holmium interlayer,” Phys. Rev. B 85, 184526 (2012).\n162J. W. A. Robinson, S. Piano, G. Burnell, C. Bell, and M. G. Blamire, “Transport and\nMagnetic Properties of Strong Ferromagnetic Pi-Junctions,” IEEE Trans. Appl. Super-\ncond. 17, 641–644 (2007).\n163B. Baek, M. L. Schneider, M. R. Pufall, and W. H. Rippard, “Anomalous Supercur-\nrent Modulation in Josephson Junctions With Ni-Based Barriers,” IEEE Trans. Appl.\nSupercond. 28, 1800705 (2018).\n164S. Piano, J. W. A. Robinson, G. Burnell, and M. G. Blamire, “0- πoscillations in nanos-\ntructured Nb/Fe/Nb Josephson junctions,” Eur. Phys. J. B 58, 123–126 (2007).\n165J. W. A. Robinson, J. D. S. Witt, and M. G. Blamire, “Controlled injection of spin-triplet\nsupercurrents into a strong ferromagnet,” Science 329, 59–61 (2010).\n166J. Kwo, M. Hong, and S. Nakahara, “Growth of rare-earth single crystals by molecular\nbeam epitaxy: The epitaxial relationship between hcp rare earth and bcc niobium,” Appl.\nPhys. Lett. 49, 319–321 (1986).\n167N. Satchell, J. D. S. Witt, G. Burnell, P. J. Curran, C. J. Kinane, T. R. Charlton,\nS. Langridge, and J. F. K. Cooper, “Probing the spiral magnetic phase in 6 nm textured\nerbium using polarised neutron reflectometry,” J. Condens. Matter Phys. 29, 055801\n(2017).\n168Y. N. Khaydukov, A. S. Vasenko, E. A. Kravtsov, V. V. Progliado, V. D. Zhaketov,\nA. Csik, Y. V. Nikitenko, A. V. Petrenko, T. Keller, A. A. Golubov, M. Y. Kupriyanov,\nV. V. Ustinov, V. L. Aksenov, and B. Keimer, “Magnetic and superconducting phase\ndiagram of Nb/Gd/Nb trilayers,” Phys. Rev. B 97, 144511 (2018).\n169C. Bell, R. Loloee, G. Burnell, and M. G. Blamire, “Characteristics of strong ferromag-\nnetic Josephson junctions with epitaxial barriers,” Phys. Rev. B 71, 180501(R) (2005).\n170J. A. Glick, M. A. Khasawneh, B. M. Niedzielski, R. Loloee, W. P. Pratt, N. O. Birge,\nE. C. Gingrich, P. G. Kotula, and N. Missert, “Critical current oscillations of elliptical\nJosephson junctions with single-domain ferromagnetic layers,” J. Appl. Phys. 122, 133906\n(2017).\n171Y. Yao, R. Cai, S.-H. Yang, W. Xing, Y. Ma, M. Mori, Y. Ji, S. Maekawa, X.-C. Xie,\nand W. Han, “Half-integer Shapiro steps in strong ferromagnetic Josephson junctions,”\nPhys. Rev. B 104, 104414 (2021).\n57172S. S. Mishra, R. M. Klaes, J. Willard, R. Loloee, and N. O. Birge, “Enhancement of\nsupercurrent through ferromagnetic materials by interface engineering,” Phys. Rev. B\n106, 014519 (2022).\n173B. M. Niedzielski, E. C. Gingrich, R. Loloee, W. P. Pratt, and N. O. Birge, “S/F/S\nJosephson junctions with single-domain ferromagnets for memory applications,” Super-\ncond. Sci. Technol. 28, 085012 (2015).\n174O. Boothby and R. Bozorth, “A new magnetic material of high permeability,” J. Appl.\nPhys. 18, 173–176 (1947).\n175M. Chen, N. Gharsallah, G. L. Gorman, and J. Latimer, “Ternary NiFeX as soft biasing\nfilm in a magnetoresistive sensor,” J. Appl. Phys. 69, 5631–5633 (1991).\n176M. A. Qader, A. Vishina, L. Yu, C. Garcia, R. Singh, N. Rizzo, M. Huang, R. Chamberlin,\nK. Belashchenko, M. van Schilfgaarde, and N. Newman, “The magnetic, electrical and\nstructural properties of copper-permalloy alloys,” J. Magn. Magn. Mater. 442, 45–52\n(2017).\n177H. Sellier, C. Baraduc, F. Lefloch, and R. Calemczuk, “Temperature-induced crossover\nbetween 0 and πstates in S/F/S junctions,” Phys. Rev. B 68, 054531 (2003).\n178V. A. Oboznov, V. V. Bol’ginov, A. K. Feofanov, V. V. Ryazanov, and A. I. Buzdin,\n“Thickness Dependence of the Josephson Ground States of Superconductor-Ferromagnet-\nSuperconductor Junctions,” Phys. Rev. Lett. 96, 197003 (2006).\n179N. Ruppelt, H. Sickinger, R. Menditto, E. Goldobin, D. Koelle, R. Kleiner, O. Vavra,\nand H. Kohlstedt, “Observation of 0– πtransition in SIsFS Josephson junctions,” Appl.\nPhys. Lett. 106, 022602 (2015).\n180V. V. Bolginov, A. N. Rossolenko, A. B. Shkarin, V. A. Oboznov, and V. V. Ryazanov,\n“Fabrication of Optimized Superconducting Phase Inverters Based on Superconductor–\nFerromagnet–Superconductor π-Junctions,” J. Low Temp. Phys. 190, 302–314 (2018).\n181T. Yamashita, A. Kawakami, and H. Terai, “NbN-Based Ferromagnetic 0 and πJoseph-\nson Junctions,” Phys. Rev. Appl. 8, 054028 (2017).\n182J. Zeng, L. Chen, X. Zhong, Y. Wang, Y. Pan, D. Zhang, S. Yu, L. Wu, L. Zhang, W. Peng,\net al. , “Nonvolatile memory cell using a superconducting-ferromagnetic πJosephson junc-\ntion,” Supercond. Sci. Technol. 35, 105009 (2022).\n183D. Pham, R. Sugimoto, K. Oba, Y. Takeshita, F. Li, M. Tanaka, T. Yamashita, and\nA. Fujimaki, “Weak spin-flip scattering in Pd 89Ni11interlayer of NbN-based ferromagnetic\n58Josephson junctions,” Sci. Rep. 12, 6863 (2022).\n184V. Stolyarov, V. Oboznov, D. Kasatonov, A. Neilo, S. Bakurskiy, N. Klenov, I. Soloviev,\nM. Kupriyanov, A. Golubov, T. Cren, et al. , “Effective Exchange Energy in a Thin,\nSpatially Inhomogeneous CuNi Layer Proximized by Nb,” J. Phys. Chem. Lett. 13, 6400–\n6406 (2022).\n185M. Weides, M. Kemmler, H. Kohlstedt, R. Waser, D. Koelle, R. Kleiner, and E. Goldobin,\n“0−πJosephson Tunnel Junctions with Ferromagnetic Barrier,” Phys. Rev. Lett. 97,\n247001 (2006).\n186F. Born, M. Siegel, E. K. Hollmann, H. Braak, A. A. Golubov, D. Y.\nGusakova, and M. Y. Kupriyanov, “Multiple 0 −πtransitions in superconduc-\ntor/insulator/ferromagnet/superconductor Josephson tunnel junctions,” Phys. Rev. B 74,\n140501(R) (2006).\n187D. Sprungmann, K. Westerholt, H. Zabel, M. Weides, and H. Kohlstedt, “Josephson\ntunnel junctions with ferromagnetic Fe 0.75Co0.25barriers,” J. Phys. D: Appl. Phys. 42,\n075005 (2009).\n188N. Satchell, T. Mitchell, P. Shepley, E. Darwin, B. Hickey, and G. Burnell, “Pt and CoB\ntrilayer Josephson πjunctions with perpendicular magnetic anisotropy,” Sci. Rep. 11,\n11173 (2021).\n189M. G. Loving, T. F. Ambrose, S. Keebaugh, D. L. Miller, R. Pownall, N. D.\nRizzo, A. N. Sidorov, and N. P. Siwak, “Magnetic Reversal and Critical Current\nTransparency of CoFeB Superconductor-Ferromagnet-Superconductor Heterostructures,”\narXiv:2104.03188 (2021).\n190S. Komori, J. E. Thompson, G. Yang, G. Kimbell, N. Stelmashenko, M. G. Blamire,\nand J. W. A. Robinson, “Enhancement of Josephson Critical Currents in Ferromagnetic\nCo40Fe40B20by Thermal Annealing,” Phys. Rev. Appl. 17, L021002 (2022).\n191A. E. Madden, Ferromagnetic and Ferrimagnetic Materials as Josephson Junction Bar-\nriers (Michigan State University, 2022).\n192R. Caruso, D. Massarotti, G. Campagnano, A. Pal, H. G. Ahmad, P. Lucignano, M. Es-\nchrig, M. G. Blamire, and F. Tafuri, “Tuning of Magnetic Activity in Spin-Filter Joseph-\nson Junctions Towards Spin-Triplet Transport,” Phys. Rev. Lett. 122, 047002 (2019).\n193K. Kang, H. Berger, K. Watanabe, T. Taniguchi, L. Forr´ o, J. Shan, and K. F. Mak, “van\nder Waals πJosephson Junctions,” Nano Lett. 22, 5510–5515 (2022).\n59194R. Lavrijsen, G. Malinowski, J. H. Franken, J. T. Kohlhepp, H. J. M. Swagten, B. Koop-\nmans, M. Czapkiewicz, and T. Stobiecki, “Reduced domain wall pinning in ultrathin\nPt/Co 100−xBx/Pt with perpendicular magnetic anisotropy,” Appl. Phys. Lett. 96, 022501\n(2010).\n195S. Ikeda, J. Hayakawa, Y. Ashizawa, Y. M. Lee, K. Miura, H. Hasegawa, M. Tsunoda,\nF. Matsukura, and H. Ohno, “Tunnel magnetoresistance of 604% at 300K by suppression\nof Ta diffusion in CoFeB/MgO/CoFeB pseudo-spin-valves annealed at high temperature,”\nAppl. Phys. Lett. 93, 082508 (2008).\n196A. D. Naylor, G. Burnell, and B. J. Hickey, “Transport spin polarization of the rare-earth\ntransition-metal alloy Co 1−xGdx,” Phys. Rev. B 85, 064410 (2012).\n197M. Jimbo, K. Komiyama, Y. Shirota, Y. Fujiwara, S. Tsunashima, and M. Matsuura,\n“Thermal stability of spin valves using amorphous CoFeB,” J. Magn. Magn. Mater. 165,\n308–311 (1997), symposium E: Magnetic Ultrathin Films, Multilayers and Surfaces.\n198A. Vecchione, R. Fittipaldi, C. Cirillo, M. Hesselberth, J. Aarts, S. Prischepa, V. Kushnir,\nM. Y. Kupriyanov, and C. Attanasio, “X-ray scattering study of interfacial roughness in\nNb/PdNi multilayers,” Surf. Sci. 605, 1791–1796 (2011).\n199Y. Khaydukov, R. Morari, O. Soltwedel, T. Keller, G. Christiani, G. Logvenov,\nM. Kupriyanov, A. Sidorenko, and B. Keimer, “Interfacial roughness and proximity\neffects in superconductor/ferromagnet CuNi/Nb heterostructures,” J. Appl. Phys. 118,\n213905 (2015).\n200M. Garcia, E. Fernandez Pinel, J. De la Venta, A. Quesada, V. Bouzas, J. Fern´ andez,\nJ. Romero, M. Gonz´ alez, and J. Costa-Kr¨ amer, “Sources of experimental errors in the\nobservation of nanoscale magnetism,” J. Appl. Phys. 105, 013925 (2009).\n201N. Satchell and N. O. Birge, “Supercurrent in ferromagnetic Josephson junctions with\nheavy metal interlayers,” Phys. Rev. B 97, 214509 (2018).\n202J. Bass and W. P. Pratt, “Spin-diffusion lengths in metals and alloys, and spin-flipping at\nmetal/metal interfaces: an experimentalist’s critical review,” J. Phys. Condens. Matter\n19, 183201 (2007).\n203C. D. Thomas, M. P. Ulmer, and J. B. Ketterson, “Superconducting tunnel junction base\nelectrode planarization,” J. Appl. Phys. 84, 364–367 (1998).\n204Y. Wang, W. P. Pratt, and N. O. Birge, “Area-dependence of spin-triplet supercurrent\nin ferromagnetic Josephson junctions,” Phys. Rev. B 85, 214522 (2012).\n60205P. Quarterman, N. Satchell, B. J. Kirby, R. Loloee, G. Burnell, N. O. Birge,\nand J. A. Borchers, “Distortions to the penetration depth and coherence length of\nsuperconductor/normal-metal superlattices,” Phys. Rev. Materials 4, 074801 (2020).\n206C. Bell, G. Burnell, C.W. Leung, E. J. Tarte, and M. G. Blamire, “Spin valve Josephson\njunctions,” IEEE Trans. Appl. Supercond. 15, 908–911 (2005).\n207B. Baek, W. H. Rippard, M. R. Pufall, S. P. Benz, S. E. Russek, H. Rogalla, and P. D.\nDresselhaus, “Spin-Transfer Torque Switching in Nanopillar Superconducting-Magnetic\nHybrid Josephson Junctions,” Phys. Rev. Applied 3, 011001 (2015).\n208B. M. Niedzielski, T. J. Bertus, J. A. Glick, R. Loloee, W. P. Pratt, and N. O. Birge,\n“Spin-valve Josephson junctions for cryogenic memory,” Phys. Rev. B 97, 024517 (2018).\n209N. Satchell, P. M. Shepley, M. Algarni, M. Vaughan, E. Darwin, M. Ali, M. C. Rosamond,\nL. Chen, E. H. Linfield, B. J. Hickey, and G. Burnell, “Spin-valve Josephson junctions\nwith perpendicular magnetic anisotropy for cryogenic memory,” Appl. Phys. Lett. 116,\n022601 (2020).\n210I. O. Kulik, “Magnitude of the critical Josephson tunnel current,” JETP 22, 841–843\n(1966).\n211D. Massarotti, A. Pal, G. Rotoli, L. Longobardi, M. G. Blamire, and F. Tafuri, “Macro-\nscopic quantum tunnelling in spin filter ferromagnetic Josephson junctions,” Nat. Com-\nmun. 6, 7376 (2015).\n212D. Massarotti, R. Caruso, A. Pal, G. Rotoli, L. Longobardi, G. P. Pepe, M. Blamire,\nand F. Tafuri, “Low temperature properties of spin filter NbN/GdN/NbN Josephson\njunctions,” Physica C Supercond. 533, 53–58 (2017).\n213J. P. Cascales, Y. Takamura, G. M. Stephen, D. Heiman, F. S. Bergeret, and J. S. Mood-\nera, “Switchable Josephson junction based on interfacial exchange field,” Appl. Phys.\nLett. 114, 022601 (2019).\n214H. Ahmad, R. Caruso, A. Pal, G. Rotoli, G. Pepe, M. Blamire, F. Tafuri, and D. Mas-\nsarotti, “Electrodynamics of Highly Spin-Polarized Tunnel Josephson Junctions,” Phys.\nRev. Appl. 13, 014017 (2020).\n215H. G. Ahmad, M. Minutillo, R. Capecelatro, A. Pal, R. Caruso, G. Passarelli, M. G.\nBlamire, F. Tafuri, P. Lucignano, and D. Massarotti, “Coexistence and tuning of spin-\nsinglet and triplet transport in spin-filter Josephson junctions,” Commun. Phys. 5, 2\n(2022).\n61216D. Razmadze, R. S. Souto, L. Galletti, A. Maiani, Y. Liu, P. Krogstrup, C. Schrade,\nA. Gyenis, C. M. Marcus, and S. Vaitiek˙ enas, “Supercurrent reversal in ferromagnetic\nhybrid nanowire Josephson junctions,” Phys. Rev. B 107, L081301 (2023).\n217L. Ai, E. Zhang, J. Yang, X. Xie, Y. Yang, Z. Jia, Y. Zhang, S. Liu, Z. Li, P. Leng, et al. ,\n“Van der Waals ferromagnetic Josephson junctions,” Nat. Commun. 12, 6580 (2021).\n218H. Idzuchi, F. Pientka, K.-F. Huang, K. Harada, ¨O. G¨ ul, Y. J. Shin, L. Nguyen, N. Jo,\nD. Shindo, R. Cava, et al. , “Unconventional supercurrent phase in Ising superconductor\nJosephson junction with atomically thin magnetic insulator,” Nat. Commun. 12, 5332\n(2021).\n219G. Hu, C. Wang, S. Wang, Y. Zhang, Y. Feng, Z. Wang, Q. Niu, Z. Zhang, and B. Xi-\nang, “Long-range skin Josephson supercurrent across a van der Waals ferromagnet,” Nat.\nCommun. 14, 1779 (2023).\n220T. T. Heikkil¨ a, F. K. Wilhelm, and G. Sch¨ on, “Non-equilibrium supercurrent through\nmesoscopic ferromagnetic weak links,” EPL 51, 434 (2000).\n221S.-K. Yip, “Magnetic-field effect on the supercurrent of an SNS junction,” Phys. Rev. B\n62, R6127–R6130 (2000).\n222C. Li, B. de Ronde, J. de Boer, J. Ridderbos, F. Zwanenburg, Y. Huang, A. Golubov,\nand A. Brinkman, “Zeeman-Effect-Induced 0 −πTransitions in Ballistic Dirac Semimetal\nJosephson Junctions,” Phys. Rev. Lett. 123, 026802 (2019).\n223T. Dvir, A. Zalic, E. H. Fyhn, M. Amundsen, T. Taniguchi, K. Watanabe, J. Linder, and\nH. Steinberg, “Planar graphene-NbSe 2Josephson junctions in a parallel magnetic field,”\nPhys. Rev. B 103, 115401 (2021).\n224J. Sch¨ onle, K. Borisov, R. Klett, D. Dyck, F. Balestro, G. Reiss, and W. Wernsdorfer,\n“Field-tunable 0- π-transitions in SnTe topological crystalline insulator SQUIDs,” Sci.\nRep.9, 1987 (2019).\n225C. T. Ke, C. M. Moehle, F. K. de Vries, C. Thomas, S. Metti, C. R. Guinn, R. Kallaher,\nM. Lodari, G. Scappucci, T. Wang, R. E. Diaz, G. C. Gardner, M. J. Manfra, and\nS. Goswami, “Ballistic superconductivity and tunable π-junctions in InSb quantum wells,”\nNat. Commun. 10, 3764 (2019).\n226D. Z. Haxell, M. Coraiola, D. Sabonis, M. Hinderling, S. C. Ten Kate, E. Cheah, F. Krizek,\nR. Schott, W. Wegscheider, and F. Nichele, “Zeeman-and Orbital-Driven Phase Shifts in\nPlanar Josephson Junctions,” ACS Nano 17, 18139–18147 (2023).\n62227A. M. Whiticar, A. Fornieri, A. Banerjee, A. C. C. Drachmann, S. Gronin, G. C. Gardner,\nT. Lindemann, M. J. Manfra, and C. M. Marcus, “Zeeman-driven parity transitions in\nan Andreev quantum dot,” Phys. Rev. B 103, 245308 (2021).\n228K. K. Likharev, “Single-electron devices and their applications,” Proceedings of the IEEE\n87, 606–632 (1999).\n229J. A. Van Dam, Y. V. Nazarov, E. P. Bakkers, S. De Franceschi, and L. P. Kouwenhoven,\n“Supercurrent reversal in quantum dots,” Nature 442, 667–670 (2006).\n230H. I. Jørgensen, T. Novotn` y, K. Grove-Rasmussen, K. Flensberg, and P. Lindelof, “Crit-\nical current 0- πtransition in designed Josephson quantum dot junctions,” Nano Lett. 7,\n2441–2445 (2007).\n231R. Maurand, T. Meng, E. Bonet, S. Florens, L. Marty, and W. Wernsdorfer, “First-order\n0−πQuantum Phase Transition in the Kondo Regime of a Superconducting Carbon-\nNanotube Quantum Dot,” Phys. Rev. X 2, 011009 (2012).\n232R. Delagrange, R. Weil, A. Kasumov, M. Ferrier, H. Bouchiat, and R. Deblock, “0- π\nquantum transition in a carbon nanotube Josephson junction: Universal phase depen-\ndence and orbital degeneracy,” Phys. Rev. B 93, 195437 (2016).\n233T. Salamone, M. B. M. Svendsen, M. Amundsen, and S. Jacobsen, “Curvature-induced\nlong-range supercurrents in diffusive superconductor-ferromagnet-superconductor Joseph-\nson junctions with a dynamic 0 −πtransition,” Phys. Rev. B 104, L060505 (2021).\n234A. J. Skarpeid, H. G. Hugdal, T. Salamone, M. Amundsen, and S. H. Jacobsen,\n“Non-constant geometric curvature for tailored spin-orbit coupling and chirality in\nsuperconductor-magnet heterostructures,” J. Condens. Matter Phys. , 10.1088/1361–\n648X/ad2e23 (2024).\n235M. Giroud, H. Courtois, K. Hasselbach, D. Mailly, and B. Pannetier, “Superconducting\nproximity effect in a mesoscopic ferromagnetic wire,” Phys. Rev. B 58, R11872–R11875\n(1998).\n236M. D. Lawrence and N. Giordano, “Proximity effects in superconductor-ferromagnet junc-\ntions,” J. Phys.: Condens. Matter 11, 1089–1094 (1999).\n237V. T. Petrashov, I. A. Sosnin, I. Cox, A. Parsons, and C. Troadec, “Giant mutual\nproximity effects in ferromagnetic/superconducting nanostructures,” Phys. Rev. Lett. 83,\n3281–3284 (1999).\n63238R. Keizer, S. Goennenwein, T. Klapwijk, G. Miao, G. Xiao, and A. Gupta, “A spin\ntriplet supercurrent through the half-metallic ferromagnet CrO2,” Nature 439, 825–827\n(2006).\n239I. Sosnin, H. Cho, V. T. Petrashov, and A. F. Volkov, “Superconducting Phase Coherent\nElectron Transport in Proximity Conical Ferromagnets,” Phys. Rev. Lett. 96, 157002\n(2006).\n240T. S. Khaire, M. A. Khasawneh, W. P. Pratt, and N. O. Birge, “Observation of spin-\ntriplet superconductivity in Co-based Josephson junctions,” Phys. Rev. Lett. 104, 137002\n(2010).\n241D. Sprungmann, K. Westerholt, H. Zabel, M. Weides, and H. Kohlstedt, “Evidence\nfor triplet superconductivity in Josephson junctions with barriers of the ferromagnetic\nHeusler alloy,” Phys. Rev. B 82, 060505 (2010).\n242M. S. Anwar, M. Veldhorst, A. Brinkman, and J. Aarts, “Long range supercurrents in\nferromagnetic CrO 2using a multilayer contact structure,” Appl. Phys. Lett 100, 052602\n(2012).\n243N. O. Birge, “Spin-triplet supercurrents in Josephson junctions containing strong ferro-\nmagnetic materials,” Philos. Trans. Royal Soc. A 376, 20150150 (2018).\n244M. Houzet and A. I. Buzdin, “Long range triplet Josephson effect through a ferromagnetic\ntrilayer,” Phys. Rev. B 76, 060504 (2007).\n245A. F. Volkov and K. B. Efetov, “Odd spin-triplet superconductivity in a multilayered\nsuperconductor-ferromagnet Josephson junction,” Phys. Rev. B 81, 144522 (2010).\n246L. Trifunovic and Z. Radovi´ c, “Long-range spin-triplet proximity effect in Josephson junc-\ntions with multilayered ferromagnets,” Phys. Rev. B 82, 020505 (2010).\n247A. F. Volkov, A. Anishchanka, and K. B. Efetov, “Odd triplet superconductivity in a\nsuperconductor/ferromagnet system with a spiral magnetic structure,” Phys. Rev. B 73,\n104412 (2006).\n248G. B. Hal´ asz, J. W. A. Robinson, J. F. Annett, and M. G. Blamire, “Critical current of\na Josephson junction containing a conical magnet,” Phys. Rev. B 79, 224505 (2009).\n249J. A. Glick, V. Aguilar, A. B. Gougam, B. M. Niedzielski, E. C. Gingrich, R. Loloee,\nW. P. Pratt, and N. O. Birge, “Phase control in a spin-triplet SQUID,” Sci. Adv. 4,\neaat9457 (2018).\n64250N. O. Birge and M. Houzet, “Spin-Singlet and Spin-Triplet Josephson Junctions for Cryo-\ngenic Memory,” IEEE Magn. Lett. 10, 4509605 (2019).\n251A. Buzdin and A. E. Koshelev, “Periodic alternating 0- and π-junction structures as\nrealization of ϕ-Josephson junctions,” Phys. Rev. B 67, 220504 (2003).\n252E. Goldobin, D. Koelle, R. Kleiner, and A. Buzdin, “Josephson junctions with second\nharmonic in the current-phase relation: Properties of ϕjunctions,” Phys. Rev. B 76,\n224523 (2007).\n253N. G. Pugach, E. Goldobin, R. Kleiner, and D. Koelle, “Method for reliable realization\nof aφJosephson junction,” Phys. Rev. B 81, 104513 (2010).\n254H. Sickinger, A. Lipman, M. Weides, R. G. Mints, H. Kohlstedt, D. Koelle, R. Kleiner,\nand E. Goldobin, “Experimental Evidence of a φJosephson Junction,” Phys. Rev. Lett.\n109, 107002 (2012).\n255S. V. Bakurskiy, N. V. Klenov, T. Y. Karminskaya, M. Y. Kupriyanov, and A. A.\nGolubov, “Josephson ϕ-junctions based on structures with complex normal/ferromagnet\nbilayer,” Supercond. Sci. Technol. 26, 015005 (2013).\n256M. A. Silaev, I. V. Tokatly, and F. S. Bergeret, “Anomalous current in diffusive ferro-\nmagnetic Josephson junctions,” Phys. Rev. B 95, 184508 (2017).\n257V. Braude and Y. V. Nazarov, “Fully developed triplet proximity effect,” Phys. Rev. Lett.\n98, 077003 (2007).\n258A. Buzdin, “Direct coupling between magnetism and superconducting current in the\nJosephson ϕ0junction,” Phys. Rev. Lett. 101, 107005 (2008).\n259A. V. Galaktionov, M. S. Kalenkov, and A. D. Zaikin, “Josephson current and Andreev\nstates in superconductor–half metal–superconductor heterostructures,” Phys. Rev. B 77,\n094520 (2008).\n260R. Grein, M. Eschrig, G. Metalidis, and G. Sch¨ on, “Spin-dependent Cooper pair phase\nand pure spin supercurrents in strongly polarized ferromagnets.” Phys. Rev. Lett. 102,\n227005 (2009).\n261I. Margaris, V. Paltoglou, and N. Flytzanis, “Zero phase difference supercurrent in fer-\nromagnetic Josephson junctions,” J. Phys. Condens. Matter 22, 445701 (2010).\n262J.-F. Liu and K. S. Chan, “Anomalous Josephson current through a ferromagnetic trilayer\njunction,” Phys. Rev. B 82, 184533 (2010).\n65263M. Alidoust and J. Linder, “ φ-state and inverted Fraunhofer pattern in nonaligned Joseph-\nson junctions,” Phys. Rev. B 87, 060503 (2013).\n264I. Kulagina and J. Linder, “Spin supercurrent, magnetization dynamics, and φ-state in\nspin-textured Josephson junctions,” Phys. Rev. B 90, 054504 (2014).\n265F. Konschelle, I. V. Tokatly, and F. S. Bergeret, “Theory of the spin-galvanic effect\nand the anomalous phase shift in superconductors and Josephson junctions with intrinsic\nspin-orbit coupling,” Phys. Rev. B 92, 125443 (2015).\n266S. Pal and C. Benjamin, “Quantized Josephson phase battery,” EPL 126, 57002 (2019).\n267Y. M. Shukrinov, “Anomalous Josephson effect,” Phys.-Usp. 65, 317–354 (2022).\n268I. V. Bobkova, A. M. Bobkov, and M. A. Silaev, “Magnetoelectric effects in Josephson\njunctions,” J. Phys. Condens. Matter 34, 353001 (2022).\n269D. B. Szombati, S. Nadj-Perge, D. Car, S. R. Plissard, E. P. A. M. Bakkers, and L. P.\nKouwenhoven, “Josephson ϕ0-junction in nanowire quantum dots,” Nat. Phys. 12, 568–\n572 (2016).\n270A. Murani, A. Kasumov, S. Sengupta, Y. A. Kasumov, V. T. Volkov, I. I. Khodos,\nF. Brisset, R. Delagrange, A. Chepelianskii, R. Deblock, H. Bouchiat, and S. Gu´ eron,\n“Ballistic edge states in Bismuth nanowires revealed by SQUID interferometry,” Nat.\nCommun. 8, 15941 (2017).\n271A. Assouline, C. Feuillet-Palma, N. Bergeal, T. Zhang, A. Mottaghizadeh, A. Zimmers,\nE. Lhuillier, M. Eddrie, P. Atkinson, M. Aprili, and H. Aubin, “Spin-Orbit induced\nphase-shift in Bi 2Se3Josephson junctions.” Nat. Commun. 10, 126 (2019).\n272H. Ren, F. Pientka, S. Hart, A. T. Pierce, M. Kosowsky, L. Lunczer, R. Schlereth,\nB. Scharf, E. M. Hankiewicz, L. W. Molenkamp, B. I. Halperin, and A. Yacoby, “Topo-\nlogical superconductivity in a phase-controlled Josephson junction,” Nature 569, 93–98\n(2019).\n273W. Mayer, M. C. Dartiailh, J. Yuan, K. S. Wickramasinghe, E. Rossi, and J. Shabani,\n“Gate controlled anomalous phase shift in Al/InAs Josephson junctions,” Nat. Commun.\n11, 212 (2020).\n274E. Strambini, A. Iorio, O. Durante, R. Citro, C. Sanz-Fern´ andez, C. Guarcello, I. V.\nTokatly, A. Braggio, M. Rocci, N. Ligato, V. Zannier, L. Sorba, F. S. Bergeret, and\nF. Giazotto, “A Josephson phase battery,” Nat. Nanotechnol. 15, 656–660 (2020).\n66275C. Baumgartner, L. Fuchs, A. Costa, J. Pic´ o-Cort´ es, S. Reinhardt, S. Gronin, G. C.\nGardner, T. Lindemann, M. J. Manfra, P. E. Faria Junior, D. Kochan, J. Fabian, N. Par-\nadiso, and C. Strunk, “Effect of Rashba and Dresselhaus spin-orbit coupling on super-\ncurrent rectification and magnetochiral anisotropy of ballistic Josephson junctions,” J.\nPhys. Condens. Matter 34, 154005 (2022).\n276D. Margineda, J. S. Claydon, F. Qejvanaj, and C. Checkley, “Observation of anomalous\nJosephson effect in nonequilibrium Andreev interferometers,” Phys. Rev. B 107, L100502\n(2023).\n277C. Bell, G. Burnell, D.-J. Kang, R. H. Hadfield, M. J. Kappers, and M. G. Blamire,\n“Fabrication of nanoscale heterostructure devices with a focused ion beam microscope,”\nNanotechnology 14, 630 (2003).\n278W. H. Meiklejohn and C. P. Bean, “New magnetic anisotropy,” Phys. Rev. 105, 904–913\n(1957).\n279C. Bell, E. J. Tarte, G. Burnell, C. W. Leung, D.-J. Kang, and M. G. Blamire, “Proximity\nand Josephson effects in superconductor/antiferromagnetic Nb/ γ– Fe 50Mn50heterostruc-\ntures,” Phys. Rev. B 68, 144517 (2003).\n280M. Weides, M. Disch, H. Kohlstedt, and D. E. B¨ urgler, “Observation of Josephson\ncoupling through an interlayer of antiferromagnetically ordered chromium,” Phys. Rev.\nB80, 064508 (2009).\n281R. M. Klaes, R. Loloee, and N. O. Birge, “Critical Current Decay in Josephson Junctions\nContaining Antiferromagnetic NiMn,” IEEE Trans. Appl. Supercond. 33, 1800903 (2023).\n282A. Vettoliere, R. Satariano, R. Ferraiuolo, L. Di Palma, H. G. Ahmad, G. Ausanio,\nG. P. Pepe, F. Tafuri, D. Massarotti, D. Montemurro, et al. , “High-Quality ferromagnetic\nJosephson junctions based on aluminum electrodes,” Nanomater. 12, 4155 (2022).\n283A. Vettoliere, R. Satariano, R. Ferraiuolo, L. Di Palma, H. Ahmad, G. Ausanio, G. Pepe,\nF. Tafuri, D. Montemurro, C. Granata, et al., “Aluminum-ferromagnetic Josephson tunnel\njunctions for high quality magnetic switching devices,” Appl. Phys. Lett. 120, 262601\n(2022).\n284S. Tavares, K. Yang, and M. A. Meyers, “Heusler alloys: Past, properties, new alloys,\nand prospects,” Prog. Mater. Sci. 132, 101017 (2023).\n285H. K¯ ono, “On the Ferromagnetic Phase in Manganese-Aluminum System,” J. Phys. Soc.\nJapan. 13, 1444–1451 (1958).\n67286T. A. Bither and W. H. Cloud, “Magnetic Tetragonal δPhase in the Mn–Ga Binary,” J.\nAppl. Phys. 36, 1501–1502 (2004).\n287X. Waintal and P. W. Brouwer, “Magnetic exchange interaction induced by a Josephson\ncurrent,” Phys. Rev. B 65, 054407 (2002).\n288E. Zhao and J. A. Sauls, “Theory of nonequilibrium spin transport and spin-transfer\ntorque in superconducting-ferromagnetic nanostructures,” Phys. Rev. B 78, 174511\n(2008).\n289F. Konschelle and A. Buzdin, “Magnetic Moment Manipulation by a Josephson Current,”\nPhys. Rev. Lett. 102, 017001 (2009).\n290J. Linder and T. Yokoyama, “Supercurrent-induced magnetization dynamics in a Joseph-\nson junction with two misaligned ferromagnetic layers,” Phys. Rev. B 83, 012501 (2011).\n291Q. Shao, P. Li, L. Liu, H. Yang, S. Fukami, A. Razavi, H. Wu, K. Wang, F. Freimuth,\nY. Mokrousov, M. D. Stiles, S. Emori, A. Hoffmann, J. ˚Akerman, K. Roy, J.-P. Wang,\nS.-H. Yang, K. Garello, and W. Zhang, “Roadmap of Spin–Orbit Torques,” IEEE Trans.\nMagn. 57, 1–39 (2021).\n292M.-H. Nguyen, G. J. Ribeill, M. V. Gustafsson, S. Shi, S. V. Aradhya, A. P. Wagner,\nL. M. Ranzani, L. Zhu, R. Baghdadi, B. Butters, et al. , “Cryogenic memory architec-\nture integrating spin Hall effect based magnetic memory and superconductive cryotron\ndevices,” Sci. Rep. 10, 248 (2020).\n293L.ˇSmejkal, J. Sinova, and T. Jungwirth, “Emerging Research Landscape of Altermag-\nnetism,” Phys. Rev. X 12, 040501 (2022).\n294S.-B. Zhang, L.-H. Hu, and T. Neupert, “Finite-momentum Cooper pairing in proximi-\ntized altermagnets,” Nat. Commun. 15, 1801 (2024).\n295J. A. Ouassou, A. Brataas, and J. Linder, “dc Josephson Effect in Altermagnets,” Phys.\nRev. Lett. 131, 076003 (2023).\n296C. W. J. Beenakker and T. Vakhtel, “Phase-shifted Andreev levels in an altermagnet\nJosephson junction,” Phys. Rev. B 108, 075425 (2023).\n297M. Papaj, “Andreev reflection at the altermagnet-superconductor interface,” Phys. Rev.\nB108, L060508 (2023).\n298Z. Feng, X. Zhou, L. ˇSmejkal, L. Wu, Z. Zhu, H. Guo, R. Gonz´ alez-Hern´ andez, X. Wang,\nH. Yan, P. Qin, et al. , “An anomalous Hall effect in altermagnetic ruthenium dioxide,”\nNat. Electron. 5, 735–743 (2022).\n68299P. Crotty, D. Schult, and K. Segall, “Josephson junction simulation of neurons,” Phys.\nRev. E 82, 011914 (2010).\n300M. Schneider, E. Toomey, G. Rowlands, J. Shainline, P. Tschirhart, and K. Segall,\n“SuperMind: a survey of the potential of superconducting electronics for neuromorphic\ncomputing,” Supercond. Sci. Technol. 35, 053001 (2022).\n301A. E. Schegolev, N. V. Klenov, S. V. Bakurskiy, I. I. Soloviev, M. Y. Kupriyanov, M. V.\nTereshonok, and A. S. Sidorenko, “Tunable superconducting neurons for networks based\non radial basis functions,” Beilstein J. Nanotechnol. 13, 444–454 (2022).\n302M. L. Schneider, C. A. Donnelly, and S. E. Russek, “Tutorial: High-speed low-power neu-\nromorphic systems based on magnetic Josephson junctions,” J. Appl. Phys. 124(2018),\n10.1063/1.5042425.\n303M. L. Schneider, C. A. Donnelly, S. E. Russek, B. Baek, M. R. Pufall, P. F. Hopkins,\nP. D. Dresselhaus, S. P. Benz, and W. H. Rippard, “Ultralow power artificial synapses\nusing nanotextured magnetic Josephson junctions,” Sci. Adv. 4, e1701329 (2018).\n304E. Ju´ e, G. Iankevich, T. Reisinger, H. Hahn, V. Provenzano, M. R. Pufall, I. W. Haygood,\nW. H. Rippard, and M. L. Schneider, “Artificial synapses based on Josephson junctions\nwith Fe nanoclusters in the amorphous Ge barrier,” J. Appl. Phys. 131, 073902 (2022).\n305E. Ju´ e, M. R. Pufall, I. W. Haygood, W. H. Rippard, and M. L. Schneider, “Perspectives\non nanoclustered magnetic Josephson junctions as artificial synapses,” Appl. Phys. Lett.\n121, 240501 (2022).\n69" }, { "title": "1310.3590v1.Unveiling_the_Origin_of_the_Insulating_Ferromagnetism_in_LaMnO3_Thin_Film.pdf", "content": "Unveiling the Origin of the Insulating Ferromagnetism in LaMnO 3 \nThin Film \nY . S. Hou , H. J. Xiang*, and X. G. Gong* \nKey Laboratory of Computational Physical Sciences (Ministry of Education), State \nKey Laboratory of Surface Physics, and Department of Physics, Fudan University, \nShanghai 200433, People’s Republic of China \n \nBy combining genetic algorithm optimizations , first -principles calculations and the \ndouble -exchange model studies, we have unveiled that the exotic insulating \nferromagnetism in LaMnO 3 thin film originates from the previously unreported \nG-type \n2 2 2 23z r x ydd orbital ordering. An insulatin g gap opens as a result of both the \norbital ordering and the strong electron -phonon coupling. Therefore , there exist two \nstrain induced phase transition s in the LaMnO 3 thin film, from the insulating A -type \nantiferromagnetic phase to the insulating ferromag netic phase and then to the metallic \nferromagnetic phase. Th ese phase transition s may be exploited in tunneling \nmagnetoresistance and tunneling electroresistance related devices. \n \n \n \nPACS number(s): 31.15.E -, 75.50.Dd , 75.70.Ak , 73.50. -h \n \n \n \n Perovskite LaMnO 3 (LMO), a fundamental s trongly correlated electron system and \nthe parent compound of the colossal magnetoresistance manganites , displays a \ncomplex correlation among structural, orbital, magnetic, and electronic degrees of \nfreedom and has been extensi vely studied o ver the past decades [1]. Recently, LMO \nthin films attract considerable attentions not only because ma ny interesting and \nemerging phenomena are discovered in LMO thin film related superlattices [2-5], but \nalso because LMO thin films display exotic behavior different from bulk . \nExperiment s discovered a surprising insulating ferromagnetic phase in LMO thin \nfilms epitaxial ly grown on the square -lattice SrTiO 3 (STO) substrate [2, 7-13]. The \norigin of this puzzling insulating FM phase is unclear. The unexpected \nferromagnetism was considered to be extrinsic and deficiency of La [10, 14] was \nsuggested as a possible explanation to it . However, it is not consistent with the fact \nthat FM phase in the LMO thin films with cation deficien cy tend s to be metallic [7, 15, \n16]. Previous models [17, 18] and recent first-principles studies [19, 20] on LMO thin \nfilms predicted the metallic FM phase, instead of the experimentally observed \ninsulating FM phase. It is important to address how the prototypical antiferromagnetic \nMott insulator transforms to an insulating ferromagnet . \nTo answer such an intriguing question, we carry out a comprehensive theoretical \nstudy on the LMO thin film by combining genetic algorithm (GA) optimizations , \nfirst-principles calculations and the orbital -degenerate double -exchange (DE) model \nstudies. Our e xtensive GA simulations give an insulating FM phase of the LMO thin \nfilm, which crystallize s in the monoclinic P21/n structure. This monoclinic P21/n \nstructure has a peculiar arrangement pattern of Jahn -Teller (JT) distortions of MnO 6 \noctahedra , giving rise to the previously unreported three -dimensional ly \n2 2 2 23alternatedz r x ydd\n orbital order. According to the Goodenough -Kanamori \nrules [21], this kind of orbital orders induces three -dimensional ferromagnetism. \nFirst-principles calculations show that the band gap of the monoclinic P21/n FM \nphase is 0.16 eV , compatible with the experimentally observed excitation energy [2] \n(\n0.14 eV ). By virtue of the orbital -degenerate DE model, it is revealed that the band gap opens as a result of both the orbital ordering and the strong electron -phonon \ncoupling. We also show that LMO thin film transform s from the insulating A -type \nantiferromagnetic (A-AFM) phase to the insulating ferromagnetic phase , and then to \nthe metallic ferromagnetic phase when the lateral lattice constant decreases. \nIn this work, the widely adopted global optimization technique GA [22-25] is \nused to search the ground state structure of the LMO thin film. The GA method we \nemployed here is similar to that described in Ref. [25] with the exception that the spin \ninformation of the magnetic ions is explicitly kept in the crossover operat ion and a \nmutation operation related to the spin direction is also added . A \n2 2 2 \nperovskite supercell containing four formula units (20 atoms ) is used. In GA searches, \nin order to match the experimentally epitaxial strain induced by the square -lattice \nsubstrate STO, the two in -plane lattice vectors are fixed at \n2LMO LMO STO a b a \nwhere \naSTO is the lattice constant (\n3.905Å ) of the cubic STO substrate. However, \nboth the length and the direction of the out -of-plane lattice vector \nLMOc and the \ninternal ionic coordinate s are fully optimized. First-principles calculations based on \ndensi ty functional theory are performed using the generalized gradient approximation \nwith the PW91 parameterization plus on -site Coulomb repulsion U method (GGA+U) \nas implemented in Vienna Ab Initio Simulation Package (V ASP) [26-30]. Because the \nexperimentally measured [31] \n3.5 eVU describe s well the bulk LMO (see Sec. I \nof the Supplemental Material) [32] and gives a band gap (0.16 eV) of the LMO thin \nfilm close to the experimentally observed excitation energy [2] (\n0.14 eV ), it is \napplied to Mn 3 d electrons in the present work . A further detailed study showed that \nour main results remain correct if a reasonable U value is used (see Sec. II of the \nSupplemental Material) [32]. \nOur calculations show that , Pbnm LMO thin film strained on STO is metallic \nFM, consistent with the recent work of Lee et al . [20]. By ex tensive GA search es, \nhowever, we find an insulating FM phase, which has a total energy lower than the metallic Pbnm FM phase by \n6.9 meV f.u. (Fig. 1). This insulating FM phase \ncrystallizes in a slightly distorted monoclinic P21/n structure with \n090.74 . Two \nkinds of Mn ions, denoted as Mn -A and Mn -B, are found in the monoclinic P21/n \nstructure (Fig. 2b) while all Mn ions are equivalent in the orthorhombic Pbnm \nstructure (Fig. 2a). Mn-A has a MnO 6 octahedron elongated along the c axis (Fig. 2c) \nwhile Mn-B has a MnO 6 octahedron stretched in the ab plane (Fig. 2d). What is most \nsignificant is that these two different kinds of Mn atoms are arranged in a \ncheckerboard G -type manner (Fig. 2b). This is rather different from the bulk LMO \nwhere the relevant \n3Q modes [1] of th e JT distortion s of all MnO 6 octahedra are \nwith their principal axes lying within the ab plane and these axes are alternatively \narranged in this plane . \nThe band structure of the majority spin (up) of the monoclinic P21/n FM phase \nis plotted in Fig. 3a. Obviously, the monoclinic P21/n FM phase is insulating. It has \nan indirect band gap of 0.16 eV , compatible with the experimentally observed \nexcitation energy [2] \n( 0.14 eV ) . Moreover , its bandwidth is narrower than that of \nthe metallic Pbnm FM phase (Fig. 3c). We expect that the insulating P21/n FM phase \ncan be confirmed by the a ngle resolved photoemission spectroscopy (ARPES) \nexperiment. \nWe now examine the orbital order in the insul ating P21/n FM phase. Since the \nMnO 6 octahedron of Mn -A is elongate d along the c axis, the \n223zrd orbital is lower \nin energy than the \n22xyd orbital and thus the single \nge electron occupies the \n223zrd \norbital, which can be evidenced by the PDOS (Fig. 4a). The notable split between the \npeak of the \n223zrd orbital and that of the \n22xyd orbital is consistent with the large \nJT distortion [1] \n22\n23 Å 0.27AQ Q Q . In contrast, the orbital occupation is just \nopposite for Mn -B. The \n22xyd orbital is lower in energy as a result of the in -plane \nstretch of MnO 6 octahedron and thus the single \nge electron mainly occupies the 22xyd orbital, which can also be verified by the PDOS (Fig. 4b) . Likewise, its \nrelative weak split between the peak of \n223zrd orbital and that of \n22xyd orbital is \ndue to the small JT distortion \n22\n23 Å0.14BQ Q Q . Fig. 3b shows the \nge charge \ndensity integrated from -1.5 eV to the Fermi level, which displays the orbital order of \nthe insulating P21/n FM phase . In this energy interval, the spectral density of Mn-A \nhas predominantly \n223zrd character while Mn-B has predominantly \n22xyd character . \nTherefore, we firstly report that the insulating P21/n FM phase has a type of \nthree -dimensional ly \n2 2 2 23alternatedz r x ydd orbital order, which is different from \nthe \n2 2 2 23 x z r ydd type [18] (Fig. 3d) and completely different from the \n2 2 2 233/x r y rdd\n type [33]. \nBased on the established orbital order, the magnetic interactions among the Mn \natoms in the insulating P21/n FM phase can be deduced according to the \nGoodenough -Kanamori rules [21]. As a result of the three -dimensional ly \n2 2 2 23alternatedz r x ydd\norbital order, the half-filled \nbond \n223zrd orbital from \nMn-A overlaps with the empty \nbond \n223zrd orbital from Mn -B along the cubic \n[001] axis through the middle \nO2zp orbital, giving rise to strong ferromagnetic \ninteractions. Besides , the half -filled \nbond \n2gt orbitals will give rise to an \nantiferromagnetic interaction between Mn -A and Mn -B. Since the overlap of the \nbonding\n electrons is greater than that of the \nbonding electrons, the \nferromagnetic interaction turns out to be stronger than the antiferromagnetic one. \nTherefore the superexchange interaction between Mn -A and Mn -B along the cubic \n[001] axis is ferromagnetic. This mechanism also applies to the magnetic interactions \nbetween Mn -A and Mn -B in the ab plane, where the mutually interacting orbitals are \nthe half -filled \nbonding \n22xyd orbital of Mn -B and the empty \nbonding \n22xyd\n orbital of Mn -A. To sum up, Mn -A and Mn -B interact ferromagnetically along [100], [010] and [001] axes (see Sec. III of the Supplemental Material) [32]. The \ndeduced magnetic interactions among Mn atoms are verified by density functional \ntheory (DFT) calculations by means of the four-states mapping method [34]. The \nconsidered nearest neighbor (NN) magnetic interaction paths \nccJ , \n1abJ and \n2abJ are \nshown in Fig. 2b. As expected, our calculations find \n9.73 meVccJ , \n13.91 meVabJ\n and \n25.94 meVabJ . All of them are ferromagnetic. The next \nnearest neighbor (NNN) magnetic interactions a re found to be much weaker than the \nNN ones. Therefore, the magnetic ground state is ferromagnetic. With the DFT \ncalculated magnetic exchange constants , our MC simulations (see Sec. IV of the \nSupplemental Material) [32] lead to a transition temperature \n446 KCT , higher than \nexperimentally measured ones [7, 11] ranging from 115K to 240K. A possible \nexplanation to this discrepancy is that the epitaxial strain is gradually relaxed from t he \nsubstrate to the film surface, thus t he FM phase locate s just near the strained \nfilm-substrate interface while the bulk -like A -AFM phase dominates at the film \nsurface, similar to the case of the s train-induced ferromagnetism in the \nantiferromagnetic LuM nO3 thin film [35]. \nIn order to unveil the mechanism of the insulating ferromagnetism, we employ \nthe orbital -degenerate DE model with one \nge electron per Mn3+ ion. An infinite \nHund coupling limit [1], i.e., \nHJ , is adopted in our study. With this useful \nsimplification, the DE model Hamiltonian reads \n \n 1 1 2 3\n2 2 2\n1 2 3H Ω ..\n12 (1)2.aa\nij i j AF i j i i i xi i zi\nij ij i\ni i i\nit d d H C J S S Q n Q Q\nQ Q Q\n \n \n\n \n\n \nA detailed description of this Hamiltonian is given in the Sec. V of S upplementary \nMaterials [32]. Note that the second term, the NN antiferromagnetic superexchange \ninteraction between Mn \n2gt spins, is left out of considerations in the present work \nsince both the P21/n and the Pbnm phases are FM. The DFT relaxed structures of the P21/n FM and the Pbnm FM phases are used. Because the breathing modes (\n1Q) are \nmuch smaller than the JT modes (\n2Q and \n3Q) in both phases, they are ignored in our \nstudy. Lastly \n00.52 eVt and \nλ1.4 (\nλ usually estimated [1] between 1.0 and \n1.6) are used [36, 37]. This set of parameters results in an energy difference between \nthe P21/n FM and the Pbnm FM phases close to that obtained from the DFT \ncalculations. By exactly diagonaliz ing the DE Hamiltonian, we show the band \nstructures of the P21/n FM and Pbnm FM phases in Fig. 3a and Fig. 3c, respectively. \nAs expected, the band structure of P21/n FM phase is characteristic of insulator and \nthat of Pbnm FM phase is characteristic of metal. The profiles of both band structures \nobtained from the model well reprod uce that obtained from the DFT calculations, \nindicating that the parameters involved in the model are appropriate ly selected. \nNow let ’s address why the P21/n FM phase is insulating while the Pbnm FM \nphase is metallic. For the P21/n FM phase, as the three -dimensionally \n2 2 2 23alternatedz r x ydd\n orbital order has almost no overlap along the c axis \n(\n0zz\nab batt ) between the occupied orbitals, its bandwidth \n21/WPn is determined \nmainly by the in -plane overlap (\n03\n4ab baxxt t t or \n03\n4ab bayyt t t ) between the \nlower -energy occupied orbital \n223zrd of Mn -A and the lower -energy occupied orbital \n22xyd\n of Mn -B. So its bandwidth \n21/WPn is proportional to \n03t . Likewise , the \nbandwidth \nWPbnm of Pbnm FM phase determined by the \n2 2 2 23typez r x ydd \norbital order is proportional to \n03t . Thus , P21/n FM phase has a narrower bandwidth \nthan Pbnm FM phase. On the other hand , the onsite energy splitting of the two \nMnge\n orbitals due to the electron -phonon coupling in an isolated MnO 6 \noctahedron is proportional to the strength of the JT distortions [19]: \n22\n23 (2). QQ \n \nThus , the onsite energy splitting in P21/n FM phase is much larger than that in Pbnm FM phase since the JT distortion s in the former are much severe r than that in the latter. \nTherefore, P21/n FM phase possesses a finite band gap as a result of the narrow \nbandwidth and the large on -site \nge level splitting (Fig. 4a and Fig. 4b). For the \nPbnm FM phase, however, its bandwidths of Mn \nge bands are so wide and the \nelectron -phonon coupling is so weak that no band gap opens (Fig. 4c) . Furthermore, \ndecompositions of the total energies of both phases (Table I) show that, the \nelectron -phonon coupling causes a much larger energy lowering in the P21/n FM \nphase than the Pbnm FM phase although P21/n FM phase has higher hopping energy \nand lattice elastic energy than the Pbnm FM phase. In conclusion, it is revealed that \nthe orbital order and the electron -phonon cooperatively make the P21/n FM phase \ninsulating while the Pbnm FM phase metallic and that the electron -phonon coupling \nplays a vit al role in stabilizing the insulating P21/n FM phase as the ground state of \nthe LMO thin film strained on STO. \nFinally, using the same strategy applied to the LMO thin film strained on STO, \nwe have systematically investigated the property dependence of LMO thin films on \nlattice constants , which can be experimentally tuned by selecting different \nsquare -lattice substrates . Here, only the P21/n and Pbnm space groups are considered. \nAs usual , several common magnetic orders in perovskite are considered , namely, FM, \nA-AFM , C-type AFM (C -AFM) and G-type AFM (G -AFM) spin ordering s. The \nresults are shown in Fig. 1. As the C -AFM and G -AFM spin orders do not appear as \nthe lowest energy phase in the considered lattice constant window, they are not shown \nin Fig. 1 and not further discussed. Our results show that LMO thin films with lattice \nconstant ranging from 3.88 Å to 4.03 Å are insulating FM with the monoclinic P21/n \nstructure and the three -dimensionally \n2 2 2 23alternatedz r x ydd orbital order. \nHowever, LMO thin films with lattice constant s smaller than 3.88 Å are metallic FM \nwith the Pbnm structure and \n2 2 2 23typez r x ydd orbital order . When the lattice \nconstant is larger than 4.04 Å, the ground state of the LMO thin films becomes the \nPbnm structure with an insulating A -AFM order and \n2 2 2 233/ typex r y rdd orbital order . The metallic Pbnm FM and the insulating Pbnm A-AFM phases are consistent \nwith the work of Lee et al. [20] on LMO thin film with large compressive and tensile \nepitaxial strain. The phase transitions from the insulating P21/n FM ( P21/n FM I) \nphase to the metallic Pbnm FM ( Pbnm FM M) phase and then to the insulating Pbnm \nA-AFM ( Pbnm A-AFM I) phase are intuitively illustrated by the obvious \ndiscontinuity of the lattice constant \nc , shown in the upper panel of Fig. 1. The phase \ndiagram shown in Fig. 1 is also confirmed by our GA optimizations . \nIn summary, the physical origin of the well -known and puzzling insulating \nferromagnetism experimentally observed in the LaMnO 3 thin film grown on the \nsquare -lattice SrTiO 3 substrate has been investigated . We find that the insulating \nferromagnetic phase is intrinsically fr om strain induced orbital ordering, instead of \nextrinsic reasons such as defects. It crystallizes in a monoclinic P21/n structure , which \nhas two different kinds of MnO 6 octahedra: one is elongated along the \nc axis and \nthe other one is stretched in the \nab plane. They are arranged in a checkerboard \nG-type manner , giving rising to a previously unreported three -dimensional ly \n2 2 2 23alternatedz r x ydd\n orbital order, which naturally leads to the ferromagnetism. \nDouble Exchange model reveals that the band gap opens due to both the orbital \nordering and the strong electron -phonon coupling. Finally, we find that epitaxially \nstrained LaMnO 3 thin film transform s from the insulating A -type antiferrom agnetic \nphase to the insulating ferromagnetic phase , and then to the metallic ferromagnetic \nphase when the lateral lattice constant decreases. If LMO thin film is epitaxially \ngrown on some specified piezoelectric material s, an electric -field-induced \nmetal -insulator transition and an electric field control of the magnetism can be \nrealized experimentally at \nFM M 21 FM I Pbnm P n and \n21 FM I A AFM IP n Pbnm \n phase boundaries, respectively. These electric -field \ninduced phase transition s may be exploited in tunneling magnetoresistance (TMR ) \nand tunneling electroresistance (TER ) related devices. \nWork was partially supported by NSFC, the Special Funds for Major State Basic \nResearch, Foundation for the Author of National Excellent Doctoral Dissertation of China, The Program for Professor of Special Appointment at Shanghai Institutions of \nHigher Learning, Research Program of Shanghai municipality and MOE. We thank \nfor H. R. Liu, X. Gu and J. H. Yang, X. F. Zhai and S. Dong for valuable discussion s. \n \n \n \n \n \n \nReferences \n[1] E. Dagotto, T. Hotta, and A. Moreo, Phys. Rep. 344, 1 (2001). \n[2] A. Bhattacharya S. J. May, S. G. E. te Velthuis, M. Warusawithana, X. Zhai, Bin \nJiang, J. -M. Zuo, M. R. Fitzsimmons, S. D. Bader, and J. N. Eckstein , Phys. Rev. Lett. \n100, 257203 (2008). \n[3] J. Garcia -Barriocanal , J.C. Cezar, F.Y . Bruno, P. Thakur, N.B. Brookes, C. Utfeld, \nA. Rivera -Calzada, S.R. Giblin, J.W. Taylor, J.A. Duffy, S.B. Dugdale, T. Nakamura, \nK. Kodama, C. Leon, S. Okamoto and J. Santamari a, Nat . Commun . 1, 82 (2010). \n[4] M. Gibert, P. Zubko, R. Scherwitzl, J. Iniguez, and J. -M. Triscone, Nat . Mater . 11, \n195 (2012). \n[5] K. Rogdakis J.W. Seo, Z. Viskadourakis, Y . Wang, L.F.N. Ah Qune, E. Choi, J.D. \nBurton, E.Y . Tsymbal, J. Lee and C. Panagop oulos, Nat . Commun . 3, 1064 (2012). \n[6] E. O. Wollan and W. C. Koehler, Phys . Rev. 100, 545 (1955). \n[7] A. Gupta, T. R. McGuire, P. R. Duncombe, M. Rupp, J. Z. Sun, W. J. Gallagher, \nand G. Xiao, Appl . Phys . Lett. 67, 3494 (1995). \n[8] W. Choi Z. Marton, S. Y . Jang, S. J. Moon, B. C. Jeon, J. H. Shin, S. S. A. Seo, T. \nW. Noh, K. Myung -Whun, H. N. Lee and Y . S. Lee, J. Phys. D: Appl. Phys. 42, \n165401 (2009). \n[9] C. Adamo, X. Ke, P. Schiffer, A. Soukiassian, M. Warusawithana, L. Maritato, \nand D. Schlom, Appl . Phys . Lett. 92, 112508 (2008). \n[10] Z. Marton, S. S. A Seo, T. Egami, and H. N. Lee, J. Cryst. Growth 312, 2923 (2010). \n[11] H.-S. Kim and H. M. Christen, J . Phys . Condens . Matter 22, 146007 (2010). \n[12] X. Zhai, C. S. Mohapatra, A. B. Shah, J. -M. Zuo, and J. N. Eckstein, J. Appl. \nPhys. 113, 173913 (2013). \n[13] W. S. Choi, D. W. Jeong, S. Y . Jang, Z. Marton, S. S. A. Seo, H. N. Lee, and Y . S. \nLee, J . Korean Phys . Soc. 58, 569 (2011). \n[14] C. Aruta et M. Angeloni, G. Balestrino, N. G. Boggio, P. G. Meda glia, A. \nTebano, B. Davidson, M. Baldini, D. Di Castro, P. Postorino, P. Dore, A. Sidorenko, G. \nAllodi, and R. De Renzi, J. Appl. Phys. 100, 023910 (2006). \n[15] T. McGuire, A. Gupta, P. Duncombe, M. Rupp, J. Sun, R. Laibowitz, W. \nGallagher, and G. Xiao, J . Appl. Phys. 79, 4549 (1996). \n[16] P. Orgiani, C. Aruta, R. Ciancio, A. Galdi, and L. Maritato, Appl. Phys. Lett. 95, \n013510 (2009). \n[17] K. H. Ahn and A. J. Millis, Phys . Rev. B 64, 115103 (2001). \n[18] B. R. K. Nanda and S. Satpathy, Phys . Rev. B 81, 174423 (2010). \n[19] B. R. K. Nanda and S. Satpathy, J. Magn. Magn. Mater. 322, 3653 (2010). \n[20] J. H. Lee, K. T. Delaney, E. Bousquet, N. A. Spaldin, and K. M. Rabe, arXiv \npreprint arXiv:1307.3347 (2013). \n[21] J. B. Goodenough, Magnetism and the chemi cal bond (Interscience Publishers \nNew York, 1963), Vol. 98. \n[22] D. M. Deaven and K. M. Ho, Phys. Rev. Lett. 75, 288 (1995). \n[23] H. J. Xiang, S. -H. Wei, and X. G. Gong, Phys . Rev. B 82, 035416 (2010). \n[24] H. J. Xiang, S. H. Wei, and X. G. Gong, J . Am. Chem . Soc. 132, 7355 (2010). \n[25] X. Lu, X. Gong, and H. Xiang, arXiv preprint arXiv:1308.0076 (2013). \n[26] J. P. Perdew, J. A. Chevary, S. H. V osko, K. A. Jackson, M. R. Pederson, D. J. \nSingh, and C. Fiolhais, Phys . Rev. B 46, 6671 (1992). \n[27] P. E. Blö chl, Phys . Rev. B 50, 17953 (1994). \n[28] A. I. Liechtenstein, V . I. Anisimov, and J. Zaanen, Phys . Rev. B 52, R5467 \n(1995). \n[29] G. Kresse and J. Furthmü ller, Phys . Rev. B 54, 11169 (1996). [30] G. Kresse and D. Joubert, Phys . Rev. B 59, 1758 (1999) . \n[31] J. H. Park, C. T. Chen, S. W. Cheong, W. Bao, G. Meigs, V . Chakarian, and Y . U. \nIdzerda, Phys. Rev. Lett. 76, 4215 (1996). \n[32] Supplementary Materials : A detailed sduty on LaMnO 3 bulk and the impact of U \non the LaMnO 3 thin film, o rigin of ferromagnetism in the P21/n FM phase, Monte \nCarlo simulation of transition temperature of LaMnO 3 thin film and a detailed \ndescription of the double exchange Hamiltonian. \n[33] S. Dong, R. Yu, S. Yunoki, G. Alvarez, J. M. Liu, and E. Dagotto, Physical \nRevie w B 78, 201102 (2008). \n[34] H. Xiang, C. Lee, H. -J. Koo, X. Gong, and M. -H. Whangbo, Dalton . Trans . 42, \n823 (2013). \n[35] J. S. White M. Bator, Y . Hu, H. Luetkens, J. Stahn, S. Capelli, S. Das, M. Dö beli, \nTh. Lippert, V . K. Malik, J. Martynczuk, A. Wokaun , M. Kenzelmann, Ch. \nNiedermayer, and C. W. Schneider, Phys. Rev. Lett. 111, 037201 (2013). \n[36] Z. Popovic and S. Satpathy, Phys. Rev. Lett. 84, 1603 (2000). \n[37] K. H. Ahn and A. J. Millis, Phys . Rev. B 61, 13545 (2000). \n[38] J. Rodriguez -Carvajal, M. Hennion, F. Moussa, A. H. Moudden, L. Pinsard, and \nA. Revcolevschi, Phys . Rev. B 57, R3189 (1998). \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n Figures \n \nFigure 1 (Color online) Calculated Phase diagram of LaMnO 3 thin films. With in the \nconsidered lattice constant window, LMO thin films transform from metallic (M) \nPbnm FM to insulating (I) P21/n FM, then to the insulating Pbnm A-AFM phases. \nLMO thin films with the nearest neighbor Mn -Mn di stance close (or equal ) to that of \nbulk LMO [38] (\n3.99a Å) do not take the bulk structure because in this case LMO \nthin film is slightly stretch ed along the a axis and seriously compressed along the b \naxis. The evolution of the lattice constant \nc of the ground state is given in the upper \npanel. The two red straight line s are guides to the eyes. The magnetic and orbital \norders of each ground state are depicted in the middle panel . Blue arrows represent \nspins. Bottom panel gives the total energies o f the P21/n A-AFM, P21/n FM, Pbnm \nA-AFM and Pbnm FM phases as a function of the square -lattice substrate lattice \nconstant \na. Black arrow indicates the lattice constant of the SrTiO3 substrate. Note \nthat P21/n FM phase merges with the Pbnm FM phase when the lattice constant \nbecomes smaller than 3.88 Å and that P21/n structure merges with the Pbnm structure \nwhen the lattice constant becomes larger than 4.03 Å. \n \n \n \nFigure 2 (Color online) Structure of LaMnO 3 thin films strained on SrTiO 3. (a) \nStructure of the metallic Pbnm FM phase. (b) Structure and magnetic interaction paths \nccJ\n, \n1abJ and \n2abJ of the insulating P21/n FM phases . MnO 6 octahedra of Mn -A (c) \nand Mn -B (d) of the insulating P21/n FM phase are shown . The ir local coordinate \nsystem s shown between (c) and (d) are chosen in such a way that the local z axis is \nnearly along the c axis and xy plane nearly in the ab plane. Crystallographic axes are \ngiven by a, b and c . All numbers give Mn -O bond lengths in units of Å. \n \nFigure 3 (Color online) Band structure and charge density of LaMnO 3 film strained on \nSrTiO 3. Band structures of the insulating P21/n FM and the metallic Pbnm FM phases \nare shown in (a) and (c), respectively. Black solid ( red dashed) lines in both (a) and (c) \nare from the DFT calculations (orbital -degenerate double -exchange model). The \nFermi level is set to zero. DFT calculated \nge charge density distributions of the \ninsulating P21/n FM and the metallic Pbnm FM phases in the energy window of 1.5 \neV widths just below the Fermi level are plotted in (b) and (d), respectively . \n \n \n \n \n \n \n \n \n \n \n \n \nFigure 4 (Color online) Partial density of states (PDOS) and onsite energy splitting of \nLaMnO 3 thin film strained on SrTiO 3. PDOS of Mn -A, Mn -B of the insulating P21/n \nFM phase and the Mn atom of the metallic Pbnm FM phase are shown in the upper \npanels of the (a), (b) and (c), respectively. The Fermi level is indicated by the black \nvertical lines and set to be zero. D iagrammatic sketch es of the onsite energy splitting \nof the two \nMnge orbitals shown in bottom panels of (a), (b) and ( c) correspond to \nMn-A, Mn -B of the insulati ng P21/n FM phase and the Mn atom of the metallic Pbnm \nFM phase, respectively. The parameter \n is the energy difference between the \n223zrd\n and the \n22xyd orbitals. The black (red) rectangular box represents the band \nderived from the \n223zrd (\n22xyd ) orbital and the green horizontal line s indicate the \nFermi level. \n \n \n \n \n \n \n \n \nTable I. Decompositions of the total energies (\ntotE ) of the insulating P21/n FM and \nthe metallic Pbnm FM phases (20 atoms, namely, four formula units) into the hopping \nenergy (\nhoppingE ), electron -phonon coupling (\nele phE ) and l attice elastic energy \n(\nlatticeE ). \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \nSupplementary Materials for \nUnveiling the Origin of the Insulating Ferromagnetism in LaMnO 3 \nThin Film \n \n Y. S. Hou, H. J. Xiang*, and X. G. Gong* \n \nKey Laboratory of Computational Physical Sciences (Ministry of Education), State \nKey Laboratory of Surface Physics, and Department of Physics, Fudan University, \nShanghai 200433, P. R. China \n \n1. A detailed study on the bulk LaMnO 3 \nWith the experimentally measured [1] \n3.5 eVU and \n0.9 eVJ applied \nto the Mn 3 d electrons, our full structural optimizations successfully reproduce the \nA-type AFM (A -AFM) magnetic ground state in the bulk L aMnO3 (LMO) with a \nlower total energy than the FM state by amount of 8.5 meV/formula unit (f.u.), \nalthough it is not a trivial task to achieve this within DFT framework [2]. The \ncalculated direct gap is 1.2 eV , consistent with the optical measurements [3]. \nBesides, the relevant magnetic exchange constants are calculated to be \n2.09 meVabJ\n and \n1.13 meVcJ , respectively . They are consistent with \nexperimentally measured [4] \n1.85 meVabJ and \n1.1 meVcJ . With these \ncalculated magnetic exchange constants , Monte Carlo (MC) simulations (Fig. S1) \nreveal that the transition temperature \n116NTK , well consistent with the \nexperimentally observed one [5] \n140 KNT . Intriguing ly, our systematical \nstudies (Fig. S2) clearly show that either too small or too large U cannot correctly \nobtain the A -AFM magnetic ground state . We find that U should be between 2.5 \neV and 4. 0 eV to reproduce the A -AFM ground state. \nFig. S1 The dependence of the energy difference between the A -AFM and FM \n(upper panel), the Jahn -Teller distortions of the A -AFM state (middle panel) and \nthe FM state (bottom panel) on the U. Jahn -Teller Q is defined as \n22\n23 Q Q Q . \n \n \nFig. S2 Specific heat capacity of bulk LaMnO 3 as a function of temperature from \nMC simulation s with different lattice sizes (up to 39200 sites). A sharp \n peak \nof the specific heat capacity locates at \n116 KNT . \n \n2. A detailed study on the impact of U upon the epitaxial ly strained LaMnO 3 film \nwith SrTiO 3 lattice constant \n Detailed study show that the monoclinic P21/n state is always more stable \nthan the Pbnm state when a reasonable U between 2.5 and 4. 0 eV that describes \nwell bulk LaMnO 3 is adopted (Fig. S3) . It is found that for U less than 2.0 eV , the \nmonoclinic P21/n state cannot be stabilized because the small U significantly \nunderestimates the JT distortions. In other words, only the Pbnm structure is \nobtained. For the U larger than 3. 0 eV , the P21/n FM phase becomes insulating \nand has a n appreciably lower total energy than the metallic Pbnm FM phase . \nMoreover, both the band gap of the insulating P21/n FM phase and the total \nenergy difference between the insulating P21/n FM and the metallic Pbnm FM \nphases increase with U. It is worthy of noting that any U ranging from 1.0 eV to \n6.0 eV can not open a band gap in the Pbnm FM phase. \n \nFig. S3 Dependence of the relative stability between the P21/n FM and the \nmetallic Pbnm FM phases , the band gap and the Jahn -Teller distortions of the \nP21/n FM phase on the U. \n \n3. Origin of ferromagnetism in the insulating P21/n phase \n In order to investigate the magnetic interactions in the insulating P21/n phase, \nferromagnetic (FM) and antiferromagnetic (AFM) spin co nfigurations between \nMn-A and Mn -B (Fig. S4) are considered. If Mn -A and Mn -B ferromagnetically \ninteract along the [001] axis, the occupied \n223zrd orbital of Mn -A can strongly \nhybridize (\ntype ) with the unoccupied \n223zrd orbital of Mn -B, and thereby \nthe total energy is lowe red significantly (upper panel of Fig. S4). However, if they \ninteract antiferromagnetically, the hybridization between the occupied \n223zrd \norbital of Mn -A and the unoccupied \n223zrd orbital of Mn -B are almost negligible \nbecause these two orbital has a large energy difference due to the Hubbard \nrepulsion U (upper panel of Fig. S4). Besides, the hybridizations (\ntype ) \nbetween \n2gt orbitals of Mn -A and Mn -B are also negligible. Thus, there is only a \ntiny energy gain for the antiferromagnetically interacting case. Therefore, we find \nthat the FM interaction between Mn -A and Mn -B along the [001] axis is preferred \nto the AFM interaction. A similar mechanism applies to the magnetic in-plane \ninteractions between Mn -A and Mn -B except that the interacting orbitals are the \nunoccupied \n22xyd orbital of Mn -A and the occupied \n22xyd orbital of Mn -B \n(bottom panel of Fig. S4). To sum up, Mn -A and Mn -B interact ferromagnetically \nalong [100], [010] and [001] axes . \nFig. S4 The orbital interaction s between Mn -A and Mn-B with FM and AFM spin \nconfigurations , respectively . Upper panel is for the Mn pair along the [001] axis \nwhile the bottom panel is for the Mn pair along the [100] and [010] axes. \n \n \n \n \n \n \n \n \n \n \n \n \n \n4. Monte Carlo simulation of the transition temperature \nCT of the epitaxially \nstrained LaMnO 3 film with SrTiO 3 lattice constant \n \n Fig. S5 Specific heat capacity of LaMnO 3 thin film grown on STO as a function of \ntemperature from MC simulation with different lattice sizes (up to 39200 sites). A \nsharp \n peak of the specific heat capacity locates at \n446 KCT . \n \n \n \n5. A detailed description of the orbital -degenerate double -exchange model \nHamiltonian \nThe model Hamiltonian in our present work reads \n \n 1 2 3\n2 2 2\n1 2 3..\n12 (1),2aa\nij i j AF i j i i i xi i zi\nij ij i\ni i i\niH t c c H c J S S Q n Q Q\nQ Q Q\n \n \n \n\n \n where \nid \n()id is the creation (annihilation) operator for the \nge electron on \nthe orbital \n22α|x y a and \n22β|3z r b , with its spin parallel to the \nlocalized \n2gt spin \niS\n ; \n,, ax y z\n is the direction of the link connecting the two \nnearest neighbor (NN) Mn3+ sites; Berry phase \nΩ2 2 2 2iji jj ii\nijcos cos sin sin e \n arises due to the infinite Hund coupling, \nwhere \nθ and \n are the polar and azimuthal angles of the \n2gt spins, \nrespectively. In the model Hamiltonian, the first term is the standard DE \ninteraction . The hopping parameters are \n033 3 34x x x x\naa ab ba bbtt t t t , \n033 3 34y y y y\naa ab ba bbtt t t t \n, \n0z z z\naa ab bat t t and \n0z\nbbtt . The second term is \nthe NN \n2gt spins interaction through the antiferromagnetic superexchange \n0AFJ\n. The third term is the electron -phonon coupling, where \nλ is a \ndimensionless constant and the \norbitalge operators are \ni ia ia ib ibn d d d d , \nxi ia ib ib iad d d d\n, and \nzi ia ia ib ibd d d d . The last term is the lattice elastic \nenergy. JT modes (\n2Q and \n3Q ) and breathing mode (\n1Q ) are defin ed in Ref. [6]. \n \n \n \n \n \n \n \n[1] J. H. Park, C. T. Chen, S. W. Cheong, W. Bao, G. Meigs, V . Chakarian, and Y . U. \nIdzerda, Phys. Rev. Lett. 76, 4215 (1996). \n[2] T. Hashimoto, S. Ishibashi, and K. Terakura, Phys . Rev. B 82, 045124 (2010). \n[3] T. Arima, Y . Tokura, and J. B. Torra nce, Phys . Rev. B 48, 17006 (1993). \n[4] R. J. McQueeney, J. Q. Yan, S. Chang, and J. Ma, Phys . Rev. B 78, 184417 \n(2008). \n[5] E. O. Wollan and W. C. Koehler, Phys . Rev. 100, 545 (1955). \n[6] E. Dagotto, T. Hotta, and A. Moreo, Phys. Rep. 344, 1 (2001). \n \n \n \n " }, { "title": "1012.0195v1.The_second_order_dense_ferromagnetic_ferromagnetic_phase_transition.pdf", "content": "arXiv:1012.0195v1 [cond-mat.stat-mech] 1 Dec 2010The second order dense ferromagnetic-ferromagnetic phase\ntransition\nAycan¨Ozkan, B¨ ulent Kutlu\nGazi¨Universitesi, Fen -Edebiyat Fak¨ ultesi, Fizik B¨ ol¨ um¨ u , 06500\nTeknikokullar, Ankara, Turkey,\ne-mail: aycan@gazi.edu.tr, bkutlu@gazi.edu.tr\nThe fcc spin-1 Ising (BEG) model has a dense ferromagnetic ( df) ground\nstate instead of the ferromagnetic ground state at low temperat ure region and\nexhibits the dense ferromagnetic ( df) - ferromagnetic ( F) phase transition for\nd=D/J= 2.9,k=K/J=−0.5,ℓ=L/J= 0 and h=H/J= 0. The critical\nbehavior of the dense ferromagnetic ( df) - ferromagnetic ( F) phase transition\nhas been investigated using the cellular automaton cooling and heatin g algo-\nrithms. The universality class and the type of the dense ferromagn etic (df) -\nferromagnetic ( F) phase transition have been researched within the framework\nof the finite - size scaling, the power law relations and the probability d istri-\nbution. The results show that the dense ferromagnetic- ferroma gnetic phase\ntransition is of the second order and the model shows universal se cond order\nIsing critical behavior at d= 2.9 parameter value through k=−0.5 line.\nPACS number(s): 05.10.-a, 05.50.+q, 64.60.-i\n1I . INTRODUCTION\nIn recent years, some of the studies indicated that the spin-1 Isin g model\nhas a ground state ordered structure which is named the dense fe rromagnetic\n(df) [1−4]. In absence of the dfordered structure, the phase diagrams were\nformed for some phase transitions which has been assumed the wea k first order\ntransition instead of the second order transition. The presence o f thedfordered\nstructure can be caused to some changes on the global phase diag rams [4]. This\ncase clarifies the differences among the results of the previous stu dies for the\n(kTC/J,d) phase diagram through the kline. [5−11]. For example, the phase\ndiagram has exhibited a tricritical point ( TCP) instead of a critical end point\n(CEP) fork=−0.5. While the MFA [7] and RG [8] studies exhibited a critical\nend point ( CEP) for k=-0.5 on the BEG model global phase diagram, CA [5 ,6],\nMCRG [9] ,TPCA [10] and CVM [11] studies show that there is the tricritical\npoint (TCP). Through the kline, the df-Fphase transition is very important\nfor determining the type of the phase boundary. The purpose of t his study is\nto define the dfordered phase and is to investigate the nature of df−Fphase\ntransition at d= 2.9 value through k=−0.5 line. This point creates the type\nof the special point of the ( kTC/J,d) phase space for k=−0.5. Therefore,\nwe have found the df-Fphase transition worthy of investigation in depth.\n2Furthermore, the universality class of the df-Fphase transition has not been\ninvestigated so far. The critical temperature and the statical cr itical exponents\nare estimated by analyzing the data within the framework of the finit e - size\nscaling theory and the power law relations.\nThe spin-1 Ising model, which is known as the generalized Blume-Emery -\nGriffiths (BEG) model, can be used to simulate many physical systems . The\nmodel firstly has been presented for describing phase separation and superfluid\norderingin He mixtures [12]. The versionsofthe model havebeen ap plied to the\nphysical systems such as the solid-liquid-gas systems [13], the mult icomponent\nfluids [14], the microemulsions [15], the semiconductor alloys [16 −18], He3-He4\nmixtures [12 ,19] and the binary alloys [20].\nThe BEG model Hamiltonian is defined as\nHI=−J/summationdisplay\nSiSj−K/summationdisplay\nS2\niS2\nj+L/summationdisplay\n(S2\niSj+SiS2\nj)+D/summationdisplay\niS2\ni+h/summationdisplay\niSi\n(1)\nwhich is equivalent to the lattice gas Hamiltonian under some transfor mations\n[21−23].∝angbracketleftij∝angbracketrightdenotes summation over all nearest-neighbor (nn) pairs of sites\nandSi=−1,0,1. The parameters J,K,L,Dandhare bilinear, biquadratic,\ndipole-quadrupole interaction terms, the single-ion anisotropy con stant and the\n3field term. The BEG model for k≥0 has been studied by mean field ap-\nproximation (MFA) [12 −14], the transfer matrix method [24], series expan-\nsion method [25], the constant coupling approximation [26], the pos ition-space\nrenormalization method [27], cluster variation method (CVM) [1], linea r-chain\napproximation[2], Monte Carlomethod (MC) [3] and Cellular Automaton (CA)\n[4−6,28,29].\nIn this paper, the fcc BEG model for d=D/J= 2.9,k=K/J=−0.5,\nℓ=L/J= 0 andh=H/J= 0 is simulated using cooling and heating algorithm\nimproved from Creutz Cellular Automaton. In the previous papers, the Creutz\ncellular automaton (CCA) algorithm and its improved versions have be en used\nsuccessfully to study the properties of the critical behaviors of t he Ising model\nHamiltonians [4 −6,28−50]. The CCA algorithm, which was first introduced\nby Creutz [51], is a microcanonical algorithm interpolating between t he conven-\ntional Monte Carlo and the molecular dynamics techniques. The Creu tz cellular\nautomaton (CCA) is faster than the conventional Monte Carlo met hod (MC).\nThe CCA does not need high quality random numbers and it is a new and a n\nalternative simulation method for physical systems. It has anothe r advantage\nallowing the specific heat to be computed from internal energy fluct uations.\nOur previous studies showed that the heating and the cooling algorit hms im-\n4proved from the Creutz Cellular Automaton algorithm are effective t o study\nthe phase space and the critical behavior of the Blume Emery Griffith s model\n[4−6,28−30,35,36].\nII . RESULTS AND DISCUSSION\nThe CA algorithm of spin-1 Ising model is a microcanonical algorithm. T he\ntotal energy H, which is conserved, is given by\nH=HI+HK (2)\nwhereHIis Ising energy which is given by equation1 and HKis kinetic energy.\nThe kinetic energy HKis an integer, equal to the change in the Ising spin\nenergy for any spin flip and its value lie in the interval (0, m).mis equal to\n24Jford=D/J= 2.9,k=K/J=−0.5,ℓ=L/J= 0 and h=H/J= 0 on\nfcc lattice. For a site to be updated, its spin is changed to one of the other two\nstates with 1 /2 probability. If this energy is transferable to or from the kinetic\nenergy variable of the site, such that the total energy H is conser ved, then this\nchange is done and kinetic energy is appropriately changed. Otherw ise the spin\nis not change [30 ,34−36].\nAt the heating and the cooling algorithms, the simulation consist of tw o\nparts, the initialization procedure and the computation of the ther modynamic\n5quantities. The initial configuration for heating and cooling algorithm s can be\nset in different shapes. In this study, the initial configurations are obtained at\nthree different shapes for heating algorithm during 20.000 CA steps . Firstly, all\nthe spins are up ( S= +1) at the absolute zero temperature for both algorithms.\nThe initial configuration of the heating algorithm has been obtained a t low\ntemperature ordered phase ( df) adding kinetic energy which is equal to the\nmaximum change in the Ising spin energy for the any spin flip to the spin\nsystem for set I and II. The another initial configuration (set III ) has been\nobtained flipping 8% of the spins to S= 0 state. The heating rate is realized by\nincreasing of 8% in the kinetic energy ( Hk) of 15% of the fcc lattice for two sets\nat the computation of the thermodynamic quantities. At set III, t he heating\nrate is realized by increasing of 8% in the kinetic energy ( Hk) of all site of the\nfcc lattice.\nOn the other hand, the initial configuration for the cooling algorithm is\nobtained adding energy to the 70% of the spin system for getting th e disordered\nphase (P) at high temperature. During the cooling cycle, the cooling rate is\nrealized by decreasing of 8% in the kinetic energy ( Hk) from 25% of the spin\nsystem. Theinitialconfigurationsarerunduringthe20.000CellularA utomaton\ntime steps. Instead of resetting the starting configuration at ea ch energy, it\n6was used the final configuration at a given energy as the starting p oint for\nthe next at both heating and cooling algorithms. The computed value s of the\nthermodynamic quantities (the order parameters ( M,Q), the susceptibility ( χ),\nthe Ising energy ( HI) and the specific heat ( C)) are averages over the lattice\nand over the number of time steps (2 .000.000 ) with discard of the first 100 .000\ntime steps during the cellular automaton develops [4 −6,28,29].\nThey have been computed on the fcc lattice with L= 8, 9, 10, 11 and 12\n(The total number of sites is N= 4L3) for periodic boundary conditions. The\nfcc lattice was formed in a simple cubic (sc) lattice. (The total numbe r of sites\nisN= 4L3= 6912 for L= 12 fcc lattice, this total site number equals to\nL= 19 for the simple cubic lattice). The presented figures are set III (heating\nalgorithm) results.\nThe order parameters, the Ising energy, susceptibility and specifi c heat are\ncalculated from\nM=1\nN/summationdisplay\niSi, Q=1\nN/summationdisplay\niS2\ni (3)\nUI= (−J/summationdisplay\nSiSj−K/summationdisplay\nS2\niS2\nj+D/summationdisplay\niS2\ni)/U0 (4)\nχ=N/angbracketleftbig\nM2/angbracketrightbig\n−∝angbracketleftM∝angbracketright2\nkT(5)\n7CI/k=N/angbracketleftbig\nU2\nI/angbracketrightbig\n−∝angbracketleftUI∝angbracketright2\n(kT)2(6)\nwhereU0is the ground state energy at kT/J= 0.\nThe ferromagnetic ( F) and the paramagnetic ( P) phases can be determined\nwith the average occupation of the states ∝angbracketleftP±1,0∝angbracketright. As the projectors for the\nstatesS= +1,−1and0are P+1=1\n2S(S+1),P−1=1\n2S(S−1)andP0= 1−S2,\nthe average occupation of the states are ∝angbracketleftP+1∝angbracketright=1\n2(Q+M),∝angbracketleftP−1∝angbracketright=1\n2(Q−M)\nand∝angbracketleftP0∝angbracketright= 1−Q, respectively. With considering the average occupation of the\nstates, another ferromagnetic phase can be determined as the d ense ferromag-\nnetic phase ( df).\nFerromagnetic ( F):∝angbracketleftP+1∝angbracketright ∝negationslash=∝angbracketleftP−1∝angbracketright ∝negationslash=∝angbracketleftP0∝angbracketright ∝negationslash= 0, (M∝negationslash=Q∝negationslash= 0)\nDense ferromagnetic ( df):∝angbracketleftP−1∝angbracketright →0;∝angbracketleftP+1∝angbracketright ∝negationslash=∝angbracketleftP0∝angbracketright ∝negationslash= 0, (M∼=Q∝negationslash= 0)\nParamagnetic ( P):∝angbracketleftP−1∝angbracketright=∝angbracketleftP+1∝angbracketright ∝negationslash=∝angbracketleftP0∝angbracketright ∝negationslash= 0, (M= 0,Q∝negationslash= 0).\nII .1 Temperature Variations of Thermodynamic Quantities f or the\ndF−F−PPhase Transitions\nThetemperaturevariationoftheorderparameters( M,Q), thesusceptibility\n(χ), theIsingenergy( HI) andthespecificheat( CI)areillustratedin figure1for\nexhibiting the general aspect of the successive df−F−Pphase transitions at\nd= 2.9 parameter value through k=−0.5 line. As it is seen in figure 1(a) and\n8figure 1(c), the order parameters and the Ising energy appear c ontinuously for\ndf−FandF−Pphase transitions. Therefore both phase transitions are of the\nsecond order as functional behavior. The susceptibility ( χ) and the specific heat\n(CI,C)exhibit two peaks at TC1andTC2temperatures corresponding to df−F\nandF−Pphase transitions(Figure 1(b), 1(d) and figure 2). It can be seen from\nfigure 1(c), the functional change of the Ising energy from orde r to order ( df\n−F) phase transition is different from the view of the order to disorder (F−P)\nphasetransition. For df-Fphase transition, the Isingenergydifference (∆ U) is\ngreater than for the F-Pphase transition. The estimated critical temperature\nfrom susceptibility and specific heat maxima is compatible with each oth er for\nF-Pphase transition. But the critical temperature values for df-Fphase\ntransition are not compatible (Figure 1(b) and 2(a)). Therefore s pecific heat\n(C) has been recalculated for only spin-spin interaction energy (Figur e 2(b)).\nThe spin-spin interaction energy Uis determined as\nU= (−J/summationdisplay\nSiSj)/U0 (7)\nThe specific heat calculated from Ucan show a sharp peak for df−Fphase\ntransition. Because, the first sum ( U) in Ising energy ( UI) distinguish the S=\n+1 and−1 states. Indeed, the obtained infinite critical temperature ( TC1(∞)=\n91.52±0.04)from the susceptibility ( χ) and specific heat ( C) peak temperatures\nare compatible with each other. TC2(∞) is obtained from the susceptibility ( χ)\nand specific heat ( CIandC) peak temperatures as 3 .20±0.02.\nThe temperature variations of the ∝angbracketleftP+1∝angbracketright,∝angbracketleftP−1∝angbracketrightand∝angbracketleftP0∝angbracketrightare given in figure\n3. The initial configuration for the heating algorithm is created as all spins are\nup (S= +1) at absolute zero temperature ( T= 0). If the enough energy is\nadded to the spin system, the S= 0 begins to arise. The excitation energy of\nthe singlespin flippingfrom S= +1toS= 0is 3.1Jwhile it is 24 Jfor the single\nspin flipping from S= +1 to S=−1. Therefore, at low temperature region,\nthe rising probability of the S=−1 state has to be lower than S= 0 state. It\ncan be seen in figure 3 that the spin system includes S= +1 and S= 0 states\npredominantly. The value of ∝angbracketleftP+1∝angbracketrightis about 1 and ∝angbracketleftP0∝angbracketrightis different from zero,\nwhile∝angbracketleftP−1∝angbracketrightappears almost zero indicating the dfphase for T < T C1(L). At the\nsame time,. So, Mis almost equal to the Q(M∼=Q∝negationslash= 0). As ∝angbracketleftP−1∝angbracketrightincreases\naboveTC1(L), thedfordered phase changes to the Fordered phase. At high\ntemperature region, the ferromagnetic - the paramagnetic phas e transition ( F\n-P) occurs . ∝angbracketleftP+1∝angbracketrightis equal to ∝angbracketleftP−1∝angbracketrightand the system is in the paramagnetic\n(P) disordered phase above TC2(L) temperature ( M∝negationslash=Q∝negationslash= 0). Therefore the\nphase space is divided into three regions ( df,FandP). It is obvious that to\n10follow the temperature variation of ∝angbracketleftP−1∝angbracketrightis a useful way to prove the existence\nof thedfordered phase.\nII .2 Probability Distribution of Order Parameter\nThe another useful procedure to distinguish the phase transition type is to\ncalculate the probability distributions of the order parameter ( P(M)). In our\nstudy the probability distribution is calculated by\nPL(M) =NM\nNCCAS(8)\nwhereNMis the number of times that magnetization Mappears, and NCCASis\nthe total number of the cellular automaton steps. The histogram w ith 200 bins\nare used for plotting the probability distribution of the magnetizatio n [24, 31].\nThe probability distribution of the order parameter ( P(M)) near the phase\ntransition temperature shows two peaks in the second phase tran sitions.\nThe probability distributions of the order parameter ( P(M)) are shown for\ndifferent temperature values in figure 4. The peaks of the order pa rameter\nprobability distribution exhibits minimum with increasing temperature a t the\nlow temperature region. This minimum corresponds to the second or derdf−F\nphase transition at the Tχ\nC1(L= 12) = 1 .479. Although the phase transition is\nof the second order, the probability distribution shows the single pe ak near the\n11phase transition temperature TC1(L). Because the system has S= +1 and 0\nspins below TC1(L) and the transition is from order ( df) to order ( F). However,\ntheprobabilitydistributionsinthe F−Pphasetransitionregionexhibitthetwo\npeaks with the contribution of the S=−1 state near the TC2(L= 12) = 3 .187.\nForT > T C2, there is a single peak focused to M= 0 indicating the disordered\n(P) phase.\nII.3 Finite - Size Scaling Analyses and the Statical Critica l Expo-\nnents\nThe values of the statical critical exponents ( ν,β,γ,α) are estimated within\nthe frameworkof the finite - size scaling theory and the power laws. The infinite\nlatticecriticaltemperature TC(∞)hasbeenobtainedfromthesusceptibilityand\nthe specific heat peak temperatures for the successive second o rderdf−F−P\nphasetransitionsandfromtheintersectionpointofBindercumulan tcurves( UL)\nforthesecondorder F−Pphasetransition[52]. Thefinite-sizescalingrelations\nof the Binder cumulant ( UL), the order parameter ( M), the susceptibility ( χ)\nand the specific heat ( C) are given by\nUL=G◦(εL1/ν) (9)\n12M=L−β/νX◦(εL1/ν) (10)\nkTχ=Lγ/νY◦(εL1/ν) (11)\nC=La/νZ◦(εL1/ν) (12)\nFor large x=εL1/ν, the finite lattice critical behaviors must be asymptoti-\ncally reproduced, that is,\nX◦(x)∝Axβ(13)\nY◦(x)∝Bx−γ(14)\nZ◦(x)∝Cx−α(15)\nAccording to the finite size scaling theory, the data for the finite - s ize\nlattices of the thermodynamic quantities should lie on a single curve fo r the\ntemperatures both below and above TC(∞) with universal critical exponents.\nThe critical exponents β,γandαhave been obtained from the Log-Log plots\nof the asymptomatic functions.\n13The temperature variation of the Binder cumulant is shown for the d ifferent\nlattice sizes in figure 5. The Binder cumulant curves intersect at the TUL\nC2(∞) =\n3.20±0.02correspondingto the F−Pphasetransition(Figure 5(a)). This value\nis compatible with TC2(∞) which is extrapolated according to the finite size\nscaling theory from the susceptibility and the specific heat peak tem peratures\n(TC(L)) of the finite lattices, respectively.\nTC(L) =TC(∞)+aL−1/ν(16)\nIt can be seen in the inset of the figure 5(a) that there is no interse ction at the\nBinder cumulant for the data of the df−Fphase transition region. However,\nthe Binder cumulant curves exhibit a plateau near the infinite lattice c ritical\ntemperature ( TC1(∞)) which is obtained from the susceptibility ( χ) and the\nspecific heat ( C) peak temperatures as Tχ\nC1(∞) = 1.52±0.04. In figure 5(b),\nthe scaling data of the Binder cumulants are shown for the second o rder phase\ntransition from FtoP. Near the TUL\nC2(∞), Binder cumulant curves have been\nscaled well for TC=TC2(∞) withν= 0.64. It can be seen in the inset of the\nfigure 5(b) that the data corresponding to dfordered phase could not be scaled\nwith the TUL\nC2(∞) critical temperature value. However, the scaling data of the\nbinder cumulant corresponding to the dfordered phase lie on a single curve for\n14ε=(T−TC1(∞))/TC1(∞) atT < TC1(∞) region using ν= 0.64 (Figure 5(c))\nand the finite size scaling relations validate for the dfordered region.\nInfigure6(a),thescalingdataoftheorderparameterisillustrate datthesuc-\ncessivedf−F−Pphase transitions for L= 8,9,10,11and 12 at TC=TC2(∞).\nThe order parameter data lie on the two different curves with slope= β/ν= 0.31\nandβ′/ν=−0.55 for the temperatures both below and above TC2(∞) respec-\ntively except for dfordered region with β= 0.31 andν= 0.64. As it is seen\nin the inset of the figure 6(a), the data of df−Fphase transition region have\nnot been scaled with TC=TC2(∞). However, the data corresponding to the\ndf−Fphase transition region have been scaled well with TC=TC1(∞) for\nT < TC1(∞) usingβ= 0.31 andν= 0.64 in figure 6(b).\nThe scaling data of the susceptibility have been shown in figure 7 with t he\nstraight lines describing the theoretically predicted behavior for lar gex(Equa-\ntion 8). The susceptibility data for the temperatures both below an d above\nTC2(∞) agrees with the asymptotic form except for the df-Fphase transition\nregion and so with the TC=TC2(∞),γ=γ′= 1.25 andν= 0.64 in figure 7(a)\nand 7(c). However, the data of dfordered phase ( T < T C1) have been scaled\nwith the TC=TC1(∞) usingγ= 1.25 andν= 0.64 in figure 7(b).\nThe finite size scaling data of the singular portion of the specific heat (C)\n15have been exhibited in figure 8. The data of F−Pphase transition region\nofCIhave been scaled well both below and above TC\nC2(∞) usingα= 0.12,\nν= 0.64 and the correction terms, b−=−70 andb+=−8 (Figure 8(a) and\n8(b)). Althoughthe dataofthe dfphaseregioncould notbeen scaledwith TC=\nTC\nC2(∞) in figure 8(a), the data lie on single curves with slope= −α/ν=−0.12\nat the both side of the TC\nC1(∞), using α= 0.12,ν= 0.64 and the correction\nterms,b−=−1 andb+=−0.3 in figure 8(c) and (d). On the other hand,\nthe specific heat ( CI) data calculated from UIscales well at df-F-Pphase\ntransitions using α= 0.12,ν= 0.64 forT < Tχ\nC1(∞)andT < TCI\nC2(∞) and\nT > TCI\nC2(∞).\nTheM,χandCdata have been analyzed within the framework of the finite\nsizescalingtheoryfor the successive df−F−Pphasetransitions. The estimated\nvalues of the statical critical exponents arein good agreementwit h the universal\nvalues (α= 0.12,β= 0.31,γ= 1.25,ν= 0.64) for the df−Fand theF−P\nphase transitions.\nII.4 Power Law Relations and the Infinite Lattice Statical Cr itical\nExponents\nOn the other hand, the critical exponent values for df−Fphase transition\ncan be obtained using the following power law relations [53].\n16M(L) =εβ(L)(17)\nχ(L) =ε−γ(L)(18)\nC(L) =ε−α(L)(19)\nwhereε=(T−TC(L))/TC(L). The finite lattice critical exponents β(L),β′(L),\nγ(L),γ′(L),α(L) andα′(L) of the order parameter ( M), susceptibility ( χ) and\nthe specific heat ( C) quantities are obtained from the slope of the log-log plot\nof the power laws relations for each finite lattices in the interval 0 .05≤ε≤0.2.\nTheinfinitelatticecriticalexponentsareobtainedusinglinearextra polationand\ntheir values are given in Table I. The estimated values for cooling algor ithm and\nthree simulation sets of heating algorithm are in good agreement with the finite\nsize scaling critical exponent estimations and the universal values f or 3dIsing\nmodel (β= 0.31,γ= 1.25,α=α′= 0.12 andν= 0.64).\nTable 1. The estimated values of the infinite lattice critica l exponents and the\ncritical temperatures ( α,β,γandTχ,C\nC1(∞)) using linear extrapolation.\n17df-Fphase transition\nHeating Cooling\nSet I Set II Set III Average of sets\nTχ\nC1(∞)1.50±0.021.50±0.031.52±0.031.51±0.031.52±0.04\nTC\nC1(∞) − − 1.52±0.01 − 1.52±0.01\nβ(TTC1)0.12±0.010.12±0.010.12±0.010.12±0.010.12±0.01\nIII. SUMMARY\nThe (kTC/J,d) phase diagrams is illustrated in figure 9 for the presence of\ndforder. The type of special point is determined by the d= 2.9 parameter. The\ncalculations show that model exhibits the phase transition from ord er to order\nford= 2.9 and the first order phase transition from order to disorder in the\n3≤d <4 parameter region. If the model doesn’t exhibit the dfordered phase\ninstead of the Fordered phase in the low temperature region, the continuous\nphase transition from Forder to Forder (F−F) is considered as the weak\nfirst order. This constitutes the part of the first order phase tr ansition line\nwhich creates the critical end point ( CEP) [7,8]. However, CA results show\nthat, the model has a dfordered phase for the parameters in the 2 .9≤d <4.0\nregion. The spin system contains S= +1 and S= 0 states. As a result of this,\nthe order parameters MandQare almost equal each other ( M∼=Q∝negationslash= 0) at\nlow temperatures. With increasing temperature, the dense ferro magnetic ( df)\nordered phase changes continuously to ferromagnetic ( F) ordered phase at the\n18d= 2.9,k=−0.5withtheenoughcontributionof S=−1state. Therefore,near\nthed= 2.9, the first order phase transition line have been changed to secon d\norder phase transition line, and there occurs the tricritical point TCP(Figure\n9). In order to determine the universality class of the successive df−F−P\nsecond order phase transitions, the static critical exponents ( α,β,γandν) are\nestimated within the frameworkofthe finite - size scaling theory. Th e estimated\nvalues of the critical exponents ( α= 0.12,β= 0.31,γ= 1.25 andν= 0.64)\nnear the TC1andTC2temperatures are in good agreement with the theoretical\nvalues for three sets. The df−Fphase transition is analyzed with the power\nlaws for comparing with the critical exponent values estimated from the finite\n- size scaling theory. The obtained values are in compatible with the fin ite -\nsize scaling analyze results and the universal values for the 3 dIsing model. The\nobtained results have shown that the df−Fphase transition is of the second\norder and it is compatible with the universal Ising critical behavior fo rd= 2.9\nparameter value through k=−0.5 line. As a result of this, the definition of the\ndfphase changes the phase transition type and the special point typ e in the\nphase space for the BEG model. This result will lead to reexamine the s tructure\nof phase spaces.\nACKNOWLEDGEMENT\n19This work is supported by a grant from Gazi University (BAP:05/200 3-07).\nReferences\n[1] Keskin M, Ekiz C, Yal¸ cın O, 1999 Physica A 267392\n[2] Albayrak E, Keskin M, 2000 J. Magn. Magn. Mater. 203201\n[3] Ekiz C, Keskin M, 2002 Phys. Rev. B 66054105\n[4]¨Ozkan A, Kutlu B, 2010 Int. J. of Mod. Phys. B accepted to publish\n[5] Sefero˘ glu N, Kutlu B, 2007 Physica A 374165\n[6]¨Ozkan A, Kutlu B, 2007 Int. J. of Mod. Phys. C 181417\n[7] Hoston W, Berker A N, 1991 Phys. Rev. Lett .671027\n[8] Netz R R, Berker A N, 1993 Phys. Rev. B 4715019\n[9] Netz R R, 1992 Europhys. Lett. 17373\n[10] Baran O R, Levitskii R R, 2002 Phys. Rev. B 65172407\n[11] Lapinskas S, Rosengren A, 1993 Phys. Rev. B 4915190\n[12] Blume M, Emery V J and Griffiths R B, 1971 Phys. Rev. A 41071\n[13] Lajzerowicz J and Siverdi˙ ere J, 1975 Phys. Rev. A 112090\n[14] Lajzerowicz J and Siverdi˙ ere J, 1975 Phys. Rev . A112101\n[15] Schick M and Shih W H, 1986 Phys. Rev. B 341797\n[16] Newman K E and Dow J D, 1983 Phys. Rev. B 277495\n[17] Gu B L, Newman K E, Fedders P A, 1987 Phys. Rev. B 359135\n20[18] Gu B L, Ni J, Zhu J L, 1992 Phys. Rev. B 454071\n[19]LawrieI D, SarbachS, Phase transitions and Critical Phenomena, edited\nby C. Domb and J. L. Lebowitz 1984 Vol 9Academic Press, New York\n[20] Kessler M, Dieterich W and Majhofer A, 2003 Phys. Rev. B 67134201\n[21] Ausloos M, Clippe P, Kowalski J M, Pekalski A, 1980 Phys. Rev. A\n222218,ibid. 1980IEEE Trans. Magnetica MAG 16233\n[22] Ausloos M, Clippe P, Kowalski J M, Pekalska J, Pekalski A, 1983 Phys.\nRev. A283080; Droz M, Ausloos M, Gunton J D, ibid.197818388\n[23] Ausloos M, Clippe P, Kowalski J M, Ekalska J P, Pekalski A, 1983 J.\nMagnet. and Magnet. Matter 3921\n[24] Koza Z, Jasuukiewicz C, Pekalski A, 1990 Physica A 164191\n[25] Saul D M, Wortis M and Stauffer D, 1974 Phys. Rev. B94964\n[26] Takahashi K, Tanaka M, 1979 J. Phys. Soc. Japan 461428\n[27] Berker A N, Wortis M, 1976 Phys. Rev. B 144946\n[28]¨Ozkan A, Kutlu B, 2009 Int. J. of Mod. Phys. C 201617\n[29]¨OzkanA,KutluB,2010 Cent. Europ. J.of Phys. , DOI:10.2478/S11534-\n010-0081-1\n[30] Demirel H, ¨Ozkan A, Kutlu B, 2008 Chineese Phys. Lett. 252599\n[31] Kutlu B, 2001 Int. J. Mod. Phys. C 121401\n21[32] Kutlu B, 2003 Int. J. Mod. Phys. C 141305\n[33] Solak A and Kutlu B, 2004 Int. J. Mod. Phys. C 151425\n[34] Kutlu B, ¨Ozkan A, Sefero˘ glu N, Solak A and Binal B, 2005 Int. J. Mod.\nPhys. C 16933\n[35]¨Ozkan A, Sefero˘ glu N and Kutlu B, 2006 Physica A 362327\n[36] Sefero˘ glu N, ¨Ozkan A and Kutlu B, 2006 Chineese Phys. Lett. 232526\n[37] Kutlu B, Aktekin N, 1994 J. Stat. Phys. 75757\n[38] Kutlu B, Aktekin N, 1995 Physica A 215370\n[39] Kutlu B, 1997 Physica A 234807\n[40] Kutlu B, 1997 Physica A 243199\n[41] Aktekin N, 2000 Annual Reviews of Computational Physics VII, ed.\nD.Stauffer, pp.1.World Scientific, Singapore\n[42] Aktekin N, 2001 J. stat. Phys. 1041397\n[43] Aktekin N, Erko¸ c S, 2001 Physica A 290123\n[44] Aktekin N, Erko¸ c S, 2000 Physica A 284206\n[45] Merdan Z, Erdem R, 2004 Phys. Lett. A 330403\n[46] Merdan Z, Bayırlı M, 2005 Applied Mathematics and Computation 167\n212\n[47] Merdan Z, Atille D, 2007 Physica A 376327\n22[48] Dress C, 1995 J. of physics A 287051\n[49] Saito K, Takesue S and Miyashita S, 1999 Phys. Rev. E 592783\n[50] Kutlu B, C ¸ivi M, 2006 Chineese Phys. Lett. 232670\n[51] Creutz M, 1986 Ann. Phys. 16762\n[52] Binder K, 1981 Z.Phys. B 43119\n[53] Huang K, 1987 Statistical Mechanics John Wiley & Sons 396\nFigure Captions\nFigure 1. For ( d= 2.9,k=−0.5), the temperature dependence of (a) the\norder parameters ( M,Q), (b) the susceptibility ( χ), (c) the Ising energy ( HI)\nand (d) the specific heat ( CI/k).\nFigure 2. For ( d= 2.9,k=−0.5), the specific heat ( C/k) calculated from\nU.\nFigure 3. For ( d= 2.9,k=−0.5), the temperature dependences of the\n∝angbracketleftP∝angbracketright.∝angbracketleftP+1∝angbracketright,∝angbracketleftP−1∝angbracketrightand∝angbracketleftP0∝angbracketrightcorrespond to the S= +1,−1 and 0 spin states,\nrespectively.\nFigure 4. The probability distribution of the Mfor (d=−0.5,k= 0.9) on\nL= 12.\nFigure 5. For ( d= 2.9,k=−0.5), (a) the temperature dependence of the\nBinder cumulant ( UL), (b) the finite - size scaling of the Binder cumulant near\n23thedf−F−Pphase transition with TUL\nC2(∞), (c) the finite - size scaling of\nthe Binder cumulant near the df−Fphase transition with TC1(∞).\nFigure 6. For ( d= 2.9,k=−0.5), the finite - size scaling plots of (a) the\norder parameter with TC2(∞), (b) the order parameter with TC1(∞) near the\ndf−Fphase transition for T < TC1(∞).\nFigure 7. For ( d= 2.9,k=−0.5), the finite - size scaling plots of the\nsusceptibility (a) with ε=(T−TC2(∞))/TC2(∞) forT < TC2(∞), (b) near the\ndf−Fphase transition with ε=(T−TC1(∞))/TC1(∞) forT < TC1(∞), (c)\nwithε=(T−TC2(∞))/TforT > TC2(∞).\nFigure 8. For ( d= 2.9,k=−0.5), the finite - size scaling plots of the specific\nheat (a) for T < TC2(∞) withTC2(∞), (b) for T > TC2(∞) withTC2(∞), (c)\nforT < TC\nC1(∞) withTC\nC1(∞), (d) for T > TC\nC1(∞) withTC\nC1(∞).\nFigure 9. The phase diagram for k=−0.5. The phase space with dfphase\ncontains a TCPatd= 2.9.\n240.0 0.2 0.4 0.6 0.8 1.0 \n0 1 2 3 4 5\nkT/J M, Q \nL=8 \nL=9 \nL=10 \nL=11 \nL=12 d=2.9 \nk=-0.5 (a) \nFigure 1(a) 0246810 12 \n0 1 2 3 4 5\nkT/J χ\nTC1 TC2 (b) \nFigure 1(b) 0510 15 \n0 1 2 3 4 5\nkT/J HI(c) \nFigure 1(c) 010 20 30 40 50 60 70 \n0 1 2 3 4 5\nkT/J CI/k TC1 \nTC2 (d) \nFigure 1(d) 0.0 0.2 0.4 0.6 0.8 1.0 1.2 \n0 1 2 3 4 5\nkT/J C/k L=8 \nL=9 \nL=10 \nL=11 \nL=12 TC1 \nTC2 \nFigure 2 Figure 3 0.0 0.2 0.4 0.6 0.8 1.0 \n0.5 2.5 4.5 6.5 \nkT/J

S=+1 \nS=0 \nS=-1 L=12 \nHeating \nAlgorithm \nTC2 TC1 F df PFigure 4 0.0 0.1 0.2 0.3 0.4 0.5 \n-0.5 0.0 0.5 1.0 \nMP(M) Tc1(L=12)=1.479 \nTc2(L=12)=3.187 \nTC1 (L) \nTC2 (L) -2.0 -1.5 -1.0 -0.5 0.0 \n1 2 3 4\nkT/J ULL=8 \nL=9 \nL=10 \nL=11 \nL=12 \n(a) \nTC2 (∞)=3.20±0.02 -2.00 -1.99 -1.98 \n0.5 1.0 1.5 2.0 Tχ\nC1 (∞)=1.52 \nFigure 5(a) 4(a) \n)=3.20±0.02 \nFigure 5(a) -2.0 -1.5 -1.0 -0.5 0.0 \n-20 -10 0 10 20 \nεL1/ν ULν=0.64 \nTC=T C2 (∞)=3.2 (b) \n-2.00 -1.98 \n-40 -30 -20 -10 dF region \nFigure 5(b) -2.00 -1.95 \n-20 -10 0 10 20 \nεL1/ν ULν=0.64 \nTT C2 df region (a) \nTC=T C2 (∞)ν=0.64 \nβ=0.31 \nFigure 6(a) slope= β/ν = 0.31 \n-0.5 0.0 0.5 1.0 \n-0.5 0.0 0.5 1.0 1.5 \nLog( εL1/ν )Log(ML β/ν )\nTT C2 (c) ν=0.64 \nγ=1.25 \nTC=T C2 (∞)\nFigure 7(c) slope= −α/ν = -0.12 \n1.5 2.0 \n-0.5 0.0 0.5 1.0 1.5 \nLog( εL1/ν )Log((C-b -)L −α/ν )\nL=8 \nL=9 \nL=10 \nL=11 \nL=12 α=0.12 \nν=0.64 \nb-=-70 TT C2 (b) \nTC=T C2 (∞)\nFigure 8(b) slope =- α/ν = -0.12 \n-0.3 0.2 \n-0.5 0.0 0.5 1.0 1.5 \nLog( εL1/ν )Log((C-b -)L −α/ν )\nTT C1 α=0.12 \nν=0.64 \nb+=-0.3 (d) \nFigure 8(d) 0.0 0.1 0.2 0.3 0.4 0.5 \n0.0 1.0 2.0 3.0 4.0 5.0 \ndkT/J SOPT \nFOPT \nF\ndf P\nTCP \nFigure 9 " }, { "title": "1412.6960v2.Possibility_of_ferromagnetic_neutron_matter.pdf", "content": "Preprint typeset in JHEP style - HYPER VERSION OU-HET-844\nRIKEN-MP-104\nPossibility of ferromagnetic neutron matter\nKoji Hashimoto\nDepartment of Physics, Osaka University, Toyonaka, Osaka 560-0043, Japan\nMathematical Physics Lab., RIKEN Nishina Center, Saitama 351-0198, Japan\nE-mail: koji(at)phys.sci.osaka-u.ac.jp\nAbstract: We study ferromagnetism at high density of neutrons in the QCD hadron\nphase, by using the simplest chiral e\u000bective model incorporating magnetic \felds and\nthe chiral anomaly. Under the assumption of spatial homogeneity, we calculate the\nenergy density as a function of neutron density, with a magnetization and a neutral\npion condensation a laDautry and Neyman. We \fnd that at a high density the energy\nof the ferromagnetic order is lower than that of the ordinary neutron matter, and\nthe reduction e\u000bect is enhanced by the anomaly. Compared to the inhomogeneous\nphase with the alternating layer structure, our ferromagnetic phase turns out to be\nunfavored. However, once an axial vector meson condensation is taken into account\nin our simplest model, the ferromagnetic energy density is lowered signi\fcantly, which\nstill leaves some room for a possible realization of a QCD ferromagnetic phase and\nferromagnetic magnetars.\nKeywords: Magnetization, Neutron, High density QCD .arXiv:1412.6960v2 [hep-ph] 13 Jan 2015Contents\n1. Motivation: QCD and ferromagnetism 1\n2. Toy model of dense neutral fermions 3\n2.1 Fermions, magnetization and constant magnetic \feld 3\n2.2 Ferromagnetism at higher density 5\n2.3 Favoring ferromagnetic phase 8\n2.4 Similarity to Nambu-Jona-Lasinio model 10\n3. Chiral model of neutrons with pion condensation, magnetic \feld\nand anomaly 11\n3.1 Dense neutrons and pions with axial anomaly 11\n3.2 Spontaneous magnetization and the pion condensation 14\n3.3 Anomaly enhancement and comparison to the ALS phase 16\n3.4 Inclusion of axial vector meson condensation 18\n3.5 AdS/CFT treatment with a large Ncapproximation 20\n4. Summary and discussion 22\n1. Motivation: QCD and ferromagnetism\nFerromagnetic order in nature always attracts interest for study as it manifests micro-\nscopic structure of matter and materials. Among observed magnetic \felds in nature,\nperhaps the strongest stable magnetic \feld is on the surface of magnetars, which\ngoes up to 1015[G] and more [ 1,2,3]. The mechanism for generating such a strong\n\feld is yet to be uncovered, and it is natural to resort the origin to the high density\nof neutrons of which the neutron stars consist. In fact, after the discovery of pulsars,\nthe possibility of ferromagnetism at neutron stars was proposed [ 4,5,6,?]. However,\nnumerical simulations of neutron matter with realistic inter-nucleon potentials have\nnot shown the ferromagnetic phase [ 7]. So the possibility of the ferromagnetic phase\nat high density neutron matter, if exists in nature, waits for a new mechanism of the\nspontaneous magnetization.\nIn this paper, we study the possibility of the ferromagnetic phase at high density\nof neutrons, by using the simplest but general chiral e\u000bective action. Low energy dy-\nnamics of neutrons is governed by the chirally symmetric interactions through pions\n{ 1 {and the spin-magnetic coupling with magnetic \felds. Our model consists of dense\nneutrons coupled with neutral pions and magnetic \felds, together with the chiral\nanomaly term. These are indispensable ingredients, and we will see the outcome for\nthe magnetic phase from this minimal model.\nThe reason for choosing the neutral pion is simply for the realization of the\nferromagnetism, as other pion condensations such as charged pion condensation [ 8,9]\nhave not been shown to exhibit a ferromagnetism. In addition, with a neutral pion\ncondensation of the form \u0005 0(x)/sink\u0001x, a neutron lattice is formed with an\nalternating layer structure (ALS) [ 10,11,12,13], then the neutron spins cancel each\nother, and macroscopic magnetization would not emerge. In this paper, instead, we\nanalyze a neutral pion condensation of the di\u000berent form \u0005 0(x) =q\u0001xfollowing\nDautry and Neyman [ 14]1, and generalize the study to include magnetic \felds and\nQCD anomaly.\nOur study is motivated by the earlier work [ 16] in which, together with M. Eto\nand T. Hatsuda, the author proposed a mechanism for a ferromagnetic phase at\nhigh density of neutrons. The mechanism utilizes a neutral pion domain wall [ 17]\ncoupled to the magnetic \feld through the QCD chiral anomaly [ 18]. A spontaneous\nmagnetization was shown in [ 16] in the approximation of a single wall and one-loop\nneutrons. In this paper, we generalize the idea, and study in the simplest chiral\nmodel the Fermi energy of the dense neutrons and its back-reaction due to the pion\ncondensation and the magnetic \felds. A successive array of the domain walls can be\napproximated by the linear pion condensation of Dautry and Neyman.\nLet us describe what we will \fnd in this paper.\n\u000fToy model of neutral fermions.\nFirst we provide a toy model of a neutral fermion with a Zeeman coupling to\nmagnetic \felds. Under the assumption of the spatial homogeneity, we calcu-\nlate the energy density of the ferromagnetic phase and show that it is favored\ncompared to the ordinary fermion matter. (Sec. 2)\n\u000fSimplest chiral model and ferromagnetic order.\nThe toy model of the neutral fermions is the essential part of the chiral model\nof neutrons and pions. We analyze the simplest chiral e\u000bective model of dense\nneutrons and neutral pions, together with the magnetic \feld coupling and the\nQCD anomaly. We \fnd that the neutral pion condensation of form proposed\nby Dautry and Neyman is precisely in the same place as the magnetization,\nunder the assumption of the spatial homogeneity. The energy density of the\nferromagnetic-pion-condensation phase is lower than the ordinary neutron mat-\nter at high density around \u001a >5\u001a0where\u001a0is the standard nuclear density.\n1For a recent review on the condensation (called chiral density wave), see [ 15].\n{ 2 {Furthermore, the chiral anomaly term actually helps the ferromagnetic order.\nThe generated magnetic \feld is \u001840 [MeV]\u0018O(1017)[G]. (Sec. 3.1, 3.2)\n\u000fComparison to ALS.\nWe compare our energy density with that of the inhomogeneous ALS phase\n(which does not exhibit a magnetization), and \fnd that the ALS phase is\nfavored. The energy gain of the ALS is by several times greater than that of\nour ferromagnetic phase. (Sec. 3.3)\n\u000fAxial vector meson condensation.\nTo seek for the possibility of the ferromagnetism, we look at the axial vector me-\nson condensation accompanied by our model. Indeed, any axial vector meson\nplays the same role as the neutral pions, and the axial vector meson condensa-\ntion further reduces the energy density of the ferromagnetic phase signi\fcantly.\nIncorporation of a higher vector meson tower and its condensation is studied\nby using the AdS/CFT correspondence. (Sec. 3.4)\nIn summary, we analyze the ferromagnetic order of our simplest chiral model of dense\nneutrons with magnetic \felds and the QCD anomaly. We \fnd that our ferromagnetic\norder, as its simplest form, is not favored compared to the ALS phase. We further\n\fnd that the axial vector meson condensation and the QCD anomaly, together with\nthe pion condensation of Dautry and Neyman, signi\fcantly helps the reduction of the\nenergy density, which suggests a necessity for further investigation for a realization\nof the ferromagnetic phase. The analysis in this paper is for the minimal model\nas we have emphasized above, so the result should be understood only qualitatively.\nIncorporation of realistic nuclear forces and nucleon contact terms, and also inclusion\nof electrons and protons, would be important for a further progress for realizing the\nQCD ferromagnetic phase at high density of neutrons.\nThe organization of this paper is as follows. In Sec. 2, we provide the toy model\nof neutral fermions and study a ferromagnetism at high density. In Sec. 3.1 and\nSec. 3.2, we analyze the simplest chiral model of neutrons with the pions, magnetic\n\felds and the chiral anomaly. In Sec. 3.3, we present our result on the energy plot\nand a comparison to the ALS phase is made. In Sec. 3.4 and 3.5, incorporation\nof the axial vector meson condensation is studied, with a help of the AdS/CFT\ncorrespondence. Sec. 4 is for a summary and discussions.\n2. Toy model of dense neutral fermions\n2.1 Fermions, magnetization and constant magnetic \feld\nWe are interested in the e\u000bect of spin and its magnetization, for a general fermion\nsystem. The magnetization is a condensation of a spin operator of fermions. Since\n{ 3 {for relativistic systems the spin operator of a fermion (x) is given by a spatial\ncomponent of an axial current,\nSi(x) =1\n2\u0016 \ri\r5 ; (2.1)\nwe can systematically write an action for the fermion with the spin magnetic coupling.\nWe consider the following general system of a neutral fermion. It is a system of a\nfree neutral fermion (x) with a mass min 4 spacetime dimensions, with a Zeeman\ncoupling under a dynamical magnetic \feld Bi,\nLfermion =\u0016 (i@\u0016\r\u0016\u0000m+i\r0\u0016) +\u000b\u0016 \ri\r5 Bi\u00001\n2B2\ni: (2.2)\nA chemical potential \u0016for the fermion number is introduced such that we can treat\nthe fermion density \u001a. Since the magnetization is a back-reaction to the spacetime\nmagnetic \feld, we have included the kinetic term for the magnetic \feld Bi. The\nsecond term in ( 2.2) is the Zeeman coupling \u000bbetween the spin of the fermion and\nthe magnetic \feld Bi. The Zeeman coupling is a part of a so-called Pauli term.2\nOur fermion does not have a charge, since we are interested in e\u000bects induced\nparticularly by the spin magnetic interaction.3So in our model there is no standard\ncanonical coupling between the gauge \feld for the magnetic \feld and the fermion\n . Normally, for a charged spin-1 =2 fermion with an electric charge e, the Zeeman\ncoupling is measured in the unit of a Bohr magneton, as\n\u000b=g\n2e\n2m(2.3)\nwheregis the \\g-factor\" and e=2mis the Bohr magneton. Our fermion does not\nhave the electric charge, so we shall treat \u000bas a general spin-magnetic coupling. The\nrelation ( 2.3) can be thought of as a reference, for example for a neutron which will\nbe treated in section 3.\nUnder a constant magnetic \feld Bi, we consider the behavior of the dense neu-\ntral fermions. We shall quantize the spin of the fermion along the direction of the\nmagnetic \feld. Then there are two Fermi seas, one is for the up spin and the other\nis for the down spin. In the presence of the background magnetic \feld Bi, due to the\nspin-magnetic Zeeman coupling, we have a Zeeman splitting for the Fermi energy for\nspin up and down states.\n2One would notice that the interaction term added in ( 2.2) does not respect the Lorentz invari-\nance, as only the spatial index iis summed. However, since we need the density for our analysis,\nthe chemical potential term already broke the Lorentz invariance in ( 2.2), so we need not worry\nabout it.\n3For charged fermions, the magnetic \feld provides Landau levels which may change the story\nquite a bit, and will bring an interesting outcome. We shall come back to the charged fermion case\nelsewhere.\n{ 4 {It is easy to evaluate the free energy of each spin sector. In the non-relativistic\nfermions where mass mis large compared to the Fermi energy of the fermions, we\nobtain\nF\"=\u0000(2m)3=2\n15\u00192(\u0016\u0000m\u0000\u000bB)5=2F#=\u0000(2m)3=2\n15\u00192(\u0016\u0000m+\u000bB)5=2:(2.4)\nThe di\u000berence is just the sign of the Zeeman coupling, due to the spins. We have\ndenotedBas the magnitude of the magnetic \feld Bi. The total free energy of the\nsystem, including the magnetic \feld energy is given by\nF=\u0000(2m)3=2\n15\u00192h\n(\u0016\u0000m\u0000\u000bB)5=2+ (\u0016\u0000m+\u000bB)5=2i\n+1\n2B2: (2.5)\nIn the following of this section, we analyze this free energy and study the ferromag-\nnetism.\n2.2 Ferromagnetism at higher density\n2.2.1 Complete polarization of the spins\nA large magnetic \feld is expected to correspond to a high density of the fermion. For\na large magnetic \feld, one of the two terms for the spins in the free energy becomes\nill-de\fned; the expression ( 2.5) is valid only when\n\u0016\u0000m\u0000j\u000bjB > 0: (2.6)\nFor a large magnetic \feld, this condition is not met. In that case, we need to use\nthe following expression for the free energy\nF=\u0000(2m)3=2\n15\u00192(\u0016\u0000m+j\u000bjB)5=2+1\n2B2: (2.7)\nThe spins are fully aligned (see Figure 1Left).\nTo turn this free energy (as a function of the chemical potential) to the energy (as\na function of the fermion density), let us make a Legendre transform. The fermion\nnumber density is given by\n\u001a\u0011\u0000@F\n@\u0016=(2m)3=2\n6\u00192(\u0016\u0000m+j\u000bjB)3=2: (2.8)\nThen the energy is given by\nE\u0011F+\u0016\u001a=m\u001a+35=3\u00194=3\n21=351\nm\u001a5=3\u0000j\u000bjB\u001a+1\n2B2: (2.9)\nThe interpretation of each term is quite clear. The \frst term is the fermion mass\nenergy, as\u001ais the number density of the fermion. The second term is the Fermi\n{ 5 {Fermion energy\t\rFermion energy\t\rSpin up Spin down Spin up Spin down Figure 1: The fermi surface and polarization of spins. The spin-magnetic coupling modi-\n\fes the depth of the dispersion relation according to the fermion spins. Left: all the spins\nare polarized. Right: there remains some density of the opposite component of the spin.\nenergy. The third term is due to the spin magnetic coupling. And the last term is\nfor the magnetic \feld self energy.\nWe would like to \fnd an energy minimum for a given fermion density \u001a. It is\nquite straightforward, since the last two terms in the energy can be written as a\nperfect squared,\nE=m\u001a+35=3\u00194=3\n21=351\nm\u001a5=3+1\n2(B\u0000j\u000bj\u001a)2\u00001\n2\u000b2\u001a2: (2.10)\nSo, to minimize the energy, a spontaneous magnetization should take place,\nB=j\u000bj\u001a; (2.11)\nat which the energy density is given by\nE=m\u001a+35=3\u00194=3\n21=351\nm\u001a5=3\u00001\n2\u000b2\u001a2: (2.12)\n2.2.2 Co-existence of both spins\nThe magnetization at the high density in the description above assumes the complete\npolarization of the fermions. At not-so-high density of the fermions, we expect that\n{ 6 {not all the fermions are polarized (see Figure 1Right). Let us see indeed this is the\ncase.\nFrom the original total free energy ( 2.5), the equilibrium condition @F=@B = 0\nis\n0 =\u0000(2m)3=2\n15\u001925\n2\u000bh\n\u0000(\u0016\u0000m\u0000\u000bB)3=2+ (\u0016\u0000m+\u000bB)3=2i\n+B: (2.13)\nNote that this can be always satis\fed at B= 0. Therefore, no magnetization is\nalways a possibility of the equilibrium, and we need to compare if magnetized phase\nhas a lower energy density to conclude the ferromagnetism. As we shall see, for\nlower density there is no magnetization, while for a high density the ferromagnetism\nis preferred.\nTo see in more detail the density dependence, we calculate the density \u001aas\n\u001a\u0011\u0000@F\n@\u0016=(2m)3=2\n6\u00192h\n(\u0016\u0000m\u0000\u000bB)3=2+ (\u0016\u0000m+\u000bB)3=2i\n: (2.14)\nWe can eliminate \u0016by using the equilibrium condition ( 2.13), to obtain the equilib-\nrium condition in terms of the density,\n\u0012\n\u001a+B\n\u000b\u00132=3\n\u0000\u0012\n\u001a\u0000B\n\u000b\u00132=3\n=4\n(3\u0019)2=3m\u000bB: (2.15)\nThis equation determines the magnitude of the spontaneous magnetic \feld B, once\nthe density \u001ais given. (Again, B= 0 is an alternative solution satisfying this\nequation.)\nWe notice here that at a density\n\u001a=1\nj\u000bjB; (2.16)\nthe equation ( 2.15) can make one term vanish. This is nothing but the point when\nwe make a transition to the fully-polarized phase which we considered earlier. Sub-\nstituting ( 2.16) into ( 2.15), we obtain the threshold density\n\u001a2=32\u00194\n241\nm3\u000b6: (2.17)\nIf the density is above this value, \u001a > \u001a 2, the system in fully polarized and the\nanalysis reduces to what we have considered earlier.\nThere is another condition for which the equation ( 2.15) can have a non vanishing\nsolution for B. Using the following expansion\n(1 +\u000f)2=3\u0000(1\u0000\u000f)2=3=4\n3\u000f+8\n81\u000f3+O(\u000f5); (2.18)\n{ 7 {we notice that ( 2.15) can have a solution only when the slope around B\u00180 can\nsatisfy the following inequality\n\u001a2=3\u00014\n31\nj\u000bj\u001a<4mj\u000bj\n(3\u00192)2=3: (2.19)\nThis condition is rephrased as\n\u001a1\u0011\u00194\n31\nm2\u000b6<\u001a: (2.20)\nWhen\u001a\u0014\u001a1, we \fnd no solution for ( 2.15), other than B= 0. So, as is expected,\nfor low density there is no ferromagnetic phase.\nIn summary, we \fnd the following possible phases in our system;\n8\n<\n:\u001a\u0014\u001a1: B= 0\n\u001a1<\u001a<\u001a 2:B= 0 orB= nontrivial solution of ( 2.15), spin mixed\n\u001a2\u0014\u001a:B= 0 orB=j\u000bj\u001a(spins fully polarized)\n2.3 Favoring ferromagnetic phase\nTo study whether this ferromagnetic order can actually occur in the system of our\nconcern, let us compare the resultant energy ( 2.12) with the energy with no magnetic\n\feld (no magnetization).\nPuttingB= 0 reduces the system to that of the ordinary free fermion, and in\nthe non relativistic case, once given the density \u001a, we know the total energy\nEB=0=m\u001a+35=3\u00194=3\n10m\u001a5=3: (2.21)\nThe \frst term is the energy contribution from the fermion mass, and the second term\nis the fermion kinetic energy integrated to the Fermi surface.\nWe compare this EB=0with the total energy density with the fully polarized\nspins ( 2.12), to have\nE\u0000EB=0=35=3\u00194=3\n10m(22=3\u00001)\u001a5=3\u00001\n2\u000b2\u001a2: (2.22)\nIt is easy to show that this is always negative for the density \u001a\u0015\u001a2which is the\ncondition for the spin full polarization, see ( 2.17). So, we conclude that indeed the\nferromagnetic phase is preferred at the high density \u001a\u0015\u001a2.\nIt is also straightforward to show that even in the range \u001a1< \u001a < \u001a 2, the\nferromagnetic phase B6= 0 is preferred. To show this, we need numerical calculations\nsince the energy for this phase is not expressed in an analytic form.\nFinally, let us see the value of the chemical potential corresponding to the ferro-\nmagnetic phase, to \fnd some consistency conditions; \frst, a thermodynamic stability\ncondition, and second, the validity of the non-relativistic approximation. When all\n{ 8 {the spins are polarized, we have ( 2.11) which can be substituted to the relation\nbetween the density and the chemical potential ( 2.8), to \fnd\n\u0016=m+(6\u00192)2=3\n2m\u001a2=3\u0000\u000b2\u001a: (2.23)\nThe thermodynamics stability condition is\n@\u0016\n@\u001a>0 (2.24)\nwhich tells just the fact that larger chemical potential provides a higher density.\nUsing our relation ( 2.23) at the high density ferromagnetic phase, we have the ther-\nmodynamic stability condition\n\u001a<\u001a 3\u001122\u00194\n31\nm3\u000b6: (2.25)\nThe value\u001a3is larger than \u001a2, so the ferromagnetic phase is stable for \u001a2<\u001a<\u001a 3.\nSecond, we check the non-relativistic approximation. If we substitute the typical\nvalue\u001a=\u001a2for the ferromagnetic phase to the relation ( 2.23), we \fnd\n\u0016\u0000m+j\u000bjB= 2\u000132\u001941\nm\u000b2: (2.26)\nThis is the hight of the Fermi sea as measured from the bottom of the dispersion\nrelation, so the non relativistic approximation is valid when this value is much smaller\nthan the mass m,\n2\u000132\u00194\nm2\u001c\u000b2: (2.27)\nSo, our non relativistic approximation is valid when this condition is met for the\nspin-spin interaction coe\u000ecient \u000b2.\nTo gain more insight on the relation ( 2.27), let us adopt hypothetically the\nexpression of the magnetic moment for a charged fermion ( 2.3) (although our fermion\nis neutral). Using ( 2.3), the relation ( 2.27) is written as\n25=23\u00192\u001cjgje: (2.28)\nFor example, the observed values for electrons are jgj\u00182 ande2=4\u0019\u00181=137, so\nthis non-relativistic condition is not met. Note however that in this paper we are\ninterested in a neutral fermions, not the electron which has a minimal coupling to\nthe magnetic \feld. In the next section, we study neutrons in more details. We will\n\fnd that, although the g-factor for the neutrons is not so large, the non-relativistic\napproximation is valid: in addition to the magnetic \feld coupling, there appears a\npion coupling which plays the same role, and the approximation is valid for the total\ninteractions. The pion condensation is the main subject of the next section.\n{ 9 {2.4 Similarity to Nambu-Jona-Lasinio model\nIn the previous subsections, we have seen that the ferromagnetic phase is preferred\ncompared to the free neutral fermions, when the density is large enough. Let us brie\ry\ndiscuss the reason why the simple model ( 2.2) is expected to favor a ferromagnetic\nphase as for homogeneous phases. Indeed, we \fnd an interesting relation to the\nfamous Nambu-Jona-Lasinio (NJL) model [ 19,20] in the following. We can naturally\nassume that the phases under consideration is spatially homogeneous, therefore there\nis no electric \feld generated. In that case, the \feld Biserves as an auxiliary \feld\nand we can integrate it out in our system ( 2.2).4The resultant Lagrangian is\nLfermion =\u0016 (i@\u0016\r\u0016\u0000m+i\r0\u0016) +1\n2\u000b2\u0000\u0016 \ri\r5 \u00012: (2.29)\nImmediately we can see a resemblance to the NJL model, the renowned model for a\nspontaneous chiral symmetry breaking. The NJL model is characterized by a four-\nfermion interaction ( \u0016 )2, which can be thought of as a squared of chiral condensate\n\u0016 . The four-fermion interaction governs the condensation of the operator \u0016 . Our\nmodel can be considered as a generalization of the NJL model by replacing the\n(\u0016 )2coupling with the spin-spin interaction5SiSi. Since fermions possess spins,\nonce we turn on a nonzero density for the fermions, the spin-spin interaction may\ncause a spontaneous magnetization, as in the case of the NJL model. In fact, the\nspin-spin interaction is a popular interaction in condensed matter physics. When\nthe coe\u000ecient \u000b2of the last term is positive, the system is expected to favor a\nspontaneous magnetization, i.e.a ferromagnetic phase.\nFor the phase to be realized, a high density would be necessary so that the\nneighboring fermions can interact.6Therefore we also expect a phase transition from\nthe normal phase to the ferromagnetic phase as we increase the density, and the\ncritical density should be a function of the coupling \u000band the mass of the fermion\nmsince these are the only parameters of our system. This is what we have seen in\nthis section, and the similarity to the NJL model allows us to intuitively understand\nthe origin of the ferromagnetism.\nBefore ending this section, we should note one thing. Our analysis in this section\nassumes the homogeneity in space. Normally one can allow inhomogeneous pro\fle\nof the matter, which results in a spontaneous emergence of a spatial modulation. A\nmodulated phase would have smaller energy density compared to the ferromagnetic\n4Note that this integration is not allowed normally, but here we ignore the electromagnetic\npropagation. However, for a discussion of only a homogeneous phase, one can make the integration\nand it provides an intuitive picture.\n5To recover the Lorentz invariance of the interaction term of the system, one can add the axial\ndensity squared term\u0000\u0016 \r0\r5 \u00012so that the interaction recovers the Lorentz invariance. It is not\nour scope of this paper.\n6Note that in the NJL model, in contrast, the fermion density is not necessary for the conden-\nsation, and the phase is unique.\n{ 10 {phase studied in this section. In the analysis in this section, we treated only a\nconstant magnetic \feld Bi. However, normally the integration of Bias a constant\nauxiliary \feld is not allowed, because photons propagate and Biis a part of the\nphoton kinetic term. Once one integrates out the electromagnetic \feld properly, one\n\fnds a non-local action of fermions. The integrated nonlocal action can be used for\nanalyses of inhomogeneous phases of the fermions, see [ 21] for example. In this paper\nwe consider a homogeneous ferromagnetic phase, and whether it is realized or not\nshould be determined by a comparison with inhomogeneous phases. As for the QCD\napplication, we shall discuss this problem later in the next section.\n3. Chiral model of neutrons with pion condensation, magnetic\n\feld and anomaly\nWe saw in the previous section that a generic neutral fermion system, with the simple\nZeeman coupling, is shown to exhibit a ferromagnetism, under the assumption of the\nspatial homogeneity. As a concrete example, in this section we investigate a neutron\nmatter at a high density. Neutrons interact with each other not only via the mag-\nnetic \feld and the spin-magnetic interaction but also a pion exchange. Interestingly,\nthe two interactions have the same structure, under a simple pro\fle for a pion con-\ndensation. The pion condensation part is a laDautry and Nyman [ 14]. In addition,\nQCD has an axial anomaly term which relates the two condensations | the magnetic\n\feld and the pion condensation, and in fact enhances each other. The enhancement\nmakes the total free energy decrease. We evaluate the total energy density of the\nferromagnetic phase. Finally we compare the resultant ferromagnetic phase with the\nwell-studied ALS (Alternating layer structure) phase for pion condensation.\n3.1 Dense neutrons and pions with axial anomaly\n3.1.1 Axial anomaly for the pion Lagrangian\nLow energy action of QCD is given by the standard Lagrangian of the linear sigma\nmodel dictated by the breaking of the chiral symmetry,\nL=\u0016 (i@\u0016\r\u0016\u0000g(\u001b+i\r5\u001c\u0001\u0019)) \n+1\n2(@\u0016\u001b)2+1\n2(@\u0016\u0019)2\u0000m2\n\u0019f\u0019\u001b\u0000V(\u001b2+\u00192): (3.1)\nHere = (p;n)Tis the nucleon \feld, and \u001band\u0019are sigma model \felds leading to\npions.f\u0019is the pion decay constant, and m\u0019is the pion mass. The global symmetry\nis the chiral symmetry U(2)L\u0002U(2)R. The chiral symmetry is broken due to the\nchiral condensate, \u001b2+\u00192=f2\n\u0019, which is realized by the potential term V. Once\nthe sigma model \feld obtains the expectation value, the nucleons acquire a mass,\n{ 11 {gf\u0019=MN. In this paper we do not consider the di\u000berence of the masses for protons\nand neutrons.\nIn the ideal case with no proton, and no charged pions, the Lagrangian is\nL\u001b=\u0016 n(i@\u0016\r\u0016\u0000g(\u001b\u0000i\r5\u00193)) n\n+1\n2(@\u0016\u001b)2+1\n2(@\u0016\u00193)2\u0000m2\n\u0019f\u0019\u001b\u0000V(\u001b2+\u00192\n3); (3.2)\nwhere nis the neutron \feld. Since we want to deal with \fnite density of neutrons,\nwe include a chemical potential term for the neutron,\nLn=i\u0016 n\r0\u0016n n: (3.3)\nIn the presence of the magnetic \feld with which the neutrons interact through their\nmagnetic moment, we add the following Lagrangian,\nLB=\u00001\n2B2\ni+1\n2\u0016 n\ri\r5 ngne\n2MNBi: (3.4)\nThe \frst term is the energy of the magnetic \feld. The second term is the Pauli term\nfor the interaction between the magnetic moment of the neutron and the magnetic\n\feld. Note that the spin density of the fermion is given by1\n2\u0016 n\ri\r5 n, andgnis the\nneutrongfactor.\nIn the presence of the magnetic \feld and the neutral pion condensation which is\nspatially dependent, there exists an axial anomaly term,\nLanom=\u0000ie\n4\u00192f2\n\u0019\u0016em\u0014\n(\u001b+i\u00193)y@i(\u001b+i\u00193)\u0015\nBi: (3.5)\nHere\u0016emis the electromagnetic chemical potential. This term is relevant for, for\nexample, the neutral pion decay \u00190!2\rvia the axial anomaly, as is seen from\nthe fact that the electromagnetic chemical potential can be thought of as a constant\nbackground electrostatic potential A(em)\n0. So, our total Lagrangian is\nL=L\u001b+Ln+LB+Lanom: (3.6)\n3.1.2 Neutral pion condensation\nWe consider a neutral pion condensate, following Dautry and Nyman [ 14],\n\u001b+i\u00193=f\u0019exp(iq\u0001x); \u0019 1+i\u00192= 0: (3.7)\nThis corresponds to a speci\fc condensation of the neutral pion in the nonlinear\nrepresentation, since the relation between the linear and non-linear representation is\n\u001b+i\u00193\u0018f\u0019exp(i\u00050(x)) where \u0005 0(x) is the physical neutral pion excitation. The\ncondensation ( 3.7) corresponds to \u0005 0(x)\u0018q\u0001x, a linear pro\fle in space. This can\n{ 12 {be regarded as a dense parallel domain walls which was considered in the context of\nanomaly-enhanced pion condensation in [ 16].\nIn this paper, we shall generalize the study of Dautry and Neyman ( 3.7), to\ninclude the magnetization and the QCD anomaly. With this condensation ( 3.7), the\nanomaly termLanom is given simply as\nLanom=e\n4\u00192\u0016emqiBi: (3.8)\nIn the following, without losing generality, we can turn on only the x3components\nof the magnetic \feld and q3, which will be denoted as Bandq.\nAccording to Dautry and Nyman, if the condensation qis large, the neutron\nspins are fully polarized. In the non-relativistic approximation for the neutron Fermi\nmomentum, the free energy for the free neutrons in the background pion condensation\nand the magnetic \feld is derived from the Lagrangian above,\nFn=\u0000(2MN)3=2\n15\u00192\u0012\n\u0016n\u0000MN+1\n2gAq\u0000gneB\n4MN\u00135=2\n: (3.9)\nHere we have introduced the axial coupling gAwhich is, at the tree level, equal to gin\nthe\u001b-model Lagrangian L\u001b. The total Free energy including the pion condensation\nand the magnetic \feld is\nF=Fn+f2\n\u0019m2\n\u0019+1\n2f2\n\u0019q2+1\n2B2\u0000e\n4\u00192\u0016emqB: (3.10)\nThe second term is from the pion mass term together with the pion condensation\n(3.7). The third term is from the pion kinetic term with ( 3.7). The last term is the\naxial anomaly term.\n3.1.3 Hamiltonian and the neutron density carried by the anomaly\nOur interest is the core of the neutron star where we have the \f-equilibrium. In\naddition to the neutrons, there exist protons and electrons. The electromagnetic\nchemical potential is given by\n\u0016em=1\n2(\u0016p\u0000\u0016e) (3.11)\nwhere\u0016pand\u0016eare proton and electron chemical potential, respectively.7Assuming\nthe\f-equilibrium, we impose \u0016n=\u0016p+\u0016e. And since we approximate the system by\nthe pure neutron matter for simplicity, we also impose the charge neutrality condition\nin a trivial manner, \u001ap=\u001ae= 0, which is equivalent to have \u0016p=MN. Then the\nanomaly term in the free energy ( 3.10) is written as\n\u0000e\n4\u00192\u0016emqB=\u0000e\n4\u001921\n2(\u0016p\u0000(\u0016n\u0000\u0016p))qB=\u0000e\n4\u00192\u0012\nMN\u00001\n2\u0016n\u0013\nqB: (3.12)\n7The factor 1 =2 should be there, because the total Free energy given with the number density\nshould be\u0016p\u001ap+\u0016e\u001ae, and the total electric charge is \u001ap\u0000\u001ae. Its canonical conjugate is1\n2(\u0016p\u0000\u0016e).\n{ 13 {We evaluate the energy density of the system as a function of the neutron density\nand the condensation qand the magnetic \feld B. The neutron density is computed\nas\n\u001an=\u0000@F\n@\u0016n=(2MN)3=2\n6\u00192\u0012\n\u0016n\u0000MN+1\n2gAq\u0000gneB\n4MN\u00133=2\n+e\n8\u00192qB: (3.13)\nThe last term is the anomaly-induced baryon charge. Using the expression, the \fnal\nresult for the energy density is given by\nE=F+\u0016n\u001an\n=35=3\u00194=3\n21=351\nMN\u0010\n\u001an\u0000e\n8\u00192qB\u00115=3\n+\u0012\nMN\u00001\n2gAq+gneB\n4MN\u0013\u0010\n\u001an\u0000e\n8\u00192qB\u0011\n+f2\n\u0019m2\n\u0019+1\n2f2\n\u0019q2+1\n2B2\u0000e\n4\u00192MNqB: (3.14)\nFor a comparison, we write the expression of Dautry and Nyman [ 14]:\nE=35=3\u00194=3\n21=351\nMN(\u001an)5=3+\u0012\nMN\u00001\n2gAq\u0013\n\u001an\n+f2\n\u0019m2\n\u0019+1\n2f2\n\u0019q2: (3.15)\nThis is obtained from ( 3.14) by just putting B= 0. The di\u000berence from just the pion\ncondensation is obvious. Let us look at the second term of ( ??), which in fact exhibits\nthe nature of our model explicitly. In the absence of the pion condensation and the\nmagnetic \feld, the second term is simply \u001anMN. This is the cost of the energy due to\nthe mass of the neutron. Now, the cost for each neutron can be reduced by the pion\ncondensation due to the axial coupling, and by the magnetic \feld times the neutron\nmagnetic moment, as MN!MN\u00001\n2gAq+gneB\n4MN. Furthermore, the anomaly term\ncan reduce e\u000bectively the density of the neutrons, \u001an!\u001an\u0000e\n8\u00192qB. In addition, the\nlast term of ( 3.14) is for the QCD anomaly, and it makes the total energy decrease\nfurther.\n3.2 Spontaneous magnetization and the pion condensation\nFor a given density \u001anof the neutrons, we can minimize the energy E(3.14). Later\nwe present our numerical results. But here, to explain the intrinsic behavior of the\nsystem, we evaluate the minimization of the energy in the absence of the anomaly\nterm. Without the anomaly term, the energy density is simpli\fed as\nE=35=3\u00194=3\n21=351\nMN\u001a5=3\nn+\u0012\nMN\u00001\n2gAq+gneB\n4MN\u0013\n\u001an\n+f2\n\u0019m2\n\u0019+1\n2f2\n\u0019q2+1\n2B2: (3.16)\n{ 14 {The energy is quadratic in qandB, so we can analytically \fnd the minimum of the\nenergy. In fact, the energy density expression ( 3.16) can be brought to the following\nform with the perfect squared,\nE=E0+1\n2f2\n\u0019\u0012\nq\u0000gA\n2f2\n\u0019\u001an\u00132\n+1\n2\u0012\nB+gne\n4MN\u001an\u00132\n; (3.17)\nwhere the minimum energy is\nE0=35=3\u00194=3\n21=351\nMN\u001a5=3\nn+MN\u001an\u0000g2\nA\n8f2\n\u0019\u001a2\nn\u0000g2\nne2\n32M2\nN\u001a2\nn: (3.18)\nThe last term in the minimum energy density E0is due to the magnetic \feld. The\n\frst three terms are that of Dautry and Neyman [ 14], and compared to that, our\nenergy is smaller by the last term.\nThe minimization of the energy is achieved when the perfect squares in ( 3.18)\nvanish,\nB=\u0012\u0000gn\n4MN\u0013\ne\u001an; (3.19)\nq=gA\n2f2\n\u0019\u001an: (3.20)\nWe have obtained the spontaneous magnetization of the neutron matter. The gener-\nated magnetic \feld is a monotonic function of the density, and in particular in this\ncase of the absence of the anomaly, it is a linear function in the density.\nThe free energy Fnfor the neutrons, ( 3.9), is for fully polarized neutrons. Let\nus check if this can be achieved for qandBwhich we obtained above for given \u001an.\nThe condition that the opposite spin state is absent is\n\u0016n\u0000MN\u00001\n2gAq+gneB\n4MN<0: (3.21)\nThis means that the Fermi sea for the opposite spin state is below the conducting\nband. At the minimum of the energy, we obtain \u001andependence of the chemical\npotential\u0016nfrom ( 3.13) with the solution ( 3.20) and ( 3.19) as\n\u0016n=(6\u00192)2=3\n2MN\u001a2=3\nn+MN\u0000\u0012g2\nA\n4f2\n\u0019+g2\nne2\n16M2\nN\u0013\n\u001an: (3.22)\nIn terms of \u001an, the condition of the full polarization is equivalent to\n(6\u00192\u001an)2=3<4MN\u0014g2\nA\n4f2\n\u0019+g2\nne2\n16M2\nN\u0015\n\u001an: (3.23)\nSubstituting values as MN= 938 [MeV], e2=4\u0019= 1=137,f\u0019= 95 [MeV], gA= 1,\ngn=\u00003:8 andm\u0019= 135 [MeV], we obtain\n\u001an>0:39 [fm\u00003]: (3.24)\n{ 15 {02468101214859095100105110\u001an=\u001a0(E0\u0000MN)=\u001an[MeV]\nFigure 2: A plot of the energy per a neutron, as a function of the neutron density \u001an.\nStraight line: ordinary neutron matter without the pion condensation. Thick curved line:\nour result with both pion condensation qand magnetization Bwith the QCD anomaly term.\nThin curved line: the result of Dautry and Neyman [ 14] with only the pion condensation\nq. Dashed line: the energy with both qandBbut without the QCD anomaly term.\nThis shows that for the density around twice of the standard nuclear density, all the\nspins are polarized.\nAnother constraint comes from a thermal stability condition. At any thermal\nequilibrium, we need to make sure\n@\u0016n\n@\u001an>0: (3.25)\nThe condition ( 3.25) can be evaluated as\n\u001an<57:6 [fm\u00003]: (3.26)\nThe bound is extremely high density and unrealistic, so this thermodynamic insta-\nbility region is far above realistic neutron density.\n3.3 Anomaly enhancement and comparison to the ALS phase\nIn the previous subsection, we found that even without the anomaly term the total\nenergy density is lowered by the magnetic \feld. In fact, the magnetic coupling works\nin the same manner as the pion coupling. Now, let us see how the anomaly term can\nhelp the condensation. The full expression for the total energy density including the\nanomaly term was given in ( 3.14), and we can \fnd the minimum energy con\fguration\nby varying qandB. Analytic analysis is not easy since the energy is not quadratic\ninqandB, so we perform a numerical analysis to \fnd the energy minimum. The\nnumerical results are summarized in Fig. 2.\nFig.2is a plot of the energy per a neutron, as a function of the neutron density\n\u001an. The neutron density is normalized by \u001a0which is the standard nuclear density.\n{ 16 {46810121410203040506070\u001an=\u001a0p\neB[MeV]\nFigure 3: A plot of the magnetic \feld spontaneously generated, as a function of the\nneutron density \u001an. The neutron density is normalized by \u001a0(the standard nuclear density).\nThe evaluation is with the fully-polarized neutrons.\nThe thick line is our result with the anomaly term. We observe that for a larger\ndensity, the energy per a neutron decreases.\nIn Fig. 2, for a comparison, we show a thin curved line which is the result\nof Dautry and Nyman [ 14] (that is, with no magnetic \feld Bbut with the pion\ncondensation q). The dashed line in Fig. 2is the energy density with both qandB\nbut without the anomaly term. We can see that the anomaly term makes the energy\nper a neutron decrease. The straight line on the left is for free neutrons without the\npion condensation qand without B. So, as a comparison to the ordinary neutron\nmatter, we see that the ferromagnetic phase is preferred at high density.\nWe plot the magnetic \feld as a function of \u001an, in Fig. 3. It is a monotonic func-\ntion of the neutron density. We \fnd that the magnitude of the generated magnetic\n\feld isO(102) [MeV] and thus it is of the QCD scale.p\neB\u001840[MeV] corresponds\ntoO(1017)[G].\nNow, let us discuss whether our ferromagnetic phase is favored or not, in reality.\nThe famous phase for a pion condensation is the ALS (Alternating layer structure)\nphase [ 10,11,12], and we can compare the result of the ALS phase with ours. See\nthe result of the ALS (Fig. 4) and compare it with our result (Fig. 2). Since already\naround\u001an=\u001a0\u00185 the energy reduction of the ALS phase compared to the ordinary\nneutron matter is 70 [MeV] (see Fig. 4), while for our ferromagnetic phase the energy\nreduction is only 10 [MeV]. So, from this comparison, we conclude that the ALS phase\nis favored against our ferromagnetic phase.\nOur analysis in this paper is with the simplest model of neutrons, and we have\nnot included full nuclear forces. Once we include them in addition to our neutral\npion coupling and the Zeeman coupling, the total free energy may change. In fact, in\nthe following subsections, we include axial vector meson condensation and it makes\nthe total energy further decrease drastically.\n{ 17 {Figure 4: An energy plot taken from [ 12], comparing the ALS phase and the Fermi gas\n(ordinary neutron matter). The horizontal axis is the neutron density \u001ain the unit of\nthe standard nuclear density \u001a0, while the vertical axis is the energy gain per a neutron.\nThe upper curve is for the Fermi gas and the lower curve is for the ALS. Solid lines are\nfor neutrons, and dashed lines are for symmetric nuclear matter. The arrows indicate the\nenergy reduction by the ALS.\nIn summary, here our observation is that the ferromagnetism is closely related\nto the neutral pion condensation, and the axial anomaly can help the total energy\nto decrease and enhance the magnetic \feld. The ferromagnetic phase has an energy\ndensity smaller than that of the ordinary neutron matter. But the energy of the ALS\nphase is smaller, in the approximation presented.\n3.4 Inclusion of axial vector meson condensation\nIn the previous subsections, we considered only the neutral pion \feld for the coupling\nto the spins of the neutrons. The spin operator of the neutron is nothing but the\nspatial components of the axial current, and in QCD, we expect in\fnite number\nof quark bound states which can couple to the axial current. They are axial vector\nmesons whose spectrum starts at the lowest with a1(1260) meson. In this subsection,\nwe shall add the contribution of this lowest axial vector meson and will \fnd that it\nwill further make the free energy decrease, together with the full spin polarization.\nThe main system which we treat in this paper is described by ( 2.2). On the\nother hand, the e\u000bective Lagrangian for the axial vector meson a\u0016(x) coupled to the\naxial current and the neutrons is\nL=\u0016 (i@\u0016\r\u0016\u0000m+i\r0\u0016) \n+1\n2gaNN\u0016 \r\u0016\r5 a\u0016+1\n4(@\u0016a\u0017\u0000@\u0017a\u0016)2\u0000m2\na\n2a2\n\u0016: (3.27)\nHere,gaNN is the coupling of the axial vector meson to the neutron axial current,\n{ 18 {andmais the mass of the meson. For the lowest a1(1260) meson, the measured value\nisma= 1230\u000640 [MeV].\nAs before, we concentrate on a homogeneous phase, and let us assume a constant\nvacuum expectation value of the spatial component of the axial vector meson,\nhaii= const.6= 0: (3.28)\nThen, re-writing bi\u0011aima, the e\u000bective Lagrangian is now\nL=\u0016 (i@\u0016\r\u0016\u0000m+i\r0\u0016) +gaNN\n2ma\u0016 \ri\r5 bi\u00001\n2b2\ni: (3.29)\nWe observe that this Lagrangian has precisely the same form as ( 2.2), so the same\nmechanism of lowering the free energy by spin alignment can work.\nThis addition of the axial vector meson to the pion system modi\fes the total\nfree energy ( 3.10) a little bit. The resultant free energy is\nF=\u0000(2MN)3=2\n15\u00192\u0012\n\u0016n\u0000MN+1\n2gAq\u0000gne\n4MNB+gaNN\n2mab\u00135=2\n+f2\n\u0019m2\n\u0019+1\n2f2\n\u0019q2+1\n2B2+1\n2b2\u0000e\n4\u00192\u0016emqB: (3.30)\nHere we determined the orientation of biin space such that it may strengthen the\nspin polarization, and denote bas the magnitude of bi. Note that the axial vector\nmesonbenters exactly in the same manner as that of the pion condensation qand the\nmagnetic \feld Bexcept for the anomaly term (the last term in ( 3.30)). So, basically\nthe addition of the axial vector condensation enhances the spin polarization of the\nneutrons, and further reduces the energy density.\nAs before, to gain an intuition of the behavior of the system, we \frst analyze\nthe system without the anomaly term. Then the total energy is\nE=35=3\u00194=3\n21=351\nMN\u001a5=3\nn+\u0012\nMN\u00001\n2gAq+gneB\n4MN+gaNN\n2mab\u0013\n\u001an\n+f2\n\u0019m2\n\u0019+1\n2f2\n\u0019q2+1\n2B2+1\n2b2: (3.31)\nCompared to ( 3.16), we \fnd that we have additional terms\n\u0001E=gaNN\n2mab\u001an+1\n2b2; (3.32)\nwhich is independent of the other variables qandB. So we can minimize it indepen-\ndent of the other terms, and \fnd\n\u0001E0=\u0000g2\naNN\n8m2\na\u001a2\nn (3.33)\n{ 19 {024681020406080100\u001an=\u001a0gaNN\u00189\ngaNN\u001818(E0\u0000MN)=\u001an[MeV]\nFigure 5: A plot of the energy per a neutron, as a function of the neutron density \u001an.\nOn the previous \fgure for the pion condensation, here we added two thick dashed lines,\nshowing the axial vector condensation. The upper thick dashed line is for gaNN = 9 and\nthe lower is for gaNN\u001818.\nwith an axial vector condensation\njhaiij=b\nma=gaNN\n2m2\na\u001an: (3.34)\nTo evaluate the energy \u0001 E0in (3.33), we need the value of the axial vector\ncouplinggaNN. We refer to a generic argument of the chiral symmetry by regarding\nthe axial vector meson as a gauge boson of the symmetry [ 22,23,24],\ngaNN\nma=2gA\nm\u0019: (3.35)\nA naive substitution gA\u00181 and the mass for the pion and the a1meson provides\ngaNN\u001818. Another estimate is as follows. We take care of one of the other equations\ncoming from the chiral symmetry argument [ 22,23,24],ma=p\n2m\u001awhich is not\nwell satis\fed by the physical masses of the \u001ameson and the a1meson. So instead\nof using the a1meson mass in the chiral symmetry formula ( 3.35) we may use the\n\u001ameson mass m\u001a= 770 [MeV]. Then we obtain gaNN\u001816. However, a lattice\nsimulation with an axial vector dominance provides gaNN\u00189 (see for example [ 25]),\nso there is uncertainty for the coupling.\nIn our numerical estimate of the energy density, we choose two typical values,\ngaNN\u001818 and 9. Our result is shown in \fg. 5. We \fnd that the energy per a nucleon\ndrastically reduces further. Compared to the ALS phase, the case with gaNN\u001818\nhas a lower energy and thus favored. The case with gaNN\u00189 is almost at the same\norder with the ALS phase.\n3.5 AdS/CFT treatment with a large Ncapproximation\nThe AdS/CFT correspondence [ 26,27,28] is a well-estabilished tool for analyzing\nstrongly coupled gauge theory in a certain limit, and its application to QCD-like\n{ 20 {gauge theories were widely studied. However the AdS/CFT tools for strongly coupled\ngauge theories work practically for large Ncgauge theories and at the limit of strong\ncoupling, so it would not be suitable for precision analysis such as the energy gain\nvia the condensation which is our interest in this paper. Nevertheless, it is important\nto \fnd what kind of couplings among hadrons and the magnetic \feld is present in\nQCD, and what is the order of magnitude of the couplings. The AdS/CFT approach,\ncalled holographic QCD, is suitable for that purpose, and in this short subsection we\nshall investigate it.\nWe use Sakai-Sugimoto model [ 29,30] which is the stringy setup closest to QCD\nat present. The nucleon meson couplings were obtained in [ 31,32,33,34], and the\nQCD anomaly term was calculated in [ 30].\nBasically in holographic QCD we have a tower of mesons, and this is true for the\na1mesons. We have in\fnite number of axial vector mesons. On the other hand, we\nhave only a single pion (that is, in the model there does not appear excited resonances\nof the pion).\nIt is easy to read from [ 30] that the axial vector mesons does not participate in\nthe QCD anomaly term, so the only contribution to the anomaly term is the pion\ncoupling which we considered in this paper. So we do not need to take care of all\nthe mixing between the axial vector mesons and the magnetic \feld in the anomaly\nterm, at the leading large Ncexpansion and at the strong coupling limit.\nOn the other hand, the contribution of the axial vector meson to the nucleon\nspins, which we considered in the previous subsection, comes to a concern. Since we\nhave in\fnite number of axial vector mesons, all piles up as a sum and would cause\npossibly a tremendous contribution. We shall discuss the issue in the following.\nFirst, in the AdS/CFT correspondence, the axial vector mesons are gauge \felds\nat higher dimensions, and their interaction terms are basically given by the Yang-\nMills action in the higher dimensions. We need to excite only the \u001c3component of the\nisospin, while the Yang-Mills action contains only a commutator-type interaction, so\nthe direct interaction among the constant axial vector mesons vanish. This means\nthat we does not need to consider the inter-level interaction of the axial vector meson\ntower.\nWe have seen in the previous subsection that a single axial vector meson reduces\nthe total energy by ( 3.33), so when there exists a tower of the axial vector mesons\nwe have an energy reduction\n\u0001E0=\u0000\u001a2\nn1X\ni=1r(i); r(i)\u0011\u0012ga(i)NN\n8ma(i)\u00132\n(3.36)\nwhereiis the label of the resonances, and i= 1 corresponds to the lowest a1(1260).\nFrom this expression, we observe that all axial vector mesons contribute additively,\nand the issue is the magnitude of the ratio ga(i)NN=ma(i)wheniincreases.\n{ 21 {The ratio can be calculated analytically by the AdS/CFT correspondence [ 31].\nHowever, the approximation of large \u0015is not good, so here we provide only the\nresulting numbers for a reference. The method developed in [ 31] can be generalized\neasily for higher axial vector mesons, and we \fnd\nr(2)=r(1)'1:06; r(3)=r(1)'1:07 (3.37)\nat the large t'Hooft coupling limit. So the ratio does not decrease for larger i. This\nwould be natural from the original idea of gauged chiral symmetry by Wess and\nZumino [ 24] which derived the relation ( 3.35). Therefore, the e\u000bect of the inclusion\nof the higher axial vector mesons is important, and it has an e\u000bect of further reducing\nthe total energy density.\nAt largeNclimit, all the axial vector meson tower reasonably contribute since\nthe meson width is narrow, and one would imagine the tremendous amount of energy\nreduction by introducing all the axial vector meson tower. However, it is unnatural\nand an artifact of the large Nclimit, since in reality the meson width gets broader\nfor higher resonances and the higher mesons participate with higher energy but also\nwith more involved chiral interactions. So, here we just point out that axial vector\nmeson condensation has a tendency to further reduce the total energy density, and\nthe contribution from the tower of the resonances would not be negligible.\n4. Summary and discussion\nFor searching a QCD ferromagnetism at high density of neutrons, we studied the\nsimplest chiral Lagrangian ( 3.6) which accommodates neutrons at high density, the\npion condensation, the constant magnetic \feld with its self energy and the QCD\nanomaly. The pion condensation is a linear spatial pro\fle of the neutral pion ( 3.7)a\nlaDautry and Neyman [ 14] which generates a neutron spin alignment.\nWe solved a self-consistent equation for the total energy density for a given neu-\ntron density, by considering the neutron Fermi energy, the pion self energy and also\nthe self energy of the constant magnetic \feld. We have shown that the minimization\nof energy under the assumption of spatial homogeneity leads to the ferromagnetic\norder preferred compared to the ordinary neutron matter without the pion conden-\nsation, at the neutron density \u001a > 5\u001a0where\u001a0is the standard nuclear density.\nThe result is summarized in Fig. 2. The generated magnetic \feld (see Fig. 3) isp\neB\u001840[MeV] which is around O(1017)[G].\nHowever, a comparison to the ALS (alternating layer structure) phase [ 10,11,12],\nwhich is with another neutral pion condensation providing a spatially alternating spin\norder, shows that our ferromagnetic order has a larger energy density and thus is not\nfavored (see Sec. 3.3).\nWe further included axial vector mesons in our model, since the axial vector\nmeson condensation has the same coupling as the Dautry-Neyman neutral pion con-\n{ 22 {densation. We found that the axial vector meson enhances the energy reduction of\nthe ferromagnetic phase signi\fcantly (see Sec. 3.4). In QCD there exists a tower\nof axial vector meson resonances, and inclusion of the tower further enhances the\nreduction, which we roughly evaluated with the use of the AdS/CFT correspondence\n(Sec. 3.5).\nWe can summarize our results as follows:\n\u000fThe simple chiral model with the linear neutral pion condensation and magnetic\n\feld accommodates a ferromagnetic order.\n\u000fThe QCD anomaly term lowers the ferromagnetic energy.\n\u000fThe axial vector meson condensation further reduces the energy signi\fcantly.\n\u000fOur analysis is among spatially homogeneous phases, and needs to be compared\nin more detail with inhomogeneous phases such as the ALS.\nOur study is based on the simple chiral model ( 3.6), so the numerical results pre-\nsented in this paper is not suitable for a detailed comparison. For example, inclusion\nof realistic nuclear forces and nucleon contact terms would give more corrections.\nNevertheless, in our analysis, in particular the axial vector meson condensation is an\ninteresting and novel feature, and a further consideration would be of worth. In the\ncondensed phase of the axial vector mesons, low energy propagation modes are of\ninterest, in view of recent progress [ 35] in non-relativistic Nambu-Goldstone theorem\n[36,37].\nIn this paper, we concentrated on the hadron phase,8not a quark phase such\nas the color superconductivity. It would be interesting to extend our calculation, if\npossible, to a hadron-quark mixed phase. For a quark matter, the possibility of the\nferromagnetism was studied in [ 41,42], while the quark-hadron mixture phase was\nstudied in the context of neutron stars [ 43,44]. The high density phase of QCD\nstill leaves a large room to be discussed [ 45], and observations of the magnetars\n[46,47] and the magnetic \felds there should reveal more about the mystery of the\nhigh density phase.\nAcknowledgment. | The author is indebted to Tetsuo Hatsuda for the collabo-\nration of this work at its early stage. The author would like to thank Teruaki Enoto,\nMinoru Eto, Deog-Ki Hong, Kei Iida, Muneto Nitta, Yudai Suwa and Toru Tama-\ngawa for valuable discussions and comments. This research was partially supported\nby the RIKEN iTHES project.\n8See for a recent attempt without a pion condensation, [ 38,39,40].\n{ 23 {References\n[1] T. Enoto et al., \\Recent Suzaku studies of the X-ray emission from magnetars,\" AIP\nConf. Proc. 1427 , 68 (2012);\n[2] A. K. Harding and D. Lai, \\Physics of Strongly Magnetized Neutron Stars,\" Rept.\nProg. Phys. 69, 2631 (2006) [astro-ph/0606674].\n[3] S. Mereghetti, \\The strongest cosmic magnets: soft gamma-ray repeaters and\nanomalous X-ray pulsars,\" Astronomy and Astrophysics Review, 15, 225 (2008).\n[4] D. H. Brownell and J. Callaway, \\Ferromagnetic transition in superdense matter and\nneutron stars\", Il Nuovo Cimento B 60, 169 (1969).\n[5] M.J. Rice, \\The hard-sphere Fermi gas and ferromagnetism in neutron stars\", Phys.\nLett. A 29, 637 (1969).\n[6] S. D. Silverstein, \\Criteria for Ferromagnetism in Dense Neutron Fermi\nLiquids-Neutron Stars\", Phys. Rev. Lett. 23,139 (1969) [Erratum: ibid. 23, 453\n(1969)]. K. Makishima, \\Measuring Magnetic Fields of Neutron Stars,\" Prog. Th.\nPhys. Supp. , 151, 54 (2003).\n[7] G. H. Bordbar and M. Bigdeli, \\Spin Polarized Asymmetric Nuclear Matter and\nNeutron Star Matter Within the Lowest Order Constrained Variational Method\",\nPhys. Rev. C77, 015805 (2008) [arXiv:0809.3498 [nucl-th]].\n[8] G. Baym, D. Campbell, R. F. Dashen and J. Manassah, \\A Simple Model\nCalculation of Pion Condensation in Neutron Matter,\" Phys. Lett. B 58, 304 (1975).\n[9] G. Baym and D. K. Campbell, \\Chiral Symmetry And Pion Condensation,\"\nLA-UR-78-395.\n[10] R. Tamagaki and T. Takatsuka, \\Pi0 Condensation Resulting from Alternating\nLayer Spin Structure in Dense Neutron Matter,\" Prog. Theor. Phys. 56, 1340 (1976).\n[11] T. Takatsuka and R. Tamagaki, \\One-Dimensional Localized (ALS) Structure in\nNuclear System and pi0 Condensation,\" Prog. Theor. Phys. 58, 694 (1977).\n[12] T. Takatsuka, K. Tamiya, T. Tatsumi and R. Tamagaki, \\Solidi\fcation and Pion\nCondensation in Nuclear Medium. Alternating Layer Spin Structure with\nOne-Dimensional Localization Accompanying pi0 Condensate,\" Prog. Theor. Phys.\n59, 1933 (1978).\n[13] T. Takatsuka, R. Tamagaki and T. Tatsumi, \\Characteristic aspects of pion\ncondensed phases,\" Prog. Theor. Phys. Suppl. 112, 67 (1993).\n[14] F. Dautry and E. M. Nyman, \\Pion Condensation And The Sigma Model In Liquid\nNeutron Matter,\" Nucl. Phys. A 319, 323 (1979).\n{ 24 {[15] M. Buballa and S. Carignano, \\Inhomogeneous chiral condensates,\" To apper ain\nProg. Part. Nucl. Phys. [arXiv:1406.1367 [hep-ph]].\n[16] M. Eto, K. Hashimoto and T. Hatsuda, \\Ferromagnetic neutron stars: axial\nanomaly, dense neutron matter, and pionic wall,\" Phys. Rev. D 88, 081701 (2013)\n[arXiv:1209.4814 [hep-ph]].\n[17] T. Hatsuda, \\E\u000bects Of Finite Density On Chiral Domain Wall In Neutron And\nQuark Matter,\" Prog. Theor. Phys. 75, 301 (1986) [Erratum-ibid. 75, 996 (1986)].\n[18] D. T. Son and M. A. Stephanov, \\Axial anomaly and magnetism of nuclear and\nquark matter,\" Phys. Rev. D 77, 014021 (2008) [arXiv:0710.1084 [hep-ph]].\n[19] Y. Nambu and G. Jona-Lasinio, \\Dynamical Model of Elementary Particles Based\non an Analogy with Superconductivity. 1.,\" Phys. Rev. 122, 345 (1961).\n[20] Y. Nambu and G. Jona-Lasinio, \\Dynamical Model Of Elementary Particles Based\nOn An Analogy With Superconductivity. Ii,\" Phys. Rev. 124, 246 (1961).\n[21] K. Maeda, T. Hatsuda and G. Baym, \\Antiferrosmectic ground state of\ntwo-component dipolar Fermi gases: An analog of meson condensation in nuclear\nmatter,\" Phys. Rev. A 87, 021604 (2013).\n[22] S. Weinberg, \\Dynamical approach to current algebra,\" Phys. Rev. Lett. 18, 188\n(1967).\n[23] J. S. Schwinger, \\Chiral dynamics,\" Phys. Lett. B 24, 473 (1967).\n[24] J. Wess and B. Zumino, \\Lagrangian method for chiral symmetries,\" Phys. Rev.\n163, 1727 (1967).\n[25] S. A. Coon, B. H. J. McKellar and V. G. J. Stoks, \\Delta I = 1 axial vector mixing\nand charge symmetry breaking,\" Phys. Lett. B 385, 25 (1996) [nucl-th/9606054].\n[26] J. M. Maldacena, \\The Large N limit of superconformal \feld theories and\nsupergravity,\" Adv. Theor. Math. Phys. 2, 231 (1998) [hep-th/9711200].\n[27] S. Gubser, I. R. Klebanov, and A. M. Polyakov, \\Gauge theory correlators from\nnoncritical string theory,\" Phys.Lett. B428,105 (1998).\n[28] E. Witten, \\Anti-de Sitter space and holography,\" Adv.Theor.Math.Phys. 2, 253\n(1998).\n[29] T. Sakai and S. Sugimoto, \\Low energy hadron physics in holographic QCD,\" Prog.\nTheor. Phys. 113, 843 (2005) [hep-th/0412141].\n[30] T. Sakai and S. Sugimoto, \\More on a holographic dual of QCD,\" Prog. Theor.\nPhys. 114, 1083 (2005) [hep-th/0507073].\n{ 25 {[31] K. Hashimoto, T. Sakai and S. Sugimoto, \\Holographic Baryons: Static Properties\nand Form Factors from Gauge/String Duality,\" Prog. Theor. Phys. 120, 1093 (2008)\n[arXiv:0806.3122 [hep-th]].\n[32] K. Hashimoto, T. Sakai and S. Sugimoto, \\Nuclear Force from String Theory,\" Prog.\nTheor. Phys. 122, 427 (2009) [arXiv:0901.4449 [hep-th]].\n[33] D. K. Hong, M. Rho, H. U. Yee and P. Yi, \\Chiral Dynamics of Baryons from String\nTheory,\" Phys. Rev. D 76, 061901 (2007) [hep-th/0701276].\n[34] D. K. Hong, M. Rho, H. U. Yee and P. Yi, \\Dynamics of baryons from string theory\nand vector dominance,\" JHEP 0709 , 063 (2007) [arXiv:0705.2632 [hep-th]].\n[35] H. Watanabe and H. Murayama, \\Englert-Brout-Higgs Mechanism in Nonrelativistic\nSystems,\" arXiv:1405.0997 [hep-th].\n[36] H. Watanabe and H. Murayama, \\Uni\fed Description of Nambu-Goldstone Bosons\nwithout Lorentz Invariance,\" Phys. Rev. Lett. 108, 251602 (2012) [arXiv:1203.0609\n[hep-th]].\n[37] Y. Hidaka, \\Counting rule for Nambu-Goldstone modes in nonrelativistic systems,\"\nPhys. Rev. Lett. 110, 091601 (2013) [arXiv:1203.1494 [hep-th]].\n[38] J. P. W. Diener and F. G. Scholtz, \\Ferromagnetic phase diagram of neutron\nmatter,\" arXiv:1006.5564 [nucl-th].\n[39] J. P. W. Diener and F. G. Scholtz, \\Ferromagnetism in neutron and charge neutral\nbeta-equilibrated nuclear matter,\" arXiv:1306.2175 [nucl-th].\n[40] J. P. W. Diener, \\Ferromagnetic Phase Transitions in Neutron Stars,\"\narXiv:1305.7346 [nucl-th].\n[41] T. Tatsumi, \\Ferromagnetism of quark liquid,\" Phys. Lett. B 489, 280 (2000)\n[hep-ph/9910470];\n[42] T. Tatsumi, \\Magnetic \feld and quark matter in the core,\" [arXiv:1107.0807\n[hep-ph]].\n[43] K. Masuda, T. Hatsuda and T. Takatsuka, \\Hadron-Quark Crossover and Massive\nHybrid Stars with Strangeness,\" Astrophys. J. 764, 12 (2013) [arXiv:1205.3621\n[nucl-th]].\n[44] K. Masuda, T. Hatsuda and T. Takatsuka, \\Hadron?quark crossover and massive\nhybrid stars,\" PTEP 2013 , no. 7, 073D01 (2013) [arXiv:1212.6803 [nucl-th]].\n[45] K. Fukushima and T. Hatsuda, \\The phase diagram of dense QCD,\" Rept. Prog.\nPhys. 74, 014001 (2011) [arXiv:1005.4814 [hep-ph]].\n[46] T. Enoto, The Astronomical Herald, Japan, 105, 431 (2012).\n{ 26 {[47] S. A. Olausen and V. M. Kaspi, Astrophys. J. Suppl., 2121, 6 (2014),\nhttp://www.physics.mcgill.ca/ pulsar/magnetar/main.html (McGill Online\nMagnetar Catalog).\n{ 27 {" }, { "title": "0811.4430v1.Anomalous_Ferromagnetism_in_TbMnO3_Thin_Films.pdf", "content": "Anomalous Ferromagnetism in TbMnO3 Thin FilmsB. J. Kirby,1,a),D. Kan,2 A. Luykx,2 M. Murakami,2 D. Kundaliya,2 and I. Takeuchi21Center for Neutron Research, National Institute of Standards and Technology, Gaithersburg, Maryland, 20899, USA2 Department of Materials Science and Engineering, University of Maryland, College Park, Maryland, 20742, USAABSTRACTMagnetometry, x-ray, and neutron scattering have been used to study the structural and magnetic properties of a TbMnO3 thin film grown on a [001] SrTiO3 substrate by pulsed laser deposition. Although bulk TbMnO3 is a low temperature antiferromagnet, magnetometry measurements indicate the presence of low temperature ferromagnetism. Depth profiling by x-ray and polarized neutron reflectometry reveals a net sample magnetization that is commensurate with the film thickness, indicating that the observed ferromagnetism is not due to an altered surface phase (such as Mn3O4), or external impurities that might give rise to an artificial magnetic signal. Instead, these results show that the ferromagnetism is an intrinsic property of the TbMnO3 film. \r1 Multiferroic materials, which feature simultaneous coupled ordering of magnetic moments and electric polarization, are the subject of widespread research interest. Multiferroics hold the promise of novel devices in which the the magnetic properties are electrically controlled, and vice versa. One example is that of an antiferromagnetic multiferroic thin film used to produce a preferred magnetization orientation in an adjacent ferromagnetic thin film via exchange bias [1,2] coupling. For such a structure, it is possible to alter the nature of the exchange coupling (and therefore the magnetic orientation of the ferromagnet) by applying an electric field to change the interfacial magnetic domain structure of the multiferroic antiferromagnet. For example, it has recently been demonstrated that the magnitude of the low temperature exchange bias field in YMnO3 / permalloy multilayers changes as a function of the electric field applied during cooling.[3] \r To further pursue the technological possibilities of exchange biasing multiferroics, other materials need to be explored. In bulk, TbMnO3 is a multiferroic material which in bulk exhibits a complex magneto-electric phase diagram, including antiferromagnetism below 30 K - making it a candidate for exchange-bias applications.[4-6] However, scaling down manganite crystals into thin film form causes strain that can drastically affect the magnetic properties,[7] and relatively little work has been done to study the magnetic properties of TbMnO3 thin films. As a first step towards determining the utility of TbMnO3,for exchange-bias applications, we have produced a thin film of the orthorhombic phase of TbMnO3, and studied it with magnetometry, x-ray, and neutron scattering measurements. These measurements show that the TbMnO3 film exhibits an intrinsic ferromagnetic order. \r A TbMnO3 film was laser ablated from a stoichiometric TbMnO3 target, onto a [001] oriented SrTiO3 substrate. Figure 1 shows results of θ-2θ x-ray diffraction (XRD) scans, performed with a commercial diffractometer.[8] The fabricated film exhibits the [00l] peak of the perovskite structure, with an out-of-plane lattice parameter c = 0.372 nm. No peaks identifiable as originating from \r2additional phases were observed. X-ray reciprocal space mapping measurements around the SrTiO3 [103] Bragg reflection (not shown) reveal that the in-plane lattice parameter of the TbMnO3 film closely matches that of the cubic SrTiO3 substrate (0.391 nm). Bulk TbMnO3 has an orthorhombic perovskite structure which (in the pseudo-cubic convention) has lattice parameters of 0.393 nm in-plane, and 0.370 nm out-of-plane.[9] Therefore, we conclude the fabricated film is compressively strained by the substrate, resulting in the tetragonally distorted orthorhombic phase of TbMnO3. \r Net magnetization measurements of the sample were performed using superconducting quantum interference device (SQUID) magnetometry. Figure 2 shows the field H dependent magnetic moment m at temperature T = 5 K after cooling from 100 K in zero field.[10] Hysteretic behavior is observed, clearly indicating ferromagnetism - a drastic departure from the antiferromagnetic order observed for bulk TbMnO3. m(T) in a 1.5 T field, is shown in the Fig. 1 inset, and reveals that a net magnetization persists above 100 K. However, standard magnetometry measurements alone cannot determine whether the observed ferromagnetism is present throughout the film, or if it is instead due to a departure from the TbMnO3 stoichiometry in a small region of the film. \r To determine if a net magnetization is truly intrinsic to the TbMnO3 film, we determined the structural and magnetic depth profiles via x-ray and polarized neutron reflectometry [11] (XRR and PNR, respectively) measurements at the NIST Center for Neutron Research. XRR measurements were conducted at room temperature with wavelength λ = 0.154 nm x-rays (Cukα), and PNR measurements were conducted at cryogenic temperatures with λ = 0.475 nm cold neutrons. The neutron beam was polarized by Fe/Si supermirror and a Mezei spin-flipper to be alternately spin-up (+) or spin-down (-) relative to an in-sample-plane magnetic field H, and was incident on the sample. The reflected beam was intercepted by an analyzer supermirror/flipper assembly so that all four PNR cross sections (non spin-flip: R+ +, R- -, and spin-flip: R- +, R+ -) could be measured with a 3He pencil detector. \r3\r In reflectometry, depth (z) dependent sample properties are determined from the scattering vector (Q) dependent reflectivities.[12] The x-ray and neutron beams are both highly penetrating, and probe the entirety of the TbMnO3 thin film. XRR is sensitive to the sample’s x-ray scattering length density ρX(z) (a function of electron density),[13] while PNR is sensitive to the sample’s nuclear scattering length density ρN(z) and volume magnetization M(z).[14,15] Specifically, the non spin-flip scattering R+ + and R- - depend on ρN(z) and the component of M(z) parallel to H, while the spin-flip scattering R+ - and R- + depend solely on the component of M(z) perpendicular to H. Thus, the structural and magnetic depth profiles of thin film structures can be determined by model-fitting XRR and PNR spectra. For this work, model fitting was done using the NIST GA_REFL software package.[16] \r Fitted Cukα XRR data are shown in Figure 3. Clear oscillations are observed, indicating a strong sensitivity to the interface between the TbMnO3 film and the SrTiO3 substrate. PNR measurements of the same sample were conducted in a 0.55 T field, after cooling the sample from above 290 K to 6 K in a 5.5 T field. The spin-flip channels were measured only at 6 K, and scattering was found to be at background levels, indicating no detectable magnetic component perpendicular to H. The fitted non spin-flip scattering data are shown in Figure 4. At 6 K (a), R+ + and R- - exhibit clear oscillations that are roughly 180 degrees out of phase from one another. In the absence of nuclear spin polarization,[18] the only way these two spin states can differ is if the sample possesses a net magnetization parallel to H. Further, the common frequency of the R+ + and R- - oscillations is essentially the same as that of the XRR oscillations shown in Figure 3, suggesting that the magnetized film thickness is similar to the total film thickness. At 100 K (b), the splitting between R+ + and R- - persists, indicating that the sample is still magnetized. The higher frequency oscillations observed in panel a) and Fig. 3 are much weaker at 100 K. This indicates that while neutrons are very sensitive to the magnetic interface between the film and the substrate, they are not very sensitive to the nuclear interface. With the high frequency oscillations \r4damped, a very weak lower frequency oscillation becomes apparent, suggesting some depth dependence of ρN within the TbMnO3 film. \r Results of model fitting the XRR and PNR data are shown in Figure 5. The structural profiles are shown in panel a) (ρN and both the real and imaginary components of ρX). The x-ray profiles feature a distinct substrate/film interface, and show that the TbMnO3 film is 70 nm thick. The nuclear profile features a slight decrease in ρN from top to bottom of the film, possibly due to a small variation in film strain. The magnetic profiles are shown in Fig 5 b), and confirm that the TbMnO3 possess a net magnetization parallel to H that is essentially constant across the entirety of the 70 nm film. Given the lattice parameters determined from XRD, the profiles shown in Fig. 5 correspond to magnetic moments of 0.6 and 0.1 μB per unit cell at 6 K and 100 K, respectively. \r These results show that the ferromagnetism observed with SQUID is an intrinsic property of the TbMnO3 thin film. While we cannot completely rule out a uniform distribution of impurities as the origin of ferromagnetism (magnetic clusters, for example), we have definitively shown that the ferromagnetism is not due to surface impurities. Specifically, observation of a uniform sample magnetization at 100 K confirms that a surface layer of ferrimagnetic Mn3O4 (TC = 42 K) [20] is not the source of the ferromagnetic signal observed with SQUID. Instead, it seems that the ferromagnetic behavior is indeed an intrinsic property of strained orthorhombic TbMnO3 films. Very recent unpublished work by Rubi et al. also shows that thin TbMnO3 films on SrTiO3 substrates can exhibit ferromagnetism, and they present evidence suggesting that the ferromagnetism arises from in-plane compressive strain. [21] This is a likely explanation for the ferromagnetism we have observed, as our film is also compressively strained in-plane. Determining how to tune the growth parameters to produce fully antiferromagnetic TbMnO3 films for exchange bias applications remains an important avenue of future research. In summary, our results reveal additional complexity to the already rich TbMnO3 \r5magneto-electric phase diagram, and have important implications for the role of TbMnO3 as a biasing multiferroic antiferromagnet.This work was supported by NSF-MRSEC under grant No. DMR 0520471, NSF DMR 0603644, and ARO W9IINF-07-1-0410. We thank J. A. Borchers of NIST for valuable discussions. \r6REFERENCES[1] W. H. Meiklejohn and C. P. Bean, Phys. Rev. 105, 904 (1957).[2] J. Nogues and I. K. Schuller, J. Magn. Magn. Mater. 192, 203 (1999).[3] V . Laukhin, V . Skumryev, X. Marti, D. Hrabovsky, F. Sanchez, M. V . Garcia-Cuenca, C. Ferrater, M. Varela, U. Luders, J. F. Bobo, and J. Fontcuberta, Phys. Rev. Lett. 97, 227201 (2006).[4] T. Kimura, G. Lawes, T. Goto, Y . Tokura, and A. P. Ramirez, Phys. Rev. B 71, 224425 (2005). [5] T. Goto, T. Kimura, G. Lawes, A. P. Ramirez, and Y . Tokura, Phys. Rev. Lett. 92, 257201 (2004).[6] T. Kimura, T. Goto, H. Shintani, K. Ishizaka, T. Arima, and Y . Tokura, Nature 426, 55 (2003).[7] D. Rubi, Sriram Venkatesan, B. J. Kooi, J. Th. M. De Hosson, T. T. M. Palstra, and B. Noheda1, Physical Review B 78, 020408(R) (2008). [8] c = λ / 2sinθ, where λ is the x-ray wavelength of 0.154 nm, and θ is the scattering angle.[9] J. A. Alonso, M. J. Martinez-Lope, M. T. Casais, and M. T. Fernandez-Diaz, Inorganic Chemistry 39, 917 (2000).[10] Magnetometry measurements of similarly prepared TMO samples show that field cooling vs. zero field cooling has no observable effect on the resulting hysteresis loop.[11] See http://www.ncnr.nist.gov/instruments/ng1refl.[12] For monochromatic radiation (as for the cases described here), Q = 4π sinθ / λ, where θ is the scattering angle, and λ is the radiation wavelength.[13] L. G. Parratt, Physical Review 95, 359 (1954). [14] G. P. Felcher, Phys. Rev. B 29, 1268 (1984).[15] C. F. Majkrzak, Physica B 221, 342 (1996).[16] P. A. Kienzle, M. Doucet, D. J. McGillivray, K. V . O’Donovan, N. F. Berk, and C. F. Majkrzak, see http://www.ncnr.nist.gov/reflpak for documentation.\r7[17] M.R. Fitzsimmons, et al., Phys. Rev. B 76, 245301 (2007).[18] Kirby Dwight and Norman Menyuk, Phys. Rev. 119, 1470 (1960). [19] D. Rubi, C. deGraaf, C. J. M. Daumont, D. Mannix, R. Broer, B. Noheda, preprint available at http://ArXiv.org, arXiv:0810.5137v1. \r8FIGURESFigure 1: The SrTiO3 and TbMnO3 [001] reflections observed using x-ray diffraction. Error bars indicate +/- 1 sigma. \n\r9Figure 2: SQUID magnetometry results. Ferromagnetism is observed at 5 K, and a net magnetization persists above 100 K (inset). \n\r10Figure 3: Fitted XRR data for a typical TbMnO3 film. Some higher Q data is omitted for clarity. Error bars indicate +/- 1 sigma.\n\r11Figure 4: Fitted PNR. Spin-dependent oscillations are present at 6 K (a), and at 100 K (b, inset). Error bars indicate +/- 1 sigma.\n\r12Figure 5: Models used to fit the data in Fig. 2-3. a) Structural profiles: real and imaginary components of the x-ray scattering length density, and the nuclear scattering length density (multiplied 10). b) Magnetic profiles at 6 K and 100 K.\n\r13" }, { "title": "1808.03966v2.Resonant_Driving_induced_Ferromagnetism_in_the_Fermi_Hubbard_Model.pdf", "content": "Resonant Driving induced Ferromagnetism in the Fermi Hubbard Model\nNing Sun,1Pengfei Zhang,1and Hui Zhai1, 2\n1Institute for Advanced Study, Tsinghua University, Beijing, 100084, China\n2Collaborative Innovation Center of Quantum Matter, Beijing, 100084, China\n(Dated: September 18, 2018)\nIn this letter we consider quantum phases and the phase diagram of a Fermi Hubbard model under periodic\ndriving that has been realized in recent cold atom experiments, in particular, when the driving frequency is\nresonant with the interaction energy. Due to the resonant driving, the e ffective Hamiltonian contains a correlated\nhopping term where the density occupation strongly modifies the hopping strength. Focusing on half filling, in\naddition to the charge and spin density wave phases, large regions of ferromagnetic phase and phase separation\nare discovered in the weakly interacting regime. The mechanism of this ferromagnetism is attributed to the\ncorrelated hopping because the hopping strength within a ferromagnetic domain is normalized to a larger value\nthan the hopping strength across the domain. Thus, the kinetic energy favors a large ferromagnetic domain\nand consequently drives the system into a ferromagnetic phase. We note that this is a di fferent mechanism\nin contrast to the well-known Stoner mechanism for ferromagnetism where the ferromagnetism is driven by\ninteraction energy.\nRecently, with the help of quantum gas microscope for\nfermions [1–6], tremendous experimental progresses have\nbeen made on quantum simulation of the Fermi Hubbard\nmodel. These progresses include the observation of equilib-\nrium properties such as short-range antiferromagnetic corre-\nlations [7–9], hidden antiferromagnetic correlations [10], in-\ncommensurate spin correlations [11], canted antiferromag-\nnetic correlations [12] and pairing correlations [13] in sev-\neral di fferent circumstances. In particular, the antiferromag-\nnetic quasi-long-range order has been successfully observed\nthrough entropy engineering [14]. These progresses also in-\nclude the study of non-equilibrium transport behaviors such as\nthe measurement of optical conductivity [15], and the spin and\ncharge transport behavior in the strongly interacting regime\n[16, 17].\nStudying Fermi Hubbard model with cold atoms also al-\nlows us to open up new avenue beyond the traditional con-\ndensed matter paradigm. One of such examples is the period-\nically driven Fermi Hubbard model [18, 19]. Since the typ-\nical parameters of a Hubbard model is the hopping strength\nJand the on-site interaction U, both of which are of the or-\nder of electron volt in strongly correlated solid-state materi-\nals, it is therefore hard to drive a solid-state material with fre-\nquency resonant with any of these two energy scales. How-\never, in cold-atom optical lattice realization of the Fermi Hub-\nbard model, the typical energy scales for these two parameters\nare both of the order of thousand Hertz, and it is quite easy\nto drive the optical lattices with such a frequency. When the\ndriving frequency is resonant with the interaction parameter\nU, the driving can strongly modify the Fermi Hubbard model.\nAs observed in a recent experiment from the ETH group, the\nshort-range antiferromagnetic correlation can be reduced, or\nenhanced, or even switch sign to become ferromagnetic cor-\nrelation [20]. Similar experiment has also been performed by\ndriving the Hubbard model with two-photon Raman transition\n[21]. Hence, by combining such a resonant driving with the\nquantum gas microscope, it is very promising to study novel\nphysics induced by periodic driving that cannot be accessed\nJ\nUShaking:A,!\nAAACA3icbVDLSgNBEJz1GeNr1ZteBoPgQcKuCIqnqBePEc0DkhBmJ51kyOzsMtMrhiXgxV/x4kERr/6EN//GyeOgiQUNRVU33V1BLIVBz/t25uYXFpeWMyvZ1bX1jU13a7tsokRzKPFIRroaMANSKCihQAnVWAMLAwmVoHc19Cv3oI2I1B32Y2iErKNEW3CGVmq6u3WEB0xvu6wnVOecDujFEa1HIXRY0815eW8EOkv8CcmRCYpN96veingSgkIumTE134uxkTKNgksYZOuJgZjxHutAzVLFQjCNdPTDgB5YpUXbkbalkI7U3xMpC43ph4HtDBl2zbQ3FP/zagm2zxqpUHGCoPh4UTuRFCM6DIS2hAaOsm8J41rYWynvMs042tiyNgR/+uVZUj7O+17evznJFS4ncWTIHtknh8Qnp6RArkmRlAgnj+SZvJI358l5cd6dj3HrnDOZ2SF/4Hz+ANWqlvk=AAACA3icbVDLSgNBEJz1GeNr1ZteBoPgQcKuCIqnqBePEc0DkhBmJ51kyOzsMtMrhiXgxV/x4kERr/6EN//GyeOgiQUNRVU33V1BLIVBz/t25uYXFpeWMyvZ1bX1jU13a7tsokRzKPFIRroaMANSKCihQAnVWAMLAwmVoHc19Cv3oI2I1B32Y2iErKNEW3CGVmq6u3WEB0xvu6wnVOecDujFEa1HIXRY0815eW8EOkv8CcmRCYpN96veingSgkIumTE134uxkTKNgksYZOuJgZjxHutAzVLFQjCNdPTDgB5YpUXbkbalkI7U3xMpC43ph4HtDBl2zbQ3FP/zagm2zxqpUHGCoPh4UTuRFCM6DIS2hAaOsm8J41rYWynvMs042tiyNgR/+uVZUj7O+17evznJFS4ncWTIHtknh8Qnp6RArkmRlAgnj+SZvJI358l5cd6dj3HrnDOZ2SF/4Hz+ANWqlvk=AAACA3icbVDLSgNBEJz1GeNr1ZteBoPgQcKuCIqnqBePEc0DkhBmJ51kyOzsMtMrhiXgxV/x4kERr/6EN//GyeOgiQUNRVU33V1BLIVBz/t25uYXFpeWMyvZ1bX1jU13a7tsokRzKPFIRroaMANSKCihQAnVWAMLAwmVoHc19Cv3oI2I1B32Y2iErKNEW3CGVmq6u3WEB0xvu6wnVOecDujFEa1HIXRY0815eW8EOkv8CcmRCYpN96veingSgkIumTE134uxkTKNgksYZOuJgZjxHutAzVLFQjCNdPTDgB5YpUXbkbalkI7U3xMpC43ph4HtDBl2zbQ3FP/zagm2zxqpUHGCoPh4UTuRFCM6DIS2hAaOsm8J41rYWynvMs042tiyNgR/+uVZUj7O+17evznJFS4ncWTIHtknh8Qnp6RArkmRlAgnj+SZvJI358l5cd6dj3HrnDOZ2SF/4Hz+ANWqlvk=AAACA3icbVDLSgNBEJz1GeNr1ZteBoPgQcKuCIqnqBePEc0DkhBmJ51kyOzsMtMrhiXgxV/x4kERr/6EN//GyeOgiQUNRVU33V1BLIVBz/t25uYXFpeWMyvZ1bX1jU13a7tsokRzKPFIRroaMANSKCihQAnVWAMLAwmVoHc19Cv3oI2I1B32Y2iErKNEW3CGVmq6u3WEB0xvu6wnVOecDujFEa1HIXRY0815eW8EOkv8CcmRCYpN96veingSgkIumTE134uxkTKNgksYZOuJgZjxHutAzVLFQjCNdPTDgB5YpUXbkbalkI7U3xMpC43ph4HtDBl2zbQ3FP/zagm2zxqpUHGCoPh4UTuRFCM6DIS2hAaOsm8J41rYWynvMs042tiyNgR/+uVZUj7O+17evznJFS4ncWTIHtknh8Qnp6RArkmRlAgnj+SZvJI358l5cd6dj3HrnDOZ2SF/4Hz+ANWqlvk=Shaking:A,!\nAAACA3icbVDLSgNBEJz1GeNr1ZteBoPgQcKuCIqnqBePEc0DkhBmJ51kyOzsMtMrhiXgxV/x4kERr/6EN//GyeOgiQUNRVU33V1BLIVBz/t25uYXFpeWMyvZ1bX1jU13a7tsokRzKPFIRroaMANSKCihQAnVWAMLAwmVoHc19Cv3oI2I1B32Y2iErKNEW3CGVmq6u3WEB0xvu6wnVOecDujFEa1HIXRY0815eW8EOkv8CcmRCYpN96veingSgkIumTE134uxkTKNgksYZOuJgZjxHutAzVLFQjCNdPTDgB5YpUXbkbalkI7U3xMpC43ph4HtDBl2zbQ3FP/zagm2zxqpUHGCoPh4UTuRFCM6DIS2hAaOsm8J41rYWynvMs042tiyNgR/+uVZUj7O+17evznJFS4ncWTIHtknh8Qnp6RArkmRlAgnj+SZvJI358l5cd6dj3HrnDOZ2SF/4Hz+ANWqlvk=AAACA3icbVDLSgNBEJz1GeNr1ZteBoPgQcKuCIqnqBePEc0DkhBmJ51kyOzsMtMrhiXgxV/x4kERr/6EN//GyeOgiQUNRVU33V1BLIVBz/t25uYXFpeWMyvZ1bX1jU13a7tsokRzKPFIRroaMANSKCihQAnVWAMLAwmVoHc19Cv3oI2I1B32Y2iErKNEW3CGVmq6u3WEB0xvu6wnVOecDujFEa1HIXRY0815eW8EOkv8CcmRCYpN96veingSgkIumTE134uxkTKNgksYZOuJgZjxHutAzVLFQjCNdPTDgB5YpUXbkbalkI7U3xMpC43ph4HtDBl2zbQ3FP/zagm2zxqpUHGCoPh4UTuRFCM6DIS2hAaOsm8J41rYWynvMs042tiyNgR/+uVZUj7O+17evznJFS4ncWTIHtknh8Qnp6RArkmRlAgnj+SZvJI358l5cd6dj3HrnDOZ2SF/4Hz+ANWqlvk=AAACA3icbVDLSgNBEJz1GeNr1ZteBoPgQcKuCIqnqBePEc0DkhBmJ51kyOzsMtMrhiXgxV/x4kERr/6EN//GyeOgiQUNRVU33V1BLIVBz/t25uYXFpeWMyvZ1bX1jU13a7tsokRzKPFIRroaMANSKCihQAnVWAMLAwmVoHc19Cv3oI2I1B32Y2iErKNEW3CGVmq6u3WEB0xvu6wnVOecDujFEa1HIXRY0815eW8EOkv8CcmRCYpN96veingSgkIumTE134uxkTKNgksYZOuJgZjxHutAzVLFQjCNdPTDgB5YpUXbkbalkI7U3xMpC43ph4HtDBl2zbQ3FP/zagm2zxqpUHGCoPh4UTuRFCM6DIS2hAaOsm8J41rYWynvMs042tiyNgR/+uVZUj7O+17evznJFS4ncWTIHtknh8Qnp6RArkmRlAgnj+SZvJI358l5cd6dj3HrnDOZ2SF/4Hz+ANWqlvk=AAACA3icbVDLSgNBEJz1GeNr1ZteBoPgQcKuCIqnqBePEc0DkhBmJ51kyOzsMtMrhiXgxV/x4kERr/6EN//GyeOgiQUNRVU33V1BLIVBz/t25uYXFpeWMyvZ1bX1jU13a7tsokRzKPFIRroaMANSKCihQAnVWAMLAwmVoHc19Cv3oI2I1B32Y2iErKNEW3CGVmq6u3WEB0xvu6wnVOecDujFEa1HIXRY0815eW8EOkv8CcmRCYpN96veingSgkIumTE134uxkTKNgksYZOuJgZjxHutAzVLFQjCNdPTDgB5YpUXbkbalkI7U3xMpC43ph4HtDBl2zbQ3FP/zagm2zxqpUHGCoPh4UTuRFCM6DIS2hAaOsm8J41rYWynvMs042tiyNgR/+uVZUj7O+17evznJFS4ncWTIHtknh8Qnp6RArkmRlAgnj+SZvJI358l5cd6dj3HrnDOZ2SF/4Hz+ANWqlvk=FIG. 1: A schematic of the Fermi Hubbard model on a two-\ndimensional square lattice. Jdenotes the hopping strength. Uis the\non-site interaction. The system is time-periodically modulated by\nshaking the lattice in a sinusoidal way with frequency ωand ampli-\ntude A. Arrows indicate the direction of shaking. Balls with di fferent\ncolors and arrows indicate fermions with di fferent spins.\nin a static system. The goal of this letter is therefore to pre-\ndict quantum phases and phase diagram of the resonant driven\nFermi Hubbard model that is newly realized in cold atom ex-\nperiments.\nModel. We consider a two-dimensional square lattice un-\nder similar driving as realized in experiments [20, 28]. The\nlattice is periodically modulated along the ˆ xand ˆydirections\nwith a frequency ωand an amplitude A, whose single particle\nHamiltonian can be written as\nˆH0(t)=p2\n2m+ˆV(x+Acos(ωt),y+Acos(ωt)) (1)\nwhere mis the mass of an atom. We can now preform a unitary\ntransformation [22]\nˆU(t)=exp(ip·r0(t)), (2)arXiv:1808.03966v2 [cond-mat.quant-gas] 17 Sep 20182\nwhere r0(t)=−Acos(ωt)(1,1) are two-dimensional vectors.\nThis unitary transformation transfers position r0(t) into the\ncomoving frame where the lattice becomes static but an extra\ntime-dependent gauge field is introduce. The resulting Hamil-\ntonian is written as\nˆH0(t)=(p−A(t))2\n2m+ˆV(r), (3)\nwith A(t)=mr0(t). In principle, the Hamiltonian Eq. 3 is\nequivalent to the one Eq. 1 but it is more convenient for later\npurpose.\nNow we consider a single-band tight-binding model with\nthe nearest neighboring tunneling coe fficient Jand on-site\nHubbard interaction strength U. The Hamiltonian in a sec-\nond quantized form can be written by the Peierls substitution\nas\nˆH(t)=−J/summationdisplay\n/angbracketlefti,j/angbracketright\nσ=↑,↓eidij·A(t)ˆc†\niˆcj+U/summationdisplay\ni/parenleftBigg\nˆni↑−1\n2/parenrightBigg/parenleftBigg\nˆni↓−1\n2/parenrightBigg\n.(4)\nwhere ˆ ciσ(ˆc†\niσ) is the fermionic annihilation (creation) opera-\ntor on site iwith spinσ, ˆniσis the density operator on site i\nwith spinσ,/angbracketleft.../angbracketrightdenotes the nearest neighboring sites, and\ndij=di−djwithdibeing the position of the ith lattice site.\nThroughout this work we focus on the half-filling case and the\nchemical potential is set to zero.\nIf the modulation frequency ωis the largest energy scale\nof the problem, one can make a high-frequency expansion\nto obtain an e ffective time-independent Hamiltonian [23, 24].\nThe e ffective Hamiltonian takes the same form as the normal\nHubbard model and the only modification is that the tunnel-\ning coe fficient is renormalized by the oscillating gauge field\nas˜J=JB0(A), where we useBlto denote the lth Bessel\nfunction andA=mAωdis the normalized shaking amplitude\nhereinafter. dis the distance of two Wannier wave packets in\nthe nearest neighboring lattice sites.\nHowever, this expansion falls down when the modulation\nfrequencyω, or lth multiple of it, is comparable to one of\nthe energy scale of the problem, say, the Hubbard interaction\nstrength U. That is to say, l/planckover2pi1ω≈U, and we call it the lth\nresonance. Note that in this case, because U−l/planckover2pi1ωis a small\nenergy scale, we should apply another unitary transformation\nˆR(t)=exp(i/summationdisplay\njlωtˆnj↑ˆnj↓), (5)\nwhich alters the interaction strength to an e ffective one ˜U=\nU−l/planckover2pi1ω. Moreover, since ˆR(t) does not commute with the\nhopping term, it introduces an additional density dependence\nto the hopping term, and e ffectively it changes the gauge field\nto a spin and density dependent one as\n˜Ai j,σ(t)=A(t)−lωt\nd2dij((1−ˆni¯σ)ˆnj¯σ−(1−ˆnj¯σ)ˆni¯σ)).(6)\nNow the high frequency expansion can be safely applied, and\nto the lowest order it again results in a time-independent ef-fective Hamiltonian written as\nˆHeff=/summationdisplay\n/angbracketlefti,j/angbracketright,σ−ˆJ/angbracketlefti j/angbracketright\neff,σˆc†\niσˆcjσ+˜U/summationdisplay\ni/parenleftBigg\nˆni↑−1\n2/parenrightBigg/parenleftBigg\nˆni↓−1\n2/parenrightBigg\n.(7)\nHere ˆJ/angbracketlefti j/angbracketright\neff,σis defined as\nˆJ/angbracketlefti j/angbracketright\neff,σ=J0ˆai j¯σ+J1ˆbi j¯σ, (8)\nwhere ¯σdenotes the complement of σ,J0=JB0(A),J1=\nJBl(ηi jA) (ηi j=±1 for ( ix,iy)=(jx±1,jy) or ( ix,iy=\njx,jy±1)), and\nˆai jσ=(1−ˆniσ)(1−ˆnjσ)+ˆniσˆnjσ, (9)\nˆbi jσ=(−1)l(1−ˆniσ)ˆnjσ+ˆniσ(1−ˆnjσ). (10)\nIn above, the site dependence of J1is made implicitly. Note,\nhowever, that for even lthe Bessel function Blis an even func-\ntion, in which case ηi jcan be simply dropped and J1becomes\na constant. Compared to the o ff-resonance case, now the hop-\nping strength depends on site occupation of fermions. As we\nwill show below, this correlated hopping plays a key role in\nthe emergent new mechanism for ferromagnetism phase.\nSymmetry. Before we discuss how to solve this e ffective\nHamiltonian, let us first comment on the symmetry of this\nproblem. Note that the original Hubbard model possesses a\nSO(4) symmetry [25], which is composed of a spin SU(2),\ngenerated by ˆSz=(1/2)/summationtext\niˆc†\ni↑ˆci↑−ˆci↓ˆci↓,ˆS+=/summationtext\niˆc†\ni↑ˆci↓\nand ˆS−=ˆS†\n+, and a charge SU(2), generated by ˆLz=\n−(1/2)/summationtext\niˆc†\ni↑ˆci↑+ˆc†\ni↓ˆci↓+Ns/2,ˆL+=/summationtext\ni(−1)iˆci↑ˆci↓and ˆL−=\nˆL†\n+.Nsis the total number of sites. The spin SU(2) ensures\nthat the direction of spin-density-wave (SDW) order parame-\nter can be taken along any direction, while the charge SU(2)\nensures the degeneracy of a charge-density-wave (CDW) or-\nder and the fermion pairing order (P).\nIn the presence of periodic modulation, considering the\ntime-dependent Hamiltonian Eq. 4, it is straightforward to\nshow that the spin SU(2) symmetry stays, yet the charge\nSU(2) symmetry no longer holds because ˆLzdoes not com-\nmute with the/summationtext\ni,σfi(t)ˆniσterm. However, considering the\ntime-independent e ffective Hamiltonian Eq. 7, one can show\nthat the charge SU(2) symmetry is recovered for even lcase\nthough not for odd lcase [26]. Hereafter we focus only on the\neven lcase which possesses the same SO(4) symmetry as the\noriginal Hubbard model. In addition, the e ffective Hamilto-\nnian also possesses particle-hole symmetry at half-filling.\nPhase Diagram. We present our results on the phase dia-\ngram following from a standard mean-field treatment based on\nthis e ffective Hamiltonian Eq. 7, which is known to be quali-\ntatively reliable for a normal Hubbard model [26, 27]. Thanks\nto the SO(4) symmetry, we can choose SDW along ˆ zdirection\n(i.e. si=/angbracketleftˆni↑−ˆni↓/angbracketright) and CDW (i.e. ci=/angbracketleftˆni↑+ˆni↓/angbracketright−1) as the\norder parameters in our mean-field theory. Note that when we\nobtain the CDW order, it means that the system can have either\nCDW order or fermion pairing order, or an arbitrary combina-\ntion of them, as the order parameter of the degenerate ground3\n-� -� -� � � � �-�-���\n�/����/��\n���/� ������/�\n+����� FM1\n2\n3\nFIG. 2: Phase diagram for the e ffective Hamiltonian Eq. 7 of a\nresonantly driven Fermi Hubbard model with even lat half filling.\nThe phase diagram is controlled by two dimensionless parameters,\nJ1/J0and ˜U/J0. “CDW” and “SDW” denote charge and spin den-\nsity wave order. “P” denote fermion pairing order. “FM” denotes\nferromagnetism, “PS” denotes phase separation into high and low\ndensity regimes. Red lines denote the first order transition and the\nblue dashes lines denote the second order transition. Dashed arrows\nmark the lines along which the order parameters are plotted in FIG.\n3.\nstates. Higher order e ffect will break the degeneracy between\nCDW and fermion pairing order, but it is beyond the scope of\ncurrent work.\nThe phase diagram is shown in Fig. 2. Setting J0as the\nenergy unit, the phase diagram is controlled by two param-\neters of J1/J0and ˜U/J0, both of which can be easily tuned\nfrom positive to negative via control of ωandA. As breach-\nmark of our calculation, first of all, note that when J1/J0=1,\nbecause ˆ ai jσ+ˆbi jσ=ˆI, the kinetic energy term becomes\n−J0/summationtext\n/angbracketlefti j/angbracketright,σˆc†\niσˆcjσand the Hamiltonian recovers the usual Hub-\nbard model. In this case (labeled by 2 in Fig. 2), the result\nfor the normal Hubbard model is retrieved where we obtain a\nCDW order of ( π,π) with attractive interaction ( ˜U<0) and a\nSDW order of ( π,π) with repulsive interaction ( ˜U>0). Ex-\nplicitly, the order parameters are chosen as si=(−1)ix+iysand\nci=(−1)ix+iyc, and s(c) gradually vanishes as ˜Uapproaches\nzero from the positive (negative) side. As a result, a second or-\nder phase transition occur at ˜U=0 (gray dot in FIG. 2). This\ncan be seen from the order parameters plotted as a function of\n˜U, shown as green curves in Fig. 3.\nSince both CDW and SDW are ordered phases, the more\ngeneric situation should be either a first order transition or a\nphase co-existence regime in between. As marked by the red\nsolid lines in Fig. 2, the phase boundary of CDW and SDW\n��/��=�����/��=�����/��=-���\n-�-�-� �������������������\n�/������� ���������\nSolid: SDW\nDahsed: CDWFIG. 3: The CDW order parameter c(dashed lines) and the SDW\norder parameter s(solid lines) as a function of ˜Ufor three represen-\ntative cases labeled as 1-3 in Fig. 2, with J1/J0=1.5, 1.0 and−2.0,\nrespectively.\nat large|J1/J0|is a first order transition. The order parame-\nters are shown with orange lines in Fig. 3 for a representative\ncase (labeled by 1 in Fig. 2), where the CDW or SDW order\nparameter jumps from a finite value to zero at ˜U=0. At cer-\ntain regime of J1/J0, a CDW and SDW co-existence regime\nshows up in between as displayed by the yellow regime in Fig.\n2. The order parameters are shown with blue lines in Fig. 3\nfor a representative case (labeled by 3 in Fig. 2).\nThe most notable feature in Fig. 2 is the green region. In\nthis region a mean-field ansatz of CDW or SDW orders with\nordering vector at ( π,π) may not yield any ordered solution.\nHowever, when we consider the case of enlarged 2 ×2, 3×3,\nup to L×Ldomains, and within each domain the CDW and\nSDW order parameters are uniformly chosen as candswhile\nin its neighboring domain they are taken as −cand−s, the\nmean-field ansatz does yield ordered solutions. It can be seen\nfrom Fig. 4 which shows that the mean-field ground state en-\nergy decreases monotonically as Lincreases. It indicates that\nthe ground state will form large domains with opposite order\nparameter values. Moreover, minimizing ground state energy\nyields c=0 and s/nequal0 at positive ˜Uandc/nequal0 and s=0\nat negative ˜U. Hence, the system at positive ˜Upossesses\nspin order, and the increasing of the domain size means the\ndecreasing of the spin ordering wave vector. Eventually, the\nwave vector decreases toward zero, and the ground state be-\ncomes a ferromagnetic state. In another word, as the domain\nsize becomes larger and larger, the system is essentially made\nof ferromagnetic domains. For negative ˜Uthe system tends\nto phase separation with high density in one domain and low\ndensity in its neighboring domain. The transition between the\nferromagnetic phase to the SDW phase, as well as the transi-\ntion from phase separation to CDW, is a first order transition.\nIt should be emphasized that both the ferromagnetism and\nthe phase separation regime occur at small ˜U. In fact, it is\npurely due to the correlated hopping e ffect in the e ffective4\nFIG. 4: The mean-field ground state energy as a function of the do-\nmain size L. The insets show the configuration for 2 ×2, 3×3 and\n4×4 blocks where the CDW and SDW order parameters are uni-\nformly distributed within a domain and take opposite values between\nneighboring domains. For comparison, the solid line is the mean-\nfield energy when order parameters are all zero.\nHamiltonian Eq. 7, which originates essentially from the res-\nonant driving. Considering the mean-field configurations as\nshown in the insets of Fig. 4, let us look at the mean-field\nvalue of the e ffective hopping strength /angbracketleftˆJ/angbracketlefti j/angbracketright\neff,σ/angbracketrightwhich quanti-\nfies how the particle occupations a ffect the hopping strength\nand thus a ffect the bandwidth. We define Jintra\neff,σas/angbracketleftˆJ/angbracketlefti j/angbracketright\neff,σ/angbracketrightwith\nboth iand jin the same domain, and Jinter\neff,σas/angbracketleftˆJ/angbracketlefti j/angbracketright\neff,σ/angbracketrightwith i\nand jacross two neighboring domains. It is straightforward\nto write down both Jintra\neff,σandJinter\neff,σas\nJintra\neff,σ=1\n2/parenleftBig\nJ0[1+(c∓s)2]+J1[1−(c∓s)2]/parenrightBig\n, (11)\nJinter\neff,σ=1\n2/parenleftBig\nJ0[1−(c∓s)2]+J1[1+(c∓s)2]/parenrightBig\n, (12)\nwhere∓corresponds to di fferent spin component. One can\nshow that when|J1|<|J0|,|Jinter\neff,σ|is always smaller than |Jintra\neff,σ|.\nHence, the size of the domain tends to increase such that\nthere are more intra-domain links than inter-domain links, and\ntherefore the e ffective bandwidth on average becomes larger.\nFor a given filling, a larger bandwidth leads to more kinetic\nenergy gain.|J1|<|J0|is precisely the regime where we find\nferromagnetism or phase separation in the phase diagram of\nFig. 3 with arbitary weak interaction. This regime can be\neasily accessed when Ais small.\nAn alternative way to understand the emergence of this fer-\nromagnetism is to consider a uniform system with ˜U=0,\nwhere the Hamiltonian contains only the correlated tunneling\nterm. Substitute ˆJ/angbracketlefti j/angbracketright\neff,σby its mean-field value, it is straight-\nforward to compute the kinetic energy of this uniform sys-\ntem that depends on ni↑andni↓, where n↑=(n+sz)/2 and\nn↓=(n−sz)/2. We plot the kinetic energy as a function of\nnforsz=0 in Fig. 5(a) and as a function of szforn=1\nin Fig. 5(b) for two representative cases with J1/J0=−0.8\nand−1.8. One can see from Fig. 5(a) that for J1/J0=−0.8,\nthere are two local minima with one at n>1 and the other at\nn<1, who locate symmetrically on two sides of n=1, while\n��/��=-�����/��=-���\n��� ��� ��� ��� ���-���-���-���-���-���-���-������\n�������� �������(� ↑+�↓)��������� ������ℰ(�)/� �\n��� ��� ��� ��� ���-���-���-���-���-���-���-������\n�������� �������(� ↑+�↓)��������� ������ℰ(�)/� �FIG. 5: The mean-field energy with ˜U=0 (a) as a function of total\ndensity nwith sz=0 fixed, and (b) as a function of szwith n=1\nfixed. Solid line is for J1/J0=−0.8 and the dashed line is for J1/J0=\n−1.8.\nforJ1/J0=−1.8 there is only one minimum located at n=1.\nSimilarly, in Fig. 5(b) for J1/J0=−0.8, there are two lo-\ncal minima with one at positive szand the other at negative sz,\nsymmetrically distributed around sz=0, and for J1/J0=−1.8\nthere is only one minimum at sz=0. Thus, when the system\nis constrained with the average n=1 and sz=0, for the\ncase with J1/J0=−0.8, it will actually phase separate into\ndomains with either di fferent nor different sz, corresponding\nto phase separation and ferromagnetism, respectively. The\nchoice is made by the sign of ˜Uwhen a small but finite ˜U\nperturbation is turned on.\nConclusion and Outlook. The most significant finding of\nthis work is to provide an alternative mechanism for the on-\nset of ferromagnetism in the model with correlated hopping,\nwhich roots in the cooperation between the spin order and the\ncorrelated hopping. It is driven by the kinetic energy term, and\noccurs in the weakly interacting regime. This is in contrast to\nthe well-known Stoner ferromagnetism mechanism which is\ndriven by large interaction energy and occurs when the inter-\naction strength is beyond certain critical value. This is also\ndifferent from the ferromagnetism due to the super-exchange\nprocesses discussed in the experiment of Ref. [20] which re-\nquires ˜Uto be negative.\nFinally we shall comment on the experimental observation\nof this ferromagnetism. First of all, the system itself has been\nrealized with cold atoms, and moreover, a very recent experi-\nment shows that the heating is insignificant in the presence of\ndriving and the life time of the system can be about one sec-\nond [28]. Second, because this ferromagnetism is driven by\nthe kinetic energy, and as one can see from Fig. 5, the energy\ngain is of the order of bandwidth, thus one expects this fer-\nromagnetism to be observed when temperature is of the order\nof bandwidth, which can be accessed now by cold atom ex-\nperiments. Thirdly, the quantum gas microscope techniques\nmentioned at the beginning can be used to detect real space\nferromagnetic domains. Hence, it is quite promising to verify\nthis theory experimentally in very near future.\nAcknowledgment. This work is supported MOST un-\nder Grant No. 2016YFA0301600 and NSFC Grant No.5\n11734010.\n[1] E. Haller, J. Hudson, A. Kelly, D. A. Cotta, B. Peaudecerf, G.\nD. Bruce and S. Kuhr, Single-atom imaging of fermions in a\nquantum-gas microscope , Nature Physics 11, 738 (2015).\n[2] L. W. Cheuk, M. A. Nichols, M. Okan, T. Gersdorf, V . V . Ra-\nmasesh, W. S. Bakr, T. Lompe, and M. W. Zwierlein, Quantum-\nGas Microscope for Fermionic Atoms , Phys. Rev. Lett. 114,\n193001 (2015).\n[3] G. J. A. Edge, R. Anderson, D. Jervis, D. C. McKay, R. Day,\nS. Trotzky, and J. H. Thywissen, Imaging and addressing of\nindividual fermionic atoms in an optical lattice , Phys. Rev. A\n92, 063406 (2015).\n[4] A. Omran, M. Boll, T. A. Hilker, K. Kleinlein, G. Salomon, I.\nBloch, and C. Gross, Microscopic Observation of Pauli Block-\ning in Degenerate Fermionic Lattice Gases , Phys. Rev. Lett.\n115, 263001 (2015).\n[5] D. Greif, M. F. Parsons, A. Mazurenko, C. S. Chiu, S. Blatt,\nF. Huber, G. Ji and M. Greiner, Site-resolved imaging of a\nfermionic Mott insulator , Science 351, 953-957 (2016).\n[6] L. W. Cheuk, M. A. Nichols, K. R. Lawrence, M. Okan, H.\nZhang, and M. W. Zwierlein, Observation of 2D Fermionic\nMott Insulators of 40K with Single-Site Resolution , Phys. Rev.\nLett.116, 235301 (2016).\n[7] M. F. Parsons, A. Mazurenko, C. S. Chiu, G. Ji, D. Greif and\nM. Greiner, Site-resolved measurement of the spin-correlation\nfunction in the Fermi-Hubbard model , Science 353, 1253-1256\n(2016).\n[8] M. Boll, T. A. Hilker, G. Salomon, A. Omran, J. Nespolo, L.\nPollet, I. Bloch and C. Gross, Spin- and density-resolved mi-\ncroscopy of antiferromagnetic correlations in Fermi-Hubbard\nchains , Science 353, 1257-1260 (2016).\n[9] L. W. Cheuk, M. A. Nichols, K. R. Lawrence, M. Okan, H.\nZhang, E. Khatami, N. Trivedi, T. Paiva, M. Rigol and M.\nW. Zwierlein, Observation of spatial charge and spin correla-\ntions in the 2D Fermi-Hubbard model , Science 353, 1260-1264\n(2016).\n[10] T. A. Hilker, G. Salomon, F. Grusdt, A. Omran, M. Boll, E.\nDemler, I. Bloch, C. Gross, Revealing hidden antiferromagnetic\ncorrelations in doped Hubbard chains via string correlators ,\nScience 357, 484-487 (2017).\n[11] G. Salomon, J. Koepsell, J. Vijayan, T. A. Hilker, J. Nespolo,\nL. Pollet, I. Bloch, C. Gross, Direct observation of incommen-\nsurate magnetism in Hubbard chains , arXiv: 1803.08892.\n[12] P. T. Brown, D. Mitra, E. Guardado-Sanchez, P. Schaub, S. S.\nKondov, E. Khatami, T. Paiva, N. Trivedi, D. A. Huse and W. S.\nBakr, Spin-imbalance in a 2D Fermi-Hubbard system , Science\n357, 1385-1388 (2017).[13] D. Mitra, P. T. Brown, E. Guardado-Sanchez, S. S. Kondov,\nT. Devakul, D. A. Huse, P. Schaub and W. S. Bakr, Quantum\ngas microscopy of an attractive Fermi-Hubbard system , Nature\nPhysics 14, 173 (2018).\n[14] A. Mazurenko, C. S. Chiu, G. Ji, M. F. Parsons, M. Kanasz-\nNagy, R. Schmidt, F. Grusdt, E. Demler, D. Greif and M.\nGreiner, A cold-atom Fermi-Hubbard antiferromagnet , Nature\n545, 462-466 (2017).\n[15] R. Anderson, F. Wang, P. Xu, V . Venu, S. Trotzky, F. Chevy,\nand J. H. Thywissen, Optical conductivity of a quantum gas ,\narXiv:1712.09965.\n[16] M. A. Nichols, L. W. Cheuk, M. Okan, T. R. Hartke, E.\nMendez, T. Senthil, E. Khatami, H. Zhang, and M. W. Zwier-\nlein, Spin Transport in a Mott Insulator of Ultracold Fermions ,\narXiv:1802.10018.\n[17] P. T. Brown, D. Mitra, E. Guardado-Sanchez, R. Nourafkan, A.\nReymbaut, S. Bergeron, A.-M. S. Tremblay, J. Kokalj, D. A.\nHuse, P. Schaub, and W. S. Bakr, Bad metallic transport in a\ncold atom Fermi-Hubbard system , arXiv:1802.09456.\n[18] M. Bukov, M. Kolodrubetz and A. Polkovnikov, Schrie ffer-\nWolffTransformation for Periodically Driven Systems: Strongly\nCorrelated Systems with Artificial Gauge Fields , Phys. Rev.\nLett.116, 125301 (2016).\n[19] A.P. Itin and M.I. Katsnelson, Effective Hamiltonians for\nRapidly Driven Many-Body Lattice Systems: Induced Exchange\nInteractions and Density-Dependent Hoppings , Phys. Rev. Lett.\n115, 075301 (2015).\n[20] F. G ¨org, M. Messer, K. Sandholzer, G. Jotzu, R. Desbuquois\nand T. Esslinger, Enhancement and sign change of magnetic\ncorrelations in a driven quantum many-body system , Nature\n553, 481-485 (2018).\n[21] W. Xu, W. Morong, H.-Y . Hui, V . W. Scarola, B. De-\nMarco, Correlated Spin-Flip Tunneling in a Fermi Lattice Gas ,\narXiv:1711.2061.\n[22] R. Shankar, Principles of Quantum Mechanics, Springer 1994.\n[23] A. Eckardt and E. Anisimovas, High-frequency approximation\nfor periodically driven quantum systems from a Floquet-space\nperspective , New J. Phys 17, 093039 (2015).\n[24] A. Eckardt, Colloquium: Atomic quantum gases in periodically\ndriven optical lattices , Rev. Mod. Phys. 89, 011004 (2017).\n[25] C. N. Yang and S. C. Zhang, SO 4symmetry in a Hubbard\nmodel, Mod. Phys. Lett. B 4, 759 (1990).\n[26] Supplementary material for (i) prove the SO(4) symmetry of the\neffective Hamiltonian for even l, and (ii) describe the mean-field\ntheory.\n[27] N. Nagaosa, Quantum Field Theory in Condensed Matter\nPhysics. Springer 1999.\n[28] M. Messer, K. Sandholzer, F. Gorg, J. Minguzzi, R. Des-\nbuquois, T. Esslinger, Floquet dynamics in driven Fermi-\nHubbard systems , arXiv:1808.00506.Supplementary Material: Resonant Driving induced Ferromagnetism in the Fermi Hubbard Model\nNing Sun,1Pengfei Zhang,1and Hui Zhai1, 2\n1Institute for Advanced Study, Tsinghua University, Beijing, 100084, China\n2Collaborative Innovation Center of Quantum Matter, Beijing, 100084, China\n(Dated: September 18, 2018)\nIn this supplementary material, we first discuss the symmetry of the resonantly driven Fermi Hubbard model.\nThe SO(4) symmetry [1] of the e ffective Hamiltonian for even lis proved. Then we establish the mean-field\ntheory adopted in this work for even l, which is consistent with the text-book results [2] for J0=J1.\nSYMMETRY\nIn this section we dicuss the symmetry of the resonantly driven Fermi Hubbard model. The symmetry is of vital significance\nsince it greatly simplifies the formulation of the mean-field description.\nSO(4) symmetry.— A usual bipartite Fermi-Hubbard model, written as ˆH=−J/summationtext\n/angbracketlefti,j/angbracketright,σˆc†\niσˆcjσ+˜U/summationtext\ni(ˆni↑−1/2)(ˆni↓−1/2),\npossesses SO(4) symmetry [1]. The SO(4) symmetry is resolved into two SU(2) symmetries, the spin SU(2) and the charge\nSU(2). The generators of spin SU(2) are given by:\nˆSz=1\n2/summationdisplay\niˆc†\ni↑ˆci↑−ˆci↓ˆci↓,ˆS+=/summationdisplay\niˆc†\ni↑ˆci↓, (1)\nand\nˆS−=ˆS†\n+,ˆSx=ˆS++ˆS−\n2,ˆSy=ˆS+−ˆS−\n2i, (2)\nwho satisfy the commutation relation of the SU(2) algebra. Due to the contraction between spin indices, the very beginning\ntime-dependent Hamiltonian ˆH(t) (Eq.(4) in the main text) is invariant under this SU(2) symmetry operations, is also the unitary\ntransformations ˆR(t) (Eq.(5) and (6) in the main text). Since the time average does not alter this attribute, one arrives at the\nconclusion that the e ffective static Hamiltonian ˆHeffin Eq.(7) of the main text also possess this spin SU(2) symmetry no matter l\neven or odd. This can also be verified directly by checking/bracketleftBigˆHeff,ˆSα/bracketrightBig\n=0 for anyα=x,y,z. The spin rotational symmetry hence\nallows us to automatically get a spin-balanced system without any additional magnetic field.\nHowever, regarding the charge SU(2) symmetry, the time-dependent Hamiltonian lacks it for the appearance of the time-\ndependent onsite energy term (the last term in Eq.(4) of the main text). As a result we do not expect the SO(4) symmetry in\ngeneral. Nevertheless, it emerges in the e ffective Hamiltonian with leven. We introduce the charge SU(2) as follows [1].\nˆLz=−1\n2/summationdisplay\niˆc†\ni↑ˆci↑+ˆc†\ni↓ˆci↓+1\n2N, (3)\nˆL+=/summationdisplay\ni(−1)iˆci↑ˆci↓=/summationdisplay\niexp(iQ·xi)ˆci↑ˆci↓ (4)\nwhere Q=(π,π),Nis the total number of lattice sites hereinafter, and\nˆL−=ˆL†\n+,ˆLx=ˆL++ˆL−\n2,ˆLy=ˆL−−ˆL−\n2i. (5)\n{ˆLz,ˆLx,ˆLy}forms an SU(2) algebra. It can be verified straightforwardly that the e ffective Hamiltonian ˆHeffwith leven commutes\nwith all these generators. This fact can also be seen from the following transformation\nˆP: ˆci↑→(−1)iˆc†\ni↑,\nwho maps ˆHeffwith leven of interaction +˜Uto the same class but of e ffective interaction−˜U, exchanging the role played by {Li}\nand{Si}. As a result, the spin SU(2) invariance for a model with −˜Uindicates the charge SU(2) symmetry for +˜U. And vice\nversa. In addition, this transformation also implies the phase diagram should be symmetric under +˜U↔− ˜Utogether with the\ninterchange of charge and spin order. However, in the lodd cases, this mapping falls down because of the additional minus sign\ninˆbi jσ.arXiv:1808.03966v2 [cond-mat.quant-gas] 17 Sep 20182\nParticle-hole symmetry.— The half-filled e ffective Hamiltonian ˆHeffholds particle-hole symmetry no matter leven or odd.\n(i) When lis even, define particle-hole transformation ˆC: ˆciσ→(−1)iˆc†\niσ. The Hamiltonian is invariant under ˆC. [ˆC,ˆHeff]=0.\n(ii) When lis odd, we further define the bipartite transformation ˆS, which switches the A/Bsublattices. Then the Hamiltonian\nis invariant under the combination of ˆCand ˆS: [ˆCˆS,ˆHeff]=0. This symmetry allow us to fix chemical potential µ=0 during\nthe mean-field calculation at half-filling situation.\nMEAN-FIELD TREATMENT\nIn this section we establish the mean-field theory adopted in this work. As explained in the main text, assuming no canted\norder, we can consider only the charge density wave (CDW) order and spin density wave (SDW) order in the zdirection because\nof the SO(4) symmetry. We focus on the half-filled spin-balanced system, in which case the total charge density and total\nmagnetic momentum is given by\n/angbracketleftˆn/angbracketright=1,/angbracketleftˆS/angbracketright=0. (6)\nPath integral approach.— Here we provide a derivation of the mean-field Hamiltonian Eq. (19) via a path integral approach.\nIn the path integral language, the real-time partition function of the system is given by:\nZ=/integraldisplay\nDψD¯ψexp(i/integraldisplay\ndtL) (7)\nL=/summationdisplay\ni,σi¯ψiσ∂tψiσ+/summationdisplay\n/angbracketlefti,j/angbracketright,σ/parenleftBig\nJ0ai j¯σ(¯ψψ)+J1bl\ni j¯σ(¯ψψ)/parenrightBig¯ψiσψjσ\n−˜U/summationdisplay\ni¯ψi↑ψi↑¯ψi↓ψi↓. (8)\nHereψcorresponds to the fermionic field operators, and ai j¯σandbl\ni j¯σare given by replacing the operators in ˆ ai j¯σandˆbl\ni j¯σwith\nfields. Since the e ffective Hamiltonian (Eq. (7) in the main text) contains six-fermion terms, the traditional decoupling based on\nHubbard-Stratonovich transformation does not applied directly. However, an alternative way of decoupling based on the similar\nspirit of Hubbard-Stratonovich transformation should be adopted here. Specifically, by inserting a delta function, we directly\nintroduce the auxiliary bosonic field:\nZ=/integraldisplay\nDψD¯ψDn/productdisplay\niσδ(ni,σ−¯ψi,σψi,σ) exp( i/integraldisplay\ndtL) (9)\nL=/summationdisplay\ni,σi¯ψiσ∂tψiσ+/summationdisplay\n/angbracketlefti,j/angbracketright,σ/parenleftBig\nJ0ai j¯σ(n)+J1bl\ni j¯σ(n)/parenrightBig¯ψiσψjσ\n−˜U/summationdisplay\nini↑ni↓. (10)\nThanks to the delta function inserted, one could replace all ψ†ψbynin the action. As an equivalence check, if we integrate out\nnfields first, we get our original action back. Now we introduce another field ηto absorb the delta function into an integral:\nZ=/integraldisplay\nDψD¯ψDnDηexp(i/integraldisplay\ndtL) (11)\nL=/summationdisplay\ni,σi¯ψiσ∂tψiσ+/summationdisplay\n/angbracketlefti,j/angbracketright,σ/parenleftBig\nJ0ai j¯σ(n)+J1bl\ni j¯σ(n)/parenrightBig¯ψiσψjσ\n−˜U/summationdisplay\nini↑ni↓−/summationdisplay\niσηiσ(ni,σ−¯ψi,σψi,σ) (12)\nAs a result, the fermion becomes quadratic. In general, one can integrate out all the fermions to get an e ffective action of wholly\nbosonic degrees of freedom. Whereas, the mean-field approximation is to say that all the bosonic fields will be replaced by their\nsaddle point solutions.\nMean-field Hamiltonian.— By doing a Legendre transformation and replacing the fermion bilinears with its expectation\nvalues, we obtain the mean-field Hamiltonian\nˆH=/summationdisplay\n/angbracketlefti,j/angbracketright,σ−/parenleftBig\nJ0ai j¯σ(n)+J1bl\ni j¯σ(n)/parenrightBigˆψ†\niσˆψjσ+˜U/summationdisplay\nini↑ni↓+/summationdisplay\niσηiσ(ni,σ−ˆψ†\ni,σˆψi,σ) (13)3\nFIG. 1: Schematic of the two cases. Left panel: case 1. Right panel: typical examples of case 2 when L=2 and L=3, respectively, as labeled.\nLattice sites of the same color have the same local particle density ciand local spin density si.\nwhere the following replacement is carried out:\nˆai jσ=(1−ˆniσ)(1−ˆnjσ)+ˆniσˆnjσ,→(1−niσ)(1−njσ)+niσnjσ (14)\nˆbi jσ=(−1)l(1−ˆniσ)ˆnjσ+ˆniσ(1−ˆnjσ)→(−1)l(1−niσ)njσ+niσ(1−njσ) (15)\nFor the half-filled spin-balanced situation, we focus on the following two cases here. Notice that, in a half-filled spin-balanced\nsystem, the local particle density can be expressed in terms of local charge density and local spin density as\nni↑=1\n2(1+ci+si),ni↓=1\n2(1+ci−si) (16)\nwhere/summationtext\nici=0 and/summationtext\nisi=0 . The two cases are then:\n1.ci=(−1)ix+iyc,si=(−1)ix+iys\n2.ci=(−1)⌈ix/L⌉+⌈iy/L⌉c,si=(−1)⌈ix/L⌉+⌈iy/L⌉s\nwhere candsare CDW and SDW order parameters introduced, and ⌈x⌉denotes the ceiling function of x. Substituting these two\nexpressions into the original mean-field Hamiltonian Eq.(13) will yield the mean-field Hamiltonian in each case.\nCase 1. The first case is actually the ( π,π) order, where the lattice is naturally divided into A/Bsublattices and the unit cell\nof the lattice consists of 2 sites, with one from Aand the other from B. The schematic is shown in FIG. 1 (Left).\nTo be specific, we explicit show here the workflow in the first case. By substitution,\nnA≡/angbracketleftˆni∈A/angbracketright=1+c,sA≡/angbracketleftˆSz\ni∈A/angbracketright=s/2, (17)\nnB≡/angbracketleftˆni∈B/angbracketright=1−c,sB≡/angbracketleftˆSz\ni∈B/angbracketright=−s/2, (18)\nand the mean-field Hamiltonian is explicitly written as:\nˆHmf=/summationdisplay\n/angbracketlefti,j/angbracketright,σ{−J0[(1−nA¯σ)(1−nB¯σ)+nA¯σnB¯σ]−J1[(1−nA¯σ)nB¯σ+nA¯σ(1−nB¯σ)]}ˆcA†\niσˆcB\njσ+H. C. +˜UN\n2(nA↑nA↓+nB↑nB↓)\n+ηc/bracketleftBigg\nc−(ˆnA↑+ˆnA↓)−(ˆnB↑+ˆnB↓)\n2/bracketrightBigg\n+ηs/bracketleftBigg\ns−(ˆnA↑−ˆnA↓)−(ˆnB↑−ˆnB↓)\n2/bracketrightBigg\n(19)\nwhere ˆ cA\niσ(ˆcB\niσ) are the annihilation operators of isite and spin σonA(B) sublattice, ˆ nµσ=1\nN/2/summationtext\ni∈µˆniσ,µ=AorB, and\nnµσ=/angbracketleftˆnµσ/angbracketright. Hereηcandηsare Lagrangian multipliers of the CDW and SDW orders, respectively, who are also variational\nparameters to optimize the ground-state energy of ˆHmfto get a mean-field solution.\nAfter some substitution and simplification, the mean-filed Hamiltonian in momentum space is written as\nˆHmf=/summationdisplay\nkσ−Pσ(c,s)Q(k)ˆcA†\nkσˆcB\nkσ+H.c.+˜UN\n2(1+c2−s2)\n+ηc/bracketleftBigg\nc−(ˆnA↑+ˆnA↓)−(ˆnB↑+ˆnB↓)\n2/bracketrightBigg\n+ηs/bracketleftBigg\ns−(ˆnA↑−ˆnA↓)−(ˆnB↑−ˆnB↓)\n2/bracketrightBigg\n(20)4\nwhere ˆ cA\nkσ(ˆcB\nkσ) are the annihilation operators on A(B) sublattice of quasi-momentum kand spinσ,Q(k)=/summationtext\niexp(ik·di),{di}\nare the lattice vectors, and\nP↑(c,s)=J0\n2/bracketleftBig\n1−(c−s)2/bracketrightBig\n+J1\n2/bracketleftBig\n1+(c−s)2/bracketrightBig\n(21)\nP↓(c,s)=J0\n2/bracketleftBig\n1−(c+s)2/bracketrightBig\n+J1\n2/bracketleftBig\n1+(c+s)2/bracketrightBig\n(22)\nThe summation of kin Eq.(20) is over the first Brillioun zone. Minimization the energy for all parameters {c,s,ηc,ηs}yields a\nset of mean-field equations that is solved by numerical iteration in this work.\nCase 2. The second case actually includes a series of circumstances of Lbeing 2,3,4,...,∞. Typical examples of L=2\nandL=3 are shown in the right panel of FIG. 1. In this case, for each independent L, doing the similar substitution as above and\nsolving optimization problem in a numerical way returns us a set of self-consistent mean-field solutions in each L. The ground\nstate should be the one with the lowest ground state energy.\nFIG. 2, 3, 4, 5 in the main text are based on the mean-field numerics of above two cases.\nCheck the formulation for J 0=J1.— Finally, we explain why our mean-field theory reduces to a traditional one appearing\nin common text books, e.g. [2]. In fact, the Pσin this case is just a constant with no candsdependence. Thus the variation of c\nandsyields:\nηc=−˜UNcηs=−˜UNs. (23)\nUsing this relation, we see the mean-field Hamiltonian reduces to a familiar one that appears in textbooks [2].\n[1] C. N. Yang and S. C. Zhang, SO4symmetry in a Hubbard model , Mod. Phys. Lett. B 4, 759 (1990).\n[2] A. Altland, and B. D. Simons. Condensed matter field theory, second edition, Cambridge University Press, 2010." }, { "title": "0808.2358v1.Canted_Magnetization_Texture_in_Ferromagnetic_Tunnel_Junctions.pdf", "content": "arXiv:0808.2358v1 [cond-mat.str-el] 18 Aug 2008Canted Magnetization Texture in Ferromagnetic Tunnel Junc tions\nIgor Kuzmenko and Vladimir Fal’ko\nDepartment of Physics, Lancaster University, Lancaster, L A1 4YB, United Kingdom\n(Dated: November 15, 2018)\nWe study the formation of inhomogeneous magnetization text ure in the vicinity of a tunnel\njunctionbetweentwoferromagnetic wires nominallyinthea ntiparallel configurationanditsinfluence\non the magnetoresistance of such a device. The texture, depe ndent on magnetization rigidity and\ncrystalline anisotropy energy in the ferromagnet, appears upon an increase of ferromagnetic inter-\nwire coupling above a critical value and it varies with an ext ernal magnetic field.\nPACS numbers: 72.25.Mk, 75.47.De, 75.60.Ch\nSpin polarized transport in multilayer ferromagnet-\nmetal-ferromagnet system and magnetic tunnel junc-\ntion is a subject of intense theoretical and experimental\nstudies.1,2The majority of such studies addresses the in-\nvestigationofthetunnelingmagneto-resistance3,4,5(MR)\nand giant MR6,7,8,9,10effects, which consist of a switch\nfrom lower to higher conductivity when polarization of\nleads in a MR device change from an antiparallel to par-\nallel configuration. In tunnel junctions, the MR effect is\nthe result of a difference between rates of tunneling of\nelectrons majority and minority bands on the opposite\nsides of a junctions, and it’s the strongest when magne-\ntization switching by an external magnetic field changes\nfrom an ideally antiparallel magnetization state in the\ntwo wires to a parallel one.11,12,13,14Any deviations of\nthe magnetization near the interface (where tunneling\ncharacteristics of the device are formed) from perfectly\nparallel/antiparallel orientation would reduce the size of\nthe effect.\nIn this paper we investigate a possibility of formation\nof inhomogeneous magnetization texture in the vicinity\nof a highly transparent tunnel junction caused by fer-\nromagnetic coupling of magnetic moments on the op-\nposite sides carried by tunneling electrons. We find\nthat a canted magnetization state can form if such fer-\nromagnetic tunneling coupling, t′, exceeds some critical\nvaluetcdetermined by the interplay between crystallineanisotropy and magnetization rigidity in the ferromag-\nnet. This means that a tunnel junction with t′< tccan\nbe viewed as an atomically sharp magnetic domain wall,\nwhereas the increase of the junction transparency above\ntcgradually transforms it into a broad texture typical\nfor bulk ferromagnetic material. For t′> tc, we study\nthe evolution of the texture upon application of an ex-\nternal magneticfield and construct a parametricdiagram\nfor distinct magnetization regimes. As an example, we\nconsider a device consisting of a tunnel junction between\ntwo easy-axis ferromagnetic wires magnetically biased at\nthe ends as illustrated in Fig.1. We show that when the\nmagnetic field exceeds somecritical value, Bc, the barrier\nbetween the states with polarization parallel/antiparallel\nto the field disappears. As a result, both wires will be\npolarized parallel to the magnetic field and the domain\nwall will be pushed away toward the end of the wire.\nWhen a magnetic field is swept back (to B < B c) and its\nsign changes, the domain wall returns back to the tunnel\njunction. The resulting hysteresis in the magnetization\nstate of the device leads to a hysteresis loop in its MR,\nwhich we analyze taking into account the formation of\nthe texture near the tunnel junction with t′> tc.\nA quasi-1D magnetization texture, S(z), near the tun-\nneljunctionbetweentwowiresoflength L,whichchanges\nslowly on the scale of the lattice constant acan be de-\nscribed using the energy functional,\nE=J\n2/parenleftBigg−0/integraldisplay\n−L+L/integraldisplay\n+0/parenrightBigg\ndz/parenleftBig\n∂zS(z)/parenrightBig2\n−w2\na3L/integraldisplay\n−Ldz/bracketleftbigg\nξ1/parenleftBig\nSz(z)/parenrightBig2\n+ξ2/parenleftBig\nSy(z)/parenrightBig2/bracketrightbigg\n+\n+gdw3\n2a3L/integraldisplay\n−Ldzdz′\nw2V(z−z′)/braceleftbigg\nSx(z)Sx(z′)+Sy(z)Sy(z′)−2Sz(z)Sz(z′)/bracerightbigg\n−µBw2\na3L/integraldisplay\n−LdzSz(z)+E′.(1)\nHereJ∼tw2/ais the magnetization rigidity (where tis\nthe exchange coupling between the neighboring atoms, w\nis the wire thickness). The crystalline anisotropy param-\neters are ξ1> ξ2>0, whereas gd=γ2/a3parameterizesdipole-dipole interaction of magnetic moments, with\nV(z) =1\n2w/integraldisplay\nd2ρd2ρ′2z2−(x−x′)2−(y−y′)2\n/parenleftbig\n(x−x′)2+(y−y′)2+z2/parenrightbig5/2,\nwhere the integration is carriedout overthe cross-section\nofthewire, d2ρ=dxdy,d2ρ′=dx′dy′. For|z|> w,V(z)2\ndecreases as w3/z3, and for |z| ≪w,V(z)∝ln(w/|z|).\nFor a smooth magnetization texture varying at a length\nscale longer than the wire thickness, we approximate the\nnon-local dipole-dipole interaction as V(z)≈V0wδ(z),\nV0=/integraltextdz\nwV(z).Finally,µis the magnetic moment per\natom, and B=Bezis an external magnetic field.\n \neV\n2\n\u0001\u0000\u0002\n\u0003\u00042\n\u0005\u0006\u0007\n\b\t\n\n\u000b (b)(a)\n()0\f \r\n()0 \u000eS ()L \u000fS()0\u0010 \u0011\n()0 \u0012S ()\u0013 \u0014S\neV\n2\n\u0015\u0016\u0017\n\u0018\u00192\n\u001a\u001b\u001c\n\u001d\u001e\n\u001f \neV\n2\n!\"#\n$%2\n&'(\n)*\n+, (b)(a)\n()0- .\n()0 /S ()L 0S()01 2\n()0 3S ()4 5S()06 7\n()0 8S ()L 9S()0: ;\n()0 S\nFIG. 1: (a)Ferromagnetic wires. Polarizations of the wire\nendsz= 0 and z=±LareS(±0) andS(±L), respectively.\nThe angle between S(±0) and the axis zisθ(±0).(b)Band\nalignment of the majority and minority bands for electrons i n\nthe vicinity of the left and right-hand side of a biased ferro -\nmagnetic junction in the antiparallel configuration. Direc tion\nof spin quantization on either side of the junction is deter-\nmined by the orientation of S(±0) of magnetization near the\ninterface.\nThe inter-wire coupling energy, E′, includes the ex-\nchange interaction due to the penetration of the polar-\nized electron wave function through the barrier, from one\nferromagnetic metal into another. For the ferromagnetic\nsign of the inter-wire coupling,\nE′=−t′w2\na2/parenleftBig\nS(+0)·S(−0)/parenrightBig\n.\nIn the following, we will focus on the magnetiza-\ntion texture formed near the tunnel junction between\ntwo ferromagnetic metals with antiparallel polarization\nS(±L) =±ezfixed by magnetic reservoirs at the dis-\ntant ends of the wires. Without coupling between wires\n(t′= 0), this would determine homogeneous magnetiza-\ntionS(z >0) =ezandS(z <0) =−ez. The exchange\ninteraction between the wires may give rise to the for-\nmation of canted magnetization texture in the vicinity of\nthe junction, with boundary values of spin S(±0) deter-\nmined by the interplay between the magnetization rigid-\nity, crystalline anisotropy, Zeeman energy and the ex-\nchange inter-wire interaction. In the case when the easy\nmagnetization direction is along zaxis, we parameterize\nS(z) =ezcos(θ(z))+eysin(θ(z)), (2)\nwithθ(−L) =πandθ(L) = 0. Then, the total energy ofthe interacting wires takes the form\nE=J\n2/parenleftbigg−0/integraldisplay\n−L+L/integraldisplay\n+0/parenrightbigg\ndz/braceleftbigg\n(θ′(z))2+α2sin2(θ(z))+\n+2α2λB/bracketleftBig\n1−cos(θ(z))/bracketrightBig/bracerightbigg\n−t′cos(θ0),(3)\nwhereθ′(z) =dθ(z)/dz,θ0=θ(−0)−θ(+0). Relevant\nparameters present in Eq.(3) are\nα=/radicalbigg\nw2\na32(ξ1−ξ2)+3gdV0\nJ, λB=2µBw2\nJa3α2≡B\nBc,\nwhere\nBc=Ja3α2\n2µw2.\nThe meaning of the parameter Bcis the following. When\nB=Bc, the energy barrier between the states of\nnoninteracting wires ( t′= 0) with polarizations paral-\nlel/antiparallel to the external magnetic field disappears\nand the polarizationantiparallelto the magneticfield be-\ncomes absolutely unstable, leading to a jump of the mag-\nnetic domain wall from the junction toward the magnet-\nically biased end of the wires. Below we assume that Bc\nis much less than the field required to switch the polar-\nization of the left/right hand side leads. The parameters\nαandλBcharacterize a typical domain wall width in an\ninfinite wire. For example, α−1is the domain wall width\nforB= 0 for a long wire with Lα≫1. An external mag-\nnetic field B=Bezmakes the domain wall asymmetric.\nIt compresses the domain wall width on the side where\nmagnetizationisalignedwith Btoα−1\n+,α+=α√1+λB,\nand it stretches the domain wall width on the side where\nSis antiparallel to Btoα−1\n−,α−=α√1−λB.\nTo minimize the energy in Eq. (3), we employ the\nfollowing procedure. First, we fix the boundary values\nθ(±0) and find the optimal form of θ(z) for given θ(±0).\nThen we minimize E(θ(+0),θ(−0)) versus the canting\nanglesθ(±0). The first step of such a procedure requires\nsolving the optimum equation\nθ′′(z) =α2\n2/bracketleftBig\nsin(2θ(z))+2λBsin(θ(z))/bracketrightBig\n.(4)\nThe latter shows that that θ(z >0) [π−θ(z <0)] takes\nits maximal value θ(+0) [π−θ(−0)] atz= 0, and that it\ndecreases as exp( −α+z) [exp(α−z)] for|z| ≫α−1. Also\nthe differential equation (4) has the first integral v±(θ) =\n−θ′(z),\nv±(θ) =α/radicalbig\n(1±λB)2−(cos(θ)+λB)2,(5)\nwhere the sign ±corresponds to positive/negative z.\nSubstituting v±into energy in Eq. (3) and changing the\nintegration variable from ztoθwe arrive at\nE=J/parenleftBiggθ(+0)/integraldisplay\n0v+(θ)dθ+π/integraldisplay\nθ(−0)v−(θ)dθ/parenrightBigg\n−\n−t′cos(θ0), (6)3\nwhich represents the explicit dependence of energy\nE(θ(+0),θ(−0)) on the cantingangles on eachside ofthe\njunction. Minimizingtheenergywithrespecttothesean-\ngles, we identify possible regimes for the magnetization\ntexture.\nFirst, we consider the magnetization texture in the\nabsence of an external magnetic field, B= 0. In this\ncase the texture is symmetric, θ(−z) =π−θ(z), so that\nv+(θ) =v−(θ) =αsin(θ) and\nE= 2Jα(1−cos(θ(+0)))+ t′cos(2θ(+0)).\nThe latter result indicates the existence ofa critical value\ntc=Jα/2 of the coupling constant t′, such that for t′<\ntcthe energy reaches its minimum when θ(+0) =π−\nθ(−0) = 0 and magnetization is homogeneous in each\nof two wires, whereas for t′> tcthe energy minimum\ncorresponds to the magnetization texture in the vicinity\nof the tunnel junction, with\nθ(+0) =π−θ(−0) = arccos( tc/t′).\nIn the presence of external magnetic field, the magne-\ntization texture becomes asymmetric. In this case, we\ndetermine canting angles θ(±0) numerically from the set\nof two equations,\n∂θ(±0)E=Jv±(θ(±0))−t′sin(θ(−0)−θ(+0)) = 0 .\n(7)\nThe results of numerical analysis of Eqs.(7) are gath-\nered in the parametric diagram in Fig. 2, where we indi-\ncate the six parametric intervals separated by lines B1,\nB′\n1,t1, and the axis B= 0 corresponding to different\nregimesofmagnetizationtexture. Below, wedescribe the\nevolution of the magnetization state of the junction upon\nsweeping the magnetic field from −Bcto +Bc(in com-\nparison to an inverse sweep from + Bcto−Bc). When\nthe exchange coupling constant t′and the magnetic field\nBare not strong enough (parametric interval I(I′) in\nFig.2), the energy functional has two minima. The first\nof them corresponds to θ(+0) =π−θ(−0) = 0 and the\nmagnetization is homogeneous in each of two wires. The\nsecondcorrespondsto the statewith both wirespolarized\nparallel to the external magnetic field and the domain\nwall pushed away towards the end z=−L(z=L) of\nthe wire. As a result, the polarizationsof the wires in the\ntunnelingareaareantiparallel/parallelonetoanotherde-\npending on whether we increase the magnetic field from\nB= 0 or decrease the field from B1(decrease the field\nfromB= 0 or increase the field from B′\n1). When one\ndecreases the magnetic field from B1(increases from B′\n1)\nso that the field changes its sign, the state with homo-\ngeneous polarization of both wires and the domain wall\nplaced at the end z=−L(z=L) of the wire becomes\nunstable, and the domain wall is pushed towardsthe tun-\nneling area z= 0.\nChanging t′and/orBandcrossingthe line t1resultsin\ncontinuous variation of the canting angles from θ(+0) =\nπ−θ(−0) = 0 on the left from the line t1to finite values\non the right from the line t1(parametric interval II(II′)inFig.2)andformationofthe magnetizationtexturenear\nthe tunnel junction. The values of the canting angles are\nfound from equations (7). The second state corresponds\ntothesituationwhenboththewiresarepolarizedparallel\nto the external magnetic field and the domain wall is\npushedawaytowardstheend z=−L(z=L)ofthewire.\nAt the line B1(B′\n1) the barrier between metastatic and\nground states disappears and there is just one minimum\nof the energy functional corresponding to the state with\nboth wires polarized parallel to the external magnetic\nfield and the domain wall pushed away towards the end\nz=−L(z=L) of the wire. As a result, the polarization\nin the vicinity ofthe tunnel junction is homogeneous,i.e.,\nθ(−0) =θ(+0) = 0 ( θ(−0) =θ(+0) =π). \n1II\nII′cB B\n05 . 0\n5 . 0−1\n1−5 . 0 5 . 1 2ct t′I\nI′III\nIII′X\n'X1B\n1B\n1B′1B′1t\nFIG. 2: Parametric diagram t′-B. There are four lines, B1,\nB′\n1,t1, andtheaxis B= 0separatingphases onefrom another\n(see the text for details). The lines B1andt1(B′\n1=−B1and\nt1) touch one another at the point X(X′). The boundary\nvaluesθ(±0) as a function of the external magnetic field are\ncalculated for t′= 0.5tc, 0.95tcand 1.2tc(dashed lines).\nTypical dependencies of θ(+0) on a magnetic field B\nare illustrated in Fig. 3 for t′= 0.5tc,t′= 0.95tc, and\nt′= 1.2tc, respectively. The dependencies of θ(−0) onB\ncan be obtained from Fig. 3 by the transformation\nθ(−0,B) =π−θ(+0,−B).\nThe formation of canted magnetization in the vicinity\nof a tunnel junction with high transparency would affect\nthe magnetoresistance characteristic of such a junction.\nThe latter can be modelled using the tunnel Hamilto-\nnian approach. When the domain wall width α−1is\nsufficiently larger then the elastic mean free path, the\ntunneling Hamiltonian can be written as\nH=H0+HT, H 0=/summationdisplay\nν=±/summationdisplay\nk,σǫσ(k)c†\nνkσcνkσ,\nwhereǫσ(k) =ǫ(k)−σ∆. Here H0is the Hamiltonian\nof the isolated ferromagnetic wires, HTis the tunneling\nHamiltonian, cνkσandc†\nνkσareannihilation and creation\noperators of electron propagating in the left ( ν=L) or\nright (ν=R) wire with wave vector kand spin parallel\n(σ= 1/2 or↑) or antiparallel ( σ=−1/2or↓) to the wire\npolarization S(±0) (2). In the following we will assume\nthat the Fermi momenta for the majority and minority4\n \n-1.2 -0.8 -0.4 0 0.4 0.8 1.2()0+θ\ncB Bπ\n2π\n-0.8 -0.4 0 0.4 0.8()0+θ\ncB Bπ\n2π\n-0.8 -0.4 0 0.4 0.8()0+θ\ncB Bπ\n2π()a\n()b\n()c\nFIG. 3: Canting angle θ(+0) as functions of λBfort′= 0.5tc\n(panel (a)), t′= 0.95tc(panel (b)), and t′= 1.2tc(panel (c)).\nbands are sufficiently larger than α±and therefore treat\nconduction band electrons as three dimensional.\nHT=/summationdisplay\nσ,σ′/summationdisplay\nk,k′/bracketleftbigg\ntσ′σ(k′,k)c†\nRk′σ′cLkσ+h.c./bracketrightbigg\n,\nwhere the tunneling matrix elements tσ′σ(k′,k) describe\nthe transfer of an electron with wave vector kand spin\nstateσfromtheleftwiretothestatewith k′andσ′inthe\nright wire and the quantization axes for the conduction\nband electrons are directed along the magnetization vec-\ntorsS(±0) (2) which are not necessarily collinear so that\nthe transitions between the majority/minority bands of\none wire and majority/minority bands of the other are\npossible even for the nominally antiparallel configuration\nofS(±L). We consider the model in which electrons tun-\nnelfromonewiretoanotherwithoutspinflipping. Inthis\ncasethe spin dependence of tσ′σ(k′,k) is determined by a\nsingle parameter, the angle θ0=θ(−0)−θ(+0) between\nthe vectors S(±0),\ntσσ′(k′,k) =˜t/vextendsingle/vextendsingle/vextendsingle/vextendsingle2π2/planckover2pi12vz(k)vz(k′)\nL2/vextendsingle/vextendsingle/vextendsingle/vextendsingle1/2\n×\n×/bracketleftBig\ncos(θ0/2)δσσ′+isin(θ0/2)τx\nσσ′/bracketrightBig\n,\nwhereτxis the Pauli matrix, and vz(k) is a component\nof electron velocity v(k) =∇kǫσ(k)//planckover2pi1perpendicular to\nthe interface.\nWhen vectors S(+0) and S(−0) are antiparallel (para-\nmetric intervals IandI′in Fig.2), electrons can tunnelonly from the majority band of one wire to the minor-\nity band of the other,15so that conductance of such a\njunction is\nG↑↓=2πe2|˜t|2\n/planckover2pi1/parenleftbigg2π/planckover2pi1\nL/parenrightbigg2\nN↑vz\n↑N↓vz\n↓,\nwhereN↑andN↓are the densities of states in the major-\nity and minority bands at the Fermi level, and vz\n↑/vz\n↓are\nthe average value of |vz(k)|over the majority/minority\nFermi surface,\nvz\nσ=1\nNσ/summationdisplay\nk/vextendsingle/vextendsingle/vextendsinglevz(k)/vextendsingle/vextendsingle/vextendsingleδ/parenleftBig\nǫ−ǫ(k)/parenrightBig\n.\nWhen the ends z=±0 of the wires have parallel magne-\ntization, i.e., θ(±0) = 0 or π, the conductance is\nG↑↑=2πe2|˜t|2\n/planckover2pi1/parenleftbigg2π/planckover2pi1\nL/parenrightbigg2(N↑vz\n↑)2+(N↓vz\n↓)2\n2,\nwhich determines the magneto-resistance (MR) ratio\nδ0=G↑↑−G↑↓\nG↑↑.\n \n()a\n()b\n()c\n-1 -0.5 0 0.5 1G\n↑↑G\n↑↓G\ncB B\n-1 -0.5 0 0.5 1G\n↑↑G\n↑↓G\ncB B-1 -0.5 0 0.5 1cB BG\n↑↑G\n↑↓G\n-1 -0.5 0 0.5 1G\n↑↑G\n↑↓G\ncB B-1.5 -1 -0.5 0 0.5 1 1.5cB BG\n↑↑G\n↑↓G\n-1.5 -1 -0.5 0 0.5 1 1.5cB BG\n↑↑G\n↑↓G\nFIG. 4: Magnetoresistance for a positive (left) and negativ e\n(right)magnetic fieldsweep, showingtheeffectofcantedmag -\nnetization texture on the form of hysteresis in the magnetic\ntunnel junction, for: (a)t′= 0.5tc,(b)t′= 0.95tc, and(c)\nt′= 1.2tc.\nTheformationofacantedmagnetizationtextureinthe\nvicinity of a junction with angle θ0between magnetiza-\ntion directions of the opposite sides of the tunnel barrier5\nproduces conductance G(θ0) and the reduced observable\nMR ratio, δ,\nG(θ0) =G↑↑cos2(θ0/2)+G↑↓sin2(θ0/2),\nδ=G↑↑−G(θ0)\nG↑↑=δ0sin2(θ0/2). (8)\nThe results of calculations for the conductance in MR\ndevices with various junction transparencies (and pos-\nitive and negative magnetic field sweep in the magne-\ntoresistance) are gathered in Fig.4. The conductance\ndip at small magnetic fields indicates the regime when\na ferromagnetic domain wall in the device in Fig.1 is\npinned to the tunnel junction. The detailed structure\nof the hysteresis loop depends on whether the magneti-\nzation texture is formed near the junction or not which\nofcoursedepends onthe value ofthe ferromagneticinter-\nwire coupling. For small inter-wire coupling, in Fig.4(a),the jump between parallel and antiparallel polarizations\nin the vicinity of the tunnel junction give rise to jumps in\nthe conductance between G↑↑and flat conductance min-\nimum equal to G↑↓. With increasing the inter-wire cou-\npling towards the critical value tc(determined by the in-\nterplaybetweencrystallineanisotropyandmagnetization\nrigidityin ferromagnet),anintervalofmagneticfieldsap-\npears where the conductance (8) increases continuously,\ndue to the formation of the canted magnetization tex-\nture and its change by a magnetic field, Fig.4(b). For\ninter-wire coupling above a critical value tc, Fig.4(c), the\nminimum of the conductance exceeds G↑↓even in the\nabsence of a magnetic field, indicating the formation of\na broad texture, which also suggests a reduction of MR\nratio in a high-transparency junction.\nThe authors thank to G. Bauer and E. McCann for\nuseful discussions. This project has been funded by EU\nSTREP Dynamax and ESFCRP “Spin current”.\n1G.A. Prinz, Phys. Today 48(4), 58 (1995); Science 282,\n1660 (1998), and references therein.\n2A. Brataas, G.E.W. Bauer, P.J. Kelly, Physics Reports,\n427, 157 (2006).\n3M. Julliere, Phys. Lett. 54A, 225 (1975).\n4J.S. Moodera, T.H. Kim, C. Tanaka, and C.H. de Groot,\nPhilos. Mag. B 80, 195 (2000).\n5G. Usaj and H. U. Baranger, Phys. Rev. B 63, 184418\n(2001).\n6A. Gerber, A. Milner, B. Groisman, M. Karpovsky, A.\nGladkikh, and A. Sulpice, Phys. Rev. B 55, 6446 (1997).\n7M.N. Baibich, J.M. Broto, A. Fert, F. Nguyen Van Dau,\nF. Petroff, P. Eitenne, G. Creuzet, A. Friederich, and J.\nChazelas, Phys. Rev. Lett. 61, 2472 (1988).\n8G. Binasch, P. Gr¨ unberg, F. Saurenbach, and W. Zinn,\nPhys. Rev. B 39, 4828(R) (1989).\n9W. P. Pratt Jr., S. F. Lee, J. M. Slaughter, R. Loloee, P. A.\nSchroeder, and J. Bass, Phys. Rev. Lett. 66, 3060 (1991).\nS. F. Lee, W. P. Pratt Jr., R. Loloee, P. A. Schroeder, andJ. Bass, Phys. Rev. B 46, 548 (1992).\n10Y. Sakuraba, M. Hattori, M. Oogane, Y. Ando, H. Kato,\nA.Sakuma, T. Miyazaki, andH.Kubota, Appl.Phys.Lett.\n88, 192508 (2006).\n11J. S. Moodera, L.R. Kinder, T.M. Wong, and R. Meservey,\nPhys. Rev. Lett. 74, 3273 (1995).\n12I.ˇZuti´ c, J. Fabian, and S. Das Sarma, Rev. Mod. Phys.\n76, 323 (2004).\n13J. Mathon and A. Umerski, Phys. Rev. B 63, 220403\n(2001).\n14Y. Li, B.Z. Li, W.S. Zhang, and D.S. Dai, Phys. Rev. B\n57, 1079 (1998)\n15E. McCann and V.I. Fal’ko, Phys. Rev. B 66, 134424\n(2002).\n16J. Shi, K. Pettit, E. Kita, S.S.P. Parkin, R. Nakatini, and\nM.B. Salamon, Phys. Rev. B 54, 15273 (1996).\n17E.Yu. Tsymbal, D.G. Pettifor, J. Shi, and M.B. Salamon,\nPhys. Rev. B 59, 8371 (1999)." }, { "title": "1108.1294v1.Ferromagnetic_spin_coupling_of_2p_impurities_in_band_insulators_stabilized_by_intersite_Coulomb_interaction__Nitrogen_doped_MgO.pdf", "content": "arXiv:1108.1294v1 [cond-mat.mtrl-sci] 5 Aug 2011Ferromagnetic spin coupling of 2 p-impurities in band insulators stabilized by intersite\nCoulomb interaction: Nitrogen-doped MgO\nI. Slipukhina,1,∗Ph. Mavropoulos,1S. Bl¨ ugel,1and M. Leˇ zai´ c1\n1Peter Gr¨ unberg Institut and Institute for Advanced Simula tion,\nForschungszentrum J¨ ulich and JARA, D-52425 J¨ ulich, Germ any\n(Dated: September 15, 2018)\nFor a nitrogen dimer in insulating MgO, a ferromagnetic coup ling between spin-polarized 2 p-holes\nis revealed by calculations based on the density functional theory amended by an on-site Coulomb\ninteraction and corroborated by the Hubbard model. It is sho wn that the ferromagnetic coupling is\nfacilitated by a T-shaped orbital arrangement of the 2 p-holes, which is in its turn controlled by an\nintersite Coulomb interaction due to the directionality of thep-orbitals. We thus conjecture that\nthis interaction is an important ingredient of ferromagnet ism in band insulators with 2 pdopants.\nPACS numbers: 75.50.Pp, 75.30.Hx, 71.70.-d, 71.15.Mb\nFerromagnetic (FM) insulators offer the potential for\nuse as active barrier materials in novel spintronic tun-\nneling devices. In the past years we have witnessed a\nnew vista in engineering FM insulators not by doping\nnonmagnetic insulators with transition-metal ions as is\ntraditionally accomplished in diluted magnetic semicon-\nductors [1, 2], but by doping with sp-elements. The\nnovel magnetic materials design was encouraged by un-\nexpected experimentalobservationsofroom-temperature\nferromagnetism in defective wide-gap oxide semiconduc-\ntors and insulators [3–7]. This phenomenon (referred to\nassp- ord0-magnetism)wasassociatedwith the partially\nfilledp-statesof the intrinsic defects like cation/anionva-\ncancies [3, 6, 8] or first-row (2 p) dopants [9].\nShortly after, numerous cation-deficient or N- and C-\ndoped oxides and sulfides were theoretically predicted\nto be FM at room temperature [10–14]. Among those,\nMgO is distinct as it is certainly the most important bar-\nrier material for magnetic tunnel junctions. Recently, N-\ndoped MgO films were experimentally shown to exhibit\nferromagnetic properties upon thermal annealing [15].\nIt is commonly concluded that the FM interaction be-\ntween the defects in d0-magnets is due to partially oc-\ncupied spin-polarized defect states, that are sufficiently\nextended to provide an exchange interaction via the\ndouble-exchange mechanism. Most likely, however, at\nlow concentrations the impurity 2 pelectrons experience\na strong on-site Coulomb repulsion Ubecause of their\nspatial localization [16], leading to insulating behavior\nthat changes the mechanism of exchange interaction and\ncould weaken or change the sign of magnetic coupling.\nCorrelation effects in defect-free oxides with partially\nfilled oxygen p-shells were studied in Refs. [17–19].\nThe importance of electron correlations in the impu-\nrityp-states was recently investigated using density func-\ntional theory (DFT) approaches amended by an on-site\nCoulomb repulsion [11, 20] for the example of N-doped\nMgO.Jahn-Teller-like(JT)distortions,capturedby’+ U’\n[21]orself-interactioncorrection[22]schemes,werefound\ntoevokeanenergysplittingbetweentheoccupiedandun-occupied nitrogen2 p-states, increasingthe localizationof\na spin-polarized hole at one of the 2 p-orbitals by pushing\nit deeper into the band gap and thus breaking the initial\nsymmetry of the electronic state.\nIn this Letter we present new insights into the role\nof electron correlations in N-doped MgO by carrying\nout DFT calculations where strong correlations are ac-\ncounted for by the GGA (generalized-gradient approx-\nimation [23])+ Uapproach, that are conceptualized on\nthe basis of a minimal Hubbard model. We demonstrate\nthat the symmetry breaking and the subsequent metal-\ninsulator transition occurs even without JT distortions,\nand is thus electronically driven. More importantly we\nfind for the insulating state of N-dimers, which are likely\ntobeformedintheannealingprocess,aFMspincoupling\nandaT-shaped2 p-orbitalarrangement(OA). Employing\na Hubbard model we show that a weak intersite Coulomb\nrepulsion, that is well accounted for by the DFT, com-\nbined with the on-site Coulomb interaction is responsi-\nble for this result. While in periodicstrongly correlated\n3dsystems the favored spin- and orbital arrangement is\nusually well explained by the Kugel-Khomskii model [24]\nwhich does not resort to the intersite repulsion, similar\narguments applied to N-N dimers in MgO yield an anti-\nferromagnetic(AFM) coupling, in contradiction with our\nfirst-principlesresults. The nearest-neighborFMinterac-\ntion is not sufficient to explain the long-range FM order,\nbut it is a necessary condition. Our aim is to shed some\nlight on the physics driving this interaction.\nWe utilized the full-potential linearized augmented\nplane wave code FLEUR[25] in our GGA+ Ucalculations.\nRelaxations were performed with an on-site Coulomb en-\nergyU=4.6 eV and Hund’s exchange J=1.2 eV applied\nonthe 2p-statesofOand N. The totalenergieswerecom-\npared[26] for a set of U-values that rangesbetween 3 and\n6 eV, reported for O 2 p-states in transition-metal oxides\nfrom photoemission and Auger experiments [27], and for\ntwovaluesof J, 0.6and1.2eV(fromHund’sexchangefor\natomic N and O [26], we expect the physically relevant\nJ-values in a solid to fall within this range). Supercells2\nof 64 atoms of host MgO, with a Brillouin-zone 2 ×2×2\nk-mesh are used to study a single N impurity and N-N\ndimers at O sites. N-O-N “dimers” are treated in a 144-\natomic supercell and sampled at the Γ-point. Different\nOAs are introduced by initiating a specific occupation of\nthep-orbitals of N in the GGA+ Udensity-matrix [28].\nCalculations are performed for all possible types of OA\nat FM and AFM spin alignment (12 different spin-orbital\nconfigurations in total) for N-N and N-O-N dimers.\nWe firstconsider the caseof asingle Nimpurity substi-\ntuting O in MgO to check whether the symmetry break-\ning and the metal-insulator transition upon applying a\nfiniteUis driven electronically or by local lattice distor-\ntions. Figure 1(a) shows the N density of states (DOS) of\nthe structurally unrelaxed supercell. Within the GGA,\nN introduces two triply degenerate states, an occupied\nspin-upp-state in the valence band and a partially unoc-\ncupied spin-down state at the Fermi level, EF. A single\nspin-polarized hole is evenly distributed among the three\nNp-orbitals. The situation changes drastically upon ap-\nplyingU: the minority 2 p-state splits into a doubly de-\ngenerateoccupied and a non-degenerateunoccupied level\n(Fig.1(b)). This symmetry breaking occurs even without\nlattice distortions, which clearly demonstrates that it is\nan electronically driven effect. Similar behavior was no-\nticed for the cation vacancy in ZnO [29]. This is different\nfromthepreviousreportsonMgO:N[20], wherethemain\neffect of both + Uor self-interaction corrections resulted\nin symmetry breaking via a JT distortion.\nNext, we examine the relative spin and p-orbital ar-\nrangement for two substitutional N impurities in MgO.\nRecent LSDA studies [30] have shown that N-N impurity\npairing in a structurally relaxed MgO supercell leads to a\nnon-magnetic insulating ground state with a fully occu-\npied bonding ppσ(as well as ppπ) and completely empty\nantibonding ppσ∗states. However, we show that this is\nno longer the case in the presence of strong correlations.\nWe consider the following relative positions of the two N\natoms: (i) at the nearest neighbor (n.n.) sites, at a dis-\ntancedN-N= 2.97˚A – a N-N dimer; (ii) connected via an\nO atom ( dN-N= 5.83˚A ) – a N-O-N dimer, and model,\ncorrespondingly,the1stand2ndn.n.interactionbetween\nN atoms in the oxygen sublattice of MgO. We omit the\nN-Mg-N configuration from the analysis because we find\nthe magnetic interactions to be negligibly small [31].\nWe define a coordinate system with the pxandpyor-\nbitals pointing toward neighboring N (or O) atoms in\nthexyplane (Fig.1(c)), while the pzorbital points out of\nplane, toward Mg. In the considered cases of N-N and N-\nO-N dimers, both N atoms are situated along the xaxis,\nthus the largest hopping is of ppσtype between neigh-\nboringpxorbitals, while pyandpzorbitalsareessentially\nnon-bonding. The sign and strengthofthe magneticcou-\nplingbetweenthetwospin-polarizedholesisfoundbythe\ntotal energy difference ∆ EAFM−FMbetween the AFM\nand FM state, calculated within GGA+ Ufor eachp-hole\nFigure 1: (color online) Spin- and orbital-resolved DOS for a\nsingle N in MgO calculated in GGA (a) and GGA+ U(b). (c)\nMinority p-holes spin-density distribution in xy-plane around\nthe N-N dimer, with ( xx), T-shaped ( xy), and (yy) OA. Blue,\nred, and grey spheres denote Mg, O, and N, respectively.\n(d) Spin- and orbital-resolved DOS of N-atom 1 (2) (solid\n(dashed) lines) of the N-N dimer for the depicted OAs, with\nFM spin alignment; red, blue, green line: px-,py-,pz-orbitals.\nOA. To be able to disentangle the electronic mechanism\nof OA from the JT one, for the N-N dimer we performed\ncalculations with and without structural relaxations.\nThe results are summarized in Table I. For each OA\nthe spin configuration (FM or AFM) with the lowest en-\nergy is shown. Clearly, there are substantial energy dif-\nferencesbetweendifferenttypesofOA.ComparingtheN-\nN and N-O-N dimers, the rapid decrease of ∆ EAFM−FM\nwith the N-Ndistance confirmspreviousstudies[31]. Im-\nportantly, for both dimers the FM T-shaped OA (( xy) or\n(xz)) is of lowest energy, moreover for any choice of U\nandJparameters (see Supplemental Material [26]).\nIt was suggested [11, 20] that the introduction of U\nwouldreducethe magneticcouplingin d0-magnets. How-\never, a precise physical value of the parameter Uis not\nknown. To check the persistence of the magnetic cou-\npling, we repeated the calculations for the N-N dimer\nvaryingUinthe rangebetween 0-6eV(neglectingatomic3\nTableI:TheGGA+ Utotal energies ∆ Erelativetothelowest-\nenergy state and exchange interaction ∆ EAFM−FMboth in\nmeV/supercell, calculated for different OA for N-N and N-\nO-N dimers in MgO. Positive (negative) ∆ EAFM−FMcorre-\nsponds to FM (AFM) coupling.\nN-N N-N, relaxed N-O-N\nOA∆E∆EAFM−FM∆E∆EAFM−FM∆E∆EAFM−FM\n(xx)156 −129 99−189 35 −8\n(xy)4 26 0 34 7 2\n(xz)0 24 13 31 0 2\n(yz)185 1 258 1 19 0\n(yy)192 2 280 1 28 0\n(zz)198 5 294 3.6 12 0\nrelaxations), with J/U=0.26. At U=J= 0, the system\nis half-metallic and ferromagnetic with ∆ EAFM−FM=\n265 meV. However, at U >2 eV, it becomes insulating\nand the interaction through the double-exchange mecha-\nnism is no longer possible. Thus, the magnetic coupling\nis much weaker and we expect AFM coupling due to ki-\nnetic exchange interaction. Instead, orbital arrangement\nsets in and despite the insulating state, the interaction\nbetween the spin-1/2 holes remains FM with total mag-\nnetic spin moment of 2 µB. In the following we provide a\nqualitative understanding of this unexpected finding.\nIn the first step we show that kinetic exchange alone\nfavors an AFM ground state. We consider a minimal\nmultiorbital Hubbard model that describes the system of\nstronglycorrelatedopen-shell2 p-electronsintermsofthe\nkinetic energy, t, of spin-conservinghoppingsbetween or-\nbitals of largest overlap (n.n. ppσ-orbitals, see Fig.2(a)),\nthe on-site Coulomb repulsion energy Uand the interor-\nbitalexchange(Hund’srule)coupling J(seealsoEq.(1)).\nFor a given OA, the favored magnetic interaction is\ndetermined by hopping of the electrons within the N-\nN-dimer. A substantial magnetic coupling occurs only\nfor those OAs that allow hopping between the two px-\norbitals of the N-dimer, where at least one of these or-\nbitals is half-filled: the ( xx) and the T-shaped ( xy) and\n(xz) OAs (see Table I). The relative coupling strength\ncan be qualitatively understood by considering that each\nspin-conserving hopping tinto half-filled orbitals lowers\nthe energy by t2/(U−J) (in 2nd order perturbation\ntheory) if the intermediate virtual atomic state satisfies\nHund’s rule for one of the atoms, and t2/Uotherwise.\nIn the (xx) OA, this gives EAFM=−2t2/U,EFM= 0,\ni.e.an AFM ground state (the factor 2 arises because\nthere is one half-filled orbital per atom participating in\nthe hopping). In the T-shaped OA there is only one half-\nfilled orbitalparticipating; then we have EAFM=−t2/U,\nEFM=−t2/(U−J),i.e., a FM ground state.\nAmong all possible types of OA, the T-shaped and\nAFM (xx) states are the ones that have the lowest en-\nergy, because these are the only ones that allow hopping\nbetween the N atoms. The hopping between O and Natoms provides a smaller total energy gain, because of (i)\nthe separation between the on-site energy levels of O and\nN, that differ by 0.5-1 eV, and (ii) the delocalization of\ntheOstatesthat formanitinerantband. Fromtheprevi-\nousdiscussion followsthat amongthe twoOAs, the AFM\n(xx) one should be lower in energy by 2 t2/U−t2/(U−J)\nas long as J < U/2. Thus, the arguments of the kinetic\nexchange interaction favor the AFM ( xx) state, contra-\ndicting our DFT finding of a FM T-shaped ground state.\nObviously, a qualitative description of the relative sta-\nbility of spin- and orbitally-arranged states in N-doped\nMgO requests an extension of the minimal Hubbard\nmodel beyond the conventional electron hopping tand\non-siteCoulomb repulsion U. The N-N dimer breaks\nthe cubic symmetry of the MgO lattice. Considering\nthe directional nature of the 2 p-orbitals, this leads to\nan orbital-dependent intersite Coulomb repulsion, that\nis well accounted for within the GGA/LDA. In terms of\nthe Hubbard model it is expressed by an additional effec-\ntive intersite Coulomb repulsion energy Vbetween n.n.\nppσ-orbitalsof N-N and N-O pairs. To capture this effect\nby such an extended Hubbard model we study an impu-\nrity cluster of 2 N and 6 n.n. O atoms (Fig.2(a)), with\n1 (2)p-orbital on each O (N) site that form n.n. ppσ-\norbitals. Thus, we restrict ourselves to 10 p-orbitals that\ncanbe occupied by18electronsand 2holesofspins ↑and\n↓. The Hamiltonian for such a system can be written as:\nH=/summationdisplay\nm,sǫmnms−t/summationdisplay\n/summationdisplay\ns[c†\nmscm′s+h.c.]\n+U/summationdisplay\nmnm↑nm↓+V/summationdisplay\nnmnm′ (1)\n+ (U′−J)/summationdisplay\n≪m,m′≫,snmsnm′s+U′/summationdisplay\n≪m,m′≫,snmsnm′−s\nHereindicates the pair of n.n. orbitals of ppσ-\noverlap, ǫmdenotes the energy level, c†\nms(cms) creates\n(annihilates) a particle of spin s=↑,↓of orbital m, while\nnms,nm=nm↑+nm↓are the spin-number and num-\nber operators, respectively. Since we have two orbitals\nat the N atoms, U′accounts for the inter-orbital on-site\nCoulomb interaction; ≪m,m′≫stands for the two on-\nsite orbital pairs (1,2) and (3,4). We neglect on-site spin-\nflip and pair-hopping terms as they are non-zero only for\nthe states with an empty and a fully occupied orbital at\nthe same site, which turned out to be of much higher\nenergy. The matrix elements of (1) are evaluated in the\n18-electron-statebasis(190 basis functions in total). The\neigenvalues of the model are obtained numerically by ex-\nactly diagonalizing (1) for a given set of parameters. We\nimpose the relations U > U′> J >0,U′=U−2J\n(for a derivation see the Supplemental Material [26]) and\nwe taket= 0.75 eV (extracted from the splitting of the\nDOS on Fig.1(d)). As the p-levels of N atoms are higher\nin energy than those of O, we take ǫm= 0.5 eV for N and\nzero for O. The pz-orbitals which are omitted from the4\nmodel are considered to be always filled and have only\nan effect of a constant shift on the on-site energies ǫm.\nFigure 2: (color online) (a) A schematic view of the clus-\nter ofp-orbitals, used in the model Hamiltonian (1). The\ndashed blue line depicts px- andpy-orbitals of the n.n. nitro-\ngen atoms (N 1and N 2), while the solid black line depicts the\ncorresponding orbitals of the O atoms. The possible hopping\nprocesses tbetween two orbitals on neighboring sites are also\nshown. (b) The ( U−2J) vs.Uphase diagram of the model\nforV= 0 and V= 0.2 eV. Orange circles denote the Uand\nJvalues, considered in our ab initio calculations [26], outlin-\ning the physically relevant area (in green). (c) The phases I ,\nII, III and IV, corresponding to different sequences of sever al\norbitally-arranged states in the order of energetic stabil ity.\nFigure 2(b) shows a ( U−2J) vs.Uphase diagram, re-\nsulting from Hamiltonian (1) for V= 0 and V= 0.2 eV.\nTheblacklinesdenotetheboundariesbetweenthephases\nI, II, III and IV, corresponding to different sequences\nof energetic stability of OAs (see Fig. 2(c)). We find\nthat the FM ( xy) state is the lowest-energy one for the\nphases II, III and IV. However, for almost all UandJ\nvalues, the AFM ( xx) state (phase I) is the ground state\nifV= 0, contradicting our ab initio results [26]. More-\nover, atV= 0 the sequence of DFT states from Table I\n(phase III) is not reproduced. In contrast, at finite Vthe\nwhole sequence of DFT states is reproduced for a wide\nrange of physically relevant values of UandJ. Hence,\nthe intersite Coulomb interaction V, which defines the\nrepulsion of electrons on n.n. p-orbitals directed towards\neachother,iscrucialforstabilizingtheFMstateindoped\ninsulators with partially occupied p-orbitals like MgO:N.\nIn summary, we have shown that the physics of spin-\npolarized holes of N-doped MgO is largely determined by\np-electron correlation effects. For a single impurity, the\nsplitting of the N 2 p-states at finite Uis mainly an elec-\ntronic effect, enhanced by the lattice distortions. More-\nover, the p-electron correlations lead to a spin and or-\nbital arrangement of the 2 p-hole states of the N-N dimer,\nresulting in a FM coupling between T-shaped orbitally-\narrangedspin-polarized p-holes. The FM state is realized\nat reasonable values of UandJ, if anintersite Coulomb\nrepulsion Vis accounted for. Considering the small valueofV≈0.2 eV, necessary to drive the phase transition\nfromanAFM toaFM groundstate, andthefact thatthe\ndirectional nature of the p-orbitals controls the strength\nof the intersite Coulomb repulsion, we conjecture that\nour finding affects all oxides where magnetism is due to\np-electrons or defects in the p-electron systems.\nWe appreciate valuable discussions with K. Rushchan-\nskii and A. Liebsch, and the support of the J¨ ulich Super-\ncomputing Centre. I.S. and M.L. gratefully acknowledge\nthe support of the Young Investigators Group Program\nof the Helmholtz Association, Contract VH-NG-409.\n∗Corresponding author: i.slipukhina@fz-juelich.de\n[1] K. Sato, L. Bergqvist, J. Kudrnovsk´ y, P.H. Ded-\nerichs, O. Eriksson, I. Turek, B. Sanyal, G. Bouzerar,\nH. Katayama-Yoshida, V.A. Dinh, T. Fukushima,\nH. Kizaki, R. Zeller, Rev. Mod. Phys. 82,1633 (2010).\n[2] S. Picozzi, M. Leˇ zai´ c, S. Bl¨ ugel, phys. stat. sol. (a) 203,\n2738 (2006).\n[3] M. Venkatesan, C.B. Fitzgerald, J.M.D. Coey, Nature\n(London) 430, 630 (2004).\n[4] N.H. Hong, J. Sakai, N. Poirot, V. Briz´ e, Phys. Rev. B\n73, 132404 (2006).\n[5] N.H. Hong, N. Poirot, J. Sakai, Phys. Rev. B 77, 033205\n(2008).\n[6] I.S. Elfimov, S. Yunoki, G.A. Sawatzky, Phys. Rev. Lett.\n89, 216403 (2002).\n[7] J. Hu, Z. Zhang, M. Zhao, H. Qin, L. Sun, X. Kong,\nM. Jiang, Appl. Phys. Lett. 93, 192503 (2008).\n[8] C. DasPemmaraju and S. Sanvito, Phys. Rev. Lett. 94,\n217205 (2005).\n[9] I.S. Elfimov, A. Rusydi, S.I. Csiszar, Z. Hu, H.H. Hsieh,\nH.-J.Lin, C.T. Chen, R. Liang, G.A. Sawatzky, Phys.\nRev. Lett. 98, 137202 (2007).\n[10] K. Kenmochi, M. Seike, K. Sato, A. Yanase,\nH. Katayama-Yoshida, Jpn. J. Appl. Phys. 43, L936\n(2004).\n[11] V. Pardo and W.E. Pickett, Phys. Rev. B. 78, 134427\n(2008).\n[12] H. Pan, J.B. Yi, L. Shen, R.Q. Wu, J.H. Yang, J.Y. Lin,\nY.P. Feng, J. Ding, L.H. Van, J.H. Yin, Phys. Rev. Lett.\n99, 127201 (2007).\n[13] A.L. Ivanovskii, Physics Uspekhi, 50, 1031 (2007).\n[14] R. Long and N.J. English, Phys. Rev. B 80, 115212\n(2009).\n[15] C.-H. Yang, PhD thesis, Stanford University (2010).\n[16] H. Peng, H.J. Xiang, S.-H. Wei, S.-S. Li, J.-B. Xia, J. Li ,\nPhys. Rev. Lett. 102, 017201 (2009).\n[17] A.K. Nandy, P. Mahadevan, P. Sen, and D.D. Sarma,\nPhys. Rev. Lett. 105, 056403 (2010).\n[18] R. Kov´ aˇ cik, and C. Ederer, Phys. Rev. B 80, 140411\n(2009).\n[19] J. Winterlik, G.H. Fecher, C.A. Jenkins, C. Felser,\nC. M¨ uhle, K. Doll, M. Jansen, L.M. Sandratskii, and\nJ. K¨ ubler, Phys. Rev. Lett. 102, 016401 (2009).\n[20] A. Droghetti, C.D. Pemmaraju, S. Sanvito, Phys. Rev.\nB.78, 140404(R) (2008).\n[21] A.I. Liechtenstein, V.I. Anisimov, J. Zaanen, Phys. Re v.\nB52, R5467 (1995).5\n[22] C.D. Pemmaraju, T. Archer, D. Sanchez-Portal, S. San-\nvito, Phys. Rev. B, 75, 045101 (2007).\n[23] J. P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett.\n77, 3865 (1996).\n[24] K.I. Kugel, D.I. Khomskii,, Zh. Eksp. Teor. Fiz. 64, 1429\n(1973) [Sov. Phys. JETP 37, 725 (1973)].\n[25] http://www.flapw.de/\n[26] See EPAPS Document No. ?????? for Supplemen-\ntal Material. For more information on EPAPS, see\nhttp://www.aip.org/pubservs/epaps.html.\n[27] A. Chainani, M. Mathew, D.D. Sarma, Phys. Rev. B 46,9976 (1992).\n[28] A.B. Shick, A.I. Liechtenstein, W.E. Pickett, Phys. Re v.\nB60, 10763 (1999).\n[29] J.A. Chan, S. Lany, A. Zunger, Phys. Rev. Lett. 103,\n016404 (2009).\n[30] H. Wu, A. Stroppa, S. Sakong, S. Picozzi, M. Scheffler,\nP. Kratzer, Phys. Rev. Lett. 105, 267203 (2010).\n[31] Ph. Mavropoulos, M. Leˇ zai´ c, S. Bl¨ ugel, Phys. Rev. B 80,\n184403 (2009)." }, { "title": "1709.09461v1.Effects_of_geometrical_frustration_on_ferromagnetism_in_the_Hubbard_model_on_the_Shastry_Sutherland_lattice.pdf", "content": "arXiv:1709.09461v1 [cond-mat.str-el] 27 Sep 2017Effects of geometrical frustration on\nferromagnetism in the Hubbard model on the\nShastry-Sutherland lattice\nPavol Farkaˇ sovsk´ y\nInstitute of Experimental Physics, Slovak Academy of Scien ces\nWatsonova 47, 040 01 Koˇ sice, Slovakia\nAbstract\nThe small-cluster exact-diagonalization calculations an d the projector\nquantum Monte Carlo method are used to examine the competing effects of\ngeometrical frustration and interaction on ferromagnetis m in the Hubbard\nmodel on the Shastry-Sutherland lattice. It is shown that th e geometrical\nfrustration stabilizes the ferromagnetic state at high ele ctron concentrations\n(n/greaterorsimilar7/4), where strong correlations between ferromagnetism and t he shape\nof the noninteracting density of states are observed. In par ticular, it is\nfound that ferromagnetism is stabilized only for these valu es of frustration\nparameters, which lead to the single peaked noninterating d ensity of states\nat the band edge. Once, two or more peaks appear in the noninte racting\ndensityof states at thebandegdetheferromagnetic stateis suppressed. This\nopens a new route towards the understanding of ferromagneti sm in strongly\ncorrelated systems.\n11 Introduction\nSince its introduction in 1963, the Hubbard model [1] has become, on e of the most\npopularmodelsofcorrelatedelectronsonalattice. Ithasbeenuse dintheliterature\nto study a great variety of many-body effects in metals, of which fe rromagnetism,\nmetal-insulator transitions, charge-density waves and supercon ductivity are the\nmost common examples. Of all these cooperative phenomena, the p roblem of fer-\nromagnetism in the Hubbard model has the longest history. Althoug h the model\nwas originally introduced to describe the itinerant ferromagnetism in narrow-band\nmetals like Fe,Co,Ni and others, it soon turned out that the single-band Hubbard\nmodel is not the canonical model for ferromagnetism. Indeed, th e existence of sat-\nurated ferromagnetism has been proven rigorously only for very s pecial limits. The\nfirst well-known example is the Nagaoka limit that corresponds to the infinite-U\nHubbardmodel withoneholein ahalf-filledband [2]. Another example, w here sat-\nurated ferromagnetism has been shown to exist, is the case of the one-dimensional\nHubbard model with nearest and next-nearest-neighbor hopping at low electron\ndensities [3]. Furthermore, several examples of the fully polarized g round state\nhave been found on special lattices as are the bipartite lattices with sublattices\ncontaining a different number of sites [4], the fcc-type lattices [5, 6], the lattices\nwith long-range electron hopping [7, 8, 9, 10, 11], the flat bands [12, 13, 14, 15] and\nthe nearly flat-band systems [16, 17, 18, 19]. This indicates that th e lattice struc-\nture, which dictates the shape of the density of states (DOS), pla ys an important\nrole in stabilizing the ferromagnetic state.\nMotivated by these results, in the current paper we focus our att ention on\nthe special type of lattice, the so-called Shastry-Sutherland latt ice (SSL). The\nSSL represents one of the simplest systems with geometrical frus tration, so that\nputting the electrons on this lattice one can examine simultaneously b oth, the ef-\n2fect of interaction as well as the effect of geometrical frustratio n on ground-state\nproperties of the Hubbard model. This lattice was first introduced b y Shastry and\nSutherland [20] asaninteresting example ofa frustrated quantum spin system with\nan exact ground state. It can be described as a square lattice with the nearest-\nneighbor links t1and the next-nearest neighbors links t2in every second square\n(see Fig. 1a). The SSL attracted much attention after its experim ental realization\n(a)\nt1t2(b)\nt1t2t3\nFigure 1: (a) The original SSL with the first ( t1) and second ( t2) nearest-neighbor\ncouplings. (b) The generalized SSL with the first ( t1), second ( t2) and third ( t3) nearest-\nneighbor couplings.\nin theSrCu2(BO3)2compound [21]. The observation of a fascinating sequence of\nmagnetization plateaus (at m/ms=1/2, 1/3, 1/4 and 1/8 of the saturated magne-\ntizationms) in this material [22] stimulated further theoretical and experimen tal\nstudies of the SSL. Some time later, many other Shastry-Sutherla nd magnets have\nbeen discovered [23, 24]. In particular, this concerns an entire gr oup of rare-earth\nmetal tetraborides RB4(R=La−Lu). These materials exhibit similar sequences\nof fractional magnetization plateaus as observed in the SrCu2(BO3)2compound.\nFor example, for TbB4the magnetization plateau has been found at m/ms=2/9,\n1/3, 4/9, 1/2 and 7/9 [23] and for TmB4atm/ms=1/11, 1/9, 1/7 and 1/2 [24].\nTo describe some of the above mentioned plateaus correctly, it was necessearry to\ngeneralize the Shastry-Sutherland model by including couplings bet ween the third\n3and even between the forth nearest neighbors [25]. The SSL with th e first, second\nand third nearest-neighbor links is shown in Fig. 1b and this is just the lattice that\nwill be used in our next numerical calculations.\nThus our starting Hamiltonian, corresponding to the one band Hubb ard model\non the SSL, can be written as\nH=−t1/summationdisplay\n/angbracketleftij/angbracketright1,σc+\niσcjσ−t2/summationdisplay\n/angbracketleftij/angbracketright2,σc+\niσcjσ−t3/summationdisplay\n/angbracketleftij/angbracketright3,σc+\niσcjσ+U/summationdisplay\nini↑ni↓,(1)\nwherec+\niσandciσare the creation and annihilation operators for an electron of\nspinσat siteiandniσis the corresponding number operator ( N=N↑+N↓=\n/summationtext\niσniσ). The first three terms of (1) are the kinetic energies correspon ding to\nthe quantum-mechanical hopping of electrons between the first, second and third\nnearest neighbors and the last term is the Hubbard on-site repulsio n between two\nelectrons with opposite spins. We set t1= 1 as the energy unit and thus t2(t3)\ncan be seen as a measure of the frustration strength.\nTo identify the nature of the ground state of the Hubbard model o n the SSL\nwe have used the small-cluster-exact-diagonalization (Lanczos) m ethod [26] and\nthe projector quantum Monte-Carlo method [27]. In both cases t he numerical\ncalculations proceed in the folloving steps. Firstly, the ground-sta te energy of the\nmodelEg(Sz) is calculated in all different spin sectors Sz=N↑−N↓as a function\nof model parameters t2,t3andU. Then the resulting behaviors of Eg(Sz) are used\ndirectly to identify the regions in the parametric space of the model, where the\nfully polarized state has the lowest energy.\n2 Results and discussion\nToreveal possible stabilityregionsoftheferromagneticstateinth eHubbardmodel\non the SSL, let us first examine the effects of the geometrical frus tration, repre-\nsented by nonzero values of t2andt3, on the behavior of the non-interacting DOS.\n4The previous numerical studies of the standard one and two-dimen sional Hub-\nbard model with next-nearest [3] as well as long-range [9, 10, 11] hopping showed\nthat just this quantity could be used as a good indicator for the eme rgence of\nferromagnetism in the interacting systems. Indeed, in both models the strong cor-\nrelation between ferromagnetism and the anomalies in the nonintera cting DOS are\nobserved. Inthefirstmodeltheferromagenticstateisfoundat lowelectronconcen-\ntrations and the noninteracting DOS is strongly enhanced at the low -energy band\nedge, while in the second one the ferromagnetic phase is stabilized at the high elec-\ntron concentrations and the spectral weight is enhanced at the h igh-energy band\nedge. This leads to the scenario according to which the large spectr al weight in\nthe noninteracting DOS that appears at the low (high) energy band edges allows\nfor a small kinetic-energy loss for a state with total spin S/negationslash= 0 in reference to one\nwithS= 0. At some finite value of interaction U, the Coulomb repulsion paid\nfor the low-spin states overcomes this energy loss and the high-sp in state becomes\nenergetically favored. The key point in this picture is the assumption that the\nshape of the DOS is only weakly modified as the interaction Uis switched on, at\nleast within its low (high) energy sector.\nThe noninteracting DOS of the U= 0 Hubbard model on the SSL of size\nL= 200×200, obtained by exact diagonalization of H(forU= 0) is shown in\nFig.2. Theleftpanelscorrespondtothesituationwhen t2>0andt3= 0,whilethe\nright panels correspond to the situation when both t2andt3are finite. One can see\nthat once the frustration parameter t2is nonzero, the spectral weight starts to shift\nto the upper band edge and the noninteracting DOSbecomes stron gly asymmetric.\nThus taking into account the above mentioned scenario, there is a r eal chance\nthat the interacting system could be ferromagnetic in the limit of high electron\nconcentrations. To verify this conjecture we have performed ex haustive numerical\n5-5 0 500.050.1DOSt2=0\nt3=0\n-5 0 500.050.1DOSt2=0.2\nt3=0\n-5 0 500.050.1DOSt2=0.4\nt3=0\n-5 0 500.050.1DOSt2=0.6\nt3=0\n-5 0 500.050.1DOSt2=0.8\nt3=0\n-5 0 5\nE00.050.1DOSt2=1\nt3=0-5 0 500.050.1DOSt2=1\nt3=0\n-5 0 500.050.1DOSt2=1\nt3=0.2\n-5 0 500.050.1DOSt2=1\nt3=0.4\n-5 0 500.050.1DOSt2=1\nt3=0.6\n-5 0 500.050.1DOSt2=1\nt3=0.8\n-5 0 5\nE00.050.1DOSt2=1\nt3=1\nFigure 2: Non-interacting DOS calculated numerically for different va lues oft2andt3\non the finite cluster of L= 200×200 sites.\nstudies of the model Hamiltonian (1) for a wide range of the model pa rameters\nU,t2andnatt3= 0. Typical results of our PQMC calculations obtained on finite\ncluster of L= 6×6 sites, in two different concentration limits ( n≤1 andn >1)\nare shown in Fig. 3. There is plotted the difference ∆ E=Ef−Eminbetween the\nferromagnetic state Ef, which can be calculated exactly and the lowest ground-\nstate energy from Eg(Sz) as a function of the frustration parameter t2. According\n60 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6\nt200.20.40.60.811.21.41.61.82∆ E(a)\nn=1/4\nn=1/2\nn=1\n0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6\nt200.10.20.30.40.50.60.70.8∆ E(b)\nn=5/4\nn=3/2\nn=7/40 0.4 0.8 1.2 1.6\nt200.20.40.60.8∆ En=5/4\nn=7/4L=36\nL=64\nFigure 3: The difference ∆ E=Ef−Eminbetween the ferromagnetic state Efand\nthe lowest ground-state energy from Eg(Sz) as a function of the frustration parameter\nt2calculated for n≤1 (a) and n >1 (b) on the finite cluster of L= 6×6 sites\n(U= 1,t3= 0). The inset shows ∆ E, calculated for two different electron densities on\nclusters of L= 6×6 andL= 8×8 sites.\nto this definition the ferromagnetic state corresponds to ∆ E= 0. It is seen that for\nelectron concentrations below the half filled band case n= 1, ∆Eis the increasing\nfunction of t2, and thus there is no sign of stabilization of the ferromagnetic stat e\nforn≤1, in accordance with the above mentioned scenario. The situation lo oks\nmore promising in the opposite limit n >1. In this case, ∆ Eis considerably\nreduced with increasing t2, however, this reduction is still insufficient to reach the\nferromagnetic state ∆ E= 0 for physically reasonable values of t2(t2<1.6) that\ncorrespond to the situation in the real materials. To exclude the fin ite-size effect,\nwe have also performed the same calculations on the larger cluster o fL= 8×8\nsites, but againno signs of stabilization the ferromagnetic state ha ve been observed\n(see inset to Fig. 4b).\nFor this reason we have turned our attention to the case t2>0 andt3>0.\nThe noninteracting DOS corresponding to this case is displayed in Fig. 1 (the right\npanels). Thesepanelsclearlydemonstratethatwiththeincreasing valueofthefrus-\ntration parameter t3, still a more spectral weight is shifted to the upper band edge.\nA special situation arises at t3= 0.6, when the spectral weight is strongly peaked\n7at the upper band edge. In this case the nonintercting DOS is pract ically identical\nwith one corresponding to noninteracting electrons with long-rang e hopping [9].\nSince the long-range hopping supports ferromagnetism in the stan dard Hubbard\nmodel for electron concentrations above the half-filled band case [9, 10, 11], we\nexpect that this could be true also for the Hubbard model on the SS L, at least for\nsome values of frustration parameters t2andt3. Therefore, we have decided to per-\nform numerical studies of the model for a wide range of t3values at fixed t2,Uand\nn(t2= 1,U= 1,n= 7/4). To minimize the finite-size effects, the numerical calcu-\nlations have been done on two different finite clusters of L= 6×6 andL= 8×8\nsites. The results of our calculations for ∆ Eas a function of t3are displayed in\nFig. 4a. In accordance with the above mentioned assumptions we fin d a relatively\n0.2 0.4 0.6 0.8 1\nt300.010.020.030.040.05∆ E(a)\nU=1\nn=7/4\nt2=1L=36\nL=640 2 4\nU0.40.50.60.70.8t3 ferroL=36\nt2=1, n=7/4\n0.2 0.4 0.6 0.8 1 1.2 1.4 1.6\nt20.40.50.60.70.80.91t3(b)\nferro\nn=7/4\nU=1L=36\n1.75 1.8 1.85\nn0.150.350.550.75t3\nt2=1, U=1ferro\nPQMC\nexact\nFigure 4: (a) The difference ∆ E=Ef−Eminas a function of the frustration parameter\nt3calculated for U= 1,t2= 1,n= 7/4 and two different finite clusters of L= 6×6\nandL= 8×8 sites. The inset shows the ground-state diagram of the mode l in thet3-U\nplane. (b) The comprehensive phase diagrams of the model in t het3-t2andt3-nplane.\nwide region of t3values around t3= 0.6, where the ferromagnetic state is stable.\nIt is seen that the finite-size effects on the stability region of the fe rromagnetic\nphase are negligible and thus these results can be satisfactorily ext rapolated to\nthe thermodynamic limit L=→ ∞. Moreover, the same calculations performed\nfor different values of the Hubbard interaction Ushowed that correlation effects\n8(nonzero U) further stabilize the ferromagnetic state and lead to the emerge nce of\nmacroscopic ferromagnetic domain in the t3-Uphase diagram (see inset to Fig. 4a).\nThis confirms the crucial role of the Hubbard interaction Uin the mechanism of\nstabilization of ferromagnetism on the geometrically frustrated lat tice. In Fig. 4b\nwe have also plotted the comprehensive phase diagrams of the mode l in thet3-n\nas well as t3-t2plane, which clearly demonstrate that the ferromagnetic state is\nrobust with respect to doping ( n/greaterorsimilar7/4) and frustration.\nTo check the convergence of PQMC results we have performed the same calcu-\nlations by the Lanczos exact diagonalization method. Of course, on such a large\ncluster, consisting of L= 6×6 sites, we were able to examine (due to high memory\nrequirements) only several electron fillings near the fully occupied b and (N= 2L).\nThe exact diagonalization and PQMC results for the width of the ferr omagnetic\nphase obtainedonfinite cluster of L= 6×6sites, for threedifferent electron fillings\nfrom the high concentration limit ( N= 66,67,68), are displayed in the inset to\nFig. 4b and they show a nice convergence of PQMC results.\nLet us finally turn our attention to the question of possible connect ion between\nferromagnetism and the noninteracting DOS that has been discuss ed at the begin-\nning of the paper. Figs. 4a and 4b show, that for each finite Uandnsufficiently\nlarge (n/greaterorsimilar7/4), there exists a finite interval of t3values, around t3∼0.6, where\nthe ferromagnetic state is the ground state of the model. To exam ine a possible\nconnection between ferromagnetism and the noninteracting DOS, we have calcu-\nlated numerically the noniteracting DOS for several different values oft3from this\ninterval and its vicinity. The results obtained for U= 1,n= 7/4 andt2= 1 are\ndisplayed in Fig. 5. Comparing these results with the ones presented in Fig. 4a for\nthe stability region of the ferromagnetic phase at the same values o fU,nandt3,\none can see that there is an obvious correlation between the shape of the noninter-\n91 2 300.050.1DOSt2=1\nt3=0.5\n1 2 300.050.1DOSt2=1\nt3=0.54\n1 2 300.050.1DOSt2=1\nt3=0.58\n1 2 300.050.1DOSt2=1\nt3=0.6\n1 2 300.050.1DOSt2=1\nt3=0.62\n1 2 3\nE00.050.1DOSt2=1\nt3=0.64\n1 2 3\nE00.050.1DOSt2=1\nt3=0.68\n1 2 3\nE00.050.1DOSt2=1\nt3=0.72\nFigure 5: Non-interacting DOS calculated numericallly for t2= 1 and different values\noft3(neart3= 0.6) on the finite cluster of L= 200×200 sites.\nacting DOS and ferromagnetism. Indeed, the ferromagnetic stat e is stabilized only\nfor these values of frustration parameters t2,t3, which lead to the single peaked\nnoninterating DOS at the band edge. Once, two or more peaks appe ar in the\nnoninteracting DOS at the band egde (by changing t2ort3), ferromagnetism is\nsuppressed.\nInsummary, thesmall-clusterexact-diagonalizationcalculationsan dthePQMC\n10method were used to examine possible mechanisms leading to the stab ilization of\nferromagnetism in strongly correlated systems with geometrical f rustration. Mod-\nelling such systems by the Hubbard model on the SSL, we have found that the\ncombined effects of geometrical frustration and interaction stro ngly support the\nformation of the ferromagnetic phase at high electron densities. T he effects of\ngeometrical frustration transform to the mechanism of stabilizat ion of ferromag-\nnetism via the behaviour of the noninteracting DOS, the shape of wh ich is deter-\nmined uniquely by the values of frustration parameters t2andt3. We have found\nthat it is just the shape of the noninteracting DOS near the band ed ge (the single\npeaked DOS) that plays the central role in the stabilization of the fe rromagnetic\nstate. Since the same signs have been observed also in some other w orks (e.g., the\nHubbard model with nearest and next-nearest neighbor hopping, or the Hubbard\nmodel with long range hopping), it seems that such a behaviour of th e noninter-\nacting DOS near the band edge should be used like the universal indica tor for the\nemergence of ferromagnetism in the interacting systems.\nThis work was supported by the Slovak Research and Development A gency\n(APVV) under Grant APVV-0097-12 and ERDF EU Grant under the c ontract\nNo. ITMS26210120002 and ITMS26220120005.\n11References\n[1] J. Hubbard, Proc. R. Soc. London A 276, 238 (1963).\n[2] Y. Nagaoka, Phys. Rev. 147, 392 (1966).\n[3] E. M¨ uller-Hartmann, J. Low. Temp. Phys. 99, 342 (1995).\n[4] E. H. Lieb, Phys. Rev. Lett. 62, 1201 (1989).\n[5] M. Ulmke, Eur. Phys. J. B 1, 301 (1998).\n[6] S. Pandey and A. Singh, Phys. Rev. B 75, 064412 (2007).\n[7] P. Pieri, Mod. Phys. Lett. B 10, 1277 (1996).\n[8] M. Salerno, Z. Phys. B 99, 469 (1996).\n[9] P. Farkaˇ sovsk´ y, Phys. Rev. B 66, 012404 (2002).\n[10] P. Farkaˇ sovsk´ y and Hana ˇCencarikov´ a, Cent. Eur. J. Phys. 11, 119 (2013).\n[11] P. Farkaˇ sovsk´ y, EPL 110, 47007 (2015).\n[12] A. Mielke, J. Phys. A 25, 4335 (1992).\n[13] H. Tasaki, Phys. Rev. Lett. 69, 1608 (1992).\n[14] A. Mielke and H. Tasaki, Commun. Math. Phys. 158, 341 (1993).\n[15] H. Katsura, I. Maruyama, A. Tanaka and H. Tasaki, EPL 91, 57007 (2010).\n[16] H. Tasaki, Phys. Rev. Lett. 73, 1158 (1994).\n[17] A. Mielke, Phys. Rev. Lett. 82, 4312 (1999).\n[18] A. Tanaka and H. Ueda, Phys. Rev. Lett. 90, 067204 (2003).\n12[19] M. Maksymenko, A Honecker, R. Moessner, J. Richter and O. D erzhko, Phys.\nRev. Lett. 109, 096404 (2012).\n[20] B.S. Shastry, B. Sutherland, Physica B and C 108, 1069 (1981).\n[21] H. Kageyama, K. Yoshimura, R. Stern, N. V. Mushnikov, K. Oniz uka, M.\nKato, K. Kosuge, C. P. Slichter, T. Goto, and Y. Ueda, Phys. Rev. Lett.82,\n3168 (1999).\n[22] K. Kodama, M. Takigawa, M. Horvatic, C. Berthier, H. Kageyam a, Y. Ueda,\nS. Miyahara, F. Becca, and F. Mila, Science 298, 395 (2002).\n[23] S. Yoshii, T. Yamamoto, M. Hagiwara, S. Michimura, A. Shigekawa , F. Iga,\nT. Takabatake, and K. Kindo, Phys. Rev. Lett. 101, 087202 (2008).\n[24] K. Siemensmeyer, E. Wulf, H. J. Mikeska, K. Flachbart, S. Gaba ni, S. Matas,\nP. Priputen, A. Efdokimova, and N. Shitsevalova, Phys. Rev. Lett .101,\n177201 (2008).\n[25] K.ˇCenˇ carikov´ a and P. Farkaˇ sovsk´ y, Physica Status Solidi b 252 , 333 (2015).\n[26] E. Dagotto, Rev. Mod. Phys. 66, 763 (1994).\n[27] M. Imada and Y. Hatsugai, J. Phys. Soc. Jpn. 58, 3752 (1989) .\n13" }, { "title": "2112.14515v1.A_unified_theory_of_ferromagnetic_quantum_phase_transitions_in_heavy_fermion_metals.pdf", "content": "arXiv:2112.14515v1 [cond-mat.str-el] 29 Dec 2021A unified theory of ferromagnetic quantum phase transitions in heavy fermion metals\nJiangfan Wang1and Yi-feng Yang1,2,3,∗\n1Beijing National Laboratory for Condensed Matter Physics,\nInstitute of Physics, Chinese Academy of Science, Beijing 1 00190, China\n2School of Physical Sciences, University of Chinese Academy of Sciences, Beijing 100190, China\n3Songshan Lake Materials Laboratory, Dongguan, Guangdong 5 23808, China\n(Dated: December 30, 2021)\nMotivated by the recent discovery of a continuous ferromagn etic quantum phase transition in\nCeRh 6Ge4and its distinction from other U-based heavy fermion metals such as UGe 2, we develop a\nunified explanation of their different ground state properti es based on an anisotropic ferromagnetic\nKondo-Heisenberg model. We employ an improved large- NSchwinger boson approach and predict\na full phase diagram containing both a continuous ferromagn etic quantum phase transition for large\nmagnetic anisotropy and first-order transitions for relati vely small anisotropy. Our calculations re-\nveal three different ferromagnetic phases including a half- metallic spin selective Kondo insulator\nwith a constant magnetization. The Fermi surface topologie s are found to change abruptly between\ndifferent phases, consistent with that observed in UGe 2. At finite temperatures, we predict the devel-\nopment of Kondo hybridization well above the ferromagnetic long-range order and its relocalization\nnear the phase transition, in good agreement with band measu rements in CeRh 6Ge4. Our results\nhighlight the importance of magnetic anisotropy and provid e a unified theory for understanding the\nferromagnetic quantum phase transitions in heavy fermion m etals.\nKeywords: Ferromagnetic Kondo lattice; Magnetic anisotropy; Quantu m phase transitions;\nKondo hybridization\nINTRODUCTION\nHeavy fermion or Kondo lattice materials are proto-\ntype strongly correlated electron systems that exhibit\nrich ground states due to competing interactions [1, 2].\nAn indirect Ruderman-Kittel-Kasuya-Yosida (RKKY)\ninteraction is typically induced and oscillates with dis-\ntance between neighboring Kondo impurities. Its re-\nsulting ferromagnetic (FM) or antiferromagnetic (AFM)\nlong-range orders can often be suppressed by pressure,\nmagnetic field, or chemical substitution and tuned to\na heavy Fermi liquid (HFL) with an enhanced effec-\ntive mass of Landau quasiparticles due to the Kondo\nscreeningby conduction electrons. However,unlike AFM\nKondo lattices, whose magnetic quantum phase transi-\ntions (QPTs) and associated quantum criticality have\nbeen intensively studied in the past decades [3, 4], the\nFM Kondo lattices are relatively less explored [5–9]. It\nhas been believed that FM QPT in clean metallic sys-\ntems should typically be of first order if not interrupted\nby other instabilities [10] because of non-analyticities in\nthe free energy induced by certain soft modes related\nto the Fermi surfaces [11, 12]. Evidences may be found\nfrom experiments on many U-based intermetallics such\nas UGe 2[13].\nThe situation, however, has been changed lately when\na pressure-tuned FM quantum critical point (QCP)\nand the associated strange metallic behaviors were ob-\nserved in the stoichiometric Kondo lattice compound,\nCeRh6Ge4[14]. Similar observations had been re-\nported previously in the non-stoichiometric compound\nYbNi4(P1−xAsx)2[15], but might be naively attributed\nto disorder effects. These observations violate theprevailing wisdom and have stimulated rapid theoreti-\ncal progress and debates on the underlying mechanism\n[14, 16]. Analogous to the AFM quantum criticality, a\nKondobreakdownscenariohasalsobeenpredictedacross\nthe FM QCP [14]. But this picture cannot explain the\nlatest observations of angle-resolved photoemission spec-\ntroscopy(ARPES)[17]andultrafastopticalspectroscopy\n[18], where strong hybridization effects have been de-\ntected at finite temperatures above the FM transition.\nCompared to CeRh 6Ge4, the intensively studied U-\nbased compound, UGe 2, exhibits very different ground\nstate properties under pressure [13]. At low pressures, it\nhastwodifferent FM phaseswith strong(FM2) andweak\n(FM1) spin polarizations, respectively, and enters a HFL\nupon further increasing pressure. Both FM2-FM1 and\nFM1-HFL transitions are of first order [19], with drastic\nchanges of Fermi surfaces reported by quantum oscilla-\ntion experiments [20, 21]. A superconducting dome oc-\ncursinside the FM phaseand the maximalsuperconduct-\ning transition temperature occurs right at the FM2-FM1\ntransition point [13].\nGiven these distinct properties, one may doubt if there\nshould exist a single mechanism for both compounds\nwithout resorting to microscopic details. In this work,\nwe propose such a unified theory and show that their\nvery different properties can in fact be explained solely\nbased on lattice anisotropy. The Ce atoms of CeRh 6Ge4\nexhibits a quasi-one-dimensional structure [14], while\nthe nearest neighbor U-U separations in UGe 2are more\nisotropic [22]. Such structural anisotropy on atomic dis-\ntances may have a sensitive influence and cause a mag-\nnetic anisotropy in the RKKY interaction, as reflected in\nan anisotropic Kondo-Heisenberg model. We will solve2\nthe model using an improved Schwinger boson large- N\napproachtakingintoconsiderationthespatialcorrelation\neffect. The method has lately been applied to the AFM\nKondo lattice [23], yielding a global phase diagram and\na unified explanation of the AFM QCP in YbRh 2Si2and\nthe quantum critical non-Fermi liquid phase (a metallic\nspin liquid) recently observed in the frustrated Kondo\nlattice CePdAl [24]. Here we further develop this ap-\nproach for the FM phase, predict a full phase diagram\nof the anisotropic FM Kondo lattice, and give a unified\nexplanation of the rather complicated and qualitatively\ndifferent behaviors of CeRh 6Ge4and UGe 2. Our meth-\nods and results are reported in detail in the following\nsections.\nMETHODS\nWe start with the Hamiltonian:\nH=/summationdisplay\n/angbracketleftij/angbracketrighttijc†\niaαcjaα+JK/summationdisplay\niSi·si−/summationdisplay\ni,δJδ\nHSi·Si+δ,(1)\nwherec†\niaαcreates a conduction electron of spin αand\nchannel (orbital) a= 1,2,···,Kon siteiof a square\nlattice (summations over repeated indices are implied),\nsi=1\n2/summationtext\naαβc†\niaασαβciaβis its spin operator, and Si\ndenotes the local spin formed by either Ce 4 f-electrons\nor U 5f-electrons beyond the Hill limit [25]. δ=x,y\ndenotes the unit vectors along the two spatial direc-\ntions. In principle, the Kondo couping will induce an\nanisotropic RKKY interaction if the electron hopping tij\nis anisotropic. Here, we apply directly an anisotropy on\nthe FM Heisenberg interaction ( Jx\nH/negationslash=Jy\nH>0) while\nkeeping the electron hopping isotropic to emphasize the\neffect of magnetic anisotropy alone.\nTo deal with FM and Kondo effect on the same ba-\nsis, we use the Schwinger boson representation of local\nspins,Si=1\n2/summationtext\nαβb†\niασαβbiβ, together with the con-\nstraintnb(i)≡/summationtext\nαb†\niαbiα= 2S, and perform large- N\ncalculationsby extending the spin groupSU(2) to SU( N)\nwithα= 1,···,Nand a fixed ratio κ≡2S/N=\nK/Nfrom the perfect Kondo screening condition [26].\nThe constraint is implemented by a Lagrange multiplier,\nλ/summationtext\ni(nb(i)−2S). The Kondo and Heisenberg terms can\nbe decomposed by two auxiliary fields:\nJKSi·si→1√\nNb†\niαciaαχia+H.c.+|χia|2\nJK,\n−Jδ\nHSi·Si+δ→∆δb†\ni+δ,αbiα+H.c.+N|∆δ|2\nJδ\nH,(2)\nwhereχ†\niacreates a spinless fermion with positive charge\ncalled holon, and ∆ δ=−Jδ\nH/summationtext\nα/angbracketleftBig\nb†\niαbi+δ,α/angbracketrightBig\n/Ndenotes\nthe hopping amplitude of bosonic spinons along the δ-\ndirection. We will use the mean-field ratio ∆ x/∆yto\nmeasure the degree of magnetic anisotropy.In the large- Nlimit, one only needs to consider the\none-loop self-energies of spinons and holons:\nΣb(p,iνn) =−κ\nβV/summationdisplay\nkmgc(p−k,iνn−iωm)Gχ(k,iωm),\nΣχ(p,iωm) =1\nβV/summationdisplay\nkngc(k−p,iνn−iωm)Gb(k,iνn),(3)\nwheregc(p,iωm) = (iωm−εc\np)−1isthebareGreen’sfunc-\ntion of conduction electrons with dispersion εc\np,Gband\nGχare the self-consistent Green’s functions of spinons\nand holons, ωm(νn) are the fermionic (bosonic) Matsub-\nara frequencies, βis the inverse temperature, and Vis\nthe total number of lattice sites.\nThe Green’s functions satisfy:\nGb(p,iνn) =1\niνn−εbp−Σb(p,iνn),\nGχ(p,iωn) =1\n−J−1\nK−Σχ(p,iωn)−b2\niωn+εc\n−p,\nGα\nc(p,iωn) =1\niωn−εcp+b2δα,1\n−J−1\nK−Σχ(−p,−iωn),(4)\nwhereεb\np≡λ−2(∆xcospx+∆ycospy) is the barespinon\ndispersion. The FM long range order is accounted for by\nthe condensation of bosonic spinons on one of its spin\ncomponent (chosen to be α= 1 here), /angbracketleftbi,1/angbracketright=√\nNb/negationslash= 0,\nso that/angbracketleftSz/angbracketright ≡N−1/summationtext\nαsgn(1+N\n2−α)/angbracketleftb†\niαbiα/angbracketright=b2[27].\nIn this formalism, the Kondo effect at large JKis\nsignaled by the negative effective Kondo coupling (or\nholon energy level), J∗\nK(k)−1≡J−1\nK+ ReΣ χ(k,0)<0,\nwhich energetically favors formation of Kondo bound\nstates. It then leaves an energy gap around the Fermi\nenergy for both holons and spinons, producing a large\nelectron Fermi surface due to the generalized Luttinger\nsum rule [23, 28]. As one increases Jδ\nH, the spinon\n(holon) gap shrinks and finally vanishes at some point\nwhere the bosonic spinons start to condense due to the\ndevelopment of long-range FM order. Inside the FM\nphase, the condensation (magnetization) provides an ef-\nfective hybridization field between the conduction elec-\ntrons (cia,α=1) and the holons, as may be directly seen\nfrom the decomposed Kondo term in Eq. (2).\nIn practice, Eqs. (3)(4) are solved self-consistently\nunder two constraints corresponding to minimizing the\nfree energy with respect to λandb. Previous studies\nof the Kondo lattice models using Schwinger boson ap-\nproach have only focused on the paramagnetic phase and\nadopted a local (or momentum-independent) approxima-\ntion for the self-energies [5, 29, 30]. Here we use the\nfast Fourier transform of the self-consistent equations\nand solve the nonlocal (or momentum-dependent) self-\nenergies efficiently in the coordinate space [23]. More\ndetails can be found in Supplementary Materials.3\n/s32/s33/s34/s35/s32/s33/s34/s36/s32/s33/s34/s32/s32/s33/s32/s37/s32/s33/s32/s38/s32/s33/s32/s35/s32/s33/s32/s36/s32/s33/s32/s32/s39/s40/s41/s42/s40/s43\n/s32/s33/s34/s32 /s32/s33/s32/s37 /s32/s33/s32/s38 /s32/s33/s32/s35 /s32/s33/s32/s36\n/s39/s44/s45/s42/s46/s47/s43FM2 FM1FM0\nHFL\ncoexisting \n region /s32/s33/s34/s34/s33/s35/s34/s33/s36/s34/s33/s37/s34/s33/s38/s34/s33/s34/s39/s40/s41/s42/s40/s43\n/s34/s33/s38/s34/s34/s33/s32/s44/s34/s33/s32/s34/s34/s33/s34/s44/s34/s33/s34/s34\n/s39/s45/s46/s42/s47/s48/s43(a)\n(b) (c)CeRh 6Ge 4\nUGe 2\n/s32\n/s33\n/s34\n/s35/s36/s37/s38/s39/s40\n/s33/s41/s35/s34/s41/s42/s34/s41/s35/s35/s41/s42/s35/s41/s35\n/s43/s37/s38/s44/s45/s46/s40/s32/s33\n/s34/s33\n/s35/s33\n/s33/s36/s37/s38/s39/s40\n/s35/s41/s33/s42/s41/s43/s42/s41/s33/s33/s41/s43/s33/s41/s33\n/s44/s37/s38/s45/s46/s47/s40FM1 \nFM2crossover \nFM Strange Metal\nHFL HFLCeRh6Ge 4UGe 2\nFIG. 1: Phase diagram of the anisotropic ferromag-\nnetic Kondo lattice. (a) Zero temperature phase diagram\nof the model described by Eq. (1). The heavy Fermi liq-\nuid (HFL) phase at large TK/Jy\nHends at the red solid line,\nwhile the ferromagnetic (FM) phases at small TK/Jy\nHend\nat the blue solid line. Between the red and blue lines (light\ngreen area), HFL and FM solutions coexist. The three differ-\nent FM phases, FM0, FM1 and FM2, are separated by two\ncontinuous transition lines (the black and grey dashed line s).\nThe two orange arrows correspond to increasing pressure of\nCeRh 6Ge4and UGe 2. The inset shows the phase diagram\nwithin the full range of anisotropy. (b)(c) The experimen-\ntal temperature-pressure phase diagram of CeRh 6Ge4and\nUGe2, respectively. Black lines are continuous transitions,\nshort dashed lines are first-order transitions, and long das hed\nlines are crossover lines. The small yellow region of (c) de-\nnotes the superconducting phase. The experimental phase\ndiagrams are reproduced based on Refs. [14] and [13].\nRESULTS\nWe first summarize the predicted zero temperature\nphase diagram in Fig. 1a in terms of the anisotropy ra-\ntio ∆x/∆yand the Doniach ratio TK/Jy\nH, whereTK=\nDe−2D/JKis the characteristic Kondo energy scale and\nDis the electron half bandwidth. The HFL phase exists\non the right hand side of the red line (larger TK/Jy\nH),\nand the FM phases appear on the left hand side of the\nblue line (smaller TK/Jy\nH). The two lines merge together\nfor large anisotropy (small ∆ x/∆y), indicating a contin-\nuous quantum phase transition and a single FM QCP.While for smaller anisotropy (larger ∆ x/∆y), there is\na coexisting region (light green area) between the two\nlines, indicating a first-order transition between FM and\nHFL. It should be mentioned that the Schwinger boson\nlarge-Ntheory suffers from the pathology of predicting\nartificial first-order transitions [31], which can be alle-\nviated via introducing a biquadratic interaction term,\n−ζ/summationtext\ni,δJδ\nH(Si·Si+δ)2[5, 14, 23]. In our case, such a\nterm with a reasonable choice of ζonly slightly shifts the\ntricritical point from ∆ x/∆y≈0.02 to 0.03 (see Sup-\nplementary Materials) and has no effect on the overall\npicture and qualitative properties.\nThe FM phase is further divided by another two con-\ntinuous transition lines into three different regions: FM0,\nFM1, and FM2. The FM1-FM0 transition line merges\nwith the FM phase boundary (the blue line) at large\n∆x/∆y, so the FM0 phase only exists at small ∆ x/∆y\n(large anisotropy) and has the smallest magnetization\namong the three FM phases. It can be suppressed con-\ntinuously by increasing TK/Jy\nHand becomes a HFL.\nAs shown in Fig. 1b, FM0 may correspond to the\nobserved FM phase of CeRh 6Ge4at ambient pressure,\nwhich also exhibits a severely reduced magnetic moment\n(0.28µB/Ce[14]ascomparedtotheestimated1.28 µB/Ce\nfrom crystalline electric field analysis [32]) and is turned\ncontinuously into a HFL through a FM QCP by apply-\ningpressure[14]. TheleftmostFM2phasehasthe largest\nmagnetization, whichdecreasesgraduallywith increasing\nTK/Jy\nH, while the intermediate FM1 phase is featured\nwith a constant magnetization insensitive to the value\nofTK/Jy\nH. Both FM2-FM1 and FM1-FM0 transitions\nare accompanied by a change of the Fermi surface topol-\nogy. For relatively larger ∆ x/∆y, the FM2-FM1 and\nFM1-HFL transitions (with the light green area collaps-\ning into a first-order transition line in reality) resemble\nthe experimental observations in UGe 2under pressure as\nreproduced in Fig. 1c, in particular the observed Fermi\nsurface changes at the two transitions [20, 21] and the\nmagnetization plateau in between [19]. For very large\n∆x/∆y(inset), where the HFL covers the whole FM1 re-\ngion and penetrates into the FM2 phase, one might find\na first-order transition directly between FM2 and HFL.\nThe FM phase transition in URhAl under pressure may\ncorrespond to this type [33, 34].\nTo see more details on these different FM phases, we\nplotinFig. 2variousquantitiesasfunctionsof TK/Jy\nHfor\ndifferent anisotropy ratios. The holon phase shift is de-\nfined as δχ≡ −1\nV/summationtext\npImln[−Gχ(p,0)−1]. The quantity\nδχ/πmeasuresthe fraction oflocal spins being effectively\nKondo screened, which reaches unity in the HFL phase,\nindicating full Kondo screening. The spinon gap ∆ sis\ndetermined by the renormalized spinon energy at the Γ\npoint, satisfying∆ s−εb\nΓ−ReΣb(Γ,∆s) = 0. In the quasi-\none dimensional case ∆ x/∆y→0,δχ/πdecreases con-\ntinuously from unity at the FM0-HFL transition point,\nmarked by the blue dashed line in Fig. 2a. Right at this4\n/s32/s33/s34\n/s32/s33/s35\n/s32/s33/s36\n/s32/s33/s32/s37/s38/s39/s40/s41/s42/s43/s44/s38/s42/s43/s45/s40\n/s32/s33/s32/s46/s32/s33/s32/s47/s32/s33/s32/s48/s32/s33/s32/s49/s32/s33/s32/s50\n/s51/s52/s53/s54/s55/s56/s32/s32/s33/s34\n/s32/s33/s35\n/s32/s33/s36\n/s32/s33/s32/s37/s38/s39/s40/s41/s42/s43/s44/s38/s42/s43/s45/s40\n/s32/s33/s32/s46/s32/s33/s32/s47/s32/s33/s32/s48/s32/s33/s32/s49\n/s50/s51/s52/s53/s54/s55/s32\n(n b,1 +Knc,1)/N \n(n b,1 +Knc,1 )/N Δx/Δ y=0.03 Δx/Δ y=0.1 /s32/s33/s34\n/s34/s33/s35\n/s34/s33/s36\n/s34/s33/s37\n/s34/s33/s38\n/s34/s33/s34/s39/s40/s41/s42\n/s34/s33/s34/s35 /s34/s33/s34/s43 /s34/s33/s34/s36\n/s44/s45/s41/s46/s47/s48/s32/s34\n/s35\n/s36\n/s37\n/s38\n/s34/s49/s50/s51/s50/s52/s32/s34/s53/s50/s32/s33/s34\n/s34/s33/s35\n/s34/s33/s36\n/s34/s33/s37\n/s34/s33/s38\n/s34/s33/s34/s39/s40/s41/s42\n/s34/s33/s34/s43/s34/s33/s34/s35/s34/s33/s34/s44/s34/s33/s34/s36/s34/s33/s34/s45\n/s46/s47/s41/s48/s49/s50/s32/s34\n/s35\n/s36\n/s37\n/s38\n/s34/s51/s52/s53/s52/s54/s32/s34/s55/s52Δx/Δ y=0.1 Δx/Δ y=0 (a) (b)\n(c) (d)\nFM2 FM1 FM0 FM2 FM1 \nFIG. 2: Holon phase shift and magnetization. (a) The\nholon phase shift ( δχ) and the spinon gap (∆ s) as functions\nofTK/Jy\nHfor ∆ x/∆y= 0. (b) Same plot as (a), but for\n∆x/∆y= 0.1. Multiple solutions exist between the red and\nblue dashed lines. (c) Evolution of the total magnetization\n/angbracketleftSz+sz/angbracketrightand the quantity ( nb,1+Knc,1)/Nwith respect\ntoTK/Jy\nHinside the FM phases. The parameters used are\n∆x/∆y= 0.03,κ= 2S/N = 0.2 andnc/N= 0.5. (d) Same\nplot as (c), but for ∆ x/∆y= 0.1.\npoint, ∆ svanishes and the spin susceptibility diverges\n(see Supplementary Materials), indicating formation of\ntheFMlongrangeorder. WehavethusacontinuousFM-\nHFL transition and a FM QCP as observed in CeRh 6Ge4\n[14]. The situation for ∆ x/∆y= 0.1 is significantly dif-\nferent. Within a certain range of TK/Jy\nHbetween the\nred and blue dashed lines of Fig. 2b, both the HFL and\nFM solutions can exist at zero temperature, so the true\nground state should correspond to a first-order transi-\ntion with a jump of δχ/πfrom 1 to 0 .5 somewhere inside\nthe coexisting region. The constant δχ= 0.5πis asso-\nciated with the half-metal nature of the FM1 phase, as\nwill be explained later. Further reducing TK/Jy\nHdrives\nthe holon phase shift away from the 0 .5πplateau and the\nsystem enters the FM2 phase for small TK/Jy\nH.\nThe distinctions between the three FM phases can\nbe further elucidated by comparing their total magne-\ntization Mz=/angbracketleftSz+sz/angbracketright, shown in Fig. 2c and 2d for\n∆x/∆y= 0.03 and 0 .1, respectively. We see that Mz\ndecreases with increasing TK/Jy\nHin the FM2 phase, but\nkeeps constant in the FM1 phase. In contrast to the wide\ncoexisting region at ∆ x/∆y= 0.1, there exists a stable\nFM0 phase next to the FM1 phase at ∆ x/∆y= 0.03,\nwhereMzfurther decreases with increasing TK/Jy\nH. We\nfind that the constant magnetization of FM1 satisfies\nMz=κ(1−nc/N), where nc=/summationtext\nαnc,αis the electron\nconcentration per channel (see Fig. S3 of SupplementaryMaterials for different κ). As shown in Figs. 2c and 2d,\nour numerical calculations also find another relation of\nthe FM1 phase, nb,1+Knc,1= 2S, which is analytically\nrelated to the magnetization plateau (see Supplementary\nMaterialsSection IV). The physical meaning ofthis iden-\ntity can be understood more directly in the case of N= 2\nandK= 2S= 1, where it becomes nb,↑+nc,↑= 1. This\nleads to the commensurability nc,↑=nb,↓due to the\nconstraint nb,↑+nb,↓= 2S= 1. Since c↑andb↓are\nboth minority components in the FM phase, they can\nfully participate in the Kondo spin-flip scattering and\nthe commensurability implies a full screening of this spin\ncomponent. Meanwhile, the remaining majority compo-\nnents (c↓,b↑) are not involved in the Kondo scattering\nand contribute to the plateau in the total magnetization.\nThese different screening properties have important\nconsequences on their respective band structures, which\ncan be best seen in Fig. 3 showing the evolution of the\nelectron Femi surface and dispersion relation inside the\nFM phases. Due to the hybridization between conduc-\ntion electrons and holons, there exist two quasiparticle\nbands for the minority component ( α= 1), which we de-\nnote asE±(p). Inside the FM2 phase (see Fig. 3a), the\nchemicalpotential intersectswith the E+(p) band, sothe\nFermi surface contains two electron-like pockets. Inside\nthe FM0 phase (see Fig. 3c), the chemical potential in-\ntersectswith the E−(p) band, givingrise to four hole-like\npockets in the Fermi surface. Right in between, the FM1\nphase is featured with an indirect energy gap separating\ntheE+(p) andE−(p) bands, and the chemical potential\nlies just within the gap. We have thus a spin selective\nKondo insulator [7–9, 35], in which the holon phase shift\ncan be calculated from the quasiparticle dispersions via\nδχ/π=1\nV/summationtext\np[1+θ(−εc\np)−θ(−E+(p))−θ(−E−(p))] =\n1\nV/summationtext\npθ(−εc\np), giving the constant δχ/π= 0.5 in Fig. 2b.\nQualitatively, the above evolution of E±(p) can be ex-\nplained by the dispersion relation given by the hybridiza-\ntion between conduction electrons and holons near the\nFermi energy:\nE±(p)≈1\n2/parenleftBig\nεc\np−εχ\n−p±/radicalBig\n(εcp+εχ\n−p)2+4Zχ\n−pb2/parenrightBig\n,(5)\nwhereεχ\np=Zχ\np/J∗\nK(p) is the holon disper-\nsion without spinon condensation, and Zχ\np=\n[−∂ωReΣχ(p,ω)|ω=0]−1>0 is its quasiparticle residue.\nAsTK/Jy\nHincreases, εχ\npgradually evolves from being\nfully above the Fermi energy, to being entirely below\nthe Fermi energy in the HFL phase. This is a general\nfeature of Kondo systems that accounts for the stability\nof forming Kondo bound states [23, 36]. As a result,\n−εχ\n−p, and hence E±(p), evolves in the opposite way,\ncausing the evolution of the Fermi surfaces from electron\nto hole pockets. In addition, because εχ\npbecomes\nincreasingly flat approaching the HFL phase boundary,\nthe quasiparticles become more and more heavy, giving\nrise to the sharp quasiparticle peak at the Fermi energy5\n/s32/s33/s34/s33 /s35/s36/s37\n/s38 /s39 /s38/s40 /s41 /s39\n/s32/s33/s34/s33 /s35/s36/s37\n/s38 /s39 /s38/s40 /s41 /s39\n/s32/s33/s34/s33 /s35/s36/s37\n/s38 /s39 /s38/s40 /s41 /s39\n/s32/s33/s34/s33/s42/s43/s36/s44\n/s32/s33 /s34 /s33\n/s42/s45/s36/s44\n/s32/s33/s34/s33/s42/s43/s36/s44\n/s32/s33 /s34 /s33\n/s42/s45/s36/s44\n/s32/s33/s34/s33/s42/s43/s36/s44\n/s32/s33 /s34 /s33\n/s42/s45/s36/s44/s33/s46/s34 /s34/s46/s47 /s34/s46/s34\n/s48/s49\n/s33/s46/s34 /s34/s46/s47 /s34/s46/s34\n/s48/s49\n/s33/s46/s34 /s34/s46/s47 /s34/s46/s34\n/s48/s49\n/s32/s33/s34/s33/s42/s43/s36/s44\n/s32/s33 /s34 /s33\n/s42/s45/s36/s44\n/s32/s33/s34/s33/s42/s43/s36/s44\n/s32/s33 /s34 /s33\n/s42/s45/s36/s44\n/s32/s33/s34/s33/s42/s43/s36/s44\n/s32/s33 /s34 /s33\n/s42/s45/s36/s44/s33/s46/s34 /s34/s46/s47 /s34/s46/s34\n/s48/s49\n/s33/s46/s34 /s34/s46/s47 /s34/s46/s34\n/s48/s49\n/s33/s46/s34 /s34/s46/s47 /s34/s46/s34/s48/s49\n/s32/s34/s46/s34/s50/s34/s46/s34/s34/s34/s46/s34/s50 /s35/s36/s37\n/s38 /s39 /s38/s40 /s41 /s39\n/s32/s34/s46/s34/s50/s34/s46/s34/s34/s34/s46/s34/s50 /s35/s36/s37\n/s38 /s39 /s38/s40 /s41 /s39\n/s32/s34/s46/s34/s50/s34/s46/s34/s34/s34/s46/s34/s50 /s35/s36/s37\n/s38 /s39 /s38/s40 /s41 /s39ΓXX' M\nE+\nE-(a) \n(b) \n(c)(d) \n(e) \n(f)\nFIG. 3: Fermi surface evolution with TK/Jy\nH.(a)-(c) The conduction electron Fermi surface, dispersion , and density of\nstates for the α= 1 component, at ∆ x/∆y= 0.03 and different TK/Jy\nH. Two quasiparticle bands can be observed, denoted as\nE+andE−. The dashed curve in (b) is a guide to the eye. (d)-(f) Same plo ts as (a)-(c), but for the α= 2,···,Ncomponents.\nA finiteN= 4 is taken to include the conduction electron self-energy i n (d)-(f). The values of TK/Jy\nHare: (a)(d) 0.044 (FM2\nphase); (b)(e) 0.061 (FM1 phase); (c)(f) 0.071 (FM0 phase).\nin the density of states shown in Fig. 3c.\nFor comparison, we also show the Green’s function\nof conduction electrons for other spin components α=\n2,···,N, whose self-energies are the order of O(1/N).\nAs an example, we take N= 4 and show its evolution\nin Figs. 3d-3f with the same values of TK/Jy\nHtaken in\nFigs. 3a-3c. Unlike that of the α= 1 component, the\nFermi surfaces for α= 2,···,Ndo not disappear in the\nFM1phase, revealingthe half-metallicnatureofthe FM1\nphase. On the other hand, we see that their Fermi sur-\nface also expands slightly with TK/Jy\nH, which may be\nattributed to the band bending caused by the precursor\neffect of Kondo resonance in the self-energy. Both Figs.\n3c and 3f suggest that the Kondo effect has partly taken\nplaceintheFM0phase, whichexplainsitshighlyreduced\nmagnetic moment. The FM0-to-HFL phase transition is\ntherefore not of a Kondo-breakdown type. Rather, our\ntheory predicts the coexistence of localized moments and\nhybridization effect within FM0, which may be respon-\nsible for the discrepancy between the measured Fermi\nsurfaces and those from density functional theory (DFT)\ncalculations with fully localized f-electrons [37].\nTo compare with spectroscopic measurements, we plot\nin Fig. 4 the Fermi volume ( VFS) evolving with tem-\nperature above the FM0 ground state. At high tem-\nperatures, the Fermi volume is small and nearly half of\nthe Brillouin zone, consistent with a half-filled conduc-\n/s32/s33/s34/s32/s35/s34\n/s32/s36/s34/s37/s38/s39\n/s37/s38/s40\n/s37/s38/s41/s42/s43/s44\n/s33/s37/s45/s46/s33/s37/s45/s47/s33/s37/s45/s33\n/s48/s49/s50/s45/s51/s52/s32/s48/s34\n/s32/s33/s34/s32/s47/s34\n/s32/s47/s34\nFIG. 4: Fermi volume evolution with temperature.\n(a) Temperature dependences of the electron Fermi volume\n(VFS=1\nV/summationtext\npθ(−˜εc\np)) above the FM0 ground state for\n∆x/∆y= 0.03. The renormalized electron dispersion ˜ εc\npis\ndetermined by the electron Green’s function with self-ener gy\nincluded at N= 4. The dashed line indicates a ln( T) behavior\nofVFSat intermediate temperatures. (b) The electron Fermi\nsurface at two different temperatures marked as (1) and (2)\nin (a).6\ntion band with negligible Kondo effect. As the temper-\nature decreases, the Fermi volume increases logarithmi-\ncally, reaches a maximal value, and then decreases while\napproaching the FM0 ground state. The logarithmic in-\ncrease of the Fermi volume indicates the development of\nKondo hybridization at intermediate temperatures, con-\nsistent with the measurements of ARPES [17] and ultra-\nfast optical spectroscopy [18], while its decrease above\nthe FM0 ground state results from a suppression of the\nKondo hybridization due to enhanced spin fluctuations,\nconfirming the “relocalization” phenomenon observed in\nmany heavy fermion materials [2, 38].\nDISCUSSION AND CONCLUSION\nTo summarize, we have predicted a phase diagram of\nthe anisotropic FM Kondo-Heisenberg model with both\ncontinuous and first-order FM quantum phase transi-\ntions. Our results provide a unified explanation of the\nvery rich and qualitatively different experimental obser-\nvations in two different types of FM Kondo lattice com-\npounds CeRh 6Ge4and UGe 2with extremely large or rel-\natively small magnetic anisotropy, respectively. More\nspecifically, for large anisotropy, we find a continuous\nquantum phase transition between the FM and HFL\nphases, consistent with that observed in the quasi-one-\ndimensional CeRh 6Ge4. As the magnetic anisotropy is\nreduced, we find a coexisting region of FM and HFL, im-\nplying afirst-orderFM-HFL phase transitionasobserved\nin UGe 2. In addition, there exist three different types of\nFM ordered phases, among which the intermediate FM1\nphase shows half-metallic property and is identified as\na spin selective Kondo insulator. Our findings suggest\nthe important role of magnetic anisotropy on interpret-\ning the experimental observations in FM Kondo lattice\nmaterials.\nIt should be noted, however, that our model studies\nhaveignoredsomepeculiarmicroscopicdetailsinrealistic\nmaterials. Forexample,CeRh 6Ge4hasastrongmagnetic\neasy-planeanisotropy,whichcorrespondstoanXXZtype\nHeisenberg interaction [14]. This easy-plane anisotropy\nmight favor a triplet resonating-valence-bond state [39],\ncause a small discontinuity of the holon phase shift at\nthe FM QCP, and further reduce the magnetic moment\nin the FM phase [14]. In UGe 2, the strong spin-orbit\ncoupling can result in FM fluctuations of Ising character,\nwhich may be important for its superconducting proper-\nties [13]. In a more realistic study, one may also want\nto consider the variation of conduction electron concen-\ntration with external parameters, which might tune the\nFM1-FM2 transition into a first-order one [40] or alter\nthe sign ofthe RKKYinteraction[41]. These details may\nresult in some additional interesting features but will not\nchange our overall theoretical picture. It is straightfor-\nward to extend our theory to include these details infuture studies.\nCONFLICT OF INTEREST\nThe authors declare that they have no conflict of in-\nterest.\nACHNOWLEDGMENTS\nThis work was supported by the National Key Re-\nsearch and Development Program of China (Grant No.\n2017YFA0303103), the National Natural Science Foun-\ndation of China (Grants No. 12174429, No. 11774401,\nand No. 11974397), and the Strategic Priority Research\nProgram of the Chinese Academy of Sciences (Grant No.\nXDB33010100).\nAUTHOR CONTRIBUTIONS\nYang YF conceived and supervised the project. Wang\nJF performed the calculations. Wang JF and Yang YF\nwrote the manuscript.\n∗yifeng@iphy.ac.cn\n[1] Si Q, Steglich F. Heavy Fermions and Quantum Phase\nTransitions. Science 2010;329:1161-1166.\n[2] Yang YF, Pines D, Lonzarich G. Quantum critical scaling\nand fluctuations in Kondo lattice materials. Proc. Natl.\nAcad. Sci. U.S.A. 2017;114:6250-6255.\n[3] Wirth S, Steglich, F. Exploring heavy fermions from\nmacroscopic to microscopic length scales. Nat. Rev.\nMater. 2016;1:16051.\n[4] Coleman P. Heavy fermions and the Kondo lattice: a 21st\ncentury perspective. arXiv:1509.05769.\n[5] Komijani Y, Coleman P. Model for a Ferromagnetic\nQuantum Critical Point in a 1D Kondo Lattice. Phys.\nRev. Lett. 2018;120:157206.\n[6] Yamamoto SJ, Si Q. Metallic ferromagnetism in\nthe Kondo lattice. Proc. Natl. Acad. Sci. U.S.A.\n2010;107:15704-15707.\n[7] Li GB, Zhang GM, Yu L. Kondo screening coexisting\nwith ferromagnetic order as a possible ground state for\nKondo lattice systems. Phys. Rev. B 2010;81:094420.\n[8] Peters R, Kawakami N, Pruschke T. Spin-Selective\nKondo Insulator: Cooperation of Ferromagnetism and\nthe Kondo Effect. Phys. Rev. Lett. 2012;108:086402.\n[9] Goleˇ z D, ˇZitko R. Lifshitz phase transitions in the ferro-\nmagnetic regime of the Kondo lattice model. Phys. Rev.\nB 2013;88:054431.\n[10] Brando M, Belitz D, Grosche FM, et al. Metallic quan-\ntum ferromagnets. Rev. Mod. Phys. 2016;88:025006.\n[11] Belitz D, Kirkpatrick TR, Vojta T. First Order Transi-\ntions and Multicritical Points in Weak Itinerant Ferro-\nmagnets. Phys. Rev. Lett. 1999;82:4707-4710.7\n[12] Belitz D, Kirkpatrick TR. Effective soft-mode the-\nory of strongly interacting fermions. Phys. Rev. B\n2012;85:125126.\n[13] Aoki D, Ishida K, Flouquet J. Review of U-based Ferro-\nmagnetic Superconductors: Comparison between UGe 2,\nURhGe, and UCoGe. J. Phys. Soc. Jpn. 2019;88:022001.\n[14] Shen B, Zhang YJ, Komijani Y, et al. Strange-metal be-\nhaviour in a pure ferromagnetic Kondo lattice. Nature\n2020;579:51-55.\n[15] Steppke A, K¨ uchler R, Lausberg S, et al. Ferromagnetic\nQuantum Critical Point in the Heavy-Fermion Metal\nYbNi4(P1−xAsx)2. Science 2013;339:936.\n[16] Kirkpatrick TR, Belitz D. Ferromagnetic Quantum Crit-\nical Point in Noncentrosymmetric Systems. Phys. Rev.\nLett. 2020;124:147201.\n[17] Wu Y, Zhang YJ, Du F, et al. Anisotropic c-fHy-\nbridization in the Ferromagnetic Quantum Critical Metal\nCeRh 6Ge4. Phys. Rev. Lett. 2021;126:216406.\n[18] Pei YH, Zhang YJ, Wei ZX, et al. Unveiling the hy-\nbridization process in a quantum critical ferromag-\nnet by ultrafast optical spectroscopy. Phys. Rev. B\n2021;103:L180409.\n[19] Pfleriderer C, Huxley AD. Pressure Dependence of\nthe Magnetization in the Ferromagnetic Superconductor\nUGe2. Phys. Rev. Lett. 2002;89:147005.\n[20] Terashima T, Matsumoto T, Terakura C, et al. Evolu-\ntion of Quasiparticle Properties in UGe 2with Hydro-\nstatic Pressure Studied via the de Haas–van Alphen Ef-\nfect. Phys. Rev. Lett. 2001;87:166401.\n[21] Haga Y, Nakashima M, Settai R, et al. A change of\nthe Fermi surface across the metamagnetic transition\nunder pressure in UGe 2. J. Phys. Condens. Matter\n2002;14:L125-L135.\n[22] Oikawa K, Kamiyama T, Asano H, et al. Crystal Struc-\nture of UGe 2. J. Phys. Soc. Jpn. 1996;65:3229-3232.\n[23] Wang JF, Yang YF. Nonlocal Kondo effect and quan-\ntum critical phase in heavy-fermion metals. Phys. Rev.\nB 2021;104:165120.\n[24] Zhao HC, Zhang JH, Lyu M, et al. Quantum-critical\nphase from frustrated magnetism in a strongly correlated\nmetal. Nat. Phys. 2019;15:1261-1266.\n[25] Hill HH. Plutonium 1970 and Other Actinides. New York:\nAIME; 1970.\n[26] Rech J, Coleman P, Zarand G, Parcollet O. Schwinger\nBoson Approach to the Fully Screened Kondo Model.\nPhys. Rev. Lett. 2006;96:016601.\n[27] Lebanon E, Rech J, Coleman P, Parcollet O. ConservingMany Body Approach to the Infinite-U Anderson Model.\nPhys. Rev. Lett. 2006;97:106604.\n[28] Coleman P, Paul I, Rech J. Sum rules and Ward identities\nin the Kondo lattice. Phys. Rev. B 2005;72:094430.\n[29] Komijani Y, Coleman P. Emergent Critical Charge Fluc-\ntuations at the Kondo Breakdown of Heavy Fermions.\nPhys. Rev. Lett. 2019;122:217001.\n[30] Wang JF, Chang YY, Mou CY, et al. Quantum\nphase transition in a two-dimensional Kondo-Heisenberg\nmodel: A dynamical Schwinger-boson large-N approach.\nPhys. Rev. B 2020;102:115133.\n[31] Silva TND, Ma M, and Zhang FC. Pathology of\nSchwinger boson mean-field theory for Heisenberg spin\nmodels. Phys. Rev. B 2002;66:104417.\n[32] Shu JW, Adroja DT, Hillier AD, et al. Magnetic or-\nder and crystalline electric field excitations of the quan-\ntum critical heavy fermion ferromagnetic CeRh 6Ge4.\narXiv:2102.12788.\n[33] Combier T. Ferromagnetic quantum criticality in the\nuranium-based ternary compounds URhSi, URhAl, and\nUCoAl. Ph. D thesis: Universit´ e de Grenoble; 2014.\n[34] Shimizu Y, Braithwaite D, Salce B, et al. Unusual strong\nspin-fluctuation effects around the critical pressure of the\nitinerant Ising-type ferromagnet URhAl. Phys. Rev. B\n2015;91:115115.\n[35] Peters R, Kawakami N. Ferromagnetic state in the\none-dimensional Kondo lattice model. Phys. Rev. B\n2012;86:165107.\n[36] Wang JF, Yang YF. Spin current Kondo effect in frus-\ntrated Kondo systems. Sci. China Phys. Mech. Astron.\n2022;65:227212.\n[37] Wang A, Du F, Zhang YJ, et al. Localized 4f-electrons\nin the quantum critical heavy fermion ferromagnet\nCeRh 6Ge4. Sci. Bull. 2021;66:1389-1394.\n[38] Shirer KR, Shockley AC, Dioguardi AP, et al. Long range\norder and two-fluid behavior in heavy electron materials.\nProc. Natl. Acad. Sci. U.S.A. 2012;109:E3067-E3073.\n[39] Coleman P, Komijani Y, K¨ onig EJ. Triplet Resonating\nValence Bond State and Superconductivity in Hund’s\nMetals. Phys. Rev. Lett. 2020;125:077001.\n[40] Wysoki´ nski MM, Abram M, Spa/suppress lek J. Criticalities\nin the itinerant ferromagnet UGe 2. Phys. Rev. B\n2015;91:081108.\n[41] Bernhard BH, Lacroix C. Coexistence of magnetic order\nand Kondo effect in the Kondo-Heisenberg model. Phys.\nRev. B 2015;92:094401.arXiv:2112.14515v1 [cond-mat.str-el] 29 Dec 2021A unified theory of ferromagnetic quantum phase transitions in heavy fermion metals\n- Supplementary Materials -\nJiangfan Wang1and Yi-feng Yang1,2,3\n1Beijing National Laboratory for Condensed Matter Physics,\nInstitute of Physics, Chinese Academy of Science, Beijing 1 00190, China\n2School of Physical Sciences, University of Chinese Academy of Sciences, Beijing 100190, China\n3Songshan Lake Materials Laboratory, Dongguan, Guangdong 5 23808, China\nI. Self-consistent equations\nWithin the SU( N) Schwinger boson representation, the anisotropic ferromagnet ic Kondo lattice model (Eq. (1) in\nthe main text) has the following Lagrangian:\nL=/summationdisplay\npaαc†\npaα(∂τ+εc\np)cpaα+/summationdisplay\npαb†\npα(∂τ+εb\np)bpα+/summationdisplay\npa|χpa|2\nJK\n+1√\nVN/summationdisplay\npkaαb†\npαckaαχp−k,a+c.c.+VN/parenleftBigg\n∆2\nx\nJx\nH+∆2\ny\nJy\nH/parenrightBigg\n−2VλS, (S1)\nwhereεb\np=λ−2(∆xcospx+∆ycospy) is the bare spinon dispersion, and λis the Lagrange multiplier associated with\nthe spinon constraint. The τ-dependence of each field is implied. In the paramagnetic phase, the Green’s functions\nof spinon, holon and conduction electron are:\nGb(p,iνn) =1\niνn−εbp−Σb(p,iνn),\nGχ(p,iωn) =1\n−J−1\nK−Σχ(p,iωn),\nGc(p,iωn) =1\niωn−εcp−Σc(p,iωn). (S2)\nBoth Σ band Σ χare order of unity, while Σ cis order of 1 /N:\nΣc(p,iωn) =1\nN/summationdisplay\nkmGχ(k−p,iνm−iωn)Gb(k,iνm). (S3)\nTherefore, one can ignore Σ cin the large- Nlimit, and simply use the bare form of Gcin the self-energy equations (Eq.\n(3) in the main text). At finite- N, one needs to include Σ cto calculate the renormalized dispersion and Fermi surface\nof conduction electrons. Minimizing the free energy with respect to λand ∆ δgives rise to the following constraints:\nκ=−1\nβV/summationdisplay\npnGb(p,iνn),\n∆δ\nJδ\nH=−1\nβV/summationdisplay\npnGb(p,iνn)cospδ. (S4)\nIn practice, we choose different values of ∆ δas input parameters, and determine the value of Jδ\nHusing the second\nconstraint of Eq. (S4) after self-consistency has been achieved .\nTo study the FM ordered phase, we allow the spinons to have a static uniform condensation on the α= 1 spin\ncomponent:\nbiα(τ)→/braceleftbigg√\nNb α= 1\nbiα(τ)α= 2,···,N. (S5)2\nThe Lagrangian can then be separated into two parts, L=L1+LN−1, with\nL1=/summationdisplay\npa(c†\npa,1,χ−p,a)/parenleftbigg∂τ+εc\np−b\n−b−J−1\nK/parenrightbigg/parenleftbiggcpa,1\nχ†\n−p,a/parenrightbigg\n+VNb2(λ−2(∆x+∆y)),\nLN−1=/summationdisplay\npa,α>1c†\npaα(∂τ+εc\np)cpaα+/summationdisplay\np,α>1b†\npα(∂τ+εb\np)bpα+1√\nVN/summationdisplay\npka,α>1b†\npαckaαχp−k,a+c.c.\n+VN/parenleftBigg\n∆2\nx\nJx\nH+∆2\ny\nJy\nH/parenrightBigg\n−2VλS. (S6)\nThe condensation provides a hybridization field for the holons and co nduction electrons of the α= 1 spin flavor, with\nthe following Green’s functions:\nGχ(p,iωn) =1\n−J−1\nK−Σχ(p,iωn)+b2\n−iωn−εc\n−p,\nGα=1\nc(p,iωn) =1\niωn−εcp+b2\n−J−1\nK−Σχ(−p,−iωn). (S7)\nThe Green’s functions of the Schwinger boson and conduction elect ron of the α >1 spin flavors remain the same as\nin Eq. (S2). Minimizing the free energy with respect to λ, ∆δandbgives the following constraints:\nκ=b2−N−1\nN1\nβV/summationdisplay\npnGb(p,iνn),\n∆δ\nJδ\nH=b2−N−1\nN1\nβV/summationdisplay\npnGb(p,iνn)cospδ,\n0 =λ−2(∆x+∆y)+ReΣ b(0,0), (S8)\nwhere (N−1)/N→1 in the large- Nlimit. The last constraint requires that the renormalized spinon spec trum\ntouches zero energy at the Γ point, consistent with our assumptio n of a uniform condensation. Again, only the first\nand the third constraints are solved self-consistently, while the se cond constraint is used to find Jδ\nHat the end of\ncalculation.\nII. The effects of biquadratic spin interaction term\nIn order to alleviate the artificial first-order transition of the Sch winger boson mean-field theory, we introduce a\nbiquadratic spin interaction term into the Hamiltonian, so that the “H eisenberg” term becomes\nHH=−/summationdisplay\niδJδ\nH/bracketleftbig\nSi·Si+δ+ζ(Si·Si+δ)2/bracketrightbig\n=−/summationdisplay\niδJδ\nH(1−ζ/2)Si·Si+δ+C, (S9)\nwhere we have used the identity ( Si·Si+δ)2= 3/16−(Si·Si+δ)/2, andCis a constant. The condition that the\neffective Heisenberg coupling Jδ\nH(1−ζ/2) remains ferromagnetic requires ζ <2. At large- N, the Heisenberg term is\nwritten as\nHH=−/summationdisplay\niδJδ\nH\nN/parenleftbigg\nTiδ+2ζ\nN2T2\niδ/parenrightbigg\n, (S10)\nwhereTiδ=/summationtext\nαβSαβ\niSβα\ni+δ, andSαβ\ni=b†\niαbiβ−κδαβis the SU( N) generator. It reduces to Eq. (S9) in the case of\nN= 2. Using the mean-field expansion Tiδ=/angbracketleftTiδ/angbracketright+δTiδ=/angbracketleftTiδ/angbracketright+(Tiδ−/angbracketleftTiδ/angbracketright), one has\nT2\niδ=−/angbracketleftTiδ/angbracketright2+2/angbracketleftTiδ/angbracketrightTiδ+O(δT2\niδ), (S11)3\n/s32/s33/s34/s35/s32/s33/s34/s36/s32/s33/s34/s32/s32/s33/s32/s37/s32/s33/s32/s38/s32/s33/s32/s35/s32/s33/s32/s36/s32/s33/s32/s32/s39/s40/s41/s42/s40/s43\n/s32/s33/s34/s32/s32/s33/s32/s44/s32/s33/s32/s37/s32/s33/s32/s45/s32/s33/s32/s38/s32/s33/s32/s46\n/s39/s47/s48/s42/s49/s50/s43/s39/s39/s51/s52/s32\n/s39\n/s39/s39/s51/s52/s36\n~\nFIG. S1: Effects of the biquadratic term. The phase diagram with ( ζ= 2) or without ( ζ= 0) the biquadratic term. The\nshadowed areas denote the coexisting regions.\nso that\nHH=−/summationdisplay\niδJδ\nH\nN/parenleftbigg\n1+4ζ\nN2/angbracketleftTiδ/angbracketright/parenrightbigg\nTiδ+C, (S12)\nwhereCis a constant. The above equation defines a modified Heisenberg cou pling\n˜Jδ\nH≡Jδ\nH/parenleftbigg\n1+4ζ\nN2/angbracketleftTiδ/angbracketright/parenrightbigg\n=Jδ\nH/parenleftBigg\n1+4ζ\n(˜Jδ\nH)2∆2\nδ/parenrightBigg\n, (S13)\nwhich suppressesthe artificialfirst-ordertransition. Fig. S1sho wsthe phasediagraminterms ofthe modified Doniach\nratioTK/˜Jy\nHforζ= 0 and ζ= 2. The size of the coexisting region decreases as one includes the b iquadratic term,\nand the tricritical point is shifted slightly from ∆ x/∆y= 0.02 to 0.03.\nIII. Quantum criticality in the 1D limit\nIn Fig. S2, we showphysical quantities asfunctions of T/TKat ∆x= 0 and different ∆ y. The uniform susceptibility\nof impurity spin is calculated as\nχu=2\nV/summationdisplay\nk/integraldisplaydω\nπ1\neβω−1ImGb(k,ω)ReGb(k,ω). (S14)\nIt follows the Curie law T−1at high temperatures, becomes constant in the heavy Fermi liquid p hase (red) at low\ntemperatures, and diverges with another power law T−γon the ferromagnetic side of the QCP (blue). The QCP\nlocates at ∆ y= 0.014 (orTK/Jy\nH= 0.067), with a critical power index γ≈0.67. The power index γincreases with\n∆yon the FM side due to the weakening of Kondo screening effect, as su ggested by the decreasing holon phase shift.\nIV. Derivation of some identities in FM1\nIn the SU( N) representation, we define the impurity spin polarization along the zdirection as\nMS\nz≡1\nV/summationdisplay\ni/angbracketleftSz(i)/angbracketright ≡1\nVN/summationdisplay\ni\n/summationdisplay\nα≤N/2/angbracketleftBig\nb†\niαbiα/angbracketrightBig\n−/summationdisplay\nα>N/2/angbracketleftBig\nb†\niαbiα/angbracketrightBig\n. (S15)4\n~T -1 ~T -γ\nFIG. S2: Quantum criticality at 1D limit. Temperature dependence of (a) the holon phase shift and (b) t he uniform\nimpurity spin susceptibility at ∆ x= 0 and different ∆ y. The black solid line indicates T−1Curie behavior at high temperature,\nwhile the dashed lines indicate T−γdivergence of the susceptibility at low temperature.\nThisdefinitioncorrectlygives MS\nz= 0intheparamagneticphase,sincethespinonconcentration nb,α≡1\nV/summationtext\ni/angbracketleftBig\nb†\niαbiα/angbracketrightBig\nis identical for all spin flavors. In the FM phase, the α= 1 spin flavor develops condensation, while all the other spin\nflavors remain equivalent. This gives us\nMS\nz=1\nN(nb,1−nb,α/negationslash=1) =1\nN(Nb2−O(1)) =b2+O(1/N). (S16)\nThe spin polarization of conduction electrons can be defined in the sa me way\nms\nz≡1\nVN/summationdisplay\nia\n/summationdisplay\nα≤N/2/angbracketleftBig\nc†\niaαciaα/angbracketrightBig\n−/summationdisplay\nα>N/2/angbracketleftBig\nc†\niaαciaα/angbracketrightBig\n\n=K\nN(nc,1−nc,α/negationslash=1), (S17)\nwherenc,α≡1\nV/summationtext\ni/angbracketleftBig\nc†\niaαciaα/angbracketrightBig\nis the electron concentration of spin flavor α, which is independent of the channel\nindex. As discussed in the main text, the FM1 phase also satisfies ano ther identity, nb,1+Knc,1= 2S, which results\ninKnc,1= (N−1)nb,α/negationslash=1when combined with the constraint/summationtext\nαnb,α= 2S. This consistently leads to the plateau\nof the total magnetization:\nMz≡MS\nz+ms\nz=nb,1+Knc,1−nb,α/negationslash=1−Knc,α/negationslash=1\nN\n=2S−K\nN−1(nc,1+(N−1)nc,α/negationslash=1)\nN\n=2S−K\nN−1nc\nN, (S18)\nwherenc=/summationtext\nαnc,αis the total electron concentration per channel. In the large- Nlimit, Eq. (S18) reduces to\nMz=κ(1−nc/N).\nWe now derive the following relation in the FM phases:\nδχ\nπ=1\nV/summationdisplay\np[1+θ(−εc\np)−θ(−E+(p))−θ(−E−(p))]. (S19)\nUsing Eq. (S7), we have\nln[−Gχ(p,ω+)−1] = ln[(J−1\nK+Σχ(p,ω+))(ω++εc\n−p)+b2]−ln[ω++εc\n−p], (S20)\nwhereω+≡ω+i0+. Near the Fermi energy, one has\nΣ′\nχ(p,ω+) = Σ′\nχ(p,0+)+∂ωΣ′\nχ(p,ω+)|0ω, (S21)5\n(a) (b) \nFIG. S3: Effects of κ.(a) Comparison of the coexisting region (light green areas) at different values of κ. The dashed lines are\nextrapolations. (b) The total magnetization and the quanti tynb,1+Knc,1as functions of TK/Jy\nHat ∆x/∆y= 0.1,nc/N= 0.5\nand different values of κ.\nwhere Σ′\nχ(Σ′′\nχ) denotes the real (imaginary) part of self-energy. Defining the q uasiparticle residue Zχ\np≡\n[−∂ωΣ′\nχ(p,ω+)|0]−1and the holon’s dispersion relation εχ\np≡Zχ\np(J−1\nK+Σ′\nχ(p,0)), we have\nln[−Gχ(p,ω+)−1] = ln[(ω−εχ\np−iZχ\npΣ′′\nχ(p,ω+))(ω++εc\n−p)−Zχ\npb2]−ln[ω++εc\n−p]−ln[−Zχ\np]. (S22)\nAt low temperature, one has lim ω→0Σ′′\nχ(p,ω) = 0−andZχ\np>0, therefore lim ω→0(−iZχ\npΣ′′\nχ(p,ω+)) =i0+. The pole\nof the holon Green’s function determines the quasiparticle spectru m:\n0 = (ω−εχ\np)(ω+εc\n−p)−Zχ\npb2\n→ω=εχ\np−εc\n−p±/radicalBig\n(εχ\np+εc\n−p)2+4Zχ\npb2\n2=−E∓(−p). (S23)\nNote that E±(p) is defined as the quasiparticle pole of Gα=1\nc(p,ω), which is exactly the pole of Gχ(p,ω) after the\nparticle-hole transformation p→ −p,ω→ −ω. Therefore, we have\nδχ\nπ=−1\nπV/summationdisplay\npImln[−Gχ(p,0+)−1]\n=−1\nπV/summationdisplay\np/parenleftbig\nImln[E+(p)+i0+]+Imln[ E−(p)+i0+]−Imln[εc\np+i0+]/parenrightbig\n+1\n=1\nV/summationdisplay\np[1+θ(−εc\np)−θ(−E+(p))−θ(−E−(p))], (S24)\nwhere we have assumed the inversion symmetry E±(p) =E±(−p),εc\np=εc\n−p.\nV. The effects of κ\nThe influence of κon the phase diagram is shown in Fig. S3(a). For smaller κ, the FM phases are suppressed by the\nHFL phase, so that the coexisting region becomes narrower and is s hifted towards smaller values of TK/Jy\nH. Roughly\nspeaking, this is because a small value of κenhances the quantum zero motion of spin, thus suppresses the F M order\nand favors the Kondo effect. We also find that the tricritical point s hifts to larger values of ∆ x/∆yasκbecomes\nlarger. Fig. S3(b) compares the plateaus of magnetization and the quantity ( nb,1+Knc,1)/Nat different values of κ,\nwith ∆ x/∆y= 0.1 andnc/N= 0.5. One can see that the two relations Mz=κ(1−nc/N) and (nb,1+Knc,1)/N=κ\nalways hold inside the FM1 phase." }, { "title": "0707.0029v1.Zooming_into_the_coexisting_regime_of_ferromagnetism_and_superconductivity_in_ErRh4B4_single_crystals.pdf", "content": "arXiv:0707.0029v1 [cond-mat.str-el] 30 Jun 2007Zooming into the coexisting regime of ferromagnetism and su perconductivity in\nErRh 4B4single crystals\nRuslan Prozorov,∗Matthew D. Vannette, Stephanie A. Law, Sergey L. Bud’ko, and Pa ul C. Canfield\nAmes Laboratory and Department of Physics & Astronomy, Iowa State University, Ames, IA 50011\n(Dated: 29 June 2007)\nHigh resolution measurements of the dynamic magnetic susce ptibility are reported for ferro-\nmagnetic re-entrant superconductor, ErRh 4B4. Detailed investigation of the coexisting regime\nreveals unusual temperature-asymmetric and magnetically anisotropic behavior. The supercon-\nducting phase appears via a series of discontinuous steps up on warming from the ferromagnetic\nnormal phase, whereas the ferromagnetic phase develops via a gradual transition. A model based\non local field inhomogeneity is proposed to explain the obser vations.\nPACS numbers: 74.25.Dw; 75.50.Cc; 74.25.Ha; 74.25.-q; 74. 90.+n\nThe coexistence of the long-range magnetic order and\nsuperconductivity was first discussed even before the ap-\npearance of the microscopic theory of superconductivity\n[1]. Since then this topic remains one of the most inter-\nesting and controversial in the physics of superconduc-\ntors with many reviews and books devoted to the subject\n[2, 3, 4, 5, 6]. Despite significant effort in new materi-\nals design and discovery, there are only few, confirmed,\nferromagnetic superconductors. Local, full-moment fer-\nromagnetic superconductors: ErRh 4B4[7] (TFM≈0.9\nK,Tc≈8.7K), Ho xMo6S8[8] (TFM≈0.7 K,Tc≈1.8\nK); the weakly ferromagnetic ErNi 2B2C [9] (TFM≈2.3\nK,TAFM≈6 K,Tc≈11 K); and more recent itin-\nerant superconducting ferromagnets, UGe 2(TFM≈33\nK,Tc≈0.95 K at P≈1.3 GPa) [10] and URhGe\n(TFM≈9.5 K,Tc≈0.27 K). Whereas all these ma-\nterials are very interesting on their own, the coexistence\nof growing, large, local moment ferromagnetism and su-\nperconductivity is most clearly presented in ErRh 4B4,\nwhich is the subject of the present Letter. In particular,\nwe are interested in the details of the narrow tempera-\nture interval ( ∼0.3 K) where the two phases coexist and\ninfluence each other.\nErRh4B4was extensively studied over past 30 years\n[2, 3, 4, 5, 6]. The ferromagnetic phase is primitive\ntetragonal with the c−axis being the hard and a−axis\nbeing the easy magnetic axes. Detailed measurements of\nanisotropic magnetization and upper critical filed, Hc2\nwere done by Crabtree et al.[11, 12], who found that\nHa\nc2(alonga−axis) peaks at 5.5 K due to large param-\nagnetic spin susceptibility in that direction [3, 11]. In the\ncontrast, Hc\nc2collapses near the onset of the long-range\nferromagnetic order.\nNeutron diffraction studies have established the ex-\nistence of a modulated ferromagnetic structure at the\nlengthscale of ∼10 nm [13, 14]. In single crystals, results\nsuggested that coexisting phases consists of a mosaic of\nnormal FM domains and SC regions larger than ∼200\nnm in size. The SC regions contain modulated FM mo-\nment with a period of ∼10 nm. These regions could be\nregular domains, spontaneous vortex lattices or laminarstructures with ≥200 nm periodicity and modulated SC\ndomains in between [14]. Thermal hysteresis is observed\nboth in the normal Bragg peak intensity and the small-\nangle peaks. For the small-angle peaks, the intensity is\nhigher on cooling than on warming. This is opposite to\nthe behavior of the regular Bragg peaks from the FM\nregions [13]. Furthermore, the first-order transition, ob-\nserved in satellite peaks temperature dependence [14], is\nconsistent with the spiralstate of Blount and Varma [15].\nHowever, a continuos transition was reported in other\nneutron diffraction [13, 16] and specific heat experiments\n[17]. Such a transition can be realized in a modulated\nstructure or via spontaneous vortex phase.\nTheoretically, some striking features of the coexist-\ning phase include an inhomogeneous, spiral, FM struc-\nture [15, 18] or a fine domain, ”cryptoferromagnetic”\nphase [2, 19], a vortex - lattice modulated spin struc-\nture [20], type-I superconductivity [2, 20, 21], a gapless\nregime and possibly, an inhomogeneous Fulde-Ferrell-\nLarkin-Ovchinnikov (FFLO) state [2]. Another interest-\ning possibility is the development ofsuperconductivity at\nthe ferromagnetic domain walls [22, 23].\nIn this Letter we report precision measurements of the\ndynamic magnetic susceptibility of ErRh 4B4with an em-\nphasisonthenarrowtemperatureregionwhereferromag-\nnetism and superconductivity coexist. We find that the\ntransition is highly asymmetric when FM →SC (heat-\ning) and SC →FM (cooling) data are compared. The\nFM↔SC transition proceeds via a series of discrete steps\nfrom FM to SC phase upon warming and proceeds via a\nsmooth crossover from the SC to FM state upon cooling.\nWith this new information we analyze relevance of some\npredictions made over years for the coexisting phase.\nSingle crystalsof ErRh 4B4were grownat high temper-\natures from molten copper flux as described in [24, 25].\nResulting samples were needle shaped with crystallo-\ngraphicc-axisalongthe needle. Transportmeasurements\ngave residual resistivity ratio of about 8, consistent with\nprevious reports. The anisotropic Hc2(T) curves (see in-\nset to Fig.3 below) are consistent with earlier reports as\nwell [11, 12].2\n/s48 /s49 /s50 /s51 /s52 /s53 /s54 /s55 /s56 /s57/s45/s49/s46/s48/s45/s48/s46/s56/s45/s48/s46/s54/s45/s48/s46/s52/s45/s48/s46/s50/s48/s46/s48/s48/s46/s50/s32/s52\n/s84/s32/s40/s75/s41/s72/s32/s61/s32/s48/s72/s61/s52/s48/s48/s48/s32/s79/s101/s72/s32 /s32/s99\nFIG. 1: Dynamic magnetic susceptibility in ErRh 4B4single\ncrystal measured along the magnetic easy axis (perpendicu-\nlar to the needle-shaped sample). Each curve corresponds to\na fixed value of the applied dc magnetic field in the range\nindicated in the figure. (color online)\nTheACmagnetic susceptibility, χ, was measured with\na tunnel-diode resonator (TDR) which is sensitive to\nchanges in susceptibility ∆ χ∼10−8. Details of the mea-\nsurement technique are described elsewhere [26, 27, 28].\nIn brief, properly biased tunnel diode compensates for\nlosses in the tank circuit, so it is self-resonating on its\nresonant frequency, ω= 1/√\nLC∼10 MHz. A sample\nis inserted into the coil on a sapphire rod. The effective\ninductance changes and this causes a change in the res-\nonant frequency. This frequency shift is the measured\nquantity and it is proportional to the sample dynamic\nmagnetic susceptibility, χ[26, 27, 28]. Knowing geomet-\nricalcalibrationfactorsofourcircuit, weobtain χ(T,H).\nAdvantages of this technique are: very small AC excita-\ntion field amplitude ( ∼20 mOe), which means that it\nonly probes, but does not disturb the superconducting\nstate; high stability and excellent temperature resolution\n(∼1 mK), which allowed detailed study of the coexist-\ning region, which is only ∼500 mK wide. Normal-state\nskin depth is largerthan the sample size, so we probe the\nentire bulk in the coexisting region, but when supercon-\nducting phase becomes dominant, there is a possibility\nthat some FM patches still exist, but are screened.\nFigure1showsthefulltemperaturescalemagneticsus-\nceptibility, χ, in single crystal ErRh 4B4for an applied\nfield oriented along the easy axis and perpendicular to\nthe needle-shaped crystal. The peak in χ(T) at the fer-\nromagnetic to superconducting boundary below 1 K is/s48/s46/s57 /s49/s46/s48 /s49/s46/s49 /s49/s46/s50/s45/s48/s46/s56/s45/s48/s46/s54/s45/s48/s46/s52/s45/s48/s46/s50/s48/s46/s48/s72/s32 /s32/s99/s52\n/s84/s32/s40/s75/s41/s72/s32/s61/s32/s48/s32/s45/s32/s51/s32/s107/s79/s101/s72/s32/s61/s32/s48\nFIG. 2: Magnetic susceptibility in a ferromagnet-\nsuperconductor transition region measured at different ap-\nplied fields. Note the temperature scale and highly assymet-\nric character of the FM-SC and SC-FM transitions. (color\nonline). Blue - warming, red - cooling.\na common feature observed in local moment ferromag-\nnets [29]. Clearly, superconductivity is fullysuppressed\nin the ferromagnetic phase. Note that at elevated fields,\nthe response is nonmonotonic on the SC side close, to\nthe FM boundary, indicative of enhanced diamagnetism\n(larger, negative χ), which may be due to suppressed\nmagnetic pairbreaking or entering into another phase,\nsuch as FFLO [2].\nFigure 2 zooms into the SC ↔FM transition region.\nMeasurements were taken after zero-field cooling, apply-\ning external field and warming up (ZFC-W) above Tc\nand then cooling back to the lowest temperature (FC-C).\nThere is striking asymmetry of the transitions - when the\nsuperconducting phase develops out of the FM state, the\nresponse proceeds with jumps in the susceptibility, which\nare clearly associated with the appearance of supercon-\nducting regions of finite size. The steps are present up\nto the largest field at which superconductivity survives.\nDecreasing temperature shows a completely different re-\nsult: the transition is smooth and gradual and proceeds\nto lower temperatures.\nTo better understand the dynamics of the transition,\nthe top frame of Fig. 3 shows measurements at H= 0 for\ndifferent temperature ramp rates. Temperature variation\nis shown in the inset. These data clearly demonstrate\nthatthishysteresisisinsensitivetoheating/coolingrates.\nIt should be noted that all the other data presented in\nthis Letter were taken with the slowest cooling rate of 60\nµK/s.\nSimilar hysteresis and steps are also present in the an-3\n/s48/s46/s57 /s49/s46/s48 /s49/s46/s49 /s49/s46/s50/s45/s49/s46/s48/s45/s48/s46/s56/s45/s48/s46/s54/s45/s48/s46/s52/s45/s48/s46/s50/s48/s46/s48/s48/s46/s50/s48/s46/s52/s52\n/s84/s32/s40/s75/s41/s72/s32/s61/s32/s48/s48 /s49/s48/s48/s48/s48 /s50/s48/s48/s48/s48 /s51/s48/s48/s48/s48 /s52/s48/s48/s48/s48/s48/s46/s52/s48/s46/s54/s48/s46/s56/s49/s46/s48/s49/s46/s50/s50/s120/s49/s48/s45/s51\n/s32/s75/s47/s115\n/s49/s120/s49/s48/s45/s51\n/s32/s75/s47/s115\n/s50/s120/s49/s48/s45/s52\n/s32/s75/s47/s115\n/s54/s120/s49/s48/s45/s53\n/s32/s75/s47/s115/s84/s32/s40/s75/s41\n/s116/s105/s109/s101/s32/s40/s115/s41/s114/s97/s109/s112/s32/s114/s97/s116/s101/s115/s58\nFIG. 3: Hysteresis of the transition at zero applied field mea -\nsured at different ramp rates. Temperature sweep profiles are\nshown in the inset. Clearly, this hysteresis and jumps are no t\ndynamic effects.\n/s48/s46/s57 /s49/s46/s48 /s49/s46/s49 /s49/s46/s50 /s49/s46/s51 /s49/s46/s52/s45/s49/s46/s48/s45/s48/s46/s56/s45/s48/s46/s54/s45/s48/s46/s52/s45/s48/s46/s50/s48/s46/s48/s52\n/s84/s32/s40/s75/s41/s50/s56/s48/s48/s32/s79/s101\n/s48/s46/s57 /s49/s46/s48 /s49/s46/s49 /s49/s46/s50 /s49/s46/s51 /s49/s46/s52/s45/s49/s46/s48/s45/s48/s46/s56/s45/s48/s46/s54/s45/s48/s46/s52/s45/s48/s46/s50/s48/s46/s48\n/s84/s32/s40/s75/s41/s72/s61/s48\n/s72/s32/s124/s124/s32/s99/s50 /s52 /s54 /s56/s50/s52/s54/s56/s49/s48\n/s78/s111/s114/s109/s97/s108 /s83/s85/s80/s69/s82/s67/s79/s78/s68/s85/s67/s84/s73/s78/s71/s72/s32/s124/s124/s32/s99\n/s32/s32/s72/s32/s40/s107/s79/s101/s41\n/s84/s32/s40/s75/s41/s108/s105/s110/s101/s115/s32/s45/s32/s84/s68/s82\n/s115/s121/s109/s98/s111/s108/s115/s32/s45/s32/s116/s114/s97/s110/s115/s112/s111/s114/s116\n/s72/s32 /s32/s99/s70/s77\nFIG. 4: Transition region measured for magnetic field ori-\nented along the needle and c-axis. Inset: summary phase di-\nagram for two orientations measured by transport (symbols)\nand TDR.\nother orientation, when external magnetic field is applied\nalongthe c−axis. Thisis showninFig. 4. Note that peak\ninχ(T) at the FM boundary is not present, which is\nconsistent with the behavior of anisotropic local-moment\nferromagnet [29]. The inset to Fig. 4 shows the phase di-\nagram obtained from resistivity and TDR measurements\nfor both orientations. There is excellent agreement be-\ntween the two techniques and, as noted earlier, this dia-/s48/s46/s56/s53 /s48/s46/s57/s48 /s48/s46/s57/s53/s45/s48/s46/s54/s45/s48/s46/s53/s45/s48/s46/s52/s45/s48/s46/s51/s45/s48/s46/s50/s45/s48/s46/s49/s48/s46/s48/s48/s46/s49\n/s48/s46/s56/s53 /s48/s46/s57/s48 /s48/s46/s57/s53/s45/s48/s46/s56/s45/s48/s46/s54/s45/s48/s46/s52/s45/s48/s46/s50/s48/s46/s48/s49/s48\n/s57\n/s55\n/s54/s44/s56/s52\n/s51/s44/s53/s50/s52\n/s84/s32/s40/s75/s41/s72/s32/s61/s32/s54/s54/s48/s32/s79/s101\n/s49/s48/s46/s49/s32/s75\n/s72/s32/s124/s124/s32/s99/s52\n/s84/s32/s40/s75/s41/s72/s32/s61/s32/s48\nFIG. 5: Details of the hysteresis with partial scans as ex-\nplained in the text. Inset shows anothe partial loop on a\ncooling part of the curve.\ngram is consistent with previous reports [11, 12].\nFinally, Fig. 5 shows so called minor hysteresis loops\n(not as function of field, but temperature). The labels\nshow the evolution of the susceptibility. It starts from\nlow temperature at (1) when sample was warmed up to\nfirst signs of superconductivity that appeared as small\njump at (2), then warmed further and reaching almost\nfull superconductivity at (3), but then cooled back down\nto (4) as indicated by arrow and warmed back to (5).\nNote that along (3) →(4)χis significantly different\nfrom (4) →(5). Another similar minor cooling-warming\nloop follows (6) →(7)→(8) after which the sample was\ncooleddowntoreturnto(10) = (1)via(9). Interestingly,\nthere are no steps or jumps observed on the minor loops\nevenonwarming. Also, theslope dχ/dtissimilaroncool-\ning and warming and is very different from the original\nsteep slope (2) →(3). This is consistent with the present\nof vortices, probably pinned by the modulated FM/SC\nstructure. The inset in Fig. 5 shows a small minor loop\non a cooling part. This loop has small slope comparable\nto the larger loops described in the main frame.\nLet us now turn to the interpretation of these results.\nClearly, the FM →SC transition proceeds via a series of\njumps in diamagnetic screening due to formation of su-\nperconducting regions of macroscopic volume, roughly\n5%−20% of the sample volume depending on the ap-\nplied field and temperature. Indeed, each observed step\nmay be a result of simultaneous formation of many in-\ndividual superconducting domains of similar size. These\nsteps in χare present both for H||candH||aaxes, al-\nthough in the latter case the steps are smaller and are\nmore pronounced, possibly due to magnetic and shape4\nanisotropies. The number of steps increases with the in-\ncreasingfield and the firststep (the first signofsupercon-\nductivity) occurs at a higher temperature for larger ap-\nplied field. Overall, FM →SC transition is apparently of\nthefirstorderandexhibitsbehaviorconsistentwith type-\nI superconductivity as predicted theoretically [2, 20, 21].\nFrom our point of view, the first jump occurs at a tem-\nperature where internal field is equal to a supercooling\nfield of a type-I superconductor. (Note that apparent su-\nperheating of the FM →SC transition [17] corresponds to\nsupercooling of a regular type-I superconductor [17, 30],\nbecause we enter normal phase on cooling). When first\nsuperconductingdomains appear, effective magneticfield\naround them increases due to the flux expulsion and in-\nternal field becomes more inhomogeneous. This net in-\ncrease in the internal field in the remaining FM regions\nstabilizes them to higher temperatures. The system now\nneeds to get farther away from the initial FM boundary,\ndeeper into the SC state to produce more superconduct-\ningpatches. Inthisscenario,theobservedjumpsin χcor-\nrespond to a cascade of supercooling transitions. If the\ntemperature is lowered before the transition is complete\ndomains stay stable to lowertemperatures due to physics\nsimilar to superheating of a type-I superconductor. It is\nalso quite possible that superconducting domains have\nthe modulated spin structure seen in neutron scattering\n[14]. Finally, itseemsthatferromagneticdomainsarenot\ndirectly related to the observed steps, because at higher\nfields, the number of these domains decrease and domi-\nnant domains (along the applied field) grow in size.\nIn a striking contrast with FM →SC transition, the\nSC→FMtransitionissmoothandproceedstomuchlower\ntemperatures. Yet the transition is hysteretic as evident\nfrom the minor loops shown in Fig. 5. It is possible\nthat Abrikosov vortices are being spontaneously created\nas the temperature is lowered and the systems crosses\nover into the normal state when vortex cores overlap.\nThe vortex state is also compatible with long-range co-\nherence observed in neutron scattering experiments [14].\nAt the same time ferromagnetic modulation with a pe-\nriod of∼10 nm may also develop between the vortices\n[2, 15, 18, 19]. This would be also be similar to FFLO\nstate in the presence of vortices [31]. Moreover, this\nwould explain different intensities of small-angle satel-\nlite peaks, because coherence volume in the vortex state\nmust be much larger compared to domain-like state on\nwarming. We alsonote that unusual enhancement of dia-\nmagnetism in the vicinity of the FM boundary from the\nSC side, could be due to an FFLO pocket as predicted by\nBulaevskii for ErRh 4B4[2]. If we plot temperature of a\nminimum in χ(T) as function applied field, and also Hc2,\nwe obtain phase diagram remarkably similar to Fig.7 of\nRef.[2].\nOverall,weconcludethat wearewitnessinganunusual\ntransition. It is definitely first order on warming with\nsigns of type-I superconductor ”supercooling”. However,it is smooth upon cooling and exhibits smooth minor\nloops similar to a type-II superconductor. A second or-\nder transition occurs between normal and FFLO phases\nas well as between normal and SC for type-II supercon-\nductor. However, transition from SC to FFLO state is\nfirst order as well as from SC to spiral state. It is possi-\nble that size of the new phase nuclei is so small that we\ncannot resolve it upon cooling.\nDiscussions with Lev Bulaevskii, Alexander Buzdin,\nVladimir Kogan and Roman Mints are appreciated.\nWork at the Ames Laboratory was supported by the De-\npartment of Energy-Basic Energy Sciences under Con-\ntract No. DE-AC02-07CH11358. R. P. acknowledges\nsupport from NSF grant number DMR-05-53285 and the\nAlfred P. Sloan Foundation.\n∗Electronic address: prozorov@ameslab.gov\n[1] V. L. Ginzburg, Soviet Phys. JETP 4, 153 (1957).\n[2] L. N. Bulaevskii, A. I. Buzdin, M. L. Kuli, and S. V.\nPanjukov, Adv. Phys. 34, 175 (1985).\n[3] O. Fischer, Magnetic superconductors , vol. 5 (Elsevier,\n1990).\n[4] M. L. Kulic, Comptes Rendus Physique 7, 4 (2006).\n[5] M. B. Maple, Physica B 215, 110 (1995).\n[6] K. P. Sinha and S. L. Kakani, Magnetic Superconduc-\ntors: Recent Developments (Nova Science Publishers,\nNew York, 1989).\n[7] W. A. Fertig, D. C. Johnston, L. E. DeLong, R. W. Mc-\nCallum, M. B. Maple, and B. T. Matthias, Phys. Rev.\nLett.38, 987 (1977).\n[8] M. Ishikawa and O. Fischer, Solid State Commun. 23, 37\n(1977).\n[9] P. C. Canfield, S. L. Bud’ko, and B. K. Cho, Physica C\n262, 249 (1996).\n[10] S. S. Saxena, P. Agarwal, K. Ahilan, F. M. Grosche,\nR. K. W. Haselwimmer, M. J. Steiner, E. Pugh, I. R.\nWalker, S. R. Julian, P. Monthoux, et al., Nature 406,\n587 (2000).\n[11] G. W. Crabtree, F. Behroozi, S. A. Campbell, and D. G.\nHinks, Phys. Rev. Lett. 49, 1342 (1982).\n[12] G. W. Crabtree, R. K. Kalia, D. G. Hinks, F. Behroozi,\nand M. Tachiki, J. Magn. Magn. Mater. 54-57, 703\n(1986).\n[13] D. E. Moncton, D. B. McWhan, P. H. Schmidt, G. Shi-\nrane, W. Thomlinson, M. B. Maple, H. B. MacKay, L. D.\nWoolf, Z. Fisk, and D. C. Johnston, Phys. Rev. Lett. 45,\n2060 (1980).\n[14] S. K. Sinha, G. W. Crabtree, D. G. Hinks, and H. Mook,\nPhys. Rev. Lett. 48, 950 (1982).\n[15] E. I. Blount and C. M. Varma, Phys. Rev. Lett. 42, 1079\n(1979).\n[16] D. E. Moncton, D. B. McWhan, J. Eckert, G. Shirane,\nand W. Thomlinson, Phys. Rev. Lett. 39, 1164 (1977).\n[17] J. M. DePuydt, E. D. Dahlberg, and D. G. Hinks, Phys.\nRev. Lett. 56, 165 (1986).\n[18] H. Matsumoto, H. Umezawa, and M. Tachiki, Solid State\nCommun. 31, 157 (1979).\n[19] P. W. Andersonand H.Suhl, Phys.Rev. 116, 898(1959).5\n[20] M. Tachiki, H. Matsumoto, and H. Umezawa, Phys. Rev.\nB20, 1915 (1979).\n[21] K. E. Gray, J. Zasadzinski, R. Vaglio, and D. Hinks,\nPhysical Review B 27, 4161 (1983).\n[22] A. I. Buzdin, L. N. Bulaevskii, and S. V. Panyukov, Zh.\nExp. Teor. Fiz. 87, 299 (1984).\n[23] A. I. Buzdin and A. S. Mel’nikov, Phys. Rev. B 67,\n020503/1 (2003).\n[24] S. Okada, K. Kudou, T. Shishido, Y. Satao, and\nT. Fukuda, Jpns. J. Appl. Phys., Part 2: Letters 35,\nL790 (1996).\n[25] T.Shishido, J.Ye,T.Sasaki, R.Note, K.Obara, T.Taka-\nhashi, T. Matsumoto, and T. Fukuda, J. Solid State\nChem.133, 82 (1997).\n[26] R. Prozorov, R. W. Giannetta, A. Carrington, and F. M.Araujo-Moreira, Phys. Rev. B 62, 115 (2000).\n[27] R. Prozorov, R. W. Giannetta, A. Carrington,\nP. Fournier, R. L. Greene, P. Guptasarma, D. G. Hinks,\nand A. R. Banks, Appl. Phys. Lett. 77, 4202 (2000).\n[28] R. Prozorov and R. W. Giannetta, Supercond. Sci. Tech-\nnol.19, R41 (2006).\n[29] M. D. Vannette, A. Safa-Sefat, S. Jia, S. A. Law,\nG. Lapertot, S. L. Bud’ko, P. C. Canfield, J. Schmalian,\nand R. Prozorov, J. Mag. Mag. Mater. in print (2007).\n[30] J. Feder, S. R. Kiser, and F. Rothwarf, Phys. Rev. Lett.\n17, 87 (1966).\n[31] M. Ichioka, H. Adachi, T. Mizushima, and K. Machida,\nJ. Magn. Magn. Mater. 310, 593 (2007)." }, { "title": "0804.1713v1.Nonergodic_thermodynamics_of_disordered_ferromagnets_and_ferroelectrics.pdf", "content": "arXiv:0804.1713v1 [cond-mat.dis-nn] 10 Apr 2008Nonergodic thermodynamics of disordered ferromagnets and ferroelectrics\nP. N. Timonin∗\nPhysics Research Institute at Southern Federal University , 344090, Rostov - on - Don, Russia\n(Dated: November 4, 2018)\nPhenomenological thermodynamic theory describing the pro perties of metastable states in dis-\nordered ferromagnets and ferroelectrics with frustrative random interactions is developed and its\nability to describe various nonergodic phenomena in real cr ystals is demonstrated.\nPACS numbers: 75.10.Nr, 77.80.-e\nI. INTRODUCTION\nMany types of disorder in ferromagnets and ferro-\nelectrics can induce spin- or dipole-glass states if emer-\ngent random interactions cause frustration and make the\nground state of a crystal highly degenerate1. Besides\nthe classic example of the solid solution (alloy) of fer-\nromagnet with antiferromagnet2it can be the random\nanisotropy3, simple dilution in the case of competing\nnearest and next-nearest interactions4or in the presence\nof strong dipole-dipole ones5as well as structural defects\nsuch as dislocations6. Now it is widely recognized that\nthe unusual nonergodic properties of disordered ferro-\nmagnets and ferroelectrics in their glassy phases result\nfrom the appearance of a number of metastable states\neach having more or less random directions of local spon-\ntaneous electric or magnetic dipole moments1. Using the\ndifferent protocols of changing temperature and external\nfield (e. g. field-cooled, FC, zero-field cooled, ZFC, etc.)\none may arrive at the metastable states with different\nnet magnetization (polarization) and differences in other\nthermodynamic functions.\nFirst adequate theoretical description of such history\n- dependence of system properties in glassy phases was\nachieved in Refs. [7, 8] using numerical simulations in\nthe local mean-field theory of random-bond magnets. It\nwas explicitly shown that various regimes of temperature\nand field variations do bring system to different minima\nof thermodynamic potential thus providing the history\n- dependence of thermodynamic functions. Under the\nassumption of the possibility of sufficiently slow (quasi-\nstatic) variations of external parameters making the pre-\ndictionsofstaticthermodynamictheoryvalidthenumber\nof nonergodic features of glassy phases were described in\nRefs. [7, 8]: the temperature dependencies of FC and\nZFC functions, thermal and isothermal remanent mag-\nnetizations, the form of hysteresis loops, FC heat capac-\nity in different fields. These results are in reasonable\nagreement with the experimental data for a series of ran-\ndom magnets. Later the method of the local mean-field\nsimulations were used to reveal the presence of noner-\ngodic reentrant spin-glass phase in 2d XY random-bond\nmodel9.\nSubsequent studies of nonergodic phenomena in spin\nsystems with random interactions were mainly focused\non numerical simulations of hysteresisloops. The smallerloops inside the main one and evolution via macroscopic\nspin avalanches on the main loop are shown to exist in\nEdwards-Anderson10and Sherrington- Kirkpatrick11,12\nspin-glass models. The existence of hysteresis loops is\nestablished in random - anisotropy model13. The new\nhard-spin mean-field method is developed for frustrated\nrandom - bond systems14,15,16which provides the evi-\ndencesfor the existenceofmultiple metastablestatesand\ncandescribeanumberofnonergodicphenomenawithless\nsimulation efforts.\nInspiteofdefinitesuccessesinnumericalsimulationsof\nnonergodic effects in glassy phases these methods are not\ndestined to provide some general picture for systemati-\nzation and qualitative explanation of these effects. It is\nrather difficult to find from the existing numerical results\nthe possible interrelations between a variety of history -\ndependent phenomena observed in different experimen-\ntal regimes, see Refs. [17-25]. Yet it seems that they\nshould be necessary present as, e. g., the hysteresis loops\nand ZFC - FC magnetization differences have the com-\nmon origin. Then one may hope to obtain the unified\npicture of such phenomena using analytical calculations\nin some simplified models. But nowadays there are no\nsimple enough microscopic models in which nonergodic\neffects in finite field could be described by the analytical\nmethods.\nThus in present situation some phenomenological ap-\nproach may help to achieve general understanding of the\nquasi-static large field response in glassy phases and the\nrole played in it by metastable states. Apparently, it\nshould be based on some realistic mechanism underlying\nthe appearance of such states. Such mechanism for the\nrandom-interaction systems has been described in Ref.\n[26]. It consists in the subsequent condensations of nu-\nmerous sparse fractal modes defined by the eigenvectors\nof the random matrix of pair-wise interactions (exchange\nmatrix in magnets or the matrix of harmonic interac-\ntion of polar atomic displacements in ferroelectrics) at\ntheir localization threshold27. In random ferroelectrics\n(relaxors) these modes play the role of the notorious\n”softmode”astheyarethedelocalizedeigenmodesofpo-\nlar atomic oscillation with the lowest frequencies. Their\namplitudes throughout the disordered crystal can be ob-\ntained by the numerical methods of Refs. [27]. The con-\ndensation (freezing) of these modes in relaxors results in\nappearance of stable polar atomic displacements same\nas in the case of ordinary soft mode albeit these dis-2\nplacements are highly inhomogeneous and sparse. Their\nanalogsinrandomIsingmagnetscanberelatedtothede-\nlocalized relaxational magnetic modes which freeze first\nat a macroscopic spin glass transition.\nIt is important to realize that only the freezing of\nthese macroscopic modes can result in emergence of sta-\nblespontaneouslocalmomentsataglasstransition. Such\nobjectsasnano-domainswillneverhavespontaneousmo-\nments stable for a macroscopic time so this notion is use-\nless for the description of the true thermodynamic tran-\nsition into the glass phase. Thus Burns temperature in\nrelaxors at which experiments reveal the appearance of\n(fluctuating) polar nano-domains marks the onset of the\nparaelectricGriffiths’phaseinwhichonlynon-analyticity\nin the field dependence of polarization28and slow non-\nexponential relaxation29appear.\nThe suggested mechanism of a macroscopic spin glass\ntransition is in sharp contrast with that of phase transi-\ntion in homogeneouscrystals. The last takesplace due to\nthe condensation of just one periodic or uniform in space\neigenmode which correspondsto the largest eigenvalue of\nexchange matrix (in magnets) or the lowest eigenvalue of\nthe matrix of effective harmonic interaction (soft mode\nin ferroelectrics). In ideal crystal it is sufficient to sta-\nbilize the other modes with eigenvalues close to the crit-\nical one throughout some low-symmetry phase regions.\nThis stabilization of the rest of near-critical modes re-\nsults from nonlinear mode couplings which are always\npresent if there is some finite spatial overlap between the\ncritical mode and near-critical ones. As all eigenmodes\nof an ideal crystal is just the plane waves strongly over-\nlapping throughout the whole sample, the condensation\nof just one critical mode makes the system stable.\nIn a strongly disordered crystal with random interac-\ntions the situation changes principally. The randomness\nof pair-wise interactions makes the eigenmodes near the\nboundary of the spectrum localized27so the transition\nmay start only with the condensation of the first de-\nlocalized eigenmode at the localization threshold some-\nwhat far from the lower boundary of the spectrum. Ow-\ning to the sparse fractal structure of the modes near\nthis threshold27, the condensation of the first delocal-\nized mode can not stabilize the other sparse near-critical\nmodes which do not overlapessentially with the first one.\nSo they proceed to condense at slightly lower tempera-\ntures until almost all crystal sites acquire spontaneous\nmoments. If the average fractal dimension of these con-\ndensing modes is df< d, the average number of sites\nparticipating in each mode is of the order N1=Ndf/d.\nSo the number ofthem needed to coverthe N-site crystal\nisN0=N/N1=N1−(df/d).\nThus strongdisorder can transformthe ordinaryphase\ntransition into the macroscopic number of them spread-\ning in some temperature interval. As each mode after\ncondensation can have (at least) two stable states with\nthe reverse directions of local moments, we may end up\nwith up to 2N0metastable states. Note that each of\nthese numerous subsequent transitions is still a macro-scopic one as the macroscopic number of local sponta-\nneous moments of order N1appears at it. Apparently,\nthe mean-field estimate of the potential barrier between\nmetastable states would also give the macroscopic value\nproportional to N1.\nHerewemaynotethatabovepicturedescribesonlythe\npossibility ofemergenceofnumerousmetastablestates in\ncrystal with random interactions. Do it actually realizes\ndepends on the details of pair-wise interaction between\nthe sparse fractal modes. If this interaction tends to ori-\nent the net moments of the modes in parallel, we may\nhave ordinary ferromagnetic (ferroelectric) phase with or\nwithout some metastable states. Principally, sketched\nabove mechanism allows for phenomenological descrip-\ntion of the thermodynamics of metastable states in the\nspirit of Landau’s theory of phase transitions. The case\nwith purely glass transition was considered in Ref. [26].\nHere we present the simplified and more general deriva-\ntion of the phenomenological Landau’s potential for the\ngeneric case of competing ferromagnetic (ferroelectric)\nand glassyinteractions in disordered crystals. For a large\nclassofsuchrandomsystemsitispossibletoobtainsemi-\nanalytical description of the properties of the emergent\nmetastable states thus providing the comprehensive ac-\ncount of possible nonergodic effects in random ferromag-\nnets and ferroelectric relaxors.\nII. LANDAU’S POTENTIAL FOR RANDOM\nFERROMAGNETS AND FERROELECTRICS\nFurther we mainly resort to the magnetic terminol-\nogy and designations just to be specific. Let us consider\nthe ferromagnet with the second-order transition and\none-component order parameter where some disorder of\nrandom-interaction type presents tending to destroy the\nferromagnetic order. According to above considerations\nwe can introduce in this case the Landau potential for\nsome specific realizationof disorderdepending on the net\nmagnetizations of sparse fractal modes, mi,i= 1,...,N0.\nIf the temperature interval ( Tg,Tg−∆T) at which\nthe condensation of these modes takes place is narrow,\n∆T≪Tgthen we can consider the small minearTgand\nexpand the potential F(m) in powers of mi. It would\ncontain only even powers of these order parameters due\nto the globalinversionsymmetry. Alsowe canretain in it\nthe terms with no more than two different modes as one\ncan suggest that their sparse spatial structure makes the\nsimultaneous interactions of three or more modes negli-\ngibly small.\nFurtherwemaynotethat∆ T≪Tgisactuallythenec-\nessary condition when one consider the frustrating disor-\nder causing the (near) degeneracy of condensing types\nof magnetic (dipole) arrangements. This means that in\nsuch case the condensing modes have nearly the same\nthermodynamic potentials, close transition temperatures\nand thermodynamic parameters. The simplest way to\nimitate this near degeneracy is to suppose that all ther-3\nmodynamic properties of all modes are identical. Then\nF(m)shouldbesymmetricunderallpermutationsof mi.\nTheimmediateeffectofthis assumptionwillbethe merg-\ning of actually subsequent transitions to just one point,\nTg. In this way we avoid the consideration of narrow\ntemperature interval ( Tg,Tg−∆T) and gain the sig-\nnificant simplification of the model which still preserve\nthe essential features of nonergodic transitions caused by\nfrustrating disorder. We may say that model with per-\nmutation symmetry is the ”minimal” one and, probably,\nthe simplest model allowing for more or less adequate ac-\ncount of real nonergodic effects in large class of random\nferromagnets and ferroelectric relaxors.\nAt last, the quadratic terms of the potential should\nreflect the competition between ferromagnetic (ferroelec-\ntric) and glassy orders. Thus we arrive at the unique\nform of potential up to forth order in miobeying the\nabove criteria,\nF(m) =τf\n2N0/parenleftBiggN0/summationdisplay\ni=1mi/parenrightBigg2\n+τg\n4N0N0/summationdisplay\ni,j=1(mi−mj)2\n+a\n4N0/summationdisplay\ni=1m4\ni+b\n4N0/parenleftBiggN0/summationdisplay\ni=1m2\ni/parenrightBigg2\n−hN0/summationdisplay\ni=1mi(1)\nHerehis external field conjugate to the net magnetiza-\ntion of a sample\n¯m=N−1\n0N0/summationdisplay\ni=1mi\nNote that this relation is the consequence of the adopted\nprincipleofmodeequivalencewhichincludes thesupposi-\ntion ofequal numbers ofsites participatingin each mode.\nN0in the denominators of Eq. (1) ensure the equal order\nof different terms’ contributions at large N0≫1. Note\nalso the absence in (1) the term\nc\nN0N0/summationdisplay\ni=1miN0/summationdisplay\nj=1m3\nj\nallowedformally by the abovecriteria. It can be removed\nby the linear transform of miresulting in rescaling of\nhandτfand appearance of two forth-order terms with\nthree-andfour-modeinteractionswhichmustbedropped\ndue to assumed smallness of such interactions. Thus the\npresence in Fof this term amounts just to the handτf\nrescaling and it can be omitted.\nThe coefficients aandbinFare some constants spe-\ncific for a given disorder realization, while τfandτgare\nlinear decreasing functions of temperature Tchanging\ntheir signs at temperatures TfandTgcorrespondingly\nalso being disorder dependent. Yet in the ferroelectric\nand ferromagnetic solid solutions all potential parame-\nters must be the self-averaging quantities. It means that\nthey depend only on the impurities’ concentrations in ac-\ncordance with the experimental data showing no notice-\nable variations of the properties of different samples withthe same composition. Here we should note that this is\ntrue only for the solid solutions with deeply frozen disor-\nder. In the relaxors with annealing mediated ordering of\ncomponents such as PSN the parameters of the potential\nwill alsodepend on the degreeof orderingachievedat the\nhigh-temperature annealing.\nGenerally, the crystal with negligible disorder should\nhaveTg≪Tfas in this case we may have just the or-\ndinary transition into homogeneous ferro-phase without\nany traces of disorder. So the growth of disorder will\nresult in the increase of Tgalong with the lowering of Tf.\nThe difference between TfandTgmust be small to\nmake the expansion of Fin small mimeaningful. Ac-\ncording to Eq. (1) below Tfferromagnetic order may set\nin, while below Tgglassy states described by the N0−1\n- dimensional order parameter composed of the indepen-\ndentmi−mjcomponents can appear favoring the an-\ntiparallel miorientations. What actually results from\nthe competition of glass- and ferro-order depends on the\nrelation between TfandTgvalues as well as the rates\nofτfandτgdecreasing. Finding the evolution of F(m)\nminima for different τfandτgrelations we can get the\npossible variants of phase sequences in crystals with ran-\ndom interactions. In spite of huge amount of glassy min-\nima that F(m) may have, to find all of them is rather\nsimple task in the present model which can be fulfilled\nby semi-analytical methods.\nIII. THERMODYNAMICS OF COMPETING\nGLASS- AND FERRO-STATES IN ZERO FIELD\nUsing the notations/bracketleftbig\nmk/bracketrightbig\n=N−1\n0N0/summationtext\ni=1mk\ni, we can rep-\nresent the potential density f(m) =F(m)/N0in the\nsimple form\nf(m) =τg\n2/bracketleftbig\nm2/bracketrightbig\n+τf−τg\n2¯m2+a\n4/bracketleftbig\nm4/bracketrightbig\n+b\n4/bracketleftbig\nm2/bracketrightbig2−h¯m\nDifferentiating f(m) with respect to miwe get the equa-\ntions of state\n/parenleftbig\nτg+b/bracketleftbig\nm2/bracketrightbig/parenrightbig\nmi+am3\ni=h+(τg−τf) ¯m(2)\nThe solutions of these equations are minima of f(m)\ndescribing possible (meta)stable states if they render the\npositive definiteness of the matrix\nGi,j≡∂2f(m)\n∂mi∂mj=δi,j/parenleftbig\nτg+b/bracketleftbig\nm2/bracketrightbig\n+3am2\ni/parenrightbig\n+N−1\n0(τf−τg+2bmimj).\nLet us consider first the homogeneously magnetized\nferro-state with the equal mi=m0. From Eq. (2) we\nhave\nτfm0+(a+b)m3\n0=h (3)4\nThus ath= 0 the ferro-state appears at τf<0 ifa+b >\n0. It is stable for\nτg>3a+b\na+bτf (4)\nTurning to the possible glassy solutions of Eq. (2) with\nunequalmi, we note that all miobey the same equation\nsosuchsolutionsexistifthereareseveralrealrootstoEq.\n(2). Being of third order with respect to miit can have\none or three real roots. In the last case mican acquire\nonly two of the root values with the largest modules: one\npositive, m+>0, and one negative, m−<0, as only\nthese two can make the coefficient at δi,jin matrix ˆG\npositive which is necessary for the stability. Thus every\nglassy state is defined by the number, N+, ofmihaving\nm+values (or the number of m−ones,N−=N0−N+).\nHence in the glassy states\n/bracketleftbig\nmk/bracketrightbig\n=n+mk\n++n−mk\n−,¯m=n+m++n−m−\nn±≡N±/N0, n ++n−= 1\nand to find a glass state with a given n+(n−) we need\nto obtain the corresponding roots, m±, of Eq. (2) which\nnow becomes\n/parenleftbig\nτg+b/bracketleftbig\nm2/bracketrightbig/parenrightbig\nm±+am3\n±=h+(τg−τf) ¯m(5)\nThus the present model may generally have the quasi-\ncontinuous set of states (local minima) defined by the\nparameter n+which changes in the interval 0 < n+<1\nin the infinitesimal steps ±N−1\n0. Note that there are/parenleftbigg\nN0\nN+/parenrightbigg\nstates with the same N+=n+N0which differ\nby the permutations of miand have the identical ther-\nmodynamic parameters due to the adopted permutation\nsymmetry.Introducing the variable\nx=−m−/m+\nwe get from Eq. (5) the following equations\n/parenleftbigg\n1−τf\nτg/parenrightbigg\n(n+−xn−)R(x,n+,β) =\nx(x−1)R(x,n+,β)3+h\na/parenleftbigga\n−τg/parenrightbigg3/2\n(6)\nm+=/radicalBig\n−τg/aR(x,n+,β) (7)\nR(x,n+,β)≡/bracketleftbig\n1−x+x2+β/parenleftbig\nn++n−x2/parenrightbig/bracketrightbig−1/2\nHereβ≡b/a.\nThus we need to solve only one Eq. (6) for x=\nx(n+,τg,τf,h) to obtain using Eq. (7) and xdefini-\ntion full description of the thermodynamic properties of\na crystal in the metastable states with a given n+. So we\nhave for the net spontaneous moment in such states\n¯m= (n+−xn−)/radicalBig\n−τg/aR(x,n+,β) (8)\nand for the Edwards-Anderson glass order parameter\nq=/bracketleftbig\nm2/bracketrightbig\n−¯m2= (−τg/a)n+n−(1+x)2R(x,n+,β)2.\nFurther we consider the simplest case a >0,b >0 so the\nglassy metastable states appear at τg<0 .\nThen we have for the susceptibility χ=∂¯m\n∂h\nχ−1=τf−τg−τgR(x,n+,β)2(1+x)(2x−1)(2−x)+2β/bracketleftbig\n(2x−1)n++x2(2−x)n−/bracketrightbig\n(2x−1)n++(2−x)n−+2βn+n−(1+x)(9)\nThe equilibrium value of the potential for a given state\nis\nfeq(n+,τg,τf,h) =/parenleftbig\nτgq+τf¯m2−3h¯m/parenrightbig\n/4\nHere we omit the term arising from the degeneracy of\nglassy states\n−T\nNSconfig=−T\nNln/parenleftbigg\nN0\nN+/parenrightbigg\n∼N0\nN∼N−df/das it vanishes at large N. Then we get for a given\nmetastable state the entropy\nS=−∂feq(n+,τg,τf,h)\n∂T=−1\n2/parenleftBig\nτ′\ngq+τ′\nf¯m2/parenrightBig\n,\nτ′\nf,g≡∂τf,g\n∂T,\nand the heat capacity5\nC\nTgχ=/parenleftbig\nτ′\nf−τ′\ng/parenrightbig2¯m2+2τ′\ng/parenleftbig\nτ′\nf−τ′\ng/parenrightbig¯mm+[n+(2x−1)−n−x(2−x)]\nn+(2x−1)+n−(2−x)+2n+n−β(1+x)\n+(τ′\ng)2m2\n+/bracketleftbig\nn+(2x−1)+n−x2(2−x)/bracketrightbig\n+n+n−(1+x)a−1(τf−τg)\nn+(2x−1)+n−(2−x)+2n+n−β(1+x)(10)\nFora >0,b >0 the stability conditions for the so-\nlutions of Eq. (6) providing the positive definetness of\nmatrixˆGare represented by the inequalities\n1/2< x <2, χ > 0 (11)\nNote that the condition χ >0 ensures the stability with\nrespect to the ferromagnetic fluctuations while at the\nboundaries of the first inequality the glassy instability\noccurs.\nTurning to the case h= 0 we can see that simple solu-\ntions of the equation of state (6) exist for n+→1,0 and\nn+=n−= 1/2. In the first case Eq. (6) tends to the\nzero-field limit of Eq. (3) and we have\n¯m=/radicalBig\n−τf\na+b,χ−1=−2τf,q= 0,feq=τf¯m2/4,\nS=−τ′\nf¯m2/2,C=Tg(τ′\ng)2\n2(a+b)\nand Eqs.(11) defines the sector on the ( τf,τg)-plane in\nwhich nearly homogeneous states with n+→1,0 are\nstable\n3+β\n1+βτf< τg<3+4β\n4(1+β)τf<0 (12)\nNote that such states exist only in the part of the region\nEq. (4) where the homogeneously magnetized ferro-state\nis stable. This just means that the set of metastable\nstates to which the states with n+→1,0 belong has\nmore narrow region of existence.\nFor the state with fully disordered moments ( n+=\nn−= 1/2 ) we have x= 1 from Eq. (6) at h= 0 so\n¯m= 0,χ−1=τf−τg3+β\n1+β,q=−τg\na+b,feq=τgq/4,\nS=−τ′\ngq/2,C=Tg(τ′\ng)2\n2(a+b)\nandχ >0 requires that\nτg<1+β\n3+βτf (13)\nalong with τg<0. It can be shown that the sector of\nstability of fully disordered states is the largest one and\nthe stability regions of all other states belong to it, the\nmost narrow sector being that of nearly homogeneous\nstates (12).\nFurther we find\n∂feq\n∂n+=−a\n4(m++m−)(m+−m−)3,Thusfeq(n+,τg,τf,h) as function of n+has one ex-\ntremum at\nm++m−= 0 ( x= 1),\nand we have at this point\n∂2feq\n∂n2\n+/vextendsingle/vextendsingle/vextendsingle/vextendsingle\nx=1= 8am4\n+χ(τf−τg)1+β\n1+4βn+n−.\nSo it is minimum at τg< τfand maximum at τg> τf.\nThus the states with n+=n−= 1/2 have the lowest\npotential at τg< τfwhile the global minimum is at the\nstates with n+→1,0 atτg> τf.\nHence the first-order transition from the fully disor-\ndered glass to ferro-state takes place at τg=τf. One\ncan see from the above equations that potentials of these\nstates indeed become equal at τg=τf. The phase dia-\ngram of the model is shown in Fig. 1. The topology of\nthe phase stability regions in it is the same that is found\nin the various microscopic models of disorderd magnets,\ncf. Refs. [13], [30-32].\nIn a specific crystal there is some linear relation be-\ntweenτfandτgboth being linear functions of T. It\ndefines a straight line on ( τf,τg) plane which crystal fol-\nlows under Tvariations and, hence, the phase sequence\nproper for a given crystal. For example, if the relation\nτf=τg+τ0holds,τ0being some constant, the crystal\nundergoesjustonesecond-ordertransition-intotheordi-\nnary ferro-phase if τ0<0 or in the glass phase if τ0>0.\nIn the first case there is another qualitative change of\nthermodynamic properties (not a phase transition) - the\nsubsequent appearance of numerous metastable states in\nadditiontoferro-statebeginswhen crystalreachesthe re-\ngions of their stability belonging to the sector (13). Yet\nin these cases the application of field causes no drastic\nchanges in the temperature dependencies of ¯ m,χand\nCapart from the usual differences in their FC and ZFC\nvalues.\nMuch more pronounced field effects occur when the\npath on ( τg,τf) plane intersects the line τg=τf. Then\nglass or ferro-phase can appear as intermediate one in\nsome finite temperature interval neighboring para-phase\nand first-order transition between them takes place at\nτg=τf. Yet there are also different temperature behav-\nior when path crosses the region of coexistence of ferro-\nand glass states (the sector in Fig. 1 bounded by the\ndashed lines) and when it always stays in this region.\nThese four principally different paths are shown in Fig.\n1. Dotted linesin itcorrespondtopaths τf= 2.5τg+1.5a6\nτf/a\nτg/aA\nB glass\nferropara-phase\n0(a) τf/a\nτg/aA1\nB1\nferroglasspara-phase\n0(b)\nFIG. 1: (Color online) Phase diagram of the model for β= 5 (a) and β= 0.1 (b). Thick lines denote the phase transitions\nbetween the phases. Upper dashed lines show the stability bo undaries for the ferro-state (Eq. 4) and lower dashed lines s how\nthat of fully disordered states with n+= 0.5 (Eq. 13). Dotted lines correspond to paths τf= 2.5τg+ 1.5a(A andA1) and\nτf= 0.5(τg−a) (B and B1).\n(A andA1) andτf= 0.5(τg−a) (B and B1) which we\nuse further to present some typical temperature varia-\ntions of thermodynamic variables which differ essentially\non them.\nThe spin-glass phase on paths B and B1is often called\n”reentrant” as crystal returns again on cooling to the\nequilibrium state with zero net magnetization. The pos-\nsible microscopic mechanism of such reentrance is con-\nsidered in Ref. [9].\nWe must note that here the first-order transition does\nnot mean that at τg=τfthe crystal do jump from one\nstate to another on the laboratory time scale. Actually\nit must traverse the rough potential landscape and to\novercome on the way a number of macroscopic barriers\nbetween the localminima(partiallyorderedstates)to ar-\nrive to the global minimum. It takes an infinite time and\nthe standard thermodynamics indeed predicts formally\nwhat will happen after infinite time when system will be\nfully relaxed. But actually one has just finite laboratory\ntime to measure the thermodynamic quantities.\nIt is quite usual situations in all nonergodic systems (i.\ne. those with metastable states) - they all have infinite\n(Arrhenius) relaxation time for the decay of metastable\nstatesintostableoneirrespectiveoftheirnumber-totra-\nverse just one macroscopic barrier also needs an infinite\ntime the same as to traverse many of them. This cir-\ncumstance causes no difficulties in the application of the\nresults of the equilibrium thermodynamics to the known\nnonergodicsystemssuchasordinaryferroelectricandfer-\nromagnets in their ferro-phases. The simple example is\nthe description of hysteresis loop. Here the infinite-time\nthermodynamic prediction that spontaneous magnetiza-\ntionshouldreverseitsdirectionjustwhenthefieldchange\nits sign is easily modified with due account of the long-\nlivingmetastablestates(inwhichfieldandmagnetization\nhave opposite signs). As they can not relax to the global\nminimum at the laboratory time and persist until they\nbecome unstable (at the coercive fields) one just includethem in the description of the response to sufficiently\nslow (quasi-static) field thus obtaining upper and lower\nbranches of hysteresis loop. Apparently, to apply the re-\nsults of static thermodynamic to the dynamic process of\nfield variation we must be sure that during it the sys-\ntem stays close to a local minimum. Hence, in this case\nthe quasi-static condition means that the characteristic\ntime of field variations is greater than the magnetization\nrelaxation time in a local minimum.\nIn the same way we can get the description of quasi-\nstatic responses from the thermodynamic results for the\npresent model. To do this one must assume that quasi-\nstatic regimes can be attained experimentally, i. e. that\nsufficiently slow temperature or field variations can be\nachieved on the laboratory time scale which ensure the\nlocation ofsystem near some minimum during these vari-\nations. Then we just should take into account that once\nthe system entered the metastable state it can not leave\nit in such quasi-static processes until that state becomes\nunstable.\nHence, the validity of this picture of the temperature\nevolution in which crystal is trapped in a local minimum\nand has the obtained above thermodynamic parameters\ncrucially depends on the rate of cooling or heating. It\nshould be sufficiently small to provide the quasi-static\nevolution of the system, that is the characteristic time of\ntemperature variations, t0=/parenleftbigdlnT\ndt/parenrightbig−1, must be greater\nthan the largest magnetic relaxation time in the specific\nminimum. The same is true for characteristic time of\nfield variations. One may expect that when we are not\nclose toTgnearly homogeneouslymagnetized states with\n¯m∼m0would have rather moderate relaxation times\ncomparable to those the crystal has deep in para-phase.\nAlso far from Tgthe magnetic susceptibility of highly\ndisordered states with ¯ m≪m0may be very small mak-\ning the amplitude of magnetization relaxation in them\nsmall too. Then we would observe their nearly static pa-7\nrameters even for t0less than relaxation times which can\nspread up to huge macroscopic values in such states. If\nχreaches large values close to Tgin the states with low\n¯m≪m0their thermodynamic description would become\nuseless due to the strong aging effects in them. Then one\nshould necessarily resort to the dynamic theory to get\nthe adequate account of their (dynamic) properties.\nProvided the above quasi-static conditons are fullfilled\nanother question inherent to such highly nonergodic sys-\ntems arises - to which of numerous stable states the sys-\ntem will go after its present state becomes unstable?\nStrictly speaking, one should again turn to the dynamics\nto answer it, yet, as we will see, there is actually no vast\nchoice of possibilities in the framework of the heuristic\nassumption that system adopts the most smooth varia-\ntions of its average magnetization.\nThus crystal can stay in metastable state for a very\nlarge (Arrhenius) time and leave it only when it becomes\nunstable. Before this the only way to bring the noner-\ngodic crystal to some other state is to apply the external\nfield which could make the present state unstable. This\ncircumstance made former investigators to believe that\nferroelectric relaxors have no phase transitions exhibit-\ning only ”field - induced” ferroelectricity17.\nAll these considerations must be taken into account\nin the interpretation of the temperature dependences of\nmetastable parameters described by the above equations\nand shown in Figs. 2 - 5. Note that to plot them one\ndoes not need to solve the equation of state (6) as it\nprovides along with the expression for a thermodynamic\nvariable the parametric representation of its temperature\ndependence, xbeing the parameter varying in the region\nof stability of a given state.\nThe temperature dependencies of metastable states’\nmagnetizations in Figs. 2 - 5 are bounded by the lines\nat which states become unstable. In the glass phase they\ncorrespond to x= 1/2,2 so their equations follow from\nEqs. 6, 8 at these values of x\n¯m2=\n1+3\n2β±/radicalBigg/parenleftbigg\n1+3\n2β/parenrightbigg2\n+2τg\nβ(τf−τg)\n3\n×τf−τg\n2a.(14)\nHere minus and plus signs are for (Figs. 2, 4) and (Fig.\n3) correspondingly. In the ferro-phase this relation is not\ngenerally valid as some states seize to exist due to the\nferromagnetic instability ( χ→ ∞). In this case it is\nhard to get the compact analog of Eq. (14)\nAs crystal leaves its state only when it becomes unsta-\nble, it is ratherevident that the quasi-staticheating (Fig.\n2) or cooling (Fig. 3) of crystal being in the ferro-state\nbeyond its stability point in the glass phase makes it to\ngo through the succession of metastable states on the\nboundary of their stability regions. So the self-organized\ncriticality shows up in these processes as system is per-\nmanently unstable in finite temperature intervals and itsevolution proceeds via small but macroscopic avalanches\nof spin upturns.\nEqs. (14) describe the temperature dependencies of\nmagnetization in these regimes. In particular, for the\nvanishing of magnetization on heating in Figs. 2, 4 we\nget atτg→0\n¯m=2a\n|τf|/parenleftbigg−τg\n3a+2b/parenrightbigg3/2\n.\nAlso the specific temperature hysteresis occurs in the\ncooling-heating-cooling cycles (Fig. 2) and cooling-\nheating ones (Fig. 3) starting on the paths A and B from\nthehigh-temperaturephases(inthepresenceofinfinitesi-\nmalpositive field tohave ¯ m >0). Thesecyclesareshown\nby the directed lines on the temperature dependencies\nof thermodynamic parameters. Their remarkable prop-\nerty is that they allow to enter a variety of metastable\nstates (actually all in the range 0 .5< n+<1 in the\ncase of intermediate glass phase in Fig. 2 and those with\n0.628< n+<1 having the limited stability regions in\nFig. 3). For example, in the case of Fig. 2 we can stop\nheating at some Twhen crystal traverses the stablility\nboundaries of metastable states and start cooling it thus\ntrapping crystal in the state it was in at the stop.\nNote the sharp rise of the susceptibilities and heat ca-\npacities of metastable states in Figs. 2, 3. Actually this\ndenotesthedivergenceof χandC(cf. Eq.(10))duetothe\nferromagnetic instability metastable states experience in\nthe ferro-phase at τf> τg(τg/a <−1 in Fig. 2 and\nτg/a >−1 in Fig. 3).\nOn the path A1the appearance of intermediate glass\nphase in zero field is marked by the jumps in heat capac-\nity andχtemperature derivative (Fig. 4) while on the\npathB1the intermediate ferro-phase (Fig. 5) has usual\nCurie - Weiss anomaly of χand the usual jump in heat\ncapacity (not shown). In these cases the appearance of\nmetastablestatescannotbeobservedinzero-fieldexperi-\nments - at all Tthe crystal will stay in the state it enters\nfirst on cooling. But the presence of metastable states\ncan be easily revealed in the glass phase. Here stopping\nat some Tand applying for some time the field (above\nsome threshold one needed to make the former state un-\nstable) onecan end up in everymetastablestate choosing\nthe value of the applied field. In the further temperature\nvariations the crystal will stay in the so prepared state\nuntil it vanishes.\nIV. THERMODYNAMICS IN FINITE FIELD\nFurther weexamine the propertiesofmetastable states\nin finite fields. In the glass phase at τg< τfall\nmetastable states always have χ >0 so they have not\nferromagnetic instabilities here. Again, substituting x=\n1/2 andx= 2 in Eqs. (6, 9) we get the parametric\nrepresentation of two lines on (¯ m,h)-plane bounding the\nregion where these states exist (0 < n+<1 being the8\nn+\n1\n0.95\n0.9\n0.85\n0.8\n0.7\n0.62\n0.57\n0.52\n-0.5 -100.20.4\nτg/am\n-1.513\nχ\n-1 0τg/a-1 -0.5τg/a01\nC/Tn+\n1\n0.95\n0.9\n0.85\n0.75\n0.7n+\n0.7\n0.75\n0.85\n0.9\n0.95\n1\nFIG. 2: (Color online) Temperature dependencies of metasta ble states’ magnetization, susceptibility and heat capaci ty in zero\nfield for β= 5 corresponding to path A in Fig.1(a). Dotted lines show the stability boundaries and dashed ones represent the\nparameter’s variations predicted by the equilibrium therm odynamics (those of the states with n+= 1/2 in the present case).\nDirected lines show the evolution of magnetization in real- time quasi-static regime.\n-3 -1χ4\n0n+\n0.8\n0.85\n0.9\n0.95\n0.99\n1\nC/T0.2\n0\n-3 -1n+\n0.8\n0.85\n0.9\n0.95\n0.99\n1n+\n1\n0.95\n0.9\n0.85\n0.8\n0.7\n0.65\n0.6\n0.550.4\n0.2m\n0 -2 -4 τg/a 1 τg/a τg/a00.6\nFIG. 3: (Color online) The same as in Fig. 2 for path B in Fig. 1( a).\nn+\n1\n0.95\n0.875\n0.8\n0.75\n0.7\n0.65\n0.6\n0.55m\nτg/a0\n-3.502.51\n0\n1 0 -1χ\nC/T\nτg/a\nFIG. 4: (Color online) The same as in Fig. 2 for path A1in Fig.1(b). Dash-dotted line corresponds to C/T.χandCare\nshown only for the n+= 1/2 state as others are inaccessible at h= 09\nn+\n1\n0.95\n0.875\n0.8\n0.75\n0.7\n0.65\n0.6\n0.55m\n-3 10\nτg/a1\nFIG. 5: (Color online) Temperature dependencies of\nmetastable states’ magnetization in zero field for β= 0.1\ncorresponding to path B1in Fig. 1(b). Dashed directed line\nshow the decay of nearly homogeneous state’s magnetization\nin the infinitesimal field. The other lines have the same mean-\ning as in Fig.2.\nparameter). Excluding n+we get the equation two real\nroots of which define the couple of these lines\n3a+b/bracketleftBigg\n2+ ¯m/parenleftbigg2a\nh+(τg−τf) ¯m/parenrightbigg1/3/bracketrightBigg\n+τg/vextendsingle/vextendsingle/vextendsingle/vextendsingle2a\nh+(τg−τf) ¯m/vextendsingle/vextendsingle/vextendsingle/vextendsingle2/3\n= 0 (15)\nThese roots have h+ (τg−τf) ¯m >0 andh+\n(τg−τf) ¯m <0 with the interval 0 < n+<1 be-\ning swept when hvaries correspondingly in the intervals\n−hf< h < h e,−he< h < h f\nhf=/radicalbigg\n−τg\n3a+b/parenleftbigg\nτf−τg1+β\n3+β/parenrightbigg\n, (16)\nhe=/radicalbigg\n−τg\n3a+4b/parenleftbigg\nτf−4τg1+β\n3+4β/parenrightbigg\n(17)\nThe magnetization curves ¯ m= ¯m(h,n+) of metastable\nstates defined by Eqs. (6, 8) fill the region bounded by\nthe line segments of Eq. (12) as Figs. 6, 7 show. Again\nthe self-organized criticality appears in these segments\nas system is permanently unstable in them and its evo-\nlution proceeds via the avalanches of spin upturns. Such\navalanches and small magnetization jumps are shown\nto exist in hysteresis loops of Sherrington-Kirkpatrick\nmodel11,12.\nAt|h|> heall metastable states vanish leaving the\nonly stable homogeneous state with magnetization curve\nm0(h) given by Eq. (3). As it was discussed above\nthe crystal will stay in a local minimum until it van-\nishes so Fig. 6 represents possible forms of hysteresis\nloops for quasi-static periodic field. If amplitude of such\nfield|h|> hethe form of quasi-static loop is completely\ndescribed by Eqs. (3, 12). For smaller amplitudes theinner loops inside the outer one will be observed formed\nby the boundaries of Eq. (12) and ¯ m= ¯m(h,n+) and\n¯m= ¯m(h,1−n+) curves with some n+defined by the\nfield amplitude.\nWe can find the analog of the coercive field hcfor the\nglassy loops putting ¯ m= 0 in Eq. (15),\nhc= 2a/parenleftbigg|τg|\n3a+2b/parenrightbigg3/2\nThe inclined hysteresis loop, see Figs. 6(b,c), 7(a),\nis the hallmark of glass phase; such loops are ubiqui-\ntous in random magnets and relaxors. They are found\nin the numerical simulation of spin glass phases of 2 d\nshort-range random-bond Ising model10, Sherrington-\nKirkpatrick model11,12, random anisotropy model13as\nwell as in the local mean-field simulations7,8. Qualita-\ntively the transition at τf=τgmanifest itself by the\nchange of loop form, inclined in the glass phase it be-\ncomes the ordinary one with the vertical sides in the\nferro-phase as disordered metastable states can not be\nreached in the quasi-static periodic field, see Fig. 6(a).\nItisshownthatsuchchangeofhysteresislooptakesplace\nin Sherrington-Kirkpatrickmodel with the growth of fer-\nromagnetic exchange11.\nAlso smaller loops vanish in the ferro-phaseas the field\nevolution of all metastable states shown on Fig. 6(a),\n7(b,c) ultimately ends by the transition into the homo-\ngeneous one. For b= 0 it happens strictly at τf=τg\nwhile for finite b >0 inclined loop transforms into rect-\nangular one already in the glass phase slightly before the\nthermodynamic transition into the ferro-phase. Gener-\nally this change takes place when hcbecomes less than\n−hf, that is at\nτf=/parenleftBigg\n1+β\n3+β+2/radicalBigg\n3+β\n(3+2β)3/parenrightBigg\nτg\nIt is also easy to get for all hthe equilibrium magne-\ntizationmeqcorresponding to the states with the lowest\npotentialhaving x= 1inglassphaseaswasshownabove.\nFrom Eqs. (6), (8) we have\n¯meq=h\nτf−τg(18)\nfor the equilibrium states with\nneq\n+=1\n2/parenleftbigg\n1+h\nhAT/parenrightbigg\nhAT=√−τg(τf−τg)√\na+b\nThe Almeida-Thouless field hATmarks the thermody-\nnamic transition between the completely ordered equi-\nlibrium phase at |h|> hATto the sequence of disordered\nones at|h|< hAT.10\nH -2 21.5\n-1.5/c109(b)\nhe\nH -1 11.5\n-1.5/c109(a)\nhf\nH -3 31.5\n-1.5/c109(c)\nhAT hc\nFIG. 6: (Color online) Field dependencies of metastable sta tes’ magnetizations for β= 0.1 on the path A1,µ≡¯mp\n−a/τg,\nH≡hp\n−a/τ3g, (a) -τg=−1.1a, (b) -τg=−0.75a, (c) -τg=−0.5a\n.\n(b)\nHµ\nhf he(a)\nHµhAT\n-0.15 0.15-0.50.5\n-0.50.5\n-0.2 0.2(c)\nHµ\n-0.60.6\n-0.3 0.3\nFIG. 7: (Color online) The same as in Fig. 6 for β= 5 and path B. (a) - τg=−1.5a, (b) -τg=−0.65a, (c) -τg=−0.5a\nFor the disordered states ¯ meqis represented in the in-\nteriors of hysteresis loops in Figs. (6, 7) by dashed lines\nwith tangents defined by the equilibrium susceptibility\nχeq=∂¯meq\n∂h=1\nτf−τg(19)\nand ending at h=±hAT.\nGenerally χeqis unobservable quantity as to follow the\nrelation in Eq. (18) under the field variations the crystal\nhastoovercomethemacroscopicbarriersinseriesoffirst-\norder phase transitions (stepping in them from the state\nwith some n+to that with n+±N−1\n0) which are not\nshown in Figs. 6, 7 (see Fig.1 in Ref. [26)]. As we\ndiscussed above, this would need a very long time and\ncrystal will stay in the local minimum on the laboratory\ntime scales. So the only susceptibility measurable under\nfield variations is that given by Eq. (9) which defines the\ntangents of the magnetization curves ¯ m= ¯m(h,n+) of\nthe specific metastable state at a given field. It is always\nless than χeq, thus in the equilibrium states with x= 1\nχ−1=τf−τg/bracketleftbigg\n1+2h2\nAT\nh2\nAT+β(h2\nAT−h2)/bracketrightbigg\nwhile at the right ( x= 1/2) and left ( x= 2) boundaries\nof hysteresis loop in glass phase we have\nχ−1=τf−τg/bracketleftbigg\n1+6β\n(3+4n±β)[3+β(1+3n±)]/bracketrightbiggwith plus and minus sign correspondingly.\nYet in the case b≪a(β≪1)χeqcan be determined\njust from the shape of hysteresis loop. Indeed, Eq. (15)\ngives then for its boundaries the almost straight lines\nwith tangent equal to χeq, see Figs. 6(b, c). But for\nlargeβthis relation is lost owing to the intricate form\nof the loop as Figs. 7(a) show. For small βthe shape\nof hysteresis loop in the glass phase is mostly defined\nby the fields heandhf. In the ferro-phase ( τf< τg)\nthe loop boundaries are no longer given by Eq. (15) as\nhere some states vanish due to ferromagnetic instability\nbefore reaching the glassy one at x= 1/2 orx= 2. Also\ninthe ferro-phasethe expressionforthe field hf, Eq. (16)\nchanges to\nhf=−2\n3/radicalBigg\n−τ3\nf\n3(a+b)(20)\nin the interval\n3+β\n3(1+β)τf< τg<3+4β\n12(1+β)τf.\nThe field in Eq. (20) is just the coercive field for the\nhomogeneous ferro-state. Note that at\nτg=3+4β\n12(1+β)τf11\nFIG. 8: (Color online) Temperature dependencies of charac-\nteristic fields; (a) - β= 0.1 on the path A1, dash-dotted line\nrepresents h= 0.2a, (b) -β= 5 and path B. Dashed lines\n-hAT, dotted ones - hc. Solid lines show the modules of hf\nandhe.\nthe field hfandhemerge at this value so at\nτg>3+4β\n12(1+β)τf\ndisordered metastable states no longer exist at all hand\nwe have the ordinary ferromagnetic loop. The temper-\nature dependencies of the loop’s characteristic fields for\npathsA1and B are shown in Fig. 8, in the ferro-phase\nhcandhewere found numerically.\nNow we can turn to the temperature dependencies of\nmetastable states’ parameteres in a constant field. In\ngeneral these states gradually vanish at larger fields, first\nthe most disordered ones, as one may expect. Fig. 9\nillustrates this process for the path A. Temperature de-\npendencies of ¯ m,χandCin constant field are closely\nrelated to that of hfandhe. Thus for paths A and\nA1hffirst grows on cooling and then diminishes after\nreaching the maximum, see Fig. 8(a) for path A1where\nhf,max≈0.28ais attained at τg≈ −0.24a. Alsohe> hf\nexhibits the similar behavior. This results in the essen-\ntial changes in χandCdependencies for field-cooling\nand field-heating in Fig. 10 and, especially, in Fig. 11\nas compared to those in Figs. 2, 4 for h= 0. In the\ninsets of Figs. 10, 11 there are slight anomalies of these\nparameters when crystal enters on cooling the interval\nwhereh < hfand leaves the homogeneous state to join\nthe sequence of glassy ones with n+<1 until it reaches\nthe first state being stable throughout all this interval.It varies for different paths and fields. In Figs. 10, 11\nthey are those with n+= 0.916 and n+= 0.921 corre-\nspondingly. This processis nothing else than the descend\nthroughout the upper branch of the growing hysteresis\nloop. In further cooling the crystal stays in this state un-\ntil it vanishes due to ferromagnetic instability near the\npoint where h≈he(at anotherbranch ofthe loop having\nnow the rectangularform, see Fig. 6(a)) for path A1). So\nχandCdiverge at this point according to Curie-Weiss\nlaw and crystal returns to the ferro-state. This is new\nfeature for path A1which is absent in zero field, Fig. 4.\nOnheating from this state n+againstartsdiminishing at\nthe lowerboundaryof the interval where h < hfreaching\nn+= 0.916(0.921) (upper branch of the loop). Further\nheating in this state ends up at the upper boundary of\nh < hfinterval in the sequence of states with n+growing\nup to 1 (lower branch of the loop). This cause the slight\nhigh-temperature spikesin χandCmanifesting the glass\ninstabilities of these states.\nHere we may note that Fig. 11 explains the origin\nof additional pronounced peak in χappearing in relaxor\nPMN in FC regime18,19. It is the consequence of the\ninstability of nearly homogeneous state the crystal en-\nters on field-cooling. Moreover, the behavior of PMN\ndielectric susceptibility observed in all field-cooling-field-\nheating cycle18,19resembles qualitatively that of χin\nFig. 11. The appearance of additional spikes of\nmagnetic susceptibility in FC regime was also regis-\ntered in random ferromagnets PrNi0.3Co0.7O320and\n(Fe0.17Ni0.83)75P16B6Al321.\nWeshouldalsonotethatmagnetizationsintheregimes\nof cooling from the para-phase shown in Figs. 10(a),\n11(a) are the field-cooled ones, mFC. These figures rep-\nresent also the zero-field-cooled magnetizations, mZFC,\nobtained at turning on the field after cooling the sam-\nple in zero field. They are the metastable curves with\nn+= 1/2 (if they exist) or the stability boundaries cor-\nrespondingto the lowerbranchof inclined hysteresisloop\nor to the homogeneous states lines when the loop is rect-\nangular as in Figs. 10(a). Indeed, in zero-field cooling\nprocess on paths AandA1the system is trapped in the\nn+= 1/2 state, see Figs. 2(a), 4(a), and stays in it af-\nter turning on the field or goes to the lower loop branch\nor to the n+= 1 state depending on the loop form and\nfield strength. So in Fig. 10(a) mZFCcorresponds to the\nlowerstabilitylinewhile in Fig. 11(a) it follows n+= 1/2\ncurve and then jump to n+= 1 curve when former seized\nto exist.\nHereweshouldnotethatofteninexperimentsandsim-\nulations another definition of mZFCis used, namely, the\nmagnetization registered in field-heating after zero-field-\ncoolingprocess. AsFig. 10(a)showsthisistheill-defined\nquantity as it will depend in this case on the temperature\natwhichtheturnfromcoolingtoheatingtakesplace. Yet\nwhen the field at the turning point leaves the n+= 1/2\nstate stable as for path A1in Fig. 11(a) the field-heating\nafter zero-field-cooling protocol gives a unique mZFCthe\nsame as in simple ZFC process described above. The de-12\n-1.4 0\n/c116g/a-0.7 -1.4 -0.2\n/c116g/a-0.80.5\n-0.5m0.5\n-0.5m(a) (b)\nFIG. 9: (Color online) Temperature dependencies of metasta ble states’ magnetizations for β= 5 corresponding to the path A\nin the fields h= 0.03a- (a) and h= 0.07a- (b).\nτg/a0.6\n0\n0 -0.5C/T\nn+\n1\n0.985\n0.967\n0.95\n0.935\n0.916-0.03 -0.120.11\n0.02\nτg/an+\n0.916\n0.935\n0.95\n0.967\n0.985\n1\n0 -0.50.41.6\nχ0.71\n0.65\n-0.1 -0.040.4\n0.2m\n-0.9 0 τg/an+\n1\n0.98\n0.95\n0.916\n0.85\n0.750.102\n0.09\n-0.13 -0.04\nFIG. 10: (Color online) Temperature dependencies of metast able states’ magnetizations, susceptibility and heat capa city in a\nfieldh= 0.13aforβ= 5 corresponding to the path A. Insets show the high-tempera ture anomalies.\ntailed picture of mZFCandmFCbehavior for path A1\nis shown in Fig.12. The qualitativly similar temperature\ndependences of these magnetizations are found in local\nmean-field simulations7and in field-heating after zero-\nfield-cooling experiments in PMN18.\nThe applied field cause also some peculiarities in tem-\nperature dependencies of ¯ m,χandCwhen ferro-phase\nis intermediate between para- and glassy ones (path B).\nAs Fig. 13 shows they appear in field-heating regime of\ncrystal being initially in the most disordered state with\nn+= 1/2. After reaching the limit of stability of this\nstate the crystal enters the sequence of states with larger\nn+on the boundaries of their existence regions until it\njoins the fully homogeneous ferro-state with n+= 1.\nThis process is accompanied by the spikes in χandC.\nAlong with it the reverse process of the demagnetizationof almost homogeneous initial state on heating is possi-\nble in the infinitesimal field. It is shown in Fig. 5 by the\ndirected dashed lines. The presence of some small field\nhere is necessary to remove the ferromagnetic instability\nof disorderedstates, without the field the inhomogeneous\nstates will jump right to the ferro-state due to this insta-\nbility.\nTheexhaustivestudyofboththeseprocesseswasmade\nin the graphite-intercalated magnet Cu0.93Co0.07Cl2\nhaving the phase sequence of path B22. The similarity of\nthe experimental data to ¯ mshown in Figs. 13, 5 is quite\nimpressive as authors has revealed the existence of nu-\nmerous metastable states using multiple cooling-heating\ncycles, see the example in Fig. 13. Also in Ref.[23]\n”thermal remagnetization” similar to that in Fig. 13\nwas observed in disordered ferromagnets Nd3(Fe,Ti)2913\nC/T\n-1.51.5\nm\n-2 0 τg/an+\n1\n0.9\n0.75\n0.65\n0.5\n0.4\n0.3\n0.2\n0.1\n0211.2n+\n1\n0.985\n0.967\n0.95\n0.935\n0.921\nχ\n0.1 -1.20.3 0.650.74\n-0.1 0τg/a-1.2τg/a0.103.5\nn+\n1\n0.985\n0.967\n0.95\n0.935\n0.9210 -0.100.55\nFIG. 11: (Color online) Temperature dependencies of metast able states’ magnetizations, susceptibility and heat capa city in a\nfieldh= 0.2aforβ= 0.1 corresponding to the path A1. Insets show the high-temperature anomalies.\nτg/a1.2\n0\n-1.5 0.5mn+\n1\n0.9\n0.8\n0.6\n0.5\n0.4FC\nZFC(A)\nZFCFC(B)\n-1.50.5 τg/a1.2\n0mn+\n1\n0.9\n0.8\n0.6\n0.5\n0.4\nFIG. 12: (Color online) Temperature dependencies of mZFCandmFCon path A1in fieldh= 0.3a(A)andh= 0.35a(B).\n-1.52\n-6 40m\nτg/a02.5\n-2.6 -1.4n+\n1\n0.9\n0.8\n0.65\n0.5\n0.4\n0.3\n0.2\n0.1\n0n+\n0.5\n0.6\n0.7\n0.8\n0.95\n1χ\nτg/a -2.6 -1.4 τg/a2.8\n0C/Tn+\n0.5\n0.6\n0.7\n0.8\n0.95\n1\nFIG. 13: (Color online) Temperature dependencies of metast able states’ magnetizations, susceptibility and heat capa city in a\nfieldh=aforβ= 0.1 corresponding to the path B.14\nandNd3(Fe,Re)29. The depolarization and polariza-\ntion on heating were also seen in PLZT ferroelectric\nceramics24,25.\nAt last we should mention the example of the relaxor\nthe thermodynamics of which is definitely distinct from\nthat of the present model. It is PMN-PT with lead ti-\ntanate (PT) content between 0.06 and 0.2 in which the\ntemperature evolution of hysteresis loop shows the grad-\nual transition from the inclined to the rectangularform33\ninstead of the sharp one as in Fig. 6. This means that\nbetween glass and ferro-phase in this compound there\nare the sequence of mixed phases where the most stable\nstates are those with partially ordered dipole moments.\nProbably the model with the three- and four - mode in-\nteractions can describe such intermediate mixed phases.\nV. DISCUSSION AND CONCLUSIONS\nThus the present thermodynamic theory can describe\nin a unified manner a wealth of nonergodic phenomena\nin disordered ferromagnets and ferroelectrics. It demon-\nstrates what the effects of numerous metastable states\ncan actually be and how they can manifest themselves\nin various regimes of real-time quasi-static experiments.\nAlso it shows what a form the results of rigorous statis-\ntical mechanics should have to describe the phase tran-\nsitions in crystals with random interactions.\nThe important conclusion can be made on the rela-\ntions between the predictions of eqiulibrium thermody-\nnamics valid in the infinite-time limit and finite-time ex-\nperiments. Contrary to naive expectations that there are\nno such relations the present study shows that they canberevieledinfinite-fieldquasi-staticexperiments. Inpar-\nticular, they are straightfordly manifested in the change\nof hysteresis loop form closely related to the form of the\nstatic thermodynamic potential. Yet its form and ther-\nmodynamic path for a given crystal can be obtained in\nsome other sufficiently full set of finite-field quasi-static\nexperiments. ’Sufficiently full’ means here that a wide\nspectrum of inhomogeneous metastable states should be\nexplored to determine the regions of their existence and\ntheir thermodynamic properties. The exellent examples\nof such studies providing the valuable data for the future\ntheory are given in Refs. [18, 22].\nApparantly the present phenomenology can be im-\nproved and expanded in many ways. Thus one may\nconsider a multicomponent order parameter as, say, in\nHeisenberg ferromagnet or cubic ferroelectric and less\nsparse modes having three- and four-mode couplings in\nLandau’s potential. Also one may include in it higher-\norder terms in mito expand the results to larger Tand\nhregions. Eventually it may be worth to explore the\nmodels with the broken in some way permutation sym-\nmetry. It seems also important to consider the role of\nfluctuations (i. e. the modes which do not condense)\nin the present approach. The quantitative description of\nthe nonergodic thermodynamics of PMN - type relaxors\nneeds also the consideration of random-fields.\nAcknowledgments\nI gratefully acknowledge useful discussions with M.P.\nIvliev, I.P. Raevski, V.B. Shirokov, V.I. Torgashev, E.D.\nGutlianskii, S.A. Prosandeev, V.P. Sakhnenko.\n∗Electronic address: timonin@aaanet.ru\n1K. Binder and A. P. Young , Rev. Mod. Phys. 58, 801\n(1986).\n2S. F. Edwards and P. W. Anderson , J. Phys. F 5, 965\n(1975).\n3D. R. Denholm and T. J. Sluckin , Phys. Rev. B 48,\n901 (1993).\n4S. Khmelevskyi , J. Kudrnovsky , B. L. Gyorffy , P.\nMohn , V. Drchal and P. Weinberger , Phys. Rev. B\n70, 224432 (2004).\n5M. J. Stephen and A. Aharony , J. Phys. C 14, 1665\n(1981).\n6P. N. Timonin , JETP 89, 525 (1999).\n7C. M. Soukoulis , K. Levin , and G. S. Grest , Phys.\nRev. B 28, 1495 (1983).\n8C. M. Soukoulis , G. S. Grest , and K. Levin , Phys.\nRev. B 28, 1510 (1983).\n9W. M. Saslow , and G. Parker , Phys. Rev. Lett. 56,\n1074 (1986).\n10H. G. Katzgraber , F. Pazmandi , C. R. Pike , Kai Liu\n, R. T. Scalettar , K. L. Verosub , and G. T. Zimanyi ,\nPhys. Rev. Lett. 89, 257202 (2002).\n11F. Pazmandi , G Zarand , and G. T. Zimanyi , Phys.Rev. Lett. 83, 1034 (1999).\n12V. Banerjee , S. Das and S. Puri , Phys. Rev. E 71,\n026105 (2005).\n13E. Vives and A. Planes , Phys. Rev. B 63, 134431\n(2001).\n14B. Yucesoy and A. N. Berker , Phys. Rev. B 76, 014417\n(2007).\n15H. Kaya and A. N. Berker , Phys. Rev. E 62, R1469\n(2000).\n16A. Kabakcioglu , A. N. Berker , and M. C. Yalabik\nPhys. Rev. E 49, 2680 (1994).\n17G. A. Smolenskii, Sov. Phys. Sol. St. 23, 784 (1981).\n18Z.- G. Ye and H. Schmid, Ferroelectrics 145, 83 (1993).\n19R. Sommer , N. K. Yushin and J. J. van der Klink ,\nPhys. Rev. B 48, 13230 (1993).\n20M T. Escote , C. H. Westphal and R. F. Jardim , J.\nAppl. Phys. 87, 5908 (2000).\n21Y. Yeshurun , M. B. Salamon , K. V. Rao and\nH. S. Chen , Phys. Rev. B 24, 1536 (1981).\n22I. S. Suzuki and M. Suzuki , Phys. Rev. B 73, 094448\n(2006).\n23S. J. Collocott , J. B. Dunlop and P. A. Watterson ,\nPhys. Rev. B 76, 054434 (2007).15\n24D. Viehland , J. F. Li , S. J. Jang , L.E. Cross and\nM. Wuttig , Phys. Rev. B 46, 8013 (1992).\n25Z. Kutnjak , C. Filipic and A. Levstik , J. Eur. Ceram.\nSoc.21, 1313 (2001).\n26P. N. Timonin, JETP 92, 1038 (2001); cond-mat/0004168.\n27M. Janssen , Int. J. Mod. Phys. B 8, 943 (1994); Phys.\nRep.295, 1 (1998); F. Evers and A. D. Mirlin , cond-\nmat/0707.4378.\n28R. B. Griffiths , Phys. Rev. Lett. 23, 17 (1969).\n29A. J. Bray , Phys. Rev. Lett. 59, 586 (1987).30J. A. Hertz, D. Sherrington and Th. M. Nieuwenhuizen,\nPhys. Rev. E 60, R2460 (1999).\n31D. H. Kim , G. J. Rodgers, B. Kahng, and D. Kim ,\nPhys. Rev. E 71, 056115 (2005).\n32S. Niidera , S. Abiko and F. Matsubara , Phys. Rev. B\n72, 214402 (2005).\n33S. M. Emelyanov , F. I. Savenko , Yu. A. Trusov ,\nV. I. Torgashev and P. N. Timonin , Phase Transitions\n45, 251 (1993)." }, { "title": "1907.12306v1.Magnetization_Dynamics_in_Holographic_Ferromagnets__Landau_Lifshitz_Equation_from_Yang_Mills_Fields.pdf", "content": "Magnetization Dynamics in Holographic\nFerromagnets:\nLandau-Lifshitz Equation from Yang-Mills Fields\nNaoto Yokoi1;2\u0003, Koji Sato2y, and Eiji Saitoh1;2;3;4;5z\n1Department of Applied Physics, The University of Tokyo, Tokyo 113-8656, Japan\n2Institute for Materials Research, Tohoku University, Sendai 980-8577, Japan\n3Advanced Institute for Materials Research, Tohoku University, Sendai 980-8577, Japan\n4Center for Spintronics Research Network, Tohoku University, Sendai 980-8577, Japan\n5Advanced Science Research Center, Japan Atomic Energy Agency, Tokai 319-1195, Japan\nAbstract\nWe introduce a new approach to understand magnetization dynamics in ferro-\nmagnets based on the holographic realization of ferromagnets. A Landau-Lifshitz\nequation describing the magnetization dynamics is derived from a Yang-Mills\nequation in the dual gravitational theory, and temperature dependences of the\nspin-wave sti\u000bness and spin transfer torque appearing in the holographic Landau-\nLifshitz equation are investigated by the holographic approach. The results are\nconsistent with the known properties of magnetization dynamics in ferromagnets\nwith conduction electrons.\n\u0003yokoi@spin.t.u-tokyo.ac.jp\nykoji.sato.c5@tohoku.ac.jp\nzeizi@ap.t.u-tokyo.ac.jparXiv:1907.12306v1 [hep-th] 29 Jul 20191 Introduction\nThe Landau-Lifshitz equation [1] is the fundamental equation for describing the dynamics\nof magnetization (density of magnetic moments) in various magnetic materials. It has been\nalso playing a fundamental role in the development of modern spintronics [2]: For instance,\nits extension to the coupled systems of localized magnetic moments and conduction electrons\nhas led to the concepts of spin transfer torque [3, 4] and spin pumping [5]. So far, the\nsymmetries and reciprocity in electronic systems have been the guiding principles to develop\nsuch extensions. In this article, we introduce another guiding principle to explore the new\nextensions and magnetization dynamics on the basis of the holographic duality.\nThe holographic duality is the duality between the quantum many body system de\fned\nind-dimensional space-time and the gravitational theory (with some matter \felds) which\nlives in (d+1)-dimensional space-time [6, 7, 8].1We constructed a holographic dual model of\nthree-dimensional ferromagnetic systems, which exhibits the ferromagnetic phase transition\nwith spontaneous magnetization and the consistent magnetic properties at low temperatures\n[10].2In the holographic duality, \fnite temperature e\u000bect in ferromagnetic systems can\nbe incorporated as the geometrical e\u000bect of black holes in higher dimensional bulk gravity,\nand the Wick rotation at \fnite temperatures is not required for the analysis in the dual\ngravitational theory. Thus, the novel analysis for real-time dynamics of quantum many\nbody systems in nonequilibrium situations can be performed using the holographic approach\n(for a review, see [14]). In addition, the holographic duality is known to be a strong-weak\nduality, which relates strongly correlated quantum systems to classical gravitational theories.\nFrom these viewpoints, the holographic approach can provide new useful tools to analyze\nnonequilibrium and nonlinear dynamics of magnetization in ferromagnets.\nIn ferromagnets, spin currents are generated by magnetization dynamics. From the holo-\ngraphic dictionary between the quantities of ferromagnets and gravitational theory [10], the\nspin currents in ferromagnets correspond to the SU(2) gauge \felds in the dual gravitational\ntheory. This correspondence indicates that the dynamics of spin currents, consequently the\ndynamics of magnetization, can be described by the Yang-Mills equation for SU(2) gauge\n\felds [15] in the holographic dual theory. In the following, we derive a Landau-Lifshitz\nequation for magnetization dynamics from the Yang-Mills equation within the holographic\nrealization of ferromagnets. This derivation can provide novel perspectives for magnetization\ndynamics from the non-abelian gauge theory.\n1See [9] for a recent review on the applications of the holographic duality to condensed matter physics.\n2Other holographic approaches to ferromagnetic systems have been also discussed in [11, 12, 13].\n2This article is organized as follows. In Section 2, we summarize the results of the mag-\nnetic properties obtained from the holographic realization of ferromagnets in thermodynamic\nequilibrium. An extension to nonequilibrium situation including the \ructuations of magne-\ntization and spin currents is discussed in the dual gravitational theory, and the holographic\nequation of magnetization dynamics is derived in Section 3. In Section 4, temperature depen-\ndences of the parameters in the resulting holographic equation are investigated by numerical\ncalculations. Finally, we summarize the results in Section 5.\n2 Holographic Dual Model of Ferromagnets\nWe begin with a brief summary on the holographic dual model of ferromagnets [10]. The\ndual model is the \fve-dimensional gravitational theory with an SU(2) gauge \feld Aa\nMand a\nU(1) gauge \feld BM, whose action is given by\nS=Zp\u0000gd5x\u00141\n2\u00142(R\u00002\u0003)\u00001\n4e2GMNGMN\u00001\n4g2Fa\nMNFaMN\n\u00001\n2(DM\u001ea)2\u0000V(j\u001ej)\u0015\n: (1)\nHere,Ris the scalar curvature of space-time, and the \feld strength is de\fned by Fa\nMN=\n@MAa\nN\u0000@NAa\nM+\u000fabcAb\nMAc\nNandGMN=@MBN\u0000@NBM, respectively. The index alabels\nspin directions in the SU(2) space (a= 1\u00183), the index Mlabels space-time directions in\n\fve dimensions ( M;N = 0\u00184), and\u000fabcis a totally anti-symmetric tensor with \u000f123= 1.\nThe model also includes a triplet scalar \feld \u001eawith the covariant derivative DM\u001ea=\n@M\u001ea+\u000fabcAb\nM\u001ec, and theSU(2)-invariant scalar potential V(j\u001ej) with the norm j\u001ej2=\nP3\na=1(\u001ea)2. Note that the scalar \feld is neutral under the U(1) gauge transformation. In\norder to guarantee asymptotic Anti-de Sitter (AdS) backgrounds, the negative cosmological\nconstant \u0003 =\u00006=`2is introduced. The \feld-operator correspondence in the holographic\nduality [7, 8] leads to the following holographic dictionary between the \felds of the dual\ngravitational theory and the physical quantities of ferromagnets:\nDual gravity Ferromagnet\nScalar \feld \u001ea() Magnetization Ma\nSU(2) gauge \feld Aa\nM() Spin current Ja\ns\u0016\nU(1) gauge \feld BM() Charge current J\u0016\nMetricgMN() Stress tensor T\u0016\u0017\nTable 1: Holographic dictionary between the dual gravitational theory and ferromagnets.\n32.1 Black Hole as Heat Bath\nIn order to establish the holographic dictionary, thermodynamical properties of the physical\nquantities of ferromagnets should be calculated in the dual gravitational theory. In Ref. [10],\nthe temperature dependences of magnetic quantities and the behavior of ferromagnetic phase\ntransition are thoroughly discussed. In the context of the holographic duality, \fnite tem-\nperature e\u000bects in the ferromagnets can be incorporated by introducing the black holes into\nthe dual gravitational theory as the background space-time. Indeed, the dual gravitational\ntheory has the charged black hole solution which is a solution to the Einstein, Yang-Mills,\nand Maxwell equations derived from the action (1):\nRMN+\u0012\n\u0003\u00001\n2R\u0013\ngMN=\u00142\n2e2\u0012\n2GKMGKN\u00001\n2GKLGKLgMN\u0013\n+\u00142\n2g2\u0012\n2Fa\nKMFaKN\u00001\n2FaKLFaKLgMN\u0013\n; (2)\nrMFaMN+\u000fabcAb\nMFcMN= 0;rMGMN= 0; (3)\nwhererMis the covariant derivative for the a\u000ene connection, and the space-time indices\nM;N are raised or lowered by the bulk metric gMN. Here, we neglect the contribution from\nthe scalar \feld and set \u001ea= 0 for the background. The metric of the black hole3is given by\nds2=gMNdxMdxN=r2\n`2\u0000\n\u0000f(r)dt2+dx2+dy2+dz2\u0001\n+`2\nf(r)dr2\nr2; (4)\nwith the radial function,\nf(r) = 1\u0000(1 +Q2)\u0010rH\nr\u00114\n+Q2\u0010rH\nr\u00116\n: (5)\nHere, we de\fne the parameter Q:\nQ2=2\u00142\n3\u0012\u00162\ne\ne2+\u00162\ns\ng2\u0013\n: (6)\nTheU(1) charge\u0016eandSU(2) charge\u0016sof the black hole are supported by the time com-\nponents of the gauge \felds,\nB0=\u0016e\u0010rH\n`\u0011\u0012\n1\u0000r2\nH\nr2\u0013\nandA3\n0=\u0016s\u0010rH\n`\u0011\u0012\n1\u0000r2\nH\nr2\u0013\n: (7)\nNote that the black hole solution (4) is asymptotically AdS at r!1 , and has the (outer)\nhorizonr=rH.\n3This type of non-abelian black holes has been discussed in the context of the holographic duality, in the\nliterature such as [11, 16].\n4For the following discussion, we make a coordinate change of the radial coordinate rinto\nubyu= 1=r, and the black hole metric becomes\nds2=1\nu2\u0012\n\u0000f(u)dt2+dx2+dy2+dz2+du2\nf(u)\u0013\n; (8)\nand the transformed function f(u) is given by\nf(u) = 1\u0000\u0000\n1 +Q2\u0001\nu4+Q2u6; (9)\nwhere we have set the coupling parameters e=g= 1 and the black hole parameters rH=\n`= 1, for simplicity.\nIn the holographic dual model, the black hole (8) plays the role of the heat bath; due to\nthe Hawking radiation, the black hole temperature is given by\nT=2\u0000Q2\n2\u0019; (10)\nand the calculations on the black hole background lead to the thermodynamical properties\nof the corresponding ferromagnet. Since we focus only on the dynamics of magnetization\nand spin current, the background space-time is \fxed to be the black hole metric (8) in the\nfollowing.\n2.2 Thermodynamics of Ferromagnets from Scalar Dynamics on Charged\nBlack Hole\nIn order to investigate the thermodynamics of magnetization, we examine the equation of\nmotion for the scalar \feld \u001ea, which is also derived from the action (1):\n1p\u0000g@M\u0000p\u0000gDM\u001ea\u0001\n+\"abcAb\nMDM\u001ec=@V\n@\u001ea: (11)\nHere, we consider a static and homogeneous solution in the boundary coordinates, x\u0016=\n(t;x1;x2;x3), which corresponds to the homogeneous magnetization in ferromagnets. With-\nout loss of generality, the ansatz for such a scalar \feld, which is invariant under the transla-\ntions on the boundary, is given by\n\u001e1=\u001e2= 0; \u001e3= \b(u)6= 0: (12)\nInserting this ansatz, the metric (8), and the gauge \felds (7) into the equation (11), we obtain\nthe following equation for \b( u):\nu2f(u)d2\b\ndu2+\u0012\nu2df(u)\ndu\u00003uf(u)\u0013d\b\ndu=@V\n@\b: (13)\n5This equation governs the thermodynamics of magnetization in the dual gravitational theory.\nWe can analyze the solution to this equation numerically with a simple quartic potential\nV(j\u001ej) =\u0015\u0000\nj\u001ej2\u0000m2=\u0015\u00012=4, and the asymptotic behavior of the numerical solution near the\nboundaryu\u00180 (orr\u00181) is obtained:\n\b(u)'H0u\u0001\u0000+M(T)u\u0001+\u0010\n\u0001\u0006= 2\u0006p\n4\u0000m2\u0011\n: (14)\nAccording to the standard recipe in the holographic duality [17, 18], the coe\u000ecients H0and\nM(T) in the asymptotic expansion correspond to an external magnetic \feld and a magne-\ntization at temperature T(underH0), respectively. In Ref. [10], the resulting temperature\ndependences of magnetization, magnetic susceptibility, and speci\fc heat have been shown\nto reproduce the ferromagnetic phase transition in the mean \feld theory. Furthermore, the\ntemperature dependences at low temperatures are also consistent with the existence of the\nspin wave excitations (magnons) and conduction electrons in low-temperature ferromagnets.\nFor later convenience, we also comment on the solutions for the gauge \felds. Assuming\nthe translational and rotational invariance on the boundary, equilibrium solutions for the\ngauge \felds are given by the following form:\nB0=b(u) andA3\n0=a3(u); (15)\nwhere all the other components vanish. Inserting this ansatz and (12), the Maxwell and\nYang-Mills equations on the black hole are reduced to the following simple forms:\nd\ndu\u00121\nudb\ndu\u0013\n= 0 andd\ndu\u00121\nuda3\ndu\u0013\n= 0: (16)\nThe general solutions are given by the forms (7) in terms of u,\nb(u) =\u0016e\u0000\n1\u0000u2\u0001\nanda3(u) =\u0016s\u0000\n1\u0000u2\u0001\n: (17)\nHere, we impose the boundary conditions B0= 0 andA3\n0= 0 at the horizon ( u= 1), which\nguarantee the regularity of the gauge \felds on the horizon. The remaining integral constants,\n\u0016eand\u0016s, correspond respectively to the electrochemical potential of underlying electrons\nand the spin chemical potential (or spin voltage), through the holographic dictionary.\nTo summarize, the solutions (14) and (17) on the charged black hole describe the ther-\nmodynamical property of the holographic dual ferromagnets in the equilibrium.\n3 Magnetization Dynamics in Holographic Ferromagnets\nIn this section, we extend the holographic analysis in the equilibrium, summarized in the\nprevious section, to more general situations including the dynamics of magnetization and\n6spin currents. In order to discuss the dynamics of magnetization and spin currents, the\nstatic and homogeneous ansatze for the scalar \feld (12) and the gauge \felds (15) need to be\ngeneralized. Here, we focus on the dynamics with the long wave length in the ordered phase\n(symmetry broken phase) below the Curie temperature, where various phenomena in modern\nspintronics are intensively studied.\n3.1 Generalized Ansatz and E\u000bective Equations of Motion\nFor the scalar \feld, following the standard derivation of the equation for magnetization\ndynamics, we consider the generalized ansatz for the scalar \feld as a factorized form:\n\u001ea(u;t;x ) = \b(u)na(t;x) with3X\na=1nana= 1; (18)\nwhere \b(u) is a solution of the equation (13) with the asymtotic behavior (14). Note that,\nsince we focus only on the dynamics of spontaneous magnetization, we \fx H0= 0 throughout\nthis article. In this ansatz, na(t;x) corresponds to the (local) direction of magnetization in\nferromagnets.\nIn ferromagnetic systems, the magnetization dynamics generates various dynamics of spin\ncurrents [2]. In the holographic dual theory, the scalar dynamics is also expected to induce\nthe dynamics of the corresponding SU(2) gauge \feld, and thus we generalize the static and\nhomogeneous ansatz for the SU(2) gauge \felds to the following factorized forms:\nAk\n0(u;t;x ) = (1\u0000u2)ak\n0(t;x);\nA?\n0(u;t;x ) = (1\u0000u2)a?\n0(t;x);\nAk\ni(u;t;x ) =Gk(u)ak\ni(t;x);\nA?\ni(u;t;x ) =G?(u)a?\ni(t;x) (i= 1\u00183); (19)\nwhere we set the radial component Aa\nu\u00110 by using the gauge degrees of freedom. Due to the\nnontrivial scalar solution \b( u), corresponding to the spontaneous magnetization, the SU(2)\ngauge symmetry is broken to U(1). The gauge \felds can be correspondingly decomposed into\nan unbroken component Ak\n\u0016and two broken components A?\n\u0016, which are de\fned by Ak\n\u0016/na\nandn\u0001A?\n\u0016= 0, respectively. As in the case of the static solutions, the time components\nof gauge \felds should satisfy the horizon boundary condition, Aa\n0= 0 atu= 1, for the\nregularity. Although the spatial components Aa\niare not required to vanish on the horizon,\nthe regularity (or \fniteness) at the horizon is required. The asymptotic solutions to the\nlinearized Yang-Mills equation near the boundary ( u\u00180) give the asymptotic expansions for\n7the radial functions Gk(u), andG?(u),\nGk(u) = 1\u0000\u001bk\nsu2+O(u4);\nG?(u) = 1 +\u001b?\nsu2+O(u4): (20)\nWe discuss the concrete numerical solutions of Ga(u) and their physical implications in the\nnext section.\nSince the scalar \feld \u001eadoes not have the U(1) charge, the \ructuation (or dynamics) of\n\u001eadoes not induce further dynamics for the U(1) gauge \feld, which implies the solution for\nB\u0016in (7) is unchanged, and we can neglect the dynamics of B\u0016.\nAt \frst, we consider the equation of motion for the scalar \feld \u001ea. Inserting the generalized\nansatz (18) into the equation (11), we obtain the following equation for na:\n\u0014\nu5@u\u0000\nu\u00003f(u)@u\b\u0001\n\u0000@V\n@\b\u0015\nna=\u0014u2\nf(u)DtDtna\u0000u2DiDina\u0015\n\b: (21)\nHere, we have used the gauge condition Aa\nu= 0, and the gauge covariant derivative is de\fned\nasD\u0016na=@\u0016na+\"abcAb\n\u0016nc. The left-hand side of the equation (21) is proportional to the\nequation (13), and thus vanishes for the solution \b( u). Since \b( u) is a non-trivial solution,\nwhich is not identically zero, we have the e\u000bective equation of motion for na:\nf\u00001DtDtna\u0000DiDina= 0: (22)\nNext, the equation of motion for the gauge \felds is considered. The Yang-Mills equation\nfor theSU(2) gauge \feld Aa\nMis derived by the variation of the holographic action (1) and\ngiven by\n1p\u0000g@N\u0000p\u0000gFNMa\u0001\n+\u000fabcAb\nNFNMc=JMa; (23)\nwhere theSU(2) current is de\fned as\nJa\nM=\"abc\u001ebDM\u001ec=\"abc\u001eb\u0010\n@M\u001ec+\"cdeAd\nM\u001ee\u0011\n: (24)\nUnlike the static case, the generalized ansatz (18) and (19) give the non-vanishing currents:\nJa\n\u0016= \b2\u0010\n\"abcnb@\u0016nc+\"abc\"cdenbad\n\u0016ne+O\u0000\nu2\u0001\u0011\n: (25)\nNote that the radial component of the currents still vanishes, Ja\nu\u00110, due to the gauge \fxing\nconditionAa\nu\u00110. With this current, we can explicitly write down the Yang-Mills equations\non the charged black hole (8), in the boundary direction :\nJa\n0=u3f@u\u0000\nu\u00001Fa\nu0\u0001\n+u2(DiFa\ni0); (26)\nJa\ni=u3@u\u0000\nu\u00001fFa\nui\u0001\n\u0000u2f\u00001(D0Fa\n0i) +u2\u0000\nDjFa\nji\u0001\n; (27)\n8where the gauge covariant derivative for the \feld strength is de\fned as D\u0016Fa\n\u0017\u001a=@\u0016Fa\n\u0017\u001a+\n\"abcAb\n\u0016Fc\n\u0017\u001a. Inserting the ansatz (19), the Yang-Mills equations give the equations for naand\naa\n\u0016. In summary, using the generalized ansatze, we have obtained the coupled equations of\nmotion for naandaa\n\u0016, (22), (26), and (27).\n3.2 Landau-Lifshitz Equation from Yang-Mills Equation\nSince it is di\u000ecult to \fnd the general solutions for the coupled non-linear partial di\u000berential\nequations, we seek simple trial solutions for naandaa\n\u0016to obtain the e\u000bective equations of\nmotion. At \frst, instead of looking for general solutions to the equation (22), we consider\nthe solutions to the simpler equations:\nDtna= 0 and Dina= 0; (28)\nwhich are explicitly given by\n@\u0016na+\u000fabcab\n\u0016nc= 0 +O(u2): (29)\nThese equations lead to the ground state solutions for the e\u000bective Hamiltonian for na:\nHe\u000b=f\n2(\u0019a)2+1\n2(Dina)2; (30)\nwhere the conjugate momentum is de\fned by \u0019a=f\u00001Dtna. In this article, we wish to\ndiscuss the dynamics of magnetization and spin currents in the boundary ferromagnetic\nsystem, which is given by the leading terms in the asymptotic expansions at u\u00180. Hence,\nthe higher order terms in the expansion with respect to uare irrelevant, and we neglect them\nin the following. Dropping the O(u2) term, we can easily obtain the solution to (29) for aa\n\u0016\nin terms of na,\naa\n\u0016=C\u0016na\u0000\"abcnb@\u0016nc; (31)\nwhere we have introduced a vector \feld C\u0016which is arbitrary at this stage. This solution\ndemonstrates the clear separation of the gauge \felds:\nak\n\u0016=C\u0016naanda?\n\u0016=\u0000\"abcnb@\u0016nc: (32)\nThe relation for the broken components, a?\n\u0016, is nothing but a non-abelian analogue of the\nrelation between the gauge \feld and the quantum phase of Cooper pair, A\u0016=@\u0016\u0012, in super-\nconductivity, and also corresponds to the Maurer-Cartan one-form of G=H\u0018SU(2)=U(1)\nin terms of the Nambu-Goldstone modes na[19, 20]. Requiring the matching condition to\n9the static solution (17), aa\n0=\u0016s\u000ea3andaa\ni= 0 forna= (0;0;1), the vector \feld C\u0016should\nsatisfy the condition:\nC0=\u0016sandCi= 0; (33)\nin the static and homogeneous limit. Note that the relation (31) and the ansatz (18) do not\ninduce new contributions of the scalar \felds to the energy-momentum tensor TMNin the\nEinstein equations, and consequently the analysis in the probe approximation remain intact.\nNext, we consider the e\u000bective Yang-Mills equations, (26) and (27). It is not di\u000ecult to\nshow that the relation (31) leads to vanishing currents Ja\n\u0016up toO(u2), using the explicit\nform (25). Furthermore, the ansatz for gauge \felds (19) with Aa\nu= 0 implies\n@u\u0000\nu\u00001Fa\nu0\u0001\n= 0;and@u\u0000\nu\u00001fFa\nui\u0001\n= 0 +O(u4): (34)\nDropping the higher order terms such as O(u4), the remaining Yang-Mills equations reduce\nto\nDiFa\ni0= 0;andD0Fa\n0i+DjFa\nji= 0: (35)\nFrom the viewpoint of the boundary theory (on the ferromagnet side), the \frst equation\ncorresponds to a non-abelian version of Gauss's law, and the second corresponds to a non-\nabelian version of Ampere's law without source and currents, for the spin gauge \felds [21].\nUsing the relation (31), we obtain the SU(2) \feld strength,4\nFa\n\u0016\u0017=nah\n(@\u0016C\u0017\u0000@\u0017C\u0016)\u0000\"bcdnb@\u0016nc@\u0017ndi\n\u0011naf\u0016\u0017: (36)\nNote that a component of the \feld strength, f\u0016\u0017, parallel to the magnetization naonly\nremains. With the \feld strength (36), the e\u000bective Yang-Mills equations (35) and the Bianchi\nidentity for the SU(2) gauge \feld are reduced to the following equations:\n@\u0016f\u0016\u0017= 0 and \u000f\u0016\u0017\u001a\u001b@\u0017f\u001a\u001b= 0: (37)\nThe above equations are the same form as the Maxwell equations, and the terms depending\nonnain the gauge \feld f\u0016\u0017actually corresponds to the so-called spin electromagnetic \feld\ndiscussed in the study on ferromagnetic metals [22, 21]. The gauge \feld (36) also corre-\nsponds to the unbroken U(1) gauge \feld upon the symmetry breaking from SU(2) toU(1),\nwith a space-dependent order parameter, which is frequently discussed in the context of\nsolitonic monopoles in non-abelian gauge theories [23]. Since the unbroken gauge \felds in\n4We used the relation \"abc@\u0016nb@\u0017nc=na\"bcdnb@\u0016nc@\u0017nddue toP\nanana= 1.\n10the holographic dual theory are identi\fed as the (exactly) conserved currents in the bound-\nary quantum system, the gauge \feld C\u0016is naturally identi\fed as the spin current with the\npolarization parallel to the magnetization na, which originates from conduction electrons.\nSo far, we have obtained the relation between the gauge \feld aa\n\u0016and the (normalized)\nmagnetization na, which implies that the gauge \feld dynamics can be solely reduced to the\ndynamics of the magnetization and the spin electromagnetic \feld C\u0016. Finally, we consider\nthe remaining Yang-Mills equation in the radial u-direction, in the holographic dual theory :\n1p\u0000g@\u0016\u0000p\u0000gF\u0016ua\u0001\n+\u000fabcAb\n\u0016F\u0016uc=Jua: (38)\nThis equation is derived by the variation of the radial u-component of the SU(2) gauge \felds\nand speci\fes the dynamics of the gauge \felds in the \fve-dimensional bulk; this equation\ncannot be seen in the ferromagnetic system on the boundary. With the ansatz (18), the\nradial component of the current also vanishes ( Ja\nu\u00110), and the gauge \fxing condition\nAa\nu\u00110 leads to the simple SU(2) \feld strength Fa\n\u0016u=\u0000@uAa\n\u0016such as\nFk\n0u= 2uak\n0(t;x); F?\n0u= 2ua?\n0(t;x);\nFk\niu= 2u\u001bk\nsak\ni(t;x); F?\niu=\u00002u\u001b?\nsa?\ni(t;x); (39)\nwhere we used the ansatz (19) and discarded the irrelevant O(u3) terms. From these forms,\nthe second term in the left-hand side of (38) automatically vanishes due to \"abcab\n\u0016ac\n\u0016= 0.\nInserting the forms of \feld strength (39) and the relation (31), the equation can be recast as\nthe following form:\n@0\u0010\nC0na\u0000\"abcnb@0nc\u0011\n\u0000@i\u0010\n\u001bk\nsCina+\u001b?\ns\"abcnb@inc\u0011\n= 0; (40)\nwhere the subleading terms are neglected. Here, we can write down the e\u000bective equation of\nmotion for the magnetization nain our holographic dual model:\nC0_na\u0000\"abcnbnc\u0000\u001b?\ns\"abcnbr2nc\u0000\u001bk\nsCi@ina= 0; (41)\nwhere the dot denotes the time-derivative and r2=@i@i.5Here, we consider the condition,\n@0C0\u0000\u001bk\ns@iCi= 0, on the unbroken gauge \feld due to the constraintP\nanana= 1. This\ncondition implies the conservation of the spin current of conduction electrons, which corre-\nsponds to the unbroken gauge \feld C\u0016, as seen below. Note that, since the Maxwell equations\n(37) forC\u0016is gauge invariant, this condition can be consistently imposed as a gauge \fxing\ncondition.\n5A simlar analysis on e\u000bective equations at the linearized level in two-dimensional magnetic systems has\nbeen also discussed in [11].\n11Considering the matching condition (33), we decompose C0intoC0=\u0016s+~C0. Finally,\nwe obtain the holographic equation for magnetization dynamics:\n\u0016s_na\u0000\"abcnbnc\u0000\u001b?\ns\"abcnbr2nc+~C0_na\u0000\u001bk\nsCi@ina= 0: (42)\nHere, we take the spin chemical potential to be \u0016s=\u0000Ms=\rwith the magnitude of spon-\ntaneous magnetization Msand the gyromagnetic ratio \r(>0),6and also identify the spin\ncurrent and spin accumulation due to conduction electrons as Jk\nsi=\u0000\u001bk\nsCiand \u0001\u0016s=~C0,\nusing the holographic dictionary. Then, the holographic equation becomes the same form as\nthe Landau-Lifshitz equation (without damping terms),\nMs\n\r_na+\"abcnbnc=\u0000\u001b?\ns\"abcnbr2nc+ \u0001\u0016s_na+Jk\nsi@ina: (43)\nThe last two terms in the right-hand side can be interpreted as the well-known terms from spin\ntransfer torque, which describes the transfer of spin angular momentum between localized\nmagnetic moments and conduction electrons [21]. Furthermore, the holographic Landau-\nLifshitz equation (42) also naturally incorporate the spin inertia term proportional to the\nsecond time-derivative of the magnetization, which is discussed in metallic ferromagnets [24].\nIt should be noted that the holographic Landau-Lifshitz equation automatically incor-\nporates the spin transfer torque due to conduction electrons without introducing the cor-\nresponding \felds to electrons in the dual gravitational theory. This is consistent with the\nthermodynamical results at low temperatures, which was obtained in the previous paper [10].\n4 Phenomenology of Holographic Magnetization Dynamics\nIn the isotropic ferromagnets su\u000eciently below the Curie temperature ( T\u001cTc), the dynamics\nof magnetization vector (or density of magnetic moments), Ma, is described by the Landau-\nLifshitz equation [1, 26]:\n@Ma\n@t=\u0000\u000b\u000fabcMbr2Mcwith3X\na=1MaMa=M(T)2= const. (44)\nIn the following discussion, the external magnetic \feld and the damping term (or relaxation\nterm) are ignored for simplicity. From the quadratic constraint, the magnetization vector\ncan be represented as Ma(x;t) =M(T)na(x;t) with the unit vector na(x;t). In terms of\nna(x;t), the Landau-Lifshitz equation becomes\nM(T)@na\n@t=\u0000\u000bM(T)2\u000fabcnbr2nc: (45)\n6The negative sign is introduced due to the negative value of the gyromagnetic ratio for electrons.\n12Note that the equation has two parameters, the magnitude of spontaneous magnetization,\nM(T), at the temperature T, and the spin sti\u000bness constant, \u000b.\nComparing the holographic equation (42) with the Landau-Lifshitz equation (45), we \fnd\nthat the spin chemical potential, \u0016s, in the gauge \feld solution (17) should be proportional\nto the magnitude of magnetization, and the spin sti\u000bness constant is given by the coe\u000bcients\n\u001b?\nsin the gauge \feld solution (20) in the following way :\n\u0016s/\u0000M(T) and \u001b?\ns/\u000bM(T)2: (46)\nIn our holographic dual model, the magnitude of magnetization, M(T), at the temperature\nTis given by the static solution of the scalar \feld \b( u) through the formula (14). The\n\frst relation between the magnitude of magnetization and the spin chemical potential in\nferromagnets is well-known, and frequently used as the starting point to analyze the various\nspintronic phenomena [2].\nAlthough the spin chemical potential in the equilibrium, \u0016s, is an integration constant,\nthe coe\u000ecient, \u001b?\ns, is the derived quantity from the gauge \feld equation, and thus the second\nrelation in (46) on the spin sti\u000bness constant is a nontrivial consequence in the holographic\ndual model. In order to obtain the coe\u000ecient, \u001b?\ns, we consider the linearized equation of\nmotion for gauge \felds on the background solution, with the static and homogeneous ansatz,\nA?\ni=kG?(u), wherek= const.7Inserting this ansatz into the Yang-Mills equation (27),\nwe have the following linearized equation for G?(u):\nu3d\ndu\u0012f(u)\nu\u0012dG?\ndu\u0013\u0013\n+\u0000\nua3(u)\u00012\nf(u)G?= 0; (47)\nwhere the metric (8) and the SU(2) gauge \feld (17) are assumed to be the background.\nNote that this is a linear equation for G?, and the constant kis irrelevant. Here, we\nimpose the \frst relation in (46), \u0016s=\u0000M(T)=M(0), which is the magnetization normal-\nized by the saturated magnetization, M(T= 0).8Using the numerical results of the holo-\ngraphic spontaneous magnetization, M(T) in [10], which is obtained using the scalar poten-\ntialV(j\u001ej) =\u0015\u0000\nj\u001ej2\u0000m2=\u0015\u00012=4 with\u0015= 1 andm2= 35=9, we can numerically solve the\nequation (47) and obtain the asymptotic expansion (20) near the boundary ( u\u00180). The\nnumerical results of temperature dependences of the spin-wave sti\u000bness, D(T)'\u001b?\ns=M(T),\nwhich appears in the dispersion relation of spin-waves, !=D(T)k2, and the spin sti\u000bness\nconstant,\u000b(T)'\u001b?\ns=M(T)2, are shown in Figure 1.\n7The nontrivial pro\fle A?\nx(u) on the background does not contribute to the energy-momentum tensor in\nthe Einstein equation at the linearized level.\n8The proportionality constant is chosen for convenience in numerical calculations.\n13Figure 1: Temperature dependence of the spin-wave sti\u000bness is shown in Figure ( a): The\ndots are numerical results for D(T)=D(0), and the bold line is the magnetization curve,\nM(T)=M(0). Temperature dependence of the spin sti\u000bness constant, \u000b(T)=\u000b(0), is shown in\nFigure (b). All the results are calculated with the parameters, \u0015= 1 andm2= 35=9.\nThe results on the spin-wave sti\u000bness in Figure 1( a) clearly show that D(T)/M(T),\nwhich is consistent with the relation (46) based on the Landau-Lifshitz equation (44). Fur-\nthermore, the results in Figure 1( b) imply the slight temperature dependence of the spin\nsti\u000bness constant, \u000b=\u000b(T), which can be attributed to the nonlinear spin-wave e\u000bects [25].\nA similar argument also holds for the unbroken (or parallel) component of the gauge\n\felds,Ak\ni, and we can obtain the coe\u000ecient \u001bk\ns, which leads to the spin torque term in the\nholographic Landau-Lifshitz equation (42). The nontrivial pro\fle of gauge \feld, Ak\nx(u), which\nis the parallel component to the spin chemical potential, Ak\n0, leads to the non-vanishing o\u000b-\ndiagonal contribution in the right-hand side of the Einstein equation (2), and thus induces the\n\ructuation of the metric gtx(u) =htx(u)=u2, wherehtx(u) parameterizes the \ructuation \fnite\non the boundary. At the linearized level, two \ructuations, Ak\nx(u) andhtx(u), form the closed\nequations, which come from the Yang-Mills equation and Einstein equation, respectively\n[27, 28]:\nud\ndu \nf(u)\nu \ndAk\nx\ndu!!\n+\u0012da3(u)\ndu\u0013d\ndu\u0000\nu2htx\u0001\n= 0; (48)\nu\u00002d\ndu\u0000\nu2htx\u0001\n+ 2\u0012da3(u)\ndu\u0013\nAk\nx= 0: (49)\nDeleting the metric \ructuation, we can obtain the equation for Gk(u):\nud\ndu \nf(u)\nu \ndGk\ndu!!\n\u00002u2\u0012da3(u)\ndu\u00132\nGk= 0: (50)\nWe can numerically solve the equation, and obtain the coe\u000ecient \u001bk\nsfrom the asymptotic\nexpansion of the solution in (20). The resulting temperature dependence of the spin torque\ncoe\u000ecient, \u001cs(T) =\u001bk\ns=M(T), is shown in Figure 2.\n14Figure 2: Temperature dependence of the spin torque coe\u000ecient: The dots are numerical re-\nsults for\u001cs(T)=\u001cs(0), and the bold line is the magnetization curve M(T)=M(0). The dashed\nline is the \ftting curve, \u001cs(T)=\u001cs(0) =c(1\u0000T=Tc)2=5withc'1:41, near the Curie temper-\nature.\nThe results on the magnitude of the spin transfer torque, \u001cs(T), show that the spin torque\ne\u000bect is approximately constant at low temperatures (in comparison with magnetization\ncurve), and is vanishing towards the Curie temperature as \u001cs(T)/(1\u0000T=Tc)2=5. This\nproperty at low temperatures is consistent with the phenomenological form of the spin transfer\ntorque, ( Jk\ns\u0001rna)=M(T), whose magnitude is independent of the norm of magnetization due\ntojJk\nsj/M(T) at the leading order [21]. In addition, the \fnite spin torque coe\u000ecient is\na consequence of the both \ructuations of the gauge \feld and metric. In accordance with\nthe holographic dictionary [27], the metric \ructuation htxcorresponds to the temperature\ngradient,rxT=T, in the ferromagnetic system. This calculation implies that the e\u000bect of\nspin transfer torque appears only in the nonequilibrium situations, where spin transfer is\naccompanied by heat (or entropy) transfer.\n5 Summary and Discussion\nWe have discussed a novel approach to understand magnetization dynamics in ferromagnets\nusing the holographic realization of ferromagnetic systems. The Landau-Lifshitz equation de-\nscribing magnetization dynamics was derived from the Yang-Mills-Higgs equations in the dual\ngravitational theory. This holographic Landau-Lifshitz equation automatically incorporates\nnot only the exchange interaction but also the spin transfer torque e\u000bect due to conduction\nelectrons. Furthermore, we numerically investigated the temperature dependences of the\nspin-wave sti\u000bness and the magnitude of spin transfer torque in the holographic dual theory,\nand the results obtained so far are consistent with the known properties of magnetization\ndynamics in ferromagnets with conduction electrons.\n15This holographic approach to magnetization dynamics can be applied to more generic sit-\nuations. For instance, the holographic Landau-Lifshitz equation can incorporate the damping\nterm by considering more generic metric \ructuations, which correspond to phonon dynamics\nin the boundary ferromagnets. Moreover, the holographic dual theory may provide geomet-\nric approaches to spin caloritronics [29], where magnetization dynamics is considered under\ntemperature gradients, from higher dimensional perspectives. We thus believe that the holo-\ngraphic approach provides useful tools to analyze nonequilibrium and nonlinear dynamics\nof magnetization in ferromagnets, and also leads to new perspectives in spintronics from\ngravitational physics.\nAcknowledgement\nThe authors thank M. Ishihara for collaboration at the early stage of this work, and also K.\nHarii and Y. Oikawa for useful discussions. The works of N. Y. and E. S. were supported\nin part by Grant-in Aid for Scienti\fc Research on Innovative Areas \"Nano Spin Conversion\nScience\" (26103005), and the work of K. S. was supported in part by JSPS KAKENHI Grant\nNo. JP17H06460. The works of N. Y. and E. S. were supported in part by ERATO, JST.\nReferences\n[1] L. D. Landau and E. M. Lifshitz, \\On the theory of the dispersion of magnetic perme-\nability in ferromagnetic bodies,\" Phys. Z. Sowjet. 8, 153 (1935).\n[2] S. Maekawa, S. O. Valenzuela, E. Saitoh and T. Kimura (Eds.), \\Spin Current (Second\nEd.),\" Oxford University Press (2017).\n[3] L. Berger, \\Emission of spin waves by a magnetic multilayer traversed by a current,\"\nPhys. Rev. B 54, 9353 (1996).\n[4] J. C. Slonczewski, \\Current-driven excitation of magnetic multilayers,\" Journal of Magn.\nMagn. Mater. 159, L1 (1996).\n[5] Y. Tserkovnyak, A. Brataas, and G. E. Bauer, \\Enhanced Gilbert damping in thin\nferromagnetic \flms,\" Phys. Rev. Lett. 88, 117601 (2002).\n[6] J. M. Maldacena, \\The Large N limit of superconformal \feld theories and supergravity,\"\nInt. J. Theor. Phys. 38, 1113 (1999) [Adv. Theor. Math. Phys. 2, 231 (1998)] [hep-\nth/9711200].\n16[7] S. S. Gubser, I. R. Klebanov and A. M. Polyakov, \\Gauge theory correlators from non-\ncritical string theory,\" Phys. Lett. B 428, 105 (1998) [hep-th/9802109].\n[8] E. Witten, \\Anti-de Sitter space and holography,\" Adv. Theor. Math. Phys. 2, 253\n(1998) [hep-th/9802150].\n[9] J. Zaanen, Y. W. Sun, Y. Liu and K. Schalm, \\Holographic duality in condensed matter\nphysics,\" Cambridge University Press (2015).\n[10] N. Yokoi, M. Ishihara, K. Sato and E. Saitoh, \\Holographic realization of ferromagnets,\"\nPhys. Rev. D 93, 026002 (2016) [arXiv:1508.01626 [hep-th]].\n[11] N. Iqbal, H. Liu, M. Mezei and Q. Si, \\Quantum phase transitions in holographic models\nof magnetism and superconductors,\" Phys. Rev. D 82, 045002 (2010) [arXiv:1003.0010\n[hep-th]].\n[12] R. G. Cai and R. Q. Yang, \\Paramagnetism-Ferromagnetism Phase Transition in a\nDyonic Black Hole,\" Phys. Rev. D 90, no. 8, 081901 (2014) [arXiv:1404.2856 [hep-th]].\n[13] R. G. Cai, R. Q. Yang, Y. B. Wu and C. Y. Zhang, \\Massive 2-form \feld and holographic\nferromagnetic phase transition,\" JHEP 1511 , 021 (2015) [arXiv:1507.00546 [hep-th]].\n[14] V. E. Hubeny and M. Rangamani, \\A Holographic view on physics out of equilibrium,\"\nAdv. High Energy Phys. 2010 , 297916 (2010) [arXiv:1006.3675 [hep-th]].\n[15] C. N. Yang and R. L. Mills, \\Conservation of Isotopic Spin and Isotopic Gauge Invari-\nance,\" Phys. Rev. 96, 191 (1954).\n[16] C. P. Herzog and S. S. Pufu, \\The Second Sound of SU(2),\" JHEP 0904 , 126 (2009)\n[arXiv:0902.0409 [hep-th]].\n[17] V. Balasubramanian, P. Kraus, A. E. Lawrence and S. P. Trivedi, \\Holographic probes\nof anti-de Sitter space-times,\" Phys. Rev. D 59, 104021 (1999) [hep-th/9808017].\n[18] I. R. Klebanov and E. Witten, \\AdS / CFT correspondence and symmetry breaking,\"\nNucl. Phys. B 556, 89 (1999) [hep-th/9905104].\n[19] H. Leutwyler, \\Nonrelativistic e\u000bective Lagrangians,\" Phys. Rev. D 49, 3033 (1994)\n[hep-ph/9311264].\n[20] J. M. Rom\u0013 an and J. Soto, \\E\u000bective \feld theory approach to ferromagnets and antiferro-\nmagnets in crystalline solids,\" Int. J. Mod. Phys. B 13, 755 (1999) [cond-mat/9709298].\n17[21] G. Tatara, \\E\u000bective gauge \feld theory of spintronics,\" Physica E: Low-dimensional\nSystems and Nanostructures 106, 208 (2019) [arXiv:1712.03489 [cond-mat.mes-hall]].\n[22] G. E. Volovik, \\Linear momentum in ferromagnets,\" J. Phys. C 20, L83 (1987).\n[23] For a review, J. A. Harvey, \\Magnetic monopoles, duality and supersymmetry,\" In\n*Trieste 1995, High energy physics and cosmology* 66-125 [hep-th/9603086].\n[24] T. Kikuchi and G. Tatara, \\Spin Dynamics with Inertia in Metallic Ferromagnets,\" Phys.\nRev. B 92, 184410 (2015) [arXiv:1502.04107 [cond-mat.mes-hall]].\n[25] U. Atxitia, D. Hinzke, O. Chubykalo-Fesenko, U. Nowak, H. Kachkachi, O. N. Mryasov,\nR. F. Evans and R. W. Chantrell, \"Multiscale modeling of magnetic materials: Temper-\nature dependence of the exchange sti\u000bness,\" Phys. Rev. B 82, 134430 (2010).\n[26] E. M. Lifshitz and L. P. Pitaevskii, \\Statistical physics: theory of the condensed state\n(Vol. 9)\", Elsevier (2013).\n[27] S. A. Hartnoll, \\Lectures on holographic methods for condensed matter physics,\" Class.\nQuant. Grav. 26, 224002 (2009) [arXiv:0903.3246 [hep-th]].\n[28] C. P. Herzog, K. W. Huang and R. Vaz, \\Linear Resistivity from Non-Abelian Black\nHoles,\" JHEP 1411 , 066 (2014) [arXiv:1405.3714 [hep-th]].\n[29] G. E. Bauer, E. Saitoh and B. J. Van Wees, \\Spin caloritronics,\" Nature materials, 11\n(5), 391 (2012).\n18" }, { "title": "1810.07831v1.Out_of_plane_auto_oscillation_in_spin_Hall_oscillator_with_additional_polarizer.pdf", "content": "arXiv:1810.07831v1 [cond-mat.mes-hall] 17 Oct 20181\nOut-of-plane auto-oscillation in spin Hall oscillator\nwith additional polarizer\nTomohiro Taniguchi\nNational Institute of Advanced Industrial Science and Tech nology (AIST), Spintronics Research Center, Tsukuba,\nIbaraki 305-8568, Japan\nAbstract —The theoretical investigation on magnetization dy-\nnamics excited by the spin Hall effect in metallic multilaye rs\nhaving two ferromagnets is discussed. The relaxation of the\ntransverse spin in one ferromagnet enables us to manipulate\nthe direction of the spin-transfer torque excited in anothe r\nferromagnet, although the spin-polarization originally g enerated\nby the spin Hall effect is geometrically fixed. Solving the\nLandau-Lifshitz-Gilbert-Slonczewski equation, the poss ibility to\nexcite an out-of-plane auto-oscillation of an in-plane mag netized\nferromagnet is presented. An application to magnetic recor ding\nusing microwave-assisted magnetization reversal is also d iscussed.\nIndex Terms —Landau-Lifshitz-Gilbert-Slonczewski (LLGS)\nequation, microwave-assisted magnetization reversal (MA MR),\nspin Hall effect, spin torque oscillator (STO), spintronic s.\nI. I NTRODUCTION\nSPIN-transfer torque [1,2] originated from the spin Hall\neffect [3,4] in nanostructured ferromagnetic/nonmagneti c\nbilayers enables us to manipulate the magnetization withou t\ndirectly injecting the electric current into the ferromagn et.\nMotivated by the demand to develop practical devices, such a s\nmagnetic memory, microwave generators, and neuromorphic\ncomputing, the magnetization switching and oscillation by the\nspin Hall effect have been experimentally demonstrated and /or\ntheoretically investigated [5-15].\nThe variety of the magnetization dynamics excited by the\nspin Hall effect is, however, limited because of the geometr ical\nrestriction of the direction of the spin-transfer torque [1 6], i.e.,\nwhen the electric current flows in the nonmagnet along the\nx-direction and the ferromagnet is set in the z-direction, the\nspin polarization of the spin current generated by the spin H all\neffect is fixed to the y-direction. Due to the restriction, the spin\nHall effect, for example, cannot excite an out-of-plane aut o-\noscillation of a magnetization in a ferromagnet in the absen ce\nof an external magnetic field [16], which is a problematic\nissue in current magnetics. This is a disadvantage of the spi n\nHall geometry to giant magnetoresistive (GMR) system and\nmagnetic tunnel junctions (MTJs), where the torque directi on\ncan be changed by controlling the magnetization direction i n\nthe pinned layer [17-22], and the out-of-plane auto-oscill ation\nof the in-plane magnetized system has already been reported\nboth experimentally and theoretically [23-32].\nThe purpose of this paper is to show that an insertion\nof another ferromagnet having a tilted magnetization to the\nspin Hall geometry enables us to manipulate the spin-transf er\ntorque direction and excite an out-of-plane auto-oscillat ion of\nan in-plane magnetized ferromagnet. The reason for such aN1F1NF2N2\npspin-transfer torque\nJsJexyzm\nFig. 1. Schematic view of the ferromagnetic/nonmagnetic mu ltilayer. The\nelectricJeand spin Jscurrent densities flow along the x- andz-directions,\nrespectively. The spin polarization of Jsflowing from N 1to F1layer points\nto they-direction. Passing through the F 1layer, however, the direction of the\nspin polarization becomes parallel or antiparallel to the m agnetization p.\nphenomenon is because the relaxation of the transverse spin\nin this additional ferromagnet results in the modification o f\nthe direction of the spin polarization. Solving the Landau-\nLifshitz-Gilbert-Slonczewski (LLGS) equation, it is foun d that\nan out-of-plane auto-oscillation of the magnetization can be\nexcited in an in-plane magnetized free layer. An applicatio n\nto magnetic recording using microwave-assisted magnetiza tion\nreversal (MAMR) is also discussed.\nII. O UT-OF-PLANE AUTO -OSCILLATION IN IN -PLANE\nMAGNETIZED FREE LAYER\nThe system we consider is schematically shown in Fig.\n1, where two ferromagnets, F 1and F2, are sandwiched by\ntwo nonmagnets, N 1and N 2. There is another nonmagnet,\nN, between the F 1and F2layers. From hereafter, we use the\nsuffixies F k, Nk, (k= 1,2) and N to distinguish quantities\nrelated to these layers and their interfaces. The unit vecto rs\npointing in the magnetization direction of the F 1and F2\nlayers are denoted as pandm, respectively. The F 1layer\nacts as the polarizer whereas the F 2layer is the free layer, as\ndescribed below. An electric current is applied to the N 1layer\nalongx-direction. The spin-orbit interaction in the N 1layer\nscatters the spin-up and spin-down electrons to the opposit e\ndirections, generating spin current flowing along z-direction\nand polarizing along y-direction. The electric and spin current\ndensities flowing along the x- andz-directions in the N 1layer\nare given by [33]\nJe=σN1Ex+ϑN1σN1\n2e∂zey·µN1, (1)2\nJs=/planckover2pi1ϑN1σN1\n2eExey−/planckover2pi1σN1\n4e2∂zµN1, (2)\nrespectively, where Exis the external electric field in the\nx-direction. The conductivity and the spin Hall angle are\nσ= 1/ρandϑ, respectively, where ρis the resistivity. The\nspin accumulation is denoted as µ, whereas eyis the unit\nvector in the y-direction. We note that the vector notation by\nboldface represents the direction of the spin polarization . The\nspin current given by Eq. (2) creates the spin accumulations\nin each layer, which obey the diffusion equation with the spi n\ndiffusion length ℓ.\nLet us explain the central idea of this paper. The spin\npolarization of the spin current flowing in the z-direction\ngenerated in the N 1layer via the spin Hall effect points to\nthey-direction [3,4,33]. However, in the F 1layer, the spin\ntransverse to the local magnetization pprecesses due to the\nexchange interaction and relaxes rapidly. As a result, only the\nspin polarization parallel or antiparallel to pdirection survives\nduring the transport through the F 1layer. When phas a finite\nz-component pz, the spin current having a finite z-component\nis injected into the F 2layer, and move its magnetization m\nto thez-direction, resulting in an excitation of an out-of-\nplane precession, as schematically shown in Fig. 1. We shoul d\nemphasize here that the magnetization palso has a finite y-\ncomponent pybecause, if py= 0, the spin polarization injected\nfrom the N 1to F1layer relaxes at the N 1/F1interface, and\nthe net spin polarization emitted from the F 1to F2layer\nbecomes zero. Therefore, the magnetization pshould have\nbothy- andz-components. In other words, pshould be tilted\nin theyzplane to maximize the efficiency of the spin injection\nhaving the finite z-component of the spin polarization. This\nis an important difference from GMR or MTJ, where the y-\ncomponent of the polarizer is unnecessary to excite an out-o f-\nplane auto-oscillation [23-32]. We note that the tilted mag netic\nanisotropy has been investigated by making use of a higher-\norder anisotropy or an interlayer exchange coupling betwee n\ntwo ferromagnets [34,35].\nWe show the spin-transfer torque formula acting on the\nmagnetization min the F 2layer. The spin current density\nat the F 2/N interface flowing from the F 2to N layer is given\nby\nJF2→N\ns=1\n4πS/bracketleftbigg(1−γ2)g\n2m·(µF2−µN)m\n−g↑↓\nrm×(µN×m)−g↑↓\niµN×m/bracketrightBig\n,(3)\nwhereg=g↑↑+g↓↓is the sum of the conductances for\nspin-up and spin-down electrons, and γ= (g↑↑−g↓↓)/g\nis its spin polarization. The conductance gis related to the\ninterface resistance rasg/S=h/(e2r), whereSis the cross-\nsectional area. The real and imaginary parts of the mixing\nconductance are denoted as g↑↓\nrandg↑↓\ni, respectively. Since\ng↑↓\nr≫ |g↑↓\ni|for typical ferromagnetic/nonmagnetic interfaces\n[36,37], we neglect the terms related to g↑↓\niin the following\ncalculations. The spin currents at the other F k/Nkand F1/N\ninterfaces are obtained in a similar manner. We assume that\nthe thickness of the N layer is sufficiently thinner than the\nspin diffusion length, and therefore, the spin current in th islayer is conserved. The spin-transfer torque excited in the\nF2layer is, according to the conservation law of the angular\nmomentum, the transverse spin current ejected from this lay er,\ni.e.,τ= [γ0/(MF2dF2)]m×[(JF2→N\ns+JF2→N2s)×m],\nwhereγ0,M, anddare the gyromagnetic ratio, saturation\nmagnetization, and thickness, respectively. Using Eqs. (1 )-(3)\nand the spin diffusion equation, the spin-transfer torque i s\ngiven by\nτ=−γ0/planckover2pi1ϑN1g↑↓\nr(F2/N)g′\nF1/NσN1Ex\n2e˜gF1[g↑↓\nr(F2/N)+g′\nF1/N]MF2dF2pym×(p×m)\n1−λ1λ2(m·p)2.\n(4)\nHere, we define\n1\ng′\nF1/N=2\n(1−γ2\nF1/N)gF1/N+1\ngsd(F1)tanh(dF1/ℓF1)\n−g′\nF1/N1\n[gsd(F1)sinh(dF1/ℓF1)]2,(5)\n1\ng′\nF1/N1=2\n(1−γ2\nF1/N1)gF1/N1+1\ngsd(F1)tanh(dF1/ℓF1)\n+1\ngsd(N1)tanh(dN1/ℓN1),\n(6)\n1\n˜gF1=g′\nF1/N1tanh[dN1/(2ℓN1)]\ngsd(F1)gsd(N1)sinh(dF1/ℓF1), (7)\ngsd(F1)\nS=h(1−β2\nF1)σF1\n2e2ℓF1, (8)\ngsd(N1)\nS=hσN1\n2e2ℓN1, (9)\nλk=g↑↓\nr(Fk/N)−g′\nFk/N\ng↑↓\nr(Fk′/N)+g′\nFk/N, (10)\n[(k,k′) = (1,2),(2,1)], respectively. The spin polarization of\nthe conductivity in ferromagnet is β= (σ↑−σ↓)/(σ↑+σ↓).\nThe spin-transfer torque excited in the F 1layer can be cal-\nculated in a similar manner [38]. The spin-transfer torques in\nsystems having two ferromagnets with different geometries\nare discussed in Refs. [39,40]. In the following, we use\ntypical values of the parameters found in experiments and fir st-\nprinciples calculations in the spin Hall geometry; ℓF= 12 nm,\nℓN= 2.5nmρF= 1600 Ω nm,ρN= 3750 Ω nm,β= 0.56,\nr= 0.28kΩnm2,γ= 0.70,ϑ= 0.1,g↑↓\nr/S= 15 nm−2,\ndF= 1 nm,dN= 4 nm,γ0= 1.764×107rad/(Oe s), and\nM= 1250 emu/c.c. [14,36,37,41,42].\nWe study the magnetization dynamics in the F 2layer under\nthe effect of the spin-transfer torque given by Eq. (4). The\nmagnetization in the F 1layer,p, is assumed to be pinned.\nThe LLGS equation of the magnetization min the F 2layer\nis given by\ndm\ndt=−γ0m×H−γ0/planckover2pi1ηjpym×(p×m)\n2e[1−λ1λ2(m·p)2]MF2dF2+αm×dm\ndt,\n(11)3\nwhere, according to Eq. (4), j=σN1Exand\nη=ϑN1g↑↓\nr(F2/N)g′\nF1/N\n˜gF1[g↑↓\nr(F2/N)+g′\nF1/N]. (12)\nUsing the values of the parameters mentioned above, η=\n0.043andλ= 0.878. The magnetic field,\nH=−4πMF2mzez, (13)\nconsists of the demagnetization (shape anisotropy) field al ong\nthez-direction. The Gilbert damping constant is α, and we use\nα= 0.01in this paper. Figure 2(a) shows typical trajectories\nof the magnetization dynamics in steady states obtained fro m\nthe numerical simulation of Eq. (11), where j=±100×106\nA/cm2andp= (0,1/√\n2,1/√\n2). This figure indicates that the\nout-of-plane precession can be excited by the spin Hall effe ct.\nThis is the main result in this paper. We should also emphasiz e\nthat the oscillation direction, i.e., clockwise or counter clock-\nwise around the z-axis, can be changed by changing the sign\nofj. The magnetization moves to the positive (negative) z-\ndirection by the negative (positive) current, as shown in Fi g.\n2(a).\nWe can obtain the analytical formulas revealing the relatio n\nbetween the current and the oscillation frequency by solvin g\nthe LLGS equation averaged over the constant energy curve\n[43]. The energy density in the present system is E=\n−MF2/integraltext\ndm·H= 2πM2\nF2m2\nz. An auto-oscillation is excited\nwhen the spin-transfer torque compensates with the damping\ntorque, and therefore, the field torque ( −γ0m×H) principally\ndetermines the magnetization dynamics. Since the field torq ue\nconserves the energy density E, the auto-oscillation can be\napproximated as occuring on a constant energy curve. As a\nresult, we can estimate current density necessary to excite an\nauto-oscillation on a constant energy curve of Efrom the\nequation/contintegraltext\ndt(dE/dt) = 0 [16], where the time integral is\nover the precession period. To simplify the discussion we\nuse the cone angle of the magnetization, θ= cos−1mz,\ninstead of E, to identify a constant energy curve in the present\ncase. Assuming that plies in the yz-plane (i.e., px= 0 and\np2\ny+p2\nz= 1), the current density necessary to excite the steady\nprecession of the magnetization with the cone angle θis found\nto be\nj(θ) =−2αeMF2dF2\n/planckover2pi1ηP(θ)4πMF2. (14)\nHere,P(θ)is given by\nP(θ) =[gz(θ)pzsin2θ+gy(θ)cosθ]py\nsin2θcosθ, (15)\nwheregz(θ)andgy(θ)are given by\ngz(θ) =1\n2\n1/radicalBig\nc2\n+−a2+1/radicalBig\nc2\n−−a2\n, (16)\ngy(θ) =1\n2√λ1λ2\nc+/radicalBig\nc2\n+−a2−c−/radicalBig\nc2\n−−a2\n, (17)withc±= 1±√λ1λ2pzcosθanda=√λ1λ2pysinθ. On the\nother hand, the precession frequency at the cone angle θis\nf(θ) =γ0\n2π4πMF2|cosθ|. (18)\nWe confirm the validities of Eqs. (14) and (18) by comparing\nthem with the numerical simulations. We add the random\ntorque,−γ0m×h, due to thermal fluctuation to Eq. (11), for\nthe sake of generality. The components of the random field\nhsatisfy the fluctuation-dissipation theorem, /angbracketlefthi(t)hj(t′)/angbracketright=\n[2αkBT/(γ0MV)]δijδ(t−t′), where the temperature Tand\nthe volume of the F 2layerVare assumed to be T= 300 K and\nV=πr2dF2withr= 50 nm. We solve Eq. (11) numerically\nfor0≤t≤1µs with the initial condition m(0) = + ex.\nRepeating such calculation 104times with random h, the\naveraged Fourier spectra of the x-component of m,|mx(f)|,\nare obtained, as shown in Fig. 2(b). As shown, each spectrum\nhave one peak at a certain frequency, which increases with\nincreasing the current magnitude. The dependence of the pea k\nfrequency on the current obtained from such simulation is\nshown by red circles in Fig. 2(c). The analytical relation\nbetween the current and frequency obtained from Eqs. (14)\nand (18) is also shown in Fig. 2(c) by the black (solid) line. W e\nobtain good agreement between the simulation and analytica l\nformulas, indicating the validity of Eqs. (14) and (18).\nIn this paper, we consider an excitation of an auto-\noscillation in an in-plane magnetized free layer by inserti ng\nan additional ferromagnet between the free layer and the\nnonmagnet having the spin Hall effect. We also notice that\nanother situation is possible [43], where an auto-oscillat ion of\na perpendicularly magnetized free layer is excited by addin g\nanother ferromagnet on the other side of the nonmagnet.\nIII. A PPLICATION TO MICROWAVE ASSISTED\nMAGNETIZATION REVERSAL\nAt the end of this paper, let us discuss the application of\nthe above result to magnetic recording. A spin torque oscill ator\n(STO) consisting of an in-plane magnetized free layer and a\nperpendicularly magnetized pinned layer [23-26] is a candi date\nfor the recording head of a hard disk drive using MAMR [28-\n32,45]. An oscillating magnetic field generated from the STO\nacts as a microwave field on the recording bit, and reduces the\nrecording field [45]. In the original design of MAMR, an STO\nhaving a current-perpendicular-to-plane (CPP) structure was\nassumed [45], where the current flows parallel to the recordi ng\nmedia, as shown in Fig. 3(a). In this case, only a linearly\nporalized field with regard to the recording bit can be obtain ed\nin the microwave field emitted from the STO.\nThis design, however, does not make full use of the idea\nembodied in MAMR. An interesting idea in MAMR is the\nchirality matching between the STO and the recording bit. As\nshown in Fig. 2(a), the magnetization in an STO consisting\nof in-plane magnetized free layer shows oscillation with bo th\nclockwise and counterclockwise chiralities, depending on the\nsign of the current. The magnetization in the recording bit\nalso has an oscillation chirality, depending on the magneti zed\ndirection. It was shown that the recording field in MAMR is\nsignificantly reduced when the chirality of the STO matches4\nmz1\n-10\n1-1\n01\n-10\ncurrent density (106 A/cm2)frequency (GHz)\n-300 -200 -100 0 100 200 300010203040\nfrequency (GHz)0 2 4 6 8 10|mx(f)| (arb. unit)\n00.51.01.5\nj=10, 20, 30 (×106 A/cm2)(a) (b)\nT=300 (K)\nmxmy(c) j=-100 (×106 A/cm2) T=0 (K)\n:theory\n:simulation\nj=100 (×106 A/cm2)\nFig. 2. (a) Steady state trajectories of the magnetization o btained by numerically solving Eq. (11). The current densit ies arej= 100 and−100×106\nA/cm2for red (mz<0) and blue ( mz>0) lines, respectively. The precession directions are indic ated by arrows. The temperature is 0 K. (b) The Fourier\ntransformations of mx(t),|mx(f)|, forj= 10 (red),20(green), and 30(blue)×106A/cm2. The temperature is 300 K. (c) The comparison between the\npeak frequencies of |mx(f)|(red circles) and the analytical solution (black line).\nCPP-STO(a)\nLP-MWSH-STOCPP-STO(b)\nCP-MW(c)\nCP-MW\nrecording media recording media recording mediacurrent\ncurrentcurrent\nFig. 3. (a) Schematic picture of recording method using MAMR . The\nmicrowave (MW) emitted from the CPP-STO acts as a linearly po larized (LP)\noscillating field on the magnetizations in the recording med ia. (b) Another\nstructure of recording method using MAMR in Ref. [46]. The mi crowave\nemitted from the CPP-STO acts as a circularly polarized (CP) oscillating\nfield. (c) A new recording method using MAMR proposed in this w ork. The\nSTO is driven by the spin Hall (SH) effect, as shown in Fig. 1.\nwith that of the recording bit [44]. According to this princi ple\nof the chirality matching, Kudo et al. proposed a concept of\nresonant switching [30,46], where the current in the STO flow s\nalong the direction perpendicular to the recording media; s ee\nFig. 3(b). In this case, the microwave field emitted from the\nSTO is a circularly polarized with regard to the recording bi t.\nTherefore, by changing the current direction in the STO, the\noscillating field having the chirality of both the clockwise and\ncounterclockwise can be generated, and applied in MAMR by\ntaking considerations of chirality matching into account.\nThe results shown in this paper may add another vital\nadvantage in designing the recording head in MAMR. For\nexample, let us consider the system shown in Fig. 3(c),\nwhere the current in the recording head flows parallel to the\nrecording media. In this structure, the current does not flow\nin the free layer, in contrast to the structure shown in Fig.\n3(a). Instead, the spin Hall effect excites an oscillation i n the\nfree layer, as schematically shown in Fig. 1. The circularly\npolarized microwave field is then emitted from the STO, as\nin the case shown in Fig. 3(b). Therefore, the structure in\nFig. 3(c) satisfies the condition to achieve MAMR combined\nwith chirality matching. In addition, the vital advantage, whichresides in this structure, is that the recording head can in\nprinciple be placed closer to the media than in the case of\nFig. 3(b) because of the absence of the electrode between the\nfree layer and the recording media. Therefore, the STO using\nthe spin Hall effect will be an interesting candidate for the\nrecording head of next generation.\nIV. C ONCLUSION\nIn conclusion, it was shown that the spin Hall effect can\nexcite the out-of-plane precession of the magnetization in a\nferromagnet by inserting another ferromagnet having a tilt ed\nmagnetization between the nonmagnetic heavy metal and the\nferromagnet. The phenomenon is due to the relaxation of the\ntransverse spin in this additional ferromagnet. Although t he\nrelaxation of the spin in the additional layer induces a loss\nof spin polarization, it enables us to manipulate the direct ion\nof the spin-transfer torque excited on the free layer. Using\nthe spin-transfer torque formula derived from the diffusiv e\nspin-transport theory and solving the LLGS equation both\nnumerically and analytically, the relation between the cur rent\nand the precession frequency was obtained. It was also shown\nthat the chirality of the precession can be reversed by rever sing\nthe current direction. An application to magnetic recordin g\nusing microwave assisted magnetization reversal was also\ndiscussed.\nNote added : After our submission, we were notified that Dr.\nSuto in Toshiba proposed another solution based on a nonloca l\ninjection of spin current to generate a circularly polarize d\nmicrowave from a spin-valve with a current flowing parallel\nto the recording media (unpublished).\nACKNOWLEDGMENT\nThe author is grateful to Takehiko Yorozu, Yoichi Shiota,\nHitoshi Kubota, and Shingo Tamaru for valuable discussions .\nThe author is also thankful to Satoshi Iba, Aurelie Spiesser ,\nHiroki Maehara, and Ai Emura for their support and encour-\nagement. This work was supported by JSPS KAKENHI Grant-\nin-Aid for Young Scientists (B) 16K17486.5\nREFERENCES\n[1] J. C. Slonczewski, ”Current-driven excitation of magne tic multilayers”,\nJ. Magn. Magn. Mater. vol. 159, p.L1, 1996.\n[2] L. Berger, ”Emission of spin waves by a magnetic multilay er traversed\nby a current”, Phys. Rev. B vol. 54, 9353, 1996.\n[3] M. I. Dyakonov and V . I. Perel, ”Current-induced spin ori entation of\nelectrons in semiconductors”, Phys. Lett. A vol. 35, 459, 1971.\n[4] J. E. Hirsch, ”Spin Hall Effect”, Phys. Rev. Lett. vol. 83, 1834, 1999.\n[5] L. Liu, O. J. Lee, T. J. Gudmundsen, D. C. Ralph, and R. A. Bu hrman,\n”Current-Induced Switching of Perpendicularly Magnetize d Magnetic\nLayers Using Spin Torque from the Spin Hall Effect”, Phys. Rev. Lett.\nvol. 109, 096602, 2012.\n[6] C.-F. Pai, L. Liu, Y . Li, H. W. Tseng, D. C. Ralph, and R. A. B uhrman,\n”Spin transfer torque devices utilizing the giant spin Hall effect of\ntungsten”, Appl. Phys. Lett. vol. 101, 122404, 2012.\n[7] K.-W. Kim, S.-M. Seo, J. Ryu, K.-J. Lee, and H.-W. Lee, ”Ma gnetization\ndynamics induced by in-plane currents in ultrathin magneti c nanostruc-\ntures with Rashba spin-orbit coupling”, Phys. Rev. B vol. 85, 180404,\n2012.\n[8] K. Garello, I. M. Miron, C. O. Avci, F. Freimuth, Y . Mokrou sov,\nS. Bl¨ ugel, S. Auffret, O. Boulle, G. Gaudin, and P. Gambarde lla,\n”Symmetry and magnitude of spin-orbit torques in ferromagn etic het-\nerostructures”, Nat. Nanotechnol. vol. 8, 587, 2013.\n[9] J. Kim, J. Sinha, M. Hayashi, M. Yamanouchi, S. Fukami, T. Suzuki,\nS. Mitani, and H. Ohno, ”Layer thickness dependence of the cu rrent-\ninduced effective field vector in Ta/CoFeB/MgO”, Nat. Mater. vol. 12,\n240, 2013.\n[10] X. Fan, J. Wu, Y . Chen, M. J. Jerry, H. Zhang, and J. Q. Xiao , Nat.\nCommun. 4, 1799 (2013). ”Observation of the nonlocal spin-o rbital\neffective field”, Nat. Commun. vol. 4, 1799, 2013.\n[11] X. Qiu, P. Deorani, K. Narayanapillai, K.-S. Lee, K.-J. Lee, H.-W. Lee,\nand H. Yang, Sci. Rep. 4, 4491 (2014). ”Angular and temperatu re depen-\ndence of current induced spin-orbit effective fields in Ta/C oFeB/MgO\nnanowires”, Sci. Rep. vol. 4, 4491, 2014.\n[12] G. Yu, P. Upadhyaya, Y . Fan, J. Alzate, W. Jiang, K. L. Won g, S. Takei,\nS. A. Bender, L.-T. Chang, Y . Jiang, M. Lang, J. Tang, Y . Wang, Y .\nTserkovnyak, P. K. Amiri, and K. L. Wang, ”Switching of perpe ndicular\nmagnetization by spin-orbit torques in the absence of exter nal magnetic\nfields”, Nat. Nanotechnol. vol. 9, 548, 2014.\n[13] M. Cubukcu, O. Boulle, M. Drouard, K. Garello, C. O. Avci , I. M.\nMiron, J. Langer, B. Ocker, P. Gambardella, and G. Gaudin, Ap pl. Phys.\nLett. 104, 042406 (2014). ”Spin-orbit torque magnetizatio n switching of\na three-terminal perpendicular magnetic tunnel junction” ,Appl. Phys.\nLett. vol. 104, 042406, 2014.\n[14] J. Torrejon, F. Garcia-Sanchez, T. Taniguchi, J. Shinh a, S. Mitani, J.-\nV . Kim, and M. Hayashi, ”Current-driven asymmetric magneti zation\nswitching in perpendicularly magnetized CoFeB/MgO hetero structures”,\nPhys. Rev. B vol. 91, 214434, 2015.\n[15] K. Kudo and T. Morie, ”Self-feedback electrically coup led spin-Hall\noscillator array for pattern-matching operation”, Appl. Phys. Express vol.\n10, 043001, 2017.\n[16] T. Taniguchi, ”Nonlinear analysis of magnetization dy namics excited by\nspin Hall effect”, Phys. Rev. B vol. 91, 104406, 2015.\n[17] J. A. Katine, F. J. Albert, R. A. Buhrman, E. B. Myers, and D. C. Ralph,\n”Current-Driven Magnetization Reversal and Spin-Wave Exc itations in\nCo/Cu/Co Pillars”, Phys. Rev. Lett. vol. 84, 3149, 2000.\n[18] J. Grollier, V . Cros, A. H. J. M. George, H. Jaffr¨ es, and A. Fert, Appl.\nPhys. Lett. 78, 3663 (2001). ”Spin-polarized current induc ed switching\nin Co/Cu/Co pillars”, Appl. Phys. Lett. vol. 78, 3663, 2001.\n[19] S. I. Kiselev, J. C. Sankey, I. N. Krivorotov, N. C. Emley , R. J.\nSchoelkopf, R. A. Buhrman, and D. C. Ralph, ”Microwave oscil lations\nof a nanomagnet driven by a spin-polarized current”, Nature vol. 425,\n380, 2003.\n[20] W. H. Rippard, M. R. Pufall, S. Kaka, T. J. Silva, and S. E. Russek,\n”Current-driven microwave dynamics in magnetic point cont acts as a\nfunction of applied field angle”, Phys. Rev. B vol. 70, 100406, 2004.\n[21] H. Kubota, A. Fukushima, Y . Ootani, S. Yuasa, K. Ando, H. Maehara,\nK. Tsunekawa, D. D. Djayaprawira, N. Watanabe, and Y . Suzuki , ”Eval-\nuation of Spin-Transfer Switching in CoFeB/MgO/CoFeB Magn etic\nTunnel Junctions”, Jpn. J. Appl Phys. vol. 44, L1237, 2005.\n[22] I. N. Krivorotov, N. C. Emley, J. C. Sankey, S. I. Kiselev , D. C. Ralph,\nand R. A. Buhrman, ”Time-Domain Measurements of Nanomagnet\nDynamics Driven by Spin-Transfer Torques”, Science vol. 307, 228,\n2005.[23] A. D. Kent, B. ¨Ozyilmaz, and E. del Barco, ”Spin-transfer-induced\nprecessional magnetization reversal”, Appl. Phys. Lett. vol. 84, 3894,\n2004.\n[24] K. J. Lee, O. Redon, and B. Dieny, ”Analytical investiga tion of spin-\ntransfer dynamics using a perpendicular-to-plane polariz er”,Appl. Phys.\nLett. vol. 86, 022505, 2005.\n[25] D. Houssameddine, U. Ebels, B. Dela¨ et, B. Rodmacq, I. F irastrau, F.\nPonthenier, M. Brunet, C. Thirion, J.-P. Michel, L. Prejben u-Buda, M.-\nC. Cyrille, O. Redon and B. Dieny, ”Spin-torque oscillator u sing a\nperpendicular polarizer and a planar free layer”, Nat. Matter. vol. 6,\n447, 2007.\n[26] U. Ebels, D. Houssameddine, I. Firastrau, D. Gusakova, C. Thirion,\nB. Dieny, and L. D. Buda-Prejbeanu, ”Macrospin description of the\nperpendicular polarizer-planar free-layer spin-torque o scillator”, Phys.\nRev. B vol. 78, 024436, 2008.\n[27] Y . Zhou, C. L. Zha, S. Bonetti, J. Persson, and J. Akerman , ”Spin-torque\noscillator with tilted fixed layer magnetization”, Appl. Phys. Lett. vol.\n92, 262508, 2008.\n[28] H. Suto, T. Yang, T. Nagasawa, K. Kudo, K. Mizushima, and R. Sato,\n”Magnetization dynamics of a MgO-based spin-torque oscill ator with\na perpendicular polarizer layer and a planar free layer”, J. Appl. Phys.\nvol. 112, 083907, 2012.\n[29] H. Suto, T. Nagasawa, K. Kudo, K. Mizushima, and R. Sato, ”Nanoscale\nlayer-selective readout of magnetization direction from a magnetic mul-\ntilayer using a spin-torque oscillator”, Nanotechnology vol. 25, 245501,\n2014.\n[30] K. Kudo, H. Suto, T. Nagasawa, K. Mizushima, and R. Sato, ”Frequency\nstabilization of spin-torque-driven oscillations by coup ling with a mag-\nnetic nonlinear resonator”, J. Appl. Phys. vol. 116, 163911, 2014.\n[31] T. Taniguchi and H. Kubota, ”Instability analysis of sp in-torque os-\ncillator with an in-plane magnetized free layer and a perpen dicularly\nmagnetized pinned layer”, Phys. Rev. B vol. 93, 174401, 2016.\n[32] T. Taniguchi, ”Crossover between fast and slow excitat ion of magneti-\nzation by spin torque”, Appl. Phys. Express vol. 9, 073003, 2016.\n[33] S. Takahashi and S. Maekawa, ”Spin Current in Metals and Supercon-\nductors”, J. Phys. Soc. Jpn. vol. 77, 031009, 2008.\n[34] J.-W. Lee, J.-R. Jeong, S.-C. Shin, J. Kim, and S.-K. Kim , ”Spin-\nreorientation transitions in ultrathin Co films on Pt(111) a nd Pd(111)\nsingle-crystal substrates”, Phys. Rev. B vol. 66, 172409, 2002.\n[35] L. Fallarino, V . Sluka, B. Kardasz, M. Pinarbasi, A. Ber ger, and A. D.\nKent, ”Interlayer exchange coupling between layers with pe rpendicular\nand easy-plane magnetic anisotropies”, Appl. Phys. Lett. vol. 109,\n082401, 2016.\n[36] A. Brataas, Y . V . Nazarov, and G. E. W. Bauer, ”Spin-tran sport in\nmulti-terminal normal metal-ferromagnet systems with non -collinear\nmagnetizations”, Eur. Phys. J. B vol. 22, 99, 2001.\n[37] M. Zwierzycki, Y . Tserkovnyak, P. J. Kelly, A. Brataas, and G. E. W.\nBauer, ”First-principles study of magnetization relaxati on enhancement\nand spin transfer in thin magnetic films”, Phys. Rev. B vol. 71, 064420,\n2005.\n[38] T. Taniguchi, S. Mitani, and M. Hayashi, ”Current-Indu ced Instability\nof a Perpendicular Ferromagnet in Spin Hall Geometry”, IEEE Trans.\nMagn. vol. 52, 1400204, 2016.\n[39] W. T. Chen, S. Takahashi, H. Nakayama, M. Althammer, S. T . B.\nGoennewein, E. Saitoh, and G. E. W. Bauer, ”Theory of spin Hal l\nmagnetoresistance”, Phys. Rev. B vol. 87, 144411, 2013.\n[40] T. Taniguchi, J. Grollier, and M. D. Stiles, ”Spin-Tran sfer Torques\nGenerated by the Anomalous Hall Effect and Anisotropic Magn etore-\nsistance”, Phys. Rev. Applied vol. 3, 044001, 2015.\n[41] J. Bass and J. W. P. Pratt, ”Spin-diffusion lengths in me tals and alloys,\nand spin-flipping at metal/metal interfaces: an experiment alistfs critical\nreview”, J. Phys.: Condens. Matter. vol. 19, 183201, 2007.\n[42] T. Taniguchi, S. Yakata, H. Imamura, and Y . Ando, ”Deter mination of\nPenetration Depth of Transverse Spin Current in Ferromagne tic Metals\nby Spin Pumping”, Appl. Phys. Express vol. 1, 031302, 2008.\n[43] G. Bertotti, I. Mayergoyz, and C. Serpico, ”Nonlinear m agnetization\nDynamics in Nanosystems” (Elsevier, Oxford, 2009).\n[44] T. Taniguchi, ”Indirect excitation of self-oscillati on in perpendicular\nferromagnet by spin Hall effect”, Appl. Phys. Lett. vol. 111, 022410,\n2017.\n[45] J.-G. Zhu, X. Zhu, and Y . Tang, ”Microwave Assisted Magn etic Record-\ning”, IEEE Trans. Magn. vol. 44, 125, 2008.\n[46] K. Kudo, H. Suto, T. Nagasawa, K. Mizushima, and R. Sato, ”Resonant\nmagnetization switching induced by spin-torque-driven os cillations and\nits use in three-dimensional magnetic storage application sAppl. Phys.\nExpress vol. 8, 103001, 2015." }, { "title": "2312.16925v1.Detecting_bulk_carbon_ferromagnetism_in_graphene_multi_edge_structure.pdf", "content": " \nDetecting bulk carbon fer romagnetism in graphene multi -edge structure \nChao Wang1,3*, Nan Jian2, Meijie Yin2, Xi Zhang2, Zhi Yang1, Xiuhao Mo1, Takashi Kikkawa4, Shunsuke \nDaimon4,5, Eiji Saitoh4,6,7, Qian Li8, Wensheng Yan8, Dazhi Ho u9, Lei Yang10, Dongfeng Diao1,2,3* \n \n1. Institute of Nanosurface Science and Engineering, Shenzhen University, Shenzhen 518060, China ; \n2. Electron Microscopy Center of Shenzhen University, Shenzhen 5180 60, China ; \n3. National Key Laboratory of Radio Frequency Heterogeneous Integration , Shenzhen University, Shenzhen \n518060, China ; \n4. Department of Applied Physics, The University of Tokyo, Tokyo 113 -8656, Japan ; \n5. Quantum Materials and Applications Research Center, National Institutes for Quantum Science and \nTechnology (QST), Tokyo 152 -8550, Japan ; \n6. Institute for AI and Beyond, The University of Tokyo, Tokyo 113 -8656, Japan ; \n7. WPI Advanced Institute for Materials Research, Tohoku University, Sendai 980 -8577, Japan ; \n8. National Synchrotron Radiation Laboratory, and School of Nuclear Science and Technology, University of \nScience and Technology of China, Hefei, Anhui 230029, China ; \n9. International Centre for Quantum Design of Functional Materials (ICQD), Hefei National Laboratory for \nPhysical Sciences at the Microscale (HFNL), and Synergetic Innovation Center of Quantum Information and \nQuantum Physics, University of Science and Technol ogy of China, 230026 Hefei, China ; \n10. Key Laboratory of Education Ministry for Modern Design and Rotor -Bearing System, Xi’an Jiaotong \nUniversity, Xi'an 710049, China . \nE-mail address: dfdiao @szu.edu.cn . \n \nAbstract: \nThe emergence of bulk carbon ferromagnetism is long -expected over years. At nano scale, carbon \nferromagnetism was detected in graphene nanoribbon segments1–3, graphene triangulene dimers4,5, and other \ngraphene -like molecules6–8 by analyzing their magnetic edge states via scanning tunneling microscop y(STM )9, \nand its origin can be explained by local redistribution of electron wave function3,5,8. Recently, there are \nexperimental efforts on discovering orbital ferromagnetism in twisted bilayer graphene10,11, which is still \nlimited in STM region . In larger scale , carbon ferromagnetism can be created by deliberately producing defects \nin graphite12–14, and detect ed by macroscopic technical magnetization. Meanwhile, it becomes crucial to \ndetermine that the detected magnetization is originated from carbon rather than from magnetic impurities. One \nsolution to eliminate this uncertainty is X -ray magnetic circular dichroism (XMCD ). A contrast of dichroism \nat * absorption energy was observed in the scanning x -ray microscope image of proto n-irradiated graphite15. \nNonetheless , a reproducible , full section of XMCD spectr um across C-1s absorption energy has not appeared \nyet, which should be decisive for assuring the indisputable existence of bulk carbon ferromagnetism. Besides , \nthe lack of direct observation on the atomic structure of the ferromagnetic carbon leaves the structural origin \nof its ferromagnetism still in mist. In this work, for detecting bulk carbon ferromagnetism , we manage d to \ngrow all -carbon film consisting of vertically aligned graphene multi -edge (VGME ), which wove into a three -\ndimensional hyperfine -porous network. Magnetization (M-H) curves and XMCD s pectra co -confirmed bulk \ncarbon ferromagnetism of VGME at room temperature , with the average unit magnetic momentum of ~0.0006 \nB/atom. The influence of magnetic impurities on magnetization was excluded by both absorption spectra and \ninductively coupled plasma mass spectrometry measurements. The spin transfer behavior also verified the \nlong-range and robust feature of the bulk carbon ferromagnetism . Our work provides direct evidence of \nelementary resolved bulk carbon ferromagnetism at room temperature and clarifies its origin from -electrons \nat graphene edge s. \n \nText: \nThe emergence of bulk carbon ferromagnetism is long -expected over years. In nanoscale, there are \ntheoretical underpinning s for carbon ferromagneti sm such as the zigzag arrangement at graphene edges1–3, \nvacancies16, and adatoms17,18, where the redistribution of electron wave function occurs and local \nferromagnetism is generated . In experimental, local ferromagnetism was detected in graphene nanoribbon \nsegments1–3, graphene triangulene dimers4,5, and other graphene -like molecules6,7. In addition, l ocal \nmagnetization in graphene was modulated by adatoms and doping19,20. Recently, there are experimental efforts \non discovering orbital ferromagnetism in twisted bilayer graphene11,21. These cases of low-dimensional carbon \nferromagnetism are at cluster scale with several hexagons in -plane, and a few layers out -of-plane. The \nferromagnetic order s were detect ed under scanning tunneling microscop y by measuring local magnetization \nor electric polarization via analyzing their edge states. \nCarbon ferromagnetism in larger scale can be created b y deliberately produc ing defects in graphite through \nirradiation12–14, oxidation22, or mechanical exploiting23. This type of ferromagnetism is detectable in \nmacroscopic technical magnetization. The hysteresis loop in magnetization curve is ascribed to magnetic \nmoments of edge states at the defected positions. Meanwhile, it becomes crucial to determine that the detected \nmagnetization is originated from carbon rather than from magnetic impurities, because carbon ferromagnetic \nsignals are several orders of magnitude lower than magnetic element , and M-H curve has no elemental \nresolution. Just out of this fact, the experimental observations of macroscopic ferromagnetis m in carbon \nmaterials were always challenged on the issue of impurities, or at least not well recognized , since it is difficult \nto rule out the influence of impurities . There are even reports against carbon ferromagnetism after careful \nexperiments24,25. \nOne solution to eliminate th is uncertainty is X-ray magnetic circular dichroism (XMCD ), which can \nrecognize carbon ferromagnetism from the change of its absorption spectra under opposite magnetic field . A \ncontrast of dichroism at 284 eV was observed in scanning X-ray microscope image of proto n-irradiated \ngraphite15. Nonetheless, a repr oducible, full section of XMCD spectrum across C-1s absorption energy has \nnot appeared yet , which should be decisive for assuring the indisputable existence of bulk carbon \nferromagnetism. \nFor ferromagnetic carbon in any form, the atomic structure is the source of its unique electronic structure \nthat produces ferro magneti c order . In the studies of low -dimensional ferromagnetic carbon allotropes, the \nvisualization of their atomic structures significantly improved the interpretation on the origin of local carbon \nferromagnetism. However , in the case of bulk ferromagnetic carbon , the lack of direct observation on the \natomic structure leaves the structural origin of its ferromagnetism still in mis t. \nIn this work, we manage d to grow all -carbon film consist ing of vertically aligned graphene edges, which \nwove into a three -dimensional hyperfine -porous network . The vertical graphene multi -edge (VGME) \nstructur al feature was identified by high resolution transmission electron microscop y (HRTEM) . M-H curves \nand XMCD spectra co -confirmed the room temperature ferromagnetism of VGME. Carefully repeated \nmeasurements were carried out to reduce the noise, and the average unit magnetic momentum of ~0.0006 \nB/atom was determined . The influence of magnetic impurities on magnetization was excluded by absorption \nspectra and inductively coupled plasma mass spectrometry measurements. Angular dependent \nmagnetor esistance of VGME was measured in room temperature , and the s pin transfer behavior verified the \nlong-range and robust feature of ferromagnetism . Our work provides direct evi dence of elementary resolved \nbulk carbon ferromagnetism at room temperature and clarifies its origin from -electrons at graphene edges. \nThe VGME structure was produced in electron cyclotron resonance (ECR) plasma via a physical vapor \ndeposition (PVD) approach with a low -energy -electron -irradiation technique, which was explained in detail \nin our previous work26,27. This approach allows us to produce VGME structure at wafer scale on variable \nsubstrate surface s include but not limited to SiO 2. Fig. 1a-c depict a sequence of plan -view high -angle annular \ndark-field (HAADF) images of the as deposited film consist of VGME structure (VGME film) , progressively \nzooming in. Fig. 1a reveal s the porous architecture of the VGME film as black void regions are distributed \nthroughout the film. The pores are of irregular shapes, and the ir average size generally does not exceed 2 nm \naccording to statistical measurements (see F ig. S1). Bright regions of the images represent the solid part of \nVGME structure, which consists of turbostratic graphene. The binding configuration of the turbostratic \ngraphene was confirmed by Raman and XPS spectra (see F ig. S2). The curving and crosslinking nature of the \ngraphene layers can be clearly observed by the line -shaped projections in Fig.1b. The dangled edges around \nsuperfine pores are highlighted in purple dashed lines. Theses vertically dangled grap hene edges are believed \ncrucial for the emergence of ferromagnetic moment. Fig.1c magnifie s the graphene layers around one of the \npores. The interlayer spacing of 0.35 nm is revealed as labeled by black arrows in the image, corresponding \nto the van der Waals distance between graphene layers. Inside each graphene layer exists the spacing of 0.21 \nnm which corresponding to the distance between neighboring zigzag edges, as labeled by yellow arrows and \nillustrated by the inset sketch. According to the viewing direction and the appearance of the zigzag edge \nprojection, it is verified that the zigzag di rection of the graphene layer is in parallel with viewing direction , \ni.e., vertical to the substrate surface. The side view TEM image of the VGME film in Fig. 1d confirms that \nmost graphene layers orient vertically to the film/substrate interface. Fig. 1e further highlights the features of \ncrosslinking, splitting, and merging between neighboring graphene layers by colored dash lines. Such features \nsuggesting the observed pores are not perpendicular linear tunnel through the thickness of the film , instead, \nthey are discontinuously distributed inside the whole film . The blue square part of the image is magnified in \nFig. 1f, in which a line -shaped projection of the graphene layer is observed with the inner spacing of 0.12 nm, \ncorresponding to the spacing between two nearest atom ic planes which are along armchair direction, as labeled \nby yellow lines and illustrated by the inset sketch. The spacing of 0.12 nm in side view co -demonstrate that \nthe vertically aligned edges are zigzag type, as Fig. 1c does. This stru cture thereby is named as vertical \ngraphene multi -edge (VGME) structure, which consist of multiple graphene layers aligned parallel to the film \ngrown direction. These graphene layers further wove into a three -dimensional hyperfine -porous network. \n \nFig. 1 Atomic image of vertical graphene multi -edge (VGME) structure. a, Plan view HADDF image in which the \nsuperfine pores in the structure are observed, as highlighted in the image by dashed circles. The bright lines in the figure \nrepresents the projections of curved graphene layers. A square area b is selected for zoom -in observation. b, zoom -in image \nof the selected area in a . The curve, dangle, and crosslink features of the graphene layers can be observed clearly, as guided \nby dashed lines. A square area c is selected for further zoom -in observation. c, zoom -in image of the selected area in b. The \natomic structural detail of the graphene layer can be identified , as indicated by dashed lines . The spacing of 0.35 nm between \ngraphene layers and 0.21 nm between zigzag edges are labeled. d-f, Side view TEM images of the VGME film, progressively \nzooming in . The vertical orientation of the graphene layers can be seen in d; the crosslink, merge and split features can be \nseen in e; the spacing of 0.35 nm between graphene layers and 0.12 nm between nearest atom lin es along armchair direction \nare labeled. \nThe magnetization (M-H) curve of the above VGME structure is shown in Fig. 2a in blue color . The \ncurve was measured at the temperature of 300 K, and the saturation magnetic field H of 0.7 T. The direction \n \nof magnetization is vertical to the VGME film, that is parallel to the graphene edges. A hysteresis loop with \nsaturation feature is clearly shown in the M -H curve. The magnetization Ms at 0.7 T, the residue magnetization \nat 0 T M R, and the coercive magnetic field H c (at 0 magnetization) are listed inside the figure. The M s value \nof VGME film at 0.7 T is 4.7 emu/g, converting to unit magnetiz ation of ~0.000 8 B per atom. This is three \norder of magnitude higher than reported in de fected HOPG23. Furthermore, the M -H curve shows rapid \nincrement at low field, and reaches 4.3 emu/g at ±0.2 T. When magnetic field further increases, the \nmagnetization slowly saturates . For the need of compare study, we prepared an amorphous carbon film (a -C) \nwith the same technique, during which the irradiating energy of electrons were at a rather low level so that \nVGME structure was not formed. Compar ing measurement results of HOPG and a-C samples are also given \nin Fig . 2a in grey and yellow curves . These curves are magnified for better resolution in Fig. 2b and c. The \nMS value of HOPG sample is 1.1 ×10-3 emu/g and the H c value is 25 Oe, close to the earlier reported value 23, \nimplying that the results of our hysteresis measurement are reasonable. The M s value of a -C film is 1.4×10-4 \nemu/g, four orders of magnitude lower than that of VGME structure. Since the VGME and a -C samples are \nprepared by the same technique, in can be inferred that potential contribution to M s from impurities could not \nexceed 1.4×10-4 emu/g. Therefore, t he detected hysteresis loop and large saturation magnetization of VGME \nstructure could be ascribed to bulk carbon ferromagnetism. To verify this interpretation, XMCD measurements \nwere introduced for elementa l-resolved magnetic detecti on. \n \nFig. 2 Magnetization curves of VGME , HOPG and a -C structures at 300 K. a, the hysteresis loop of VGME with magnetic \n \nfield H perpendicular to film surface , as indicated inside the figure . The saturation magnetization Ms, residue Magnetization \nMR and coercive field H c derived from the loop are listed. The Ms reaches 4.7 emu/g, equal s to 0.0008 B/atom. M-H curves \nof and a -C are presented in the same scale in the figure. b, magnified hysteresis loop of HOPG with H perpendicular to \ngraphene layer, as indicated inside the figure. c, magnified hysteresis loop of a -C with H perpendicular to the film surface. \nFig.3a summarizes the X -ray absorption near edge spectra (XANES) collected from VGME and a-C film \nas well as HOPG in vertical and horizontal positioning directions . The near -edge detailed spectra are \nhighlighted in yellow background, and the X -ray inject direction are presented above each spectrum. All \nsample s show clear C -1s K -edge absorption signals . The XANES of the VGME sample shows a strong peak \nat * position of ~284.5 eV , which is more similar to that of vertical HOPG than to horizontal HOPG, \nsuggesting the strong X -ray absorption from 2p orbital. This is due to the reason that the incident direction of \nthe X-ray was perpendicular to the film surface, and the graphene layers in VGME are vertically grown as \nwere observed in TEM images. It should be po inted out that the * peak fro m VGME sample is extremely \nstronger than from the other three. Such feature can be ascribed to the vertical graphene multi -edge structure , \nwhich produces abundant edge states those are more active under X-ray excitation . In order t o check whether \nour results were affected by magnetic impurities, t he absorption spectra at the energy ranges of L -edge of Fe, \nCo, and Ni were acquired , and no signals were detected, as can be seen in Fig 2b -d. This indicates that the \nobserved magnetization signal of the sample does not originate from magnetic impurities other than carbon. \nIn addition, the trace of metallic impurities in VGME film were analyzed by inductively coupled plasma mass \nspectrometry (ICP -MS), and the results showed that the total content of metallic impurities (mainly Al, Cr and \nFe) are less than 30 ppm (see Fig. S 3). Such content level coincides with the above interpretation. Therefore, \nthe influence of metallic impurities on the electronic structure of total VGME sample, for example, via \nalloying or proximity effect , can be neglected. \n \nFig.3 X-ray absorption spectroscopy of the VGME, HOPG and a -C structures. a, C-1s K-edge absorption spectra of \nVGME, a -C and HOPG (vertical and horizontal ) samples. The near -edge detailed spectra are highlighted in yellow \nbackground, and the X -ray incident direction are presented beside each spectrum. b-d, absorption spectra of VGME sample \nat the energy ranges of L -edge of Fe, Co, and Ni , respectively . No signal is observed, suggesting the influence of magnetic \nimpurities can be neglected. \nFig. 4 compa res the XMCD spectra of VGME and a -C structure, which are detected at 300 K with the \nexternal , out-of-plain magnetic field of 0.7 T. In order to guarantee the reproducibility and increase signal -to-\nnoise ratio, the spectra were repeatedly taken for 10 times, and the average signals from increasing scanning \ntimes are exhibited in Fig 4a -f. As sh own in the figure s, the dichroism of * peak at ~284.5 eV becomes more \nprofound after multiple repeating m easurements on VGME sample , and the noise level is effectively reduced, \nso that the dichroism peak can be ide ntified at rather low amplitude. The profound feature of the dichroism at \n* excitation en ergy of VGME structure provides a direct evidence of bulk carbon ferromagnetism originated \nfrom 2p orbital. Since the dichroism appears at the same energy as in XA NES spectra, its origin can be ascribed \nto the edge states in the VGME structure, as mentioned above. The amplitude of VGME -XMCD peak is \n~0.005 and stabled after repeatedly mea surements, corresponding to the effective magnetic moment of \n~0.000 6 B per atom by relative area ratio metho d (see description in method section) . This value is close to \n \nits M s that calculated from the hysteresis loop. On the other hand, after 10 times of repeating measurements, \nthe XMCD spectrum of a -C film only shows fluctuations at the amplitude of ~0.002 , such amplitude is \ncomparable to background noise level of the measurement s, as indicated in each panel with parallel dashed \nlines . We note that for the L -edge absorption spectra of transition metals, spin and orbital angular momentum \ncan be separate d by introducing sum rule28. Yet so far, this methodology is not suitable for the K -edge \nabsorption spectra as in the case of carbon. Recent breakthrough in the analysis of K-edge spectra shed light \non extracting the orbital contribution to the s -p dichroism signal29. Therefore, we anticipate the XMCD of C -\n1s K -edge to unveil more information about the origin of bulk carbon ferromagnetism and benefits deeper \nunderstanding of carbon electronic structures. \n \nFig. 4 XMCD spectra of VGME and a -C structure s detected at 300 K with the external out -of-plain magnetic field of \n0.7 T. a -g, the average results of XMCD spectra of VGME structure with increasing scan times from 1 to 10 scans. The \n \ndirhosim peak s at ~28 4.5 eV are indicated by solid curve s in each spectrum. h, the average result XMCD spectra of a -C \nstructure after 10 scans. The noise range s between ±0.002 a.u. are indicated by parallel dashed lines in each panel . \nFor a bulk material with long -range ferromagnetic order, one should expect multiple magnetic dynamic \nbehaviors other than a classic hysteresis loop. For such purpose, we investigated the spin trans port behavior \nin the VGME film. Fig. 5a-b illustrate the principle of the measurement setup. A platinum layer Pt is coated \non top of the VGME or a -C film to build a double layer junction, being referred as Pt/VGME or Pt/a -C. When \nan electric current J c is injected through Pt layer along X-axis, a spin current J s is naturally formed along Z-\naxis due to spin -Hall effect (SHE) , in which the spin polarized electrons diffuse to the interface of the junction. \nSimultaneously, a reversed bounced back spin current J’s1 at the interface is generated, and transfer in to J c due \nto inverse spin -Hall effect (ISHE) , keeping the overall J c barely changed. Under this condition, an external \nmagnetic field B is further applied along different directions, by which the magneti c moment in VGME is \ndetermined if B is large enough. When B is along Y -axis as shown in Fig . 5a, the magnetic moment in VGME \nM1 is with the same direction as in the spin polarized electron, therefore, no spin moment is absorbed and J c \nis unchanged. When B is along X -axis as Fig . 5b illustrates, the magnetic moment in VGME M2 is \nperpendicular to that of the spin polarized electron, the spin moment will be absorbed by VGME and the \nbounced back spin current decreases to J’ s2. Therefore, the overall charge current after ISHE is also lowered \nto J’ c. Such angular dependence of can be detected by measuring the change of longitudinal resistance xx in \nPt/VGME or Pt/ a-C with respect to the rotation angle of B. Fig. 5c-e summarize the angular dependences of \nxx of Pt/VGME and Pt/ a-C on the rotation of magnetic field B in X-Y ( angle), Y-Z( angle) and X-Z \n( angle) plane s, respectively. A periodical response of xx with the angular change is clearly observed in \nPt/VGME , whereas in Pt/a-C the resistance barely ch anges (xx maintains near zero). It can be inferred that \nthe angular dependent magnetoresistance is relate d to VGME film. The xx along angle shows a cosine \nfeature, where the resistance is lower when B||Y-axis, and higher when B ⊥Y-axis, in accordance to the \nbehavior of spin Hall magnetoresistance. We also notice that when B rotates in and angle, same angular \ndependence appears with larger amplitude. This could be from the negative magnetoresistance of the VGME \nfilm, which is not our focus here . By subtracting xx along angle from xx along angle, we obtained a \nchange of xx exactly equals to xx along angle. The angular dependent magnetoresistance behavior \nalong -angle imply s a classic spin Hall magnetoresistance ( SMR ) process30,31. The observation of such SMR \nbehavior at 300 K further proves the robust of the magnetization in VGME structure. \n \nFig. 5 Spin Hall magnetore sistance in VGME structure. a -b, The principle of current J c variation in response to external \nmagnetic field B in parallel and perpendicular to Y axis . Js represents the spin current generated in Pt layer via spin Hall \neffect. J’ s1 and J’ s2 represent the bounced spin current when B||Y and B ⊥Y, respectively. M 1 and M 2 represent the \nmagnetization moment of VGME by B in different directions. c-e, the change of longitudinal resistance xx in response to \nrotati ng B of 9 T around , and angles at 300 K . The definition of each angle is illustrated inside each figure. Blue \ncurves represent xx of Pt/VGME bilayer, and orange c urves represent xx of Pt/a-C bilayer . \nIn conclusion, w e produced an all-carbon film consisting of vertically aligned graphene edges, which \nwove into a three -dimensional hyperfine -porous network. A reproducible, full section of XMCD spectr um \nacross C-1s K-edge absorption energy was detected for the first time, and the average amplitude of dichroism \nat 28 4.5 eV reaches ~0.0 05 at the 300 K, equals to uni t magnetic moment of ~0.000 6 B per atom. The \nmagnetization mea sured by XMCD is co -confirmed by hysteresis loop of M -H curve . The influence of \nmagnetic impurities on magnetization was excluded by XA NES and ICP -MS measurements. Spin transfer \n \nbehavior of SMR was observed in the VGME film, which verified the long -range and robust features of the \nbulk ferromagnetic order at room temperature . Our work provides direct evidence of elementary resolved bulk \ncarbon ferromagnetism at room temperature and clarifies its origin from -electrons at graphene edges. \n \nMethods: \nSample preparation: \nA silico n wafer of 0.5 mm thickness with a 300 nm oxide layer was used as the substrate for the film growth. \nThe thickness of the VGME was controlled via deposition time. For spin transport measurement s, Pt layers \nwith the thickness of 5 nm were sputtered via thermal evaporation. \nTEM characterization: \nFor TEM characterization, the carbon film with 20nm thickness was transferred from Si substrate to a lacey \ncarbon TEM grid for plan view characterization. The side view lamella was produced by a Focused \nIon/Electron Dual beam system (Thermo -Fisher Scios). Both plan and side view characterization were \nperformed by a double aberration -corrected electron microscope (Thermo -Fisher, Titan Themis G2) with \nCeta2 CMOS camera and HAADF detector in TEM and HAADF STEM modes. The acceleration voltage of \nelectron beam w as set to 80kV to minimize the beam damage to the carbon film. \nRaman spectroscopy : \nFor Raman measurement, the carbon film was deposited on Si substrate with the thickness of 20 nm. The \nspectra were obtained on H ORIBA HR Evolution. An excitation laser with the wavelength of 532 nm was \nused with the emitting power of 1 µW, and the intensity coefficient of 3.2% was used for measurement. The \nnumerical hole -diameter of the project lens was 1 m. The measurement on each sample was repeated three \ntimes at different locations randomly selected on the film surface. The irradiating time of focused laser bea m \non each detecting area was less than 5 min. \nM-H curves: \nFor M -H measurement, the carbon film was deposited on Si substrate with the thickness of 300 nm. The \nsample containing carbon film and S i substrate was cut into the in -plane size of 2 ×4 mm2, and weighted on a \ndigital weighing balance. The result was marked as m 1. The sample was glued onto a plastic sample holder \nin the position that the film surface was perpendicular to the external magnetic line. Hysteresis curve of the \nsample was obtained with a physical property measurement system (on Quantum Design, MPMS3). T he raw \nmagnetization result was recorded as M 1. The Si substrate without carbon film was measured with the same \nmethod. The weight and magnetization of the Si was recorded as m 2 and M 2. Since the thickness of the \ncarbon film (300 nm) was much smaller than that of the substrate (0.5 mm), the contribution of the carbon \nsample to m 2 can be neglected (less than 1‰). Therefore, the magnetization response of carbon film can be \ncalculated as M 1- m1×M2/m2. \nLaser ablation inductively coupled plasma mass spectrometry (LA -ICP-MS): \nFor LA-ICP-MS measurement s, the carbon film was deposited on Si substrate with the thickness of 300 nm. \nThe as -deposited carbon film on Si substrate was cleaned by ultrasonic cleaning in acetone bath and dried \nnaturally. The cleaned sample was mounted flatly in the sample chamber o f laser -ablation chamber \n(NWR193HE ) which was coupled with the ICP-MS system . The chamber was closed and Ar was introduced \nas background. A pulse laser with the frequency of 5 Hz was introduced to evaporate the film. The sampling \ntime at each position was 40 s, and the process repeated at 24 different positions in total. A NIST -610 sample \nwas used for concentration calibration of the most possible metallic impurities, including Al, P, Ti, V , Cr, Mn, \nFe, Co, Ni, Cu, Zn, and Pb. \nX-ray photon spectroscopy (XPS): \nFor XPS measurement, the carbon film was deposited on Si substrate with the thickness of 20 nm. The as \ndeposited sample was cleaned in acetone bath and dried naturally before measurement. The film surface was \ntested firstly, followed by Ar plasma etching for 2 min to remove the part that contacted with atmosphere, and \nthe same test was carried out on the fresh -exposed surface. An X -ray source was used, and the whole spectr um \nwere collected in the energy range from 0 to 1,350 eV . O 1s and C 1s spectra was magnified to check the \ncomposition difference between film body and surface. \nX-ray magnetic circular dichroism \nFor XMCD measurement, 10 times of repeatedly scan were carried out continuously. Each time a MXCD \ncurve was derived from a pair of XA NES spectra taken under opposite magnetic fields. The integrated area of \nXMCD curve AXMCD was calculated by summing up the normalized amplitude of all data points. The area of \nXANES spectra AXANES was derived with the same process. The ratio of the area of XMCD to the area of \nXANES was calculated , and the value r=AXMCD\nAXANES is considered as the effective magnetic moment from each \ncarbon atom. \nDevice fabrication: \nFor transport measurement, the carbon film was deposited on Si substrate with the thickness of 20 nm. The \nsample was cut into pieces (in-plane size 4 × 4 mm2) with a diamond pencil, followed by 3 min ultrasonic \ncleaning in acetone bath. For Hall measurements, a Hall -bar shaped photoresist coating was fabricated on the \ncarbon film by lithography method, and Ar etching was introduced to remove the photoresist and the \nuncovered carbon film, what left was a Hall -bar shaped carbon film. The line width of the Hall bar is 100 m \n(see Fig. S4) . The electrodes were connected to a standard sample puck with Pt wires (diameter of 20 m) by \nusing a wiring bonding machine. For spin transport measurements, a Pt film w as sputtered on the whole \nsurface of the carbon film by DC sputtering in Ar atmosphere, followed by Hall -bar fabrication process as for \nHall measurements, with the Ar etching time slightly longer. \nWe also prepared the devices by directly attaching Pt wire on carbon and Pt/carbon film surfaces with silver \npaste instead of fabricating the Hall -bar sample. The results showed that the influence of different techniques \non measurement results can be neglected . \n \nData availability : \nThe data that support the findings of this study are available from the corresponding author on reasonable \nrequest. \nReferences: \n1. Son, Y . -W., Cohen, M. L. & Louie, S. G. Half -metallic graphene nanoribbons. Nature 444, 347 –349 (2006). \n2. Magda, G. Z. et al. Room -temperature magnetic order on zigzag edges of narrow graphene nanoribbons. \nNature 514, 608 –611 (2014). \n3. Blackwell, R. E. et al. Spin splitting of dopant edge state in magnetic zigzag graphene nanoribbons. Nature \n600, 647 –652 (2021). \n4. Mishra, S. et al. Collective All‐Carbon Magnetism in Triangulene Dimers**. Angew. Chem. Int. Ed. 59, \n12041 –12047 (2020). \n5. Mishra, S. et al. Topological frustration induces unconventional magnetism in a nanographene. Nat. \nNanotechnol. 15, 22–28 (2020). \n6. Pizzochero, M. et al. Edge Disorder in Bottom -Up Zigzag Graphene Nanoribbons: Implications for \nMagnetism and Quantum Electronic Transport. J. Phys. Chem. Lett. 12, 4692 –4696 (2021). \n7. Friedrich, N. et al. Magnetism of Topological Boundary States Induced by Boron Substitution in Graphene \nNanoribbons. Phys. Rev. Lett. 125, 146801 (2020). \n8. Slota, M. et al. Magnetic edge states and coherent manipulation of graphene nanoribbons. Nature 557, 691 –\n695 (2018). \n9. Rizzo, D. J. et al. Inducing metallicity in graphene nanoribbons via zero -mode superlattices. Science 369, \n1597 –1603 (2020). \n10. Tschirhart, C. L. et al. Imaging orbital ferromagnetism in a moiré Chern insulator. Science 372, 1323 –\n1327 (2021). \n11. Sharpe, A. L. et al. Emergent ferromagnetism near three -quarters filling in twisted bilayer graphene. \nScience 365, 605 –608 (2019). \n12. Esquinazi, P. et al. Induced Magnetic Ordering by Proton Irradiation in Graphite. Phys. Rev. Lett. 91, \n227201 (2003). \n13. Yazyev, O. V . Magnetism in Disordered Graphene and Irradiated Graphite. Phys. Rev. Lett. 101, 037203 \n(2008). \n14. Chen, J. -H., Li, L., Cullen, W. G., Williams, E. D. & Fuhrer, M. S. Tunable Kondo effect in graphene with \ndefects. Nat. Phys. 7, 535 –538 (2011). \n15. Ohldag, H. et al. π -Electron Ferromagnetism in Metal -Free Carbon Probed by Soft X -Ray Dichroism. \nPhys. Rev. Lett. 98, 187204 (2007). \n16. Zhang, Y ., Gao, F., Gao, S. & He, L. Tunable magnetism of a single -carbon vacancy in graphene. Sci. \nBull. 65, 194 –200 (2020). \n17. Nair, R. R. et al. Dual origin of defect magnetism in graphene and its reversible switching by molecular \ndoping. Nat. Commun. 4, 2010 (2013). \n18. Hu, W. et al. Embedding atomic cobalt into graphene lattices to activate room -temperature \nferromagnetism. Nat. Commun. 12, 1854 (2021). \n19. Zhang, X., Qi, S. & Xu, X. Long -range and strong ferromagnetic graphene by compensated n –p codoping \nand π–π stacking. Carbon 95, 65–71 (2015). \n20. Ugartemendia, A., Garcia−Lekue, A. & Jimenez−Izal, E. Tailoring magnetism in silicon -doped zigzag \ngraphene edges. Sci. Rep. 12, 13032 (2022). \n21. Park, J. M., Cao, Y ., Watanabe, K., Taniguchi, T. & Jarillo -Herrero, P. Tunable strongly coupled \nsuperconductivity in magic -angle twisted trilayer graphene. Nature 590, 249 –255 (2021). \n22. Wang, Y . et al. Room -Temperature Ferromagnetism of Graphene. Nano Lett. 9, 220 –224 (2009). \n23. Červenka, J., Katsnelson, M. I. & Flipse, C. F. J. Room -temperature ferromagnetism in graphite driven \nby two -dimensional networks of point defects. Nat. Phys. 5, 840 –844 (2009). \n24. Sepioni, M. et al. Limits on Intrinsic Magnetism in Graphene. Phys. Rev. Lett. 105, 207205 (2010). \n25. Nair, R. R. et al. Spin-half paramagnetism in graphene induced by point defects. Nat. Phys. 8, 199 –202 \n(2012). \n26. Wang, C., Diao, D., Fan, X. & Chen, C. Graphene sheets embedded carbon film prepared by electron \nirradiation in electron cyclotron resonance plasma. Appl. Phys. Lett. 100, 231909 (2012). \n27. Wang, C., Zhang, X. & Diao, D. Nanosized graphene crystallite induced strong magnetism in pure carbon \nfilms. Nanoscale 7, 4475 –4481 (2015). \n28. Chen, C. T. et al. Experimental Confirmation of the X -Ray Magnetic Circular Dichroism Sum Rules for \nIron and Cobalt. Phys. Rev. Lett. 75, 152 –155 (1995). \n29. N’Diaye, A. et al. Interplay between Transition -Metal K -edge XMCD and Magnetism in Prussian Blue \nAnalogs. ACS Omega 7, 36366 –36378 (2022). \n30. Avci, C. O. et al. Unidirectional spin Hall magnetoresistance in ferromagnet/normal metal bilayers. Nat. \nPhys. 11, 570 –575 (2015). \n31. Oyanagi, K. et al. Paramagnetic spin Hall magnetoresistance. Phys. Rev. B 104, 134428 (2021). \n \nAcknowledgements: \nThe authors would like to thank Prof. Kai Chen for discussion on XMCD technique , and thank Dr. Duo Zhao \nfor discussion on SMR measurement. We also thank BL12B beamline of National Synchrotron Radiation \nLaboratory (NSRL, China) and BL07U beamline of Shanghai Synchrotron Radiation Facility (SSRF). The \nwork is supported by National Natural Science Foundation of China (grant No. 51875364 ) \nAuthor contributions: \nC.W, D.D and E.S. designed the research. N.J. , M.Y . and X. Z. performed the transmission electron \nmicroscopy analysis. Z.Y. and X. M. prepared the sample and assisted with all measurements. T.K. and S. D. \ndesigned the M -H and SMR measurements, Q. L, W. Y . and D. H. assisted with XMCD measurements. L.Y . \nassisted with M -H measurements. D. D. and E.J. supervised the research. C.W. and D.D. co -wrote the paper. \nAll authors discussed the results and commented on the manuscript. \nCompeting interests: \nThe authors declare no competing interests. \n " }, { "title": "0710.5477v1.Thermoelectrical_manipulation_of_nano_magnets.pdf", "content": "arXiv:0710.5477v1 [cond-mat.mes-hall] 29 Oct 2007Thermoelectrical manipulation of nano-magnets\nA. M. Kadigrobov,1,2R. I. Shekhter,1M. Jonson,1,3and V. Korenivski4\n1Department of Physics, G¨ oteborg University, SE-412 96 G¨ o teborg, Sweden\n2Theoretische Physik III, Ruhr-Universit¨ at Bochum, D-448 01 Bochum, Germany\n3School of Engineering and Physical Sciences, Heriot-Watt U niversity, Edinburgh EH14 4AS, Scotland, UK\n4Nanostructure Physics, Royal Institute of Technology, SE- 106 91 Stockholm, Sweden\n(Dated: October 29, 2018)\nWe propose a device that can operate as a magneto-resistive s witch or oscillator. The device\nis based on a spin-thermo-electronic control of the exchang e coupling of two strong ferromagnets\nthrough a weakly ferromagnetic spacer. We show that the loca l Joule heating due to a high con-\ncentration of current in a magnetic point contact or a nanopi llar can be used to reversibly drive the\nweak ferromagnet through its Curie point and thereby exchan ge-decouple the strongly ferromag-\nnetic layers, which have an antiparallel ground state. Such a spin-thermionic parallel-to-antiparallel\nswitching causes magnetoresistance oscillations where th e frequency can be controlled by proper\nbiasing from essentially DC to GHz.\nManipulation of the magnetic state on the nanome-\nter scale is the central problem of applied magneto-\nelectronics. The torque effect1,2, which is based on the\nexchange interaction of electrons injected into a ferro-\nmagnetic region, is one of the key phenomena leading\nto current-induced magnetic switching. Current-induced\nprecession and switching of the orientation of magnetic\nmoments due to this effect has been observed in many\nexperiments3,4,5,6,7,8,9,10,11,12,13.\nTheefficiencyofcurrent-inducedswitchingis,however,\nlimited by the necessity to work with high current densi-\nties. Anaturalsolutiontothisproblemistouseelectrical\npoint contacts (PCs). This is because now the current\ndensity is high only near the PC, where it can reach14,15\nvalues of order 109A/cm2. The characteristic energies\nsupplied to the electronic system are determined by the\nvoltage drop V, and are for V∼0.1÷1 V (which is eas-\nily reached in experiments) comparable to the exchange\nenergy in magnetic materials.\nOne of the characteristic features of such high current-\ndensity states is the possibility to create local heating\nin regions with a size of a few nanometers. The energy\nsupplied to the electronic system in this restricted region\nresults in an enhanced local temperature, which to a high\ndegree of accuracy can be controlled by the applied volt-\nage. Electrical manipulation of nanomagnetic conduc-\ntors by such controlled Joule heating is a new principle\nfor current-induced magnetic switching. In this Letter\nwe discuss one of the possibilities of the thermoelectri-\ncal magnetic switching causedby a non-linearinteraction\nbetween the spin-dependent electrical transport and the\nmagnetic sub-system of the conductor due to the Joule\nheating effect. We predict that a special design of mag-\nnetic PC can provide both a fast switching and a smooth\nchange of the magnetization direction in nano-size re-\ngions of the magnetic material controlled by the voltage.\nWe also predict temporal oscillations of the magnetiza-\ntion direction in such regions (which are accompanied\nby electrical oscillations) under an applied DC voltage.\nThese phenomena are potentially useful for microelec-\ntronic applications such as memory devices and voltagecontrolled oscillators.\nEquilibrium magnetization distribution in a magnetic\nstack.The system under consideration is shown in Fig.1\nwhere three ferromagnetic layers are coupled to a non-\nmagnetic conductor. We assume the following conditions\nto be satisfied: The Curie temperature T(1)\ncof region 1\nis smaller than the Curie temperatures T(0,2)\ncof regions\n0 and 2; in region 2 there is a magnetic field directed\nopposite to the magnetization of the region, which can\nbe an external field, the fringing field from layer 1, or a\ncombination of the two. We require this combined mag-\nnetostatic field to be weak enough so that at low temper-\natures the magnetization of layer 2 is kept parallel to the\nmagnetization of layer 0 due to the exchange interaction\nbetween them via region 1 (we assume the magnetization\ndirection of layer 0 to be fixed). In the absence of an ex-\nternalfieldandifthetemperature T > T(1)\nc,theproposed\nF/f(N)/F tri-layer is similar to the antiparallel spin-flop\n’free layers’ widely used in memory applications16.\nAsTapproaches T(1)\ncthe magnetic moment of layer\n1 decreases and the exchange coupling between layers 0\nand 2 weakens. This results in an inhomogeneous distri-\nbution of the stack magnetization. Euler’s equation for\nthe magnetization distribution that minimizes the free\nenergy of the system can be written as follows (see, e.g.,\nRef. 17):\nd\ndx/parenleftbigg\nα(x)M2(x)dθ\ndx/parenrightbigg\n−β\n2M2sin2θ+HM\n2sinθ= 0; (1)\nHere the x- andz-axes are directed along the stack and\nthe magnetization direction in the region 0, respectively;\nθ(x) is the angle between the magnetic moment /vectorM(x) at\npointxand the z-axis;M=|/vectorM|;α∼I/aM2, where\nI∼kTcandais the exchange energy and the lattice\nspacing, respectively, kis the Boltzmann constant; βis\nthe dimensionless constant of the anisotropy energy. Be-\nlow we assume the lengths L1,2of regions 1 and 2 to be\nsmaller than the domain wall lengths in these regions.\nUsing Eq. (1) and the boundary condition at the inter-\nfaceSbetween magnetic region 2 and the non-magnetic2\nL2H\nL10 1 2 3\nFigure 1: Orientation of the magnetic moments in a stack of\nthreeferromagnetic layers (0, 1, 2); thepresence ofamagne tic\nfield directed opposite to the magnetization is indicated by an\narrow outside the stack.\nFigure 2: Magnetic moment orientations in the stack at tem-\nperatures T1< T < T(1)\nc.\nregion 3 dθ/dx|S= 0 (see, e.g., Ref. 18) one finds the\nmagnetization in region 1 to be inhomogeneous θ1(x) =\nθ2x/L1,0≤x≤L1, while the magnetic moments in\nregion 2 are approximately parallel with the accuracy\nα1M2\n1(T)/α2M2(T)≪1 (the subscripts 1 and 2 refer to\nthe region 1 and 2, respectively); the dependence of their\ntilt angle θ2onHandTis given by the equation\nθ2=D(H,T)sinθ2;D(H,T) =L1(L2HM2(T))\n4α1M2\n1(T)(2)\nWhile writing Eq.(2) we neglected the anisotropy energy\nin comparison with the magnetic energy in region 2.\nAt low temperatures ( D(T,H)<1) Eq.(2) has the\nsingle root θ2= 0 and hence a parallel orientation of all\nthe magnetic moments in the stackis thermodynamically\nstable. At temperatures T≥T1(hereT1is the tempera-\nture at which D(T,H) = 1) one has D≥1 and a second\nrootθ2/negationslash= 0 appears in addition to the root θ2= 0. The\nparallel magnetization corresponding to θ2= 0 is now\nunstable and the magnetization direction of region 2 tilts\nwith an increase of T(see Fig.2) until at T≥T(1)\ncthe\nexchange coupling between layers 0 and 2 vanishes and\ntheir magnetic moments become antiparallel.\nCurrent-voltage characteristics of the stack under\nJoule heating. If the stack is Joule heated by the cur-\nrentJits temperature T(V) is determined by the heat-\nbalance condition\nJV=Q(T), J=V/R(θ2) (3)\nand Eq. (2) that determines the temperature dependence\nofθ2(T(V)). Here Q(T) is the heat flux from the stack,\nR(θ2) is the stack resistance. Eqs. (2) and (3) define the\ncurrent-voltage characteristics (IVC) of the stack J=\nR(θ2(V))Vin a parametric form.\nAs follows from the previous paragraph, in the whole\nintervals T(V)< T1andT(V)> T(1)\ncthe stack re-\nsistances are R(0) and R(π), respectively, that is IVC1.2\n1.41.4\n1.20.8\n0.80.4\n0.40J/J\nV/VC\nC..aa’\nbb’\nFigure 3: Numerical results for the I−Vcharacteristics\n(IVC) of the magnetic stack. Here R(θ2) =R+−R−cosθ2,\nR−/R+= 0.2,D0= 0.2;Jc=Vc/R+. The branches 0 −a′\nandb−b′of the IVC correspond to the parallel and antipar-\nallel orientations of the stack magnetization, respective ly; the\nbrancha−bcorresponds tothe inhomogeneous magnetization\ndistribution shown in Fig.2\nbranches J=R(0)VandJ=R(π)Vare linear at\nV < V 1=/radicalbig\nR(0)Q(T1) andV > V c=/radicalBig\nR(π)Q(T(1)\nc),\nrespectively. At V1≤V≤Vcthe stack temperature is\nT1≤T(V)≤T(1)\nc, and the magnetization direction of\nregion 2 changes with a change of Vand hence IVC is\nnon-linear there. Differentiating Eqs.(2,3) with respect\ntoVone finds that inside this interval the differential\nconductance Rd(V) =dJ/dVis negative if d[G(θ2)(1−\nD0sinθ2/θ2)]/dθ2<0(hereD0=D(H,T= 0)); e.g., for\nthe stack resistance of the form R(θ2) =R+−R−cosθ2)\n(hereR±= (R(π)±R(0))/2) one has dJ/dV < 0 if\nD0(H)<3r/(1+3r) wherer=R−/R+, and hence IVC\nof the stack is N-shaped as shown in Fig.3.\nSelf-exciting electric, thermal and magnetic direction\noscillations . The set of equations describing thermal and\nelectrical processes in the stack connected in series with\nan inductance Land DC voltage drop ¯V(see Fig.4) is\nCVdT\ndt=J2R(θ2)−Q(T);LdJ\ndt+JR(θ2) =¯V(4)\nwhereCVistheheatcapacity. Forthe casethat themag-\nnetic moment relaxation is the fastest process the tem-\nperature dependence of θ2=θ2(T(t)) is given by Eq.(2).\nThe set of equation Eq.(4) has always a steady-state\nsolutionEqs.(2,3). Astudy ofitsstability with respectto\nsmall perturbations shows that inside the interval V1≤\n¯V≤Vcthis solution is unstable if L>Lcr=tv|Rd(¯V)|\nwheretv∼(TCVR+/¯V2) isthe characteristictime ofthe\nvoltageevolutionand Rd=dV/dJ < 0isthe stackdiffer-\nential resistance. As a result periodic oscillations of the\ncurrentJ(t) and the voltagedrop on the stack V(t) spon-\ntaneous arise: the current J(t) and the voltage drop on\nthe stack V(t) periodically move along the limiting cycle\nin the plane ( J,V) as is shown in Fig.5. The temper-\nature of the stack T(t) and the magnetization direction3\nFigure 4: Equivalent circuit for a Joule-heated magnetic\nstack. A differential resistance Rd(V) biased by a fixed DC\nvoltage¯Vis connected in series with an inductance L;V(t) is\nthe voltage drop over the stack and J(t) is the total current\n1.2J/J\nV/V\n0.90.95 0.9 1.05 1.11.1\ncc\n. .\nFigure 5: Numerical results showing spontaneous oscillati ons\nofthecurrent J(t)andthevoltagedrop V(t)overthestackfor\nthe parameters R−/R+= 0.2,D0= 0.1 and (L−L cr)/Lcr=\n0.25;Jc=Vc/R+. The time development of the current and\nthe voltage drop over the stack follows the dashed line or\nthe thin solid line towards the limiting cycle (thick solid l ine)\ndepending on the initial state.\nθ2(t) adiabatically follows these electric oscillations ac-\ncording to the following equalities: Q(T(t)) =V(t)J(t)\nandθ2(t) =θ2(T(t)) (see Eq.(2)) For ( L−Lcr)/Lcr≪1\nthe oscillation frequency is ω= 1/tv.\nWith afurther increaseofthe inductance the character\nof the periodic motion changes its character: at L ≫Lcrthe current and the voltage slowly move along the\nbranches of the IVC 0 aandbb′with the velocity ˙J/J≈\n1/tj(heretJ=L/R+is the characteristic time of the\ncurrent evolution) fast switching between them at points\naandbwith the velocity ∼1/tv. Therefore, in this case\nthe stack periodically switches between the parallel and\nantiparallel magnetic states.\nIn conclusion , we have shown that Joule heating of\nthe magnetic stack presented in Fig.1 allows electrical\nmanipulations of the mutual magnetization orientation\nof layers 0 and 2 as follows:\n1) In the regime of a controlled current there is a hys-\nteresis loop: with an increase of the current along the\na-a’ IVC branch the magnetic directions are parallel up\nto point a. At this point the stack is heated up to the\ntemperature T1at which D= 1 and the parallel ori-\nented state looses its stability. As a result the system\nswitches to the b−b′branch of IVC at which T > T(1)\nc\nand the magnetic moment /vectorM2switches to the antiparal-\nlel orientation. With a further change of the current the\nstack remains on the b-b’ branch of IVC until the current\nreaches point bwhere the stack is cooled below T(1)\ncand\nsystem switches again to the a−a′branch at which the\nmagnetic moments are parallel.\n2) At the regime of a controlled voltage and a low\ninductance one may smoothly re-orient the magnetiza-\ntion directions from the parallel to the antiparallel (see\nFig.2) changing the applied voltage inside the interval\nV1≤V≤Vcthat is moving along a−bbranch of IVC.\n3) An increase of the inductance of the circuit under a\nfixed DC voltage(see Fig.4) allowsto excite electrical pe-\nriodic oscillations accompanied by periodical switchings\nof the magnetization directions in the stack from parallel\nto the antiparallel orientation.\nAcknowledgement. FinancialsupportfromtheSwedish\nKVA, VR, and SSF is gratefully acknowledged.\n1J.C. Slonczewski, J. Magn. Magn. Mater. 159, L1 (1996);\nibid.195, L261 (1999).\n2L. Berger, Phys. Rev. B 54, 9353 (1996).\n3M. Tsoi, A.G.M. Jansen, J. Bass, W.-C. Chiang, M. Seck,\nV. Tso, and P.Wyder, Phys. Rev. Lett. 80, 4281 (1998).\n4M. Tsoi, A.G.M. Jansen, J. Bass, W.-C. Chiang, V. Tsoi,\nand P.Wyder, Nature (London), 406, 46 (2000).\n5E.B. Myers, D.C. Ralph,J.A. Katine, R.N. Louie, and R.A.\nBuhrman Science 285, 867 (1999).\n6J.A. Katine, F.J. Albert, R.A. Buhrman, E.B. Myers, and\nD.C. Ralph, Phys. Rev. Lett. 84, 3149 (2000).\n7S.I.Kiselev, J.C. Sankey,I.N.Krivorotov, N.C.Emley, R.J .\nSchoelkopf, R.A. Buhrman, and D.C. Ralph, Nature (Lon-\ndon)425, 380 (2003).\n8W.H. Rippard, M.R. Pufall, and T.J. Silva, Appl. Phys.\nLett.82, 1260 (2003).\n9R.S. Beech, J.A. Anderson, A.V. Pohm, and J.M.\nDaughton, J. Appl. Phys. 87, 6403 (2000).10J.M. Daughton and A.V. Pohm, J. Appl. Phys. 93, 7304\n(2003).\n11I.L. Prejbeanu, Kula, W.; Ounadjela, K.; Sousa, R.C.; Re-\ndon, O.; Dieny, B.; Nozieres, J.-P., IEEE Trans. Magn. 40,\n2625 (2004)\n12Jianguo Wang and P.P. Freitas, Appl. Phys. Lett. 84, 945\n(2004).\n13M.Kerekes, R.C. Sousa, I.L. Prejbeanu, O. Redon, U.\nEbels, C. Baraduc, B. Dieny, J-PNozeres, P.P. Freitas, and\nP. Xavier, J. Appl. Phys. 97, 10P501 (2005); R.C. Sousa,\nM.Kerekes, I.L. Prejbeanu, O. Redon, J. Appl. Phys. 99,\n08N904 (2006). B. Dieny, J.P Nozeres,\n14A.V. Khotkevich and I.K. Yanson, Atlas of Point Contact\nSpectraof Electron-Phonon Interactions in Metals, Kluwer\nAcademic Publishers, Boston/Dordrecht/London (1995)\n15J.J. Versluijs, M.A. Bari, and M.D. Coey, Phys. Rev. Lett.\n87, 026601 (2001).\n16D. C. Worledge, IBM Journal of Research and Develop-4\nment50, 69 (2006).\n17L.D. Landau, E.M. Lifshits, and L.P. Pitaevski, Elec-\ntrodynemics of Continuous Media, 43, Butterworth-\nHeinemann, Oxford, Pergamon Press 1998.18A.I. Akhiezer, V.C. Bar’yakhtar, S.V. Peletminskii, Spin\nWaves, 5.4, p.42, North-Holland Publishing Company -\nAmsterdam John Wiley & Sons, INC. - New York, 1968." }, { "title": "1402.4346v1.The_Complexity_of_Ferromagnetic_Two_spin_Systems_with_External_Fields.pdf", "content": "arXiv:1402.4346v1 [cs.CC] 18 Feb 2014The Complexity of Ferromagnetic Two-spin\nSystems with External Fields\nJingcheng Liu1, Pinyan Lu2, and Chihao Zhang1\n1Shanghai Jiao Tong University, {liuexp, chihao.zhang }@gmail.com\n2Microsoft Research. pinyanl@microsoft.com\nAbstract. We study the approximability of computing the partition\nfunction for ferromagnetic two-state spin systems. The rem arkable algo-\nrithmbyJerrumandSinclairshowedthatthereisafullypoly nomial-time\nrandomized approximation scheme (FPRAS) for the special fe rromag-\nnetic Ising model with any given uniform external field. Late r, Goldberg\nand Jerrum proved that it is # BIS-hard for Ising model if we allow incon-\nsistent external fields on different nodes. In contrast to the se two results,\nwe prove that for any ferromagnetic two-state spin systems e xcept the\nIsing model, there exists a threshold for external fields bey ond which the\nproblem is # BIS-hard, even if the external field is uniform.\n1 Introduction\nSpin systems are well studied in statistical physics and applied proba bility. We\nfocus on two-state spin systems in this paper. An instance of a spin system is\ndescribed by a graph G(V,E), where vertices are particles and edges indicate\nneighborhood relation among them. A configuration σ:V→{0,1}assigns one\nof the two states to every vertex. The contribution of local inter actions between\nadjacent vertices is quantified by a matrix A=/bracketleftbigg\nA0,0A0,1\nA1,0A1,1/bracketrightbigg\n=/bracketleftbigg\nβ1\n1γ/bracketrightbigg\n, where\nβ,γ≥0. The contribution of vertices in different spin states is quantified b y\na vector b=/bracketleftbiggb0\nb1/bracketrightbigg\n=/bracketleftbiggµ\n1/bracketrightbigg\n, where µ >0. Thisµis also called the external\nfield of the system, which indicates a priori preference of an isolate vertex. The\npartition function Z(β,γ,µ)(G) of a spin system G(V,E) is defined to be the\nfollowing exponential summation:\nZ(β,γ,µ)(G) =/summationdisplay\nσ∈{0,1}V/productdisplay\nv∈Vbσv/productdisplay\n(u,v)∈EAσu,σv.\nWe call such a spin system parameterized by ( β,γ,µ). If the parameters are\nclear from the context, we shall write Z(G) for short. Although originated from\nstatistical physics, the spin model is also accepted in computer scie nce as a\nframework for counting problems. For example, with β= 0,γ= 1 and µ= 1,\nZ(β,γ,µ)(G) is the number of independent sets (or vertex covers) of the gra phG.Given a set of parameters ( β,γ,µ), it is a computational problem to com-\npute the partition function Z(β,γ,µ)(G) where graph Gis given as input. We\ndenote this computation problem as Spin(β,γ,µ) and want to characterize their\ncomputational complexity in terms of β,γandµ. Forexactcomputation, poly-\nnomial time algorithms are known only for the very restricted settin gs that\nβγ= 1 or ( β,γ) = (0,0), and for all other settings the problem is proved to\nbe #P-hard [2]. Therefore, the main focus is to study their approxim ability.\nFor any given parameter ε >0, the algorithm outputs a number ˆZsuch that\nZ(G)exp(−ε)≤ˆZ≤Z(G)exp(ε) and runs in time poly(n,1/ε), where nis the\nsize of the graph G. This is called a fully polynomial-time approximationscheme\n(FPTAS). The randomized relaxation of FPTAS is called fully polynomial- time\nrandomized approximation scheme (FPRAS), which uses random bits and only\nrequires the final output be within the required accuracy with high p robability.\nThe spin systems ( β,γ,µ) are classified into two families with distinct phys-\nical and computational properties: ferromagnetic systems ( βγ >1) andanti-\nferromagnetic systems ( βγ <1). We shall denote the corresponding computa-\ntion problems respectively by Ferro(β,γ,µ) andAnti-Ferro (β,γ,µ), so as to\nemphasize which family these parameters belong to. Systems with βγ= 1 are\ndegenerate and trivial both physically and computationally. As a res ult, we only\nstudy systems with βγ/ne}a⊔ionslash= 1.\nGreat progress has been made recently for approximately comput ing the\npartition function for anti-ferromagnetic two-spin systems: it ad mits an FPTAS\nup to the uniqueness threshold [19,12,16,13], and is NP-hard to appro ximate in\nthe non-uniqueness range [17,6]. The uniqueness threshold is a phas e transition\nboundary in physics. It is widely conjectured that the computation al difficulty\nis related to the phase transition point in many problems; this is one of the very\nfew examples where a rigorous proof was obtained.\nFor ferromagnetic systems, the picture is quite different. The uniq ueness\ncondition does not coincide with the transition of computational diffic ulty and it\nis not clear whether it plays any role in the computational difficulty. In a seminal\npaper [10], Jerrum and Sinclair gave an FPRAS for ferromagnetic Isin g model\nβ=γ >1withanyexternalfield µ.Thus,thereisnotransitionofcomputational\ndifficulty for ferromagnetic Ising model, which contrasts the situat ion for anti-\nferromagnetic Ising model β=γ <1. For general ferromagnetic spin systems\nwith external field, the approximability is less clear. Since the Ising mo del (β=\nγ) is solved, we focus on the case β/ne}a⊔ionslash=γand always assume β < γby symmetry.\nBy transferring to Ising model, an FPRAS was known for the range o fµ≤/radicalbig\nγ/β[9].\nOn the other hand, a hardness result was obtained for Ising model with in-\nconsistent external fields [7]. This is a generalization of the spin syst em where\nthe external fields for different vertices can be different and take n from a setV.\nWe useSpin(β,γ,V) (Ferro(β,γ,V) orAnti-Ferro (β,γ,V) ) to denote this\ncomputation problem. It is proved that the Ising model with arbitra ry external\nfieldsFerro(β,β,(0,+∞)) is #BIS-hard,namelythe problemis atleast ashard\nas counting independent sets on bipartite graphs (# BIS). #BISis a problem ofintermediate hardness and has been conjectured to admit no FPRA S [5]. The\nreduction used here is called approximation-preserving reduction a s introduced\nin [4]: Let A,B:Σ∗→Rbe two functions. An approximation-preserving reduc-\ntionfromAtoBis a randomized algorithm that approximates Awhile using\nan oracle for B. We write A≤APBfor short if an approximation-preserving\nreduction exists from AtoB. To get that # BIS-hardness result, one need to use\n(or simulate) both arbitrarily small and large external fields. As β < γ, we can\nalways simulate some arbitrarily small external fields with gadgets. H owever,\nsimulating arbitrarily large external fields is only possible when βµ+1> µ+γ,\nin which case one gets a # BIS-hardness result similarly. If this is not the case,\nand in particular if β≤1< γ, no hardness result was known for any bounded\nexternal fields. These systems havecertain monotonicity proper ty, so all external\nfields that can be simulated by gadgets are inherently bounded by ab ove. It was\nnot even clear whether there is any hardness result or not. As a fir st result, we\nshow that the problem is already hard as long as we allow sufficiently larg e (yet\nstill bounded by above) and vertex-dependent external fields.\nTheorem 1. For anyβ < γwithβγ >1, there exsits a bounded set Vsuch that\nFerro(β,γ,V)is#BIS-hard.\nThe main difficulty is for the case of β≤1, for which we cannot simulate\nany external field larger than the upper bound of V. We overcome this difficulty\nby making use of a recent beautiful result in [3]. Instead of startin g with the\nindependent set problem on arbitrary bipartite graphs, we start w ith a soft\n(βγ >0)anti-ferromagnetictwo-spinsystemon boundeddegreebipart itegraphs\nfrom [3]. Then all the external fields needed for the reduction are b ounded.\nHowever, in the above reduction, we do need vertices to have differ ent ex-\nternal fields to make the reduction go through. This gives a hardne ss result for\nFerro(β,γ,V) but not Ferro(β,γ,µ) for a single µ. It is more interesting and\nintriguing (both physically and computationally) to understand the c omputa-\ntional complexity of a uniform spin system ( β,γ,µ) with the same external field\nµon all the vertices. As our main result of this paper, we also prove # BIS-\nhardness on this uniform case for sufficiently large single external fi eldµ. We\nprove that when µis sufficiently large, we can realize by sufficient precision of\nall the external fields which is smaller than µ∗(µ,β,γ), where µ∗(µ,β,γ) is a\nfunction of µ,βandγ, and approaches infinity as µgoes to infinity. Then by\nchoosinglargeenough µand makinguse ofTheorem1, weget ourmain theorem.\nTheorem 2. For anyβ < γwithβγ >1,there exist a µ0such that Ferro(β,γ,µ)\nis#BIS-hard for all µ≥µ0.\nOur main technical contribution is the construction of a family of gad gets\nto simulate a given target external field. We use a reverse idea of co rrelation\ndecay to do that. Correlation decay is proved to be a very powerfu l technique\nto design FPTAS for counting problems (see for examples [19,1,13,16,1 4,15]). In\ncorrelationdecaybasedFPTAS, one firstestablishesa treestruc ture andhope tocompute the marginal probability of the root. With a recursive relat ion, one can\nwrite the marginal probability of the root as a function of that of its sub-trees,\nthen truncate the computation tree at certain depth and do a rou gh guess at\nthe leaf nodes. A correlation decay property ensures that the er ror for the root\nis exponentially small although there are constant error for the lea ves. Here, we\nuse a similar idea to construct a tree gadget so that the marginal pr obability for\nthe root is very close to a target value. Using the same recursion, o ne translates\nthe target marginal probability for the root to that of its sub-tre es. In the leaf\nnodes, we simply use some basic gadgets to approximate the target marginal\nprobabilities. Again, although these approximation for leaves may ha ve constant\ngap, the error at the root is exponentially small thanks to the corr elation decay\nproperty. We believe that this idea of using an algorithm design techn ique to\nbuild gadgets and get hardness result is of independent interest an d may find\napplications in other problems.\nWe also make some improvements on the algorithm side showing that th ere\nis an FPRAS if µ≤γ/β. We remark that all the computational problem\nFerro(β,γ,µ) andFerro(β,γ,V) is no more difficult than # BIS, as we can\nuse the standard transformation to transform any ferromagne tic two-spin sys-\ntem to ferromagnetic Ising model with possibly different external fi elds and use\nthe #BIS-easiness result in [7]. Thus, the two # BIS-hardness theorems can also\nbe stated as # BIS-equivalent. We believe that the conjecture here is that for\nany fixed β < γ, there exists a critical µcsuch that it admits an FPRAS if the\nexternal field µ < µ c, and it is # BIS-equivalent if µ > µ c. The result of this\npaper is an important step towards this dichotomy.\n1.1 Related works\nThe approximationfor partition function has been studied extensiv ely with both\npositive and negative results [19,1,13,16,10,11,18,8,17,6]. For the algor ithm side\nof ferromagnetic two-spin systems, besides the FPRASes, there is also a recent\ndeterministic FPTAS for certain range of the parameters based on correlation\ndecay and holographic reduction [15].\n2 Bounded Local Fields\nIn the section, we show that bounded local fields are sufficient to es tablish a\nhardness result. The following theorem is a formal statement of Th eorem 1.\nTheorem 3. Letβ < γ,βγ >1,∆=⌊√βγ+1√βγ−1⌋+1andµ >/parenleftBig/radicalBig\nγ\nβ/parenrightBig∆\n. Then\nFerro(β,γ,[1,µ])is#BIS-hard.3\n3Technically, we should only define the problem by a finite set o f external fields. In\nthis paper and as in many others, we adopt the following conve ntion: when we say\na problem with an infinite set of external fields is hard, it mea ns that there exists a\nfinite subset of external fields to make the problem hard alrea dy.We first introduce our starting point from anti-ferromagnetic Isin g model on\nbipartite graphs, and show the reduction in the second subsection .\n2.1 Anti-ferromagnetic Spin Systems on Bipartite Graphs\n#BISis a special anti-ferromagnetic two-state spin system. Similar to # BIS,\none can also study other anti-ferromagnetic two-state spin syst ems on bipartite\ngraphs. We use a prefix Bi-to emphasize that input graphs are bipartite, and\na subscript ∆to indicate that maximum degree is ∆. For instance, the prob-\nlem ofAnti-Ferro (β,γ,µ) on bipartite graphs with maximum degree ∆, is\ndenoted shortly by Bi-Anti-Ferro ∆(β,γ,µ). The following theorem from [3]\nis the starting point of our reduction.\nTheorem 4 ([3]). Suppose a set of anti-ferromagnetic parameters (β,γ,µ)lies\nin the non-uniqueness region of the infinite ∆-regular tree T∆and that√βγ≥√∆−1−1√∆−1+1, andβ/ne}a⊔ionslash=γorµ/ne}a⊔ionslash= 1. ThenBi-Anti-Ferro ∆(β,γ,µ)is#BIS-hard.\nFor simplicity, we use the special anti-ferromagnetic Ising model β=γ <1\nin our reduction, where the non-uniqueness condition is easy to sta te.\nProposition 5. Ifβ <∆−1\n∆+1, then there exists a critical activity µc(β,∆)>1\nsuch that the Gibbes measure of Ising model (β,β,µ)on infinite ∆-regular tree\nT∆is unique if and only if |logµ|≥logµc(β,∆).\nProposition 5 is folklore, a proof can be found in, e.g. [16]. Combining t hese\ntwo results, we can get\nCorollary 6. For all0< β <1, there is an ε >0such that for any µ∈\n(1,1+ε),Bi-Anti-Ferro ∆(β,β,µ)is#BIS-hard, where ∆=⌊1+β\n1−β⌋+1.\nProof.As∆=⌊1+β\n1−β⌋+1, we know that β <∆−1\n∆+1. Then by Proposition 5, we\ncan choose ε=µc(β,∆)−1 and get that ( β,β,µ) is in the non-uniqueness region\nof the infinite ∆-regular tree T∆for allµ∈(1,1+ε). In order to make use of\nTheorem 4 and conclude our proof, we only need to verify that β≥√∆−1−1√∆−1+1.\nOur choice of ∆is the smallest integer to satisfy β <∆−1\n∆+1. As a result, we have\nβ≥∆−1−1\n∆−1+1≥√∆−1−1√∆−1+1. ⊓ ⊔\n2.2 The Reduction\nLemma 7. For any β < γwithβγ >1,µ >1and integer ∆ >1, we have\nBi-Anti-Ferro ∆/parenleftbigg1√βγ,1√βγ,µ/parenrightbigg\n≤APBi-Ferro ∆/parenleftBigg\nβ,γ,/bracketleftBigg\n1\nµ/radicalbiggγ\nβ,µ/parenleftbigg/radicalbiggγ\nβ/parenrightbigg∆/bracketrightBigg/parenrightBigg\n.Proof.Let bipartite graph G(L∪R,E) be an instanceofanti-ferromagneticIsing/parenleftBig\n1√βγ,1√βγ,µ/parenrightBig\nwith maximum degree ∆. We construct an instance of ferromag-\nnetic system with exactly the same graph. Each vertex u∈Lwith degree du\nhas weight µ/parenleftBig/radicalBig\nγ\nβ/parenrightBigdu\n, and each vertex v∈Rhas weight1\nµ/parenleftBig/radicalBig\nγ\nβ/parenrightBigdv\n. Then the\nmaximum possible external field is µ/parenleftBig/radicalBig\nγ\nβ/parenrightBig∆\nwhile the minimum one is1\nµ/radicalBig\nγ\nβ.\nTherefore, it is indeed an instance of Bi-Ferro ∆/parenleftbigg\nβ,γ,/bracketleftbigg\n1\nµ/radicalBig\nγ\nβ,µ/parenleftBig/radicalBig\nγ\nβ/parenrightBig∆/bracketrightbigg/parenrightbigg\n.\nLetZ1(G) be the partition function of the anti-ferromagnetic Ising instanc e,\nandZ2(G) be that for the ferromagnetic system. We shall prove that Z1(G) =\nγ−|F|µ|R|Z2(G). LetV/definesL∪R,A=/bracketleftBigg1√βγ1\n11√βγ/bracketrightBigg\n,A′=\n/radicalBig\nγ\nβγ\nγ/radicalBig\nγ\nβ\n,ˆA′=\n/bracketleftbigg\n1β\nγ1/bracketrightbigg\nandˆA=/bracketleftbigg\nβ1\n1γ/bracketrightbigg\n. Then\nZ2(G) =/summationdisplay\nσ∈{0,1}V/productdisplay\n(u,v)∈EˆAσu,σv/productdisplay\nu∈L/parenleftBigg\nµ/parenleftbigg/radicalbiggγ\nβ/parenrightbiggdu/parenrightBigg1−σu/productdisplay\nv∈R/parenleftBigg\n1\nµ/parenleftbigg/radicalbiggγ\nβ/parenrightbiggdv/parenrightBigg1−σv\n=/summationdisplay\nσ∈{0,1}V/productdisplay\n(u,v)∈EˆA′\nσu,σv/productdisplay\nu∈L/parenleftBigg\nµ/parenleftbigg/radicalbiggγ\nβ/parenrightbiggdu/parenrightBigg1−σu/productdisplay\nv∈R/parenleftBigg\n1\nµ/parenleftbigg/radicalbiggγ\nβ/parenrightbiggdv/parenrightBiggσv\n=/summationdisplay\nσ∈{0,1}V/productdisplay\n(u,v)∈EA′\nσu,σv/productdisplay\nu∈Lµ1−σu/productdisplay\nv∈R1\nµσv\n=µ−|R|γ|F|/summationdisplay\nσ∈{0,1}V/productdisplay\n(u,v)∈EAσu,σv/productdisplay\nu∈Lµ1−σu/productdisplay\nv∈Rµ1−σv\n=µ−|R|γ|F|Z1(G).\nThus we can get an approximation for the anti-ferromagnetic Ising model by an\noracle call to the ferromagnetic two-spin system. This concludes t he proof.⊓ ⊔\nNow,given the target µ >/parenleftBig/radicalBig\nγ\nβ/parenrightBig∆\nin Theorem3, we simply choosea µ′close\nenough to 1 in Lemma 7 and Corollary 6, such that/bracketleftbigg\n1\nµ′/radicalBig\nγ\nβ,µ′/parenleftBig/radicalBig\nγ\nβ/parenrightBig∆/bracketrightbigg\n⊆[1,µ]\nand #BIS≤APBi-Anti-Ferro ∆/parenleftBig\n1√βγ,1√βγ,µ′/parenrightBig\n. Then we can conclude that\n#BIS≤APBi-Ferro ∆(β,γ,[1,µ]) and finish the proof of Theorem 3.\n3 Uniform Local Field\nWe establish Theorem 2 in this section. If β >1, then one can use external field\nofµ >γ−1\nβ−1to simulate any external fields and get the # BIS-hardness result.This follows from a similar argument as that in [7]. To be self-contained , we also\ninclude a formal proof in the appendix. So we assume β≤1 in this section. We\nalso introduce a function h(x) =βx+1\nx+γwhich is used throughout this section.\nNote that since βγ >1,h(x) is monotonically increasing and1\nγ< h(x)< β≤1\nforx∈(0,+∞). We shall prove the following key reduction.\nLemma 8. Letβ≤1,βγ >1,dbe an integer such that β(βγ)d>1,µ∗be the\nlargest solution of xtox=µh(x)d, andµ >γd(βγ−1)\nβ/parenleftBig\n1+d+1\nln(β(βγ)d)/parenrightBig\n. Then\nFerro(β,γ,[1,µ∗])≤APFerro(β,γ,µ).\nAsµ∗=µh(µ∗)dand1\nγ< h(µ∗)< β, we have the following bound for µ∗.\nProposition 9.µ\nγd< µ∗< βdµ.\nWith this bound and Lemma 8, we can choose sufficiently large µso that this\nµ∗islargeenoughtoapplythehardnessresult(Theorem3)of Ferro(β,γ,[1,µ∗])\nto get the hardness result for Ferro(β,γ,µ). Formally, we have\nTheorem 10. Letβ≤1,βγ >1,dbe an integer such that β(βγ)d>1,\n∆=⌊√βγ+1√βγ−1⌋+ 1, andµ > γdmax/braceleftbigg/parenleftBig/radicalBig\nγ\nβ/parenrightBig∆\n,βγ−1\nβ/parenleftBig\n1+d+1\nln(β(βγ)d)/parenrightBig/bracerightbigg\n. Then\nFerro(β,γ,µ)is#BIS-hard.\nWe remark that there always exists such integer dsinceβ >0 andβγ >1.\nDifferent ds give different bounds for µand it is not necessarily monotone. For\na givenβ,γ, one can choose a suitable dto get the best bound4.\nIn the remaining of this section, we prove the key reduction stated in Lemma\n8. The main idea is to simulate any external field in [1 ,µ∗] by a vertex weight\ngadget.Inthefirstsubsection,westatethegeneralframewor kofsuchsimulation.\nThen in the second subsection, we present the detailed construct ion of a gadget.\n3.1 Vertex Weight Gadget\nDefinition 11 (Vertex weight gadget). LetG(V,E)be a graph with a special\noutput vertex v∗, defineµ(G) =ZG(v∗=0)\nZG(v∗=1)whereZG(v∗= 0)(resp.ZG(v∗= 1))\nis the partition function of G(V,E)in(β,γ,µ)-system conditioned on v∗= 0\n(resp.v∗= 1). We call Gavertex weight gadget that realizes µ(G).\nWe also use a family ofgraphsto approacha given externalfield. Let {Gi}i≥1\nbeafamilyofvertexweightgadgets.We say {Gi}realizesµiflimi→∞µ(Gi) =µ.\nVertex weight gadgets can be used to simulate external fields. For mally, we\nhave the following reductions.\n4We give one numerical example here to get some idea of this bou nd: ifβ= 1 and\nγ= 2, we can get ∆= 6 and choose d= 1; then the theorem tell us that the problem\nFerro(1,2,µ) is #BIS-hard ifµ >12.Lemma 12. LetGbe a vertex weight gadget of (β,γ,V). ThenSpin(β,γ,V∪\n{µ(G)})≤APSpin(β,γ,V).\nLet{Gi}be a sequence of vertex weight gadget of (β,γ,V)to realize µsuch\nthat for any ε >0there is a Giof sizepoly/parenleftbig\nε−1/parenrightbig\nwithexp(−ε)≤µ(Gi)\nµ≤exp(ε).\nThenSpin(β,γ,V∪{µ})≤APSpin(β,γ,V).\nProof.The proof of the first part is straightforward. For any instance Hof\nSpin(β,γ,V∪{µ(G)}) and a vertex of Hwith external field µ(G), we use one\ncopyofGandidentify the output vertexof Gwith that chosenvertexof H. After\nthe identification, the external field in that vertex is that of outpu t vertex of G.\nTherefore,afterthe modification,the new instanceisaninstance ofSpin(β,γ,V)\nand the partition function is equal to the partition function of Hscaled by a\npolynomial-time computable global factor/parenleftBig\nZ(G)\n1+µ(G)/parenrightBigj\n, wherejis the number of\nvertices with external field µ(G) inH.\nFor the second part, for an instance HofSpin(β,γ,V∪{µ}) and required\napproximation parameter ε, choose a gadget Giwhich is ε′=ε\n2nclose to realize\nµ; dothe samemodificationasaboveusingthis Giand callthe oracleforthe new\ninstance with approximation parameter ε′. This gives the desired approximation\nfor the original instance. ⊓ ⊔\n3.2 The Construction\nWe first define a gadget operation combas follows: for a given list of graphs\nG={G1,...,G k}, each with output v∗\nifori∈[k],comb(G) is a new graph\nG(V,E) thatcombinesthe graphsandjoinstheiroutputs.Fig.1isanillustr ation\nofcomb. Formally, we define V={u}∪/uniontext\ni∈[k]V(Gi) andE={(u,v∗\ni)|i∈\n[k]}∪/uniontext\ni∈[k]E(Gi), where uis the output of G. It is easy to verify that µ(G) =\nµ/producttext\ni∈[k]h(µ(Gi)).\n⇒\nFig.1.Result ofS5⇒comb({S5,S5}), the output vertex is marked as unfilled.\nWealsodefinetwobasicgadgets.Let Swbeaw-stargraph,withoutputbeing\nitscenter.Inparticular, S0isthesingletongraph.Notethat µ(Sw) =µh(µ)w.We\nalso defineTtbe ad-ary tree with depth t. For any external field ˆ µ∈(0,µ∗], we\nshallconstructalistofgadgetstosimulateit.Thetwoboundariesa reapproached\nbySwandTtrespectively.\nProposition 13. LetTtbe ad-ary tree with depth tandSwbe aw-star. Then(1){Sw}w≥1realizes0, or formally, µ(Sw) =µh(µ)w< µβw.\n(2){Tt}t≥0realizesµ∗, or formally, there exist two positive constants ιand\nc <1depending on µ,β,γanddsuch that 1<µ(Tt)\nµ∗≤exp(ctι).\nProof.(1) is obvious, we only prove (2).\nNote that µ(Tt) =µh(µ(Tt−1))d, we denote f(x) =µh(x)d. Recall that µ∗\nis the largest fixed point of f(x) andf(µ)< µ, we have 0 < f′(µ∗)<1. Define\ng(x) =xf′(x)\nf(x), theng(µ∗) =f′(µ∗). Sinceg(x) is a continuous function, we can\nchoose some η >0 such that 0 < g(x)≤c <1 for allx∈(µ∗−η,µ∗+η).\nWe now define a sequence {xi}i≥0such that x0=µandxi=f(xi−1)\nfor alli≥1. We claim that {xi}converges to µ∗asiapproaches infinity.\nTo see this, note that xi+1=f(xi)< xiandxi> µ∗for alli≥0. This\nimplies{xi}converges to some z≥µ∗. Moreover, since fis continuous, the\nsequence{f(xi)}i≥0also converges to z. These two facts together imply z=\nlimi→∞f(xi) =f(limi→∞xi) =f(z). In other word, zis a fixed point of fand\nthusz=µ∗. The claim implies that for some integer t0,xt0∈(µ∗,µ∗+η).\nWe define another sequence {yi}i≥0such that y0=µ(Tt0) andyi=f(yi−1)\nfor alli≥1. It holds that yi∈(µ∗,µ∗+η) and thus g(yi)≤c <1 for all i≥0.\nTherefore for all t≥1,\nlnyt−lnµ∗= lnf(yt−1)−lnf(µ∗)\n=˜yf′(˜y)\nf(˜y)·|lnyt−1−lnµ∗|for some y∈[µ∗,yt−1]\n=g(˜y)·|lnyt−1−lnµ∗|\n≤c·|lnyt−1−lnµ∗|\n≤ctη.\nWe denote ι= max{lnµ,ηc−t0}and conclude the proof. ⊓ ⊔\nOur main idea to realize a target external field ˆ µis to construct a list of\ngadgetsG={G1,...,G k}such that µ(comb(G))≈ˆµor more concretely ˆ µ≈\nµ/producttext\ni∈[k]h(µ(Gi)). All but one of these Giare basic gadgets of the following\nthree types: (1) isolate point S0withµ(S0) =µ; (2)Swwith large enough w\nsuch that µ(Sw)≈0; and (3)Ttwith large enough tsuch that µ(Tt)≈µ∗.\nThe remaining one Giis recursively constructed with a new target ˆ µ′so that\nideally ˆµ=µ/producttext\ni∈[k]h(µ(Gi)) holds. The combination of these basic gadgets\nare carefully chosen so that the new target ˆ µ′is also in the range (0 ,µ∗]. Then\nwe recursively construct this ˆ µ′by a subtree. We terminate the recursion after\nenough steps, and use a basic star gadget which is closest to the de sired value as\nan approximation in the leaf. With a correlation decay argument, we s how that\nthe error in the root can be exponentially small in terms of the depth , although\nthere may be a constant error in the leaf. A detailed construction w ith special\ntreatment for the boundary cases are formally given in Algorithm 1.Algorithm 1: Constructing Gℓ\nfunction construct (ℓ,^µ) :\ninput: Recursion depth ℓ; Target 0 <ˆµ≤µ∗to simulate;\noutput: Graph Gℓconstructed.\nbegin\nifℓ= 0then\nLetkbe the positive integer such that µh(µ)k+1<ˆµ≤µh(µ)k;\nreturn Sk;\nelse\nLetkbe the non-negative integer with µ∗h(µ)k+1<ˆµ≤µ∗h(µ)k;\nY′←k·S0;// a set of kcopies ofS0.\nµ1←ˆµ\nh(µ)k;\n// Invariant: µh(x′)d−i+1=µihas a solution 0< x′≤µ∗.\nfori←1tod−1do\nifµh(µ∗)h(0)d−i≥µithen\nyi←0;w←⌊ℓ·lnα−ln(dµ)\nlnβ⌋+1;Yi←Sw;\nelse\nyi←µ∗;t←⌊ℓ·lnα−lnd−lnι\nlnc⌋+1;Yi←Tt;\nµi+1←µi\nh(yi);\nLet ˆµ′be the solution of µh(x) =µdin (0,µ∗];\nY←Y′∪{Yi}d−1\ni≥1;\nδ←exp(−lnγlnα\nlnβℓ+lnγln(dµ)\nlnβ+lnµ\nγ);\nifˆµ′≤δthen\nChoose the largest integer wsuch that µ/parenleftBig\n1\nγ/parenrightBigw\n> δ;\nreturncomb(Y∪{S w});\nelse\nreturn comb(Y∪construct (ℓ−1,^µ′));\nBeforewe provethat the construction is correct,weobtain a few observations\nwhich is used in our proof. The condition on µin the key Lemma 8 is due to the\nfollowing property we need.\nProposition 14. Letµ >γd\nβ(βγ−1)/parenleftBig\n1+d+1\nln(β(βγ)d)/parenrightBig\n, for any µ1withµ∗h(µ)<\nµ1≤µ∗, the equation µh(x)d=µ1always has a solution with 0< x≤µ∗.\nProof.It suffices to show µ·h(0)d≤µ∗h(µ) andµ·h(µ∗)d≥µ∗. Sinceµ∗=\nµh(µ∗)d, the second part is trivial. As for the first part, it is sufficient to sho w/parenleftBig\nh(µ∗)\nh(0)/parenrightBigd\nh(µ)>1. Note that/parenleftBig\nh(µ∗)\nh(0)/parenrightBigd\nh(µ)> γdh(µ∗)d+1> γd/parenleftBig\nβ−βγ−1\nµ∗/parenrightBigd+1\n,\nγd/parenleftbigg\nβ−βγ−1\nµ∗/parenrightbiggd+1\n>1⇐⇒ln/parenleftbig\nβ(βγ)d/parenrightbig\n+(d+1)ln/parenleftbigg\n1−βγ−1\nβµ∗/parenrightbigg\n>0,(d+1)ln/parenleftbigg\n1−βγ−1\nβµ∗/parenrightbigg\n(♣)\n>−(d+1)βγ−1\nβµ∗\n1−βγ−1\nβµ∗(♠)\n>−ln/parenleftbig\nβ(βγ)d/parenrightbig\n,\nwhere (♣) is due to ln(1−x)>−x\n1−xforx∈(0,1), and (♠) is by the fact that\nβ(βγ)d>1 and the choice of µsuch that−βγ−1\nβµ∗>ln(β(βγ)d)+d+1\nβln(β(βγ)d).⊓ ⊔\nProposition 15. For every x,t≥0, it holds that h(x+t)≤(1 +t)h(x)and\nh((1+t)x)≤(1+t)h(x).\nProof.Note that x,t≥0,\nh(x+t)≤(1+t)h(x)⇐⇒/parenleftbiggβ(x+t)+1\nx+t+γ/parenrightbigg\n≤(1+t)/parenleftbiggβx+1\nx+γ/parenrightbigg\n⇐⇒t2(1+βx)+t/parenleftbig\n1+γ(1+β(x−1))+x+βx2/parenrightbig\n≥0.\nSince/parenleftbig\n1+γ(1+β(x−1))+x+βx2/parenrightbig\n>0, the inequality always holds.\nh((1+t)x)≤(1+t)h(x)⇐⇒x(1+t)β+1\nx(1+t)+γ≤(1+t)βx+1\nx+γ\n⇐⇒t2(x+βx2)+t(γ+2x+βx2)≥0\nAgain every term is non-negative, the last inequality is always true. ⊓ ⊔\nInthefollowing,westarttoverifythecorrectnessoftheconstr uction.Wefirst\nverify that the algorithm is well defined, namely µh(x) =µddoes havea solution\nˆµ′in (0,µ∗]. This can be done by verifying the loop invariant “ µh(x′)d−i+1=µi\nhas a solution 0 < x′≤µ∗” inductively.\nInitialization. Fori= 1, by Proposition 14, for some 0 <˜x≤µ∗it holds that\nµh(˜x)d−i+1=µi.\nMaintenance. Assuming µh(˜x)d−i+1=µihas solutions ˜ x∈(0,µ∗], we verify\nthatµh(x′)d−i=µi+1≡µi\nh(yi)has solutions x′∈(0,µ∗] fori∈[1,d−1].\nCaseµh(µ∗)h(0)d−i≥µi. By assumption we have µh(0)d−i+1< µi, also\nnote that µi≤µh(µ∗)h(0)d−i≤µh(0)h(µ∗)d−i, henceµh(0)d−i<µi\nh(0)≤\nµh(µ∗)d−i. Then by continuity, µh(x′)d−i=µi\nh(0)has solutions 0 < x′≤µ∗.\nCaseµh(µ∗)h(0)d−i< µi.Byassumption µh(µ∗)d−i+1≥µi,thusµh(0)d−i<\nµi\nh(µ∗)≤µh(µ∗)d−i, henceµh(x′)d−i=µi\nh(µ∗)has solutions 0 < x′≤µ∗.\nTermination. After the loop completes, µh(x′) =µdhas solutions0 < x′≤µ∗.\nNow we verify the vertex weight gadget returned by the construc tion satisfies\nour requirement by choosing ℓ=O(−logε).\nLemma 16. For0<ˆµ≤µ∗(β,γ,µ), and let G(V,E)be the graph returned\nbyconstruct (ℓ,ˆµ), we have the following: (1) exp(−(c+ℓ)·αℓ)≤µ(G)\nˆµ≤\nexp((c+ℓ)·αℓ), wherec= lnγandα=√βγ−1√βγ+1<1; (2)|G|= exp(O(ℓ)).Proof.We apply induction on ℓfor both statements. We prove for (1) first. For\nbase case ℓ= 0, we have\n|lnµ(G)−ln ˆµ|≤/vextendsingle/vextendsinglelnµh(µ)k−lnµh(µ)k+1/vextendsingle/vextendsingle=−lnh(µ)≤lnγ.\nAssume the statement holds for smaller ℓ. Letk,{yi}1≤i≤d−1and{Yi}1≤i≤d−1\nbe parameters chosen in the algorithm. Define\nF(z) = ln/parenleftBigg\nµh(µ)kd−1/productdisplay\ni=1h(yi)h(exp(z))/parenrightBigg\n,˜F(z) = ln/parenleftBigg\nµh(µ)kd−1/productdisplay\ni=1h(µ(Yi))h(exp(z))/parenrightBigg\n.\nWe note that F(z) is thecorrectrecursion to compute ln( µ(G)) and˜F(z) is our\napproximate recursion used in the algorithm.\nIn the following, we distinguish between ˆ µ′≤δand ˆµ′> δ.\n–If ˆµ′≤δ, then ln µ(G) =˜F(lnµ(Sw)) and ln ˆ µ=F(lnˆµ′). We have\nF(ln ˆµ′)≤˜F(lnµ(Sw)) = ln/parenleftBigg\nµh(µ)kd−1/productdisplay\ni=1h(µ(Yi))h(µ(Sw))/parenrightBigg\n(♥)\n≤αℓ+ln/parenleftBigg\nµh(µ)kd−1/productdisplay\ni=1h(yi)h(ˆµ′)/parenrightBigg\n=αℓ+F(ln ˆµ′)),\nwhere (♥) follows from following facts derived from Proposition 15:\n(i) Ifyi= 0, then 0≤µ(Yi)≤αℓ\nd, which implies h(µ(Yi))≥h(yi) and\nh(µ(Yi))≤/parenleftBig\n1+αℓ\nd/parenrightBig\nh(yi)≤exp/parenleftBig\nαℓ\nd/parenrightBig\nh(yi).\n(ii) Ifyi=µ∗, thenµ∗≤µ(Yi)≤exp/parenleftBig\nαℓ\nd/parenrightBig\nµ∗, which implies h(µ(Yi))≥\nh(yi) andh(µ(Yi))≤exp/parenleftBig\nαℓ\nd/parenrightBig\nh(yi).\n(iii) We claim that ˆ µ′< µ/parenleftBig\n1\nγ/parenrightBigw\n≤µ(Sw)≤µβw≤αℓ\nd. The only nontrivial\npart is to verify that µβw≤αℓ\nd. Sincewis the largest integer that\nˆµ′< µ/parenleftBig\n1\nγ/parenrightBigw\n, we have µ/parenleftBig\n1\nγ/parenrightBigw+1\n≤ˆµ′, which gives w≥lnµ−lnδ\nlnγ−1.\nPlug this into µβw≤αℓ\ndand letδ= exp(−lnγlnα\nlnβℓ+lnγln(dµ)\nlnβ+lnµ\nγ),\nthe inquality holds. Thus h(µ(Sw))≥h(ˆµ′) andh(µ(Sw))≤h(αℓ\nd)≤/parenleftBig\n1+αℓ\nd/parenrightBig\nh(ˆµ′)≤exp/parenleftBig\nαℓ\nd/parenrightBig\nh(ˆµ′).\n–If ˆµ′> δ, definex=µ(construct (ℓ−1,^µ′)), then by induction hypothesis,\nit holds that|lnx−ln ˆµ′|≤(c+(ℓ−1))·αℓ−1.ThensimilarlybyProposition15andthechoiceof wandt,wehave F(lnx)≤\n˜F(lnx)≤F(lnx)+αℓ. Thus by construction, we have\n|lnµ(G)−ln ˆµ|=/vextendsingle/vextendsingle/vextendsingle˜F(lnx)−F(ln ˆµ′)/vextendsingle/vextendsingle/vextendsingle\n≤αℓ+|F(lnx)−F(lnˆµ′)|\n≤αℓ+|F′(ln˜x)|·|lnx−ln ˆµ′|(for some ˜ x∈[ˆµ′,x].)\n≤αℓ+(ℓ−1)|F′(ln˜x)|αℓ−1+c|F′(ln˜x)|αℓ−1\nThus it is sufficient to show that |F′(ln˜x)|≤α.In fact,F′(lnx) =x·h′(x)\nh(x)=\n(βγ−1)x\n(x+γ)(βx+1)≤βγ−1\n(√βγ+1)2=α.\nNow we prove (2) of the Lemma. We denote s(ℓ) = max ˆµ|construct (ℓ,ˆµ)|\nand show that s(ℓ) =ℓexp(O(ℓ)) = exp( O(ℓ)).\nIfℓ= 0, since ˆ µis either the eventual external field (which is a constant\nbounded away from 0), or ˆ µ > δ, we have s(ℓ) =|Sk|=O(1).\nIfℓ >0, then|Yi|= exp(O(ℓ)) and thus|Y|= exp(O(ℓ)). By our choice of δ,\nit holds that w=O(ℓ) and thus|Sw|=O(ℓ). Therefore,\ns(ℓ) = exp(O(ℓ))+max{s(ℓ−1),O(ℓ)}=ℓexp(O(ℓ)) = exp( O(ℓ)).\nThis concludes the proof. ⊓ ⊔\nReferences\n1. Antar Bandyopadhyay and David Gamarnik. Counting withou t sampling: Asymp-\ntotics of the log-partition function for certain statistic al physics models. Random\nStructures & Algorithms , 33(4):452–479, 2008.\n2. Andrei A. Bulatov and Martin Grohe. The complexity of part ition functions.\nTheoretical Computer Science , 348(2-3):148–186, 2005.\n3. Jin-Yi Cai, Leslie Ann Goldberg, Heng Guo, and Mark Jerrum . Approximating\nthe partition function of two-spin systems on bipartite gra phs.arXiv preprint\narXiv:1311.4451 , 2013.\n4. Martin Dyer, Leslie Ann Goldberg, Catherine Greenhill, a nd Mark Jerrum. The\nrelative complexity of approximate counting problems. Algorithmica , 38(3):471–\n500, 2004.\n5. Martin E. Dyer, Leslie Ann Goldberg, and Mark Jerrum. An ap proximation tri-\nchotomyfor boolean #csp. Journal of Computer and System Sciences , 76(3-4):267–\n277, 2010.\n6. A. Galanis, D. Stefankovic, and E. Vigoda. Inapproximabi lity of the partition\nfunction for the antiferromagnetic ising and hard-core mod els.Arxiv preprint\narXiv:1203.2226 , 2012.\n7. Leslie Ann Goldberg and Mark Jerrum. The complexity of fer romagnetic ising\nwith local fields. Combinatorics, Probability & Computing , 16(1):43–61, 2007.\n8. Leslie Ann Goldberg and Mark Jerrum. Approximating the pa rtition function of\nthe ferromagnetic potts model. Journal of the ACM , 59(5):25, 2012.9. Leslie Ann Goldberg, Mark Jerrum, and Mike Paterson. The c omputational com-\nplexityof two-state spin systems. Random Structures & Algorithms , 23(2):133–154,\n2003.\n10. Mark Jerrum and Alistair Sinclair. Polynomial-time app roximation algorithms for\nthe ising model. SIAM Journal on Computing , 22(5):1087–1116, 1993.\n11. MarkJerrum, Alistair Sinclair, andEric Vigoda. Apolyn omial-time approximation\nalgorithm for the permanent of a matrix with nonnegative ent ries.Journal of the\nACM, 51:671–697, July 2004.\n12. Liang Li, Pinyan Lu, and Yitong Yin. Approximate countin g via correlation decay\nin spin systems. In Proceedings of SODA , pages 922–940, 2012.\n13. Liang Li, Pinyan Lu, and Yitong Yin. Correlation decay up to uniqueness in spin\nsystems. In Proceedings of SODA , pages 67–84, 2013.\n14. Chengyu Lin, Jingcheng Liu, and Pinyan Lu. A simple FPTAS for counting edge\ncovers. In Proceedings of SODA , 2014.\n15. Pinyan Lu, Menghui Wang, and Chihao Zhang. FPTAS for weig hted fibonacci\ngates and its applications. Submitted .\n16. Alistair Sinclair, Piyush Srivastava, and Marc Thurley . Approximation algorithms\nfor two-state anti-ferromagnetic spin systems on bounded d egree graphs. In Pro-\nceedings of SODA , pages 941–953, 2012.\n17. Allan Sly and Nike Sun. The computational hardness of cou nting in two-spin\nmodels on d-regular graphs. In Proceedings of FOCS , pages 361–369, 2012.\n18. Eric Vigoda. Improved bounds for sampling coloring. In Proceedings of FOCS ,\npages 51–59, 1999.\n19. Dror Weitz. Counting independent sets up to the tree thre shold. In Proceedings\nof STOC , pages 140–149, 2006.\nA The β >1 case\nWe prove the hardness result for the case of β >1. This follows a similar argu-\nment as that in [7] and is known as a folklore. We include a formal proof here\nto be self-contain.\nTheorem 17. Letγ > β > 1andµ >γ−1\nβ−1. ThenFerro(β,γ,µ)is#BIS-hard.\nWe follow the same idea of simulating external field and making use of Th e-\norem 3 to conclude the proof. In this case, we can simulate all positiv e external\nfields.\nLemma 18. For every ˆµ >0, there is afamily of vertex weight gadgets {Gm}m≥1\nthat realizes ˆµ. Moreover, Gmis constructible in time mO(1)and\nexp(−1\nm)≤µ(Gm)\nˆµ≤exp(1\nm). (1)\nProof.For any m≥1, we add xself-loops and ybristles to a single vertex v,\nwherexandyare integers to be determined. Let vbe the output of Gm, then\nµ(Gm) =µ/parenleftBig\nβ\nγ/parenrightBigx/parenleftBig\nµβ+1\nµ+γ/parenrightBigy\n. Denote a= lnγ\nβ,b= lnµβ+1\nµ+γandc=ln ˆµ\nlnµ, then (1)\nis equivalent to\n|(y·b−x·a)−c|≤1\nm.We can use a procedure similar to extended Euclidean algoroithm to fin d such\nintegersx,yin timeO(lnm), such that it also guarantees x,y=mO(1).\nB Improved Tractable Result\nIn this section, we establish the following tractable result:\nTheorem 19. Letβ < γ,βγ >1andµ≤γ/β. Then there is an FPRAS for\nFerro(β,γ,µ).\nThe proof of this theorem follows by refining the proof in [9], where t hey\nestablish the tractable result for µ≤(γ/β)δ/2forδbeing the minimum degree\nof vertices in the graph. Specifically, we first contract all vertices with degree one\nand modify the external fields of their neighboring vertices, this on ly scales the\npartition function by a constant. Next, just as in [9], we shall reduc e a (β,γ,µ)\ninstance to a ferromagnetic Ising instance and apply the following ce lebrated\nresult, which is first introduced in [10] for uniform external fields an d refined for\nnon-uniform external fields in [9]:\nTheorem 20 ([10] and [9]). There is an FPRAS for Ising system (a,a,V)\nprovided that a >1and all external fields in Vare at most one.\nLetG(V,E) be an instance of ( β,γ,µ) system, we repeatedly apply the fol-\nlowing operations until no degree one vertices can be found:\n1. Pick a vertex uof degree one. Denote its incident edge by e= (u,v). Letµu\nandµvbe external fields on uandvrespectively.\n2. Remove uand edge ( u,v), update µv←µvh(µu).\nLetG′(V′,E′) be the remaining graph. G′either has no vertices of degree\none, or it only contains a single vertex. Moreover, for every v∈V′, the external\nfieldsµ′\nvsatisfiesµ′\nv≤µ. This can be easily verified given that µ≤γ/β. Let\nU={µ′\nv|v∈V′}, consider G′as an instance of ( β,γ,U) system, clearly\nZ(β,γ,µ)(G) =Z∗·Z(β,γ,U)(G′)whereZ∗isaneasilypolynomial-timecomputable\nfactor.\nLetV=/braceleftbigg\nµ′\nv/parenleftBig\nβ\nγ/parenrightBigdv/2/vextendsingle/vextendsingle/vextendsinglev∈V′/bracerightbigg\nwheredvis the degreeof vinG′. LetˆG(ˆV,ˆE)\nbe a copy of G′with ˆµv=µ′\nv/parenleftBig\nβ\nγ/parenrightBigdv/2\nfor every v∈ˆV. We are going to verify\nthatZ(β,γ,U)(G′) =/radicalBig\nγ\nβ|E′|\n·Z(a,a,V)(ˆG) fora=√βγ.\nDefineA=/bracketleftbiggβ1\n1γ/bracketrightbigg\n,A′=/bracketleftbiggγ/radicalbig\nγ/β/radicalbig\nγ/β γ/bracketrightbigg\nandˆA=/bracketleftbigg√βγ1\n1√βγ/bracketrightbigg\n. Then,Z(β,γ,U)(G′) =/summationdisplay\nσ∈{0,1}V′/productdisplay\n(u,v)∈E′Aσu,σv/productdisplay\nv∈V′µ′1−σv\nv\n=/summationdisplay\nσ∈{0,1}V′/productdisplay\n(u,v)∈E′A′\nσu,σv/productdisplay\nv∈V′\n/parenleftBigg/radicalBigg\nβ\nγ/parenrightBiggdv\nµ′\nv\n1−σv\n=/radicalbiggγ\nβ|E′|/summationdisplay\nσ∈{0,1}V′/productdisplay\n(u,v)∈E′ˆAσu,σv/productdisplay\nv∈V′\n/parenleftBigg/radicalBigg\nβ\nγ/parenrightBiggdv\nµ′\nv\n1−σv\n=/radicalbiggγ\nβ|E′|/summationdisplay\nσ∈{0,1}ˆV/productdisplay\n(u,v)∈ˆEˆAσu,σv/productdisplay\nv∈V′ˆµ1−σv\nv\n=/radicalbiggγ\nβ|E′|\n·Z(a,a,V)(ˆG)\nFinally, to apply Theorem 20, we only need ˆ µv≤1 for allv∈ˆV. Recall that\nˆGhasδ≥2, hence µ≤γ/βimplies ˆµv≤1. To sum up, this concludes the proof." }, { "title": "1405.1951v2.Peculiar_long_range_supercurrent_in_SFS_junction_containing_a_noncollinear_magnetic_domain_in_the_ferromagnetic_region.pdf", "content": "arXiv:1405.1951v2 [cond-mat.supr-con] 19 May 2014Peculiar long-range supercurrent in SFS junction containi ng a noncollinear magnetic\ndomain in the ferromagnetic region\nHao Meng∗and Xiuqiang Wu\nNational Laboratory of Solid State Microstructures and Dep artment of Physics, Nanjing University, Nanjing 210093, Ch ina\n(Dated: August 27, 2018)\nWe study the supercurrent in a superconductor-ferromagnet -superconductor heterostructure con-\ntaining a noncollinear magnetic domain in the ferromagneti c region. It is demonstrated that the\nmagnetic domain can lead to a spin-flip process, which can rev erse the spin orientations of the\nsinglet Cooper pair propagating through the magnetic domai n region. If the ferromagnetic layers\non both sides of magnetic domain have the same features, the l ong-range proximity effect will take\nplace. That is because the singlet Cooper pair will create an exact phase-cancellation effect and\ngets an additional πphase shift as it passes through the entire ferromagnetic re gion. Then the\nequal spin triplet pair only exists in the magnetic domain re gion and can not diffuse into the other\ntwo ferromagnetic layers. So the supercurrent mostly arise s from the singlet Cooper pairs and the\nequal spin triplet pairs are not involved. This behavior is q uite distinct from the common knowl-\nedge that long-range supercurrent induced by inhomogeneou s ferromagnetism stems from the equal\nspin triplet pairs. The result we presented here provides a n ew way for generating the long-range\nsupercurrent.\nPACS numbers: 74.78.Fk, 73.40.-c, 74.50.+r, 73.63.-b\nThe interplay between superconductivity and ferro-\nmagnetism in mesoscopic structures has been extensively\nstudied because of the underlying rich physics and po-\ntential applications in spintronics and quantum informa-\ntion1–4. When a homogeneous ferromagnet ( F) is sand-\nwiched between two s-wave superconductors ( S) to form\na Josephson junction, the magnetic configuration of the\nFlayer may substantially modify the spatial properties\nof the superconducting order parameter. This behav-\nior is induced by the different action of the ferromag-\nnetic exchange field hon the spin-up and spin-down elec-\ntrons that form the Cooper pair. Then this pair inside\ntheFlayer acquire a relative phase shift Q·R, where\nQ≃2h//planckover2pi1vF,vFis the Fermi velocity and Ris the thick-\nness of the Flayer. This phase shift changes with Rand\nresults in an oscillation of critical current accompanied\nwith a rapid decay. The above oscillation will lead to the\ntransition from the so-called 0 state to the πstate1–4.\nTwo kinds of approaches have been proposed to pro-\nduce long-range supercurrent in SFSjunction. The first\napproachrequiresinhomogeneousmagnetism in Fregion\nso that the interested equal spin triplet pairs can be gen-\nerated2. One way to achieve this purpose is by arranging\nferromagnetictrilayerwith noncollinearmagnetizations5.\nInthisgeometry,thespin-flipprocessesattheinterface F\nlayers can convert the singlet Cooper pairs into the equal\nspin triplet pairs4,6. The triplet pairs can penetrate into\nthe central Flayer over a long distance unsuppressed by\nthe exchange interaction so that the proximity effect is\nenhanced. The second approach requires the Flayer to\nbe arrangedantiparallel. This situation was described by\nBlanteret al.for aSFFSjunction7. The physical origin\nof the enhanced proximity effect is described as a com-\npensation of the relative phase shift of a Cooper pair as\nit passes from the first Flayer into the second one. If the\ntwoFlayers have the same thickness, the net change inthe relative phase of the Cooper pair is zero in the clean\nlimit. This enhanced Josephson current has been proved\nby recent experiment8.\nIn this paper, we predict the third approach to gener-\nate the long-rangesupercurrent in SFSstructure, shown\nschematically in figure 1(a). The ferromagnetic region\nconsists of two ferromagnetic layers ( FLandFR) with\nmagnetizationsorientedin same directions. The FLlayer\nandFRlayer are separated by a clean magnetic domain\n(FM), whose magnetization is misaligned with direction\nof theFLlayer andFRlayer. The magnetic domain FM\ncaninduceaspin-flipprocess, whichreversesthe spinori-\nentations of the singlet Cooper pair propagating through\ntheFMregion. Thisprocesswill makethe singletCooper\npair create a phase-cancellation effect and obtain an ad-\nditionalπphase shift. If the FLlayer andFRlayer have\nthe same thickness and exchange field, the net change\nin the relative phase of a singlet Cooper pair is πwhen\nit passes through the entire Fregion. In this case, the\ncontribution to the long-rangesupercurrentmostly arises\nfrom the singlet Cooper pairs. This is because the equal\nspin triplet pairs only display in FMregion and can not\ndiffuse into the FLandFRlayers.\nIn our numerical calculation, the transport direction\nis along the yaxis, and the system is assumed to be\ninfinite in the x-zplane. The BCS mean-field effective\nHamiltonian1,9is\nHeff=/integraldisplay\nd/vector r{/summationdisplay\nα,βψ†\nα(/vector r)[He(ˆ1)αβ−(/vectorh·/vector σ)αβ]ψβ(/vector r)\n+1\n2[/summationdisplay\nα,β(iσy)αβ∆(/vector r)ψ†\nα(/vector r)ψ†\nβ(/vector r)+h.c.]},(1)\nwhereHe=−/planckover2pi12∇2/2m−EF,ψ†\nα(/vector r) andψα(/vector r) are cre-\nation and annihilation operators with spin α. ˆσand\nEFare Pauli matrices and the Fermi energy, respec-2\ndL dR(a)\ndMFMFL\nz\nxyθ s−wave Ss−wave SFR\n−2 024681012−4−3−2−101234\n(b)FM\nFLFRs−wave S s−wave S\nFIG. 1. (color online) (a) Schematic diagram of SFSstruc-\nture with two ferromagnetic layers FLandFRoriented along\nthezaxis and separated by a noncollinear magnetic domain\nFM. The lengths of FL,FRandFMare denoted by dL,dR\nanddM, respectively. The phase difference between the two\nSs isφ=φR−φL. (b) The transmission of electron and hole\nin above structure.\ntively. The superconducting gap is given by ∆( /vector r) =\n∆(T)[eiφLΘ(−y) +eiφRΘ(y−dF)] withdF=dL+\ndM+dR. Here, ∆( T) accounts for the temperature-\ndependent energy gap. It satisfies the BCS relation\n∆(T) = ∆0tanh(1.74/radicalbig\nTc/T−1) withTcthe supercon-\nducting critical temperature. Θ( y) is the unit step func-\ntion, andφL(R)is the phase of the left (right) S. The\nexchange field /vectorhdue to the ferromagnetic magnetizations\nin theFregion is described by\n/vectorh=\n\nhLˆz, 00[u↑(y)v∗\n↓(y)+u↓(y)v∗\n↑(y)]η(t),(6)\nf1(y,t) =1\n2/summationdisplay\nE>0[u↑(y)v∗\n↑(y)−u↓(y)v∗\n↓(y)]η(t),(7)3\n0 50 100 15000.20.40.60.8\nkFdL¯hIc/e∆\n00.25 0.50.75 100.10.20.3\nhL/EF¯hIc/e∆\n08162400.10.20.3\nkFdM¯hIc/e∆\n0 0.2 0.400.10.20.3\nhM/EF¯hIc/e∆(b) (a)\nFIG. 2. (Color online) (a) Critical current as a function of\nlengthkFdL(=kFdR) for exchange field hM/EF= 0 (black\nline) and hM/EF= 0.17 (magenta line) then kFdM= 10,\nand inset shows the critical current versus kFdMforkFdL=\nkFdR= 100. Parameters used in (a): hL/EF=hR/EF= 0.1\nandhM/EF= 0.17. (b) Critical current as a function of\nexchange field hL/EF(=hR/EF) forhM/EF= 0.17, and\ninset shows the critical current versus hM/EFforhL/EF=\nhR/EF= 0.1. Parameters used in (b): kFdL=kFdR= 100\nandkFdM= 10. In all plots θ=π/2.\nwhereη(t) = cos(Et)−isin(Et)tanh(E/2kBT). The\nabove singlet and triplet pair amplitudes are all normal-\nized to the value of the singlet pairing amplitude in a\nbulkSmaterial. The LDOS is given by17\nN(y,ǫ) =−/summationdisplay\nα[u2\nα(y)f′(ǫ−E)+v2\nα(y)f′(ǫ+E)],(8)\nwheref′(ǫ) =∂f/∂ǫis the derivative of the Fermi func-\ntion. The LDOS is normalized by its value at ǫ= 3∆0\nbeyond which the LDOS is almost constant.\nUnless otherwise stated, we use the superconducting\ngap ∆ 0as the unit of energy. The Fermi energy is\nEF= 1000∆ 0and temperature is T/Tc= 0.1. In self-\nconsistent field method, we consider the low tempera-\nture limit and take kFdS1=kFdS2=400,ωD/EF=0.1, the\nother parameters are the same as the ones mentioned\nabove.\nIn figure2(a), wepresent the dependence ofthe critical\ncurrentIcon the length kFdL(=kFdR) for two different\nexchange fields of FMregion. It is well known that, if\nhM/EF= 0 the spin-flip process does not occur in FM\nregion, the critical current Icexhibits oscillations with\na period 2πξFand simultaneously decays exponentially\non the length scale of ξF1. Here,ξFis the magnetic\ncoherence length. The main reason is described below:\na Cooper pair entering into the Flayer receives a finite\nmomentum Qfrom the spin splitting of up and down\nbands. So the spin singlet pairing state |↑↓∝an}b∇acket∇i}ht− |↓↑∝an}b∇acket∇i}ht in\nSwill be converted into the mixed state |↑↓∝an}b∇acket∇i}hteiQ·R− |↓↑\n∝an}b∇acket∇i}hte−iQ·RinFlayer, thus leading to a modulation of the\npair amplitude with the thickness RofFlayer. Then the\n0−πtransition will arise due to spatial oscillations of the\npair amplitude. In this case the phase shift induced by\ntheFlayer is additive and the relative phase is generally\nnonzero so that the supercurrents are suppressed.In contrast, when hM/EF= 0.17,Icwill slowly de-\ncrease with the thickness kFdL. The electron and hole\ntransport process is shown in figure 1(b). Because the\nmagnetization direction of the FMregion is along the x\naxis (θ=π/2), which is orthogonal to the magnetization\ninFLlayerandFRlayer,the spin-flip processwill appear\nin theFMregion. As a result, when a electron |↑∝an}b∇acket∇i}htetrans-\nmits fromFLlayer toFRlayer, the spin-flip can convert\n|↑∝an}b∇acket∇i}hteinto|↓∝an}b∇acket∇i}hte. Subsequently, the |↓∝an}b∇acket∇i}hteis Andreev reflected\natFRSinterface and will be further converted into hole\n|↑∝an}b∇acket∇i}hth. The|↑∝an}b∇acket∇i}hthmoving to left is consequently inverted to\n|↓∝an}b∇acket∇i}hth. Finally,this |↓∝an}b∇acket∇i}hthwillpropagatetothe SFLinterface\nandbereflectedbackastheoriginal |↑∝an}b∇acket∇i}hte. Hence, thespin-\nflip process occurring in the FMregion can reverse the\nspin orientations of the electron and Andreev-reflected\nhole transporting between FLlayer andFRlayer. While\nif the mixed state |↑↓∝an}b∇acket∇i}hteiQ·R− |↓↑∝an}b∇acket∇i}hte−iQ·Rpasses from FL\nlayer intoFRlayer, it will be converted into a new state:\n|↓↑∝an}b∇acket∇i}hteiQ′·R′− |↑↓∝an}b∇acket∇i}hte−iQ′·R′\n=−(|↑↓∝an}b∇acket∇i}hte−iQ′·R′− |↓↑∝an}b∇acket∇i}hteiQ′·R′)\n=|↑↓∝an}b∇acket∇i}htei(−Q′·R′+π)− |↓↑∝an}b∇acket∇i}hte−i(−Q′·R′+π).(9)\nTherefore, when the Cooper pair passes through the FL\nlayer, it will acquire a relative phase shift δχ1=QdL.\nSimilarly, traversing through the FRlayer, it gets the\nother phase shift δχ2=−Q′dR+π. As a result, this sit-\nuation can be described as a superposition of the phase\nshift of the Cooper pair as it travels through the entire\nFregion:χ=δχ1+δχ2. If theFLandFRlayer have\nthe same exchange field and thickness, the net change\nin the relative phase of the Cooper pair is χ=π, and\nit will not turn to 0 with increase of the length of FL\nlayer andFRlayer. Provided one does not take abso-\nlute value for Ie(φ) to define the critical current Ic,Ic\nis always negative and is correspond to the πstate. As\nthe magenta line shown in figure 2(a), Icalso displays\nan oscillatory behavior with increasing the length kFdL.\nFrom this behavior, it is easier to deduce that the spin\norientations of the part of Cooper pairs will be inverted\nby spin-flip in FMregion, so that these pairs can lead\nto a substantially enhanced critical current but do not\nprovide the oscillations for this current. However, the\nrest of Cooper pairs pass through the FMregion without\nspin-flipandcannotgeneratethecancellationoftherela-\ntive phase. Consequently, the transmission of these pairs\nmakes the critical current oscillate with the length of en-\ntireFregion. In addition, another interesting property\nis the nonmonotonic dependence of the critical current\nIcas a function of the length kFdM(see the inset in\nfigure 2(a)). We find that the maximum of Icis nearly\nlocated atkFdM= 10. It means that the spin-flip ratio\nreaches their maximal value in this condition.\nNext, we discuss the dependence of critical current\nIcon the exchange field hL/EF(=hR/EF). As plot-\nted in figure 2(b), Icalmost decreases monotonically to\n0 with the increase of hL/EF. This is easily under-\nstood, because normal Andreev reflection occurred at4\n−40−20 0204000.10.20.3\nkF(dL−dR)¯hIc/e∆\n−0.04−0.02 00.020.0400.10.20.3\n(hL−hR)/EF¯hIc/e∆\n00.25 0.50.75 1−0.3−0.2−0.100.10.2\nφ/π¯hIe(φ)/e∆\n \n−2−1 0120.511.52\nǫ/∆LDOS\n kF(dL−dR)=0 [A]\nkF(dL−dR)=10 [B]\nkF(dL−dR)=20 [C]A\nBC\n(c)(a) (b)\n(d)\nFIG. 3. (Color online) Critical current (a) as a function of\nthe length difference kF(dL−dR) forhL/EF= 0.1, and (b)\nas a function of the exchange field difference ( hL−hR)/EF\nforkFdL= 100. The current-phase relation Ie(φ) (c) and the\naveraged LDOS in all Fregion (d) corresponding to the point\nA, B and C in panel (a). The parameters in FMandFRlayers\nhave the fixed values kFdM= 10,kFdR=100,hM/EF= 0.17,\nhR/EF= 0.1 andθ=π/2. Here the LDOS is calculated at\nkBT= 0.001.\ntheSFLandFRSinterface will be suppressed by ex-\nchange splitting of FLlayer andFRlayer. Especially, if\nhL/EF=hR/EF= 1, theFLlayer andFRlayer are\nall converted into half metal and only one spin band can\nbe occupied, then the Andreev reflections at the inter-\nfaces will be completely prohibited. In this case, none of\nCooper pairs can transmit from the left Sto the right\none, so the Josephson current would be complete sup-\npressed. This feature further demonstrates that the cur-\nrent mostly arises from the contribution of the singlet\nCooper pairs but not the equal spin triplet pairs. That\nis because the equal spin triplet pairs can penetrate over\na long distances into the half metal and will scarcely be\naffected by exchange splitting6,19. In addition, the inset\nin figure 2(b) shows the hM/EFdependence of the criti-\ncal current Ic. It also displays a nonmonotonic behavior\nas thehM/EFis increased, and the maximum is nearly\nseated athM/EF= 0.17.\nTo further demonstrate the conclusion mentioned be-\nforehand, we now discuss intriguing influence of the\nlength and exchange field on the critical current Icwhen\ntheFLlayer andFRlayer have nonidentical physical fea-\ntures. As illustrated in figure 3(a) and (b), we present\nthe dependence of IconkF(dL−dR) and (hL−hR)/EF\nrespectively on condition that the exchange field and\nlength ofFRlayer are all fixed. Take first one for ex-\nample, we find that the dependence of Icon length dif-\nferencekF(dL−dR) look like a “Fraunhofer pattern”.\nWith the length difference close to 0, Icwill increase\nand also accompanies the transition between the 0 and π\nstates. Asmentionedabove,if FLlayerandFRlayerhavethe identical exchange fields ( hL/EF=hR/EF= 0.1)\nbut different lengths, the Cooper pair passing through\ntheFLlayer andFRlayer could acquire the phase shift\nχ=Q(dL−dR) +π. So the variation of length dif-\nference can lead to the oscillation of Ic. On the other\nhand, ifFLlayer andFRlayer have the same length\n(kFdL=kFdR= 100) but different exchange fields, the\nCooper pair can get the phase shift χ= (Q−Q′)dL+π.\nIcwill also oscillate with ( hL−hR)/EFbecause ofQ∝h\n(see figure 3(b)).\nIn addition, the current-phase relations Ie(φ) in par-\nticular points are illustrated in figure 3(c), and corre-\nsponding LDOSs are plotted in figure 3(d). If one take\nthe identical parameter in FLandFRlayer, such as\nkFLL=kFLR= 100 and hL/EF=hR/EF= 0.1,\nthe Cooper pairs will obtain a net phase shift χ=π.\nThen we could observe a negative Josephson current and\na shark zero energy conductance peak in LDOS, which\nindicate the junction is located in πstate. When the\nlength difference kF(dL−dR) = 10, it corresponds to\nthe transition point between the 0 and πstates of the\njunction. At this critical point, the harmonic I1sinφof\nthe current ( I(φ) =I1sinφ+I2sin(2φ) +...) vanishes,\nand theI2sin(2φ) will be fully revealed. Subsequently,\nthe sign of Icis changed with the increase of the length\ndifference. For kF(dL−dR) = 20, the junction is in the\n0 state, and the LDOS at ǫ= 0 will be converted from\nthe peak to a valley.\nFinally, we discuss the dependence of Icon the misori-\nentation angle θ, and the spatial distributions of the spin\nsinglet pair and the spin triplet pair for two different θ.\nOne can see from the inset in figure 4(a) that, when the\norientation of the magnetization in the FMregion is per-\npendicular to the direction of FLlayer andFRlayer, the\ncritical current reaches the maximum. However, it de-\ncreases to minimum on condition that the magnetization\nofFMregion is parallel or antiparallel to the one in FL\nlayer andFRlayer. For given thickness of the FMregion,\nit is possible to find the exchangefield at which switching\nbetween parallel and perpendicular orientations will lead\nto switching of Icfrom near-zero to a finite value. This\neffect may be used for engineering cryoelectronic devices\nmanipulating spin-polarized electrons. Furthermore, we\nfind the spin singlet pair amplitude f3oscillates in all F\nregion atθ= 0. But it will be coherent counteracted\natθ=π(see the main plot in figure 4(a)). In contrast,\nthe spin triplet pair amplitude f0are almost identical in\nabove two cases. For θ= 0, the equal spin triplet pair\namplitudef1is zero in entire Fregion. However, f1only\nsurvives in FMregion and can not exist in FLlayer and\nFRlayer under the condition of θ=π. From these con-\nsequences, we can derived that the long-range supercur-\nrent mostly arises from the coherent propagation of f3,\nand the contributions of f0andf1can be ruled out. To\nfurther uncover this contribution, we take into account\nf3induced by only one S. According to above theory,\nthe spin mixed state state in FLlayer is expressed as\n|↑↓∝an}b∇acket∇i}hteiQ·R− |↓↑∝an}b∇acket∇i}hte−iQ·R= (|↑↓∝an}b∇acket∇i}ht− |↓↑∝an}b∇acket∇i}ht )cos(Q·R) +i(|↑↓5\n050100150200−0.0500.050.10.15\nkFyf3\n0 100 200−0.0400.040.08\nkFyf0\n \n0 100 200048x 10−3\nkFyf1\n050100150200−0.0400.040.08\nkFyf3\n 00.250.50.75100.10.20.3\nθ¯hIc/e∆θ=0\nθ=π/2\nkFdS1=400, kFdS2=0\nkFdS1=0, kFdS2=400(a) (b)\n(c) (d)\nFIG. 4. (color online) Spatial distributions of the spin sin glet\npair amplitude f3(a), the real parts of spin triplet pair am-\nplitudef0(b) andf1(c) for two angles θ= 0 and θ=π/2 at\nωDt= 12. The panels (a), (b) and (c) utilize the same legend.\nInset in (a) shows the critical current as a function of the an -\ngleθ. (d) The f3plotted as a function of kFyfor two cases\nkFdS1= 400,kFdS2= 0 and kFdS1= 0,kFdS2= 400 when\nθ=π/2. Parameters used in all figures: kFdL=kFdR= 100,\nkFdM= 10,hL/EF=hR/EF= 0.1,hM/EF= 0.17 and\nφ= 0.\n∝an}b∇acket∇i}ht+|↓↑∝an}b∇acket∇i}ht)sin(Q·R). This state in FRlayer can be\nconverted as |↑↓∝an}b∇acket∇i}htei(−Q′·R′+π)− |↓↑∝an}b∇acket∇i}hte−i(−Q′·R′+π)= (|↑↓\n∝an}b∇acket∇i}ht− |↓↑∝an}b∇acket∇i}ht)cos(Q′·R′+π) +i(|↑↓∝an}b∇acket∇i}ht+|↓↑∝an}b∇acket∇i}ht)sin(Q′·R′). We\ncan see the spin singlet pair has a phase shift πin theexpression of the FLlayer andFRlayer. But the spin\ntriplet pair with zero spin projection shows the same\ndescription in these two layers. These inferences are in\nagreement with our numerical results. As demonstrated\nin figure 4(d), f3shows a antisymmetric configuration\naround the middle FMregion. As a result, the superpo-\nsition of two f3stemming from the left and right Swill\nmake their amplitudes cancel each other.\nIn conclusion, we have studied numerically the long-\nrange supercurrent in a SFSstructure including a non-\ncollinear magnetic domain in the ferromagnetic region.\nWe find the magnetic domain could induce a spin-flip\nprocess, which can reverse the spin orientations of the\nsinglet Cooper pair when this pair propagate through\nthe magnetic domain region. This process will make the\nsinglet Cooper pair generate a phase-cancellation effect\nand acquire an additional πphase shift. If the ferromag-\nnetic layers on both sides of magnetic domain have the\nsame features (such as thickness and exchange field), the\nnet change in the relative phase of the singlet Cooper\npair isπwhen the pair passes through the entire Fre-\ngion. In this case, the long-range supercurrent mostly\nstems from the singlet Cooper pairs. The reason is that\nthe equal spin triplet pairs are only present in the mag-\nnetic domain region and can not spread to the other two\nferromagnetic layers. It is hoped that our results could\npropose a new way to generate the long-range Josephson\ncurrent.\nThis work is supported by the State Key Pro-\ngram for Basic Research of China under Grants No.\n2011CB922103andNo. 2010CB923400,andtheNational\nNatural Science Foundation of China under Grants No.\n11174125 and No. 11074109.\n∗menghao1982@shu.edu.cn\n1A. I. Buzdin, Rev. Mod. Phys. 77, 935 (2005).\n2F. S. Bergeret, A. F. Volkov, and K. B. Efetov, Rev. Mod.\nPhys. 77, 1321 (2005).\n3A. A. Golubov, M. Yu. Kupriyanov, and E. Ilichev, Rev.\nMod. Phys. 76, 411 (2004).\n4M. Eschrig, Phys. Today 64, No. 1, 43 (2011).\n5M. Houzet and A. I. Buzdin, Phys. Rev. B 76, 060504(R)\n(2007).\n6M. Eschrig, J. Kopu, J. C. Cuevas, and G. Schon, Phys.\nRev. Lett. 90, 137003 (2003); M. Eschrig and T. Lofwan-\nder, Nature Physics 4, 138 (2008).\n7Y. M. Blanter and F.W. J. Hekking, Phys. Rev. B 69,\n024525 (2004).\n8J. W. A. Robinson, Gabor B. Halasz, A. I. Buzdin, and M.\nG. Blamire, Phys. Rev. Lett. 104, 207001 (2010).\n9P.G. de Gennes, Superconductivity of Metals and Alloys,\nBenjamin, New York, 1966 (Chap.5).\n10Hao Meng, Xiuqiang Wu, and Zhiming Zheng, Europhys.Lett. 104, 37003 (2013).\n11Jun-Feng Liu and K. S. Chan, Phys. Rev. B 82, 184533\n(2010).\n12G. E. Blonder, M. Tinkham, and T. M. Klapwijk, Phys.\nRev. B 25, 4515 (1982).\n13A. Furusaki andM. Tsukada, Solid State Commun. 78, 299\n(1991).\n14Z. M. Zheng and D. Y. Xing, J. Phys.: Condens. Matter\n21, 385703 (2009).\n15Y. Tanaka and S. Kashiwaya, Phys. Rev. B 56, 892 (1997).\n16J. B. Ketterson and S. N. Song, Superconductivity, Cam-\nbridge University Press, 1999 (Part III).\n17Klaus Halterman, Oriol T. Valls, andPaul H.Barsic, Phys.\nRev. B 77, 174511 (2008).\n18L. D. Landau, E.M. Lifshitz, Quantum Mechanics, Non-\nRelativistic Theory (third ed.) Pergamon, Elmsford, NY\n(1977).\n19R. S. Keizer, S. T. B. Goennenwein, T. M. Klapwijk, G.\nMiao, G. Xiao and A. Gupta, Nature 439, 825 (2006)." }, { "title": "0810.4081v2.Itinerant_Ferromagnetism_in_the_electronic_localization_limit.pdf", "content": "arXiv:0810.4081v2 [cond-mat.str-el] 2 Feb 2009Itinerant Ferromagnetism in the electronic localization l imit\nN. Kurzweil, E. Kogan, and A. Frydman\nThe Department of Physics, Bar Ilan University, Ramat Gan 52 900, Israel\n(Dated: October 30, 2018)\nWe present Hall effect, Rxy(H), and magnetoresistance, Rxx(H), measurements of ultrathin films\nof Ni, Co and Fe with thicknesses varying between 0.2-8 nm and resistances between 1 MΩ - 100 Ω .\nBoth measurements show that films having resistance above a c ritical value, RC, (thickness below\na critical value, dC) show no signs for ferromagnetism. Ferromagnetism appears only for films with\nR < R C, whereRCis material dependent. We raise the possibility that the rea son for the absence\nof spontaneous magnetization is suppression of itinerant f erromagnetism by electronic disorder in\nthe strong localization regime.\nPACS numbers: 72.15.Rn; 75.70.Ak; 73.61.At\nFerromagnetism in the transition metals (Fe, Co and\nNi) relies, at least partially, on the itinerancy of the con-\nduction band electrons. The balance between kinetic en-\nergy loss and exchange energy gain leads to polarization\nof the band electrons and to magnetic order. It is inter-\nesting to askwhat happens when the mobility ofthe elec-\ntrons is strongly suppressed by disorder. In the extreme\ncase electrons can be localized with localization length\nξ, that may be of atomic scale. Under these conditions\none may expect itinerant ferromagnetism to be entirely\nsuppressed since the electronic functions do not overlap\nand exchange energy is no longer relevant. This may be\nanalogous to the situation in uniform disordered super-\nconductors, where the superconductivity is suppressed\nby strong disorder as a result of disorder induced pair\nbreaking (see for example [1, 2, 3]).\nIn order to achieve high enough disorder for significant\nlocalizationinhomogeneousmetallicmaterialsitisneces-\nsary to use very thin amorphous films having thicknesses\nof a few mono-atomic layers in which the sheet conduc-\ntance can be of the order ofe2\nhor less. For this purpose\nthe samples studied in this work were Ni, Co and Fe thin\nfilms fabricated using ”quench condensation” (evapora-\ntiononacryo-cooledsubstrate). Thistechniqueallowsto\ndeposit sequential layers of ultrathin films and measure\ntransport without thermally cycling the sample orexpos-\ning it to atmosphere. If a thin underlayer of Ge or Sb is\npre-deposited before quench condensing a metal film, the\nunderlayerwetsthesubstratethusenablingthegrowthof\ncontinuousultrathinamorphouslayersevenatmonolayer\nthickness [4]. The underlayer being an insulator at low\ntemperatures is assumed to have negligible effect on the\nelectricpropertiesofthemetal[2, 4]. Theabilitytostudy\na single sample while driving it from strong localization\nto weak localization has been vastly used in the context\nof the superconductor-insulator transition. In the cur-\nrent work we used this technique to drive a ferromag-\nnetic amorphous film from strong to weak localization.\nWe studied Ni, Co and Fe films having resistance in the\nrange of 1 MΩ - 100 Ω and thicknesses of 0.2-8 nm. The\nfilmswereevaporatedona2nmthicklayerofSborGe, in0 2 4 6 8 10 0.0 0.2 0.4 0.6 0.8 1.0 σ/σ 0\nd(mfp) 1 222k 23k 24k R4k =0.3k Ω\n RS(Ω)\nlog(T) (a) (b) \n0.020 0.025 0.030 0.035 -2 -1 01R4k =150 k Ω\n ln(R s)\n1/T 0.5 [1/K 0.5 ]\nFIG. 1: Relative conductivity versus the thickness in mean\nfree path units for a Fe sample . The solid line is a fit to eq.\n1. The insets show resistance as a function of temperature in\nthe weak (a) and strong (b) localization regime (for 3.5 and\n0.3 nm thick samples respectively).\na Hall bar geometry allowing 4 probe resistance and Hall\neffect measurements. Both sample resistance and thick-\nness were monitored during the growth and the process\nwas stopped at different evaporation stages. Room tem-\nperature AFM measurements of these films (performed\ninambientafterheatingthem up) revealthicknessrough-\nness smaller than 1nm.\nFigure 1 shows the normalized conductance versus\nthickness of a Fe sample. The data is fitted the following\nexpression extracted from the scaling theory of localiza-\ntion [5, 6]:\nσ/σ0= 1−3A/(2k2\nfdl)ln[τϕ/3τ], (1)\nwhereσ0is the Drude conductivity, kfis Fermi wave\nlength of the bulk, d is the thickness of the film and\nτandτϕare the transport and dephasing time respec-\ntively. Eq. 1. is well known in the theory of weak local-\nization. We have introduced a phenomenological factor\nA≈3 which enables us to extend the fit well beyond the\nexpected validity range into the strong disorder regime.\nThe discussion on justification of this procedure will be\npresented elsewhere [6]. The fitted relevant lengthscales:2\nl(nm) lϕ(nm) RC(Ω)dC(nm) µ(µB)\nFe 0.4 46 80k 0.5 2.2\nCo 0.29 21 10k 0.8 1.7\nNi 0.27 25 2k 1.8 0.6\nTABLE I: Mean free path, l, dephasing length, lϕ, critical\nresistance and thickness for the appearance of ferromagnet ic\nsignatures, RCanddC, and bulk magnetic moments, µB, for\nthe Ni,Co and Fe films use in this study.\nl-the mean free path and lϕ- the dephasing length are\nlisted in table 1. The insets depict the resistance versus\ntemperature curves for 0.3 nm and 3.5nm thick Fe film\nshowing Efros-Shklovskii like hopping behavior and log-\narithmic dependence typical of weak localization behav-\nior respectively. Extracting the localization lengths from\nsuch R(T) curves allows us to determine the crossover\nfrom strong localization ( ξ < L ϕ) to weak localization\n(ξ > Lϕ) at a sheet resistance of ∼10kΩ and thickness\nof∼0.8 nm.\nIn order to probe the magnetic state of the films we\nmeasured both Hall effect (HE) and magnetoresistance\n(MR) for each evaporation stage by applying a magnetic\nfield perpendicular to the films. All presented results\nwere obtained at T=4K. Hall resistivity in magnetic ma-\nterials is combined of the ordinary part, ρ0, observed\nin all normal metals due to the Lorentz force which is\nproportional to the magnetic field, H, and the extraor-\ndinary Hall effect (EHE), ρxy, which is proportional to\nthe magnetization of ferromagnetic films, M, so that\nρxy∝REHEµ0M,REHEbeing the extraordinary Hall\neffect coefficient. The two processes, which are believed\nto have the major contribution to the EHE, skew scat-\ntering and side jump, are both scattering processes, and\nthe EHE is expected to depend on the longitudinal re-\nsistivity (which is also due to scattering) in the following\nway:\nREHE=aρxx+bρ2\nxx (2)\nwhereais the skew scattering coefficient and bis the side\njump coefficient. Gerber et-al [7] showed that for very\nthin films the dominant factor is skew scattering and ρxx\nwas found to be proportional to ρxyfor Ni films between\n4 to 20 nm thick. For high magnetic fields the magneti-\nzation saturates and only the ordinary part contributes\nto the magneticfield dependence ofthe HE. Studying the\nEHE is an elegant way to measure magnetic properties\nof thin films using transport measurement, in particular,\nthe saturation magnetization can be obtained by extrap-\nolating the Hall resistance to zero field.\nFig 2a. shows the Hall effect of sequential quench con-\ndensed layers of Ni. Down to a sheet resistance of 2 kΩ\n(1.8 nm thick) only the ordinary Hall effect is observed.\nAs more material is added and the resistance further de-\ncreasedan EHE contribution develops and increaseswithH(T) (b) (a) \n-0.75 -0.50 -0.25 0.00 0.25 0.50 0.75 \n 6k Ω\n 4k Ω\n 2k Ω\n 1.5k Ω\n 1k Ω\n 0.7k Ω\n 0.3k ΩρH(µΩ cm) \n-4 -2 0 2 4-0.10 -0.05 0.00 0.05 ∆R/R( %)\nFIG. 2: Hall effect (a) and magnetoresistance (b) measure-\nments of sequential quench condensed Ni films. T=4.2K. The\ndash line in the negative MR of the 0.3 kΩ stage is the fit\nfor to the AMR phenomenological expression in eq. 3. The\nsolid lines in the positive MR curves show fits to the anti\nweak-localization theory of ref [9] with l=0.265 nm, lϕ= 25\nnm andlsoequal to 20, 21.6 and 22.7 nm for films with sheet\nresistance 6, 4, and 2 kΩ respectively.\nincreasing thickness. This defines a critical resistance,\nRC, and critical thickness, dC, for the appearance of the\nEHE. Fig 3 shows both the longitudinal resistivity, ρxx,\nandthe transverse(Hall) resistivity(determined fromex-\ntrapolating to H=0), ρxy, as a function of thickness for a\ntypical Ni sample. ρxxdecreasesmonotonously while ρxy\nis zero for the thinnest films. For films thicker than 2 nm\nρxyincreases with thickness until at d=4-5 nm it changes\nits trend. For thick enough samples (d >4 nm)ρxy∝ρxx\nin accordance with [7]. If the REHE∝ρxxexpression is\nextrapolated to smaller d, our results indicate that the\nmacroscopic magnetization, M, is zero for very resistive\nsamples in the strong localization regime, it grows in an\nintermediate regime until it reaches saturation for films\nhaving sheet resistance below 500 Ω.\nInterestingly, the magnetoresistance also undergoes a\nunique crossover at the same critical region. For high re-\nsistance (small thickness) a positive magnetoresistance is\nobserved, whereas for R < R C(d > dC) the MR changes\nsign and becomes negative, saturating at fields above a\nsaturationfield, HS. Thisisseeninfigure2bwhichshows\nthat the MR exhibits a change of MR sign for resistances\nbelow 2 kΩ ( R < R C).\nSimilar behavior is observedin Co and Fe as well, how-\neverRCanddCvary from material to material. Table\n1 summarizes the critical thicknesses and resistances for3\nwhich EHE appears and the MR changes sign for the\nthree transition metals.\nJust like the emergence of EHE, we interpret the MR\nsign change as a signature for the appearance of ferro-\nmagnetism in the film. Positive MR such as seen in our\nthin layers is also observed in quench condensed non-\nmagnetic layers such as Ag, where the MR remains pos-\nitive for films with resistance as low as 100 Ω. This be-\nhavior is understood as being due to weak localization\nin the presence of strong spin orbit. Since the films are\nultrathin, one can indeed expect strong Rashba spin or-\nbit scattering [8] on the surface leading to weak anti-\nlocalization and positive MR. The solid lines in fig 2b\nshows fits of our curves to the spin-orbit weak localiza-\ntion expression [9]. The spin orbit length, lSO, grows as\nthe film thickens as can be expected if Rashba spin orbit\nis the dominant factor.\nThe crossover to negative MR occurs only in ferro-\nmagnetic materials. Dugaev [10] showed that weak anti-\nlocalization due to spin orbit interaction does not oc-\ncur in ferromagnetic systems so that if ferromagnetism\nis present in a film, the MR should be always negative.\nHence, the MR sign change is also an indication for the\nappearance of magnetization in the film. We have tried\nto fit the negative MR curve to the Dugaev expression\n[10] but could obtain reasonable fits only for very thin\n(d <0.5nm) Fe layers [6]. For thicker samples we find\nthat the curves are well described by anisotropic magne-\ntoresistance (AMR) typical to ferromagnetic films. Since\nthe field is applied perpendicular to the current direc-\ntion one can expect a negative contribution that should\ndepend quadratically on the angle between the current\nand the magnetization [11]. We have fitted our curves to\nthe following phenomenological expression that assumes\nquadratic dependance of ∆ Ron H for low fields and sat-\nuration for fields larger than the saturation field, HS:\n∆R(H) = ∆R(∞)H2\nH2+H2\nS(3)\nA typical fit is seen in the 0.3 kΩ stage in fig. 2b. The\nextracted saturated fields from these fits match those ex-\ntracted from the Hall effect curves for all stages of evap-\noration.\nHence both HE and MR measurements imply that\nfor strong enough disorder, R > R C, the sample does\nnot show signs of spontaneous magnetization and that\nferromagnetism emerges only for films characterized by\nsmaller resistances. One may wonder whether this ef-\nfect can be due to granularity in the film. While the\nsamples are electrically continuous for films with thick-\nnesses of 0.2 nm we can not rule out the presence of\nsome film granularity. For comparison we have produced\ngranular films of Ni with grain sizes of about 20 nm in\ndiameter and 2 nm in height [12]. These are achieved by\nquench condensing Ni on a bare Si/SiO substrate with-\nout depositing a Sb or Ge wetting layer [13, 14]. For0 2 4 6 8 10 12 01000 2000 \n0.0 0.2 0.4 0.6 0.8 1.0 \nρxy (µΩ cm) ρxx (µΩ cm) \nd(nm) 10 210 310 410 510 624\nRXX (Ω)RXY (Ω)\n \nFIG. 3: Longitudinal, ρxx, and saturated transverse (Hall)\nresistance, ρxyof a homogenous Ni sample versus the thick-\nness of the films. The inset shows the saturated transverse\n(Hall) resistance, Rxyversus longitudinal sheet resistance of\na Nigranular film.\nthe thinnest films of this type the grains are superpara-\nmagnetic showing no hysteresis in the magnetoresistance\ncurve [13]. In the granular case, EHE is observed even\nfor the highest-resistance measured samples R∼1MΩ.\nAs material is added the resistance drops considerably,\nhowever, the Hall effect resistance hardly changes until\nR <10kΩ, as can be seen in the inset of fig 3. Such\nbehavior has been also reported for granular Fe prepared\nby different methods [15] whereit wasunderstood that in\nthe granularcase the EHE is due to scattering within the\ngrains rather than being due to hopping between grains.\nForR <10kΩRxyreduces linearly with Rxx. This is\nthe regime in which the grains are expected to coalescing\nand the behavior approaches that of a continuous film.\nIn any case, a granular film behaves very differently than\nour uniform films.\nWe have also considered the possibility that appear-\nance of the magnetization is a function of the film thick-\nness. A number of experiments on epitaxially grown Ni,\nCoandFethinfilmsonCusubstrates[16, 17,18, 19]have\nshown that the Curie temperature, TCdrops sharply to\nzero for thicknesses approaching an atomic monolayer.\nThesesampleswereorderedcrystallinefilms andthe sub-\nstrates were metallic, hence, no effect of electron local-\nization could be expected. The suppression of TCwas\ninterpreted as the formation of magnetic dead layers due\nto electronic hybridization with the states in the Cu sub-\nstrate. This reduces the density of states and the Stoner\ncriterion is defied. A similar conclusion was drawn by\nBergman [20] who saw no signs for magnetization in the\nHall effect of quench condensed Ni evaporated on an\namorphousmetallic Pb75Bi25substratesforfilms thinner\nthan 2.5 monolayers. It seems that our results are differ-\nent from all the above. We note that our films are grown\non insulating substrates, hence electronic hybridization\nis not relevant. Consequently we are unable to detect\nmagnetization in Ni films that are 1 .8 nm thick ( ∼7\nmonolayers), which is much thicker than any of the pre-4\n-5 -4 -3 -2 -1 0246 1nm-Co \n 0.1nm-Ag \n 0.2nm-Ag \n 0.3nm-Ag \n 0.4nm-Ag \n 0.5nm-Ag RH(Ω)\nH(T) 0 2 4 6 80.04 0.06 0.08 0.10 \n \n0 1 2 3 4 5 60.08 0.12 0.16 0.20 0.24 \n \ndAg ( Å)RXy (Ω)(a) \n(b) RH(Ω)\nH(T) \nFIG. 4: Hall effect measurements of a series of a 1.1 nm Co\nfilm coated by Ag overlayers. The inset shows the saturated\ntransverse (Hall) resistance, Rxyof a 2nm thick Ni film (a)\nand a 1nm thick Co sample (b) versus the thickness of the Ag\noverlayer.\nvious experiments.\nIn order to test the notion that localization of the con-\nduction electrons and not the thickness of the films is\nthe cause for absence of magnetization we performed an\nexperiment designed to vary the degree of localization\nwithout changing the ferromagnetic layer thickness. For\nthis purpose we deposited thin layers of normal (non-\nmagnetic) Ag on top of a thin ferromagnetic layer in\nthe vicinity of RC. The normal metal is not expected\nto increase the amount of magnetic material but it can\nreduce electron localization and screen out electronic in-\nteractions. This is similar to an experiment performed\non ultrathin layers of superconductors in which the crit-\nical temperature is suppressed by disorder [21, 22]. Such\nfilms exhibit an ”inverse proximity effect”; Addition of\nultrathin normal metal layers acts to increase TCrather\nthan reduce it as expected from the proximity effect.\nIn our ferromagnetic films we find a similar effect. For\nthick films, adding any amount of an Ag overlayer re-\nduces the EHE signal. This is the case for Ni films for\nwhichdC∼1.8 nm. Indeed one expects that electronic\nhybridization with the Ag electrons should suppress fer-\nromagnetism. On the other hand, for 1 nm thick Co\nfilms, the first few Ag overlayers have an opposite effect.\nAs can be seen in fig. 4, adding ultrathin normal metal\nlayers plays the same role as adding ferromagnetic layers\nin that it increases the EHE signal. Thick enough Ag\noverlayers result in a change of trend and suppression of\nthe EHE. This is seen in the inset of fig 4 where we com-\nparethe outcomeofaddingultrathinlayersofAgto2nm\nthick and 1 nm thick ferromagnetic films. The results of\nthe 1nm Co film demonstrate that reducing the disorder\nwithout changing the ferromagnet thickness results in an\nincrease in the magnetization despite the fact that there\nis a competing effect of electron hybridization.\nIn summary, our results show that no ferromagnetism\ncan be detected by Hall effect or magnetoresistance mea-\nsurements for films having resistances larger than RC.\nWe note that RC, which is material dependent, seems tocorrelate with the material atomic magnetic moment µ,\nas can be seen in table 1. Hence it is possible that the\ncritical point at which magnetization appears in the film\nis related to the fraction of magnetization that is due to\nlocalized moments on the atoms. Clearly, more theoreti-\ncal work is required to clarify the interplay between itin-\nerant ferromagnetism and Anderson localization. Such\nresearch may prove to be very useful in shedding light on\nthe oldpuzzleofthe originofmagnetismin the transition\nmetals.\nWearegratefulforfruitfuldiscussionswithD.Golosov,\nA.M. Goldman and R. Berkovits. This research was sup-\nported by the Israeli academy of science (grant number\n249/05).\n[1] J.M. Valles Jr., R.C. Dynes and J.P. Garno, Phys. Rev.\nLett.69, 3567 (1992).\n[2] A. M.Goldman andN.Markovic, PhysicsToday, Novem-\nber, 39 (1998) and references therin.\n[3] A.M. Finkelstein, Physica B, 197, 636 (1994) and ref.\ntherin.\n[4] M. Strongin, R. Thompson, O. Kammerer and J. Crow,\nPhys. Rev. B1, 1078 (1970).\n[5] E. Abrahams, P. W. Anderson, D. C. Licciardello, and\nT. V. Ramakrishnan, Phys. Rev. Lett. 42, 673 (1979).\n[6] N. Kurzweil, A. Frydman and E. Kogan, in preparation.\n[7] A. Gerber, A. Milner, A. Finkler, M. Karpovski, L. Gold-\nsmith, J. Tuaillon-Combes, O. Boisron, P. M´ elinon, and\nA. Perez, Phys. Rev. B, 69, 224403 (2004).\n[8] Yu. A. Bychkov and E. I. Rashba, J. Phys. C 17, 6093\n(1984); G. Dresselhaus, Phys. Rev. 100, 580 (1955).\n[9] S. Chakravarty and A. Schmid, Phys. Rep. 140, 195\n(1986).\n[10] V. K. Dugaev, P. Bruno, and J. Barna´ s, Phys. Rev. B\n64, 144423 (2001).\n[11] T. R. McGuire and R. I. Potter, IEEE Trans. Magn. 11,\n1018 (1975).\n[12] A. Frydmanand R.C. Dynes, Phil. Mag. 81, 1153 (2001).\n[13] A. Frydman and R.C. Dynes, Sol. State Comm. 110, 485\n(1999).\n[14] A.Frydman,T.L.KirkandR.C.Dynes, SolidStateCom-\nmun.114, 481 (2000).\n[15] P. Mitra, R. Misra, A. F. Hebard, K. A. Muttalib and P.\nWolfle Phys. Rev. Lett., 99, 046804 (2007)\n[16] F. Huang, M. T. Kief, G. J. Mankey, and R. F. Willis,\nPhys. Rev. B 49, 3962 (1994).\n[17] F. Huang, G. J. Mankey, M. T. Kief, and R. F. Willis, J.\nAppl. Phys. 73, 6760 (1993).\n[18] X. Y. Lang, W. T. Zheng, and Q. Jiang, Phys. Rev. B\n73, 224444 (2006).\n[19] R. Zhang and R. F. Willis, Phys. Rev. Lett. 86, 2665\n(2001).\n[20] G. Bergmann, Phys. Rev. Lett. 41, 264 (1978).\n[21] O. Bourgeois, A Frydman and R. C. Dynes, Phys. Rev.\nLett.88, 186403 (2002).\n[22] O. Bourgeois, A Frydman and R. C. Dynes, Phys. Rev.\nB,68, 092509 (2003)." }, { "title": "2208.12799v3.Mechanics_of_a_ferromagnetic_domain_wall.pdf", "content": "Mechanics of a ferromagnetic domain wall\nSe Kwon Kim1and Oleg Tchernyshyov2\n1Department of Physics, Korea Advanced Institute of Science and Technology,\nDaejeon 34141, Republic of Korea\n2William H Miller III Department of Physics and Astronomy and Institute for\nQuantum Matter, Johns Hopkins University, Baltimore, MD 21218, USA\nE-mail:1sekwonkim@kaist.ac.kr,2olegt@jhu.edu\nAbstract. This paper gives a pedagogical introduction to the mechanics of\nferromagnetic solitons. We start with the dynamics of a single spin and develop all\nthe tools required for the description of the dynamics of solitons in a ferromagnet.\nSubmitted to: J. Phys.: Condens. MatterarXiv:2208.12799v3 [cond-mat.mes-hall] 24 Dec 2022CONTENTS 2\nContents\n1 Introduction 3\n2 Single magnetic dipole 3\n2.1 Magnetic moment and angular momentum . . . . . . . . . . . . . . . . 3\n2.2 Spherical geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4\n2.3 Precessional dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . 4\n2.4 Example: Larmor precession . . . . . . . . . . . . . . . . . . . . . . . . 5\n2.5 Conservative and gyroscopic forces . . . . . . . . . . . . . . . . . . . . 5\n2.6 Lagrangian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6\n2.7 Example: precession near a potential minimum . . . . . . . . . . . . . 7\n2.8 Dissipative force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7\n2.9 Example: Larmor precession with dissipation . . . . . . . . . . . . . . 8\n2.10 Transformation to canonical variables . . . . . . . . . . . . . . . . . . 9\n2.11 Transformation to general variables . . . . . . . . . . . . . . . . . . . . 10\n2.12 Rayleigh's dissipation function . . . . . . . . . . . . . . . . . . . . . . 10\n2.13 Example: strong easy-plane anisotropy . . . . . . . . . . . . . . . . . . 11\n2.14 Recap: equations of motion . . . . . . . . . . . . . . . . . . . . . . . . 12\n3 Domain wall in a ferromagnetic wire 13\n3.1 Heisenberg spin chain . . . . . . . . . . . . . . . . . . . . . . . . . . . 13\n3.2 Continuum theory of a ferromagnet . . . . . . . . . . . . . . . . . . . . 15\n3.3 Model with local interactions . . . . . . . . . . . . . . . . . . . . . . . 16\n3.4 Linear spin waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17\n3.5 Nonlinear spin waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17\n3.6 Static domain wall . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18\n3.7 Domain wall in a magnetic \feld . . . . . . . . . . . . . . . . . . . . . . 20\n3.8 Weakly perturbed domain wall . . . . . . . . . . . . . . . . . . . . . . 21\n3.9 Normal modes of a domain wall . . . . . . . . . . . . . . . . . . . . . . 22\n3.10 Angular momentum of a domain wall . . . . . . . . . . . . . . . . . . . 24\n3.11 Linear momentum of a domain wall . . . . . . . . . . . . . . . . . . . 25\n3.12 Counting states of a domain wall . . . . . . . . . . . . . . . . . . . . . 26\n3.13 Hamiltonian dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . 26\n4 Walker's model of a domain wall 27\n4.1 Energy functional and domain-wall con\fguration . . . . . . . . . . . . 28\n4.2 Dynamics of collective coordinates . . . . . . . . . . . . . . . . . . . . 29\n4.3 Steady-state motion in a weak \feld . . . . . . . . . . . . . . . . . . . . 29\n4.4 Walker's breakdown . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30\n4.5 Oscillatory motion in a very strong \feld . . . . . . . . . . . . . . . . . 31\n4.6 D oring's mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32\n5 Discussion 33\nAcknowledgements 33\nAppendix A 33CONTENTS 3\n1. Introduction\nThis article aims to provide a self-contained introduction to theory of ferromagnetic\nsolitons. The subject has been actively researched for more than 50 years and there\nare a number of excellent reviews and books [1, 2]. The last decades have seen new\ndevelopments in the \feld such as the use of collective coordinates to describe the\ndynamics of solitons [3, 4, 5, 6] as well as in technological applications such as soliton-\nbased racetrack memory [7, 8]\nA ferromagnet contains a large number of atomic magnetic dipoles that tend\nto line up with one another at su\u000eciently low temperatures. The interaction most\ncommonly responsible for the parallel alignment is Heisenberg's exchange, which\nrepresents a quantum-mechanical e\u000bect related to the fermionic statistics of electrons.\nOne of the most frequently used models of a ferromagnet is the Heisenberg model\nwith quantum spins of length S=~=2 on a lattice with nearest-neighbor exchange\ninteractions. The ground state has all spins pointing in the same direction. Elementary\nexcitations are magnons|quasiparticles carrying spin Sz=\u0000~along the direction of\nmagnetization. A magnon mode in a coherent state with a large amplitude represents\na classical wave of magnetization (a spin wave) [9].\nThis widespread approach, starting with a quantum model of spins on a lattice,\nis inconvenient for our purposes. Although it is suitable for obtaining weakly excited\nstates such as magnons, the treatment of topological solitons at the quantum level\nand on a lattice is too complicated. Instead, we shall start with continuum models of\na classical ferromagnet. This approach allows one to obtain both linear and nonlinear\nexcitations at the classical level and then quantize them.\nIn Section 2, we discuss the dynamics of a single spin and introduce various\nmathematical tools. The same tools, suitably generalized, are used for the description\nof soliton dynamics in a one-dimensional ferromagnet in Section 3. Sec. 4 deals with\na more complex model of a domain wall of Schryer and Walker [10].\n2. Single magnetic dipole\n2.1. Magnetic moment and angular momentum\nThe building block of a macroscopic magnet is an atom with angular momentum J\nand magnetic dipole moment \u0016related by the gyromagnetic ratio \r,\n\u0016=\rJ: (1)\nIf the angular momentum comes solely from the electron spins, the gyromagnetic ratio\nis\r=e=mec, wheree <0 is the electron charge and meits mass. More generally,\n\r=ge=2mec, where the Lande g-factor is determined by the lengths of spin, orbital,\nand total angular momenta S,L, andJ[11].\nWe shall treat both the angular momentum Jand the magnetic moment \u0016as\nclassical physical variables of \fxed lengths Jand\u0016. Often both are expressed in terms\nof a unit vector mparallel to J:\nJ=Jm;\u0016=\rJm;jmj= 1: (2)\nBecause of the linear proportionality between the three quantities, we will casually\nrefer to mas the magnetic moment or spin.CONTENTS 4\nem\ne✓e\u0000\nFigure 1. The local frame de\fned by the unit vectors (4).\n2.2. Spherical geometry\nTo resolve the constraint of the unit length jmj= 1, it is convenient to introduce two\nspherical angles: the polar angle \u0012(co-latitude) and the azimuthal angle \u001e(longitude):\nm= (sin\u0012cos\u001e;sin\u0012sin\u001e;cos\u0012): (3)\nThey de\fne a local frame with three mutually orthogonal unit vectors\nem=m= (sin\u0012cos\u001e;sin\u0012sin\u001e;cos\u0012);\ne\u0012=@m\n@\u0012= (cos\u0012cos\u001e;cos\u0012sin\u001e;\u0000sin\u0012); (4)\ne\u001e=1\nsin\u0012@m\n@\u001e= (\u0000sin\u001e;cos\u001e;0);\npointing up, south, and east, respectively (Fig. 1).\nSmall deviations of magnetization from a given direction mcan be written in the\nlocal frame as follows:\n\u000em=e\u0012\u000em\u0012+e\u001e\u000em\u001e; \u000em \u0012=\u000e\u0012; \u000em\u001e= sin\u0012\u000e\u001e: (5)\n2.3. Precessional dynamics\nThe dynamics of an atomic magnetic dipole is very di\u000berent from that of a Newtonian\nparticle. For a particle, an external force generates an acceleration proportional to the\nforce and inversely proportional to its mass. In contrast, a magnetic dipole does not\nhave inertia. Like a fast-spinning gyroscope, it precesses with a velocity proportional\nto the external torque.\nIt is helpful to picture a magnetic dipole whose south pole is pivoted and the\nnorth pole is allowed to move on a sphere of radius rso that its position is r=rm.\nThe force from the potential energy may then be calculated as F=\u0000dU\ndr=\u00001\nrdU\ndm\nand its torque as \u001c=r\u0002F=\u0000m\u0002dU\ndm. The torque determines the rate of change\nof angular momentum dJ=dt=Jdm=dt, hence the equation of motion\nJ_m=\u0000m\u0002dU\ndm; (6)CONTENTS 5\nwhere _mstands fordm=dt. Note that the geometrical size of the dipole rcanceled\nout.\nIt is worth keeping in mind that the derivative dU=dmis only taken in the\ndirections locally tangential to the sphere jmj= 1. In a formal sense, this derivative\nis de\fned as the coe\u000ecient in the Taylor expansion\nU(m+dm) =U(m) +dU\ndm\u0001dm+:::; (7)\nwheredmis an in\fnitesimal displacement tangential to the sphere and thus transverse\ntom. Even ifdU=dmhad a component parallel to m, it would not contribute to the\n\frst order in the expansion (7). For this reason, dU=dmshould be understood as the\ntransverse component of the energy derivative.\n2.4. Example: Larmor precession\nConsider the motion of a magnetic dipole in an external magnetic \feld h. Its potential\nenergy isU=\u0000\u0016\u0001h=\u0000\u0016m\u0001h. Upon substituting Uinto the equation of motion\n(6) we obtain\nJ_m=\u0000\u0016h\u0002m: (8)\nThe magnetic moment rotates uniformly about the direction of the \feld at the Larmor\nprecession frequency\n!L=\u0016h=J =\rh: (9)\nThe angle between the magnetic moment and the \feld remains constant, implying\nconservation of energy U=\u0000\u0016\u0001h.\nTo see this explicitly, rewrite the equation of motion in terms of spherical angles\n\u0012and\u001e(3):\n\u0000Jsin\u0012_\u001e\u0000@U\n@\u0012= 0; J _\u0012\u00001\nsin\u0012@U\n@\u001e= 0: (10)\nFor the \feld along the zaxis, h= (0;0;h), the potential energy U=\u0000\u0016hcos\u0012is\nindependent of the azimuthal angle \u001e. Thus the polar angle \u0012remains unchanged\nduring precession.\n2.5. Conservative and gyroscopic forces\nConservation of energy is a general property of the precession equation (6). After\nsome simple algebra, it can be transformed into the form\n\u0000J_m\u0002m\u0000dU\ndm= 0: (11)\nReturning to the picture of a dipole pivoted at the south pole, we may interpret this\nequation of motion as Newton's second law for a massless particle con\fned to move\non a unit sphere (hence no inertial term proportional to m). The second in Eq. (11)\nobviously represents a conservative force.\nWhat does the \frst term represent? It is similar to the Lorentz force acting on\na moving electric charge in a magnetic \feld as it is proportional to the velocity _m\nand is orthogonal to it. We thus may interpret the \frst term in Eq. (11) as a Lorentz\nforce acting on a particle with unit electric charge moving on the unit sphere jmj= 1\nin a radial \\magnetic \feld\" b(m) =\u0000Jm, as if there were a magnetic monopole ofCONTENTS 6\nstrengthJat the sphere's center. (We use the scare quotes to distinguish the \fctitious\n\\magnetic \feld\" bon the unit sphere jmj= 1 from the physical magnetic \feld h.)\nThe \\Lorentz force\" in Eq. (11) does no work and thus cannot be described in\nterms of potential energy. It belongs to a class of gyroscopic forces \frst discussed by\nKelvin [12]. Precession of a spin is indeed reminiscent of the motion of a fast-spinning\ngyroscope.\nEq. (11) can be used to establish conservation of potential energy during motion:\ndU\ndt=_m\u0001dU\ndm=_m\u0001(\u0000J_m\u0002m) = 0: (12)\n2.6. Lagrangian\nThe equation of motion (11) can be obtained by minimization of action S=R\nLdt.\nThe Lagrangian\nL=a(m)\u0001_m\u0000U(m) (13)\nhas no kinetic energy [9]. The \frst term gives rise to the gyroscopic (\\Lorentz\") force\nand contains a gauge potential a(m) whose curl gives the \\magnetic \feld\" of the\nmonopole,\nrm\u0002a(m) =b(m) =\u0000Jm: (14)\nThe \frst term in the Lagrangian (13) is peculiar for a number of reasons.\n(i) Its action\nSg=Z\na(m)\u0001_mdt=Z\na(m)\u0001dm (15)\nis independent of how fast the vector mmoves on the unit sphere and depends\njust on the geometry of its path. Hence the name geometric action forSgand\ngeometric (or Berry) phase for Sg=~[13].\n(ii) This term is not uniquely de\fned. A gauge transformation\na(m)7!a(m) +rm\u001f(m); (16)\nwhere\u001f(m) is an arbitrary scalar function on the unit sphere, changes the vector\npotential a(m) but leaves the \\magnetic \feld\" b(m) the same.\n(iii) Strictly speaking, the magnetic \feld of a monopole b(m) cannot be represented\nby a gauge potential because this \feld con\fguration is not divergence-free: it\nhas a source with magnetic charge \u0000Jat the origin m= 0 [9, 14]. Although\nthe magnetic monopole is not accessible, we can still observe that there is a net\n\\magnetic \rux\"\u00004\u0019Jby integrating the \\\feld\" over the unit sphere.\n(iv) As a result of this problem, any choice of the gauge potential a(m) contains a\nsingularity. For example, the gauge potential [14]\na(m) =Jms\u0002m\n1\u0000ms\u0001m; (17)\nhas a singularity at m=ms. It is associated with a localized \\magnetic \rux\"\nof strength +4 \u0019J, which compensates the uniform \rux \u00004\u0019Jspread over the\nsphere. If we were allowed to explore the entire 3-dimensional space of m, we\nwould discover that there is a Dirac string carrying \\magnetic \rux\" +4 \u0019Jfrom\nthe origin to in\fnity in the direction ms. With the Dirac string attached, the\n\\magnetic \feld\" becomes solenoidal, which makes it possible to describe it in\nterms of a vector potential.CONTENTS 7\n(v) Vector potential (17) describes the \\magnetic \feld\" of a monopole b(m) =\u0000Jm\nalmost everywhere on the unit sphere m, with the exception of point ms, where\nthe \\magnetic \rux\" of the Dirac string pierces the sphere. As long as the\ntrajectory m(t) stays away from the Dirac string, we can use the gauge potential.\nTo complete this section, we write down the gauge term in spherical angles for\nthe azimuthally symmetric gauge choices with the Dirac string attached to the north\nand south poles, ms= (0;0;\u00061):\na(m)\u0001_m=J(cos\u0012\u00061)_\u001e: (18)\nThese are the two most commonly used gauge choices [9].\n2.7. Example: precession near a potential minimum\nChoose the axes so that the energy has a minimum at the north pole, m0= (0;0;1).\nIn the vicinity of the energy minimum,\nm= (mx; my;q\n1\u0000m2x\u0000m2y); (19)\nwe may expand the potential energy in powers of the transverse deviations mx\u001c1\nandmy\u001c1. To the second order,\nU(m) =U(m0) +Kxm2\nx\n2+Kym2\ny\n2+::: (20)\nHereKxandKyare positive constants and the omitted terms are of higher orders in\nthe transverse components of magnetization. Choose the gauge potential (17) with the\nsingularity at the south pole, ms= (0;0;\u00001), so that a\u0019(Jmy=2;\u0000Jmx=2;0) +:::\nThe resulting Lagrangian\nL\u0019\u0002\nJ( _mxmy\u0000_mymx)\u0000Kxm2\nx\u0000Kym2\ny\u0003\n=2 +::: (21)\nyields the equations of motion for mxandmy:\n\u0000J_my\u0000Kxmx= 0; J _mx\u0000Kymy= 0: (22)\nThe spin follows an elliptical trajectory Kxm2\nx+Kym2\ny= const precessing at the\nfrequency!=\u0000p\nKxKy=J.\n2.8. Dissipative force\nIn the real world, a magnetic dipole will not precess about the magnetic \feld forever\nand will eventually line up with it, settling into a minimum of potential energy. Energy\ndissipation can be caused by various processes, e.g., radiation of electromagnetic waves\nor friction. We shall add a simple phenomenological term to the equations of motion\nthat represents|in the analogy with a particle on a sphere|a dissipative force \u0000\u000bJ_m\nproportional to the velocity _mand directed against it. Eq. (11) now reads\n\u0000J_m\u0002m\u0000dU\ndm\u0000\u000bJ_m= 0: (23)\nIt expresses the balance of gyroscopic, conservative, and dissipative forces acting on\nthe dipole. The dimensionless constant \u000b>0 introduced by Gilbert [15] characterizes\nthe strength of energy dissipation. It is usually small, \u000b\u001c1, with typical values\nranging from 10\u00004in insulating magnets to 10\u00002in metallic ones.CONTENTS 8\nFigure 2. Larmor precession of a spin in an external magnetic \feld in the\npresence of dissipation. The energy minimum is at the north pole (black point).\nThe spin trajectory is shown for 1 =(2\u000b) Larmor periods with the starting point\non the equator. The Gilbert damping \u000b= 0:05; the gyromagnetic ratio \r >0.\nThe equation of motion, expressed in spherical angles, reads\n\u0000Jsin\u0012_\u001e\u0000@U\n@\u0012\u0000\u000bJ_\u0012= 0;\nJ_\u0012\u00001\nsin\u0012@U\n@\u001e\u0000\u000bJsin\u0012_\u001e= 0: (24)\nIn the presence of the dissipative force, potential energy is dissipated at the rate\ndU\ndt=_m\u0001dU\ndm=_m\u0001(\u0000J_m\u0002m\u0000\u000bJ_m) =\u0000\u000bJj_mj2\u00140: (25)\n2.9. Example: Larmor precession with dissipation\nWe return to the problem of a dipole in a magnetic \feld, this time with dissipation.\nWith energy U=\u0000\u0016hcos\u0012, Eq. (24) reads\n\u0000Jsin\u0012_\u001e\u0000\u0016hsin\u0012\u0000\u000bJ_\u0012= 0;\nJ_\u0012\u0000\u000bJsin\u0012_\u001e= 0: (26)\nThe introduction of dissipation leads to a slight slowdown of the Larmor precession\nand to a gradual approach to the energy minimum:\n\u001e(t) =\u001e(0)\u0000~!Lt;~!L=!L\n1 +\u000b2;\ncos\u0012(t) = tanh (\u000b~!L(t\u0000t0)): (27)\nHere!L=\rhis the bare Larmor frequency (9) and t0is the time when the spin lies\nin the equatorial plane, cos \u0012= 0. We see that the azimuthal motion is precessional\nand fast, with the frequency of order \u000b0, whereas the polar motion is relaxational and\nslow, with the relaxation rate of order \u000b. Fig. 2 shows the trajectory of a spin for\n1=(2\u000b) Larmor periods starting at the equator and approaching the energy minimum\nat the north pole.CONTENTS 9\n2.10. Transformation to canonical variables\nAlthough our physical quantities|magnetization and angular momentum|are\nrepresented by a three-dimensional vector m, its three components are not\nindependent because of the \fxed length jmj= 1. It may sometimes be convenient to\nwork with two independent variables, which can be introduced in a number of ways.\nOne possible route is to use a pair of canonical variables ( q;p), one of which is a\ncoordinate and the other its conjugate momentum. To obtain them, we may use the\nLagrangian in one of the standard gauges (18),\nL=J(cos\u0012\u00061)_\u001e\u0000U(\u0012;\u001e): (28)\nWe choose the azimuthal angle \u001eas the coordinate and obtain the corresponding\ncanonical momentum in the usual way:\nq=\u001e; p =@L\n@_\u001e=J(cos\u0012\u00061) =Jz\u0006J: (29)\nAs one might expect, the momentum conjugate to \u001eis (up to an additive constant)\nthe angular momentum Jz=Jcos\u0012. It is somewhat alarming that the polar angle \u0012\ndoes not have a conjugate momentum (no _\u0012term in the Lagrangian). We will address\nthis problem in a later section. For now, we shall dismiss this concern by noting that\n\u0012is present in the canonical momentum so that both angles are represented in the\ncanonical pair ( q;p).\nWe now realize that our two-dimensional space is a phase space . The number of\nstates in a phase space is proportional to phase-space volume:\nd\u0000 =dqdp\n2\u0019~=J\n2\u0019~sin\u0012d\u0012d\u001e: (30)\nIntegrating over the sphere, we obtain the total number of states\n\u0000 =J\n2\u0019~Z\u0019\n0sin\u0012d\u0012Z2\u0019\n0d\u001e=2J\n~= 2j: (31)\nwherej=J=~is the length of angular momentum in the units of ~. The exact answer,\nwhich we have learned in a quantum mechanics course, is \u0000 = 2 j+1. Our approximate\nanswer, obtained within the scope of classical mechanics, is correct in the limit of a\nlarge length of angular momentum, j\u001d1.\nWith canonical variables identi\fed, we have the Poisson bracket. In particular,\nf\u0012;\u001eg=\u0000f\u001e;\u0012g=@\u0012\n@q@\u001e\n@p\u0000@\u0012\n@p@\u001e\n@q=1\nJsin\u0012: (32)\nThe Hamiltonian is obtained in a standard way,\nH=p_q\u0000L=U(\u0012;\u001e): (33)\nIt contains potential energy only. The _\u001eterm in the Lagrangian (28) is notkinetic\nenergy. Hamilton's equations of motion,\n_q=fq;Hg=@H=@p; _p=fp;Hg=\u0000@H=@q; (34)\nreproduce the equations of motion for the spherical angles (10).CONTENTS 10\n2.11. Transformation to general variables\nWe may not always want to use pairs of canonically conjugate variables and instead\nperform a more general transformation from mto a pair of coordinates ( q1;q2), e.g.,\nq1=\u0012andq2=\u001e. Equations of motion for the new variables can be obtained by\nstarting from the equation (23) expressing force balance.\nTo obtain the equations of motion for the new variables fqig, wherei= 1;2, we\n\frst note that the time evolution of m(q1;q2) comes solely through the time evolution\nof the coordinates: _m= _qj@m=@qj, where summation is implied over the doubly\nrepeated index j. We substitute this expression into Eq. (23) and multiply the result\nby@m=@qito obtain an equation of motion for the new coordinates:\nGij_qj\u0000@U=@qi\u0000Dij_qj= 0: (35)\nIt expresses the balance of the gyroscopic, conservative, and dissipative forces for\neach coordinate qi. The coe\u000ecients GijandDijof the antisymmetric gyroscopic and\nsymmetric dissipative tensors are de\fned as follows:\nGij=\u0000Jm\u0001\u0012@m\n@qi\u0002@m\n@qj\u0013\n; Dij=\u000bJ@m\n@qi\u0001@m\n@qj: (36)\nThe gyroscopic tensor Gijis the inverse of the of Poisson tensor \u0005ij=fqi;qjg, i.e.,\nGij\u0005jk=\u000ek\ni[16]. It can be obtained as the curl of a gauge potential Ai:\nGij=@Aj\n@qi\u0000@Ai\n@qj; Ai=a(m)\u0001@m\n@qi: (37)\nThe Lagrangian, expressed in terms of the new coordinates, reads\nL=Ai(q) _qi\u0000U(q): (38)\nThe \frst term of the Lagrangian gives rise to the geometrical action\nSg=Z\nAi_qidt=Z\nAidqi: (39)\nThe number of states in an in\fnitesimal rectangle with sides ( dq1;dq2) is\nd\u0000 =1\n2\u0019~jG12jdq1dq2=p\ndetG2Y\ni=1dqi\np\n2\u0019~: (40)\nFor the choice of coordinates q1=\u0012andq2=\u001e, the nonzero components of the\ntensors are\nG\u0012\u001e=\u0000G\u001e\u0012=\u0000Jsin\u0012; D\u0012\u0012=\u000bJ; D\u001e\u001e=\u000bJsin2\u0012: (41)\nUpon substituting these tensor components into Eq. (35), we obtain the equations of\nmotion for the spherical angles (24).\n2.12. Rayleigh's dissipation function\nDissipative forces cannot be described as a term in a Lagrangian. It is an emergent\nforce, arising from interaction of a macroscopic object with numerous microscopic\ndegrees of freedom. In thermodynamics, it plays a major role in the relaxation of the\nphysical system toward thermal equilibrium. In that context, the dissipative force can\nbe obtained from the Rayleigh dissipation function Rquadratic in velocities [17],\nR=1\n2\u000bJj_mj2=1\n2Dij_qi_qj: (42)CONTENTS 11\nThe modi\fed Euler-Lagrange equations are\n@L\n@qi\u0000d\ndt@L\n@_qi\u0000@R\n@_qi= 0: (43)\nThe last term is the dissipative force.\n2.13. Example: strong easy-plane anisotropy\nLet us discuss the dynamics of a spin whose potential energy\nU(\u0012;\u001e) =\u0000K\u0012\n2sin2\u0012+V(\u0012;\u001e) (44)\nis dominated by the easy-plane anisotropy term with K\u0012>0. Whatever the starting\nstate, after a period of initial relaxation the spin will \fnd itself near the equatorial\ncircle\u0012=\u0019=2. Its long-term dynamics will be in\ruenced by the azimuthal potential\nlandscape de\fned by V(\u0019=2;\u001e).\nTo get a \frst look at the long-term dynamics, consider the motion in the vicinity\nof the potential energy minimum at \u0012=\u0019=2 and some \u001e=\u001e0, where the potential\nenergy is approximately quadratic in the deviations \u000e\u0012and\u000e\u001efrom the minimum:\nU(\u0019=2 +\u000e\u0012;\u001e 0+\u000e\u001e)\u0019K\u0012\u000e\u00122\n2+K\u001e\u000e\u001e2\n2; (45)\nwhereK\u001e\u001cK\u0012. Equations of motion (24), linearized in the deviations from\nequilibrium, read\n\u0000J\u000e_\u001e\u0000K\u0012\u000e\u0012\u0000\u000bJ\u000e _\u0012= 0; J\u000e _\u0012\u0000K\u001e\u000e\u001e\u0000\u000bJ\u000e_\u001e= 0: (46)\nIn the absence of dissipation, \u000b= 0, we obtain elliptical precession near\nthe equatorial plane with the frequency !=p\nK\u0012K\u001e=Jand the amplitude ratio\n\u000e\u00120=\u000e\u001e0=p\nK\u001e=K\u0012\u001c1. Turning on dissipation at \frst makes the precession\nunderdamped, with the relaxation rate \u0000 \u0019\u000bK\u0012=2Jin the limit \u000b\u001c2p\nK\u001e=K\u0012.\nWhen the damping constant exceeds the critical value \u000bc= 2p\nK\u0012K\u001e=(K\u0012\u0000K\u001e)\u001c1,\nthe spin dynamics becomes overdamped, with two distinct relaxation rates, \u0000 1\u0019\n\u000bK\u0012=Jand \u0000 2\u0019K\u001e=\u000bJ\u001c\u00001in the limit 2p\nK\u001e=K\u0012\u001c\u000b\u001c1.\nOur simple analysis of the \fnal approach to equilibrium demonstrates that, even\nwhen dissipation is nominally weak, \u000b\u001c1, spin dynamics does not necessarily\nretain its precessional character and can be purely relaxational. An apparently\nsmall damping constant \u000b\u001c1 may represent relatively strong dissipation when it\nis compared to another small parameter, the ratio of azimuthal and polar anisotropiesp\nK\u001e=K\u0012. Put another way, increasing the strength of the easy-plane anisotropy K\u0012\nat a \fxed\u000b\u001c1 will eventually bring us into a regime where the spin motion separates\ninto two independent components, fast relaxation, during which the spin approaches\nthe equatorial plane, and slow relaxation within the equatorial plane. (The fast mode\ninclude azimuthal as well as polar motion.)\nNext we shall derive the long-term dynamics of spin motion in a more general\nsetting, when the polar angle is already near equilibrium, \u0012=\u0019=2 +\u000e\u0012with a small\n\u000e\u001e, but\u001emay not be. The Lagrangian can then be approximated as follows:\nL\u0019\u0000J\u000e\u0012 _\u001e\u0000K\u0012\u000e\u00122=2\u0000V(\u0019=2;\u001e): (47)\nIn the Rayleigh dissipation function,\nR\u0019\u000bJ(\u000e_\u00122+_\u001e2)=2\u0019\u000bJ_\u001e2=2; (48)CONTENTS 12\nwe have neglected the contribution of polar motion to damping because it is restricted\nby the strong easy-plane anisotropy, so \u000e_\u0012\u001c_\u001e. This is an important general point,\nso let us emphasize it: dissipation is dominated by soft modes; dissipation from hard\nmodes can often be neglected .\nFrom the Lagrangian (47) and Rayleigh function (48) we obtain the equations of\nmotion:\n\u0000J_\u001e\u0000K\u0012\u000e\u0012= 0; J\u000e _\u0012\u0000@V=@\u001e\u0000\u000bJ_\u001e= 0: (49)\nThe \frst equation lets us express the polar angle (a hard mode of no interest to us)\nin terms of the azimuthal angle (the soft mode we are interested in). Then the second\nequation will be for the azimuthal angle alone:\nI\u001e\u0000\u000bJ_\u001e\u0000@V=@\u001e = 0; (50)\nwhere we introduced a \\moment of inertia\" I=J2=K\u0012.\nBy integrating out the hard mode \u000e\u0012, we have generated inertia (mass) for the\nsoft mode\u001e. This is a su\u000eciently frequent occurrence in ferromagnetic dynamics to\nmerit a proper name, the D oring mass [18]. The Lagrangian of the azimuthal angle\nalone (with the polar angle integrated out) acquires kinetic energy:\nL(\u001e) =I_\u001e2=2\u0000V(\u0019=2;\u001e): (51)\nThis elegant result hinges on neglect of dissipation for the hard mode \u000e\u0012. Had we\nnot neglected it, the equation relating \u0012to\u001ewould be more involved|non-local in\ntime|and this simple picture would not have emerged. The non-locality in time is\nnegligible on time scales longer than \u001c1= 1=\u00001.\nReturning to the equation of motion for the azimuthal angle (50), we see that it\ndescribes a massive rotator subject to an external potential and dissipative friction.\nOn long time scales, we may neglect inertia and obtain a simpler equation of dissipative\ndynamics,\n\u0000\u000bJ_\u001e\u0000@V=@\u001e = 0: (52)\nThe azimuthal velocity _\u001eis proportional to the potential torque \u0000@V=@\u001e . Doing\nso is permissible on time scales longer than the characteristic acceleration time\n\u001c1=I=\u000bJ =J=\u000bK\u0012= 1=\u00001.\nWe should check that our long-term dynamics reproduces the previously obtained\napproach to equilibrium, when \u001e=\u001e0+\u000e\u001e. For small \u000e\u001e, the equatorial torque\n\u0000@V=@\u001e\u0019\u0000k\u000e\u001eand we recover exponential relaxation of \u000e\u001eat the rate \u0000 2=K\u001e=\u000bJ,\nin agreement with our preliminary analysis.\nOn a \fnal note, we observe that the long-term dynamics of the soft mode \u001egiven\nby Eq. (52) can be obtained directly from the Lagrangian (47) and Rayleigh dissipation\nfunction (48) by merely ignoring the hard mode \u000e\u0012. This is a general principle that we\nwill apply in many contexts in what follows: the response of a system to weak external\nperturbations is dominated by its soft modes .\n2.14. Recap: equations of motion\nWe end the discussion of a single magnetic dipole by gathering in one place di\u000berent\nforms of its equations of motion, all of which are equivalent to one another.\n\u000fThis form of the equation expresses the balance of the gyroscopic, conservative,\nand dissipative forces acting on a magnetic dipole:\n\u0000J_m\u0002m\u0000dU\ndm\u0000\u000bJ_m= 0: (53)CONTENTS 13\n\u000fThis equation expresses the rate of change of angular momentum in terms of\nconservative and dissipative torques acting on the dipole:\nJ_m=\u0000m\u0002dU\ndm\u0000\u000bJm\u0002_m: (54)\n\u000fThe previous equation is often written in terms of an \\e\u000bective \feld\" he\u000b=\n\u0000dU=d\u0016as follows:\n_m=\u0000\rhe\u000b\u0002m\u0000\u000bm\u0002_m;he\u000b=\u00001\n\u0016dU\ndm: (55)\nThis form of the equation of motion is known as the Landau-Lifshitz-Gilbert\n(LLG) equation.\n\u000fThe LLG equation (55) contains two terms with a time derivative _m, which\nis inconvenient when the equation is integrated numerically. Through simple\nalgebra, it can be converted into the form used originally by Landau and Lifshitz,\n(1 +\u000b2)_m=\u0000\rhe\u000b\u0002m\u0000\u000bj\rjhe\u000b: (56)\nThis form of the equation contains a single term with the time derivative _m\nand is used in numerical micromagnetic solvers such as OOMMF and MuMax.\nOne cautionary remark is on order. The e\u000bective \feld he\u000bde\fned in Eq. (55)\nis understood as a derivative dU(m)=dmwith respect to transverse variations of\nm, see Eq. (7). A sloppy evaluation may produce a longitudinal component of\nhe\u000bparallel to mand create a problem for the Landau-Lifshitz equation (56),\nchanging the length of the unit vector m. Just to be on the safe side, one may\nreplace he\u000bin the second term of Eq. (56) with its transverse part m\u0002(he\u000b\u0002m).\nThe same applies to Eq. (53).\n\u000fWritten in terms of spherical angles, the equations of motion are\n\u0000Jsin\u0012_\u001e\u0000@U\n@\u0012\u0000\u000bJ_\u0012= 0;\nJ_\u0012\u00001\nsin\u0012@U\n@\u001e\u0000\u000bJsin\u0012_\u001e= 0: (57)\n\u000fFor arbitrary coordinates q1andq2parametrizing the unit vector m, the equations\nof motion for these coordinates again expresses the balance of the gyroscopic,\nconservative, and dissipative forces associated with each coordinate:\nGij_qj\u0000@U=@qi\u0000Dij_qj= 0: (58)\nThe antisymmetric gyroscopic and symmetric dissipative tensors are\nGij=\u0000Jm\u0001\u0012@m\n@qi\u0002@m\n@qj\u0013\n; Dij=\u000bJ@m\n@qi\u0001@m\n@qj: (59)\n3. Domain wall in a ferromagnetic wire\n3.1. Heisenberg spin chain\nTheory of a single magnetic dipole can be readily extended to that of many dipoles.\nA convenient way to accomplish that is to start with the Lagrangian of a single spin\n(13) and to generalize it to many spins fmng:\nL=X\nna(mn)\u0001_mn\u0000U(fmng); (60)CONTENTS 14\nwhere once again the vector potential a(m) describes the \feld of a magnetic monopole,\nrm\u0002a(m) =\u0000Jm, andU(fmng) is the potential energy of the system. The Rayleigh\nfunction is generalized in the same way,\nR=1\n2X\nn\u000bJj_mnj2: (61)\nThen the Euler-Lagrange equations of motion for every spin n,\n@L\n@mn\u0000d\ndt@L\n@_mn\u0000@R\n@_mn= 0; (62)\nyields an obvious generalization of single-spin dynamics (54),\nJ_mn=\u0000mn\u0002@U\n@mn\u0000\u000bJmn\u0002_mn; (63)\nor its variants, Eqs. (53) and (55). The e\u000bective \feld acting on spin mnishe\u000b\nn=\n\u0000\u0016\u00001@U=@ mn.\nHere we examine a Heisenberg chain with magnetic dipoles arranged periodically\nwith lattice constant a. Adjacent dipoles mnandmn+1interact with energy\n\u0000Amn\u0001mn+1, whereAis the exchange constant:\nU=\u0000AX\nnmn\u0001mn+1: (64)\nThe equation of motion for dipole mnis\nJ_mn=Amn\u0002(mn\u00001+mn+1)\u0000\u000bJmn\u0002_mn: (65)\nAn equilibrium state, _mn= 0, is reached when all dipoles line up along the same\ndirection, mn=m. Such a state clearly minimizes the potential energy (64) for A>0.\nThe common direction of magnetization mis arbitrary.\nIt is instructive to investigate the dynamics of magnetization close to equilibrium,\nmn=m+\u000emn, where in\fnitesimal deviations from equilibrium \u000emnare transverse\ntomin order to preserve the unit length of mn.\nIt is convenient to introduce a reference frame with three unit vectors ex,ey, and\nezforming a right triple, ei\u0002ej=\u000fijkek. We align one of them with the equilibrium\ndirection, ez=m. Then\u000emn=mn;xex+\u000emn;yey. After expanding Eq. (65) to the\n\frst order in \u000emn, we obtain the dynamical equation for the deviations:\nJ_mn;x =\u0000A(mn\u00001;y\u00002mn;y+mn+1;y) +\u000bJ_mn;y;\n\u0000J_mn;y=\u0000A(mn\u00001;x\u00002mn;x+mn+1;x) +\u000bJ_mn;x: (66)\nThe two transverse components can be conveniently combined into a complex number\n n=mn;x+imn;y. These complex variables have the following equation of motion:\ni(J+i\u000bJ)_ n=\u0000A( n\u00001\u00002 n+ n+1): (67)\nSolutions of this equation have the form of waves \u000emn(t) =Ceikna\u0000i!t, whereais the\nlattice period and kis the wavenumber; the frequency is\n!=2A(1\u0000coska)\nJ+i\u000bJ: (68)\nTwo things are worth noting, First, the frequency of a spin wave !vanishes when\nthe wavenumber k!0. This is a consequence of spontaneous breaking of a continuous\nsymmetry, in this case of global spin rotations, by the ground state with spontaneousCONTENTS 15\nmagnetization m. Second, in the presence of dissipation ( \u000b > 0), the frequency has\nan imaginary part,\nIm!=\u00002\u000bA(1\u0000coska)\nJ(1 +\u000b2): (69)\nBecause Im !<0, a spin wave n(t) =Ceikna\u0000i!tdecays exponentially in time.\nThe mathematics can be simpli\fed for waves whose wavelength is long compared\nto the lattice spacing, ka\u001c1. For such waves, magnetizations of adjacent dipoles\nare almost the same, mn+1\u0019mn. In this limit, we may treat discrete variables mn\nas a slowly varying and continuous function m(z), wherez=nais the coordinate\nalong the chain. Then \fnite di\u000berences can be transformed into spatial derivatives via\nTaylor expansion:\nmn+1\u0000mn=m(z+a)\u0000m(z)\u0019@m(z)\n@za;\nmn+1\u00002mn+mn\u00001\u0019@2m(z)\n@z2a2; (70)\nand so on.\nIn the continuum approximation, the equation of motion for transverse\n\ructuations (z;t) =mx(z;t) +imy(z;t) (67) reads\niJ@ \n@t=\u0000Aa2@2 \n@z2: (71)\nFor simplicity, we switched o\u000b dissipation, \u000b!0. The resulting wave equation\nresembles the Schr odinger equation for a free particle with a nonrelativistic energy-\nmomentum relation E=p2=2mand the mass m=J~=(2a2A). These particles were\n\frst introduced by Felix Bloch [19] and are now known as magnons, the quanta of\nspin waves.\n3.2. Continuum theory of a ferromagnet\nTo develop a continuum description of a ferromagnetic medium, on a formal level,\nwe promote the unit vector mencoding the magnetic moment of a single dipole to\na unit-vector \feld m(r) varying smoothly in space r. The mathematical apparatus\ndeveloped in the previous section is generalized accordingly.\nFor example, the equation of motion (53) undergoes two very minor changes:\n\u0000J_m\u0002m\u0000\u000eU\n\u000em\u0000\u000bJ_m= 0: (72)\nFirst, the \\length\" of angular momentum Jhas been replaced with the corresponding\nintensive quantity, the angular momentum Jper unit volume. Second, the potential\nenergy function U(m) has become the potential energy functional U[m(r)] and thus\nthe ordinary derivative dU=dm(7) has turned into the functional derivative \u000eU=\u000em(r).\nThe Landau-Lifshitz equations (55) and (56) retain their forms, albeit with an\ne\u000bective \feld\nhe\u000b(r) =\u00001\nM\u000eU\n\u000em(r)(73)\ncontaining the functional derivative.CONTENTS 16\nEquations of motion for the spherical-angle \felds \u0012(r) and\u001e(r) are\n\u0000Jsin\u0012_\u001e\u0000\u000eU\n\u000e\u0012\u0000\u000bJ_\u0012= 0;\nJ_\u0012\u00001\nsin\u0012\u000eU\n\u000e\u001e\u0000\u000bJsin\u0012_\u001e= 0: (74)\nA \feld m(r) has in\fnitely many degrees of freedom, so if we attempt to describe\nit in terms of some coordinates qi, the number of these coordinates will be in\fnite.\nNonetheless, the equations of motion for these coordinates remain unchanged,\nGij_qj\u0000@U=@qi\u0000Dij_qj= 0; (75)\nand only the gyroscopic and dissipative tensors are promoted to functionals of the\nmagnetization \feld:\nGij=\u0000JZ\ndVm\u0001\u0012@m\n@qi\u0002@m\n@qj\u0013\n; Dij=\u000bJZ\ndV@m\n@qi\u0001@m\n@qj:(76)\nThe gyroscopic tensor can be obtained from a gauge potential Ai:\nGij=@Aj\n@qi\u0000@Ai\n@qj; Ai=Z\ndVa(m)\u0001@m\n@qi: (77)\nThis is entirely analogous to Eq. (37) for a single spin. Equations of motion (75) can\nbe obtained from the Lagrangian\nL=Ai(q) _qi\u0000U(q) (78)\nsupplemented by the Rayleigh dissipation function [17]\nR=1\n2Dij_qi_qj: (79)\nOur model is a ferromagnet in one spatial dimension with coordinate z. It\ncan describe a ferromagnetic wire whose transverse dimensions are smaller than the\ncharacteristic length \u00150de\fned in Eq. (82). Exchange interaction tends to align spins,\nso a ground state should have a uniform magnetization \feld m(z) = const. However,\nsome of these uniform states will have lower energies than others. In particular, long-\nrange dipolar interactions prefer that the dipoles line up along the direction of the\nwire, leaving just two ground states,\nm(z) =\u0006ez= (0;0;\u00061): (80)\n3.3. Model with local interactions\nWe will use a simple phenomenenological form of the potential energy functional\nminimized by these two ground states:\nU[m(z)] =Z\ndz\u0000\nAjm0j2+Kjm\u0002ezj2\u0001\n=2: (81)\nHere m0\u0011dm=dz. The \frst term models the exchange interaction and penalizes\nspatially non-uniform states. It can be derived as a continuum approximation to the\nHeisenberg chain. The second term penalizes states with magnetization deviating\nfrom the direction of the wire ez.\nBoth the exchange constant Aand anisotropy strength Kare positive. Together\nwith the density of angular momentum J, they de\fne the characteristic scales of\nlength\u00150, time\u001c0, and energy \u000f0:\n\u00150=p\nA=K; \u001c 0=J=K; \u000f 0=p\nAK: (82)CONTENTS 17\nWe stress that this model is not very realistic as it relies on a local anisotropy term\nto select the two ground states (80). A more realistic model with dipolar interactions\nwould require a non-local energy functional as spins would interact over large distances.\nThe advantage of our toy model is in its relative simplicity: with not too much e\u000bort,\nwe can study its linear and nonlinear excitations.\n3.4. Linear spin waves\nTo obtain spin waves near one of the uniform ground states, m0=ez, we consider a\nweakly excited state with m(z) =m0+\u000em(z), where\u000em=exmx+eymyrepresents\nsmall transverse deviations of magnetization. Expand the Lagrangian density to the\nsecond order in \u000emas in Sec. 2.7:\nL=1\n2Z\ndz\u0010\n_mxmy\u0000_mymx\u0000m0\nx2\u0000m0\ny2\u0000m2\nx\u0000m2\ny\u0011\n+::: (83)\nHere we have switched to natural units (82), in which J=A=K= 1. The\nLagrangian (83) is quadratic in the transverse spin components, The resulting\nequations are linear in them, hence the term linear spin waves .\nIn the absence of dissipation, \u000b= 0, the equations of motion for the transverse\nmagnetization components are\n\u0000_my+m00\nx\u0000mx= 0;_mx+m00\ny\u0000my= 0: (84)\nIt is convenient to combine the two transverse components into a complex \feld\n =mx+imy, whose equation of motion,\ni_ =\u0000 00+ ; (85)\nresembles the Schr odinger equation i_ =H with the Hamiltonian\nH=\u0000d2\ndz2+ 1; (86)\nwhose energy spectrum is \u000fk= 1+k2. A traveling spin wave (t;z) = (0;0)e\u0000i!t+ikz\nhas the frequency\n!=K+Ak2\nJ; (87)\nwhere we have restored the units. The frequency spectrum has a gap with a minimal\nexcitation frequency lim k!0!=K=J= 1=\u001c0.\n3.5. Nonlinear spin waves\nTo describe spin waves with a large amplitude, we introduce an Ansatz\nm(z;t) = (sin \u0002 cos ( kz+ \b);sin \u0002 sin (kz+ \b);cos \u0002): (88)\nThe polar angle \u0002 between mand the easy axis ezis the wave's amplitude; the\nazimuthal \b is its phase, Fig. 3. We will treat the amplitude \u0002( t) and the phase \b( t)\nas two collective coordinates of the nonlinear wave (88). Eq. (75) determines their\ndynamics.\nThe energy of the wave (88) scales linearly with the wire length and diverges in an\nin\fnite wire. It is therefore convenient to assume a \fnite wire length `and to imposeCONTENTS 18\nzxy\nFigure 3. Snapshot of a large-amplitude spin wave. The spins have a \fxed polar\nangle \u0002. The azimuthal angle varies in space as \u001e=kz+ \b.\nperiodic boundary conditions to eliminate edge e\u000bects. The wire length `must then\nbe commensurate with the wavelength \u0015= 2\u0019=k. The energy of the wave (88) is then\nU(\u0002;\b) =1\n2(Ak2+K)`sin2\u0002: (89)\nThe lack of a \b dependence is related to a symmetry, in this case the symmetry of\nglobal spin rotations about the zaxis. We say that \b is a zero mode .\nThe gyroscopic and dissipative tensors are obtained with the aid of Eq. (76).\nTheir nonvanishing coe\u000ecients are\nG\u0002\b=\u0000G\b\u0002=\u0000J`sin \u0002; D \u0002\u0002=\u000bJ`; D \b\b=\u000bJ`sin2\u0002:(90)\nThe following equations of motion result for the collective coordinates:\n\u0002 :\u0000J`sin \u0002 _\b\u00001\n2(K+Ak2)`sin 2\u0002\u0000\u000bJ`_\u0002 = 0;\n\b : +J`sin \u0002 _\u0002\u0000\u000bJ`sin2\u0002_\b = 0: (91)\nThe equation for \b shows that \u0002 is the slower variable of the two as _\u0002 =\u000bsin \u0002 _\b\nand\u000b\u001c1. It would be conserved in the absence of dissipation ( \u000b= 0). After some\nsimpli\fcation, we \fnd that the spin rotation frequency depends on the amplitude \u0002:\n_\b =\u0000K+Ak2\nJ(1 +\u000b2)cos \u0002: (92)\nEq. (92) reproduces the spectrum of linear spin waves (87) in the limit of a small\namplitude (\u0002!0) and weak dissipation ( \u000b!0).\n3.6. Static domain wall\nThe uniform ground states m(z) =\u0006ezare global minima of the potential energy\n(81). There are also local minima, in which magnetization interpolates between \u0000ez\non one end of the wire and + ezon the other. In such a state, magnetization m(z)\nforms two (nearly) uniform domains, in which the energy density is (nearly) zero. The\nregion in between has non-uniform magnetization deviating from the easy axis has a\npositive energy density and is known as a domain wall.\nTo obtain a formal solution for a domain wall, it is convenient to write the\npotential energy (81) in terms of spherical angles,\nU[\u0012(z);\u001e(z)] =Z\ndzh\nA(\u001202+ sin2\u0012\u001e02) +Ksin2\u0012i\n=2: (93)CONTENTS 19\nzxyσ=+1σ=−1\nFigure 4. Domain walls in a ferromagnetic wire with Z2topological charges\n\u001b=\u00061.\nMinimization of the energy \u000eU= 0 yields \feld equations\n\u0000A\u001200+1\n2(A\u001e02+K) sin 2\u0012= 0; A (sin2\u0012\u001e0)0= 0: (94)\nSolutions with a uniform azimuthal angle, \u001e0= 0 automatically satisfy the second of\nthese equations. The \frst one can be integrated once to obtain \u0000A\u001202+Ksin2\u0012=C.\nThe constant of integration can be obtained by noting that at z!\u00061 we recover\nuniform states with \u00120= 0 and sin \u0012= 0. Therefore C= 0. Integrating this equation\nyields domain-wall solutions (Fig. 4),\ncos\u0012(z) =\u001btanhz\u0000Z\n\u00150; \u001e(z) = \b: (95)\nSolutions (95) have three parameters, one discrete and two continuous.\n\u000fTopological charge \u001b=\u00061 determines the ground states on the two sides of\nthe domain wall: m=\u0000\u001bezatz=\u00001 andm=\u001bezatz= +1. The\nword \\topological\" indicates that this characteristic cannot be changed by small\ncontinuous deformations of the soliton.\n\u000fPositionZdetermines the location of the domain wall. The change from\nm=\u0000\u001bezto\u001bezhappens around z=Z, in a region of width \u00150(82).\n\u000fAngle \b de\fnes an azimuthal plane in which the spins interpolate between\nm=\u0000\u001bezand\u001bez.\nNote that the domain-wall solution (95) contains two parameters, Zand\n\b, de\fning the position of the domain wall and its azimuthal plane. These\ncollective coordinates of a domain wall quantify its rigid translations and rotations,\ntransformations that leave the energy (81) invariant and thus known as zero modes .\nIt is helpful to picture the con\fguration space\n\u00001

γ1are of type II [10, 12]. The quantum\nphasetransitionsat P0candPchaveremarkableproperties, asshownin[11]. Depending\non the system properties, TC– the temperature locating point C, can be either positive\n(when a direct N-FS first order transition is possible), zero, or neg ative (when the\nFM-FS and N-FM phase transition lines terminate at different zero-t emperature phase\ntransition points). The last two cases correspond to Ts<0. All these cases are possible\nin spin-triplet ferromagnetic superconductors. The zero temper ature transition at Pc0\nis found to be a quantum critical point, whereas the zero-tempera ture phase transition\natPcis of first order. As noted, the latter splits into two first order pha se transitions.\nThis classical picture may be changed through quantum fluctuation s [10].\nThe quantum phase transitions at P0candPchave remarkable properties. An investi-\ngation [11] performed by renormalization group methods revealed a fluctuation change\nin the order of the zero temperature first order phase transition atPcto a continuous\nphase transition belonging to anentirely new class of universality. Ho wever, thisoption\n4exists only for magnetically isotropic order (Heisenberg symmetry) and is unlikely to\napply in the known spin-triplet ferromagnetic superconductors, w hich are magnetically\nanisotropic.\nThe application of the theory was demonstrated on the example of U Ge2, but the\nsame theory has ample volume of options to describe various real su perconductors of\nthe same type and is not restricted to a particular compound, or, t o a particular group\nof such materials.\nEven in its simplified form, this theory has been shown to be capable of accounting\nfor a wide variety of experimental behavior. A natural extension t o the theory is\nto add aM6term which provides a formalism to investigate possible metamagnetic\nphase transitions and extend some first order phase transition line s, as required by\nexperimental data. Another modification of this theory, with rega rd to applications to\nother compounds, is to include a Pdependence for some of the other GL parameters.\nAmong the outstanding problems are: local gauge effects on the vo rtex phase and the\nphasetransitions, theoutlineoftheuppercriticalmagneticfield Hc2(T,P),thermaland\nmagnetic properties, and the description of T−Pdiagrams with topologies, which are\nobserved in experiments. Such studies require an extension of the theory by including\nM6term, the magnetic induction B=H+4πM, and a more precise P-dependence\nof some theory parameters.\nAcknowledgements: The author is grateful to A. Harada and S. M. Hayden for\nvaluable discussions of experiments.\nReferences\n[1] D. Vollhardt and P. W¨ olfle (1990) The Superfluid Phases of Helium 3 , Taylor &\nFrancis, London.\n[2] G. E. Volovik and L. P. Gor’kov (1985) Sov. Phys. JETP 61843 [Zh. Eksp. Teor.\nFiz.88(1985) 1412].\n[3] E. J. Blagoeva, G. Busiello, L. De Cesare, Y. T. Millev, I. Rabuffo, a nd D. I.\nUzunov (1990) Phys. Rev. B426124.\n[4] D. I. Uzunov (1990)in Advances in Theoretical Physics , ed. by E. Caianiello; World\nScientific, Singapore, p. 96.\n[5] D. I. Uzunov (2010) Theory of Critical Phenomena , World Scientific, Singapore\n(Second edition); First edition (1993).\n[6] S. S. Saxena, P. Agarwal, K. Ahilan, F. M. Grosche, R. K. W. Hase lwimmer, M.J.\nSteiner, E. Pugh, I. R. Walker, S.R. Julian, P. Monthoux, G. G. Lonz arich, A.\nHuxley. I. Sheikin, D. Braithwaite, and J. Flouquet (2000) Nature406587.\n[7] D. Aoki, A. Huxley, E. Ressouche, D. Braithwaite, J. Flouquet, J -P.. Brison, E.\nLhotel, and C. Paulsen (2001) Nature413613.\n5[8] N. T. Huy, A. Gasparini, D. E. de Nijs, Y. Huang, J. C. P. Klaasse, T. Gortenmul-\nder, A. de Visser, A. Hamann, T. G¨ orlach, and H. v. L¨ ohneysen( 2007) Phys. Rev.\nLett.99067006 .\n[9] T. Akazawa, H.Hidaka, H.Kotegawa, T. C.Kobayashi, T. Fujiwa ra, E. Yamamoto,\nY. Haga, R. Settai, and Y. Onuki (2005) Physica B 359-361 1138.\n[10] D. V. Shopova and D. I. Uzunov (2003) Phys. Lett. A 313139; D. V. Shopova\nand D. I. Uzunov (2005) Phys. Rev. B 72024531; D. V. Shopova and D. I. Uzunov\n(2009)Phys. Rev. B 79064501.\n[11] D. I. Uzunov (2006) Phys. Rev. B 74134514; D. I. Uzunov (2007) Europhys. Lett.\n7720008.\n[12] D. I. Uzunov (2012) in Superconductors , ed. by A. Gabovich; Intech, Rijeka, Ch.\n17, pp 415-440.\n6" }, { "title": "1502.00118v1.Room_temperature_local_ferromagnetism_and_nanoscale_domain_growth_in_the_ferromagnetic_semiconductor_GeFe.pdf", "content": " 1 \nRoom -temperature local ferromagnetism and nanoscale domain growth in the \nferromagnetic semiconductor Ge 1 – xFex \n \nYuki K. Wakabayashi,1 Shoya Sakamoto,2 Yukiharu Takeda,3 Keisuke Ishigami,2 Yukio \nTakahashi ,2 Yuji Saitoh,3 Hiroshi Yamagami,3 Atsus hi Fujimori,2 Masaaki Tanaka ,1 and \nShinobu Ohya1 \n1Department of Electrical Engineering and Information Systems, The University of \nTokyo, 7 -3-1 Hongo, Bunkyo -ku, Tokyo 113 -8656, Japan \n2Department of Physics, The University of Tokyo, Bunkyo -ku, Tokyo 113 -0033, Japan \n3Synchrotron Radiation Research Unit, JAEA, Sayo, Hyogo 679 -5148, Japan \n \nAbstract \nWe investigate the local electronic structure and magnetic properties of the group -IV-\nbased ferromagnetic semiconductor, Ge 1 – xFex (GeFe), using soft X -ray magnetic circular \ndichroism. Our results show that the doped Fe 3 d electrons are strongly hybridize d with \nthe Ge 4 p states, and have an unusually large orbital magnetic moment relative to the spin \nmagnetic moment; i.e., morb/mspin ≈ 0.3. We f ind that local ferromagnetic domains , which \nare formed through ferromagnetic exchange interactions in the high -Fe-content regions \nof the GeFe films , exist at room temperature, well above the Curie temperature of 20 – \n100 K. We demonstrate the first observation of the intriguing nanoscale domain growth \nprocess in which ferromagnetic domains expand as the temperature d ecreases, followed \nby a transition of the entire film into a ferromagnetic state at the Curie temperature .\n 2 \nA major issue that must be addressed for the realiz ation of semiconductor spintronic \ndevices is to achieve room -temperature ferromagnetism in ferromagnetic semiconductors \n(FMS s) based on the widely used III -V and group -IV materials. In Ga1 – xMn xAs \n(GaMnAs) , which is a particularly well -studied FMS, the highest Curie temperature ( TC) \never reported is 200 K [1]. In GaMnAs, TC is limited by the presence of interstitial Mn \natoms, which are antiferromagnetically coupled to the substitutional Mn atoms [2]. \nRecently, however, the group -IV-based FMS, Ge 1 – xFex (GeFe), has been reported to \nexhibit several attractive features [3 – 5]. It can be grown epitaxially on Si and Ge \nsubstrates without the formation of intermetallic precipitates, and is therefore compatible \nwith mature Si process technology. Unlike GaMnAs, with GeFe, interstitial Fe atoms do \nnot lead to a decrease in TC [6], and TC can be easily increased to above 200 K by thermal \nannealing [7]. Furthermore, TC does not depend on the carrier concentration, and thus TC \nand resistivity can be controlled separately [8], which is a unique feature that is only \nobserved with G eFe and is a considerable advantage in overcoming the conductivity \nmismatch problem between ferromagnetic metals and semiconductors in spin -injection \ndevices. Despite these attractive features, a detailed microscopic understanding of the \nferromagnetism in GeFe , which is vitally important for room -temperature applications , is \nlacking. Here, we investigate the local magnetic behavior of GeFe using X -ray magnetic \ncircular dichroism (XMCD), which is a powerful technique for element -specific detection \nof magnetic moments [9 – 12]. We find that local ferromagnetic domains remain in the \nGeFe films even at room temperature, i.e., well ab ove TC; it follows that GeFe potentially \nhas strong ferromagnetism, which persist s even at room temperature. Furthermore, we \npresent the first observation s of the intriguing feature that ferromagnetic domains, which \nare formed above TC via the ferromagneti c exchange interaction in high -Fe concentration \nregions of the films, develop and grow as the temperature decreases, and that all of them \ncoalesce at temperatures below TC. Such a nanoscale domain growth process is a key \nfeature in understanding materials that exhibit homogeneous ferromagnetism (i.e., where \nthe film is free from any ferromagnetic precipitates) despite the inhomogeneous \ndistribution of magnetic atoms in the film [6,7]. \nWe carried out XMCD measurements on two samples (labeled A and B) consist ing \nof a 2-nm-thick Ge capping layer, and a 120-nm-thick Ge0.935Fe0.065 layer with a 30-nm-\nthick Ge buffer layer grown on a Ge(001) substrate by low-temperature molecular beam \nepitaxy (LT -MBE) [Figs. 1(a) and 1(b) ]. The Ge buffer and Ge cap layer s were grown at 3 \n200°C, and the Ge0.935Fe0.065 layer of sample A was grown at 160°C, whereas that of \nsample B was grown at 240°C [6]. From the Arrott plots of the magnetic field (H) \ndependence of the magnetic circular dichroism (MCD) measured with visible li ght with \na photon energy of 2.3 eV (corresponding to the L-point energy gap of bulk Ge), we found \nTC = 20 K and 100 K for samples A and B, respectively. Detailed crystallographic \nanalyses showed that the GeFe films are single crystalline, with a diamond -type structure \nand nanoscale spatial Fe concentration fluctuations of 4% – 7% (sample A) and 3% – \n10% (sample B) [6]. We found that TC is higher when the fluctuations in the Fe \nconcentration are larger. In addition, channeling Rutherford backscattering and \nchanneling particle -induced X -ray emission measurements showed that ~85% (~15% ) of \nthe doped Fe atoms exist at the substitutional (tetrahedral interstitial) sites in both samples \nA and B, and that the interstitial Fe concentration is not related to TC [6]. This indicates \nthat there are no ferromagnetic precipitates with different crystal structures in our films. \nWe performed X -ray absorption spectroscopy (XAS) and XMCD measurements at the \ntwin-helical undulator beamline BL23SU of SPring -8 [13]. The XAS spect ra were \nobtained in total electron yield mode. To remove the oxidized surface layer, the samples \nwere briefly etched in dilute hydrofluoric acid ( HF) prior to loading into the XAS \n(XMCD) vacuum chamber. \nWe measured XAS spectra [ μ+, μ–, and (μ+ + μ–)/2] at the L2 (~721 eV) and L3 (~708 \neV) absorption edges of Fe in sample A [Fig. 1(c)] and B [Fig. 1(d)] at 5.6 K with μ0H = \n5 T applied perpendicular to the film surface. Here, μ+ and μ– refer to the absorption \ncoefficients for photon helicity parallel and antiparallel to the Fe 3 d majority spin \ndirection, respectively. In both films, the three peaks a, b, and c are observed at the Fe L3 \nedge in the XAS spectra [ see also the insets in Figs. 1(c) and 1(d)]. We found that the \nsmall peak c was suppressed by etching the surface with dilute HF, indicating that this \npeak, which can be assigned to the Fe3+ state, originates from a small quantity of surface \nFe oxide [14], which remains even after surface cleaning. Meanwhile, p eaks a and b are \nassigned to the Fe atoms in GeFe [15,16]. \nWe measured the XMCD ( = μ+ – μ–) spectra at the Fe L2 and L3 absorption edges of \nsamples A [Fig. 1(e)] and B [Fig. 1(f)] at 5.6 K with various H applied perpendicular to \nthe film surface . Here, we discuss the XMCD intensit ies at 707.66 eV (X) and 708.2 eV \n(Y), which correspond to the photon energies of peaks a and b in the XAS spectra, \nrespectively. When normalized to 707.3 eV , the XMCD spectra with various H differ, and 4 \nthe intensity at X gr ows faster than that at Y as H increase s, as shown in the insets of Figs. \n1(e) and 1(f). As shown in F igs. 1(c) and (d), the shapes of the XAS spectra at the Fe L3 \nedge are similar between samples A and B, which have almost the same interstitial Fe \nconcentrations (i.e., 15% of the total Fe content [6]) ; therefore, we can assign the XMCD \nintensity at X to the substitutional Fe atoms and the paramagnetic component of the \nXMCD intensity at Y to the interstitial Fe atoms. We do not observe fine structure s due \nto multiplet splitting at the Fe L3 edge in both samples, which would be observed if the \n3d electrons were localized and were not strongly hybridized with other orbitals [17]. \nThese observations indicate that the Fe 3 d electrons are strongly hybridize d with the Ge \n4p states [18]. \nWe determine the orbital magnetic moment, morb, and the spin magnetic moment, \nmspin, of the substitutional Fe atoms from the XAS and XMCD spectra at the L2,3 edge \nregion of Fe using the XMCD s um rules [19 – 23] [see Section I of Supplemental Material \n(SM)]. As shown in Figs. 2(a) and 2(b), both mspin and morb (and therefore the total \nmagnetic moment M = mspin + morb) are larger in sample B ( TC = 100 K) than in sample A \n(TC = 20 K) over the entire temperature region when 𝜇0H = 5 T. For sample A, morb/mspin \n= 0.31 ± 0.02, and for sample B , morb/mspin = 0.30 ± 0.03, both of which are positive and \nsignificantly larger than that of bulk Fe (where morb/mspin ~ 0.043 [19]); the o rbital angular \nmomentum in GeFe is not quenched. The observation that the spin and orbital angular \nmomentum are parallel suggests that the Fe 3 d shell is more than half filled. This implies \nthat the Fe atoms are in the 2+ state rather than in the 3+ state, in which the Fe 3 d shell is \nhalf-filled and the orbital angular momentum vanishes. This large morb is a characteristic \nproperty of GeFe, and excludes the possibility of the existence of ferromagnetic Fe metal \nprecipitates in our films. \nFigure 2(c) shows the H dependence of the XMCD intensity at energy X and a \ntemperature of 5.6 K (blue curve), the MCD intensity measured with visible light of 2.3 \neV at 5 K (red dotted curve), and the magnetization measured using a superconducting \nquantum interference device (SQUID) at 5 K (green dotted curve) for sample B. The \nshapes of these curves show excellent agreement with each other; it follows that the spin \nsplitting of the valence band composed of the Ge 4 p orbitals is induced by the Fe 3 d \nmagnetic moment, wh ich originates from the substitutional Fe atoms , through the p-d \nhybridization. The lower panels of Fig. 2 show the effective magnetic -field (Heff) \ndependence of the XMCD intensity measured at X for samples A (d) and B (e) at various 5 \ntemperatures. Here, M is also plotted (filled red symbols), and μ0Heff is obtained by \nsubtracting the product of M and the sheet density of the substitutional Fe atoms from \nμ0H to eliminate the influence of the demagnetization field. The scale of the vertical axis \nof the XMCD intensity is adjusted so that it represents M at each temperature. The insets \nshow clear hysteresis below TC in both samples. The XMCD – Heff curves show large \ncurvature above TC in both samples [see the main panel s of Figs. 2(d) and 2(e)], indicating \nthat part of the film is superparamagnetic (SPM) above TC. It indicates that ferromagnetic \ndomains form in nanoscale high -Fe concentration regions at temperatures above TC, and \nthus M can be described by \n𝑀=5.2𝑓SPML(𝑚SPM𝜇0𝐻eff\n𝑘B𝑇) + (1 -𝑓SPM)𝐶\n𝑇𝜇0𝐻eff, (1) \nwhere fSPM and mSPM are fitting parameters expressing the fraction of substitutional Fe \natoms which participat e in the SPM component , and the magnetic moment per \nferromagnetic domain, respectively. Also, C is the Curie constant per substitutional Fe \natom (see Section III of the SM), and L is the Langevin function . Here, 5.2 is the ideal \nsaturated value of M; i.e., M = mspin + (morb/mspin) × mspin, where mspin = 4 μB (for Fe2+) \nand morb/mspin ≈ 0.3 [Figs. 2(a) and 2(b)] when all the substitutional Fe atoms are \nmagnetically active. The first and second terms in Eq. (1) correspond to the SPM and \nparamagnetic components, respectively. In Figs. 2(d) and 2(e), the black dashed curves \ncorrespond to the best fit obtained with Eq. (1). For sample B, the M – Heff curves at \ntemperatures in the range 100 – 300 K are well reproduced by Eq. (1), which indicates \nthat the ferromagnetic – SPM transition occurs at TC = 100 K. With sample A, the M – \nHeff curves at temperatures above TC (i.e., T > 20 K) are well reproduced by Eq. (1), \nexcept fo r T = 20 K, which is probably due to the onset of ferromagnetism. These good \nfits up to room temperature indicate that ferromagnetic interactions within the nanoscale \nhigh-Fe concentration region s remain at room temperature in both samples. \nThe residual M, which is obtained from a linear extrapolation of M from the high \nmagnetic field region to Heff = 0 at 5.6 K, is 1.2 μB per Fe atom in sample A, and 1.5 μB \nper Fe atom in sample B. This result suggests that only ~23% (= 1.2/5.2) and ~29% (= \n1.5/5.2) of the substitutional Fe atoms are magnetically active in samples A and B, \nrespectively. In Figs. 2(d) and 2(e), t he high -field magnetic susceptibilities ∂𝑀/\n𝜕(𝜇0𝐻eff) (μB/T per Fe atom) at 4 T and 5.6 K are 0.15 in sa mple A and 0.10 in sample \nB. Because ∂𝑀/𝜕(𝜇0𝐻eff) at 4 T per substitutional parama gnetic Fe atom should be 0.37 6 \n(see Section IV of the SM), this result indicates that the ratios of paramagnetic Fe atoms \nto the total number of Fe atoms are only ~41% (= 0.15/0.37) and ~27% (= 0.10/0.37) , \nrespectively. This means that some fraction of the moment of the Fe atoms is missing, \nand thus suggests that there are Fe atoms th at couple antiferromagnetically with the \nferromagnetic Fe atoms in the films. This is also supported by the weak spin -glass \nbehavior observed in GeFe at very low temperatures [7]. \nWe see a similar trend in the temperature dependence of the fitting paramete rs fSPM \nand mSPM in both films; i.e., fSPM and mSPM both increase with decreasing temperature (Fig. \n3). This result means that the ferromagnetic domains, which form only in the nanoscale \nhigh-Fe concentration regions at room temperature [Fig. 4(a)] , expand toward lower Fe \nconcentration regions with decreasing temperature [Fig. 4(b)], and finally the entire film \nbecomes ferromagnetic at TC [Fig. 4(c)]. This appears to be a characteristic feature of \nmaterials that exhibit homogeneous ferromagnetism, d espite the inhomogeneous \ndistribution of magnetic atoms in the film [6,7]. As shown in Fig. 3, fSPM and mSPM are \nlarger in sample B than in sample A, which is attributed to the difference in spatial \nfluctuations of the Fe concentration, which are 4% – 7% i n sample A and 3% – 10% in \nsample B [6]. The larger the nonuniformity of the Fe distribution is , the larger \nferromagnetic domains, fSPM , and mSPM become, and the domains can more easily connect \nmagnetically , resulting in a higher TC. \nIn summary, we have in vestigated the local electronic structure and magnetic \nproperties of the doped Fe atoms in the Ge0.935Fe0.065 films using XAS and XMCD. The \nFe atoms appear in the 2+ state, with the 3 d electrons strongly hybridized with the 4 p \nelectrons in Ge; this results in a delocalized 3 d nature and long -range ferromagnetic \nordering, leading to the excellent agreement between the H dependence of magnetization, \nMCD, and XMCD. Using the XMCD sum rules, we obtained the M – Heff curves, which \ncan be explained by the coexistence of SPM and paramagnetic ordering at temperatures \nabove TC. The fitting results clearly show that the local ferromagnetic domains , which \nexist at room temperature, expand with decreasing temperature, leadin g to a \nferromagnetic transition of the entire system at TC. The nonuniformity of the Fe \nconcentration plays a crucial role for the formation of the magnetic domains, and our \nresults indicate that strong ferromagnetism is inherent to GeFe, and persists at r oom \ntemperature. \n 7 \nACKNOWLEDGEMENTS \nWe would like to thank T. Okane for support with the experiments. This work was partly \nsupported by Grants -in-Aid for Scientific Research (22224005, 23000010, and \n26249039) including the Specially Promote d Research , and the Project for Developing \nInnovation Systems from MEXT. This work was performed under the Shared Use \nProgram of JAEA Facilities (Proposal No. 201 4A-E31) with the approval of the \nNanotechnology Platform Project supported by MEXT . The synchr otron radiation \nexpe riments were performed at the JAEA beamline BL23SU in SPring -8 (Proposal No. \n2014A3881 ). Y. K. Wakabayashi and Y. Takahashi acknowledge financial support from \nJSPS through the Program for Leading Graduate Schools (MERIT). S. Sakamoto \nacknowledge s financial support from JSPS through the Program for Leading Graduate \nSchools (ALPS). 8 \nReferences \n[1] L. Chen , X. Yang , F. Yang , J. Zhao , J. Misuraca , P. Xiong , and S. von Molnár , Nano \nLett. 11, 2584 (2011). \n[2] K. M. Yu, W. Walukiewicz, T. Wojtowicz, I. Kuryliszyn, X. Liu, Y . Sasaki, and J. K. \nFurdyna, Phys. Rev. B 65, 201303 (2002). \n[3] Y . Shuto, M. Tanaka, and S. Sugahara, phys. stat. sol. 3, 4110 (2006). \n[4] Y . Shuto, M. Tanaka, and S, Sugahara, Appl. Phys. Lett. 90, 132512 (2007). \n[5] Y . Shuto, M. Tanaka, and S. Sugahara, Jpn. J. Appl. Phys. 47, 7108 (2008). \n[6] Y . K . Wakabayashi, S. Ohya, Y . Ban, and M. Tanaka, J. Appl. Phys. 116, 173906 \n(2014). \n[7] Y . K. Wakabayashi, Y . Ban, S. Ohya and M. Tanaka, Phys. Rev. B 90, 205209 (2014). \n[8] Y . Ban, Y . Wakabayashi, R. Akiyama, R. Nakane, and M. Tanaka, AIP Advances 4, \n097108 (2014). \n[9] D. J. Keavney, D. Wu, J. W. Freeland, E. J. Halperin, D. D. Awschalom, and J. Shi, \nPhys. Rev. Lett. 91, 187203 (2003). \n[10] K. W. Edmonds, N. R. S. Farley, T. K. Johal, G. V . D. Laan, R. P. Campiom, B. L. \nGallegher, and C. T. Foxon, Phys. Rev. B 71, 064418 (2005). \n[11] D. J. Keavney, S. H. Cheung, S. T. King, M. Weinert, and L. Li, Phys. Rev. Lett. 95, \n257201 (2005). \n[12] V . R. Singh, K. Ishigami, V . K. Verma, G. Shibata, Y . Yamazaki, T. Kataoka, A. \nFujimori, F. -H. Chang, D. -J. Lin, C. T. Che n, Y . Yamada, T. Fukumura, and M. \nKawasaki, Appl. Phys. Lett. 100, 242404 (2012). \n[13] Y . Saitoh, Y . Fukuda, Y . Takeda, H. Yamagami, S. Takahashi, Y . Asano, T. Hara, K. \nShirasawa, M. Takeuchi, T. Tanaka, and H. Kitamura, J. Synchrotron Rad. 19, 388 \n(2012) \n[14] T. J. Regan, H. Ohldag, C. Stamm, F. Nolting, J. Luning, J. Stohr, and R. L. White, \nPhys. Rev. B 64, 214422 (2001). \n[15] R. Kumar, A. P. Singh, P. Thakur, K. H. Chae, W. K. Choi, B. Angadi, S. D. Kaushik, \nand S. Panaik, J. Phys. D: Appl. Phys. 41, 155 002 (2008). \n[16] E. Sakai, K. Amemiya, A. Chikamatsu, Y . Hirose, T. Shimada, and T. Hasegawa, J. \nMagn. Magn. Mater, 333, 130 (2013). \n[17] G. V . D. Laan and I. W. Kirkman, J. Phys.: Condens. Matter 4, 4189 (1992). \n[18] I. A. Kowalik, A. Persson, M. A. Nino, A. Na, A. Navarro -Quezada, B. Faina, A. \nBonanni, T. Dietl, and D. Arvanits, Phys. Rev. B 85, 184411 (2012). \n[19] C. T. Chen, Y . U. Idzerda, H. -J. Lin, N. V . Smith, G. Meigs, E. Chaban, G. H. Ho, \nE. Pellegrin, and F. Sette, Phys. Rev. Lett. 75, 152 (1995). \n[20] Y . Takeda, M. Kobayashi, T. Okane, T. Ohkochi, J. Okamoto, Y . Saitoh, K. \nKobayashi, H. Yamagami, A. Fujimori, A. Tanaka, J. Okabayashi, M. Oshima, S. \nOhya, P. N. Hai, and M. Tanaka, Phys. Rev. Lett. 100, 247202 (2008). \n[21] K. Mamiya , T. Koide, A. Fujimori, H. Tokano, H. Manaka, A. Tanaka, H. Toyosaki, \nT. Fukumura, and M. Kawasaki, Appl. Phys. Lett. 89, 062506 (2006). \n[22] J. Stohr and H. Konig, Phys. Rev. Lett. 75, 3748 (1995). \n[23] C. Piamonteze, P. Miedema, and F. M. F. de Groot, P hys. Rev. B 80, 184410 (2009). 9 \nFig. 1. (Color) Schematic diagrams showing the structures of sample A (a) and sample B \n(b). (c), (d) XAS spectra of μ– (blue curve), μ+ (red curve), and (μ+ + μ–)/2 (black curve) \nat the L2 (~721 eV) and L3 (~708 eV) absorption edges of Fe for sample A (c) and sample \nB (d). The measurements were made with a magnetic field of μ0H = 5 T applied \nperpendicular to the film surface at a temperature of 5.6 K. The insets show a magnified \nplot of the spectra at the Fe L3 edge. (e), (f) XMCD (= μ+ – μ–) spectra at the L2 and L3 \nabsorption edges of Fe for sample A (e) and sample B (f) measured at 5.6 K with μ0H = \n0.1 T (red curve), 1 T ( brown curve), 3 T (green curve), and 5 T (blue curve) applied \nperpendicular to the film surface . The insets show a magnified plot of the spectra at the \nFe L3 edge, in which the XMCD data are normalized to 707.3 eV . \n \n 10 \nFig. 2. (Color) (a ), (b) The temperature dependence of mspin + morb (green squares), mspin \n(blue triangles), morb (violet circles), and morb/mspin (red rhombuses) for sample A (a) and \nsample B (b) with an applied magnetic field of μ0H = 5 T. (c) The H dependence of the \nXMCD intensity (blue solid curve) at X shown in Fig. 1 (707.66 eV) at 5.6 K, the MCD \nintensity at 5 K with a photon energy of 2.3 eV corresponds to the L-point energy gap of \nbulk Ge (red dotted curve), and the magnetization measured using a SQUID at 5 K (green \ndotted curve) for sample B. (d),(e) The dependence of the XMCD intensity measured at \nX on the effective magnetic field Heff for sample A (d) and sample B (e) at various \ntemperatures. The total magnetization (M = mspin + morb) obtained using the XMCD sum \nrules is also plotted (filled red symbols ). We scaled the vertical axis of the XMCD \nintensity so that it represents M at each temperature. In all measurements, H was applied \nperpendicular to the film surface . \n \n \n \n \n \n 11 \nFig. 3. (Color online ) (a), (b) The temperature dependence of the best -fit parameters fSPM \nand mSPM obtained for sample A (a) and sample B (b). The red (or dark gray) , gray, and \nwhite areas indicate ferromagnetic (FM), FM + SPM + paramagnetic (PM), and SPM + \nPM regions, respectively. \n \n \n \n \n \n \n \n \n \n \n \n 12 \n \nFig. 4 . (Color) (a ) – (c) Schematic diagrams showing the magnetic states in the \nGe0.935Fe0.065 films with zero magnetic field at room temperature (i.e., T = 300 K ) (a), TC \n< T < 300 K (b), and T < TC (c). The small black , red, and blue arrows correspond to the \nmagnetic moments of the paramagnetic , ferromagnetic , and antiferromagnetically \ncoupled substitutional Fe atoms , respectively. The red areas indicate ferromagnetic \nregions. Antiferromagnetically coupled Fe atoms are thought to exist all over the film at \ntemperatures below TC. \n 13 \nSupplemental Material for Room -temperature local ferromagnetism and nanoscale \ndomain growth in the ferromagnetic semiconductor Ge 1 - xFex \n \nYuki K. Wakabayashi,1 Shoya Sakamoto,2 Yukiharu Takeda,3 Keisuke Ishigami,2 Yukio \nTakahashi,2 Yuji Saitoh,3 Hiroshi Yamagami,3 Atsushi Fujimori,2 Masaaki Tanaka,1 and \nShinobu Ohya1 \n \n1Department of Electrical Engineering and Information Systems, The University of Tokyo, \n7-3-1 Hongo, Bunkyo -ku, Tokyo 113 -8656, Japan \n2Department of Physics, The University of Tokyo, Bunkyo -ku, Tokyo 113 -0033, Japan \n3Synchrotron Radiation Research Unit, JAEA, Sayo, Hyogo 679 -5148, Japan \n \nI. Estimation of the spi n magnetic moment and the orbital magnetic moment of \nsubstitutional Fe atoms using the X -ray m agnetic -circular -dichroism sum rules \n \nWe obtain the spin magnetic moment mspin and the orbital magnetic moment morb \nfrom the spectra of the X -ray absorption spectroscopy (XAS) and X -ray magnetic circular \ndichroism (XMCD) in the energy region near the L2 and L3 absorption edges of Fe using \nthe XMCD sum rules.S3 Figure S1(a) shows the XAS spectra (solid curves) and the \nXAS signals integrated from 690 eV (dashed curves) of sample A. Figure S1(b) shows \nthe XMCD spectra (solid curves) and the XMCD signals integrated from 690 eV (dashed \ncurves) of sample A. Here, the measurements were carried out with a magnetic field \n𝜇0H = 5 T applied perpendicular to the film surface at 5.6 K (black curves), 20 K (blue \ncurves), 50 K (lig ht blue curves), 100 K (green curves), 150 K (orange curves), 250 K \n(pink curves), and 300 K (red curves). Figure S2 shows the same data measured for \nsample B. For the XMCD sum -rules analyses, we define th e values of r, p, and q as the \nfollowing equations at each temperature. \n𝑟= ∫(𝜇++𝜇−)\n2𝑑𝐸𝐸3+𝐸2, (S1) \n𝑝= ∫(𝜇+−𝜇−)𝑑𝐸𝐸3, (S2) \n𝑞= ∫ (𝜇+−𝜇−)𝑑𝐸𝐸3+𝐸2, (S3) \nwhere E3 (690-718 eV) and E2 (718-760 eV) represent the integration energy range s for \nthe L3 and L2 absorption edges, respectively. Here, μ+ (μ-) and E represent the absorption 14 \ncoefficient for the photon helicity parallel (antiparallel) to the Fe 3 d majority spin \ndirection and the i ncident photon energy, respectively. We can obtain mspin and morb of \nsubstitutional Fe atoms using the XMCD sum rules, which are expressed as follows: \n𝑚orb= −2𝑞\n3𝑟(10 − 𝑛3𝑑), (S4) \n 𝑚spin+7𝑚T= −3𝑝−2𝑞\n𝑟(10 − 𝑛3𝑑), (S5) \nwhere 𝑛3𝑑 and 𝑚T are the number of 3 d electrons on the Fe atom and the expectation \nvalue of the intra -atomic magnetic dipole operator , respectively . Because the \nparamagnetic component observed at Y in Figs. 1(e) and 1(f) in the main text, which \noriginates from the intersti tial Fe atoms, is negligibly small (See Sec tion II of the \nSupplemental Material), the integrated values of the XMCD spectra p [Eq. ( S2)] and q \n[Eq. ( S3)] can be attributed only to the substitutional Fe atoms. Meanwhile, because the \nXAS signals have both c ontributions of the substitutional and interstitial Fe atoms, we \nreduced the integrated XAS intensity r [Eq. ( S1)] to 85% of its raw value (85% is the \napproximate ratio of the substitutional Fe atoms to that of the total Fe atoms in both \nsamples A and BS4) when using the XMCD sum rules. We neglect the expectation value \nof the intra -atomic magnetic dipole operator, because it is negligibly small for Fe atoms \nat the Td symmetry site.S5 Also, we assume 𝑛3𝑑 to be 6 and the correction factor for \nthe mspin to be 0.88 for Fe2+.S6 \nBy the above calculations using the temperature dependence of XAS and XMCD \nspectra shown in Figs. S1 and S2, we obtain ed the temperature dependence of mspin and \nmorb of substitutional Fe atoms as shown in Figs. 2(a) and 2(b) in the main text. \n \n \n \n \n 15 \n \n \nFig. S1. (a) XAS [= ( μ+ + μ-)/2] spectra (solid curves) and the XAS signals integrated \nfrom 690 eV (dashed curves) of sample A. (b) XMCD (= μ+ - μ-) spectra (solid curves) \nand the XMCD signals integrated from 690 eV (dashed curves) of sample A. These \nmeasurements were carried out with a magnetic field 𝜇0H = 5 T applied perpendicular \nto the film surface at 5.6 K (black curves), 20 K (blue curves), 5 0 K (light blue curves), \n100 K (green curves), 150 K (orange curves), 250 K (pink curves), and 300 K (red curves). \n \n \nFig. S2. (a) XAS [= ( μ+ + μ-)/2] spectra (solid curves) and the XAS signals integrated \nfrom 690 eV (dashed curves) of sample B. (b) XMCD (= μ+ - μ-) spectra (solid curves) \nand the XMCD signals integrated from 690 eV (dash ed curves) of sample B. These \nmeasurements were carried out with a magnetic field 𝜇0H = 5 T applied perpendicular \n 16 \nto the film surface at 5.6 K (black curves), 50 K (light blue curves), 100 K (green curves), \n150 K (orange curves), and 250 K (pink curves). \n \n \nII. Influence of the paramagnetic XMCD component on the XMCD sum -rules \nanalyses \n \nFigures S3 show s the XMCD spectra of samples A (a) and B (b) normalized to 707.3 \neV measured at 5.6 and 300 K with magnetic fields of 0.1 and 5 T applied perpendicular \nto the film surface . In both films, all the line shape s of the XMCD spectr a are almost \nthe same, which means that the paramagnetic component observed at Y in Figs. 1( e) and \n1(f) of the main text is negligibl y small for the XMCD sum -rules analyses and that the \nintegrated XMCD signal can be attributed only to the substitutional Fe atoms. \n \n \nFig. S3. XMCD spectra of samples A (a) and B (b) normalized to 707.3 eV measured at \n5.6 a nd 300 K with magnetic fields of 0.1 and 5 T applied perpendicular to the film surface . \n \nIII. Estimation of the Curie constant per substitutional Fe atom \n \nThe Curie constant per substitutional Fe atom is obtained using the following \nequations: \n𝐶= 𝜇B2\n3𝑘B𝑛B2, (S6) \n 17 \n𝑛B= [3\n2+ 𝑆(𝑆+1) − 𝐿(𝐿+1)\n2𝐽(𝐽 + 1)]√𝐽(𝐽+1), (S7) \nwhere 𝜇B, 𝑘B, 𝑛B, S, L, and 𝐽 represent the Bohr magneton, the Boltzmann constant , \nthe effective Bohr magneton number, the spin angular momentum, the orbital angular \nmomentum, and the total angular momentum, respectively. We obtained 𝑛B assuming \nS = 2 (for Fe2+), L = 1.2 ( L = 2S × morb/mspin, where morb/mspin ≈ 0.3 as shown in Figs. \n2(a) and 2(b) in the main text), and J = 3.2 (= L + S because the spin and orbital angular \nmomenta of a substitutional Fe atom are parallel) in Eq. ( S7). Thus, 𝑛B is estimated to \nbe 5.96. \n \nIV . Estimation of the high -field magnetic susceptibility of a substitutional \nparamagnetic Fe atom \n \nAt very low temperature below ~20 K, the effective magnetic -field Heff dependence \nof the total magnetization M (= mspin+morb) of one substitutional paramagnetic Fe atom is \nexpressed by the Langevin function. Thus, the Heff dependence of M of one \nsubstitutio nal paramagnetic Fe atom at 5.6 K is obtained by substituting 5.2 μB, 1, and 5.6 \nK in mSPM, fSPM, and T of Eq. (1) in the main text, respectively (Fig. S4). We \napproximated the high -field magnetic susceptibility ∂𝑀/𝜕(𝜇0𝐻eff) (μB/T per Fe) at 4 T \nby th e slope of the M-Heff line from 4 T to 5 T (black dashed line in Fig. S4). In this \nway, ∂𝑀/𝜕(𝜇0𝐻eff) at 4 T is estimated to be 0.37 μB/T per one substitutional \nparamagnetic Fe. In Figs. 2(d) and 2(e) in the main text, t he ∂𝑀/𝜕(𝜇0𝐻eff) value s at \n4 T and 5.6 K in samples A and B are 0.15 and 0.10, respectively ; it follows that the ratios \nof paramagnetic Fe atoms to the total number of Fe atoms are ~41% (= 0.15/0.37) in \nsample A and ~27% (= 0.10/0.37) in sample B. \n 18 \n \nFig. S4. Effective magnetic -field Heff dependence of the total magnetization M (= \nmspin+morb) per one substitutional paramagnetic Fe at 5.6 K obtained using Eq. (1) in the \nmain text. \n \nReferences \nS3 C. T. Chen, Y . U. Idzerda, H. -J. Lin, N. V . Smith, G. Meigs, E. Chaban, G. H. Ho, E. \nPellegrin, and F. Sette, Phys. Rev. Lett. 75, 152 (1995). \nS4 Y . K. Wakabayashi, S. Ohya, Y . Ban, and M. Tanaka, J. Appl. Phys . 116, 173906 \n(2014). \nS5 J. Stohr and H. Konig, Phys. Rev. Lett. 75, 3748 (1995). \nS6 C. Piamonteze, P. Miedema, and F. M. F. de Groot, Phys. Rev. B 80, 184410 (2009). \n" }, { "title": "2110.06478v1.Theoretical_Study_of_Temperature_Dependence_of_Spin_Susceptibility_in_Anisotropic_Itinerant_Ferromagnets.pdf", "content": "Theoretical Study of Temperature Dependence of Spin \nSusceptibility in Anisotropic Itinerant Ferromagnets \n \n \nDaisuke Miura * and Akimasa Sakuma \n \nDepartment of Applied Physics, Tohoku University , Sendai 980 -8579 , Japan \n \nWe develop ed a framework for directly calculating the spin \nsusceptibility of anisotropic itinerant ferromagnets in the full \ntemperature range within the coherent potential approximation in the \ndisordered local moment picture . As a test of our formulation, we \ndemonstrate the computation for the temperature ( 𝑇) dependence of \nthe spin susceptibility for Rashba -type ferromagnets with the Curie \ntemperature 𝑇C . In a certain parameter, we found that the inverse \ntransverse su sceptibility 1/𝜒⊥≃const. for 𝑇<𝑇C and 1𝜒⊥⁄ ∝\n𝑇−𝑇C for 𝑇>𝑇C , which reflects characteristics of itinerant \nferromagnets including a spin -orbit interaction. \n \nThe magnetic susceptibility 𝜒 is one of the most fundamental properties of magnetic \nmaterials, whose microscopic description has attracted the interest of many researchers [1–5]. \nHistorically, demonstrating the Curie –Weiss (CW) law has been a touchstone issue to confirm \nthe reasonableness of models and approximations used in theoretical stud ies on finite -\ntemperature magnetism. It has long been known that the transverse spin fluctuation plays a \ncentral role in the CW behavior in the localized electron systems such as ferromagnetic \ninsulators, which can easily be recognized by the Heisenberg model [6]. On the other hand, \nderiv ing the CW law on itinerant electron systems such as ferromagnetic metals was a lengthy \nprocess . Primarily , this is for the itinerant electrons ; the magnitude of local moment on each \nsite is not a conserved quantity against the temperature change and the e xternal field, which is \ngoverned by the electron -electron correlation effects. For weak ferromagnets, in 1972, Murata \nand Doniach [7] have first presented the phenomenological description of 𝜒 in which the self-\nconsistent mean field theory was improved to include the mode -mode coupling, and the first \nmicroscopic one was given by Moriya and Kawabata in 1973 [8,9] . Furthermore, a unified \npicture including a treatment of strong fe rromagnets was proposed by Moriya and Takahashi in \n1978 [10]. In addition to the transverse spin fluctuation considered in the Heisenberg model, \nthese theories considered the longitudinal spin fluctuation through the mode -mode coupling of \nspin-waves and successfully demonstrated the qualitative characteristic of itinerant electron \nsystem s, especially of weak ferromagnets. \nRecently, in both practical a nd industrial viewpoint s, strong ferromagnets having a large 2 \n magnetization and a high 𝑇C are required. The disordered local moment (DLM) analysis [11–\n16], by using the first -principles calculation , has shown at quantitative level s that many strong \nferromagnetic metals retain robust local magnetic moments even above the 𝑇C, implying that \nthe transverse spin fluctuation greatly governs the CW behavior in these systems. In fact, first -\nprinciples calculations for the magnetic susceptibility of Fe [17–20] and Ni [18,20] have been \nperformed on the basis of combined theory [17,21] between the density functional theory and \nthe coh erent potential approximation (CPA) in the DLM picture in finite temperatures . There , \nthe CW law was successfully observed for both Fe and Ni , although the calculations for the \nCurie temperature of Ni resulted in an underestimation indicating the failure o f the DLM picture. \nIn addition to the CW law in 1/𝜒 , the importance of finite temperature magnetism of \npractical materials lies in a temperature dependence of magnetic anisotropy constants (MACs) \nbelow 𝑇C . In this temperature range, several authors have performed first-principles \ncalculations for magnetocrystalline anisotropy (MA) in the DLM picture for metallic \nferromagnets for FePt [22,23] , FePd [23,24] , CoPt, MnAl, and FeCo [23]. In addition, the \nauthors demonstrated that the temperature dependence of the MACs are roughly proportional \nto the second power of the magnetization , as indicated by experiments for FePt [25]. The second \npower law cannot be understood by single ion models predicting the third power law (so -called \nAkulov –Zener –Callen –Callen law) [26–31], which suggests a characteristic of ferromagnetic \nmetals that the DLM picture appears to capture well. Again, the transverse susceptibility 𝜒⊥ \ncan reflect the temperature dependence of MA below the 𝑇C. For example, if a material has \nuniaxial MA, then the 𝜒⊥ may ha ve a finite value depending on the temperature, whereas 𝜒⊥ \ndiverges below the 𝑇C in the absence of MA. However , theoretical stud y of 𝜒⊥ in the full \ntemperature range has not yet been reported for anisotropic itinerant ferromagnetism. \nThe purpose of this work is to provide a practical framework for directly computing the \ntemperature dependence of the magnetic susceptibility of anisotropic ferromagnetic metals in \nthe full temperature range from the microscopic viewpoint based on the DLM and CPA theory . \nFor confirming the validity of our formulation, we also show the temperature dependence of \nmagnetization curves , from whose gradient we can estimate the magnetic susceptibilit y. \nHowever, the computational load fo r obtaining the magnetization curves is overly heavy ; thus, \nwe employ the two -dimensional tight -binding Rashba model [32] for describing itinerant \nelectrons with MA, which has been valued as a theoretically tractable model for investigating \nanisotropic magnetic metals [33–35]. The usage of the CPA -Green function method enables us \nto easily extent the present theory to first -principles calculation s in the future, which has been 3 \n widely used for evaluating various physical quantities such as electrical conductivities [36,37] , \nGilbert damping constants [38,39] , magnetic anisotropies [24,40] , and so on. \nLet us consider the tight -binding Rashba Hamiltonian in the presence of an external uniform \nstatic magnetic field 𝑯=(𝐻x,𝐻y,𝐻z) on the two -dimensional square lattice given by \n ℋ≔∑𝑐𝒌†(𝜖̂𝒌+𝑣̂𝑯)𝑐𝒌\n𝒌−𝛥ex∑𝒆𝑖⋅𝑐𝑖†𝛔̂𝑐𝑖𝑁\n𝑖=1, (1) \nwhere 𝑐𝒌 is the annihilation operator of the electron with a wave vector 𝒌=(𝑘x,𝑘y) in the \nspinor representation , the 𝒌 summation is perfor med in the first Brillouin zone, 𝑐𝑖≔\n(1\n√𝑁)∑𝑐𝒌exp(i𝒌⋅𝑹𝑖)𝒌 , 𝛔̂≔(σ̂x,σ̂y,σ̂z) for the Pauli matrices σ̂𝛼 (a quantity with the hat \nis a 2×2 matrix in the spin space ), and 𝛥ex denotes the strength of the exchange field whose \ndirection is a classical unit vector 𝒆𝑖 allocated at a lattice vector 𝑹𝑖 in the 𝑁 sites. In the \nlanguage of the functional integral method under the Hubbard -Stratonovich transformation, \n𝛥ex corresponds to the saddle point concerning the amplitude of auxiliary exchange field s. The \nkinetic and Zeeman parts in the first term in Eq. (1), respectively, are defined by \n 𝜖̂𝒌≔−2𝑡(cos𝑘x𝑎+cos𝑘y𝑎)1̂+𝜆(σ̂xsin𝑘y𝑎−σ̂ysin𝑘x𝑎), (2) \n 𝑣̂𝑯≔𝑔μB\n2 𝝈̂⋅μ0𝑯≡𝑡𝝈̂⋅𝒉, (3) \nwhere 𝑡 is the hopping integral between the two nearest -neighbor sites, 𝑎 is the lattice \nconstant, 1̂ is the unit matrix, 𝜆 is the strength of the Rashba spin-orbit interaction ( SOI), 𝑔 \nis the g factor of an electron , μB is the Bohr magneton , μ0 is the permeability in vacuum , and \n𝒉 stands for the d imensionless external magnetic field . In the DLM -CPA theory , it is assume d \nthat a local moment at each site is located in a configuration {𝒆𝑖}, and {𝒆𝑖} is determined by \na functional integral method based on the single -site approximation at each temperature ; \nsubsequently , a Green ’s function is obtained from the CPA condition [11,15] . In the present \nmodel , the single -electron Green ’s function [35] between sites 𝑖 and 𝑗 is given by \n 𝐺̂𝑖𝑗𝑯(𝑧):=1\n𝑁∑ei𝒌⋅(𝑹𝑖−𝑹𝑗)𝑔̂𝒌𝑯(𝑧)\n𝒌, (4) \nwhere 𝑧 is a complex variable , and 𝑔̂𝒌𝑯(𝑧)≔[𝑧1̂−𝜖̂𝒌−𝑣̂𝑯−𝛴̂𝑯(𝑧)]−1 . The self -energy \n𝛴̂𝑯(𝑧) is determined by the CPA condition: \n ⟨𝑡̂𝒆𝑯(𝑧)⟩𝒆𝑯=(00\n00), (5) \nwhere 4 \n ⟨⋯⟩𝒆𝑯≔∫d𝒆 𝑤𝒆𝑯⋯, (6) \n ∫d𝒆≔∫d𝜃sin𝜃𝜋\n0∫ d𝜙2𝜋\n0 in terms of 𝒆≔(sin𝜃cos𝜙,sin𝜃sin𝜙,cos𝜃), (7) \n 𝑡̂𝒆𝑯(𝑧)≔1\n1̂−[−𝛥ex𝒆⋅𝛔̂−𝛴̂𝑯(𝑧)]𝐺̂00𝑯(𝑧)[−𝛥ex𝒆⋅𝛔̂−𝛴̂𝑯(𝑧)], (8) \n 𝑤𝒆𝑯≔e−𝛽(Ω𝒆𝑯−𝐹𝑯), (9) \n Ω𝒆𝑯≔−1\n𝛽∑Trln{1̂−[−𝛥ex𝒆⋅𝛔̂−𝛴̂𝑯(𝑧𝑛𝑯)]𝐺̂00𝑯(𝑧𝑛𝑯)}\n𝑛, (10) \n 𝐹𝑯≔−1\n𝛽ln∫d𝒆 e−𝛽Ω𝒆𝑯, (11) \nwhere 𝛽≔(kB𝑇)−1 for the temperature 𝑇 , 𝑧𝑛𝑯≔iℏ𝜔𝑛+𝜇𝑯 , 𝜔𝑛≔𝜋(2𝑛+1)/(𝛽ℏ) , \nthe 𝑛 summation is performed over all integers, and 𝜇𝑯 is the chemical potential . \n The spin magnetization 𝑴𝑯, in units of magnetic field, is represented regarding the single -\nelectron Green’s function as \n 𝑴𝑯=−𝑔μB\n2𝑎3𝛽∑Tr𝛔̂𝐺̂00𝑯(𝑧𝑛𝑯)\n𝑛≡𝑔μB\n2𝑎3𝝈𝒉, (12) \nwhere 𝝈𝒉 denotes the spin concentration vector (dimensionless) . Here w e assume the form of \n𝝈𝟎=(0,0,𝜎s) with 𝜎s>0 for 𝑇<𝑇C, that is , we set the model parameters s uch that the \nsystem exhibits uniaxial ly anisotropic ferromagnets . In addition, we must refer to the problem \nof which be fixed, the chemical potential or the electron concentration . Although the magnetism \nand the MA of the ground state , at least, must be ferromagnetic and perpendicular to the lattice \nplane, both strongly depend on the electron concentration as shown by previous theoretical \nstudies on Rashba magnets [33–35]. Therefore, we adjust 𝜇𝑯 such that the electron \nconcentration is fixed to an appropriate value 𝑛e: \n 𝑛e𝑯=𝑛e, (13) \nwhere 𝑛e𝑯 is represented by \n 𝑛e𝑯=1\n𝛽∑Tr𝐺̂00𝑯(𝑧𝑛𝑯)\n𝑛. (14) \nFrom Eq. (12), we obtain the formally microscopic expressions for the longitudinal and \ntransverse spin susceptibilities (dimensionless) as \n 𝜒∥:=lim\n𝑯→𝟎𝜕𝑀 z𝑯\n𝜕𝐻z=𝑔2μB2μ0\n4𝑡𝑎3𝜒̃z, (15) \nand \n 𝜒⊥:=lim\n𝑯→𝟎𝜕𝑀 x𝑯\n𝜕𝐻x=𝑔2μB2μ0\n4𝑡𝑎3𝜒̃x, (16) 5 \n respectively, and the normalized susceptibility has been defined by \n 𝜒̃𝛼:=lim\n𝒉→𝟎𝜕𝜎𝛼𝒉\n𝜕ℎ𝛼=−lim\n𝒉→𝟎1\n𝛽∑Trσ̂𝛼𝜕𝐺̂00𝑯(𝑧𝑛𝑯)\n𝜕ℎ𝛼𝑛. (17) \nAs indicated , we simply must compute the on -site Green function 𝐺̂00𝑯(𝑧𝑛𝑯) in principle , but \nthe high-precision computation is required for reducing an error owing to the numerical \ndifferentia tion, especially whose compu tational load for 𝜒⊥ is heavy . Therefore, developing a \nmethod to directly comput e the derivative of 𝐺̂00𝑯(𝑧𝑛𝑯) is desired for future realistic \ncalculations. It can be achieved by the following pro cedure. \n Firstly, from the definition (4), the derivative of 𝐺̂00𝑯(𝑧𝑛𝑯) is represented by \n 𝜕𝐺̂00𝑯(𝑧𝑛𝑯)\n𝜕𝒉=−1\n𝑁∑𝑔̂𝒌𝑯(𝑧𝑛𝑯)(𝜕𝜇𝑯\n𝜕𝒉1̂−𝑡𝝈̂−𝜕𝛴̂𝑯(𝑧𝑛𝑯)\n𝜕𝒉)𝑔̂𝒌𝑯(𝑧𝑛𝑯)\n𝒌, (18) \nand 𝜕𝜇𝑯/𝜕𝒉 is determined by \n 𝜕𝜇𝑯\n𝜕𝒉=(1\n𝛽𝑁∑Tr𝑔̂𝑘𝑯(𝑧𝑛𝑯)2\n𝑛𝒌)−11\n𝛽𝑁∑Tr𝑔̂𝑘𝑯(𝑧𝑛𝑯)2(𝑡𝝈̂+𝜕𝛴̂𝑯(𝑧𝑛𝑯)\n𝜕𝒉)\n𝑛𝒌, (19) \nwhich is derived by 𝜕𝑛e𝑯/𝜕𝒉=𝟎 from the condition (13) and the definition (14). The final \nform of the equation to be solved is obtained by differentiating both sides of the CPA condition \n(5) with respect to 𝒉: \n ∂𝛴̂𝑯(𝑧𝑛𝑯)\n∂𝒉−⟨𝑡̂𝒆𝑯(𝑧𝑛𝑯)(𝜕𝐺̂00𝑯(𝑧𝑛𝑯)\n𝜕𝒉−𝐺̂00𝑯(𝑧𝑛𝑯)𝜕𝛴̂𝑯(𝑧𝑛𝑯)\n𝜕𝒉𝐺̂00𝑯(𝑧𝑛𝑯))𝑡̂𝒆𝑯(𝑧𝑛𝑯)⟩\n𝒆𝑯\n=−⟨𝑡̂𝒆𝑯(𝑧𝑛𝑯)Tr∑(𝜕𝐺̂00𝑯(𝑧𝑚𝑯)\n𝜕𝒉−𝐺̂00𝑯(𝑧𝑚𝑯)𝜕𝛴̂𝑯(𝑧𝑚𝑯)\n𝜕𝒉𝐺̂00𝑯(𝑧𝑚𝑯))𝑡̂𝒆𝑯(𝑧𝑚𝑯)\n𝑚⟩\n𝒆𝑯\n. (20) \nConsequently, we can obtain a closed equation for ∂𝛴̂𝑯(𝑧𝑛𝑯)∂𝒉⁄ from Eqs. (18), (19), and \n(20), which is free from the numerical differentiation . After the limit of 𝒉→𝟎, the equation \nfor ∂𝛴̂𝑯(𝑧𝑛𝑯)∂𝒉⁄ is consisted of only the quantities based on the zero -magnetic field CPA . \nHere , in Eq. (19) for 𝑇<𝑇C and 𝒉→0, we notice that 𝜕𝜇𝑯𝜕ℎx→0 ⁄ but 𝜕𝜇𝑯𝜕ℎz ⁄ does \nnot vanish as noted in Ref. [41]. \n Let us consider a st rong Rashba SOI case [42] in which the parameters are set as 𝛥ex𝑡⁄=\n0.2 and 𝜆𝑡⁄=1 , and 𝑛e=0.25 ; the Curie temperature was estimated at kB𝑇C𝑡⁄=\n3.7×10−4 from the computation of the temperature dependence of 𝜎s. Fortunately, because \nof the simplicity of the present model , we can directly perform the CPA computation for finite \n𝒉 . Figure 1 shows the calculated spin-concentration curves along the hard axis , i.e. , 𝒉=\n(ℎx,0,0), by Eq. (12) for several temperatures . For each temperature, we can define a magnetic \nfield ℎA such that 𝜎z(ℎA,0,0)=0. In the range of ℎx<ℎA, the magnetization rotates in the 𝑧-6 \n 𝑥 plane as indicated by |𝝈𝒉|≃const . On the other hand , in the range of ℎx≥ℎA , the \nmagnetization is perfectly parallel to the external magnetic field and the magnitude slowly \ngrows with increasing ℎx primarily because of suppressing the spin fluctuation . Here, we \nshould refer the ℎx dependence of 𝜎x(ℎx,0,0) at a near zero temperature , 𝑇=0.005 =𝑇1 in \nFig. 1 . Although 𝜎x(ℎx,0,0) appears to be in a constant in the high -field range of ℎx>ℎA in the \npresent scale of the vertical axis in Fig. 1, the gradient — the high -field spin susceptibility \n𝜒high ≔𝜕𝜎x(ℎx,0,0)/𝜕ℎx — is no n-zero, whose value is evaluated at 𝜒high ≃0.179 by the \nlinear fitting to 𝜎x(ℎx,0,0) with respect to ℎx . This finite value can be understood as a \ncharacteristic of the itinerant electron system. More specifically, ℎx enhances the spin \npolarization of the density of states of the electron via the Zeeman energy , and its effect on \n𝜒high is represented by 𝜒high ≃4𝜌↑𝜌↓(𝜌↑+𝜌↓) ⁄ for ℎx≪𝛥ex/𝑡 where 𝜌𝜎≔\n−(𝑡𝜋⁄)ℑ [𝐺̂00𝟎(𝜇𝟎+i0)]𝜎𝜎 at 𝑇=0. By the expression, we can obtain 𝜒high ≃0.175 from \nthe calculated values of 𝜌↑=0.0909 and 𝜌↓=0.0841 , and it is in good agreement with the \nabove fitting value . Next, we can find a high linearity on the 𝜎x𝒉-ℎx curves in the cases of 𝑇=\n𝑇1,𝑇2,𝑇3,𝑇4(<𝑇C) and ℎx<ℎA, and can observe that its gradient at ℎx=0, i.e., 𝜒̃x, does \nnot depend on temperature , in which 𝜒̃x describes the response regarding the magnetization \nrotatio n. In the para magnetic case (𝑇=𝑇5,𝑇6,𝑇7(≥𝑇C)), the external magnetic field, even at \nnear zero strength , acts on the magnetization as a suppresser for the spin fluctuatio n (not as a \ndriving force of the rotation ), and thus 𝜒̃x directly reflects the spin fluctuation , and \nconsequently i ts temperature dependence is led to the CW law. Figure 2 shows the spin \nsusceptibilities calculated on the basis of Eqs. (15)–(20), for 𝜆𝑡⁄=1 and 1/√2. As expected \nfrom the above discussion, 𝜒̃x is a constant for 𝑇<𝑇C and exhibits the CW law for 𝑇≥𝑇C. \nThe comparison between the results for 𝜆𝑡⁄=1 and 1/√2 suggests 𝜒̃x−1∝𝜆2, which may \nbe interpret ed as below discussion. Now , we approximate ℎA by the equation , 𝜎s=\n𝜎x(ℎA,0,0)≃𝜒̃xℎA, that is, \n ℎA=𝜎s\n𝜒̃x, (21) \nand if estimating the MAC from 𝐾u=𝑡𝜎sℎA(2𝑎3) ⁄ by regarding ℎA as the magnetic \nanisotropy field, which correspond s to the area of the triangle formed by 𝑀z𝑯-μ0𝐻x and 𝑀x𝑯-\nμ0𝐻x curves , then we have \n 𝐾u=𝑡\n2𝑎3𝜎s2\n𝜒̃x. (22) \nIgnoring the effect of the Rashba SOI on the temperature dependence of t he saturation \nmagnetization , we may estimate the 𝜆 dependency of 𝜒̃x−1 from one of 𝐾u . Because t he \nperturbative treatment for 𝐾u with respect to 𝜆 gives that 𝐾u∝𝜆2 [33,34] , we can accept the \nrelation 𝜒̃x−1∝𝜆2. Furthermore, let us consider what Eq. (22) suggests from the viewpoint of 7 \n the temperature dependence of the MAC. Although the Akulov –Zener –Callen –Callen law \npredicts 𝐾u∝𝜎s3 for single -site uniaxial MA systems [26–31], several previous studies have \nindicated 𝐾u∝𝜎s2 in ferromagnetic metals [22–25]. Here, r ecalling that t he present model \ndescribes anisotropic ferromagnetic metals , it is rea sonable that the temperature dependence of \n𝜒̃x−1 becomes a constant , i.e., 𝐾u∝𝜎s2. In the anisotropic Heisenberg model, the temperature \ndependenc ies of 𝜒̃x−1 and 𝜒̃z−1 have been demonstrated by several authors [43–45]. \nEspecially, Fröbrich and Kuntz have obtained a constant 𝜒̃x and a slightly decreasing 𝜒̃x in \ninter-site and single -site anisotrop ic models , respectively, for 𝑇<𝑇C. This suggests that the \nfinite -temperature magnetism of the present model involves the situation presented by the \nHeisenberg model with the inter-site anisotropy. Lastly, we remark the effect of the SOI on 𝜒̃z−1 \nin this model. The longitudinal susceptibility is a response for the external magnetic field along \nthe easy axis yielded by the SOI , thus the Zeeman effect and the SOI do not compete. \nConsequently , the SOI acts as an additional effective magnetic field, only affecting the Curie \ntemperature as indicated by the shift of the zero points of 𝜒̃z−1. \n In conclusion, we have develope d a computational method for directly calculating the spin \nsusceptibility of itinerant ferromag nets with the uniaxial MA. It is based on the DLM -CPA \ntheory and includes the finite temperature effect over the full range. To demonstrate , we have \nperformed the computation for the temperature dependence of the spin susceptibility in the \nRashba -type ferromagnet , and especially, we have confirmed that the transverse susceptibility \n𝜒̃x has a finite value reflecting the MAC for 𝑇≤𝑇C and obeys the CW law for 𝑇>𝑇C. For \nthe longitudinal susceptibility 𝜒̃z, on the other hand, the effect of the SOI is limited to the \nchange of 𝑇C and it show s an ordinary temperature dependence as well as isotropic \nferromagnets . \n \nAcknowledgment \nThis work was supported by JSPS KAKENHI Grant Num bers JP17K14800, JP19H056 12, \nand JP21K04624 in Japan. \n \n*E-mail: dmiura 1222 @gmail.com \n \n[1] T. Moriya, J. Magn. Magn. Mater. 14, 1 (1979). \n[2] B. L. Gyorffy, A. J. Pindor, J. Staunton, G. M. Stocks, and H. Winter, J. Phys. F Met. Phys. 15, \n1337 (1985). \n[3] J. B. Staunton, Rep. Prog. Phys. 57, 1289 (1994). \n[4] F. J. Ohkawa, Phys. Rev. B 65, 174424 (2002). \n[5] N. Singh, ArXiv 1807.11291 (2018). \n[6] W. Heisenberg, Z. Phys. 49, 619 (1928). \n[7] K. K. Murata and S. Doniach, Phys. Rev. Lett. 29, 285 (1972). 8 \n [8] T. Moriya and A. Kawabata, J. Phys. Soc. Jpn. 34, 639 (1973). \n[9] T. Moriya and A. Kawabata, J. Phys. Soc. Jpn. 35, 669 (1973). \n[10] T. Moriya an d Y. Takahashi, J. Phys. Soc. Jpn. 45, 397 (1978). \n[11] H. Hasegawa, J. Phys. Soc. Jpn. 46, 1504 (1979). \n[12] H. Hasegawa, J. Phys. Soc. Jpn. 49, 178 (1980). \n[13] H. Hasegawa, J. Phys. Soc. Jpn. 49, 971 (1980). \n[14] J. Hubbard, Phys. Rev. B 19, 2626 (1979). \n[15] J. Hubbard, Phys. Rev. B 20, 4584 (1979). \n[16] J. Hubbard, Phys. Rev. B 23, 5974 (1981). \n[17] T. Oguchi, K. Terakura, and N. Hamada, J. Phys. F Met. Phys. 13, 145 (1983). \n[18] J. Staunton, B. L. Gyorffy, A. J. Pindor, G. M. Stocks, and H. Winter, J. Magn. Magn. Mater. \n45, 15 (1984). \n[19] J. Staunton, B. L. Gyorffy, G. M. Stocks, and J. Wadsworth, J. Phys. F 16, 1761 (1986). \n[20] J. B. Staunton and B. L. Gyorffy, Phys. Rev. Lett. 69, 371 (1992). \n[21] A. J. Pindo r, J. Staunton, G. M. Stocks, and H. Winter, J. Phys. F Met. Phys. 13, 979 (1983). \n[22] J. B. Staunton, S. Ostanin, S. S. A. Razee, B. L. Gyorffy, L. Szunyogh, B. Ginatempo, and E. \nBruno, Phys. Rev. Lett. 93, 257204 (2004). \n[23] N. Kobayashi, K. Hyodo, and A. Sakuma, Jpn. J. Appl. Phys. 55, 100306 (2016). \n[24] J. B. Staunton, L. Szunyogh, A. Buruzs, B. L. Gyorffy, S. Ostanin, and L. Udvardi, Phys. Rev. \nB 74, 144411 (2006). \n[25] S. Okamoto, N. Kikuchi, O. Kitakami, T. Miyazaki, Y. Shimada, and K. Fukamichi, Phys. Rev. \nB 66, 024413 (2002). \n[26] N. Akulov, Z. Phys. 100, 197 (1936). \n[27] C. Zener, Phys. Rev. 96, 1335 (1954). \n[28] H. B. Callen and E. Callen, J. Phys. Chem. Solids 27, 1271 (1966). \n[29] D. Miura, R. Sasaki, and A. Sakuma, Appl. Phys. Express 8, 113003 (2015). \n[30] D. Miura and A. Sakuma, AIP Adv. 8, 075114 (2018). \n[31] D. Miura and A. Sakuma, J. Phys. Soc. Jpn. 88, 044804 (2019). \n[32] T. Ando and H. Tamura, Phys. Rev. B 46, 2332 (1992). \n[33] K. W. Kim, K. J. Lee, H. W. Lee, and M. D. Stiles, Phys. Rev. B 94, 184402 (2016). \n[34] J. Ieda, S. E. Barnes, and S. Maekawa, J. Phys. Soc. Jpn. 87, 053703 (2018). \n[35] A. Sakuma, J. Phys. Soc. Jpn . 87, 034705 (2018). \n[36] I. Turek, J. Kudrnovský, and V. Drchal, Phys. Rev. B 89, 064405 (2014). \n[37] K. Hyodo, A. Sakuma, and Y. Kota, Phys. Rev. B 94, 104404 (2016). \n[38] H. Ebert, S. Mankovsky, D. Ködderitzsch, and P. J. Kelly, Phys. Rev. Lett. 107, 066603 \n(2011). \n[39] A. Sakuma, J. Phys. Soc. Jpn. 81, 084701 (2012). \n[40] A. Deák, E. Simon, L. Balogh, L. Szunyogh, M. Dos Santos Dias, and J. B. Staunton, Phys. 9 \n Rev. B 89, 224401 (2014). \n[41] K. Kuboki and H. Yamase, Phys. Rev. B 96, 064411 (2017). \n[42] Although these parameters set were referred in our previous study [35], the value of λt⁄=1 \nmay be too large as typical 3d ferromagnetic metals [for example, S. E. Barnes, J. Ieda, and S. \nMaekawa, Sci. Rep. 4, 4105 (2014); C.-H. Chang , K.-P. Dou , G.-Y. Guo , and C. -C. Kaun, \nNPG Asia Materials 9, e424 (2017)] . In the present work, t his large SOI has been introduced \nonly for easily showing both the CW law and a finite value of χ̃x in the vertical -axis scale in \nFig. 2 . Furthermore, we notice that , with respect to even such large λ, the purturbative \ntreatment for the spin susceptibility remains valid as discussed later . \n[43] P. J. Jensen, S. Knappmann, W. Wulfhekel, and H. P. Oepen, Phys. Rev. B 67, 184417 (2003). \n[44] P. Fröbrich and P. J. Kuntz, Phys. Stat. Sol. 241, 925 (2004). \n[45] P. Fröbrich and P. J. Kuntz, J. Phys. Condens. Matter 16, 3453 (2004). \n \n \n 0510152025303540\n0123456Fig. 1 The calculated spin concentrations as functions of external magnetic fields ݄୶\tfor several temperatures. The\nsolid, dashed, and dotted lines are ߪ୶,ߪ,a n d||, respectively. Each number reflects the index of the\ntemperatures given by ܶଵൌ 0.005 ,ܶଶൌ0 . 5 ,ܶଷൌ0 . 8 ,ܶସൌ 0.095 ,ܶହൌ1,ܶൌ1 . 5 ,a n dܶൌ3in units of ܶେ.\nNote that ߪൌ0and||ൌߪ୶\tforܶହ,ܶ,a n dܶ.Spin concentrations×10-3\n×10-31\n1\n22\n3\n3\n44\n5\n6\n7\n୶\n0.00.10.20.30.40.50.60.70.80.91.0\n01234567Inverse spin susceptibilities×10-4୶\n\nFig. 2 The calculated inverse spin susceptibilities as functions of\ntemperature ܶ. The solid and dotted lines correspond to the results\nfor⁄ݐߣൌ1and1/2, respectively." }, { "title": "1908.07652v1.Novel_universality_class_for_the_ferromagnetic_transition_in_the_low_carrier_concentration_systems_UTeS_and_USeS_exhibiting_large_negative_magnetoresistance.pdf", "content": "arXiv:1908.07652v1 [cond-mat.str-el] 21 Aug 2019APS/123-QED\nNovel universality class for the ferromagnetic transition in the low carrier\nconcentration systems UTeS and USeS exhibiting large negat ive magnetoresistance∗\nNaoyuki Tateiwa1,†Yoshinori Haga1, Hironori Sakai1, and Etsuji Yamamoto1\n1Advanced Science Research Center,\nJapan Atomic Energy Agency, Tokai,\nNaka, Ibaraki 319-1195, Japan\n(Dated: August 22, 2019)\nWe report the novel critical behavior of magnetization in lo w carrier concentration systems UTeS\nand USeS that exhibit the large negative magnetoresistance around the ferromagnetic transition\ntemperatures TC∼85 and 23 K, respectively. UTeS and USeS crystallize in the sa me orthorhom-\nbic TiNiSi-type crystal structure as those of uranium ferro magnetic superconductors URhGe and\nUCoGe. We determine the critical exponents, βfor the spontaneous magnetization Ms,γfor the\nmagnetic susceptibility χ, andδfor the magnetization isotherm at TCwith several methods. The\nferromagnetic states in UTeS and USeS have strong uniaxial m agnetic anisotropy. However, the\ncritical exponents in the two compounds are different from th ose in the three-dimensional Ising\nmodel with short-range magnetic exchange interactions. Si milar sets of the critical exponents have\nbeen reported for the uranium ferromagnetic superconducto rs UGe 2and URhGe, and uranium\nintermetallic ferromagnets URhSi, UIr and U(Co 0.98Os0.02)Al. The universality class of the ferro-\nmagnetic transitions in UTeS and USeS may belong to the same o ne for the uranium compounds.\nThe novel critical phenomenon associated with the ferromag netic transition is observed not only in\nthe uranium intermetallic ferromagnets with the itinerant 5felectrons but also in the low carrier\nconcentration systems UTeS and USeS with the localized 5 felectrons. The large negative magne-\ntoresistance in UTeS and USeS, and the superconductivity in UGe2and URhGe share the similarity\nof their closeness to the ferromagnetism characterized by t he novel critical exponents.\nPACS numbers:\nI. INTRODUCTION\nMuch interest has been focused on novel physical phe-\nnomena in uranium metallic compounds with 5 felec-\ntrons such as “hidden order”in URu 2Si2, unconventional\nsuperconductivity in UPt 3or UBe 13, and ferromagnetic\nsuperconductivity in UGe 2, URhGe, and UCoGe[1–4].\nMeanwhile, relatively few studies have been conducted\nfor the magnetism and the electrical conductivity in ura-\nnium semimetals or semiconductors. This is in contrast\nwith rare earth magnetic semiconductors, for example,\neuropium chalcogenides EuX (X = O, S, Se, and Te)\nwhere the interplay between 4 fand conduction electrons\nplaysanimportantroleforanomalousphysicalproperties\nsuch as a negative magnetoresistance[5].\nVery recently, the superconductivity has been discov-\nered in uranium compound UTe 2[6]. We have studied\nuranium dichalcogenides UTeS, USeS, and β-US2that\nshow the large magnetoresistanceat low temperatures[7–\n11]. Figure 1 (a) represents the orthorhombic TiNiSi-\ntype crystal structure ( Pnma) of the uranium dichalco-\ngenides. The structure is the same as those of uranium\nferromagnetic superconductors URhGe and UCoGe[4].\nNote that UTe 2crystalizes in a different orthorhombic\nstructure ( Immm)[12]. Figure 1(b) shows the tempera-\n∗Phys. Rev. B 100, 064413 (2019)\n†Electronic address: tateiwa.naoyuki@jaea.go.jpturedependencies ofthe electricalresistivity ρin the ura-\nnium dichalcogenides under magnetic fields of 0 T and 7\nT applied parallel to the crystallographic c-axis[10, 11].\nThe electrical current Jwas applied along the b-axis.\nUTeS is a semimetal with a low carrier density in the\norder of 1025m−3[10]. USeS and β-US2are narrow-gap\nsemiconductors[7,9,11]. UTeSandUSeS showferromag-\nnetic transitionsat TC= 85 and 23 K, respectively[9, 10].\nβ-US2does not order magnetically down to 0.5 K[13].\nFigs. 1(c) and 1(d) show the magnetization at 5.0 K in\nUTeS and at 2.0 K in USeS, respectively. The ferromag-\nnetic states have strong uniaxial magnetic anisotropy.\nThe effect of the magnetic field on the electrical resis-\ntivity is strong in the uranium dichalcogenides as shown\nin Fig. 1(b). In particular, the resistivity values in\nUSeS and β-US2decrease largely with increasing field\nat low temperatures[11]. The magnitudes of the trans-\nverse magnetoresistance are comparable with those in\nperovskite-typemanganeseoxides[14]. Theapplicationof\nthe pressure above 1 GPa induces a ferromagnetic state\ninβ-US2[15]. The large magnetoresistance in the ura-\nnium dichalcogenides can be regarded as a novel phys-\nical phenomenon that appears around a ferromagnetic\nphase boundary. The mechanism of the magnetoresis-\ntance has not been fully understood yet[11]. In this pa-\nper, we report the novel critical behavior of the mag-\nnetization in UTeS and USeS, and its similarity to those\nin the uranium ferromagneticsuperconductors UGe 2and\nURhGe[16].2\nabcUST(a)\nFIG. 1: (a)Representation of the orthorhombic TiNiSi-type\ncrystal structure in UTS (T: Te, Se, S). (b)Temperature de-\npendencies of the electrical resistivity ρunder magnetic fields\nof 0 and 7 T in UTeS[10], USeS, and β-US2[11]. Magnetic\nfield dependencies of the magnetization in magnetic fields ap -\nplied along the a,b, andc-axes (c) at 5.0 K in UTeS[10] and\n(d) at 2.0 K in USeS.\nII. EXPERIMENT\nSingle crystals of UTeS and USeS were grown by\nthe chemical transport using bromine as a trans-\nport agent[10, 11]. Neither impurity phase nor off-\nstoichiometric chemical composition larger than about\n1%wasdetected intheX-raydiffractionandthe Electron\nProbe Micro Analysis (EPMA). Magnetization was mea-\nsuredinacommercialsuperconductingquantuminterfer-\nence device (SQUID) magnetometer (MPMS, Quantum\nDesign). Themagneticfield µ0Hextwasappliedalongthe\nmagnetic easy caxis in the orthorhombic structure. We\ndetermine the internal magnetic field µ0Hby subtract-\ning the demagnetization field DMfromµ0Hext:µ0H=\nµ0Hext-DM. Thedemagnetizationfactors D=0.50and\n0.46 were estimated from the macroscopic dimensions of\nthe single crystals of UTeS and USeS, respectively.\nIII. RESULTS\nIn the asymptotic critical region near TCwhere the\nmean field theory fails, the magnetic correlation length\nξ=ξ0|(T−TC)/TC|−νdiverges. Here, νis the critical\nexponent. The spontaneous magnetization Ms, the ini-\ntial susceptibility χ, and the magnetization at TCfollow\nuniversal scaling laws[17].\nMs(T)∝ |t|β(T < T C) (1)\nχ(T)−1∝ |t|γ′\n(T < T C),|t|γ(TC< T) (2)\nM(µ0H)∝(µ0H)1/δ(T=TC) (3)/s49/s46/s53/s120/s49/s48/s49/s54\n/s48/s46/s48 /s77/s49/s47/s98/s32/s91/s65/s47/s109/s93/s49/s47/s98\n/s56 /s52 /s48\n/s40/s72/s47/s77/s41/s49/s47/s103/s51/s68/s32/s73/s115/s105/s110/s103/s40/s100/s41/s32/s32/s85/s83/s101/s83\n/s84/s32/s61/s32/s50/s49/s46/s52/s32/s75/s32\n/s50/s52/s46/s54/s32/s75/s32/s68/s84/s32/s32/s61\n/s32/s48/s46/s56/s32/s75/s50/s120/s49/s48/s49/s48\n/s48/s77/s49/s47/s98/s32/s91/s65/s47/s109/s93/s49/s47/s98\n/s52/s48 /s50/s48 /s48\n/s40/s72/s47/s77/s41/s49/s47/s103/s40/s97/s41/s32/s32/s85/s84/s101/s83\n/s77/s70/s84/s32/s61/s32/s56/s49/s46/s48/s32/s75/s32\n/s56/s57/s46/s48/s32/s75/s32/s68/s84/s32/s32/s61/s32\n/s32/s32/s49/s46/s48/s32/s75\n/s49/s120/s49/s48/s49/s54\n/s48/s77/s49/s47/s98/s32/s91/s65/s47/s109/s93/s49/s47/s98\n/s50/s48 /s49/s48 /s48\n/s40/s72/s47/s77/s41/s49/s47/s103/s40/s99/s41/s32/s32/s85/s84/s101/s83\n/s32/s32/s51/s68/s32/s73/s115/s105/s110/s103/s84/s32/s61/s32/s56/s49/s46/s48/s32/s75/s32\n/s56/s57/s46/s48/s32/s75/s32/s68/s84/s32/s32/s61/s32\n/s32/s32/s49/s46/s48/s32/s75/s51/s120/s49/s48/s49/s48\n/s48/s77/s49/s47/s98/s32/s91/s65/s47/s109/s93/s49/s47/s98\n/s56 /s48\n/s40/s72/s47/s77/s41/s49/s47/s103/s77/s70/s40/s98/s41/s32/s32/s32/s85/s83/s101/s83\n/s84/s32/s61/s32/s50/s49/s46/s52/s32/s75/s32\n/s50/s52/s46/s54/s32/s75/s32/s68/s84/s32/s32/s61\n/s32/s48/s46/s56/s32/s75\nFIG. 2: Magnetic isotherms in the form of M1/βvs.\n(H/M)1/γwith the mean field theory ( β= 0.5 and γ= 1.0)\n(a) for 81.0 K ≤T≤89.0 K in UTeS and (b) for 21.4 K ≤\nT≤24.6 K in USeS. Isotherms with the short-range (SR) 3D\nIsing model ( β= 0.325 and γ= 1.241) in (c) UTeS and in (d)\nUSeS.\nHere,tis the reduced temperature t= (T−TC)/TC.β,\nγ,γ′andδare the critical exponents.\nFigures 2(a) and 2(b) show the magnetic isotherms in\nthe form of M1/βversus (H/M)1/γwith the mean field\n(MF) theory ( β= 0.5 and γ= 1.0) for 81.0K ≤T≤89.0\nK in UTeS and for 21.4 K ≤T≤24.6 K in USeS, respec-\ntively. The data do not form straightlines. This suggests\nthat the mean field theory is not suitable to describe the\nmagnetization around TC. Figs. 2(c) and 2(d) represent\nthe isotherms in the form of M1/βvs. (H/M)1/γwith\nthe 3D Ising model with short-range (SR) exchange in-\nteractions( β= 0.325and γ= 1.241)for UTeS and USeS,\nrespectively. The isotherms are curved, suggesting that\nthe 3D Ising model is also not appropriate.\nWeanalyzedthedatawiththefollowingArrott-Noakes\nequation of state[18]:\n(H/M)1/γ= (T−TC)/T1+(M/M1)1/β(4)\n, whereT1andM1are material constants.\nFigures 3(a) and 3(b) show the modified Arrott plot\n(MAP) in the form of M1/βvs. (H/M)1/γfor UTeS and\nUSeS, respectively. The isotherms become straight if the\nappropriate values of TC,β, andγare chosen. We deter-\nmine these parameters from fits of Eq. (4) to the data\nfor 81.0 K ≤T≤89.0 K and 1.2 T ≤µ0H≤7.0 T in\nUTeS, and those for 21.4 K ≤T≤24.6 K and 0.4 T ≤\nµ0H≤3.0 T in USeS. The values of TC,β, andγare3\n/s51/s120/s49/s48/s49/s55\n/s48/s77/s49/s47/s98/s32/s91/s65/s47/s109/s93/s49/s47/s98\n/s49/s53 /s49/s48 /s53 /s48\n/s40/s72/s47/s77/s41/s49/s47/s103/s40/s98/s41/s32/s32/s85/s83/s101/s83\n/s84/s32/s61/s32/s50/s49/s46/s52/s32/s75/s32\n/s50/s52/s46/s54/s32/s75/s32/s68/s84/s32/s32\n/s61/s32/s48/s46/s56/s32/s75/s54/s120/s49/s48/s49/s54\n/s48/s77/s49/s47/s98/s32/s91/s65/s47/s109/s93/s49/s47/s98\n/s52/s48 /s50/s48 /s48\n/s40/s72/s47/s77/s41/s49/s47/s103/s40/s97/s41/s32/s32/s85/s84/s101/s83\n/s84/s32/s61/s32/s56/s49/s46/s48/s32/s75/s32\n/s56/s57/s46/s48/s32/s75/s32/s68/s84/s32/s32/s61/s32\n/s32/s32/s49/s46/s48/s32/s75\n/s49/s48/s53/s77/s32/s91/s65/s47/s109/s93\n/s48/s46/s49 /s49\n/s109/s48/s72/s32/s91/s84/s93/s84/s32/s61/s32/s56/s49/s46/s48/s32/s75/s32\n/s68/s84/s32/s32\n/s61/s32/s48/s46/s53/s32/s75\n/s56/s57/s46/s48/s32/s75/s40/s99/s41/s32/s85/s84/s101/s83\n/s56/s57/s49/s48/s53/s50/s77/s32/s91/s65/s47/s109/s93\n/s48/s46/s49 /s49\n/s109/s48/s72/s32/s91/s84/s93/s84/s32/s61/s32/s50/s49/s46/s52/s32/s75/s32\n/s68/s84/s32/s32\n/s61/s32/s48/s46/s50/s32/s75\n/s50/s52/s46/s54/s32/s75/s40/s100/s41/s32/s85/s83/s101/s83\nFIG. 3: Modified Arrott plot (MAP) of magnetic isotherms\n(a) for 81.0 K ≤T≤89.0 K in UTeS and (b) for 21.4 K\n≤T≤24.6 K in USeS. Data represented as close circles in\n(a) and (b) are analyzed with the Arrott-Noakes equation\nof state [Eq. (4)]. Solid lines show fits to the data in the\nhigher magnetic field region with a linear function. Magneti c\nfield dependencies of the magnetization in (c) UTeS and in\n(d) USeS. Bold circles indicate the critical isotherms at 85 .0\nK and 23.2 K for UTeS and USeS, respectively. Solid lines\nrepresent fits to the data represented as closed bold circles\nwith Eq. (3).\ndetermined as TC= 84.88±0.05 K,β= 0.309±0.003,\nandγ= 0.998±0.003 for UTeS, and TC= 23.09±0.03\nK,β= 0.300±0.003, and γ= 1.00±0.02for USeS. The\nparameters are shown in Table I. The data used for the\nanalyses are represented as closed circles in Figs. 3(a)\nand 3(b). The data points in the MAPs generally form\nstraight lines but those in the low magnetic field region\ndeviate from the lines. There are several reasons for the\ndeviation such as the movement of domain walls or sam-\nple inhomogeneities. In addition, there might be an error\nin the calculated value of the demagnetization factor D.\nThe origins of the deviation have been discussed in Ref.\n19, although it has not been completely understood yet.\nSolidlinesinFigs. 3(a)and3(b)representfitstothedata\nin the high magnetic field region with a linear function\nin order to obtain the spontaneous magnetic moment Ms\nand the magnetic susceptibility χ. The temperature de-\npendencies of the quantities will be used in the analysis\nwith the Kouvel-Fisher method.\nWe determine the critical exponent δfrom the critical\nisotherm at TC. The value of δis determined as δ= 4.21/s49/s48\n/s48 /s99/s32/s45/s49/s49/s120/s49/s48/s53\n/s48/s77/s115/s32/s91/s65/s47/s109/s93/s84/s67/s40/s97/s41/s32/s85/s84/s101/s83\n/s45/s49/s52\n/s48/s77/s115/s32/s40/s100/s77/s115/s47/s100/s84/s41/s45/s49/s32/s91/s75/s93\n/s56/s56 /s56/s52\n/s84/s32/s91/s75/s93/s52\n/s48/s99/s45/s49/s32/s40/s100/s99/s45/s49/s47/s100/s84/s41/s45/s49/s32/s91/s75/s93/s84/s67/s32\n/s85/s84/s101/s83/s45/s54\n/s48/s77/s115/s32/s40/s100/s77/s115/s47/s100/s84/s41/s45/s49/s32/s91/s75/s93\n/s50/s52 /s50/s50\n/s84/s32/s91/s75/s93/s50\n/s48/s99/s45/s49/s32/s40/s100/s99/s45/s49/s47/s100/s84/s41/s45/s49/s32/s91/s75/s93/s84/s67\n/s85/s83/s101/s83/s50\n/s48/s99/s32/s45/s49/s49/s120/s49/s48/s53\n/s48/s77/s115/s32/s91/s65/s47/s109/s93/s84/s67/s40/s98/s41/s32/s85/s83/s101/s83\nFIG. 4: Temperature dependencies of Ms(T) andχ−1de-\ntermined from the MAPs (upper panels) and Kouvel-Fisher\nplots (lower panels) in (a) UTeS and in (b) USeS.\n±0.04for UTeS and 4.34 ±0.04for USeS from fits to the\nisothermsat85.0KforUTeSandat23.2KforUSeS with\nEq. (3)asshowninFigs. 4(c)and4(d), respectively. The\ndata shown as closed circles are analyzed. The exponents\nβ,γ, andδshould be related by the Widom scaling law\nδ= 1+γ/β[20]. The value of δis estimated as 4.23 ±\n0.06 for UTeS and 4.33 ±0.10 for USeS using the βand\nγvalues in the MAPs. These values are consistent with\nthose determined from the critical isotherms.\nNext, we analyze the data with the Kouvel-Fisher\nmethod where the critical exponents can be determined\nmore accurately[21]. The solid lines in Figs 3(a) and\n3(b) intersect with the vertical axis at M1/β=M1/β\nsfor\nT < T Cand with the transverse axis at ( H/M)1/γ=\n(1/χ)1/γforTC< T. The values of Ms(T) andχ(T) can\nbe obtained by inserting the βandγvalues. Figures 4(a)\nand 4(b) show the temperature dependencies of Ms(T)\nandχ−1(T) for UTeS and USeS, respectively. Solid lines\nrepresent fits to the data with Eqs. (1) and (2). We de-\ntermine the critical exponents βandγwith the Kouvel-\nFisher (KF) method where temperature-dependent ex-\nponentsβ(T) andγ(T) are introduced as follows[21]:\nMs(T)[dMs(T)/dT]−1= (T−TC−)/β(T) (5)\nχ−1(T)[dχ−1(T)/dT]−1= (T−TC+)/γ(T) (6)\nThe quantities β(T) andγ(T) are equal to the critical\nexponents βandγ, respectively, in the limits T→TC\nandH→0. The slopes of Ms(T)[dMs(T)/dT]−1and4\n/s55/s120/s49/s48/s53\n/s48/s77/s47/s124/s116/s124/s32/s98/s32/s91/s65/s47/s109/s93\n/s52/s48/s48/s48 /s50/s48/s48/s48 /s48\n/s40/s109/s48/s72/s41/s47/s124/s116/s124/s40/s32/s98/s43/s103/s41/s32/s91/s84/s93/s84/s32/s60/s32/s84/s67/s84/s67/s32/s60/s32/s84/s85/s84/s101/s83 /s40/s97/s41\n/s54/s120/s49/s48/s53\n/s48/s77/s47/s124/s116/s124/s32/s98/s32/s91/s65/s47/s109/s93\n/s49/s48/s48/s48 /s48\n/s40/s109/s48/s72/s41/s47/s124/s116/s124/s40/s32/s98/s43/s103/s41/s32/s91/s84/s93/s84/s32/s60/s32/s84/s67/s84/s67/s32/s60/s32/s84/s85/s83/s101/s83/s40/s98/s41\n/s49/s48/s53/s50/s52/s54/s56/s77/s47/s124/s116/s124/s32/s98/s32/s91/s65/s47/s109/s93\n/s49/s48 /s32/s49/s48/s48/s48\n/s40/s109/s48/s72/s41/s47/s124/s116/s124/s40/s32/s98/s43/s103/s41/s32/s91/s84/s93/s84/s32/s60/s32/s84/s67\n/s84/s67/s32/s60/s32/s84\n/s50/s120/s49/s48/s53/s51/s52/s53/s54/s55/s77/s47/s124/s116/s124/s32/s98/s32/s91/s65/s47/s109/s93\n/s49/s48/s48 /s49/s48/s48/s48\n/s40/s109/s48/s72/s41/s47/s124/s116/s124/s40/s32/s98/s43/s103/s41/s32/s91/s84/s93/s84/s32/s60/s32/s84/s67\n/s84/s67/s32/s60/s32/s84\nFIG. 5: Renormalized magnetization mas afunction ofrenor-\nmalized field hfollowing Eq. (7) below and above TCfor (a)\nUTeS and for (b) USeS. Solid lines represent best-fit polyno-\nmials.\nχ−1(T)[dχ−1(T)/dT]−1-Tplots at TCyield the βand\nγvalues, respectively. The fits to the data with Eqs.\n(5) and (6) are shown as solid lines in the low panels of\nFigs. 4(e) and 4(f), respectively. The parameters are\ndetermined as β= 0.315±0.003 and γ= 0.996±0.003,\nandTC= 85.09 ±0.04 K for UTeS, and β= 0.293 ±\n0.003,γ= 0.989 ±0.003, and TC= 23.18 ±0.03 K for\nUSeS. Here, TCis defined as TC= (TC++TC−)/2.\nIt may be possible to speculate that the 3D Ising uni-\nversality class below TCis changed to the mean field one\naboveTCin UTeS and USeS. Scaling theory enables us\nto determine separately the values of γ’ forT < T Cand\nγforTC< T. A reduced equation of state close to TC\nwas predicted in the scaling theory as follows[17]:\nm=f±(h) (7)\nHere,f+andf−are regular analytical functions for\nTC< TandT < T C, respectively. The renormalized\nmagnetization mis defined as m≡ |t|−βM(µ0H,t) and\nthe renormalized field hash≡µ0H|t|−(β+γ). Two uni-\nversal curves are formed in the plot of mvs.hwhen the\ncorrect values of β,γ’,γ, andTCare chosen. The data\nin the temperature ranges t=|(T−TC)/TC|<0.05\nfor UTeS and 0.08 for USeS are shown in Figs. 5(a) and\n5(b), respectively. The analyses yield the values of TC\nand the critical exponents as TC= 85.09±0.03 K,β=\n0.318±0.002,γ′= 1.03±0.02, and γ= 1.04±0.02 in\nUTeS, and TC= 23.18 ±0.02 K,β= 0.300 ±0.002,γ′\n= 1.00±0.02, and γ= 1.02±0.02 in USeS. This result\nsuggests that the sets of the critical exponents in the two\ncompounds are common below and above TC. Note that\nthe magnetic isotherms in the forms of the mean field\ntheory and the 3D Ising model with SR exchange inter-\nactions do not form straight lines below and above TCas\nshown in Figs. 2(a)-2(d). The strongly asymmetric crit-\nical region or the change of the universality class across\nTCcan be ruled out.IV. DISCUSSION\nTable I shows the critical exponents β,γ′,γ, andδin\nUTeS and USeS, and those in mean field theory and var-\nious theoretical models with SR exchange interactions\nof a form J(r)∼e−r/b[17, 22, 23]. The exponents in\nthe uranium ferromagnetic superconductors UGe 2and\nURhGe[16], and some uranium ferromagnets URhSi[24],\nUIr[25,26], andU(Co 0.98Os0.02)Al[27,28]arealsoshown.\nThe sets of the exponents in UTeS and USeS are similar\nto those of the uranium ferromagnets. The ferromag-\nnetic states of these ferromagnets have strong uniaxial\nanisotropy. However, the critical exponents of the com-\npounds differ from those in the 3D Ising model with SR\nexchange interactions. The βvalues are relatively close\nto those of the 3D models. Meanwhile, the values of γ\nare close to unity, expected one in the mean field theory.\nWe discuss the mean-field behavior of the magne-\ntization in the uranium ferromagnetic superconductor\nUCoGe[29]. The extent of the asymptotic critical region\n∆TGwhere the mean field theory fails can be estimated\nby the Ginzburg criterion[30]. ∆ TGin three dimensions\nis given as ∆ TG/TC=k2\nB/[32π2(∆C)2ξ06][31, 32]. Here,\n∆Cis the jump of the specific heat at TCin units of\nerg·cm−3K−1andξ0is the bare correlation length. The\nvalue of ∆ TGfor UCoGe was estimated as less than 1\nmK using reported ∆ Candξ0values[16, 29, 33]. It is\nnatural that the mean field behavior of the magnetiza-\ntion is observed because most of magnetic data might be\ntaken outside the very narrow region around TC. The\nlonger magnetic correlation length may be originated\nfrom the strong itinerant character of the 5 felectrons\nin UCoGe[24]. We have previously reported the critical\nexponents in UGe 2[16]. The value of ∆ TGwas estimated\nas∼100K.Itcanbeconcludedthatthedatausedforthe\ndetermination of the critical exponents were taken inside\nthe asymptotic critical region in UGe 2. Meanwhile, it is\nimpossible to estimate ∆ TGfor UTeS and USeS since the\nmagnetic correlation length ξhas not been reported so\nfar. In this study, the data for 81.0 K ≤T≤89.0 K and\n1.2 T≤µ0H≤7.0 T in UTeS, and those for 21.4 K ≤T\n≤24.6 K and 0.4 T ≤µ0H≤3.0 T in USeS are analyzed\nto determine the critical exponents. If the data up to 7.0\nT in USeS are analyzed, they do not form straight lines\nin the MAP and nor do universal curves in the scaling\nanalysis for any values of the critical exponents. We re-\npeated the analysis and found that the upper limit of the\ncritical region is about 3.0 T for USeS. The consistency\nin the obtained exponents determined by different meth-\nods suggests the reliability of them. We conclude that\nthe data used for the analyses were collected inside the\nasymptotic critical regions of each compound.\nFigures 6(a) and 6(b) show the normalized sponta-\nneous magnetic moment Ms/M0and the magnetic sus-\nceptibility χ/Cas a function of the reduced tempera-\nture|t|(=|(T−TC)/TC|) determined from the MAPs\nin UTeS, USeS, UGe 2[16], URhGe[16], and URhSi[24].\nHere, constants M0andCare obtained from fits to5\nTABLE I: Critical exponents β,γ,γ′, andδof UTeS and USeS, and those in the mean field theory and various theoretical\nmodels with short-range (SR) exchange interactions[17, 22 , 23]. The exponents in UGe 2[16], URhGe[16], URhSi[24], UIr[25],\nand U(Co 0.98Os0.02)Al[27] are also shown. Abbreviations: RG- φ4, renormalization group φ4field theory; MAP, Modified Arrott\nplot; CI, Critical isotherm; KF, Kouvel-Fisher method.\nMethod TC(K) β γ′(T < T C)γ(TC< T) δ Reference\nMean field 0.5 1.0 3.0\nd= 2,n=1 Onsager solution 0.125 1.75 15.0 [17, 22]\nd= 3,n=1 RG- φ40.325 1.241 4.82 [17, 23]\nd= 3,n=2 RG- φ40.346 1.316 4.81 [17, 23]\nd= 3,n=3 RG- φ40.365 1.386 4.80 [17, 23]\nUTeS MAP, CI 84.88 ±0.05 0.309 ±0.003 0.998 ±0.003 4.21 ±0.04 this work\nKF 85.09 ±0.04 0.315 ±0.003 0.996 ±0.003\nScaling 85.09 ±0.03 0.318 ±0.002 1.03 ±0.02 1.04 ±0.02\nUSeS MAP, CI 23.09 ±0.03 0.300 ±0.003 1.00 ±0.02 4.34 ±0.04 this work\nKF 23.18 ±0.03 0.293 ±0.003 0.989 ±0.003\nScaling 23.18 ±0.02 0.300 ±0.002 1.00 ±0.02 1.02 ±0.02\nUGe2 Scaling, CI 52.79 ±0.02 0.329 ±0.002 1.00 ±0.02 1.02 ±0.02 4.16 ±0.02 [16]\nURhGe Scaling, CI 9.47 ±0.01 0.302 ±0.001 1.00 ±0.01 1.02 ±0.01 4.41 ±0.02 [16]\nURhSi Scaling, CI 10.12 ±0.02 0.300 ±0.002 1.00 ±0.02 1.03 ±0.02 4.38 ±0.04 [24]\nUIr MAP, CI 45.15 ±0.2 0.355 ±0.05 1.07 ±0.1 4.04 ±0.05 [25]\nU(Co0.98Os0.02)Al MAP, CI 25 0.33 1.0 4.18 [27]\nthe|t|-dependencies of Msandχwith formulas Ms(t)\n=M0|t|βandχ(t) =C/|t|γ, respectively. The data\nfor the latter three compounds are from our previous\nstudies[16, 24]. Although the spontaneous magnetiza-\ntionMsshows the critical behavior expected for the 3D\nmagnets, the magnetic susceptibility χdoes the mean-\nfield-like behavior. Many theoretical studies have been\ndone for critical phenomena of ferromagnetic transitions\nin ferromagnetic materials. However, this unusual be-\nhavior cannot be explained with previous theoretical ap-\nproaches. We discuss this issue from following six view-\npoints.\n(1) It is well known that the long-range nature of mag-\nnetic exchange interactions affects the critical behavior\nof the magnetization around TC. The theoretical values\nof the critical exponents in Table I are those of theoret-\nical models with short-range (SR) exchange interactions\nofa form J(r)∼e−r/b[17, 22, 23]. The strengthofthe ex-\nchange interaction J(r) decreases rapidly with increasing\ndistance r. When the range of the exchange interaction\nbecomes longer, the critical exponents ofthe models shift\ntoward those of the mean field theory. This problem was\nstudied by Fischer et al.with a renormalization group\napproach for systems with the exchange interaction of a\nformJ(r)∼1/rd+σ[34]. Here, σis the range of exchange\ninteraction and dis the dimension of the system. They\nshowed that the model is valid for σ <2 and derived\na theoretical formula for the exponent γ= Γ{σ,d,n}.\nHere,nis the dimension of the order parameter and the\nfunction Γ is given in Ref. 34. We calculate the criti-\ncal exponents using the formula and scaling relations for\ndifferent sets of {d:n}(d,n= 1, 2, 3) in order to\nreproduce the exponents in the uranium ferromagnets.\nHowever, there is no reasonable solution of σ.\n(2) The critical phenomenon of the magnetization isalso affected by the classical dipole-dipole interaction as\nhas been studied for rare earth metal gadolinium ( TC=\n292.7K and the spontaneousmagnetic moment ps= 7.12\nµB/Gd)[35]. Theinteractionmaynothaveastrongeffect\non critical phenomena in uranium ferromagnets since the\nstrength of the effect is proportional to p2\ns[36]. The value\nofpsis determined as 1.62 µB/U at 5.0 K for UTeS and\n1.09µB/U at 2.0 K for USeS. These are much smaller\nthanthatofGd. Moreover,the exponentsintheuranium\nferromagnets are not consistent with those of theoretical\nstudies for the critical phenomenon associated with the\nisotropic or anisotropic dipole-dipole interaction[37, 38].\n(3) We discuss the critical exponents from the view-\npointofthelocalmomentmagnetism. Theferromagnetic\nstates in the uranium ferromagnets can be regarded as\nthe anisotropic 3D Ising system or the anisotropic next-\nnearest-neighhor 3D Ising (ANNNI) system. However,\nthe critical exponents in the uranium ferromagnets are\nnot consistent with those obtained in numerical calcula-\ntions for the two systems[39, 40].\n(4) The temperature dependencies of the spontaneous\nmagnetic moment psand the magnetic susceptibility χ\nobtained analytically or numerically in the spin fluctua-\ntion theoriesarenot consistentwith thosein the uranium\nferromagnets[41, 42]. The spin fluctuation theories can-\nnot be applied to physical phenomena in the asymptotic\ncritical region.\n(5) The unconventional critical phenomenon in UGe 2\nand URhGe has been discussed by Singh, Dutta, and\nNandy with a nonlocal Ginzburg-Landau model focus-\ning on magnetoelastic interactions[43]. It was claimed\nthat their calculated results are comparable with those\nof UGe 2and URhGe. It is hoped that the almost mean-\nfield behavior of χis completely reproduced.\nThe itinerant picture of the 5 felectrons is basically6\n/s49/s48/s49/s48/s48/s99/s47/s67\n/s48/s46/s48/s49 /s48/s46/s49\n/s124/s116/s124/s32/s40/s61/s32/s124/s40/s84/s45/s84/s67/s41/s47/s84/s67/s124/s41/s51/s68/s32/s72/s101/s105/s115/s101/s110/s98/s101/s114/s103\n/s32/s32/s32/s51/s68/s32/s88/s89\n/s32/s32/s32/s32/s32/s51/s68/s32/s73/s115/s105/s110/s103\n/s77/s70/s32/s40/s98/s41/s48/s46/s48/s49/s48/s46/s49/s49/s77/s115/s47/s77/s48\n/s48/s46/s49 /s48/s46/s48\n/s124/s116/s124/s32/s40/s61/s32/s124/s40/s84/s45/s84/s67/s41/s47/s84/s67/s124/s41/s32/s50/s68/s32/s73/s115/s105/s110/s103\n/s51/s68/s32/s73/s115/s105/s110/s103\n/s51/s68/s32/s88/s89\n/s51/s68/s32/s72/s101/s105/s115/s101/s110/s98/s101/s114/s103\n/s77/s70\n/s32/s85/s71/s101/s50\n/s32/s85/s82/s104/s71/s101\n/s32/s85/s82/s104/s83/s105/s32\n/s32/s85/s84/s101/s83\n/s32/s85/s83/s101/s83/s32/s40/s97/s41\nFIG. 6: (a) Normalized spontaneous magnetic moment\nMs/M0and (b) magnetic susceptibility χ/Cas a function of\nreduced temperature |t|(=|(T−TC)/TC|) for UTeS, USeS,\nUGe2[16], URhGe[16], and URhSi[24]. Dotted lines indicate\ncalculated curves for various theoretical models.\nappropriate to describe the ferromagnetism in uranium\nintermetallic compounds[44]. Meanwhile, the dual na-\nture of the 5 felectrons in UGe 2has been experimentally\nsuggestedin the Muonspin rotationspectroscopy[45,46].\nThe concept of the duality of the 5 felectrons has been\na basis in theoretical studies for the superconductivity in\nUGe2[47], URhGe[48], and UPd 2Al3[49]. Previously, we\npointed out relevance between the dual nature of the 5 f\nelectrons and the novel critical behavior of the magne-\ntization in UGe 2, URhGe, and URhSi[16, 24]. A novel\ncritical phenomenon can be expected due to two corre-\nlation lengths of the localized and itinerant components\nof the 5felectrons and a Hund-type coupling between\nthem. However, this scenario cannot be applied to UTeS\nand USeS with the localized 5 felectrons. The soft X-\nray photoelectron spectroscopy showed that the 5 flevelis situated about 750 meV below the Fermi energy in\nUTeS[50]. Thepresentstudyshowsthatthenovelcritical\nphenomenon of the ferromagnetic transition is observed\nnot only in the uranium intermetallic compounds where\nthe ferromagnetism is carried by the itinerant 5 felec-\ntrons but also in UTeS and USeS with the localized 5 f\nelectrons. The large negative magnetoresistance in UTeS\nand USeS, and the uranium ferromagnetic superconduc-\ntivity in UGe 2and URhGe are observed in the vicinity\nof the ferromagnetism characterized by the novel critical\nexponents.\nIt has been long thought that the p-wave supercon-\nductivity in the uranium ferromagnetic superconductors\nis driven by longitudinal spin fluctuations developed in\nthe vicinity of the ferromagnetic state described with the\n3D Ising model. Meanwhile, our studies suggest that\nthe ferromagnetic correlation between the 5 felectrons\ndiffers from that of the 3D Ising system in the ura-\nnium ferromagnets including the superconductors. Re-\ncent uniaxial experiment on URhGe and its theoretical\ninterpretation suggest that a pairing mechanism other\nthan that driven by Ising-type longitudinal fluctuations\ntake a certain role for the superconductivity[51, 52].\nIt would be interesting to study the dynamical mag-\nnetic property of the uranium dichalcogenides. It was\nclaimed that the superconductor UTe 2is on the verge of\nthe ferromagnetism since the critical exponents are close\nto values expected for a ferromagnetic quantum critical\npoint[6,53]. Magneticfluctuationsobservedinmuonspin\nrelaxation/rotation ( µSR) measurements on UTe 2may\ntake an important role for anomalous behaviors of the\nupper critical field Hc2or the unconventional supercon-\nductingorderparameterwithpointnodessuggestedfrom\nthe thermal transport, heat capacity and magnetic pen-\netration depth measurements[54–56]. The group of the\nuranium dichalcogenides would be an interesting plat-\nform for the study of both the large magnetoresistance\nand the superconductivity in terms of the ferromagnetic\ncorrelation between the 5 felectrons.\nV. SUMMARY\nIn summary, we study the novel critical behavior of\nmagnetization in uranium semimetal UTeS and semicon-\nductor USeS exhibiting a large transverse magnetoresis-\ntance around the ferromagnetic transition temperatures.\nThe critical exponents in the two compounds differ from\nthoseinthe3DIsingmodelwithshort-rangeexchangein-\nteractions in spite of uniaxial magnetic anisotropy in the\nferromagnetic states. The critical exponents are similar\ntothoseinuraniumferromagneticsuperconductorsUGe 2\nandURhGe, andsomeuraniumferromagnetsURhSi, UIr\nand U(Co 0.98Os0.02)Al. The universalityclassforthe fer-\nromagnetic transition in UTeS and USeS may belong to\nthe same one for the uranium ferromagnets. The novel\ncritical phenomenon of the ferromagnetic transition ap-\npears not only in the uranium intermetallic ferromag-7\nnets with the itinerant 5 felectrons but also in UTeS\nand USeS with the localized 5 felectrons. There is simi-\nlarity between the large magnetoresistance in UTeS and\nUSeS, and the superconductivity in UGe 2and URhGe\nof their closeness to the ferromagnetism characterized by\nthe novel critical exponents.VI. ACKNOWLEDGMENTS\nThis work was supported by Japan Society for the\nPromotion of Science (JSPS) KAKENHI Grant No.\nJP16K05463, JP16KK0106, and JP17K05522.\n[1] J. A. Mydosh and P. M. Oppeneer, Rev. Mod. Phys. 83,\n1301 (2011).\n[2] C. Pfleiderer, Rev. Mod. Phys. 81, 1551 (2009).\n[3] G. R. Stewart, Advances in Physics 66, 75 (2017).\n[4] A. D. Huxley, Physica C 514, 368 (2015).\n[5] A. Mauger and C. Godart, Phys. Rep. 141, 51 (1986).\n[6] S. Ran, C. Eckberg, Q-P. Ding, Y. Furukawa, T. Metz,\nS. R. Saha, I-L. Liu, M. Zic, H. Kim, J. Paglione, and N.\nP. Butch, Science 365, 684 (2019).\n[7] W. Suski, T. Gibi´ nski, A. Wojakowski, and A. Czopnik,\nPhys. Stat. Solidi (a) 9, 653 (1972).\n[8] L. Shlyk and R. Troc, Physica B 262, 90 (1999).\n[9] R. Tro´ c, D. Kaczorowski, L. Shlyk, M. Potel, and H.\nNo¨ el, J. Phys. Chem. Solids. 55, 815 (1994).\n[10] S. Ikeda, H. Sakai, T. D. Matsuda, D. Aoki, Y. Homma,\nE. Yamamoto, A. Nakamura, Y. Shiokawa, Y. Haga, and\nY.¯Onuki, J. Phys. Soc. Jpn. 75, 124706 (2006).\n[11] S. Ikeda, H. Sakai, D. Aoki, Y. Homma, E. Yamamoto,\nA. Nakamura, Y. Shiokawa, Y. Haga, and Y. ¯Onuki, J.\nPhys. Soc. Jpn. 78, 114704 (2009).\n[12] V. Hutanu, H. Deng, S. Ran, W. T. Fuhrman, H. Thoma,\nand N. P. Butch, arXiv:1905.04377.\n[13] Y. Haga, E. Yamamoto, and S. Ikeda, unpublished.\n[14] M. B. Salamon, Rev. Mod. Phys. 73, 583 (2001).\n[15] N. Tateiwa, Y. Haga, E. Yamamoto, and Z. Fisk, JPS\nConf. Proc. 3, 011086 (2014).\n[16] N. Tateiwa, Y. Haga, T. D. Matsuda, E. Yamamoto, and\nZ. Fisk, Phys. Rev. B 89, 064420 (2014).\n[17] V. Privman, P. C. Hohenberg, and A. Aharony, in Phase\nTransitions and Critical Phenomena , edited by C. Domb\nand J. L. Lebowitz (Academic, New York, 1991), p. 1.\n[18] A. Arrott and J. E. Noakes, Phys. Rev. Lett. 19, 786\n(1967).\n[19] M. Seeger and H. Kronm¨ uller, J. Magn. Magn. Mater.\n78, 393 (1989).\n[20] B. Widom, J. Chem. Phys. 43, 3892 (1965).\n[21] J. S. Kouvel and M. E. Fisher, Phys. Rev. 136, A1626\n(1964).\n[22] M. E. Fisher, Rev. Mod. Phys. 46, 597 (1974).\n[23] J. C. Le Guillou and J. Zinn-Justin, Phys. Rev. B 21,\n3976 (1980).\n[24] N. Tateiwa, Y. Haga, and E. Yamamoto, Phys. Rev. B\n99, 094417 (2019).\n[25] W. Knafo, C. Meingast, S. Sakarya, N. H. v. Dijk, Y.\nHuang, H. Rakoto, J-M. Broto, and H. v. L¨ ohneysen, J.\nPhys. Soc. Jpn. 78, 043707 (2009).\n[26] S. Sakarya, N. H. van Dijk, A. de Visser, E. Br¨ uk, Y.\nHuang, J. A. A. J. Perenboom, H. Rakoto, and J. -M.\nBrot, J. Magn. Magn. Mater. 310, 1564 (2007).\n[27] M. Maeda, A. V. Andreev, K. Shirasaki, T. Yamamura,\nand N. Kimura, J. Phys. Soc. Jpn. 87, 094713 (2019).\n[28] A. V. Andreev, K. Shirasaki, J. ˇSebek, J. Vejpravov´ a, D.\nI. Gorbunov. L. Havela, S. Daniˇ s, and T. Yamamura, J.Alloys Compd. 681, 275 (2016).\n[29] N. T. Huy, A. Gasparini, D. E. de Nijs, Y. Huang, J. C.\nP. Klaasse, T. Gortenmulder, A. de Visser, A. Hamann,\nT. G¨ orlach, and H. v. L¨ ohneysen, Phys. Rev. Lett. 99,\n067006 (2007).\n[30] V. L. Ginzburg, Sov. Phys. Solid State 2, 1824 (1961).\n[31] P. M. Chaikin and T. C. Lubensky, Principles of Con-\ndensed Matter Physics (Cambridge University Press,\nCambridge, 1995).\n[32] E. A. Yelland, S. J. C. Yates, O. Taylor, A. Griffiths, S.\nM. Haydem, and A. Carrington, Phys. Rev. B 72, 184436\n(2005).\n[33] C. Stock, D. A. Sokolov, P. Bourges, P. H. Tobash, K.\nGofryk, F. Ronning, E. D. Bauer, K. C. Rule, and A. D.\nHuxley, Phys. Rev. Lett. 107, 187202 (2011).\n[34] M. E. Fisher, S-k. Ma, and B. G. Nickel, Phys. Rev. Lett.\n29, 917 (1972).\n[35] S. Srinath, S. N. Kaul, and H. Kronm¨ uller, Phys. Rev. B\n59, 1145 (1999).\n[36] M. E. Fisher and A. Aharony, Phys. Rev. Lett. 30, 559\n(1973).\n[37] E. Frey and F. Schwabl, Phys. Rev. B 43, 833 (1991).\n[38] E. Frey and F. Schwabl, Phys. Rev. B 42, 8261 (1990).\n[39] M. A. Yurishchev, Phys. Rev. B 50, 13533 (1994).\n[40] A. K. Murtazev and Zh. G. Ibaev, J. Exp. Theor. Phys.\n113, 106 (2011).\n[41] T. Moriya, Spin Fluctuations in Itinerant Electron Mag-\nnetism(Springer-Verlag, New York, 1985).\n[42] Y. Takahashi, Spin Fluctuations Theory of Itinerant\nElectron Magnetism (Springer-Verlag, New York, 2013).\n[43] R. Singh, K. Dutta, and M. K. Nandy, Phys. Rev. E 95,\n012133 (2017).\n[44] N. Tateiwa, J. Posp´ ıˇ sil, Y. Haga, H. Sakai, T. D. Mat-\nsuda, andE. Yamamoto, Phys.Rev.B 96, 035125 (2017).\n[45] A. Yaouanc, P. Dalmas de R´ eotier, P. C. M. Gubbens,\nC. T. Kaiser, A. A. Menovsky, M. Mihalik, and S. P.\nCottrell, Phys. Rev. Lett. 89, 147001 (2002).\n[46] S. Sakarya, P. C. M. Gubbens, A. Yaouanc, P. Dalmas\nde R´ eotier, D. Andreica, A. Amato, U. Zimmermann, N.\nH. van Dijk, E. Br¨ uck, Y. Huang, and T. Gortenmulder,\nPhys. Rev. B 81, 024429 (2010).\n[47] R. Tro´ c, Z. Gajek, andA. Pikul, Phys. Rev.B 86, 224403\n(2012).\n[48] K. Hattori and H. Tsunetsugu, Phys. Rev. B 87, 064501\n(2013).\n[49] P. Thalmeier, Eur. Phys. J. B 27, 29 (2002).\n[50] Y. Takeda, T. Okane, T. Okochi, M. Kobayashi, S. Fuji-\nmori, Y. Saito, H. Yamagami, A. Fujimori, S. Ikeda, H.\nSakai, and Y. Haga, unpublished.\n[51] D. Braithwaite, J-P. Brison, J. Flouquet, G. Knebel, A.\nNakamura, and A. Pourret, Phys. Rev. Lett. 120, 037001\n(2018).\n[52] V. P. Mineev, Phys. Rev. B 96, 104501 (2017).8\n[53] T. R. Kirkpatrick and D. Belitz, Phys. Rev. B 91, 214407\n(2015).\n[54] S. Sundar, S. Gheidi, K. Akintola, A. M. Cˆ ot´ e, S. R.\nDunsiger, S. Ran, N. P. Butch, S. R. Saha, J. Paglione,\nand J. E. Sonier, arXiv:1905.06901\n[55] S. Ran, I-L. Liu, Y. S. Eo, D. J. Campbell, P. Neves, W.T. Fuhrman,S.R.Saha, C.Eckberg, H.Kim, J.Paglione,\nD. Graf, J. Singleton, and N.P. Butch, arXiv:1905.04343.\n[56] T. Metz, S. Bae, S. Ran, I-L. Liu, Y. S. Eo, W. T.\nFuhrman, D. F. Agterberg, S. Anlage, N. P. Butch, and\nJ. Paglione, arXiv:1908.01069." }, { "title": "1807.10387v1.Phase_separation_and_proximity_effects_in_itinerant_ferromagnet_superconductor_heterostructures.pdf", "content": "Phase separation and proximity e\u000bects in itinerant ferromagnet - superconductor\nheterostructures\nC. Martens,1A. Bill,2,\u0003and G. Seibold1,y\n1Institut F ur Physik, BTU Cottbus{Senftenberg, PBox 101344, 03013 Cottbus, Germany\n2Department of Physics & Astronomy, California State University Long Beach, Long Beach, CA 90840, USA\n(Dated: November 28, 2021)\nHeterostructures made of itinerant ferromagnets and superconductors are studied. In contrast\nto most previous models, ferromagnetism is not enforced as an external Zeeman \feld but induced\nin a correlated single-band model (CSBM) that displays itinerant ferromagnetism as a mean-\feld\nground state. This allows us to investigate the in\ruence of an adjacent superconducting layer on\nthe properties of the ferromagnet in a self-consistent Bogoliubov-de Gennes approach. The CSBM\ndisplays a variety features not present in the Zeeman exchange model that in\ruence the behavior\nof order parameters close to the interface, as e.g. phase separation and the competition between\nmagnetism and superconducting orders.\nPACS numbers: 74.45+c, 75.70.Cn, 74.78.Fk, 75.30.Cr\nI. INTRODUCTION\nSuperconductivity and ferromagnetism were long be-\nlieved to be mutually exclusive phases because in a\nconventional superconductor the spin part of the wave-\nfunction is a singlet which is easily broken by the strong\nmagnetization of the ferromagnet. However, there are sit-\nuations where the coexistence is possible as in an itiner-\nant ferromagnet where the spin-up and spin-down bands\nare split by an e\u000bective exchange \feld. Following the idea\nof Fulde, Ferrell, Larkin and Ovchinnikov1one can con-\nstruct a Cooper pair which is composed of two electrons\nwith opposite spins and momenta,but acquires a \fnite\ntotal momentum due to the exchange \feld splitting and\nresulting di\u000berent Fermi momenta. Another possibility\nis the formation of 'spin triplet' Cooper pairs where the\nassociated orbital part is antisymmetric either in the ex-\nchange of the electrons position or the time coordinates,\nthe latter being called 'odd frequency pairing'.2\nAn ideal playground to study these possibilities and\nmany related questions are heterostructures composed\nof superconducting and magnetic layers. Besides their\nrelevance for studying fundamental properties of the in-\nterplay between superconductivity and magnetism such\nsystems have also a strong relevance for applications in\nspintronics due for example to the possibility of acting as\nspin-polarized current sources.3\nThe attractive feature of nanohybrid structures is that\nthe two phases are spatially separated and interact solely\nvia the proximity e\u000bect. In a S/F/S junction (S being\na superconductor, F a ferromagnet) Cooper pairs tun-\nnel through the S/F interfaces and, for a thin enough\nferromagnetic layer, may realize a Josephson junction.\nFirst studies on such systems where conducted by\nBuzdin and collaborators4who predicted for example the\nexistence of \u0019-junctions. Here the oscillatory behavior\nof the superconducting state wave function leaking into\nthe ferromagnet may lead to a reversal of the Joseph-\nson current (the \u0019\u0000state) with respect to the ordinaryJosephson e\u000bect (correspondingly called the 0 \u0000state) if\nthe thickness of the ferromagnet is chosen such that there\nis a sign change of the wave function on either side of the\njunction.\nIn a series of papers Bergeret, Efetov and Volkov,5\npredicted that under certain conditions, such as for ex-\nample the presence of inhomogeneities in the magnetiza-\ntion, a triplet component with non-zero spin projection\nalong the quantization axis, m=\u00061 of the pair amplitude\nmight arise, even though the Cooper pairs generated in\nthe superconducting parts of the junction are singlets.\nA similar conclusion was reached in Ref. 6. Singlet and\nspin-zero triplet projection pairs ( m= 0) entering a fer-\nromagnet are subject to the pair breaking e\u000bect and are\nthus short-range components decaying exponentially over\na few nanometers, except at every change of direction of\nthe magnetization where they are regenerated through\nthe cascade e\u000bect.7,8In contrast, non-zero spin projection\ntriplet components are comparatively long range because\nuna\u000bected by magnetism and have in fact been observed\nin a number of experiments (see e.g. Refs. 9{16).\nThe creation of such long-ranged triplet components\nout of a singlet superconductor requires a rotation of the\nmagnetization which can be either realized via magnetic\nmultilayers,11,13conical magnets as Ho10,11or Heusler\nalloys.17On the theoretical side the conditions for which\na long range triplet component of the order parameter\ncan be observed in superconductor { magnetic nanojunc-\ntions has been worked out in some detail in the di\u000busive\nregime (i.e. on the basis of Eilenberger and/or Usadel\nequations; see e.g. Refs.5, 7, 8, 18{22 and references\ntherein) and in the clean limit within the Bogoliubov-de\nGennes (BdG) approach (see e.g.Refs. 23{34).\nMore recently other sources for the generation of a\nlong-ranged triplet component have been analyzed, in\nparticular it has been shown that the physical mecha-\nnism of the singlet- triplet conversion can be linked to\nthe local SU(2) invariance of magnetized systems with\nspin-orbit interaction.35\nIn most of the studies of S/F junctions a strong fer-arXiv:1807.10387v1 [cond-mat.supr-con] 26 Jul 20182\nromagnet is considered with a N\u0013 eel temperature much\nlarger than Tcso that the in\ruence of superconductiv-\nity on the magnetism is negligible. However, it was\nsuggested5,36that under certain conditions for a ferro-\nmagnetic \flm attached to a S it is favorable to be in\na 'cryptoferromagnetic state',37i.e. a segregation of the\nferromagnet into small-size domains, smaller than the su-\nperconducting coherence length.\nIn this paper we aim to investigate the in\ruence of S on\nthe magnetism in a S/F system in the clean limit on the\nbasis of the Bogoliubov-de Gennes approach. We will\ndescribe both the superconductor and the ferromagnet\non equal footing i.e. the latter is not implemented by an\nexternal exchange \feld but also described by an itinerant\nmodel with ferromagnetic exchange Jbetween the charge\ncarriers (cf. also Refs. 38 and 39). Our model is therefore\nimplemented on a \fnite lattice in contrast to previous\ncontinuum studies of proximity e\u000bects within the BdG\napproach.28{34\nThe approach has the advantage that it allows for the\nstudy of both proximity e\u000bects, S order parameter in the\nF and ferromagnetic order parameter in the S as a func-\ntion of the exchange J, thus covering the range from soft\nto hard F's. Moreover, we will investigate the inverse\nproximity e\u000bect, in particular the weakening of the fer-\nromagnetic order in the F close to the interface.\nPrevious studies of heterostructures on discrete lattices\nwithin BdG have addressed the subgap conductance at\nFS interfaces,23the decay of d-wave correlations inside\nthe F within an extended Hubbard model,24or the gen-\neration of parallel spin triplets in conical magnetizations\nsuch as Holmium.26\nOur paper is organized as follows. The next section\noutlines the system, model and approximations of our\ninvestigations. The resulting charge density, magnetiza-\ntion and order parameter pro\fles together with the cor-\nresponding spectral properties are discussed in Secs. III\nand IV. Finally, the \fndings are summarized in Sec. V.\nII. MODEL AND FORMALISM\nWe consider a two-dimensional superconducting sys-\ntem sandwiched between two ferromagnetic layers as\nshown in Fig. 1. We use periodic boundary conditions\nalong x- and y-directions so that the S system is actually\nconnected to the same ferromagnet on the left and right.\nTranslational invariance is assumed along the y-direction.\nThe calculations are done at zero temperature.\nOur model hamiltonian reads\nH=H0+HU+Hint; (1)\nFSFxyLSLxLyFIG. 1. Sketch of the SF bilayer proximity system with pe-\nriodic boundary conditions along x- and y-directions. Lx;y=\nNx;yawhere adenotes the lattice constant which in the fol-\nlowing is set to a\u00111.\nwhere\nH0=X\nij;\u001btijcy\ni;\u001bcj;\u001b+X\ni;\u001b(Vloc\ni\u0000\u0016)cy\ni;\u001bci;\u001b\n+X\ni;\u001bhi;z\u001bcy\ni;\u001bci;\u001b (2)\nis the single-particle part composed of the kinetic energy\nand a local term. The latter consists of the chemical po-\ntential\u0016and a local energy Vloc\niwhich is implemented\nin order to account for the local orbital energy but also\nto tune the charge density in the S layer. We consider\nelectrons on a Nx\u0002Nysquare lattice and c(y)\ni;\u001bannihi-\nlates (creates) a particle with spin \u001b=\u00061 at site Ri;\nthe site index is a two-dimensional vector, i= (ix;iy).\nThe hopping matrix element of the \frst term is set to\ntij=\u0000t <0 for nearest neighbors and tij= 0 other-\nwise. We also consider a constant Zeeman term \u0018hi;z\nto describe the local magnetization at site i. This com-\nponent of the Hamiltonian is used below to compare the\nresults for the itinerant F system studied in this paper,\nwith previous investigations where F was modeled with\nan external exchange \feld h.\nThe second part of Eq. (1),\nHU=X\niUini;\"ni;#; (3)\nis a local interaction with ni;\u001b=cy\ni;\u001bci;\u001b. Within the\nsuperconducting layers we take Ui\u0011US<0, i.e. an\nattractive Hubbard interaction, in order to support sin-\nglet superconductivity. In the ferromagnetic layers Ui\narises as the \frst term in an expansion of the long-range\nCoulomb interaction in Wannier functions and therefore\nUi\u0011UF\u00150 in these regions.\nFinally, we model the magnetic interaction in the fer-\nromagnetic layer as\nHint=\u0000X\nhijiJij[sisj+ninj]; (4)\nwhereni=ni;\"+ni;#,s\u000b\ni=P\n\u001b\u001b0cy\ni;\u001b\u001c\u000b\n\u001b\u001b0ci;\u001b0with the\nPauli matrices \u001c\u000b, andhijilimits the summations to\nnearest neighbor sites. The interaction Eq. (4) orig-\ninates from the nearest neighbor Coulomb interaction3\ntermsJ=hijj1=rjjii=hiij1=rjjji>0 and has been\nderived in Refs. 40{42 from the expansion of the long-\nrange Coulomb interaction in Wannier functions. In the\nexpansion we have neglected a density dependent 'corre-\nlated hopping' term arising from Coulomb matrix con-\ntributionshiij1=rjijibetween sites RiandRj, which\nhas been shown to further stabilize ferromagnetism.43,44\nBased on a variational Ansatz it has been argued that the\nmodel explains the occurrence of weak metallic ferromag-\nnetism in materials with a partially \flled nondegenerate\nband, as in Sc 3In45and metallic hydrogen.46Note, how-\never, that the model does not capture ferromagnetism\nin transition metals, where magnetism is usually gener-\nated from localized electrons.43,46Nevertheless, as shown\nbelow, the theory contains many features that are also\npresent in double exchange models. No speci\fc material\nis considered here and Eq. (4) is chosen as a convenient\nway to describe itinerant ferromagnetism.\nIn the heterostructure of Fig. 1, the ferromagnetic cou-\nplingJijis only \fnite in the F layer. Furthermore, mag-\nnetism and charge density solely vary along the xdirec-\ntion and are constant along the ydirection. Therefore,\nEq. (4) is rewritten in the form\nHint=\u0000X\ni=(ix;iy)Jx\ni[sisi+^x+nini+^x] (5)\n\u0000X\ni=(ix;iy)Jy\ni[sisi+^y+nini+^y]:\nIn the calculations of later sections it is assumed that\nJx\ni=Jy\ni=Jis a constant within the F layer since a\nmoderate anisotropy did not seem to a\u000bect the results\nsigni\fcantly.\nSince the system is translationally invariant along the\nydirection we perform the corresponding Fourier trans-\nform\nci;\u001b=1p\nNyX\nixcix;\u001b(ky) exp(\u0000ikyiy); (6)\nso that the kinetic term Eq. (2) reads ( t>0)\nH0=\u0000tX\nix;ky;\u001bh\ncy\nix;\u001b(ky)cix+1;\u001b(ky) +h:c:i\n(7)\n+X\nix;ky;\u001b[\u00002tcos(ky)\u0000\u0016]cy\nix;\u001b(ky)cix;\u001b(ky):\nWe apply the transformation, Eq. (6), to the inter-\naction terms, Eqs. (3-5). We then approximate these\nterms in mean-\feld. This includes the anomalous singlet\n(Gor'kov) correlations f0(i) =hci;#ci;\"ithat are induced\nin the S regions where Ui<0 but leak into the F layer\ndue to the proximity e\u000bect. The problem then can be\ndiagonalized by means of the Bogoliubov-Valatin trans-\nformation\ncix;\u001b(ky) =X\nph\nuix;\u001b(p;ky)\rp;ky\u0000\u001bv\u0003\nix;\u001b(p;ky)\ry\np;kyi\n;\n(8)and the integer plabels the eigenvalue. In-\ntroducing the basis vector ~\tn(p;ky) =\n[un;\"(p;ky);un;#(p;ky);vn;\"(p;ky);vn;#(p;ky)] one has to\nsolve the following eigenvalue problem for each value of\nky\nHij(ky)~\tj(p;ky) =\"p(ky)~\ti(p;ky); (9)\nwhere the hamiltonian is composed of a local and an\nintersite part\nHixjx(ky) =Tix;jx(ky) (\u000ejx;ix+1+\u000ejx;ix\u00001)+Vix(ky)\u000ejx;ix:\n(10)\nThe explicit structure of these operators is given in ap-\npendix A.\nIII. FERROMAGNETIC SYSTEM\nBefore presenting the results for the SF heterostructure\nwe brie\ry discuss the homogeneous magnetic system, i.e.\nUi\u0011UF>0 in Eq. (3) and Jx\ni=Jy\ni\u0011Jin Eq.\n(5). This case is instructive for the later analysis of the\ncompetition between F and S in the interface regions. A\nmore extensive discussion of the magnetic system can be\nfound in Refs. 40{42, and 46. Details of the calculations\nare provided in appendix B.\nThe chosen values of parameters, UF=tandJ=t, are\ngeneric but describe realistic systems. For example, UF\nis up to the order of the bandwidth, while Jis typically\nless thanUF.40Calculations are performed on a lattice\nwithNx\u0002Ny= 420\u0002420 sites.\nThe model already displays rich physics for a sin-\ngle magnetic layer. Depending on the parameter values\nthe system is paramagnetic, ferromagnetic, antiferromag-\nnetic or shows electronic phase separation.40{42,46\nPanels (a,b) of Fig. 2 report the ground state energy\nE(q) for a spiral modulation Si=S0exp (iq\u0001Ri) as a\nfunction of the spiral wave-vector qwhich is taken along\nthe diagonal direction ( qx=qy).S0is a variational pa-\nrameter.\nAt half-\flling and UF=t> 0,J=t= 0 the system shows\nantiferromagnetic spin-density wave order [ q=QAF=\n(\u0019;\u0019)] which due to perfect nesting occurs for in\fnitesi-\nmally small values of the repulsive interaction UF=t.47\nUpon increasing the ferromagnetic exchange J=t> 0 a\nsecond energy minimum develops at q= (0;0) which\nabove some critical J=tthat depends on UF=tcorre-\nsponds to the ferromagnetic ground state [cf. Fig. 2(a)].\nAs can be seen from panel (b) of Fig. 2 the same holds for\ndoping away from half-\flling where for su\u000eciently large\nrepulsionUF=tandJ=t= 0 the commensurate AF is re-\nplaced by a spiral, but with some incommensurate mod-\nulation Qspiral = (q;q). In the regime of small doping\n(n\u001c1) and small UF=t(<1=N(EF)) the system would\nbe a paramagnet for J=t= 0 and ferromagnetism can\nbe induced above some critical J=t. The corresponding4\n01 2 3 qx=qy-1.7-1.65-1.6-1.55-1.5E(q)/N [t]0\n1 2 3 qx=qy-1.8-1.75-1.7-1.65-1.6-1.55E(q)/N [t]0.2\n0.4 0.6 0.8 1 n-2-1.5-1-0.50chem. pot. [t]0\n2 4 6 8 U\nF/t00.20.40.60.8J [t]a) n=1, UF/t=5J/t=0.1\nJ/t=0.15J/t=0.2J/t=0.2 J/t=0.1\nb) n=0.9, UF/t=5J/t=0.15\nc)\nferro\npara\nd) UF/t=2J/t=0.7\nJ/t=0.5n=0.2n=0.6\nFIG. 2. Ground state energy vs. spin-spiral modulation vec-\ntorq(Si=S0exp (iq\u0001Ri)) along the diagonal direction of\nthe Brillouin zone at half-\flling (a) and density n= 0:9 (b).\nNote that for other UF=tresults are qualitatively similar but\nenergy variations decrease with decreasing UF=t. (c) Phase\ndiagram in the ( UF=t; J=t )-plane for density n= 0:2. (d)\nChemical potential vs. density. The horizontal dotted lines\nare determined by the Maxwell construction. The compress-\nibility diverges at the local extrema.\nphase diagram is displayed in Fig. (2c) for concentrations\nn= 0:2 andn= 0:6. Upon increasing UF=tthe transi-\ntion line approaches the value for the standard Stoner\ncriterionUF= 1=N(EF) atJ= 0 (N(EF) is the density\nof states at the Fermi energy EF).\nIn Figure 2(d) we also demonstrate that the model\nhas an instability region with respect to phase separa-\ntion which can be deduced from the dependence of the\nchemical potential \u0016on the density n. The compressibil-\nity\u0014=@n=@\u0016 diverges at the local extrema of \u0016(n) and\nbecomes negative in between. The phase separation re-\ngion innis determined by the Maxwell construction (dot-\nted horizontal line). The two curves in Fig. 2d indicate\nthat the phase separation region decreases with decreas-\ningJ=t. Note that the occurrence of phase separation is\nnota peculiar feature of the present model. This phe-\nnomenon also appears in double exchange models that\nare for example used for the description of magnetism in\nmanganites (cf. Ref. 48 and references therein).\nIV. SUPERCONDUCTING-MAGNETIC\nHETEROSTRUCTURE\nThe results of this section have been obtained on lat-\ntices with 120\u000280 sites and periodic boundary conditions\nin both directions. In the S region (40 \u0014ix\u001480) singlet\nsuperconductivity is generated with a negative US=\u00002t.\nFor this value the coherence length can be estimated as\u0018S\u00194 in units of the lattice spacing, i.e. much smaller\nthan the linear size of the system.\nThe remaining sites pertain to the F region with local\nonsite repulsion UF>0 and ferromagnetic exchange in-\nteractionsJx\ni\u0011Jx,Jy\ni\u0011JyandJx=Jy=J. This\ndescription of the F layer will be referred to as the cor-\nrelated ferromagnetic model (CFM). For comparison, we\nalso use an exchange \feld hzto model the F. The latter\nis referred to as the exchange \feld model (EFM).\nSince we consider an itinerant F we treat supercon-\nductivity and magnetism on equal footing. Hence, we\npresent in this section results for the charge density, the\nmagnetization and pair correlations in both the F and\nthe S.\nIn addition to the normal proximity e\u000bect, two distinct\nphenomena appear in these hybrid structures: the inverse\nproximity e\u000bect and phase separation. In the \frst, the\nS correlations suppress the magnetization inside the F\nnear the SF interface. In general, ni6=miin such sit-\nuation; the F is nowhere fully polarized. Moreover, the\ncoupling between magnetic and charge degrees of free-\ndom leads to a concomitant reduction of the nnear the\nSF interface. By contrast, when the system undergoes\nphase separation, the system is fully polarized ( ni=mi)\ndeep in the F and the superconducting state is a\u000bected\nby the itinerant electrons of the F.\n1. Charge density, magnetization and pair correlations\nFigure 3 reports the charge density n(panel a), mag-\nnetizationm(panel b), and singlet pair correlations f0\n(panel c) in the heterostructure for varying exchange con-\nstantJ=t. Since this parameter also in\ruences the local\nHartree-potential in the F layers a change of Jalters the\ncharge distribution between the S and F regions. To be\nable to compare results for di\u000berent values of Jwe there-\nfore adjust the local potential Vlocin the S regions in\nsuch a way that the charge density is the same for all\nJ=t-values deep inside the S layer; hence, in panel (a) of\nFig. 3 the charge densities noverlap in S for all J=t. Since\nmost of the physics occurs close to the interface between\nthe S and the F layer, Fig. 4 zooms into this region to\nshow the behavior of ni,miandf0(i). In addition, the\n\fgure also reports the result for the EFM (dashed lines)\nforhi;z= 3t. The value of hzis \fxed in such a way that\nit reproduces the same magnitude of the magnetization\nas the CFM deep inside the F region for J=t= 0:5.\nFrom Figs. 3 and 4 one can distinguish three di\u000ber-\nent regimes. For the parameters of the system, these are\nJ=t.0:3, 0:3.J=t.0:55 andJ=t&0:55. At low\nJ=t(.0:3) ferromagnetism disappears; this paramag-\nnetic regime was discussed in the previous section. Above\nthis transition the inverse proximity e\u000bect regime [black\nopen circles and red squares in panels (a-d)] is e\u000bective.\nThe S correlations completely suppress the magnetiza-\ntion in F over a signi\fcant distance from the interface.\nThis regime is also characterized by a partial depletion of5\n0,20,40,60,81charge niJ/t=0.35\nJ/t=0.4\nJ/t=0.5\nJ/t=0.6\n0\n0,20,40,60,8magnet. miJ/t=0.7\n20\n40 60 80 100 120 x-position x\ni00,020,040,06singlet OP f0(i)90\n100110 a)b)\nc)\nFF S \nFIG. 3. Charge density (a), magnetization (b) and singlet\npair correlations (c) as a function of xand various values of\nthe exchange coupling J=tin the ferromagnetic region. For\nall cases the local potential Vlocis adjusted in such a way as\nto obtain a similar charge density deep in the S layer. The\nthin dashed green line reports the result within the EFM for\nhi;z=t= 3 (see text). Parameters: UF=t= 2, US=t=\u00002,\nn= 0:625.\ncharge density over similar depth in the F, resulting from\nthe coupling between charge and spin. The third regime\nis the phase separation regime found at high values of\nJ=t&0:55 (blue triangles and gold stars). Phase sepa-\nration was found in the homogeneous system of Sec. III\n[see Fig. 2(d)]. It is characterized by full polarization\ndeep in the F ( i.e.ni=mi) and a concomitant deple-\ntion of the charge density and the magnetization in the\nF over moderate distance away from the interface.\nWe note that for intermediate values, J=t\u00190:5 (green\ndiamonds), the charge and magnetic pro\fles adjust to\nreach equal value already within the charge/spin corre-\nlation length\u00181=kF(\u00192\u00003 lattice constants) from\nthe SF interface. This steep rise is close to the result of\nthe conventional EFM (dashed green line in Figs. 3,4),\nwhere the transition is driven by the abrupt onset of the\nmagnetization inside the ferromagnetic layer.\nFigure 4(c) reveals the behavior of the magnetization\nin the S. The magnetization decays exponentially with a\ncorrelation length (naturally) independent of J=t. The\noverall magnitude is determined by the value of miat\n8085 90 95 100 00,20,40,60,8charge ni75\n80 85 90 95 100 00,20,40,60,8magnetization m\ni60\n70 80 site i\nx1e-061e-050,00010,0010,010,1magnetization |mi|75\n80 85 90 95 100 site i\nx1e-101e-091e-081e-071e-061e-050,00010,0010,01singlet OP |f\n0(i)|J/t=0.35J/t=0.4\nJ/t=0.5\nJ/t=0.6\nJ/t=0.7\na)b) c)\nd)\nFIG. 4. Close look near the interface of the charge den-\nsity (a), magnetization (b,c) and singlet pair correlations (d)\nshown in Fig. 3 for various values of the exchange coupling\nJ=tin the F. For all cases the local potential Vlocis ad-\njusted as described in Fig. 3. Shown for comparison as a\nthin dashed green line is also the pair correlation in the EFM\nforhi;z=t= 3. The S correlations inside the F in panel d have\nbeen \ftted with Eq. (11) (light red and black lines; see text\nfor \ftting parameters). Parameters: UF=t= 2, US=t=\u00002,\nn= 0:625.\nthe interface, which is largest for intermediate values of\nJ=t, wheremiis not suppressed by S correlations and\nphase separation is not relevant (diamond green lines in\nFig. 4).\nFigs. 3c and 4d display the decay of the superconduct-\ning order parameter f0(i) in the F. Panel 3(c) shows the\noverall behavior of the singlet order parameter for the\nsame range of J=tvalues while the decay inside the F\nis detailed in the inset, and in panel (d) of Fig. 4 on a\nlogarithmic scale. The latter \fgure shows the stark con-\ntrast between the inverse proximity ( J=t.0:55) and the\nphase separated ( J=t> 0:55) regimes. Both cases can be\nmodeled by the following expression\nf0\u0018e(\u0000x=\u0018N)\nxcos\u0012x\n\u0018F\u0013\n; (11)\nwhere\u0018N=vF=2\u0019Tand\u0018F=vF=2\u0019m(x) are the para-\nmagnetic and ferromagnetic coherence lengths. A close\nlook at the curves in Fig. 4 shows that the magnetization\ntakes non-zero values from the interface on; for exam-\nple, forJ=t= 0:35 the value of m(x) is small but \fnite\nalready for 80 < x\u001491. The above expression for f0,\nEq. (11), can be used to obtain an excellent piecewise \ft\nof the numerical data; one divides the space into x x 0withx0\u001882 (x0\u001890) forJ=t= 0:35\n(J=t= 0:4). ForJ=t= 0:35 for example, the curve in\nregions 80 < x\u0014x0can be \ftted with \u0018N> LFand\n\u0018F= 30, whereas for x > x 0we have\u0018N= 33 but\n\u0018F= 0:6. Forx < x 0the pair correlation undergoes a\nsmooth exponential decay that is expected of a paramag-6\nnet (the magnitude of m(x) is very small in this region),\nwhereas farther away the behavior is characteristic of a\nhomogeneous ferromagnet. These results are consistent\nwith previous \fndings for singlet pair correlations in the\nEFM, in a paramagnet (Ref. 27) and a homogeneous fer-\nromagnet (see for example Ref. 4). Refs. 28 and 29 pro-\nvided a complementary analysis of the decay of f0within\nthe EFM. Two almost identical length scales were intro-\nduced that inversely scale with the polarization of the F\nand are\u0018(kF;\"\u0000kF;#)\u00001. The oscillatory behavior is\ndue to the interference of up- and down excitations in\nthe pair amplitude.\nIn the opposite case of large exchange coupling J=t\n(phase separation regime) the pair correlations \frst fol-\nlow the small- J=tbehavior up to some distance x0away\nfrom the interface, followed by a much stronger decay\ndeeper inside the F ( x > x 0). The length x0is deter-\nmined by the point where the magnetization reaches full\npolarization and therefore increases with J=tdue to the\nincreasing low density domain inside the F.\n8085 90 95 100 00,20,40,60,8charge ni75\n80 85 90 95 100 00,20,40,60,8magnetization m\ni60\n70 80 site i\nx1e-061e-050,00010,0010,010,1magnetization |mi|75\n80 85 90 95 100 site i\nx1e-101e-091e-081e-071e-061e-050,00010,0010,01singlet OP |f\n0(i)|J/t=0.2J/t=0.5\nJ/t=0.7\na)b) c)\nd)\nFIG. 5. Charge density (a), magnetization (b,c) and sin-\nglet pair correlations (d) close to the interface as in Figs. 3,4\nbut for a larger value of the Coulomb potential in the F,\nUF=t= 5. The phase separation regime is substantially ex-\ntended when compared to Figs. 3, 4. Both pair correlations\nin the F and the charge and magnetic con\fgurations in S\nare rapidly suppressed away from the interface. Parameters:\nUF=t= 5,US=t=\u00002,n= 0:625.\nThe results of Figs. 3, 4 were obtained for UF=t= 2. A\nlarger local correlation stabilizes the F and increases the\nrange of values of J=tfor which the F is fully polarized\n(see Fig. 5). As a result, even for values of J=t\u00190:2\nclose to the onset of ferromagnetism, pair correlations\nare not able to signi\fcantly suppress magnetism; the in-\nverse proximity e\u000bect is almost absent. Nevertheless, this\nreduction of the intermediate regime does not imply a\ncorresponding extension of the phase separation regime\nto lower values of J=t. AtJ=t\u00180:5 one still observes a\nbehavior similar to the EFM. The phase separation insta-\nbility for large J=t&0:5 persists and the pair amplitudedecay is again shifted away from the interface (panel (d)\nof Fig. 5). The distance from the interface over which\nthe magnetization is suppressed is about the same than\nobserved in Fig. 4. As expected, the behavior of the\nmagnetization in the S is una\u000bected by the change in\nUF=tin the F.\n2. Spectral properties and pair correlations\nThe proximity e\u000bect is also re\rected in spectral prop-\nerties such as the local density of states (LDOS),\n\u001aloc(xi;!) =1\nNX\np;ky;\u001b[jui;\u001b(p;ky)j\u000e(!\u0000\"p(ky))\n\u0000jvi;\u001b(p;ky)j\u000e(!+\"p(ky))];\nwhich within the BdG formalism and the EFM has\nbeen analyzed in Refs. 28 and 29.\nFig. 6a shows the LDOS deep in the S and the F. No-\nticeable are the standard BCS coherence peaks at the\ngap edges in the S (near != 0, black curve at x= 60);\na small numerical 'pair-breaking parameter' \u000f= 0:02t\nhas been introduced for numerical reasons which is re-\nsponsible for the small \fnite LDOS inside the gap. The\noverall structure of the LDOS is otherwise characteristic\nof a two-dimensional square lattice with its logarithmic\nvan-Hove singularity at the band center.\nDeep inside the ferromagnet [red curve at x= 110 in\nFig. 6(a)] the van Hove singularity is split due to the\nformation of subbands (peaks near !=t=\u00063). Note\nthat the apparent \"noise\" in the data is not due to the\nlack of precision of the calculation, but are oscillations\noriginating from the discreteness of the lattice.\nIt is instructive to investigate the dynamical singlet\npair correlations (Gor'kov function)\nf0(ix;t) =1\n2[hcix;\"(t)cix;#(0)i\u0000hcix;#(t)cix;\"(0)i] (12)\ninside the ferromagnet for di\u000berent exchange parameters\nJ=t. The imaginary part of the Fourier transform reads\nF(ix;!) = ImZ1\n\u00001dtei!tf0(ix;t) (13)\n=\u0019X\np;ky\u0002\nui;\"(p;ky)v\u0003\ni;#(p;ky) +ui;#(p;ky)v\u0003\ni;\"(p;ky)\u0003\n\u0002h\rp;ky\ry\np;kyi\u000e(!+\"p(ky))\n\u0000\u0019X\np;ky\u0002\nui;\"(p;ky)v\u0003\ni;#(p;ky) +ui;#(p;ky)v\u0003\ni;\"(p;ky)\u0003\n\u0002h\ry\np;ky\rp;kyi\u000e(!\u0000\"p(ky)):\nFigure 7 shows the position and frequency dependence\nof the singlet correlations, Eq. (13), as an intensity plot\nwhich visualizes the decay of the pair correlations inside\nthe ferromagnet. Note that the Gor'kov functions are\nasymmetric in !,f0(\u0000!) =\u0000f0(!), and in a clean S7\n-4-20246800.20.40.6LDOS [1/t]x=60x=110\n0\n-0.2 -0.4 -0.6 -0.8 -10.150.1\n0.05\n0\n-0.05\n-0.1\n0\n-0.2 -0.4 -0.6 -0.8 -10.20.1\n0\nF(x,ω) [1/t]J/t=0.35J/t=0.5\nJ/t=0.7\n0\n-0.2 -0.4 -0.6 -0.8 -10.030\n-0.03\n0\n-0.2 -0.4 -0.6 -0.8 -1ω \n[t]0.050\nF(x,ω) [1/t]J/t=0.35J/t=0.5\nJ/t=0.7\n0\n-0.2 -0.4 -0.6 -0.8 -1ω\n [t]0.020.01\n0\n-0.01\n-0.02\nb) x=80c) x=81\nd) x=85e) x=90\nf) x=100\na)\nFIG. 6. Local density of states (a) inside the S (black;\nx= 60) and the F (red; x= 110) for J=t= 0:5. Panels (b-f):\nimaginary part of the singlet Gor'kov pair correlation F(ix; !)\ndependence on frequency at speci\fc points in the heterostruc-\nture and for various exchange couplings as indicated in the\npanels. The correlations at x= 80 are in the S while those\nforx\u001581 are in the F. Note that the singlet order parameter\nforJ=t= 0:5 and x= 81 has opposite sign as compared to the\nother couplings (cf. panel (c) of Fig. 3). For better compari-\nson we have therefore multiplied the J=t= 0:5 in panel (c) by\n\u00001, symbolized by the dashed line. Parameters: UF=t= 2,\nUS=t=\u00002,n= 0:625.\nsystem show a\u00061=p\n!\u0000\u0001 singularity at the gap edges\n[\u0001(x) =f0(ix;t= 0) being the superconducting gap].\nOnly the! <0 part ofF(ix;!) is shown in Figs. 6 and\n7.\nA main di\u000berence between the inverse proximity\nregime and the phase separated regime is immediately\napparent at small !=twhen comparing Figs. 7(a) and\n7(b): pair correlations extend deep into the F for small\nJ=t[Fig. 7(a)]. The region in the F where these pair\ncorrelations are present also shrinks with increasing !=t.\nIn the region where the magnetization is suppressed, ei-\nther because of the inverse proximity e\u000bect (small J=t)\nor phase separation (large J=t) the low energy F(ix;!)\ncontinuously extends from the S region into the F. This\nis clearly visible in Fig. 7(a) at low !=twhere the inten-\nsity plot shows pronounced pair correlations (indicated\nby solid red color) close to the interface. By contrast,\nthe Gor'kov function starts oscillating within a partially\npolarized region of the ferromagnet as indicated by the\nalternating red-blue pattern of Fig. 7(a). Similarly, os-\ncillations of F(ix;!) are seen at \fxed position ixas a\nfunction of !=tand we now analyze this !-dependence\nofF(ix;!) in more detail at various distances from the\ninterface in the F, shown in panels (b-f) of Fig. 6.\nThe singlet pair correlations F(ix;!) shown in panel\n(b) of Fig. 6 are calculated at the interface, on the S side,\nx= 80. The peak position, indicating the size of the S\ngap is largest (and sharpest) for J=t= 0:7 in agreement\nwith the larger singlet order parameter (cf. panel (c) of\nFig. 3).\nFIG. 7. Position and frequency dependent singlet correla-\ntions, Eq. (13), for (a) J=t= 0:35 and (b) J=t= 0:7. Pair\ncorrelations oscillate both in space and frequency and extend\ndeeper into the F for small J=t. Parameters US=t=\u00002,\nUF=t= 2n= 0:625.\nOn the other side of the interface, in the F ( x= 81,\npanel (c) of Fig. 6) one still observes a sizable gap for all\ncouplings, which however, is now largest for J=t= 0:35\nwhere the magnetization is suppressed close to the in-\nterface. Further away from the interface, at x= 85,\npanel (d) shows that for J=t= 0:5 the magnetic sys-\ntem is already completely polarized, and the pair cor-\nrelations are suppressed on the scale of the plot. In-\nterestingly, at this same location, the phase separated\nsolution (J=t= 0:7) and the inverse proximity solution\n(J=t= 0:35) still reveal a low energy peak and thus the\noccurrence of a proximity induced gap. Moreover, at\nlarger energies ( !=t\u0018\u00000:3) a second broad peak ap-\npears with opposite sign. For J=t= 0:35 this peak turns\nout to be related to the onset of frequency oscillation in\nF(ix;!) observed further away from the interface, as seen\nin panels (e) and (f). In the di\u000busive limit and within\nthe EFM4,5it was shown that the occurrence of such\noscillations in frequency is a direct consequence of the\nexchange \feld h. Similarly, in the BdG approach they\narise from the superposition of the di\u000berent excitations\nin the spin-up and down bands and thus disappear when\nthe system is completely polarized. This explains why in\npanel (e) and (f), at x= 90;100, only pair correlations\nin the inverse proximity regime ( J=t= 0:35) are \fnite.\nNote that the frequency integration of F(ix;!) yields the\nlocal (equal time) Gor'kov function \u0001( x) =f0(ix;t= 0)\natxwhich vanishes in the F region and thus requires a\ncancellation of the \fnite contributions to F(ix;!).\nFinally, we note in panels (d-f) that there are smaller\noscillations, both in magnitude and frequency, super-\nposed to the large oscillation of F(ix;!) just mentioned.\nThese are the same small oscillations found in all curves\nof panels (b,c) and are related to the discrete spatial lat-\ntice.\nWe used everywhere minstead ofSzso i changed it in\nthe following paragraph and removed the sentence that\nrefers to the z-axis choice. The results of this section IV\nwere obtained for singlet pair correlations generated in8\nthe S and leaking into the F. A similar behavior is also\nfound for the m= 0 triplet correlations brought about\nin the F region. In the SF hybrid structure with periodic\nboundary conditions studied here, there are no m=\u00061\ntriplet components since the magnetic inhomogeneities\nare found in the magnitude of the magnetization while its\norientation is \fxed. This contrasts the magnetic and su-\nperconducting inhomogeneities discussed here from those\nof previous work.7,8,20{22\nV. CONCLUSION\nWe have analyzed proximity e\u000bects in a FS het-\nerostructure in which the F is described within an ex-\ntended Hubbard-type model where the ferromagnetic ex-\nchange arises from intersite contributions of the Coulomb\ninteraction. Such a description allows the self-consistent\ntreatment of both superconducting and magnetic order\nparameters which gives rise to features not present in ap-\nproaches where the exchange \feld is \fxed inside the F. In\nparticular, we have found that for small exchange inter-\nactions and onsite correlations the magnetization close\nto the interface may be suppressed by the S correlations\nwhich signi\fcantly alters the decay of the pair correla-\ntions inside the F. Similar for large exchange couplings\nthe system shows an instability towards phase separa-\ntion which is also realized close to the interface with a\nconcomitant suppression of the magnetization. As a con-\nsequence the S correlations extend far inside the phase\nseparated region and only get suppressed when the mag-netization recovers at some distance from the interface.\nThe correlations leak much deeper than for the e\u000bec-\ntive \feld model; this may signi\fcantly a\u000bect the Joseph-\nson current and is being investigated. Such an instabil-\nity towards phase separation is also inherent in double-\nexchange models for ferromagnetism which are usually\nconsidered to be appropriate for transition metals48.\nTherefore, we expect that these aspects of our results\nare also valid in such systems and are planned in future\ninvestigations. In the present paper we have restricted\nourselves to collinear magnetic structures. However, mi-\ncroscopic magnetic models usually show also more com-\nplex magnetic structures in some part of the phase di-\nagram, as for example the antiferromagnetic spirals in\npanel (b) of Fig. 2 for J=t= 0:1. A heterostructure where\nthe ferromagnet displays a spiral rotation would also in-\nducem= 1 triplet components inside the F. As a result,\nwe expect that the pair correlations inside the F have a\npronounced in\ruence on the periodicity of such magnetic\nstructures. Work in this direction is in progress.\nVI. ACKNOWLEDGEMENTS\nCM and GS thank the DAAD for \fnancial support.\nAB gratefully acknowledges funding provided by the Na-\ntional Science Foundation (DMR-1309341).\nAppendix A\nThe matrices de\fned in Eq. (10) are given by\nTi;i+x(ky) =0\nBB@\u0000t+ 2Jx\ni(kx\ni;\"+kx\ni;#)\u00030 0 2 Jx\nipx(i)\n0\u0000t+ 2Jx\ni(kx\ni;\"+kx\ni;#)\u00032Jx\nipx(i+x) 0\n0 2 Jx\nipx(i)t\u00002Jx\ni(kx\ni;\"+kx\ni;#) 0\n2Jx\nipx(i+x) 0 0 t\u00002Jx\ni(kx\ni;\"+kx\ni;#)1\nCCA(A1)\nand\nVi(ky) =0\nBB@\u00002tcos(ky) +v\";\"(i)v\";#(i) 0 Uif0(i) + 2Jy\nipy(i)\nv#;\"(i)\u00002tcos(ky) +v#;#(i)Uif0(i) + 2Jy\nipy(i) 0\n0 Ui\u0001\u0003\ni+ 2Jy\ni[py(i)]\u00032tcos(ky)\u0000v\";\"(i)v\u0003\n\";#(i)\nUi\u0001\u0003\ni+ 2Jy\ni[py(i)]\u00030 v\u0003\n#;\"(i) 2tcos(ky)\u0000v#;#(i)1\nCCA(A2)\nwith the following abbreviations\nv\u001b;\u001b(i) =Ui\n2(ni\u0000\u001bmi) +Vloc\ni\u0000\u0016+hi;z\u001b\n\u0000Jx\ni\u0000x(ni\u0000x+\u001bmi\u0000x)\u0000Jx\ni(ni+x+\u001bmi+x)\n\u00002Jy\ni(ni+\u001bmi) + 4Jy\niRe[(ky\ni;\"+ky\ni;#)eiky]\nv\";#(i) =\u0000UihS\u0000\nii\u00002Jx\ni\u0000xhS\u0000\ni\u0000xi\u00002Jx\nihS\u0000\ni+xi\n\u00004Jy\nihS\u0000\nii\nni=X\n\u001bhni;\u001bi\nmi=X\n\u001b\u001bhni;\u001bi(\u001b=\u00061)\nhS\u0000\nii=hcy\ni;#ci;\"i=1\nNyX\nkyhcy\ni;#(ky)ci;\"(ky)ikx\ni;\u001b=hcy\ni;\u001bci+x;\u001bi=1\nNyX\nkyhcy\ni;\u001b(ky)ci+x;\u001b(ky)i\nky\ni;\u001b=hcy\ni;\u001bci+y;\u001bi=1\nNyX\nkye\u0000ikyhcy\ni;\u001b(ky)ci;\u001b(ky)i\npx(i) =hci;#ci+x;\"i=1\nNyX\nkyhci;#(ky)ci+x;\"(ky)i\npy(i) =hci;#ci+y;\"ie\u0000iky+hci+y;#ci;\"ieiky:9\nAppendix B\nIn Sec. III we discuss the magnetic state of the F alone,\nin the correlated single band model, Eq. (9). Spiral mag-\nnetic solutions with the Ansatz hS\u0006\nii=S0exp(\u0006iqRi)\nare obtained by factorizing Eqs. (3, 4) with respect to\nthe operators\nS+(\u0000)\nq =X\nkcy\nk+q\"(#)ck#(\"):Since the charge density for these solutions is constant\nHartree terms are neglected as they only shift the en-\nergy by a constant value. For a given momentum qthe\nresulting energy is given by\nE(q) =NX\nk\u001c\u0010\ncy\nk+q\"cy\nk#\u0011\nH\u0012ck+q\"\nck#\u0013\u001d\n+ 4JNS2\n0(cos(qx) +cos(qy))\u00002JN(v2\nx+v2\ny) +UNS2\n0; (B1)\nwhere\nH=\u0012\n\"k+q+2J\nt[vxcos(kx+qx) +vycos(ky+qy)]\u00004JS0\nt[cos(qx) + cos(qy)]\u0000US0\nt\n\u00004JS0\nt[cos(qx) + cos(qy)]\u0000US0\nt\"k+2J\nt[vxcos(kx) +vycos(ky)]\u0013\n: (B2)\nThe quantities\nvx=y=1\nNX\nk\u001bcos(kx=y)hnk\u001bi:renormalize the kinetic energy via the magnetic interac-\ntion and have to be determined self-consistently.\n\u0003andreas.bill@csulb.edu\nygoetz.seibold@tu-cottbus.de\n1P. Fulde and R. A. Ferrell, Phys. Rev. 135, A550 (1964);\nA. I. Larkin and Y. N. Ovchinnikov, Sov. Phys. JETP 20,\n762 (1965).\n2V. L. Berezinskii, JETP Lett. 20, 287 (1975).\n3M. Eschrig, Physics Today 64, 43 (2011); F. Giazotto and\nF. Taddei, Phys. Rev. B 77, 132501 (2008); M. Eschrig,\nRep. Prog. Phys. 78, 104501 (2015).\n4A.I. Buzdin, L.N. Bulaevskii, S.V. Panyukov, JEPT Lett.\n35, 147 (1982); A.I. Buzdin, Rev. Mod. Phys. 77, 935\n(2005).\n5F. S. Bergeret, A. F. Volkov, and K. B. Efetov, Phys. Rev.\nLett. 86, 4096 (2001); F. S. Bergeret, K. B. Efetov, and A.\nI. Larkin, Phys. Rev. B 62, 11872 (2000); F. S. Bergeret,\nA. F. Volkov, and K. B. Efetov, Rev. Mod. Phys. 77, 1321\n(2005).\n6A. Kadigrobov, R. I. Shekhter, and M. Jonson, Europhys.\nLett. 54, 394 (2001).\n7T. E. Baker, A. Richie-Halford, O. E. Icreverzi, and A.\nBill, Europhys. Lett. 107, 17001 (2014).\n8T. E. Baker, A. Richie-Halford, and A. Bill, Phys. Rev. B\n94, 104518 (2016).\n9R. S. Keizer, S. T. B. Goennenwein, T. M. Klapwijk, G.\nMiao, G. Xiao, and A. Gupta, Nature 439, 825 (2006).\n10I. Sosnin, H. Cho, V. T. Petrashov, and A. F. Volkov, Phys.\nRev. Lett. 96, 157002 (2006);\n11J. W. A. Robinson, J. D. S. Witt, and M. G. Blamire,\nScience 329, 59 (2010); G.B. Hal\u0013 asz, J.W.A. Robinson,\nJ.F. Annett, and M.G. Blamire, Phys. Rev. B 79, 224505\n(2009); G.B. Hal\u0013 asz, M.G. Blamire, and J.W.A. Robin-son,Phys. Rev. B 84, 024517 (2011).\n12M. S. Anwar, F. Czeschka, M. Hesselberth, M. Porcu, and\nJ. Aarts, Phys. Rev. B 82, 100501 (2010).\n13T. S. Khaire, M. A. Khasawneh, W. P. Pratt, Jr., and N.\nO. Birge, Phys. Rev. Lett. 104, 137002 (2010); C. Klose, T.\nS. Khaire, Y. Wang, W. P. Pratt Jr., and N. O. Birge, B. J.\nMcMorran, T. P. Ginley, J. A. Borchers, B. J. Kirby, B. B.\nMaranville, and J. Unguris, Phys. Rev. Lett. 108, 127002\n(2012); M. A. Khasawneh, T. S. Khaire, C. Klose, W. P.\nPratt, Jr., and N. O. Birge, Supercond. Sci. Technol. 24,\n024005 (2011); E. C. Gingrich, P. Quarterman, Y. Wang,\nR. Loloee, W.P. Pratt Jr, and N.O. Birge, Phys. Rev. B\n86, 224506 (2012).\n14J. Zhu, I. N. Krivorotov, K. Halterman, and O. T. Valls,\nPhys. Rev. Lett. 105, 207002 (2010).\n15P. V. Leksin, N. N. Garifyanov, I. A. Garifullin, Y. V.\nFominov, J. Schumann, Y. Krupskaya, V. Kataev, O. G.\nSchmidt, and B. B uchner, Phys. Rev. Lett. 109, 057005\n(2012).\n16L. Wen, L.-J. Jin, Y. Chen, G.-Q. Zha, and S.-P. Zhou,\nEurophys. Lett. 105, 27007 (2014).\n17D. Sprungmann, K. Westerholt, H. Zabel, M. Weides, and\nH. Kohlstedt, Phys. Rev. B 82, 060505 (R) (2010).\n18F. S. Bergeret, A. F. Volkov, and K. B. Efetov, Phys. Rev.\nB64, 134506 (2001);\n19W. Belzig, F.K. Wilhelm, C. Bruder, and G. Sch on, Su-\nperlatt. and Microstr. 25, 1251 (1999).\n20T. E. Baker, A. Richie-Halford, and A. Bill, New J. Phys.\n16, 093048 (2014)\n21J. de Rojas, T.E. Baker, A. Richie-Halford, and A. Bill J.\nof Supercond. Nov. Magn. 25, 2177 (2012).10\n22T. E. Baker and A. Bill, arXiv:1802.08196.\n23M. Leadbeater, C. J. Lambert, K. E. Nagaev, R. Raimondi,\nand A. F. Volkov, Phys. Rev. B 59, 12264 (1999).\n24J.-X. Zhu and C. S. Ting, Phys. Rev. 61, 1456 (1999).\n25M. Eschrig, T. L ofwander, T. Champel, J. C. Cuevas, J.\nKopu, and Gerd Sch on, J. Low Temp. Phys. 147, 457\n(2007).\n26D. Fritsch and J. F. Annett, New J. Phys. 16, 055005\n(2014), ibid. J. Phys.: Cond. Mat. 26, 274212 (2014).\n27P. G. de Gennes, Superconductivity of Metals and Alloys ,\nAddison-Wesley (1989).\n28K. Halterman and O. T. Valls, Phys. Rev. B 65, 014509\n(2001).\n29K. Halterman and O. T. Valls, Phys. Rev. B 66, 224516\n(2002).\n30K. Halterman and O. T. Valls, Phys. Rev. B 70, 104516\n(2004).\n31K. Halterman and O. T. Valls, Phys. Rev. B 72,\n060514(R).\n32K. Halterman, P. H. Barsic, and O. T. Valls, Phys. Rev.\nLett. 99, 127002 (2007).\n33K. Halterman, O. T. Valls, and P. H. Barsic, Phys. Rev. B\n77, 174511 (2008).\n34Chien-Te Wu, Oriol T. Valls, and Klaus Halterman, Phys.\nRev. B 86, 014523 (2012).35F. S. Bergeret and I. V. Tokatly Phys. Rev. Lett. 110,\n117003 (2013).\n36A. I. Buzdin and L. N. Bulaevskii, Sov. Phys. JETP 67,\n576 (1988).\n37P. W. Anderson and H. Suhl, Phys. Rev. 116, 898 (1959).\n38G. Annunziata, M. Cuoco, C. Noce, A. Romano, and P.\nGentile, Phys. Rev. B 80, 012503 (2009); G. Annunziata,\nM. Cuoco, P. Gentile, A. Romano, and C. Noce, Phys.\nRev. B 83, 094507 (2011).\n39M. Cuoco, P. Gentile, C. Noce, A. Romano, Z.-J. Ying,\nand H.-Q. Zhou, J. Phys.: Condens. Matter 21, 254203\n(2009).\n40J. E. Hirsch, Phys. Rev. B 40, 2354 (1989).\n41J. C. Amadon and J. E. Hirsch, Phys. Rev. B 54, 6364\n(1996).\n42J. E. Hirsch, Phys. Rev. B 59, 6256 (1999).\n43M. Kollar and D. Vollhardt, Phys. Rev. B 63, 045107\n(2001).\n44A. Schiller, Phys. Rev. B 60, 15660 (1999).\n45J. E. Hirsch, Phys. Rev. B 44, 675 (1991).\n46A. M. Oles and W. Grzelka, Phys. Rev. B 46, 5412 (1992).\n47Y. Nagaoka, Phys. Rev. B 147, 392 (1966); M. Kollar, R.\nStrack and D. Vollhardt, Phys. Rev. B 53, 9225 (1996).\n48E. Dagotto, Nanoscale Phase Separation and Colossal\nMagnetoresistance , Springer (2003)." }, { "title": "1010.1075v1.Ferromagnetic_Resonance_in_Spinor_Dipolar_Bose__Einstein_Condensates.pdf", "content": "arXiv:1010.1075v1 [cond-mat.quant-gas] 6 Oct 2010Ferromagnetic Resonance in Spinor Dipolar Bose–Einstein C ondensates\nMasashi Yasunaga and Makoto Tsubota∗\nDepartment of Physics, Osaka City University, Sumiyoshi-k u, Osaka 558-8585, Japan\n(Dated: April 27, 2022)\nWe used the Gross–Pitaevskii equations to investigate ferr omagnetic resonance in spin-1 Bose–\nEinstein condensates with a magnetic dipole-dipole intera ction. By introducing the dipole interac-\ntion, we obtained equations similar tothe Kittel equations used torepresent ferromagnetic resonance\nin condensed matter physics. These equations indicated tha t the ferromagnetic resonance originated\nfrom dipolar interaction, and that the resonance frequency depended upon the shape of the conden-\nsate. Furthermore, spin currents driven by spin diffusions a re characteristic of this system.\nPACS numbers: 03.75.Mn, 03.75.Nt\nI. INTRODUCTION\nMagnetic resonance (MR) as a physical concept has\nbeen applied in various fields, enabling physical, chem-\nical, and medical experiments to obtain information on\nnuclear spin and electron spin systems. The concept has\nalso provided valuable information to help understand\nthe unknown structures of many condensed matter sys-\ntems [1].\nThe use of MR in the study of ferromagnets, e.g.\nNickel, Cobalt, and Iron, began in the 1940s. Grif-\nfiths observed that the Land´ e’s g-factor of electrons in\nferromagnets was far from the well known value, 2 [2].\nIn order to understand these anomalous results, Kittel\ntheoretically introduced a demagnetizing field into the\nequation representing the motion of the magnetization\nM= (Mx,My,Mz), obtaining an equation valid in an\nexternal magnetic field H0ˆz, withMz0=H0/Nzand de-\nmagnetizing fields [3], thereby obtaining the Kittel equa-\ntion,\ndM\ndt=γn[M×H]. (1)\nHere,γnis the nuclear gyromagnetic ratio, and H=\n(−NxMx,−NyMy,H0−NzMz) is given by the demagne-\ntizing factors Ni. By linearizing the magnetization M=\nM0+δMfromthestationarymagnetization M0=Mz0ˆz,\nKittel obtained a precession of the magnetization and a\nprecessing frequency, i.e.resonance frequency,\nω2=γ2\nn{H0+(Ny−Nz)Mz0}{H0+(Nx−Nz)Mz0},(2)\nwhich explained the anomalous g-factor. Furthermore,\nhe found that the resonance frequency depends on the\nshape of a ferromagnet because Nidepends on the shape\n[3]. Thus, ferromagnetic resonance (FMR) was estab-\nlished, and the workenabled numerousadditionalstudies\n[4].\nMR also plays an important role in quantum conden-\nsate systems. In superfluid3He, the dynamics of the spin\n∗Electronic address: tsubota@sci.osaka-cu.ac.jpvector and the d-vector are represented by the Leggett\nequation, which couples these vectors through magnetic\ndipole-dipole interactions [5]. The equation also shows\nnot only an MR typical of condensed matter, but also a\nnew MR that cannot be described using the equations of\nmotion for general paramagnets and ferromagnets. This\nMR was used to find AandBphases [6]. Parallel ring-\ning, which is an oscillation of longitudinal magnetization,\nwas also observed [7].\nSince the discovery of atomic Bose–Einstein conden-\nsates(BECs)[8,9], BECshavebeenstudiedinopticsand\natomic and condensed matter physics. We have intro-\nducedMRintoBECstorealizemagneticresonanceimag-\ning, a popular method of nondestructive testing. Spinor\nBECs are expected to be suitable for MR, since they\nhave not only internal degrees of freedom but also mag-\nnetic properties. In particular, we are interested in mag-\nnetic dipole-dipole interactions (MDDI) in spinor BECs,\nwhich have been actively studied. The interaction be-\ntween spins has a characteristic symmetry of rotation\nand spin, which is expected to result in a new quan-\ntum phase [10–12] and Einstein–de Haas effects [13]. Ex-\nperimentally, Griesmaier et al.realized spinor dipolar\ncondensates using52Cr atoms, which have a larger mag-\nnetic moment than alkali atoms [14]. The shape of the\ncondensates clearly represented the anisotropy of the in-\nteraction [15, 16]. Thus, MDDI has opened new areas of\nspinor condensate research.\nAs an introduction to MR in BECs, we numerically\nstudied spin echo in dipolar BECs with spin-1 [19]. The\nspin echo is a typical phenomenon of MR, discovered by\nHahn [17] and developed by Carr and Purcell [18]. Previ-\nously, we calculated the transition from Rabi oscillations\nto internal Josephson oscillations in spinor condensates\n[20]. In this paper, we consider MDDI in spin-1 BECs,\nexamining FMR by analyzing the Gross–Pitaevskii (GP)\nequations.\nIn section II, we derive Kittel-like equations from the\nGP equations, and analyze them. In section III, using\na single-mode approximation, we derive Kittel equations\nfrom the Kittel-like equations. The MDDI of the Kittel\nequationsisconsideredasthe originofthe demagnetizing\nfield, which is phenomenologically introduced in Eq. (1).\nIn section IV, we numerically solve the GP equations,2\nobtaining resonance frequencies that depend upon the\nshape of the condensates, and spin currents driven by\nspin diffusion which is given by the MDDI. Finally, Sec.\nV is devoted to our conclusions.\nII. FORMULATION\nIn this section, we derive the equations of motion for\nspinsfromthespin-1GPequationswithanexternalmag-\nnetic field and an MDDI [19].\ni/planckover2pi1∂ψα\n∂t=/parenleftbigg\n−/planckover2pi12\n2M∇2+V−µ+c0n/parenrightbigg\nψα\n−gµBHiFi\nαβψβ+c2FiFi\nαβψβ\n+cdd/integraldisplay\ndr′δij−3eiej\n|r−r′|3Fi(r′)Fj\nαβψβ.(3)\nHere,Vis the trapping potential, µis the chemical\npotential, and the total density n=/summationtext\niniis given\nbyni=|ψi|2. The external magnetic field is H=\n(Hx,Hy,Hz), and the components Fi\nαβof the spin ma-\ntricesˆFiare for spin-1. The interaction parameters are\nc0= (g0+2g2)/3 andc2= (g2−g0)/3 forgi= 4π/planckover2pi12ai/M\nrepresented by s-wave scattering lengths ai. The dipolar\ncoefficient is cdd=µ0g2\neµ2\nB/4π, and the unit vector is\ne= (ex,ey,ez) = (x−x′,y−y′,z−z′)/|r−r′|.\nUnder the homogeneous magnetic field H=Hˆz, the\nequations can be rewritten as,\ni/planckover2pi1∂ψ1\n∂t=/parenleftbigg\n−/planckover2pi12∇2\n2M+V−µ+c0n/parenrightbigg\nψ1−gµBHψ1\n+c2{(n1+n0−n−1)ψ1+ψ∗\n−1ψ2\n0}+D1,\n(4a)\ni/planckover2pi1∂ψ0\n∂t=/parenleftbigg\n−/planckover2pi12∇2\n2M+V−µ+c0n/parenrightbigg\nψ0\n+c2{(n1+n−1)ψ0+2ψ∗\n0ψ1ψ−1}+D0,\n(4b)\ni/planckover2pi1∂ψ−1\n∂t=/parenleftbigg\n−/planckover2pi12∇2\n2M+V−µ+c0n/parenrightbigg\nψ−1+gµBHψ−1\n+c2{(n−1+n0−n1)ψ−1+ψ∗\n1ψ2\n0}+D−1.\n(4c)\nThese dipolar terms are represented as,\nD1=/parenleftbiggψ0√\n2d−+ψ1dz/parenrightbigg\n,\nD0=/parenleftbiggψ1√\n2d++ψ−1√\n2d−/parenrightbigg\n,\nD−1=/parenleftbiggψ0√\n2d+−ψ−1dz/parenrightbigg\n,\nwith the integrations d±=dx±idyanddzgiven by,\ndi=cdd/integraldisplay\ndr′Fi(r′)\n|r−r′|3{1−3ei/summationdisplay\njej}.(5)The spin density vectors Fiare defined as,\nFx=Ψ†ˆFxΨ\n=/planckover2pi1√\n2{ψ∗\n0(ψ1+ψ−1)+ψ0(ψ∗\n1+ψ∗\n−1)},(6a)\nFy=Ψ†ˆFyΨ\n=i/planckover2pi1√\n2{ψ∗\n0(ψ1−ψ−1)−ψ0(ψ∗\n1−ψ∗\n−1)},(6b)\nFz=Ψ†ˆFzΨ=/planckover2pi1(|ψ1|2−|ψ−1|2). (6c)\nHere,Ψ= (ψ1,ψ0,ψ−1)Tis the spinor wave function.\nDifferentiating Eq. (6) with respect to time and utiliz-\ning Eq. (4), we can obtain the Kittel-like equation,\n∂F\n∂t=K+γe[F×Heff] (7)\nwith the gyromagnetic ratio γe=gµB//planckover2pi1of an electron.\nThe first term K= (Kx,Ky,Kz) becomes,\nKx=/planckover2pi1\n2Mi1√\n2{(ψ1+ψ−1)∇2ψ∗\n0−ψ∗\n0∇2(ψ1+ψ−1)\n+ψ0∇2(ψ∗\n1+ψ∗\n−1)−(ψ∗\n1+ψ∗\n−1)∇2ψ0},\nKy=/planckover2pi1\n2Mii√\n2{(ψ1−ψ−1)∇2ψ∗\n0−ψ∗\n0∇2(ψ1−ψ−1)\n−ψ0∇2(ψ∗\n1−ψ∗\n−1)+(ψ∗\n1−ψ∗\n−1)∇2ψ0},\nKz=/planckover2pi1\n2Mi(ψ1∇2ψ∗\n1−ψ∗\n1∇2ψ1\n+ψ∗\n−1∇2ψ−1−ψ−1∇2ψ∗\n−1).\nThe effective magnetic fields Heff=H+Hdd=\n(Hx\neff,Hy\neff,Hz\neff)consistoftheexternalmagneticfieldand\nthe dipolar field Hdd, given by,\nHx\neff=−cdd\ngµBdx,\nHy\neff=−cdd\ngµBdy,\nHz\neff=H−cdd\ngµBdz.\nNote that Eq. (7) does not depend on spin exchange\ninteraction, which refers to the second term with c2in\nEq. (3). Generally, the interaction affects a spin through\nthe effective magnetic fields of the other spins. However,\nexchange interaction does not appear in Heff. Therefore,\nthe isotropic exchange interaction does not affect MR in\nthese condensates.\nWe can redefine Eq. (7) as,\n∂Fk\n∂t=/planckover2pi1\n2Mi∇2Fk−∇·jk+γe[F×Heff]k,(8)\nwhere,\njx=/planckover2pi1√\n2Mi(ψ∗\n0∇(ψ1+ψ−1)+(ψ∗\n1+ψ∗\n−1)∇ψ0),\njy=/planckover2pi1√\n2M(ψ∗\n0∇(ψ1−ψ−1)−(ψ∗\n1−ψ∗\n−1)∇ψ0),\njz=/planckover2pi1\nMi(ψ∗\n1∇ψ1−ψ∗\n−1∇ψ−1).3\nThe equation of motion (8) for spins describes the prop-\nerties of spin dynamics in a ferromagnetic fluid. The\nfirst, second, and third terms of Eq. (8) represent spin\ndiffusion, spin current, and spin precession around Heff,\nrespectively.\nComparing Eq. (8) with Eq. (1), we noticed several\ndifferences. First, Eq. (8) was directly derived from\nthe GP equations, whereas Eq. (1) is a phenomenologi-\ncal equation of magnetization. The spin density vectors\nin Eq. (8) are microscopically affected by other spins\nthrough the dipolar fields in the effective magnetic fields.\nOn the other hand, the magnetization in Eq. (1) is af-\nfected by demagnetizing fields originating from macro-\nscopicallypolarizedmagnetizationin the condensed mat-\nter. Namely, Eq. (8) can describe the macroscopic de-\nmagnetizing field resulting from the microscopic dipolar\nfield. This is a very important difference between these\nequations.\nWe initially investigated the physics of the first and\nsecond terms of Eq. (8). To simplify the discussion, we\nconsidered the equation under the condition Heff=0.\nThus, we derived the continuity equations,\n∂Fi\n∂t+∇·Ji= 0, (9)\nwhereJk=jk−/planckover2pi1/(2Mi)∇Fkisaneffectivecurrentterm,\nJx=−i/planckover2pi12\n2√\n2M{ψ∗\n0∇(ψ1+ψ−1)+(ψ∗\n1+ψ∗\n−1)∇ψ0\n−ψ0∇(ψ∗\n1+ψ∗\n−1)−(ψ1+ψ−1)∇ψ∗\n0},(10a)\nJy=/planckover2pi12\n2√\n2M{ψ∗\n0∇(ψ1−ψ−1)−(ψ∗\n1−ψ∗\n−1)∇ψ0\n+ψ0∇(ψ∗\n1−ψ∗\n−1)−(ψ1−ψ−1)∇ψ∗\n0},(10b)\nJz=−i/planckover2pi12\n2M(ψ∗\n1∇ψ1−ψ1∇ψ∗\n1−ψ∗\n−1∇ψ−1+ψ−1∇ψ∗\n−1).\n(10c)\nEquation (9) can also be rewritten as,\nd\ndt/integraldisplay\nVFidV=/integraldisplay\nV∇·JidV=/integraldisplay\nSJi·ndS,\nby using the volume integral and the surface integral,\nwhose unit vector nis vertical to the surface for Stokes’\ntheorem. The equation indicates that the expectation\nvalue of the spin matrix ∝an}b∇acketle{tˆFi∝an}b∇acket∇i}ht=/integraltextdVFiin the volume V\nis conserved for the spin probability flux Jileaving and\nentering the surface.\nUnderHeff∝ne}ationslash= 0, the Kittel-like equation can be re-\nduced to the following equation,\n∂Fi\n∂t+∇·Ji= [F×Heff]i, (11)\nwhere the right side of the equation breaks the conserva-\ntion law of spin density. Therefore, the Kittel-like equa-\ntionshavetwodynamics: spin precessionswith frequency\ngiven by the effective magnetic field and spin currents\nwithout spin conservation. The spin currents of the sys-\ntem will be discussed in Sec. IVBIII. FMR UNDER SINGLE-MODE\nAPPROXIMATION\nIn order to study the basic properties of the second\nterm in Eq. (7), we introduced the single-mode approxi-\nmation,\nψi(r,t) =√\nNξi(t)φ(r)exp/parenleftbigg\n−iµt\n/planckover2pi1/parenrightbigg\n,(12)\nwhereφsatisfies the eigenvalue equation ( −/planckover2pi12∇2/2M+\nV+c0n)φ=µφwith the relation/integraltextdr|φ|2= 1. The\napproximation is effective when the shapes of the con-\ndensates are determined by the spin-independent terms,\nnamely|c0| ≫ |c2|[21]. For87Rb and23Na, the re-\nlation is satisfied. Under this approximation, the first\nterm of Eq. (7) vanishes, and we obtain the Kittel equa-\ntion for the spatially independent spin density vector\nS= (Sx,Sy,Sz),\ndS\ndt=γe[S×HSMA\neff], (13)\nwhere,\nSx=/planckover2pi1√\n2{ξ∗\n0(ξ1+ξ−1)+ξ0(ξ∗\n1+ξ∗\n−1)},\nSy=i/planckover2pi1√\n2{ξ∗\n0(ξ1−ξ−1)−ξ0(ξ∗\n1−ξ∗\n−1)},\nSz=/planckover2pi1(|ξ1|2−|ξ−1|2),\nand the effective magnetic field HSMA\neff =\n(−Nx\nddSx,−Ny\nddSy,H−Nz\nddSz) is given by\nNi\ndd=cdd\ngµBN/integraldisplay /integraldisplay\ndrdr′|φ(r)|2|φ(r′)|2\n|r−r′|3{1−3ei/summationdisplay\njej}.\n(14)\nEquation (13) also indicates that the spin vector Spre-\ncesses around HSMA\neff. The precession frequency reveals\nthe characteristic dynamics. Next, we consider a small\ndeviationδS= (δSx,δSy,δSz) around the stationary so-\nlution,S0=S0ˆzwithS0=H0/Nz\ndd, of Eq. (13), namely\nS=S0+δS. Introducing this representation into Eq.\n(13) and linearizing the equation, we derived the follow-\ning equations,\nd\ndtδSx=γe{H+(Ny\ndd−Nz\ndd)S0}δSy,\nd\ndtδSy=−γe{H+(Nx\ndd−Nz\ndd)S0}δSx,\nd\ndtδSz= 0,\nwhich give the resonance frequency,\nω2=γ2\ne{H+(Nx\ndd−Nz\ndd)S0}{H+(Ny\ndd−Nz\ndd)S0}(15)\nThespinprecesseswiththe resonancefrequency ω, which\ndepends on the dipolar terms Ni\ndd.4\nHere, we consider the single particle density distri-\nbution|φ(r)|2∝e−(x2+y2+λzz2)/a2, whereλzis the as-\npect ratio, and discuss simple situations. For the spher-\nical case of λz= 1, the integration (14) results in\nNx\ndd=Ny\ndd=Nz\ndd, givingω=γeH. The dipolar\nfields are canceled because of the isotropy, so that the\nspin precesses with Larmor frequency. For the circular\nplane (infinite cylinder) case of λz=∞(0), we obtain\nω=γe{H−(Nx\ndd−Nz\ndd)S0}forNx\ndd=Ny\ndd.\nIn this representation, it seems that the microscopic\ndipolar fields, Eq. (14), act as a macroscopic demagne-\ntizingfield tocompareEq. (2) with (15). We believethat\nthe origin of the demagnetizing field is an MDDI. If the\nabove discussion is correct, the dipolar coefficients Ni\ndd\nshoulddepend ontheshapeofthecondensates. However,\nthe single-mode approximation in spinor dipolar BECs\nis not effective in large-aspect-ratio condensates, as dis-\ncussed by Yi and Pu [22]. Therefore, we must consider\nthe spin dynamics beyond the approximation.\nIV. FMR FOR NUMERICAL CALCULATION\nA. Precession dependence on the aspect ratio λ\nIn this section, we discuss FMR by numerically calcu-\nlating the two-dimensional Eq. (3) under the condition\nof87Rb, namely c0≫ −c2>0. We begancalculatingthe\nspin precessions by applying a π/20 pulse to the ground\nstate, whose spins were polarized to the uniform mag-\nnetic field H=Hˆztrapped by V=Mω2\nx(x2+λ2y2)/2\nwithgµBH//planckover2pi1ωx= 20 and an aspect ratio λ=ωy/ωx.\nWe investigated the dynamics of ∝an}b∇acketle{tFx∝an}b∇acket∇i}htforλ= 0.5,1,\nand 1.5 with and without the MDDI. From t= 0 to\nπ/(20γeH), aπ/20 pulse was applied. Then, the spins\nwere tilted by π/20 radians from the zaxis with pre-\ncession. After turning off the pulse, the spins precessed\naround the zaxis, conserving ∝an}b∇acketle{tFz∝an}b∇acket∇i}ht. We define the nota-\ntion∝an}b∇acketle{tFi∝an}b∇acket∇i}htdd\nλ=λaand∝an}b∇acketle{tFi∝an}b∇acket∇i}htλ=λaas indicating the expectation\nvalues ofFiwith and without an MDDI in the trap with\nλ=λa.\nFirst, the typical motions of spins are shown in Fig. 1.\nInvestigatingthe time development of ∝an}b∇acketle{tFi∝an}b∇acket∇i}htdd\nλ=0.5,∝an}b∇acketle{tFi∝an}b∇acket∇i}htdd\nλ=1,\nand∝an}b∇acketle{tFi∝an}b∇acket∇i}htdd\nλ=1.5, we obtained the differences between their\nprecession frequencies, as shown in Fig. 1 (a) and (b).\nThe differences appeared at frequencies below the Lar-\nmor frequency, given by H. For 0 ≤t≥2, no devi-\nation between the precessions was observed, but devia-\ntions clearly appeared as more time elapsed. In order\nto demonstrate that the λdependence was given not by\nHbut byHdd, we show precessions for the same aspect\nratios without the MDDI in Fig. 1 (c) and (d). The\nprecession frequency did not change without the MDDI\nfor different values of λ. Therefore, the dipolar frequency\nωdd=γeHdddepends upon the shape of the condensate.\nNext, we examined the effects of the MDDI on the\nprecessions in Fig. 2. Comparing ∝an}b∇acketle{tFx∝an}b∇acket∇i}htdd\nλwith∝an}b∇acketle{tFx∝an}b∇acket∇i}htλ,!\"#$!\"#%\"\"#%\"#$\n\" % $!\"#$!\"#%\"\"#%\"#$\n&\" &% &$π/20pulse/angbracketleftFx/angbracketrightdd\nλ=0.5/¯h/angbracketleftFx/angbracketrightdd\nλ=1/¯h\n/angbracketleftFx/angbracketrightdd\nλ=1.5/¯h!\"# !$#\nωxt ωxt!\"#$!\"#%\"\"#%\"#$\n\" % $!\"#$!\"#%\"\"#%\"#$\n&\" &% &$!%# !&#\n/angbracketleftFx/angbracketrightλ=1/¯h/angbracketleftFx/angbracketrightλ=1.5/¯h\n/angbracketleftFx/angbracketrightλ=0.5/¯h\nFIG. 1: (Color online) The time development of /angbracketleftFx/angbracketrightdd\nλ, (a)\nand (b), and /angbracketleftFx/angbracketrightλ, (c) and (d). The red solid, blue dashed,\nand green dotted lines show the results of λ= 0.5, 1, and 1 .5\nrespectively. The gray zone represents the duration of a π/20\npulse.\nwe observed that the MDDI caused an effective mag-\nnetic field, because the frequency of the precession with\nthe MDDI deviated from that without the MDDI in\nFig. 2 (a) to (f). Assuming that ∝an}b∇acketle{tFi∝an}b∇acket∇i}htdd\nλ=1− ∝an}b∇acketle{tFi∝an}b∇acket∇i}htλ=1\nis represented approximately to Acosγe(H+Hdd)t−\nAcosγeHtwith an amplitude A, we extracted the dipole\nfrequency from the waveform. Since the waveform be-\ncame−2Asinωddt/2sin(ωL+ωdd/2)t, the beat consisted\nof the large frequency ωL+ωdd/2 and the small fre-\nquencyωdd/2. From Fig. 2 (h), we estimated these\nfrequencies to obtain ωdd/ωL≃6.5,9,and11×10−3for\nλ= 0.5,1,and1.5 respectively.\nFigure 3 shows the λdependence of ωdd/ωL. From the\nresults, however, we cannot safely conclude that the λ\ndependence of the frequencies is given by changing the\nshape of the condensates, since the dipolar frequencies\nmay be given by change of the density with the shape.\nFMR in condensed matters has been discussed in con-\ndensed matter of uniform density, even with changing\nshape. On the other hand, atomic BECs have tunable\ndensity and shape. Therefore, our calculations indicate\ncharacteristic of FMR in atomic cold gases.\nB. Spin current\nWe observed spin currents driven by spin diffusion,\nwhich was caused by a rdependence of the dipolar field.\nFigure 4 shows the projections of Fonto thex-yplane5\n!\"#$!\"#%\"\"#%\"#$\n&\" &% &$!\"#$!\"#%\"\"#%\"#$\n\" % $π/20pulse\nωxt ωxt!\"#\n!\"#$!\"#%\"\"#%\"#$\n&\" &% &$!$#\n!\"#$!\"#%\"\"#%\"#$\n\" % $!%#\n!\"#$!\"#%\"\"#%\"#$\n&\" &% &$!&#\n!\"#$!\"#%\"\"#%\"#$\n\" % $!'# !(#\n!\"#$!\"#%\"\"#%\"#$\n\" &\" %\" '\" $\" (\"!)# !*#\n!\"#$!\"#%\"\"#%\"#$\n\" & % ' $ ( )λ=0.5\nλ=1\nλ=1.5λ=0.5\nλ=1\nλ=1.5\nFIG. 2: (Color online) Comparing the precession with and\nwithout the MDDI. (a) and (b), (c) and (d), and (e) and (f)\nshow the precession for λ= 0.5, 1, and 1 .5, respectively. The\nsolid anddashedlines are /angbracketleftFx/angbracketrightλand/angbracketleftFx/angbracketrightdd\nλ. (g)and(h)repre-\nsent (/angbracketleftFx/angbracketrightdd\nλ=0.5−/angbracketleftFx/angbracketrightλ=0.5)//planckover2pi1(solid), ( /angbracketleftFx/angbracketrightdd\nλ=1−/angbracketleftFx/angbracketrightλ=1)//planckover2pi1\n(dot), and ( /angbracketleftFx/angbracketrightdd\nλ=1.5−/angbracketleftFx/angbracketrightλ=1.5)//planckover2pi1(dashed), respectively.\nforλ= 1.5 andωxt= 12.7. The precession with the\nMDDI lost homogeneity of the spin directions, whereas\nthe precession without the MDDI maintained this ho-\nmogeneity. This is because the precession frequency\nhas anrdependence, specifically, ω(r) =γeHeff(r) =\nγe(H+Hdd(r)).\nThe dipole interaction drives the spin diffusion, which\nis shown in Fig. 5. The figure shows Fx/|Fxy|= cosφas\na function of xaty= 0, where φis the angle between\nthe spin vector and the xaxis. In the dynamics with!\"!#!\"!#$!\"!%!\"!%$!\"!&\n!!\"$##\"$%%\"$&\nλωdd\nωL\nFIG. 3:λdependence of ωdd/ωL.\nxy\n03.637×10−3\n0!\"# !$#1.964×10−3\nFIG. 4: (Color online) Projection of Fonto the x-yplane\nforλ= 1.5 andωxt= 12.7. The figures show the results (a)\nwith MDDI and (b) without that. The vectors are nondimen-\nsionalized.\nthe dipole interaction for λ= 1.5 (a) and 1 (b), the\nspin densities lost their angular coherence, whereas the\ndynamicswithout the dipole interactionsmaintained this\ncoherence ( (c) and (d)).\nThe spin diffusion drives the spin current Jkin Eq.\n(10), which is shown in Fig. 6. In order to explain how\nthe spin current is driven by the spin diffusion, we con-\nsidered the amplitudes of the wave functions ψj=fjeiϕj\nas,\nψ1(r,t) =/radicalbig\nn(r,t)\n2(1+cosθ(r,t))eiϕ1(r,t),\n(16a)\nψ0(r,t) =/radicalbigg\nn(r,t)\n2sinθ(r,t)eiϕ0(r,t),(16b)\nψ−1(r,t) =/radicalbig\nn(r,t)\n2(1−cosθ(r,t))eiϕ−1(r,t),\n(16c)6\n!\"!#$%##$%\"\n!\"#!%# % \"#!\"#!\"!#$%##$%\"\n!\"#!%# % \"#!$#\n!%#\nx xλ=1λ=1.5\nωxt=0.128\n163240\n4848\n!\"!#$%##$%\"\n!\"#!%# % \"#λ=1\n0.128\n1632\n40\n0.12 4048 32 16\n8!&#\nλ=1.5\n!\"!#$%##$%\"\n!\"#!%# % \"#0.12 40832 16 48cosφ cosφ\nFIG. 5: Dynamics of a cross-section of Fx/|Fxy|aty= 0,\nwhere|Fxy|=/radicalbig\nF2x+F2y. From the relation Fx=|Fxy|cosφ,\nthe parameter represents cos φ. The results with the MDDI\n(a) and without it (b) are shown for λ= 1.5, and (c) and (d)\nshow results for λ= 1. The xaxis are nondimensionalized by/radicalbig\n/planckover2pi1/Mωx\nwhere the forms show the ground state of the ferromag-\nnetic state [23]. The amplitude is represented by nand\nthe angleθbetween the spin and the zaxis. We intro-\nduced this representation to demonstrate that the spin\ncurrent is derived from the spin diffusion. Of course, we\nconfirmed the validity of the ferromagnetic representa-\ntion under the pulse and magnetic field by calculating θ\ndirectly. Therefore, it can be utilized for the polarized\nspin state studied in our work. The amplitudes f±1were\nformed to represent Fz=n/planckover2pi1cosθ, andf0was deter-\nmined to satisfy the relation n=/summationtext\nj|ψj|2. For example,\n(n1,n0,n−1) = (n,0,0) led toFz=n/planckover2pi1withθ= 0, and\n(n1,n0,n−1) = (n/4,n/2,n/4) resulted in Fz= 0 with\nθ=π/2. The wave function can only express the fer-\nromagnetic states, i.e.the form cannot represent the\nantiferromagnetic state ( n1,n0,n−1) = (n/2,0,n/2) or\nthe polar state ( n1,n0,n−1) = (0,n,0). This restriction\nof the wave function is caused by the first representation\nFz=n/planckover2pi1cosθ.\nBy introducing this representation into Eqs. (6) and\n(10), we can redefine as follows,\nFx=n/planckover2pi1sinθ(cosϕrcosϕ−cosθsinϕrsinϕ),\nFy=−n/planckover2pi1sinθ(cosϕrsinϕ+cosθsinϕrcosϕ),and,\nJx=n/planckover2pi12\n4M/braceleftbigg\nsinθ(1+cosθ)cos(ϕ1−ϕ0)∇ϕ1\n+ sinθ(1−cosθ)cos(ϕ−1−ϕ0)∇ϕ−1\n−2sinθ(cosϕrcosϕ−cosθsinϕrsinϕ)∇ϕ0\n+ 2(cosϕrsinϕ+cosθsinϕrcosϕ)∇θ/bracerightbigg\n,(17a)\nJy=−n/planckover2pi12\n4M/braceleftbigg\nsinθ(1+cosθ)sin(ϕ1−ϕ0)∇ϕ1\n−sinθ(1−cosθ)sin(ϕ−1−ϕ0)∇ϕ−1\n+ 2sinθ(cosϕrsinϕ+cosθsinϕrcosϕ)∇ϕ0\n+ 2(cosϕrcosϕ−cosθsinϕrsinϕ)∇θ/bracerightbigg\n,(17b)\nJz=n/planckover2pi12\n4M{(1+cosθ)2∇ϕ1−(1−cosθ)2∇ϕ−1},\n(17c)\nwhereϕr= (ϕ1+ϕ−1−2ϕ0)/2 andϕ= (ϕ1−ϕ−1)/2 are\nrelative phases. Since the relation ϕr= 0 was satisfied in\nour calculations, we used the relation in Eqs. (17), and\nthe spin density vectorformed an azimuthal angle ϕwith\nthexaxis. Then, we derived the spin components Fx=\nn/planckover2pi1cosϕsinθ,Fy=n/planckover2pi1cosϕsinθ, andFz=n/planckover2pi1cosθ. We\ncan therefore rewrite the spin density currents,\nJx=n/planckover2pi12\n4M(4cosϕsinθ∇ϕ0\n+2cosϕsinθcosθ∇ϕ−2sinϕ∇θ),(18a)\nJy=−n/planckover2pi12\n4M(4sinϕsinθ∇ϕ0\n+2sinϕsinθcosθ∇ϕ+2cosϕ∇θ),(18b)\nJz=n/planckover2pi12\n4M{4cosθ∇ϕ0+4(1+cos2θ)∇ϕ},(18c)\nwhich are driven by the gradients of the angles, ϕandθ,\nandthephase ϕ0. IntheprecessionswithMDDI,thegra-\ndients occurred because of the dipolar fields Hdd(r). As\na result, the spin currents were clearly driven, as shown\nin Fig. 6. For ωxt= 0.12, the spin vectors were coher-\nent just after the applied π/20 pulse (Fig. 6 (a)). The\nspin densities, FxandFy, then flowed to the center of\nthe condensates from Fig. 6 (b) to (c). Then, the den-\nsities reversed, and diffused outward from Fig. 6 (d) to\n(e). This oscillation was repeated. Of course, we cannot\nobtain the spin current without the dipolar interactions,\nsince the gradients of θandϕwere not caused; the dy-\nnamics are shown in Fig. 7.\nIn order to investigate the spin fluid dynamics, we cal-\nculated the spin current Jxfor Eq. (17), as shown in\nFigs. 8 and 9. These figures represent Jxfrom the pre-\nvious calculations with λ= 1 and 1.5 respectively. De-\nspite the difference in the ratio, we observed two com-\nmon properties in these figures. The direction of the\ncurrents changed rapidly, corresponding to the large pre-\ncession frequency, and the magnitudes changed slowly7\n1.961×10−3\n0\n!\"# !$# 3.152×10−3\n0\n 0\n3.447×10−3!%#\n3.168×10−3\n0\n!.113×10−3\n0\n!'# !(#3.043×10−3\n0\nωxt=0.12 ωxt=8 ωxt=16\nωxt=32 ωxt=40 ωxt=48\nFIG. 6: (Color online) Dynamics of Fprojected onto the x-\nyplane for λ= 1.5 with dipolar interaction.\nwith the small dipolar frequency, as shown in Fig. 10,\nwhich shows the time development of the xcomponent\nofJx(x= 4,y= 0). This figure indicates that the oscil-\nlation of the current direction occurred with the preces-\nsion frequency, which varied in magnitude with changing\ndipolar frequency. Eq. (11) also indicates that the spin\ndensity was not conserved because of the effective mag-\nnetic field. Therefore, the spin currents can be driven\nfrom a source and sink in the center of the condensates,\nas in Figs. 8 and 9. The two common properties were\ninsensitive to the value of λ. However, the change in spin\ndensity for λ= 1.5 exhibited quadratic pole motion in\na scissors-like mode for mass density [24], which can be\nunderstood as an oscillationbetween the spin density mi-\ngrating to the yaxis from the xaxis and back again, as\nshown in Figs. 6 (a) to (c). Therefore, the spin collec-\ntive mode was caused by spin diffusions induced by the\nMDDI. Therefore, the spin current causes the dynam-\nics of spin scissors-like mode, which was observed as a\nshrinking and expansion of the spin density in Fig. 6.\nThe shrinking and expansion were common features for\nλ= 1 and 1.5. However, the spin currents were affected\nby the symmetry of the traps, as shown in Figs. 8 and 9.\nFrom the calculations, we expected that the spin cur-\nrent would be observable when using the spinor BECs.\nRecently, spin current is focused from fields of spintron-\nics. However, it is difficult to observe the spin current in\nmetals and condensed matter. Atomic BECs, a macro-\nscopic quantum phenomenon, can show the spin current\nclearly and directly in the dynamics of the spinor den-\nsities. Therefore, we should attempt to observe various\nspin currents utilizing tunable experimental parameters,\ni.e.interaction parameters, trap frequencies, and the\nnumber of particles.\nV. CONCLUSION\nWe investigated the properties of magnetic resonance\nin spinor dipolar BECs by calculating the GP equations,\nobtainingKittel-likeequationsastheequationsofmotion!\"# !$# !%#\n!&# !'# !(#0\n1.965×10−3\n0\n1.965×10−3\n0\n1.965×10−3\n0\n1.965×10−3\n0\n1.965×10−3\n0\n1.965×10−3ωxt=0.12 ωxt=8 ωxt=16\nωxt=32 ωxt=40 ωxt=48\nFIG. 7: (Color online) Dynamics of Fprojected onto the x-\nyplane for λ= 1.5 without dipolar interaction.\nωxt=0.12 ωxt=8 ωxt=16\nωxt=32 ωxt=40 ωxt=481.95×10−4\n0\n!\"# !$# !%#\n!&# !'# !(#0\n2.69×10−4\n0\n9.8×10−5\n0\n4.6×10−5\n02.73×10−4\nFIG. 8: (Color online) Dynamics of the spin currents Jxpro-\njected onto the x−yplane for λ= 1 with dipolar interaction.\nThe vectors are nondimensionalized.\nfor the spin density vector. The equations revealed two\nproperties. One is the dynamics of the spin fluid, and\nthe other is precession under the effective magnetic field\nconsisting of the external magnetic fields and the dipolar\nfields. The magnetic resonancewith the properties of the\nspin fluid was characteristic of this system.\nIn order to extract properties from the GP equations,\nwe studied the law of conservation of spin density cur-\nrent without effective magnetic fields by first deriving\nthe continuity equations from the GP equation, obtain-\ning representations of the spin current. Second, we ana-\nlytically evaluated the precession dynamics described by\nthe Kittel equations derived from the GP equations us-\ning a single-mode approximation, where the Kittel equa-\ntions show conventional FMR. The analysis clearly in-\ndicated that the origin of the FMR in the BECs is like\nthe dipolar field, whereas the origin of the resonance in\nthe Kittel equations for condensed matter is the demag-\nnetizing field. Comparing the FMR of the BEC with\nthat of the condensed matter, we concluded that the ori-\ngin of the resonance was not the spin exchange interac-\ntion that causes magnetism in condensed matter, but the8\nωxt=0.12 ωxt=8 ωxt=16\nωxt=32 ωxt=40 ωxt=48!\"# !$# !%#\n!&# !'# !(#0\n3.59×10−4\n0\n2.76×10−4\n0\n2.13×10−4\n0\n8.5×10−5\n05.5×10−5\nFIG. 9: (Color online) Dynamics of the spin currents Jxpro-\njectedontothe x−yplanefor λ= 1.5withdipolarinteraction.\n!\"#\"\"\"$\"\"#\"\"\"$\n\" %\" $\" &\" '\" (\"!\"#\"\"\"$\"\"#\"\"\"$\n\" %\" $\" &\" '\" (\"{Jx}x(x=4,y=0)\nωxtλ=1 λ=1.5\nωxt!\"# !$#\n!\n! \" #!\n! \" #\nFIG. 10: Dynamics of the xcomponent of Jxatx= 4 and\ny= 0. The inter figures are the results for ωxt= 0 to 4.anisotropy of the MDDI. Finally, we numerically calcu-\nlated the GP equations, representing the dynamics with\nthe twocommon properties. The characteristicdynamics\nshowed that the effective magnetic field introduced spin\ndiffusion into the Larmor precession, driving the spin-\ncurrent-like scissors modes.\nThe relation between the spin current and FMR has\nnot yet been discussed for typical FMR. Therefore, it is\nimportant to study spin current in condensates. We also\nbelieve that the study of spin current will be useful for\nthe development of spintronics, because it is difficult to\ndirectly observe spin currents in condensed matter spin-\ntronics.\nVI. ACKNOWLEDGMENT\nM. Y. acknowledges the support of a Research Fellow-\nship of the Japan Society for the Promotion of Science\nfor Young Scientists (Grant No. 209928). M. T. ac-\nknowledges the support of a Grant-in Aid for Scientific\nResearch from JSPS (Grant No. 21340104).\n[1] C. P. Slichter, Principles of Magnetic Resonance , (Berlin:\nSpringer–Verlag, 1990).\n[2] J. H. E. Griffiths, Nature 158, 670 (1946).\n[3] C.Kittel, Phys.Rev. 71, 270(1947) ; ibid.73, 155(1948).\n[4] C. Kittel, Introduction to Solid State Physics, 8th ed.\n(John Wiley and Sons Inc., USA, 2005)\n[5] A. J. Leggett, Rev. Mod. Phys. 47, 331 (1975).\n[6] D. Vollhard and P. W¨ olfle, The Superflluid Phasess of\nHelium 3 (Taylor and Francis, London, 1990).\n[7] R. A. Webb, R. L. Kleinberg, and J. C. Wheatley, Phys.\nRev. Lett. 33, 145 (1974).\n[8] K. B. Davis, M. -O. Mewes, M. R. Andrews, N. J. van\nDruten, D. S. Durfee, D. M. Kurnand W. Ketterle, Phys.\nRev. Lett. 75, 3969 (1995).\n[9] M. H. Anderson, J. R. Ensher, M. R. Matthews, C. E.\nWieman and E. A. Cornell, Science 269, 198 (1995).\n[10] S. Yi, L. You and H. Pu, Phys. Rev. Lett. 93, 040403\n(2004)\n[11] Y. Kawaguchi, H. Saito and M. Ueda, Phys. Rev. Lett.\n97, 130404(2006).\n[12] H.M¨ akel¨ aandK.-A.Suominen, Phys.Rev.A 75, 033610\n(2007).\n[13] Y. Kawaguchi, H. Saito and M. Ueda, Phys. Rev. Lett.96, 080405 (2006).\n[14] A. Griesmaier, J. Werner, S. Hensler, J. Stuhler, and T.\nPfau, Phys. Rev. Lett. 94, 160401 (2005).\n[15] J. Stuhler, A. Griesmaier, T. Koch, M. Fattori, T. Pfau,\nS. Giovanazzi, P. Pedri, and L. Santos, Phys. Rev. Lett.\n95, 150406 (2005).\n[16] T. Lahaye, T. Koch, B. Fr¨ ohlich, M. Fattori, J. Metz, A.\nGriesmaier, S. Giovanazzi, and T. Pfau, Nature 448, 672\n(2007).\n[17] E. L. Hahn, Phys. Rev. 80, 580 (1950).\n[18] H. Y. Carr and E. M. Purcell, Phys. Rev. 94, 630 (1954).\n[19] M. Yasunaga and M. Tsubota, Phys. Rev. Lett. 101\n220401 (2008).\n[20] M. Yasunaga and M. Tsubota, J. Low Tem. Phys. 158,\n51 (2010).\n[21] W. Zhang, D. L. Zhou, M. -S. Chang, M. S. Chapman,\nand L. You, Phys. Rev. A 72, 013602 (2005).\n[22] S. Yi and H. Pu, Phys. Rev. Lett. 97, 020401 (2006).\n[23] T. -L. Ho, Phys. Rev. Lett. 81, 742 (1998).\n[24] C. J. Pethick and H. Smith, Bose-EInstein Condensation\nin Dilute Gases, 2nd ed. (Cambridge, New York, 2008)." }, { "title": "2011.05566v1.Reduction_of_back_switching_by_large_damping_ferromagnetic_material.pdf", "content": "arXiv:2011.05566v1 [cond-mat.mes-hall] 11 Nov 2020Reduction of back switching by large damping ferromagnetic material\nTomohiro Taniguchi1, Yohei Shiokawa2, and Tomoyuki Sasaki2\n1National Institute of Advanced Industrial Science and Tech nology (AIST),\nResearch Center for Emerging Computing Technologies, Tsuk uba, Ibaraki 305-8568, Japan,\n2TDK Corporation, Advanced Products Development Center, Ic hikawa, Chiba 272-8558, Japan\nRecent studies on magnetization dynamics induced by spin-o rbit torque have revealed a weak\ndependence of the critical current for magnetization switc hing on the damping constant of a ferro-\nmagnetic free layer. This study, however, reveals that the d amping constant nevertheless plays a key\nrole in magnetization switching induced by spin-orbit torq ue. An undesirable switching, returning\nto an initial state, named as back switching, occurs in a ferr omagnet with an easy axis parallel\nto the current direction. Numerical and theoretical analys es reveal that back switching is strongly\nsuppressed when the damping constant of the ferromagnet is l arge.\nLow-damping ferromagnetic materials has been inves-\ntigated for spintronics applications [1–6]. These materi-\nals are interesting because the critical current for excit-\ning magnetization dynamics by spin-transfer torque[7, 8]\nin two-terminal devices is typically proportional to the\ndampingconstant[9,10], andtherefore, low-dampingfer-\nromagnetic materials help to reduce power consumption.\nThe value of the damping constant in typical ferromag-\nnets, such as Fe, Co, Ni, and their alloys, is on the order\nof 10−3−10−2, even after the effect of spin pumping\n[11, 12] is taken into account [5, 6]. Studies on three-\nterminal devices manipulated by spin-orbit torque [13–\n24], however, have revealed that the dependence of the\ncritical current on the damping constant is weak when\nthe easy axis of the ferromagnet is perpendicular to the\nfilm plane [25, 26] or parallelto the currentdirection [27];\nthese systems are named type Z and type X, respectively,\nin Ref. [19]. Then, a question arises as to what the role\nof the damping constant is in three-terminal devices.\nThe purpose of this study is to investigate the rela-\ntion between the magnetizationstate manipulated by the\nspin-orbit torque and the magnetic damping constant by\nsolving the Landau-Lifshitz-Gilbert (LLG) equation. We\nfocus onthe type-Xsystembecause the switchingmecha-\nnism in the system is not fully understood yet, unlike the\nothersystems[9,10,25,26]. Anundesirablereturnofthe\nmagnetization,calledbackswitchinginthispaper,occurs\nin a ferromagnet, where, although the spin-orbit torque\nbrings the magnetization close to the switched state, it\nreturns to the initial state after turning off the current.\nAs a result, the phase diagramofthe magnetization state\nas a function of the current and external magnetic field\nalternately shows switched and non-switched states. It\nis revealed that the back switching appears as a result of\nthe magnetization precession around the perpendicular\naxis after the current is turned off. The back-switching\nregion is strongly suppressed by using large-damping fer-\nromagnetic materials.\nThe system studied in the work is an in-plane mag-\nnetized ferromagnet placed on a nonmagnetic metal and\nschematically shown in Fig. 1(a). Electric current flow-\ning in the nonmagnet in the xdirection generates pure\nspincurrentthatisinjectedintotheferromagnetwiththez\nxy\njm\nHappl\nj=300 MA cm-2Happl=500 Oemx1\n0\n-1\nTime (ns)0 1(a) (b)\nFIG. 1: (a) Schematic illustration of the system. The spin\nHall effect in the bottom nonmagnet generates spin current\nwith a polarization along the ydirection and excites the mag-\nnetization ( m) dynamics in the top ferromagnet. The easy\naxis of the ferromagnet is parallel to the xaxis. An external\nmagnetic field Happlis applied in the zdirection. (b) Exam-\nple of the time evolution of mxforj= 300 MA cm−2and\nHappl= 500 Oe.\nspin polarization in the ydirection. The spin current ex-\ncitesthespin-transfertorqueactingonthemagnetization\nin the ferromagnet and induces magnetization dynamics.\nThezaxis is perpendicular to the film plane. We denote\nthe unit vectors pointing in the magnetization direction\nin the ferromagnet and along the k-axis (k=x,y,z) as\nmandek, respectively. The magnetization dynamics are\ndescribed by the LLG equation\ndm\ndt=−γm×H−γHsm×(ey×m)+αm×dm\ndt,(1)\nwhereγandαarethe gyromagneticratioandthe Gilbert\ndamping constant, respectively. The magnetic field H\nconsists of the in-plane magnetic anisotropy field HKin\nthexdirection, the demagnetization field −4πMin thez\ndirection, and an external field Happlin thezdirection:\nH=HKmxex+(Happl−4πMmz)ez.(2)\nThe strength of the spin-transfer torque is given by\nHs=/planckover2pi1ϑj\n2eMd, (3)\nwhereϑis the spin Hall angle in the nonmagnet, whereas\nManddare the saturation magnetization and thickness\nof the ferromagnet. The electric current density is de-\nnoted as j. The values of the parameters used in this2\n\"mx.dat\" u 1:2:3 \"mx_relax.dat\" u 1:2:3 (a)\nApplied field, Happl (kOe)Current density, j (MA cm -2 )\n0 1 2 0 1 202004006008001000 1\n0\n-1 mx\nApplied field, Happl (kOe)1\n0\n-1 mx\nS0 S1 S2\njc(b)\nCurrent density, j (MA cm -2 )\n02004006008001000\nFIG. 2: Phase diagrams of mx(a) at a fixed point in the\npresence of current and (b) in a relaxed state after turning o ff\nthe current. The initial state is m0−. The white dotted line in\n(a) represents the theoretical formula of the critical curr ent\ndensity derived in Ref. [27]. Switching regions are distin-\nguished by the labels S n(n= 0,1,2) in (b). The damping\nconstant αis 0.005.\nstudy were taken from typical experiments [13–23] as\nM= 1500 emu c.c.−1,HK= 200 Oe, γ= 1.764×107\nrad Oe−1s−1,ϑ= 0.4, andd= 1.0 nm.\nThe magnetic field His related to the magnetic energy\ndensityEviaE=−M/integraltext\ndm·H,\nE=−MHapplmz−MHK\n2m2\nx+2πM2m2\nz.(4)\nThe energy density Ehas two minima at m0±=\n±/radicalbig\n1−m2\n0zex+m0zez, wherem0z=Happl/(HK+4πM).\nThroughout this paper, the initial state is set to be m0−,\nwhich points in the negative xdirection. Accordingly, we\ncall the other stable state, m0+pointing in the positive x\ndirection, the switched state. Thus, we are interested in\nexperiments where the initial state is reset to m0−dur-\ning each trial ofmagnetization switching. By convention,\nwe will focus on switching by a positive current and field\nregion.\nFigure 1(b) shows an example of the time evolution of\nmxin the presence of current. The damping constant is\nα= 0.005. The results indicate that the magnetization\nsaturates to a fixed point. Figure 2(a) is a phase diagram\nsummarizing the fixed point of mxin the presence of cur-\nrent, where the vertical and horizontal axes represent the\ncurrent density jand the external magnetic field Happl.\nThe magnetization stays near the initial state [ mx≃ −1\nshown in black in Fig. 2(a)] in a relatively small cur-\nrent region, where its boundary is well explained by the\ncritical-currentformula jc[27] shownby the white dotted\nline in Fig. 2(a). On the other hand, the magnetization\nstate above the critical current satisfies mx>0. There-\nfore, one might suppose that the magnetization relaxes\nto the switched state, mx≃+1, after turning off the\ncurrent.\nHowever, the phase diagram after turning off the cur-\nrentshowninFig. 2(b) revealsacomplicateddependence\nof the relaxed state on the current and external magnetic\nfield. Therearestripesdistinguishingtheswitched( m0+,\nyellow) and non-switched ( m0−, black) states. This re-\nsult indicates that the magnetization returns to the ini-tial state under certain conditions. We name this phe-\nnomenon back switching and investigate its relation to\nthe damping constant in the following.\nLet us briefly comment onanotherphenomenon, called\nbackhopping [28–32], to avoid any confusion. Back-\nhopping is a phenomenon in which, after magnetization\nswitchinginafreelayerisachieved,anundesirablereturn\nto the original resistance-state occurs at a high-bias volt-\nage. There are two differences between back switching\nandbackhopping. First, backhoppingintwo-terminalde-\nvices originates from magnetization switching in the ref-\nerence layer. On the other hand, back switching relates\nto the magnetization dynamics in the free layer. Second,\nbackhopping has often been investigated by sweeping the\ncurrent. On the other hand, we reset the initial state of\nthe magnetization at each trial because we are interested\nin designing the operation conditions of memory devices.\nBack switching originates from the fact that the fixed\npoint of the magnetization in the presence of current is\nnot parallel to the easy axis of the magnetization, as dis-\ncussed below, and thus, it will occur in not only type-X\ndevices, but also type-Z devices [25, 26].\nNote that the horizontal stripe in the relatively low\ncurrent region of Fig. 2(b), slightly above jc, was ana-\nlyzed in Ref. [27], and therefore, it will be excluded from\nthe following discussion. We label the other switched\nregions shown by the vertical stripes as the S n-region\n(n= 0,1,2,···); see Fig. 2(b). The role of the integer n\nwill be clarified below.\nFigure 3(a) shows the dynamic trajectory of the mag-\nnetization obtained by numerically solving Eq. (1),\nwherej= 300 MA cm−2andHappl= 500 Oe correspond\nto the S 0region. The red line represents the trajectory\nfrom the initial state to the fixed point in the presence of\nthe current, whereas the blue line shows the relaxation\ndynamics after turning off the current. The fixed point\nsatisfying dm/dt=0in the presence of the current is\nindicated by a yellow circle. Note that the fixed point is\nfar away from the energetically stable states, m0±. This\nis because the spin-transfer torque drives the magnetiza-\ntion in the ydirection. The black solid line in Fig. 3(b)\nshows the time evolution of mxafter turning off the cur-\nrent. Starting from the fixed point located in the region\nofmx>0, the magnetization relaxes to the switched\nstatem0+after showing several precessions around it.\nFigures 3(c) and 3(d), on the other hand, show an exam-\nple of back switching, where the parameter Happl= 750\nOe corresponds to the region sandwiched by the S 0and\nS1-regions. In this case, the magnetization after turning\noff the current shows a precession around the zaxis. As\na result, even though the fixed point in the presence of\nthe current is in the region of mx>0, the magnetization\nreturns to the initial state ( m0−). Let us then move to\nthe next switching region, the S 1-region. The dynamic\ntrajectory, as well as mx, shown in Figs. 3(e) and 3(f),\nindicates that, starting from the fixed point in the region\nofmx>0, the magnetization returns once to the region\nofmx<0 before relaxing to the switched state. Now3\n(a) (b)\nj=300 MA cm-2Happl=500 Oe\nj=300 MA cm-2Happl=750 Oe\nj=300 MA cm-2Happl=1000 Oemz1\n0\n-1\n0-1\n-111\n0mx mymx1\n0\n-1\nTime (ns)0 1m0+m0-\n(e) (f)\nmz1\n0\n-1\n0-1\n-111\n0mx mymx1\n0\n-1\nTime (ns)0 1m0+m0-(c) (d)\nmz1\n0\n-1\n0-1\n-111\n0mx mymx1\n0\n-1\nTime (ns)0 1m0+m0-\nNormalized energy (×10-3) Normalized energy (×10-3) Normalized energy (×10-3)1\n0\n-1\n-2\n-3\n-4\n12\n0\n-1\n-2\n-3\n-4\n1234\n0\n-1\n-2\n-3\n-4\nFIG. 3: (a) Dynamictrajectory in theS 0-region;Happl= 500\nOe, and j= 300 MA cm−2. The red line represents the\ntrajectory from the initial state to the fixed point (yellow\ncircle) in the presence of the current, whereas the blue line\ncorresponds to the relaxation dynamics after turning off the\ncurrent. The symbols m0±indicate the locations of the initial\n(m0−) and switched ( m0+) states. (b) Time evolutions of mx\n(black solid line) and normalized energy density ε(dashed red\nline) after turning off the current. The red triangle indicat es\nthe time at which εbecomes zero. (c) Dynamic trajectory\nand (d) time evolution of mxandεin the region sandwiched\nby the S 0- and S 1-regions; Happl= 750 Oe, and j= 300 MA\ncm−2. (e) Dynamic trajectory and (f) time evolution of mx\nandεin the S 1-region; Happl= 1000 Oe, and j= 300 MA\ncm−2.\nthe meaning of the integer n(n= 0,1,2,···) we used to\ndistinguish the switched regions becomes clear: it repre-\nsents how many times the magnetization returns to the\nregion of mx<0 before relaxing to the switched state.\nThese results imply that the precession around the z\naxis after turning off the current is the origin of back\nswitching. Such a precession is induced by the precession\ntorque,−γm×H. In particular, the precession around\nthezaxis occurs when the energy density at the fixed\npoint of the magnetization in the presence of the current\nis larger than the saddle-point energy given by\nEd=−MH2\nappl\n8πM. (5)In fact, a fixed point of the LLG equation is given by\nmx=HapplHs\nH2s−HK4πM, mz=−HapplHK\nH2s−HK4πM.(6)\nSubstituting Eq. (6) into Eq. (4), the energy density at\nthis fixed point is\nEj=MH2\napplHK\n2(H2s−HK4πM). (7)\nNote that H2\ns−HK4πM >0 because the back switch-\ning appears in the current region above jcforHappl=\n0, which is given by jc= [2eMd/(/planckover2pi1ϑ)]√HK4πM[27].\nTherefore, Eq. (7) is alwayspositive, whereasthe saddle-\npoint energy density given by Eq. (5) is negative. Ac-\ncordingly, the fixed point in the presence of the current is\nin the unstable region corresponding to an energy larger\nthan the saddle-point energy. As a result, after turn-\ning off the current, the magnetization starts to precess\naround the zaxis, as mentioned above. Simultaneously,\nthe magnetization loses energy due to damping torque,\nwhere the dissipated energy density is given by\n∆E(t) =αγM\n1+α2/integraldisplayt\n0dt′/bracketleftBig\nH2−(m·H)2/bracketrightBig\n.(8)\nHere,tis the time after the current is turned off. When\nthe condition,\nEj−∆E(t) =Ed, (9)\nis satisfied, the magnetization relaxes to the nearest sta-\nble state. The point here is that the magnetization al-\nternately comes close to two stable states, m0±, due to\nthe precession around the zaxis before Eq. (9) is sat-\nisfied. As a result, both m0+andm0−can be the final\nstate, which leads to back switching. The conclusions of\nthe discussion are confirmed by the dashed red lines in\nFigs. 3(b), 3(d), and 3(f), where are the time evolutions\nof the energy density. Here, we introduce the normalized\nenergy density,\nε(t) =E(t)−Ed\n4πM2. (10)\nTheenergydensity Eisafunction oftime with the initial\ncondition of E(t= 0) =EjbecauseE, given by Eq. (4),\ndepends on the magnetization direction, and the magne-\ntization changes direction in accordance with the LLG\nequation. The condition given by Eq. (9) is ε(t) = 0\nin terms of the normalized energy density. The times at\nwhichεbecomes zero are indicated by the red triangles\nin the figures. It is clear from Figs. 3(b), 3(d), and 3(f)\nthat the final state of the magnetization is determined by\nwhether mxis positive or negative when εbecomes zero.\nIt is expected that using a large-damping ferromag-\nnetic material will result in a reduction of the back-\nswitching region. Remember that back switching occurs\ndue tothe precessionaroundthe zaxis. When the damp-\ning constant is large, the energy dissipation given by Eq.4\n\"mx_relax.dat\" u 1:2:3 \"mx.dat\" u 1:2:3 mx1\n0\n-1 \nTime (ns)0 1\nNormalized energy (×10 -3 )\nNormalized energy (×10 -3 )\n2\n0\n-2 \n-6 -4 mx1\n0\n-1 \nTime (ns)0 124\n0\n-2 \n-6 -4 \n(c)(a)\nApplied field, Happl (kOe)Current density, j (MA cm -2 )\n0 1 2 0 1 202004006008001000 1\n0\n-1 mx\nApplied field, Happl (kOe)1\n0\n-1 mx(d)(b)\nCurrent density, j (MA cm -2 )\n02004006008001000\nFIG. 4: Time evolutions of mx(black solid lines) and nor-\nmalized energy density ε(red dashed lines) after turning off\nthe current, where j= 300 MA cm−2and (a)Happl= 750 Oe\nand (b) 1000 Oe. The damping constant αis 0.050. The red\ntriangles indicate the time at which εbecomes zero. Phase\ndiagrams of mx(c) at a fixed point in the presence of current\nand (d) in a relaxed state after turning off the current.\n(8) becomes large within a short time t, and the condi-\ntion given by Eq. (9) is immediately satisfied before the\nmagnetizationreturnstothe regionof mx<0bythepre-\ncession around the zaxis. Accordingly, back switching\ndoes not occur. On the other hand, when the damping\nconstant is small, it takes a long time to dissipate the\nenergy to satisfy Eq. (9), during which time the mag-\nnetization shows the precession around the zaxis. As a\nresult, back switching appears.\nTo verify this picture, we evaluated the phase diagram\nof the magnetization state for a relatively large damping\nconstant, α= 0.050, which is ten times large than that\nused above. Figures 4(a) and 4(b) show the time evo-\nlutions of mxandεafter turning off the current, where\nHapplis (a) 750 Oe and (b) 1000 Oe. Contrary to the\ndynamics shown in Figs. 3(d) and 3(f), the magnetiza-tion in Figs. 4(a) and 4(b) does not return to the region\nofmx<0 because the energy dissipates quickly due to\nthe large damping torque. As a result, the magnetiza-\ntion immediately relaxes to the switched state. Figures\n4(c) and 4(d) summarize the magnetization state in the\npresence of current and after turning off the current, re-\nspectively. Comparing Fig. 4(c) with Fig. 2(a), it is\nclear that the phase diagram of the magnetization state\nin the presence of the current is approximately indepen-\ndent of the damping constant. On the other hand, the\nback-switching region is strongly suppressed, as can be\nseen by the comparing Fig. 2(b) and 4(d). This is be-\ncause the large damping torque immediately dissipates\nthe energy from the ferromagnet and leads to a fast re-\nlaxationtothe switched state, due to whichthe S 0-region\ndominates in the phase diagram.\nThe existence of the back switching gives an upper\nlimit of a write-current margin for memory applications,\nand therefore, it restricts the device design and/or ma-\nnipulationconditions. Ontheotherhand, backswitching\nmight be applicable to other devices such as a random\nnumber generator. The analysis in this study provides\nfruitful insights for the development of spintronics appli-\ncations utilizing spin-orbit torque.\nIn conclusion, the phase diagram of the magnetization\nstate in a type-X spin-orbit torque device was calculated\nasafunction ofthe electriccurrentdensity and the exter-\nnal magneticfield. The magnetizationstate afterturning\noff the current has stripe structures alternately showing\nswitched and non-switched states. Such a non-switched\nstate, named back switching here, occurs as a result of\nthe magnetization precession around the perpendicular\naxis after the current is turned off. The back-switching\nregion is reduced by using large-damping ferromagnetic\nmaterials, whereas the critical current inducing the mag-\nnetization switching is approximately independent of the\ndamping constant.\nThe authors thank to Masamitsu Hayashi, Shinji\nIsogami, Toru Oikawa, Takehiko Yorozu, and Seiji Mi-\ntani for valuable discussion. This work was supported by\nfunding from TDK Corporation.\n[1] M. Oogane, T. Wakitani, S. Yakata, R. Yilgin, Y. Ando,\nA. Sakuma, and T. Miyazaki, Jpn. J. Appl. Phys. 45,\n3889 (2006).\n[2] C. Bilzer, T. Devolder, J.-V. Kim, G. Counil, C. Chap-\npert, S. Cardoso, and P. P. Freitas, J. Appl. Phys. 100,\n053903 (2006).\n[3] S. Mizukami, H. Abe, D. Watanabe, M. Oogane,\nY. Ando, and T. Miyazaki, Appl. Phys. Express 1,\n121301 (2008).\n[4] S. Iihama, Q. Ma, T. Kubota, S. Mizukami, Y. Ando,\nand T. Miyazaki, Appl. Phys. Express 5, 083001 (2012).\n[5] M. Konoto, H. Imamura, T. Taniguchi, K. Yakushiji,H. Kubota, A. Fukushima, K. Ando, and S. Yuasa, Appl.\nPhys. Express 6, 073002 (2013).\n[6] S. Tsunegi, H. Kubota, S. Tamaru, K. Yakushiji,\nM. Konoto, A. Fukushima, T. Taniguchi, H. Arai,\nH. Imamura, and S. Yuasa, Appl. Phys. Express 7,\n033004 (2014).\n[7] J. C. Slonczewski, J. Magn. Magn. Mater. 159, L1\n(1996).\n[8] L. Berger, Phys. Rev. B 54, 9353 (1996).\n[9] J. Z. Sun, Phys. Rev. B 62, 570 (2000).\n[10] J. Grollier, V. Cros, H. Jaffr´ es, A. Hamzic, J. M. George ,\nG. Faini, J. B. Youssef, H. LeGall, and A. Fert, Phys.5\nRev. B67, 174402 (2003).\n[11] Y. Tserkovnyak, A. Brataas, and G. E. W. Bauer, Phys.\nRev. Lett. 88, 117601 (2002).\n[12] S. Mizukami, Y. Ando, and T. Miyazaki, J. Magn. Magn.\nMater.239, 42 (2002).\n[13] L. Liu, O. J. Lee, T. J. Gudmundsen, D. C. Ralph, and\nR. A. Buhrman, Phys. Rev. Lett. 109, 096602 (2012).\n[14] C.-F. Pai, L. Liu, Y. Li, H. W. Tseng, D. C. Ralph, and\nR. A. Buhrman, Appl. Phys. Lett. 101, 122404 (2012).\n[15] G. Yu, P. Upadhyaya, Y. Fan, J. G. Alzate, W. Jiang,\nK. L. Wong, S. Takei, S. A. Bender, L.-T. Chang,\nY. Jiang, et al., Nat. Nanotechnol. 9, 548 (2014).\n[16] M. Cubukcu, O. Boulle, M. Drouard, K. Garello, C. O.\nAvci, I. M. Miron, J. Langer, B. Ocker, P. Gambardella,\nand G. Gaudin, Appl. Phys. Lett. 104, 042406 (2014).\n[17] L. You, O. L. nd D. Bhowmik, D. Labanowski, J. Hong,\nJ. Bokor, and S. Salahuddin, Proc. Natl. Acad. Sci. 112,\n10310 (2015).\n[18] J. Torrejon, F. Garcia-Sanchez, T. Taniguchi, J. Shinh a,\nS. Mitani, J.-V. Kim, and M. Hayashi, Phys. Rev. B 91,\n214434 (2015).\n[19] S. Fukami, T. Anekawa, C. Zhang, and H. Ohno, Nat.\nNanotechnol. 11, 621 (2016).\n[20] S. Fukami, T. Anekawa, C. Zhang, and H. Ohno, Nat.\nMater.15, 535 (2016).\n[21] Y.-C. Lau, D. Betto, K. Rode, J. M. D. Coey, and P. Sta-\nmenov, Nat. Nanotechnol. 11, 758 (2016).\n[22] A. van den Brink, G. Vermijs, A. Solignac, J. Koo, J. T.Kohlhepp, H. J. M. Swagten, and B. Koopmans, Nat.\nCommun. 7, 10854 (2016).\n[23] Y.-W. Oh, S. h. C. Baek, Y. M. Kim, H. Y. Lee, K.-\nD. Lee, C.-G. Yang, E.-S. Park, K.-S. Lee, K.-W. Kim,\nG. Go, et al., Nat. Nanotechnol. 11, 878 (2016).\n[24] Y. Shiokawa, E. Komura, Y. Ishitani, A. Tsumita,\nK. Suda, Y. Kakinuma, and T. Sasaki, AIP Adv. 9,\n035236 (2019).\n[25] K.-S. Lee, S.-W. Lee, B.-C. Min, and K.-J. Lee, Appl.\nPhys. Lett. 102, 112410 (2013).\n[26] T. Taniguchi, Phys. Rev. B 100, 174419 (2019).\n[27] T. Taniguchi, Phys. Rev. B 102, 104435 (2020).\n[28] J. Z. Sun, M. C. Gaidis, G. Hu, E. J. O’Sullivan,\nS. L. Brown, J. J. Nowak, P. L. Trouilloud, and D. C.\nWorledge, J. Appl. Phys. 105, 07D109 (2009).\n[29] S.-C. Oh, S.-Y. Park, A. Manchon, M. Chshiev, J.-H.\nHan, H.-W. Lee, J.-E. Lee, K.-T. Nam, Y. Jo, Y.-C.\nKong, et al., Nat. Phys. 5, 898 (2009).\n[30] W. Skowro´ nski, P. Ogrodnik, J. Wrona, T. Stobiecki,\nR.´Swirkowicz, J. Barna´ s, G. Reiss, and S. van Dijken, J.\nAppl. Phys. 114, 233905 (2013).\n[31] C. Abert, H. Sepehri-Amin, F. Bruckner, C. Vogler,\nM. Hayashi, and D. Suess, Phys. Rev. Applied 9, 054010\n(2018).\n[32] C. Safranski and J. Z. Sun, Phys. Rev. B 100, 014435\n(2019)." }, { "title": "2304.10709v1.Soft_Magnons_in_Anisotropic_Ferromagnets.pdf", "content": "Soft Magnons in Anisotropic Ferromagnets\nG.E.W. Bauer,1, 2, 3P. Tang,1M. Elyasi,1Y.M. Blanter,4and B.J. van Wees5\n1WPI Advanced Institute for Materials Research,\nTohoku University, 2-1-1, Katahira, Sendai 980-8577, Japan\n2Institute for Materials Research and CSIS, Tohoku University, 2-1-1 Katahira, Sendai 980-8577, Japan\n3Kavli Institute for Theoretical Sciences, University of the Chinese Academy of Sciences, Beijing 10090, China\n4Kavli Institute of NanoScience, Delft University of Technology, 2628 CJ Delft, the Netherlands\n5Physics of Nanodevices, Zernike Institute for Advanced Materials,\nUniversity of Groningen, 9747 AG Groningen, the Netherlands\n(Dated: April 24, 2023)\nWe discuss spin-wave transport in anisotropic ferromagnets with an emphasis on the zeroes of\nthe band edges as a function of a magnetic \feld. An associated divergence of the magnon spin\nshould be observable by enhanced magnon conductivities in non-local experiments, especially in\ntwo-dimensional ferromagnets.\nI. INTRODUCTION\n\\Magnonics\" is the study of the elementary excitations\nof the magnetic order, i.e. spin waves and their quanta\ncalled \\magnons\"[1{3]. It is believed to be competitive\nin future information, communication, and thermal man-\nagement technologies [4, 5].\nIn the exchange interaction-only continuum model the\nspin wave dispersion is a parabola that shifts linearly\nwith an applied magnetic \feld [1{3]. The gain of angu-\nlar momentum of the ground state by \ripping a single\nelectron is ~. The associated change of the magnetic mo-\nmentum is\u00002\u0016B, where\u0016Bis the Bohr magneton. The\nexchange energy cost of a single spin \rip is minimized by\nspreading this excitation over the whole system, forming\na spin wave or its quantum, the magnon.\nMagnetic dipolar interactions and spin-orbit interac-\ntions strongly a\u000bect the spin wave dispersion of ferro-\nmagnets. Crystal anisotropies are the main consequence\nof the latter, and cause, for example, magnon gaps in\nthe absence of an applied magnetic \feld. Only quite re-\ncently have researchers realized that the magnon spin is\nnot a universal constant. Ando et al. [6] reported en-\nhanced spin pumping by an elliptic magnetization pre-\ncession, which, as we show below, can be interpreted as\nan enhanced magnon spin. Flebus et al. [7] reported\nthat the exchange magnon polaron, i.e. the hybrid state\nof a magnon and a phonon, carries a spin between 0 and\n~. Kamra and Belzig [8] predict super-Poissonian shot\nnoise in the spin pumping from ferromagnets into metal-\nlic contacts based on a magnon spin that is enhanced\nfrom its standard value of ~by anisotropy \\squeezing\".\nThese authors start from a lattice quantum spin Hamil-\ntonian with local and magneto-dipolar anisotropies and\npredicted magnon spins of around 4 ~for the fundamental\n(Kittel) mode of an iron \flm. Kamra et al. [9] address\nthe emergence of the magnon spins in antiferromagnets\nat weak applied magnetic \felds. Neumann et al. [10]\nintroduced an \\orbital contribution\"to the magnon mag-\nnetic moment. Yuan et al. [11] review the history of the\nmagnon spin concept and its enhancement by quantumsqueezing\nMagnon currents can be injected into ferromagnetic in-\nsulators by heavy metal contacts, electrically by means\nof the spin Hall e\u000bect or by thermal gradients (spin See-\nbeck e\u000bect). Vice versa , magnons can pump a spin cur-\nrent into a heavy metal contact and be detected by an\ninverse-spin-Hall voltage. Both e\u000bects may be combined\nto study magnon transport in magnetic insulators [12].\nFilms of ferrimagnetic yttrium iron garnet are well suited\nfor magnon transport studies and can be grown with high\nquality down to a few monolayers [13]. The same tech-\nnique also works well for antiferromagnetic hematite [14{\n16].\nde Wal et al. [17] studied non-local magnon transport\nin an antiferromagnetic van der Waals \flm with a per-\npendicular N\u0013 eel vector. An in-plane magnetic \feld cants\nthe two sublattices until the material becomes ferrimag-\nnetically ordered at the \\spin-\rip\" transition, not unlike\nthe in-plane spin texture of hematite at high \felds. In\nthe absence of additional anisotropies, the band gap of\nthe spin wave dispersion vanishes at this point, i.e. the\nmagnons become \\soft\".\nHere we make a step back by realizing that magnons\ncan be \\soft\" in simple ferromagnets as well. We \fnd\nthat the magnon spin can be strongly a\u000bected by the as-\nsociated non-analyticities of the dispersion relation and\nmay even diverge. We illustrate the general concept of\n\\soft magnons\" and \\magnon spin\" for the three generic\nmagnetic con\fgurations in Figure 1 in which applied\nmagnetic \felds and uniaxial anisotropies compete. We\npredict observable enhancements of magnon transport in\nthe proximity of the soft magnon con\fgurations. These\nresults may partly explain the strong enhancement of\nthe magnon transport at the spin-\rip transition in anti-\nferromagnets [18].\nSections II and III set the stage by re-deriving results\nfor the ground and excited states of long energy excita-\ntions in magnetic systems. In Section II we analyze the\nground state by minimizing the classical magnetic free\nenergy. We solve the linearized Landau-Lifshitz equa-\ntions in Section 3 for the spin wave dispersion relations,\n\fnding results that agree with those obtained from quan-arXiv:2304.10709v1 [cond-mat.mes-hall] 21 Apr 20232\nFIG. 1. Three con\fgurations of ferromagnets with uniaxial\nmagnetic anisotropies parameterized by the constant Kand\ntheir magnetization MBin the direction of the applied mag-\nnetic \feldB. Case (A) is an easy-plane ferromagnet with a\nmagnetic \feld along z(red). In case (B) the \feld lies in the\neasy plane (green). In case (C) the \feld is normal to the easy\nz-axis and\u0012is the tilt angle of the magnetization.\ntum spin models. In Section IV we address the magnon\nspin in a way we have not found in the literature. We\nshow that the Hellman-Feynman theorem can be a use-\nful tool to get hands on the magnon magnetic moment,\nbut only when \feld and magnetization are collinear. In\nSection V we discuss possible experimental signatures of\nthe large magnon spins close to the kinks in the magnon\ndispersion. Section VI contains a critical discussion of\nthe model and recommendations for future work.\nII. SPIN HAMILTONIANS AND\nFERROMAGNETIC GROUND STATES\nWe consider Hamiltonians for local spins ^Siand mag-\nnetic moments ^Mi=\u0000\r^Sion lattice sites i:\nH=\r~X\ni^Si\u0001B\u0000X\nijJij(^Si\u0001^Sj)\u0000\u0016KX\ni(^Sz\ni)2;(1)\nwhereJijis the exchange integral between spins on sites\ni6=j,\u0000\ris the gyromagnetic ratio of an electron, and B\nis a constant external magnetic \feld. We focus here on\nJij>0 that leads to ferromagnetic order. \u0016Kis a uniax-\nial anisotropy parameter along the Cartesian z-direction,\n\u0016K > 0\u0000\u0016K < 0\u0001\ncorresponds to an easy-axis (easy-plane)\nferromagnet.\nThe ground state that minimizes the total energy of\nthe macroscopic system is ferromagnetic. The total\nspin of the system ^S=P\ni^Sithen becomes a classi-\ncal vector Scorresponding to a magnetization density\nM=\u0000(\r~=\n)S, where \n is the crystal volume. The as-\nsociated energy density without irrelevant constant termsreads\nE(M) =\u0000M\u0001B\u0000\u0016K\nM(Mz)2+D\n2M(rM)2+Edmag(M);\n(2)\nwhereM=jMj= (\r~S=\n) andS=jSj. The spin-\nwave sti\u000bness Drepresents the exchange energy cost of\nspatial deformations in the continuum limit that depends\non crystal structure and the exchange parameters Jij.\nThe magnetizations of the ground states are constant in\nspace with zero exchange energy. The demagnetization\nenergy in magnetic \flms with surface normal nalong\nzcan then be absorbed into the anisotropy \feld K=\n(S\u0016K)=(\r~) +M. We discuss here three con\fgurations:\n(A) Easyxy-plane anisotropy ( K < 0) and the magnetic\n\feld normal to the plane, (B) easy xy-plane anisotropy\nwith an in-plane magnetic \feld (Kittel problem), and (C)\neasyz-axis anisotropy ( K > 0) with a magnetic \feld in\nthey-direction. Other con\fgurations such as in-plane\neasy axis etc. give similar results.\nWe are interested in the discontinuities that emerge at\ncritical \felds Bc= 0 for case (B) and Bc= 2jKjfor (A)\nand (B).\nIn case (A) B=Bz;where zis the unit vector along\nthe anisotropy axis, the energy and magnetizations are\ndiscontinuous at Bc= 2jKj\nE(A)\n0\nM=(\n\u0000B2\n4jKj\njKj\u0000BforB < 2jKj\nB > 2jKj(3)\nMz\nM=\u001aB\n2jKj\n1forB < 2jKj\nB > 2jKj(4)\nIn case (B) the energy\nE(B)(M) =\u0000MyB+jKj\nMM2\nz (5)\nis minimal for M0= (0;M;0) for allB6= 0:Finally,\nin (C) a \feld B=Byalong they-axis tilts the magne-\ntization into the yz-plane with M0=M(0;sin\u0012;cos\u0012),\nwhere\u0012is the angle with the out-of-plane direction. Eq.\n(2)\nE(C)(\u0012)\nM=\u0000Kcos2\u0012\u0000Bsin\u0012: (6)\nis minimized for\nsin\u0012=\u001aB\n2K\n1forB B c: (7)\nAboveBc= 2K;the \feld and magnetization are aligned.\nIII. SPIN WAVE DISPERSION\nHere we consider the frequency dispersion relation\nfor the elementary excitations in homogeneous extended3\nFIG. 2. Spin wave frequency of the magnon band edges for\nthe con\fgurations (A)-(C) in Fig. 1. The dashed black line\n!0=\r=B\u0000Kis the asymptotic form of the case (B) at large\n\felds.\nmagnets that can be bulk crystals, thin \flms, or two-\ndimensional systems. The excitation frequencies are\nsharply de\fned in the limit of small amplitude oscilla-\ntions, in which spin waves can be mapped on a set of non-\ninteracting harmonic oscillators. The magnon Hamilto-\nnian is the lowest-order term in the Holstein-Primako\u000b\nexpansion of the spin Hamiltonian, which can subse-\nquently be diagonalized by a Bogoliubov transformation\n[1{3].\nHere we chose to start from the Landau-Lifshitz\nequation _M =\u0000\r(M\u0002Be\u000b), in which Be\u000b=\n\u0000@E(M)=@MandE(M) is the classical magnetic free\nenergy introduced above. Writing Mq=M0+mqand\nto leading order in the small transverse excitation ampli-\ntudes mq\u0001M0= 0, the spin wave frequencies !qfor wave\nvector qare the solutions of\n(i!q\u0000Be\u000b\u0002)Mq=0: (8)\nFor simplicity, we compute !qwithout dipolar correc-\ntions that cause well-known anisotropic corrections for\nq6= 0, but do not a\u000bect the band edges in Figure 3. The\nresults agree with those obtained from the lowest-order\nHolstein-Primakov expansion of the corresponding spin\nHamiltonians.\nWhen the magnetic \feld is normal to the easy plane\n(A), the spin wave dispersion reads\n!(A)\nq=\r(q\nDq2\u0000\n2jKj\u0000B2\n4K2+Dq2\u0001\nB\u00002jKj+Dq2forB < 2jKj\nB > 2jKj\n(9)\nThe spin wave frequency of the uniform precession\n(q= 0) vanishes for magnetic \felds below Bcbecause thetorques generated by the anisotropy and applied \felds\ncancel. The energies for all magnetization directions that\nlie on a cone with opening angle \u0012are the same, so the\nground state is highly degenerate.\nThe Kittel problem (B) leads to\n!(B)\nq=\rp\n(B+Dq2+ 2K) (B+Dq2): (10)\nThe anisotropy qualitatively changes the dispersion at\nsmall magnetic \felds from !(B)\n0=\rBforK= 0 to\n!(B)\n0\u0018\rp\nBwhenK > B . The anisotropy breaks the\naxial symmetry and mixes the anti-clockwise circular pre-\ncession mode with positive frequencies and the forbidden\nclockwise one with negative frequencies, which leads to\nthe square root dependence rather than 2 \rK, the linearly\nextrapolated value from the high-\feld region.\nIn con\fguration (C) (as in (A)), magnetization and\n\feld are not collinear for \felds B 2jKj: (11)\nWe observe that the B-dependence of the collinear con-\n\fguration equals that of the Kittel mode shifted by 2 jKj\nto higher magnetic \felds.\nIV. MAGNON SPIN AND\nHELLMANN-FEYNMAN THEOREM\nAn excited state jqiwith energy \"q=~!qof an arbi-\ntrary spin Hamiltonian carries a spin magnetic moment\n\u0016q=hqj^Mjqi\u0000M0;where ^M=\u0000\r~P\ni^Si:Consider\na spin system with zero-\feld Hamiltonian HSthat in-\nteracts with constant applied magnetic \feld Bby the\nZeeman interaction\nH=HS\u0000^M\u0001B: (12)\nThe Hellmann-Feynman theorem then states that for\nsimplicity in the absence of textures,\n\u0016q=\u0000@\"q\n@BkM0\nM(13)\nwhereBk=B\u0001M0=M. For our classical spin system,\nwe replace\"qwith \u0001Eq=E(M0+\u0016q)\u0000E(M0). When\nthe magnetization and the applied \feld are parallel, we\ncan simply read o\u000b the spin of the excited state from\nthe magnon spectrum as a function of the applied \feld.\nWhen M,B, the situation is more complicated and the4\nFIG. 3. Magnon spin at the band edges for the con\fgurations\n(A)-(C) in Fig. 1. In (C) the magnon spin is tilted for B <\n2jKjwith a component in the y- (full curve) and z-directions\n(dashed curve).\nHellmann-Feynman theorem requires additional calcula-\ntions. Con\fgurations (A) and (C) at \felds B < 2K;for\nexample, acquire an \\orbital correction\" [10]\n\u0016q=\u0000d\"q\ndB+@\"q\n@\u0012@\u0012\n@B; (14)\nwhere\u0012is the equilibrium tilt angle. The implicit de-\npendence on the \feld enters here with an opposite sign\ncompared to Ref. [10].\nTo leading order, the solutions of the LL equation are\ntransverse, i.e. mq\u0001M0=O(m2\nq). If we transform back\ninto the time domain, the transverse components oscil-\nlate with the spin wave frequency and average out to\nzero. A magnon moment along the magnetization direc-\ntion persists in the time-average as\n\u0016q\u0019\u00001\n2\u0012jmqj\nM\u00132M0\nM: (15)\nThe amplitudes of the solutions of the LL equation are\ncontinuous but can be quantized by requiring that the\nminimum excitation energy of a mode with energy \"qis\ndiscrete:\n\u0001Eq=E(Mq)\u0000E0=~!q: (16)\nBy normalizing the mode amplitudes in this way we can\ncompute the spin of a single magnon for our three con-\n\fgurations (A-C).\nIn case (A), the moment of a single magnon \u0016(A)\nq=\n\u00002\u0016BforB > B cfrom the Hellmann-Feynman. How-\never, spin and magnetization are not collinear for B <\nBc. The LL equation then leads to a magnon magneticmoment along the equilibrium magnetization direction of\n\u0016(A)\nq\n2\u0016B=\u0000jKj\u0000B2\n4jKj+Dq2\nq\nDq2(2jKj\u0000B2\n2jKj+Dq2): (17)\nthat diverges when q!0 because the degeneracy of all\nmagnetization directions on the cone with angle \u0012allows\ncoherent precession of the magnetization without an en-\nergy cost.\nIn case (B), we can compute the magnon spin simply as\na derivative of the frequency with respect to the applied\nmagnetic \feld from the Hellmann-Feynman theorem Eq.\n(13) since BkM:\n\u0016(B)\nq\n2\u0016B=\u0000B+K+Dq2\np\n(B+Dq2+ 2K) (B+Dq2): (18)\nIn the zero \feld limit \u0016(B)\n0=\u00002\u0016B=p\nK=(2B) diverges.\nThe anisotropy mixes right and left precession modes to\ncreate an elliptic motion that in the limit of zero mag-\nnetic \feld leads to a linear x-polarized motion in the\neasy plane. As in case (A) the restoring torque vanishes\nin this limit, which allows the magnon amplitude to be-\ncome large.\nAlso in the high \feld regime B >B cof the case (C), we\ncan use the Hellmann-Feynman theorem without having\nto worry about \\orbital\" terms\n\u0016(C)\nq\n2\u0016B=\u0000\u0000\nB+Dq2\u0001\n+\u0000\nB+Dq2\u00002K\u0001\np\n(B+Dq2\u00002K) (B+Dq2)forB >B c;\n(19)\nwhich is a shifted version of \u0016(B)\nq.\u0016(C)\n0diverges for\nB#Bcbecause the torque in the out-of-plane z-direction\nvanishes. The magnon spin for the canted con\fguration\n(B 0 causes a logarithmic divergence at the\ncritical \feld B#Bcthat can be regulated by a low-\nmomentum cut-o\u000b Q0that can be rationalized by a resid-\nual in-plane anisotropy or disorder. To leading orders for\nlargeQ1and smallQ2\n0:\n\u001b(3d)\nm!e2\n~22\n3\u00192\u001c(3d)\nrkBT \nQ1\u0000p\n2\n16r\nK\nDlnDQ2\n0\n8K!\n(26)\nThe last term is the conductivity enhancement by the\nsoft mode magnon.\nIn two dimensions and K= 0\n\u001b(2d)\nm=e2\n~21\n\u0019\u001c(2d)\nrkBTe2\u001c(2d)\nrkBT\n\u00191\n2\u0014\nln\u0012\n1 +DQ2\n1\nB\u0013\n+B\nB+DQ21\u0015\n(27)\nthe divergence is logarithmic, but with vanishing magnon\ngapB!0 there is now also a logarithmic infrared di-\nvergence for a constant magnon spin, which is caused6\nby the enhanced magnonic density of states at the band\nedge. Including the divergent magnon spin when K > 0,\nto leading order in a small Q0and largeQ1, and for\nB#Bc\n\u001b(2d)\nm!e2\u001c(2d)\nrkBT\n\u0019\"\nln r\n2D\nKQ1!\n+r\nK\n2D1\n4Q0#\n(28)\nThe high-momentum divergence is still logarithmic, but\nthe low-momentum one becomes algebraic. This implies\na dimensionally enhanced magnon transport around the\ncritical magnetic \feld in van der Waals ferromagnets with\nperpendicular magnetization.\nVI. DISCUSSION\nSeveral processes regulate the enhanced \ructuations\nthat cause the divergence of the magnon spin reported\nhere, but should not destroy the predicted enhancement\nof spin pumping and spin transport. The situation is not\nunlike that of ferrimagnets with singular points in the\nphase diagram caused by angular or magnetic momen-\ntum compensation that causes interesting enhancements\nof e.g. domain wall velocities, in their vicinity [20].\nThe magnon spin of the Kittel mode can be observed\nby spin pumping under ferromagnetic resonance. Ando\net al. [6] indeed reported that the ellipticity of the mag-\nnetization precession under ferromagnetic resonance con-\nditions increases the spin pumping, a \fnding that we in-\nterpret as evidence for an enhanced magnon spin. How-\never, the resonance frequency also enters the pumped\nspin current, and the product spin \u0002frequency remains\nwell-behaved for soft magnons. Similarly, the spin See-\nbeck e\u000bect is hardly a\u000bected.\nMagnon transport using heavy metal contacts for spin\ninjection and detection can be carried out only on mag-\nnets that are also good electrical insulators. The magnon\nconductivity can be measured reliably when the injector\nand detector separation does not exceed the magnon dif-\nfusion length, which requires high magnetic quality. YIG\n\flms can be tuned to a perpendicular magnetization by\ndoping and epitaxial strain [21]. The hexagonal barium\nferrite (BaFe 12O19) has a perpendicular anisotropy \feld\nof 1.7 teslas and damping constant \u000b= 10\u00003[22]. Other\noptions are electrically insulating van der Waals ferro-\nmagnets with perpendicular magnetization such as CrI 3\n[23]. With minor adaptions, the present formalism can be\nused as well for magnetic \flms (strips) with an in-plane\neasy crystal (demagnetization) anisotropy axis and an\nin-plane magnetic \feld applied at right angles to it.\nWe estimate the magnitude of the expected e\u000bects\nby adopting the spin wave sti\u000bness of YIG D= 5\u0001\n10\u000017Tm2;a perpendicular anisotropy K= 1 T;a high\nmomentum cut-o\u000b frequency of 1 THz and an in-plane\nanisotropy of 1 mT :This leads to an enhancement of themagnon conductivity at the soft magnon point of \u00182 for\nthree-dimensional and \u00185 for two-dimensional magnets.\nThe divergences reported here occur only in materi-\nals with weak residual anisotropies, i.e. a su\u000eciently\nsmall cut-o\u000b Q0. Moreover, non-parabolicities render\nthe magnon approximation invalid when the spin wave\namplitudes become large by large magnon spins and/or\nnumbers. The implied magnon interactions may act as\na brake on the dynamics. The predicted numbers at the\ncritical points should therefore be taken with a grain of\nsalt, but the enhancement of the magnon thermal con-\nductivity close to the softening of the magnon modes\nshould persist even when these factors are taken into ac-\ncount.\nVII. CONCLUSIONS\nWe considered the spin waves of ferromagnetic\nmagnons as a function of an applied magnetic \feld, fo-\ncusing on the singular points at which the band edges\n(nearly) vanish. The more general conclusions such as\nthe relation between the Hellmann-Feynman theorem for\ncollinear systems also hold for antiferromagnets. We pre-\ndict enhanced non-local magnon transport caused by the\ndivergence of the magnon spin for magnets with a mag-\nnetic \feld applied perpendicular to a uniaxial magnetic\nanisotropy, which is stronger in two than in three dimen-\nsions. The e\u000bect should contribute to the observation\nof magnon transport above and down to the \\spin-\rip\"\ntransition at which the spin sublattices are forced to align\nferromagnetically [17] and we expect enhanced non-local\nsignals carried by the soft acoustic magnon as in case (C).\nHowever, more work is necessary to fully understand the\nexperiments [18].\nThe present calculations of the transport properties ad-\ndress the linear response at elevated temperatures, disre-\ngarding the e\u000bect of non-linear terms that are likely to be-\ncome important. Higher harmonic generation is easier for\n\roppy magnetic order but is at vanishing applied \felds\ncomplicated by spontaneous magnetic textures [24, 25],\nwhich may not interfere when the soft magnon is shifted\nto su\u000eciently large magnetic \felds. It should also be in-\nteresting to study the non-linearities of propagating soft\nmagnons by microwave spectroscopy [26].\nACKNOWLEDGMENTS\nP.T. and G.B. acknowledge the \fnancial support\nby JSPS KAKENHI Grants No. 19H00645 and No.\n22H04965. BW was supported by the European Union's\nHorizon 2020 program under Grant Agreement No.\n785219 and 881603, the NWO Spinoza prize, and ERC\nAdvanced Grant 2DMAGSPIN (Grant agreement No.\n101053054). Akash Kamra kindly commented on the\nmanuscript.7\n[1] A.G. Gurevich and G.A. Melkov, Magnetization Oscilla-\ntions and Waves, (CRC Press, London, 1996).\n[2] D.D. Stancil and A. Prabhakar, Spin Waves, (Springer,\nNew York, 2009).\n[3] S.M. Rezende, Fundamentals of Magnonics (Springer,\nCham, 2020).\n[4] A. Barman et al ., The 2021 Magnonics Roadmap, J.\nPhys.: Condens. Matter 33, 413001 (2021).\n[5] A. V. Chumak et al., Advances in Magnetics: Roadmap\non Spin-Wave Computing, IEEE Trans. Mag. 58, 1\n(2022).\n[6] K. Ando, S. Takahashi, J. Ieda, Y. Kajiwara, H.\nNakayama, T. Yoshino, K. Harii, Y. Fujikawa, M. Mat-\nsuo, S. Maekawa, and E. Saitoh, Inverse spin-Hall e\u000bect\ninduced by spin pumping in metallic system, J. Appl.\nPhys. 109, 103913 (2011).\n[7] B. Flebus, K. Shen, T. Kikkawa, K. Uchida, Z. Qiu,\nE. Saitoh, R.A. Duine, and G. E. W. Bauer, Magnon-\npolaron transport in magnetic insulators, Phys. Rev. B\n95, 144420 (2017).\n[8] A. Kamra and W. Belzig, Super-Poissonian Shot Noise of\nSqueezed-Magnon Mediated Spin Transport, Phys. Rev.\nLett. 116, 146601 (2016).\n[9] A. Kamra, U. Agrawal, and W. Belzig, Noninteger-spin\nmagnonic excitations in untextured magnets, Phys. Rev.\nB96, 020411(R) (2017).\n[10] R. R. Neumann, A. Mook, J. Henk, and I. Mertig, Orbital\nMagnetic Moment of Magnons, Phys. Rev. Lett. 125,\n117209 (2020).\n[11] H.Y. Yuan, Y. Cao, A. Kamra, R.A. Duine, and P. Yan,\nQuantum magnonics: When magnon spintronics meets\nquantum information science, Phys. Rep. 965, 1 (2022).\n[12] L. J. Cornelissen, J. Liu, R. A. Duine, J. Ben Youssef, and\nB. J. van Wees, Long-distance transport of magnon spin\ninformation in a magnetic insulator at room temperature,\nNature Phys. 11, 1022 (2015).\n[13] X.-Y.Wei, O. A. Santos, C. H. S. Lusero, G. E. W. Bauer,\nJ. Ben Youssef, and B. J. van Wees, Giant magnon spin\nconductivity in ultrathin yttrium iron garnet \flms Nat.\nMat. 21, 1352 (2022).\n[14] R. Lebrun, A. Ross, S.A. Bender, A. Qaiumzadeh, L.\nBaldrati, J. Cramer, A. Brataas, R.A. Duine, M. Kl aui,\nLong-Distance Spin Transport in a Crystalline Antifer-\nromagnetic Iron Oxide, Nature 561, 222 (2018).\n[15] T. Wimmer, A. Kamra, J. G uckelhorn, M. Opel, S.\nGepr ags, R. Gross, H. Huebl, and M. Althammer, Obser-vation of Antiferromagnetic Magnon Pseudospin Dynam-\nics and the Hanle E\u000bect, Phys. Rev. Lett. 125, 247204\n(2020).\n[16] J. Han, P. Zhang, Z. Bi, et al ., Birefringence-like\nspin transport via linearly polarized antiferromagnetic\nmagnons. Nat. Nanotech. 15, 563 (2020)..\n[17] D.K. de Wal, A. Iwens, T. Liu, P. Tang, G.E.W.\nBauer, and B.J. van Wees, Long distance magnon\ntransport in the van der Waals antiferromagnet CrPS 4,\narXiv:2301.03268\n[18] P. Tang, unpublished.\n[19] Y. Tserkovnyak, A. Brataas, G.E.W. Bauer, B.I.\nHalperin, Nonlocal magnetization dynamics in ferromag-\nnetic heterostructures, Rev. Mod. Phys. 77, 1375 (2005).\n[20] L. Caretta, M. Mann, F. B uttner, K. Ueda, B. Pfau,\nC. M. G unther, P. Hessing, A. Churikova, C. Klose, M.\nSchneider, D. Engel, C. Marcus, D. Bono, K. Bagschik, S.\nEisebitt, and G. S. D. Beach, Fast current-driven domain\nwalls and small skyrmions in a compensated ferrimagnet,\nNat. Nanotechnol. 13, 1154 (2018).\n[21] L. Soumah, N. Beaulieu, L. Qassym, C. Carr\u0013 et\u0013 ero, E.\nJacquet, R. Lebourgeois, J.B. Youssef, P. Bortolotti,\nV. Cros, and A. Anane, Ultra-low damping insulating\nmagnetic thin \flms get perpendicular, Nat. Comm. 9,\n3355(2018).\n[22] L. Alahmed and P. Li, Chapter 3 in: Magnetic Materials\nand Magnetic Levitation, edited by D.R. Sahu and V.\nN. Stavrou (IntechOpen Limited, London, 2021), DOI:\n10.5772/intechopen.92277 .\n[23] B. Huang, G. Clark, E. Navarro-Moratalla, et al. , Layer-\ndependent ferromagnetism in a van der Waals crystal\ndown to the monolayer limit, Nature 546, 270 (2017).\n[24] C. Koerner, R. Dreyer, M. Wagener, N. Liebing, H. G.\nBauer, and G. Woltersdorf, Frequency multiplication by\ncollective nanoscale spin-wave dynamics, Science 375,\n1165 (2022).\n[25] G. Lan, K. Liu, Z. Wang, F. Xia, H. Xu, T. Guo, Y.\nZhang, B. He, J. Li, C. Wan, G.E.W. Bauer, P. Yan,\nG. Liu, X. Pan, X. Han, and G. Yu, Coherent frequency\nmultiplication by interacting magnons in spin textures\n(unpublished).\n[26] L. Sheng, M. Elyasi, J. Chen, W. He, Y. Wang, H. Wang,\nH. Feng, Y. Zhang, I. Medlej, S. Liu, W. Jiang, X. Han,\nD. Yu, J.-P. Ansermet, G.E.W. Bauer, and H. Yu, Non-\nlocal Detection of Interlayer Three-Magnon Coupling,\nPhys. Rev. Lett. 130, 046701 (2023)." }, { "title": "1009.0643v1.Josephson_current_and__π__state_in_ferromagnet_with_embedded_superconducting_nanoparticles.pdf", "content": "arXiv:1009.0643v1 [cond-mat.supr-con] 3 Sep 2010Josephson current and π−state in ferromagnet\nwith embedded superconducting nanoparticles\nA. V. Samokhvalov(1), A. I. Buzdin(2)\n(1)Institute for Physics of Microstructures,\nRussian Academy of Sciences, 603950 Nizhny Novgorod, GSP-1 05, Russia\n(2)Institut Universitaire de France and Universit´ e Bordeaux I,\nCPMOH, UMR 5798, 33405 Talence, France\nAbstract\nOn the basis of Usadel equations we investigate superconduc tor/ferromagnet/superconductor\n(S/F/S) hybrid systems which consist of superconductingna nostructures (spheres, rods)embedded\nin ferromagnetic metal. The oscillations of the critical cu rrent of the S/F/S Josephson junctions\nwith the thickness of ferromagnetic spacer between superco nducting electrodes are studied. We\ndemonstrate that the πstate can be realized in such structures despite of a dispers ion of the\ndistances between different parts of the electrodes. The tran sitions between 0 and πstates at some\nthickness of ferromagnetic spacer can be triggered by tempe rature variation.\nPACS numbers:\n1I. INTRODUCTION\nThe particularity of the proximity effect in superconductor/ferro magnet (S/F) hybrid\nstructures is the damped oscillatory behavior of the Cooper pair wa ve function inside the\nferromagnet1,2(for the reviews see3,4). In some sort it is a manifestation of the Larkin-\nOvchinnikov-Fulde-Ferrell (LOFF) state induced in ferromagnet ( F) near the interface with\nsuperconductor (S). In contrast with original LOFF, which is poss ible only in the clean\nsuperconductors, the damped oscillatory S/F proximity effect is ve ry robust and exists also\nin the diffusive limit. This special type of the proximity effect is at the or igin of the π\nJosephson S/F/S junction1,2which has at the ground state the opposite sign of the super-\nconducting orderparameter inthebanks. Firstlythe π-junctionwasobserved atexperiment\nin Ref.5and since then a lot of progress has been obtained in the physics of π- junctions and\nnow they are proven to be promising elements of superconducting c lassical and quantum\ncircuits6. Different manifestations of unusual proximity effect and the πstates have been\nobserved experimentally in various layered S/F hybrids7–9. The proximity induced switching\nbetween the superconducting states with different vorticities in mu ltiply connected hybrid\nS/F structures was suggested recently in10,11. Theoretical studies and experiments12–15both\ndemonstrated that in the diffusive limit the spin-flip and spin-orbit sca ttering lead to the\ndecrease of the decay length and the increase of the oscillating per iod. In addition the spin-\nflip scattering may generate the temperature induced transition f rom 0 to πstate of the\njunction14,15.\nNaturally at the first stage of the work on the S/F/S junctions the systems with planar\ngeometry and well controlled F-layer thickness were considered. H owever, now when the π\nstate proven to be very robust vs different types of the impurities scattering13,15(magnetic\nand non-magnetic), interface transparency16,17it may be of interest to address a question\nhow the πjunction could realize in the S/F/S systems with a bad defined thickne ss of F-\nspacer, in particular for two superconducting particles imbedded in ferromagnet or between\na flat superconducting electrode and a small superconducting nan oparticle (such situation\ncouldbeof interest for theSTM-like experiments with asupercondu cting tip). This question\nis non-trivial because the transition from 0 to πstate occurs at the very small characteristic\nlength3ξf=/radicalbig\nDf/h, whereDfis the diffusion constant in ferromagnetic metal and his the\nferromagnetic exchange field, and the typical values of ξfare in the nanoscopic range. We\n2could expect that when the variation of the distance between differ ent parts of S-electrodes\nis of the order of ξftheπstate would disappear. Our calculations show that it is not the\ncase and once again the πstate occurs to be very robust and the transition between 0 and π\nstates is always present at some distance and also can be triggered by temperature variation.\nIn this paper we present the results of a theoretical study of the peculiarities of the\nproximity effect and Josephson current in S/F hybrids which consist of superconducting\nnanostructures placed in electrical contact with a ferromagnetic metal. The paper is orga-\nnized as follows. In Sec. II we briefly discuss the basic equations. In Sec. III we calculate\nthe Josephson current in two model hybrid S/F/S systems. The fir st system consists of two\nsuperconducting rod-shaped electrodes imbedded in ferromagne t. The second one is a S/F\nbilayer with a superconducting spherical particle at the surface of the ferromagnetic layer.\nWe examine the temperature dependence of the critical current o f S/F/S junction between\ntheflatsuperconductingelectrodeandtheS-particletakingintoa ccountthespin-flipscatter-\ning. For both cases the S/F interface transparency between sup erconducting nanoparticles\nand ferromagnet is assumed to be low to prevent from supercondu ctivity destruction due to\nproximity. We summarize our results in Sec. IV\nII. MODEL AND BASIC EQUATIONS\nSince the models of S/F/S junctions we are going to study consist of superconducting\nparticles embedded in a ferromagnetic matrix or placed on a ferroma gnetic substrate, we\nstart from a description of the damped oscillatory behavior of the C ooper wave function\ninduced by such particle in a ferromagnet.\nWe assume the elastic electron-scattering time τto be rather small, so that the critical\ntemperature Tcand exchange field hsatisfy the dirty-limit conditions Tcτ≪1 andhτ≪1.\nIn this case a most natural approach to calculate Tcis based on the Usadel equations18for\nthe averaged anomalous Green’s functions FfandFsfor the F- and S-regions, respectively\n(see3for details). These equations are nonlinear but can be simplified when the tempera-\nture is close to the critical temperature Tcor at any temperature of the F-layer when the\ntransparency of S/F interface is low. In the F-region the linearized Usadel equations take\nthe form\n−Df\n2∇2Ff+(|ω|+ıhsgnω+1/τs)Ff= 0, (1)\n3whereDfis the diffusion coefficient in ferromagnetic metal and ωare the Matsubara fre-\nquencies, ω= 2πT(n+ 1/2), andτsis the magnetic scattering time. We consider the\nferromagnet with strong uniaxial anisotropy, in which case the mag netic scattering does not\ncouple the spin up and spin down electron populations. Restricting ou rselves to the case of\nsuperconducting inclusions with cylindrical or spherical symmetry ( cylindrical rod-shaped\nor spherical particles of radius Rs), one can easily find the following solutions of Eq. (1),\nwhich describe the distribution of anomalous Green’s function Ffin ferromagnet ( r≥Rs)\nsurrounding a superconducting cylinder\nFc\nf(r) =AK0(qr), (2)\nor superconducting sphere\nFs\nf(r) =Aexp(−qr)/qr, (3)\nwhereK0(z) is the Macdonald function,\nq=/radicalBig\n2/Df/radicalbig\n|ω|+ıhsgnω+1/τs (4)\nis the characteristic wave number of the order parameter variatio n in the F-metal, and the\namplitude Ais determined by the boundary conditions at the S/F interface ( r=Rs)19:\nσs∂rFs=σn∂rFf, Fs=Ff−γbξn∂rFf. (5)\nThe S/F interface between a particle and ferromagnet is assumed t o be characterized by\nthe dimensionless parameter γb=Rbσn/ξnrelated to the boundary resistance per unit area\nRb. Hereξs(n)=/radicalbig\nDs(f)/2πTcis the superconducting (normal-metal) coherence length,\nσsandσnare the normal-state conductivities of the S- and F-metals, and ∂rdenotes a\nderivative taken in the radial direction. We will assume that the rigid b oundary condition\nγb≫min{ξsσn/ξnσs,1}is satisfied, when the inverse proximity effect and the suppression\nof superconductivity is S-metal can be neglected20,21. As a result, the pair amplitude Fs(r)\nat the S/F interface is equal to the one far from the boundary:\nFs(Rs) =∆√\nω2+∆2=∆\nωGn, (6)\nwhere ∆ is the superconducting order parameter, and\nGn=ω√\nω2+∆2(7)\n4is normal Green’s function. Using the solution (2) or (3) we obtain fr om the Eqs. (5) and (6)\nthe value of Ffat the S/F boundary r=Rs, and the amplitudes Afor both cases. Finally,\nthe expressions\nFc\nf(r) =∆\nωGnK0(qr)\nK0(qRs)+γbξnqK1(qRs), (8)\nFs\nf(r) =∆\nωGnRse−q(r−Rs)\nr(1+γbξn(q+1/Rs)). (9)\ndescribe the the damped oscillatory behavior of anomalous Green’s f unctionFfin ferromag-\nnet surrounding the superconducting cylinder or sphere, respec tively.\nThe general expression for the supercurrent density is given by\n/vectorJs=iπTσn\n4e/summationdisplay\nω,σ=±/parenleftBig\n˜Ff∇Ff−Ff∇˜Ff/parenrightBig\n, (10)\nwhere˜Ff(r,ω) =F∗\nf(r,−ω).\nIII. CRITICALCURRENT OFJUNCTIONS WITHSUPERCONDUCTING PAR-\nTICLES\nNow we proceed with calculations of the Josephson critical current for two examples of\nmesoscopic hybrid S/F systems. The first one is two identical super conducting cylindrical\nrod-shaped electrodes surrounded by a ferromagnetic metal (s ee Fig. 1). The second one is\na S/F bilayer with a superconducting particle at the surface of the f erromagnetic layer (see\nFig. 3).\nA. S/F/S junction between two superconducting rod\nConsider two superconducting cylinders of a radius Rsembedded in ferromagnet as it\nis shown in Fig. 1. The distance between the cylinder axes is d >2Rs. These rod-shaped\nelectrodes form a Josephson junction in which the weak link between two superconductors\nis ensured by ferromagnetic neighborhood. The supercurrent\nIs(ϕ) =Icsin(ϕ) (11)\nflowing acrossthis structure depends onthephase difference ϕbetween theorder parameters\nof the rods:\n∆1,2= ∆e±iϕ/2. (12)\n5Rs-d/2 d/2 ay\nx\n/c68ei /2/c106/c68e-i /2/c106\nF\nFIG. 1: (Color online) Schematic representation of the F/S h ybrid system under consideration:\ntwo identical superconducting cylindrical rod-shaped ele ctrodes of radius Rssurrounded by a fer-\nromagnetic metal. The axes of superconducting cylinders ar e assumed to be parallel. Figure shows\nthe cross section of the structure by the plane ( x,y) perpendicular to the cylinder axis.\nFor large enough distance between the superconducting cylinders (a=d−2Rs>2ξf), the\ndecay of the Cooper pair wave function in ferromagnet in the first a pproximation occurs\nindependently near either of the electrodes and can be described b y the solution (8). There-\nfore the anomalous Green function Ff(r) in ferromagnet nearby the plane x= 0 may be\ntaken as the superposition of the two decaying functions (8), tak ing into account the phase\ndifference ϕ22:\nFf(x,y) =∆\nωGnK0(qr+)eiϕ/2+K0(qr−)e−iϕ/2\nK0(qRs)+γbξnqK1(qRs), (13)\nwherer±=/radicalbig\n(x±d/2)2+y2. Using the expression (10), we obtain the sinusoidal current-\nphase relation (11) in the S/F/S Josephson junction between two s uperconducting rod-\nshaped electrodes for the case of low transparent S/F interface s. For the critical current of\nsuch Josephson structure, we have\nIc=2πTσn\ne/summationdisplay\nω>0∆2\nω2G2\nn\n×Re\n\naq\n[K0(qRs)+γbξnqK1(qRs)]2∞/integraldisplay\n−∞dyK0(qr0)K1(qr0)\nr0\n\n, (14)\nwherer0=/radicalbig\ny2+d2/4. In the limit of large Rs≫ξfa curvature of the electrodes is not\nessential and the formula (14) coincide with the corresponding exp ressions for the critical\ncurrent of S/F/S layered structures with a large interface trans parency parameter γbprevi-\n6/s48 /s50 /s52 /s54 /s56/s49/s48/s45/s51/s49/s48/s45/s49/s49/s48/s49\n/s48/s46/s53 /s49/s46/s48/s49/s48/s45/s50/s49/s48/s48\n/s49/s48\n/s49/s73\n/s99/s32/s47/s32/s73\n/s48\n/s97/s32/s47/s32\n/s102/s49/s48/s48\n/s97\n/s48\n/s32/s73\n/s99/s32/s47/s32/s73\n/s48\n/s100/s32\n/s102/s32/s49/s48/s49/s48/s48\n/s49\nFIG. 2: (Color online) Influence of the electrode radius Rson the dependenceof the critical current\nIc(15) on the distance abetween two superconducting rod-shaped electrodes embedd ed in F-metal\n(1/τs= 0). The numbers near the curves denote the corresponding va lues of the radius Rsin the\nunits ofξf.. The inset gives the zoomed part of the Ic(a) line , marked by the dashed box.\nously obtained in Refs.15. The critical current equation (14) can be simplified for h≫πTc\nandRs,r0≫ξfand may be written as\nIc=I0dRs\nξf∞/integraldisplay\n−∞dye−2(√\ny2+d2/4−Rs)/ξf\ny2+d2/4cos/parenleftBigg\n2/radicalbig\ny2+d2/4−Rs\nξf+π\n4/parenrightBigg\n, (15)\nI0=πσn∆ξ2\nf\n2√\n2eγ2\nbξ2ntanh/parenleftbigg∆\n2T/parenrightbigg\n. (16)\nNote that our approach is valid for large enough distance between t he superconducting\ncylinders a >2ξfand the first thansition into the πstate ata0is described only qualita-\ntively. It has been demonstrated23that for the planar S/F/S junction with low interface\ntransparency the first transition into the πstate occurs at F layer thickness smaller than\nξf( its actual value depends on the the exchange field hand transparency parameter γb\n).The similar situation is expected for the embedded superconductin g particles and to find\nthe corresponding interparticle distance we need to solve our prob lem exactly.\nThe dependence of the critical current Icas a function of the distance abetween the\n7superconducting cylindrical electrodes calculated from Eq. (15) is presented in Fig. 2 for\nseveral values of the radius Rs. From the figure, we see that with increasing the distance\na, the S/F/S junction undergoes the sequence of 0- πandπ-0 transitions when the value of\nIcchanges its sign from positive to negative and vice versa. We may rou ghly estimate that\nthe first transition from 0 to πstate in S/F/S junction formed by rod-shaped electrodes\noccurs at the thickness of F-layer a0∼0.5ξf, similar to a S/F/S junction with a low S/F\ninterface transparency in the ordinary layered geometry16,23. We observe that the distances\nacorresponding to 0- πandπ-0 transitions grow slightly with decrease of the cylinders radius\nRsdue to a dispersion of the distances between different parts of the electrodes (see the inset\nin Fig. 2).\nB. S/F/S junction in S/F bilayer with a superconducting particle\nAsa second example we consider ferromagnetic film of a thickness don asuperconducting\nplate with a transparent S/F interface. The S/F/S Josephson jun ction is assumed to be\nformed between the flat superconducting electrode and a small su perconducting half-sphere\nof radius Rsembedded into ferromagnet, as it is shown in Fig. 3. The center of th e sphere\nis placed at the surface of the F-film. As before, there is the tunne l barrier ( γb≫1) at the\nS/F interface between the superconducting particle and ferroma gnetic metal.\nSince we consider a S/F bilayer with a transparent interface then th e complete nonlinear\nUsadel equation in the F-layer has to be employed. Using the usual p arametrization of the\nnormal and anomalous Greens functions Gf= cosΘ fandFf= sinΘ f, the Usadel equation\nis written as\n−Df\n2∇2Θf+/parenleftbigg\n|ω|+ihsgnω+cosΘf\nτs/parenrightbigg\nsinΘf= 0. (17)\nNote that Eq. (17) transforms into the linear equation (1) in the limit of small Θ f≪1.\nFor simplicity we restrict ourselves to the case of thick F-layer ( d≫ξf) then the decay\nof superconducting order parameter occurs independently near each S/F interface. In that\ncase, the behavior of the anomalous Green’s function near each int erface can be treated\nseparately, assuming that the F-layer thickness is infinite. Following Ref.1415, the analytical\nsolution of the equation (17) for flat transparent interface at z=dcan be written as\n/radicalbig\n1−ε2sin2(Θf/2)−cos(Θf/2)/radicalbig\n1−ε2sin2(Θf/2)+cos(Θ f/2)=f0e2q(z−d), (18)\n8/c71/c114\nz\nzc/c68e-i /2/c106\n/c68ei /2/c106rn\nRs\nFd\nFIG. 3: (Color online) Schematic representation of the S/F/ S Josephson junction between the flat\nsuperconducting electrode and a small superconducting sph erical particle of radius Rsembedded\ninto ferromagnet. Dashed line shows the cross-section of th e paraboloid Γ by the plane ( ρ, z) (24):\nρ2=x2+y2. Here/vector nis a unit vector along the normal to the surface Γ. Figure show s the cross\nsection by the plane.\nwhere\nε2= (1/τs)(|ω|+ıhsgnω+1/τs)−1.\nTheintegrationconstant f0shouldbedeterminedfromtheboundaryconditionatthesurface\nz=d. As before the rigid boundary conditions is assumed to be valid at z=d:\nΘf(d) = arctan∆\nω. (19)\nFrom Eqs. (18),(19) we get\nf0=(1−ε2)F2\nn/bracketleftBig/radicalbig\n(1−ε2)F2\nn+1+1/bracketrightBig2, (20)\nFn=|∆|\nω+/radicalbig\nω2+|∆|2. (21)\nLinearizing the solution (18) for Θ f≪1 we obtain the anomalous Green’s function in\nferromagnet (0 ≤z≤d) induced by flat superconductig electrode:\nΘf≃4Fn/radicalbig\n(1−ε2)F2\nn+1+1eq(z−d). (22)\n9The total anomalous Green’s function in F-layer far from both the S /F interfaces may be\ntaken as superposition of the two decaying functions (9), (22), t aking into account the phase\ndifference in each superconducting electrode\nFf=∆\nωGnRse−q(r−Rs)+iϕ/2\nr(1+γbξn(q+1/Rs))+4Fneq(z−d)−iϕ/2\n/radicalbig\n(1−ε2)F2n+1+1, (23)\nwherer=/radicalbig\nx2+y2+z2.\nTo derive the general expression for the critical current Icwe have to calculate the total\nJosephson current flowing through a virtual surface Γ: the point s of the surface Γ are\nequidistant from both electrodes of the junction. The surface Γ is a paraboloid which form\nis described by the equations:\nz=zc−(x2+y2)/4zc, zc= (Rs+d)/2. (24)\nUsing the solution (23) and Eq.(10), one can arrive at a sinusoidal cu rrent-phase relation\n(11) with the critical current (see A for details):\nIc=I0zcRsT\nξfTcRe∞/summationdisplay\nn=0/braceleftBigg\nFn(∆Gn/ω)e−q(d−Rs)\n[/radicalbig\n(1−ε2)F2n+1+1][1+ γbξn(q+1/Rs)]\n×\nqzc/integraldisplay\n0due−2qu\nu+zc+1\n2zc/integraldisplay\n0due−2qu\n(u+zc)2\n\n\n, (25)\nwhereI0= 64π2Tcσnξf/e.\nThe dependence of the critical current Ic(25) as a function of the thickness a=d−Rsof\nthe ferromagnetic spacer separating the superconducting plate and the particle is presented\nin Fig. 4 for several values of the particle radius Rsand the magnetic scattering time τs. It\nis clearly seen from Fig. 4 that with a decrease of the particle radius Rs, the position a0of\nthe first zero of the critical current is shifted towards larger valu es of the distance abetween\nsuperconducting electrodes. Figure 4b demonstrates the influen ce of magnetic scattering on\nthe proximity effect and the critical current in S/F bilayer with the pa rticle: decrease of\nthe magnetic scattering time τsleads to the decrease of decay length and increase of the\noscillation period of the anomalous Green’s function Ff14. This results in much stronger\ndecrease of the critical current in the S/F/S junction with increas e of the thickness a, if the\nmagnetic scattering time τsbecomes relatively small τ−1\ns≥h.\nFigure 5 shows the temperature dependence of the S/F/S junctio n critical current Ic(25)\nat several values of the thickness of the ferromagnetic spacer b etween the superconducting\n10/s48 /s50 /s52 /s54 /s56 /s49/s48/s49/s48/s45/s55/s49/s48/s45/s53/s49/s48/s45/s51/s49/s48/s45/s49\n/s48/s49/s49/s48/s73\n/s99/s32/s47/s32/s73\n/s48\n/s40/s32/s100/s32/s45/s32/s82\n/s115/s32/s41/s32/s47/s32\n/s102/s49/s48/s48/s40/s97/s41\n/s48\n/s48 /s50 /s52 /s54 /s56 /s49/s48/s49/s48/s45/s49/s49/s49/s48/s45/s56/s49/s48/s45/s53/s49/s48/s45/s50\n/s49/s49/s48/s73\n/s99/s32/s47/s32/s73\n/s48\n/s40/s32/s100/s32/s45/s32/s82\n/s115/s32/s41/s32/s47/s32\n/s102/s49/s48/s48/s40/s98/s41\n/s48\n/s97\n/s48\nFIG. 4: (Color online) Thedependenceof thecritical curren tIc(25) on thedistance d−Rsbetween\nthe superconducting plate and the particle of radius Rsembedded in F-metal for different values of\nthe radius Rsand the magnetic scattering time τs(T/Tc= 0.5,h= 3πTc,γb= 10): (a) hτs= 100;\n(b)hτs= 0.5. The numbers near the curves denote the corresponding valu es of the radius Rsin\nthe units of ξf.\nelectrodes. The F-spacer thickness ais chosen close to the first transition from 0 to πstate:\na∼a0≃3ξf. The nonmonotonic dependences Ic(T) demonstrate 0 - πtransition due to a\nchange of the temperature. The transition temperature T∗(Ic(T∗) = 0) seems to be very\nsensitive to the size of the superconducting particle. It should be n oted however that the\ntemperature T∗is determined rather by the scale a=d−Rsthen by the scales dorRs,\n11/s48/s46/s48 /s48/s46/s50 /s48/s46/s52 /s48/s46/s54 /s48/s46/s56 /s49/s46/s48/s48/s46/s53/s49/s46/s48\n/s48/s46/s57/s53/s49/s49/s46/s48/s53/s73\n/s99/s32/s47/s32/s73\n/s48/s32/s91/s49/s48/s45/s55\n/s93\n/s84/s32/s47/s32/s84\n/s67/s40/s97/s41\n/s48/s46/s48 /s48/s46/s50 /s48/s46/s52 /s48/s46/s54 /s48/s46/s56 /s49/s46/s48/s48/s46/s53/s49/s46/s48/s49/s46/s53/s50/s46/s57\n/s51/s46/s48\n/s51/s46/s49/s73\n/s99/s32/s47/s32/s73\n/s48/s32/s91/s49/s48/s45/s55\n/s93\n/s84/s32/s47/s32/s84\n/s67/s40/s98/s41\nFIG. 5: (Color online) (a) The dependence of the critical cur rentIc(25) on the temperature T\nfor different values of the radius Rs/ξf= 0.95,1.0,1.05 and the fixed F-spacer thickness a= 3ξf.\n(b) The dependence of the critical current Ic(25) on the temperature Tfor different values of the\nF-spacer thickness a/ξf= 2.9,3.0,3.1 and fixed radius Rs=ξf. The calculation parameters are\nh= 3πTc,hτs= 0.5,γb= 10.\nseparately.\n12IV. CONCLUSION\nTo sum up, we have analyzed the Josephson effect in S/F/S hybrid st ructures with a bad\ndefined thickness of F-spacer. As an example, we have calculated t he Josephson current\nbetween two rod-shaped superconducting electrodes embedded in ferromagnet or between\nflat superconducting electrode and the small superconducting na noparticle at the surface of\nthe F-layer. For the both cases we have demonstrated the possib ility of the realization of π\njunctions in such hybrid systems. We have studied dependence of t he transitions between\n0 andπstates both on the size of superconducting particles and the temp erature. The π\nstate has been proven to be very robust with respect to a geomet ry of the S/F/S junction.\nIn the dirty limit the transition into πstate is determined rather by the thickness of the F-\nspacer between superconducting electrodes then by a shape of t he electrodes. Naturally our\ncalculations can be easily generalized to the different shape of the S p articles (for example\nspherical) with similar conlusions.\nA set of the superconducting particles embedded in a ferromagnet ic matrix realize a\nJosephson network. Depending on the geometry of this network a nd its state (0 or π) it\nmay reveal a spontaneous current similar to that observed in supe rconducting arrays of π\njunctions24. For example the equilibrium phase difference for triangular 2D πjunctions\nnetwork is equal to 2 π/3 which corresponds to the current state. For typical paramete rs\nNb/CuNi hybrid system ( Tc= 9K,ξf≈2nm,ρn= 1/σn≈60µΩcm14one can get from\n(25) the following estimate of the Josephson energy EJ=φ0Ic/2πcof the S/F/S junction:\nEJ/Tc∼104(Ic/I0)<1, i.e. an observation of spontaneous currents near Tcis expected\nto be masked by strong temperature fluctuations. Despite of this restriction we believe\nthat intrinsically–frustrated superconducting networks induced by the proximity effect can\nbe experimentally observable in such S/F/S composites. In particula r, a two-dimensional\nJosephson network of πjunctions may serve as a laboratory to study the phase transition s\nwith continuous degeneracy25.\nThe possibility to fabricate the regular 2D and 3D arrays of Josephs onπjunctions and\nmonitor the transitions between 0 and πstates simply varying the temperature open inter-\nesting perspectives to study a very reach physics of different pha se transition in such systems\ndue to interplay between fluctuations, frustration, disorder and dimensionality.\n13V. ACKNOWLEDGMENTS\nWe are indebted to A. S. Mel’nikov for useful discussions. This work w as supported, in\npart, by the Russian Foundation for Basic Research, by Russian Ag ency of Education under\nthe Federal Program ”Scientific and educational personnel of inn ovative Russia in 2009-\n2013”, byInternational Exchange ProgramofUniversite Bordea uxI, byFrenchANR project\n”ELEC-EPR”, and by the program of LEA Physique Theorique et Mat iere Condensee.\nAppendix A: Josephson current in S/F bilayer with a superconducting particle\nThe general expression for the supercurrent is given by the Eq. ( 10), where the anomalous\nGreens function Ffnearby the surface Γ (see Fig. 3) may be written as\nFf=A1eqz−iϕ/2+A2e−q√\nρ2+z2+iϕ/2, (A1)\nA1=4Fne−qd\n/radicalbig\n(1−ε2)F2n+1+1, A 2=(∆Gn/ω)eqRs\n1+γbξn(q+1/Rs). (A2)\nwherer2=ρ2+z2, and the functions Gn,Fnare determined by expressions (7) and (21),\nrespectively. At the surface Γ the function Ffand the projection of the vector ∇Ffalong\nthe normal\n/vector n=ρ/radicalbig\nρ2+4z2c/vector ρ0+2zc/radicalbig\nρ2+4z2c/vector z0\nto the surface are\nFf|Γ=/parenleftbigg\nA1eqzc−iϕ/2+A24zcRs\nρ2+4z2ce−qzc+iϕ/2/parenrightbigg\ne−qρ2/4zc, (A3)\n(∇Ff,/vector n)|Γ=2zce−qρ2/4zc\n/radicalbig\nρ2+4z2c\n×/bracketleftbigg\nqA1eqzc−iϕ/2−A24zcRs\nρ2+4z2c/parenleftbigg\nq+4zc\nρ2+4z2c/parenrightbigg\ne−qzc+iϕ/2/bracketrightbigg\n. (A4)\nSubstitution of Eqs. (A3), (A4) into the expression for the super current (10) and taking into\naccount the symmetry relations q(−ω) =q∗(ω),A1,2(−ω) =A∗\n1,2(ω) leads to the following\nformula\n(/vectorJs,/vector n)/vextendsingle/vextendsingle/vextendsingle\nΓ=Jc(ρ) sinϕ, (A5)\nJc(ρ) =32πTσn\neRe/summationdisplay\nω>0/braceleftbiggz2\ncRsA1A2\n(ρ2+4z2\nc)3/2/parenleftbigg\nq+4zc\nρ2+4z2\nc/parenrightbigg\ne−qρ2/4zc/bracerightbigg\n. (A6)\n14Further integration of supercurrent density (A6) over the surf ace Γ\nIc=/integraldisplay\nΓdSJc=π\nzc2zc/integraldisplay\n0dρρ/radicalbig\nρ2+4z2\ncJc(ρ). (A7)\nresults in the expressions (25) for the critical current Icof the S/F/S Josephson junction\nbetween superconducting plate and particle.\nReferences\n1A. I. Buzdin, L. N. Bulaevskii, and S. V. Panyukov, Pis’ma Zh. Eksp. Teor. Fiz. 35, 147 (1982)\n[JETP Lett. 35, 178 (1982)].\n2A. I. Buzdin and M. Y. Kuprianov, Pis’ma Zh. Eksp. Teor. Fiz. 5 3, 308 (1991) [JETP Lett. 53,\n321 (1991)].\n3A. I. Buzdin, Rev. Mod. Phys., 77, 935 (2005).\n4A. A. Golubov, M. Yu. Kupriyanov, and E. Il’ichev, Rev. Mod. P hys. 76, 411 (2004).\n5V. V. Ryazanov, V. A. Oboznov, A. Yu. Rusanov, A. V. Veretenni kov, A. A. Golubov, and J.\nAarts, Phys. Rev. Lett. 86, 2427 (2001)\n6A. K. Feofanov, V. A. Oboznov, V. V. Bol’ginov, J. Lisenfeld1 , S. Poletto, V. V. Ryazanov, A.\nN. Rossolenko, M. Khabipov, D. Balashov, A. B. Zorin, P. N. Dm itriev, V. P. Koshelets and A.\nV. Ustinov, Nature Physics, 6, 593 (2010).\n7J. S. Jiang, D. Davidovic, D. H. Reich, and C. L. Chien, Phys. R ev. Lett. 74, 314 (1995).\n8A. S. Sidorenko, V. I. Zdravkov, A. A. Prepelitsa, C. Helbig, Y. Luo, S. Gsell, M. Schreck, S.\nKlimm, S. Horn, L. R. Tagirov, and R. Tidecks, Ann. Phys. 12, 37 (2003).\n9I. A. Garifullin, D. A. Tikhonov, N. N. Garifyanov, L. Lazar, Yu. V. Goryunov, S. Ya. Khleb-\nnikov, L. R. Tagirov, K. Westerholt, and H. Zabel, Phys. Rev. B66, R020505 (2002).\n10A. V. Samokhvalov, A. S. Melnikov, and A. I. Buzdin, Phys. Rev . B76, 184519 (2007).\n11A. V. Samokhvalov, A. S. Melnikov, J-P. Ader, and A. I. Buzdin , Phys. Rev. B 79, 174502\n(2009).\n12L. Cretinon, A. K. Gupta, H. Sellier, F. Lefloch, M. Faur´ e, A. Buzdin, and H. Courtois, Phys.\nRev. B 72, 024511 (2005).\n1513E. A. Demler, G. B. Arnold, and M. R. Beasley, Phys. Rev. B 55, 15174 (1997)\n14V. A. Oboznov, V. V. Bolginov, A. K. Feofanov, V. V. Ryazanov, and A. I. Buzdin, Phys. Rev.\nLett. 96, 197003 (2006).\n15M. Faure, A. I. Buzdin, A. A. Golubov, and M. Yu. Kupriyanov, P hys. Rev. B 73, 064505\n(2006).\n16A. Buzdin and I. Baladie, Phys. Rev. B 67, 184519 (2003).\n17M. Weides, M. Kemmler, E. Goldobin, D. Koelle, R. Kleiner, H. Kohlstedt, and A. Buzdin,\nAppl. Phys. Lett. 89, 122511 (2006).\n18L. Usadel, Phys. Rev. Lett. 25507 (1970).\n19M. Yu. Kuprianov and V. F. Lukichev, Zh. Eksp. Teor. Fiz. 94, 139 (1988) [Sov. Phys. JETP\n67, 1163 (1988)].\n20V. N. Krivoruchko and E. A. Koshina, Phys. Rev. B 66, 014521 (2002).\n21F. S. Bergeret, A. F. Volkov, and K. B. Efetov, Phys. Rev. B 69, 174504 (2004).\n22K. K. Likharev, Rev. Mod. Phys. 51, 101 (1979).\n23A. Buzdin, JETP Lett 42, 583 (2003).\n24S. M. Frolov, M. J. A. Stoutimore, T. A. Crane, D. J. Van Harlin gen, V. A. Oboznov, V. V.\nRyazanov, A. Ruosi, C. Granata, and M. Russo, Nature Physics 4, 32 (2008).\n25S. E. Korshunov, Usp. Fiz. Nauk 49, 225 (2006).\n16" }, { "title": "0805.3925v2.Magnetic_anisotropy_in_ferromagnetic_Josephson_junctions.pdf", "content": "arXiv:0805.3925v2 [cond-mat.supr-con] 15 Jul 2008Magnetic anisotropy in ferromagnetic Josephson junctions\nM. Weides\nCenter of Nanoelectronic Systems for Information Technolo gy and Institute of Solid State Research,\nResearch Centre J¨ ulich, D-52425 J¨ ulich, Germany\n(Dated: December 3, 2018)\nMagnetotransport measurements were done on Nb /Al2O3/Cu/Ni/Nb superconductor-insulator-\nferromagnet-superconductor Josephson tunnel junctions. Depending on ferromagnetic Ni interlayer\nthickness and geometry the standard (1 d) magnetic field dependence of critical current deviates\nfrom the text-book model for Josephson junctions. The resul ts are qualitatively explained by a\nshort Josephson junction model based on anisotropy and 2 dremanent magnetization.\nPACS numbers: 74.25.Fy 74.45.+c 74.50.+r, 74.70.cn\nSuperconductivity (S) and ferromagnetism (F) in thin\nlayered films have now been studied during some decades\n[1]. In SF bilayers the superconductivity may be non-\nuniform[ 2], i.e. the Cooperpairwavefunction extends to\nthe ferromagnet with an oscillatory behavior. In Joseph-\nson junctions (JJs) based on s-wave superconductors the\nphase coupling between the superconducting electrodes\ncan be shifted by πwhen using a ferromagnetic barrier\nwith an appropriate chosen thickness dF, i.e. SFS or\nSIFS-type junctions (I: insulating tunnel barrier). Only\nin recent years the experimental realization of πJJs was\nsuccessful. In particular, the πcoupling was demon-\nstratedbyvaryingthe temperature[ 3,4,5], the thickness\nof the F-layer [ 5,6,7,8] or measuring the current-phase\nrelation of JJs incorporated into a superconducting loop\n[9,10,11]. The coupling can also change within a single\nJJ by a step-like F-layer, i.e. one half is a 0 JJ and the\nother half is a πJJ [12,13].\nFor useful classical or quantum circuits based on\nSFS/SIFS JJs a large critical current density jc(small\nJosephson penetration depth λJ) and a high IcRprod-\nuct are needed [ 14,15]. Up to now the limiting fac-\ntor is the low jcdue to strong Cooper pair breaking in-\nside F-layer. Alloys of magnetic and non-magnetic atoms\nsuch as NiCu face problems of clustering [ 16] and strong\nmagnetic scattering [ 5,8]. Promising experiments using\nstrong ferromagnet transition metals [ 6,17,18,19,20]\nwere published.\nShape anisotropy of magnetic interlayer may provoke a\nnot flux-closeddomainstructure andconsequentlya shift\nof critical current diffraction pattern Ic(H) [19]. In ex-\nperiments [ 3,5,7,19,20] the SFS/SIFS JJs had nearly\nmirror-symmetrical Ic(H), i.e. the effective shift along\nH-axis is small, usually less than one flux quantum Φ 0.\nThis is explained by a multi-domain state of F-layer with\na very small net magnetization. However, up to now the\n2dnature of thin-film magnetism was disregarded.\nIn this Letter the Ic(H) dependence for remanent 2 d\nmagnetization of F-layer is systematically studied. First,\nthe maximal flux from F-layer is estimated. Second, the\nIc(H) pattern considering 2 din-plane magnetization is\ncalculated for different aspect ratios. Third, the Ic(H)\npattern is measured alongboth magnetic axes for various\njunction geometries and dF./s48 /s49 /s50/s45/s48/s46/s48/s53/s48/s46/s48/s48/s48/s46/s48/s53/s48/s46/s49/s48\n/s48 /s49 /s50/s45/s48/s46/s48/s53/s48/s46/s48/s48/s48/s46/s48/s53/s48/s46/s49/s48\n/s48 /s49 /s50/s45/s48/s46/s48/s53/s48/s46/s48/s48/s48/s46/s48/s53/s48/s46/s49/s48\n/s48 /s49 /s50/s45/s48/s46/s48/s53/s48/s46/s48/s48/s48/s46/s48/s53/s48/s46/s49/s48/s48/s46/s48 /s48/s46/s53/s45/s48/s46/s49/s48/s46/s48/s48/s46/s49/s48/s46/s50/s48/s46/s51\n/s48/s46/s48 /s48/s46/s53/s45/s48/s46/s49/s48/s46/s48/s48/s46/s49/s48/s46/s50/s48/s46/s51\n/s48/s46/s48 /s48/s46/s53/s48/s46/s48/s48/s46/s50\n/s48/s46/s48 /s48/s46/s53/s48/s46/s48/s48/s46/s50/s48/s46/s48 /s48/s46/s53/s48/s46/s48/s48/s46/s53\n/s48/s46/s48 /s48/s46/s53/s48/s46/s48/s48/s46/s53\n/s48/s46/s48 /s48/s46/s53/s48/s46/s48/s48/s46/s53\n/s48/s46/s48 /s48/s46/s53/s48/s46/s48/s48/s46/s53\n/s77 /s32/s32\n/s50/s53/s32/s120/s32/s50/s48/s48/s32 /s109/s50/s53/s48/s32/s120/s32/s49/s48/s48/s32 /s109/s50/s51/s48/s32/s120/s32/s51/s48/s32 /s109/s50/s109/s97/s103/s110/s101/s116/s105/s99/s32/s102/s105/s101/s108/s100/s32 /s72\n/s120/s32/s40 /s109/s84 /s41\n/s109/s97/s103/s110/s101/s116/s105/s99/s32/s102/s105/s101/s108/s100/s32 /s72\n/s121 /s32/s40/s109/s84 /s41/s61/s48/s46/s49/s52/s50/s32 /s84 /s47/s109/s61/s48/s46/s48/s55/s49/s32 /s84 /s47/s109/s61/s48/s46/s50/s56/s52/s32 /s84 /s47/s109/s61/s48/s32 /s84 /s47/s109/s110/s111/s114/s109/s97/s108/s105/s122/s101/s100/s32/s99/s114/s105/s116/s105/s99/s97/s108/s32/s99/s117/s114/s114/s101/s110/s116/s32 /s73\n/s99/s47/s73/s48 /s99\n/s48/s48/s46/s50/s53/s48/s48/s48/s46/s53/s48/s48/s48/s48/s46/s55/s53/s48/s48/s49/s46/s48/s48/s48\nFIG. 1: (Color online) Calculated surface plot of Ic(Hx,Hy)\nfor different geometries and remanent magnetizations α. The\nmagnetization vector /vectorMpoints from top right to bottom left\ncorner (arrow), I0\ncis shifted in opposite direction.\nThe maximal shift of Ic(Hy) is estimated for a strong\nmagnet, i.e. Ni, being magnetized fully in-plane and\nalongy-axis. The atomic magnetic momentum is 0 .6µB\n[21], the specific density ρis 8.9g/cm3(bulk) and mag-\nnetization µ0M= 0.64 T. A cross-section of length\nLx= 100µm and F-layer thickness dF= 3nm encloses\na magnetic flux Φ M=dFLxµ0M. The total magnetic\nflux Φ through the JJ is the applied field flux Φ H=\n(2λL+dF)LxHy(Londonpenetrationdepth λL= 90nm)\nplus Φ M, i.e. Φ = Φ H±ΦM= 8.85Φ0Hy/mT±92Φ0.\nTheIc(Hy) pattern is shifted by 92 periods from the\ncenter. This simple calculation neglects dead magnetic\nlayer [22], as found in SFS/SIFS JJs [ 5,8,20], and de-\nmagnetizingbydomains. RealNi filmstendtoformcom-\nplex magnetization profiles (in/out-of-plane) and domain\nstructures as function of dF[23]. Integral magnetization\nmeasurements in SF layers show a complex behavior as a\nfunctionoftemperature,appliedfield, andsamplehistory\n[24,25],e.g. SFstructuresspontaneouslyaltertheirstray\nfield by changingmagnetic domain distribution [ 26]. The\nlocal magnetization depends on stray fields from neigh-\nbor domains, flux focusing from S-electrodes [ 19,27] or\non bias induced spin accumulation at F/S interface [ 28].\nIt is generally agreed that the average magnetization in\nSFS/SIFS JJs is much smaller than the maximum mag-2\n/s45/s50 /s45/s49 /s48 /s49 /s50/s48/s46/s48/s48/s46/s50/s48/s46/s52/s48/s46/s54/s48/s46/s56/s49/s46/s48\n/s45/s48/s46/s54 /s45/s48/s46/s52 /s45/s48/s46/s50 /s48/s46/s48 /s48/s46/s50 /s48/s46/s52 /s48/s46/s54/s48/s46/s48/s48/s46/s50/s48/s46/s52/s48/s46/s54/s48/s46/s56/s49/s46/s48\n/s45/s48/s46/s54 /s45/s48/s46/s52 /s45/s48/s46/s50 /s48/s46/s48 /s48/s46/s50 /s48/s46/s52 /s48/s46/s54/s48/s46/s48/s48/s46/s50/s48/s46/s52/s48/s46/s54/s48/s46/s56/s49/s46/s48/s45/s48/s46/s49/s48 /s45/s48/s46/s48/s53 /s48/s46/s48/s48 /s48/s46/s48/s53 /s48/s46/s49/s48/s48/s46/s48/s48/s46/s50/s48/s46/s52/s48/s46/s54/s48/s46/s56/s49/s46/s48\n/s45/s48/s46/s50 /s45/s48/s46/s49 /s48/s46/s48 /s48/s46/s49 /s48/s46/s50/s48/s46/s48/s48/s46/s50/s48/s46/s52/s48/s46/s54/s48/s46/s56/s49/s46/s48\n/s45/s48/s46/s54 /s45/s48/s46/s52 /s45/s48/s46/s50 /s48/s46/s48 /s48/s46/s50 /s48/s46/s52 /s48/s46/s54/s48/s46/s48/s48/s46/s50/s48/s46/s52/s48/s46/s54/s48/s46/s56/s49/s46/s48\n/s72\n/s120/s61/s48/s72\n/s121/s61/s48\n/s72\n/s121/s61/s48\n/s72\n/s121/s61/s48\n/s109/s97/s103/s110/s101/s116/s105/s99/s32/s102/s105/s101/s108/s100/s32 /s72\n/s121/s32/s40/s109 /s84/s41 /s32/s109/s97/s103/s110/s101/s116/s105/s99/s32/s102/s105/s101/s108/s100/s32 /s72\n/s121/s32/s40/s109 /s84/s41/s32 /s61/s48/s32 /s84/s47/s109 /s32/s32 /s32 /s61/s48/s46/s48/s55/s49/s32 /s84/s47/s109 /s32 /s32 /s61/s48/s46/s49/s52/s50/s32 /s84/s47/s109 /s32/s32 /s32 /s61/s48/s46/s50/s56/s52/s32 /s84/s47/s109/s32/s110/s111/s114/s109/s97/s108/s105/s122/s101/s100/s32/s99/s114/s105/s116/s105/s99/s97/s108/s32/s99/s117/s114/s114/s101/s110/s116/s32 /s73\n/s99/s47 /s73/s48 /s99\n/s32/s109/s97/s103/s110/s101/s116/s105/s99/s32/s102/s105/s101/s108/s100/s32 /s72\n/s121/s32/s40/s109 /s84/s41/s32/s109/s97/s103/s110/s101/s116/s105/s99/s32/s102/s105/s101/s108/s100/s32 /s72\n/s120/s32/s40/s109 /s84/s41\n/s72\n/s120/s61/s48/s72\n/s120/s61/s48/s32/s109/s97/s103/s110/s101/s116/s105/s99/s32/s102/s105/s101/s108/s100/s32 /s72\n/s120/s32/s40/s109 /s84/s41/s32/s32/s110/s111/s114/s109/s97/s108/s105/s122/s101/s100/s32/s99/s114/s105/s116/s105/s99/s97/s108/s32/s99/s117/s114/s114/s101/s110/s116/s32 /s73\n/s99/s47 /s73/s48 /s99\n/s32/s109/s97/s103/s110/s101/s116/s105/s99/s32/s102/s105/s101/s108/s100/s32 /s72\n/s120/s32/s40/s109 /s84/s41/s51/s48/s32/s120/s32/s51/s48/s32 /s109/s50\n/s53/s48/s32/s120/s32/s49/s48/s48/s32 /s109/s50\n/s50/s53/s32/s120/s32/s50/s48/s48/s32 /s109/s50\nFIG. 2: (Color online) Calculated Icas function of 1 dmag-\nnetic field, Ic(Hx,0) (top) and Ic(0,Hy) (bottom) for the\nthree geometries and various α. A substantial deviation from\nideal (α= 0T/m) pattern appears for already small magne-\ntization vector /vectorM=α(Lx,Ly).\n/s45/s48/s46/s54 /s45/s48/s46/s53 /s45/s48/s46/s52 /s45/s48/s46/s51 /s45/s48/s46/s50 /s45/s48/s46/s49 /s48/s46/s48 /s48/s46/s49 /s48/s46/s50 /s48/s46/s51 /s48/s46/s52 /s48/s46/s53 /s48/s46/s54/s48/s50/s48/s48/s52/s48/s48/s54/s48/s48/s56/s48/s48/s49/s48/s48/s48/s49/s50/s48/s48\n/s106\n/s99/s61/s50/s51/s46/s51/s177/s48/s46/s52/s32 /s65/s47/s99/s109/s50\n/s32/s40 /s61/s49/s46/s55/s37/s41/s100\n/s70/s61/s50/s46/s50/s32 /s110/s109\n/s32/s32/s32/s32 /s76\n/s120/s32/s120/s32 /s76\n/s121/s32/s40 /s109/s50\n/s41/s32/s32/s32 /s76\n/s120/s47\n/s74/s32/s120/s32 /s76\n/s121/s47\n/s74\n/s32/s51/s48/s32/s120/s32/s51/s48/s32/s32/s32/s32/s32/s40/s48/s46/s52/s32/s120/s32/s48/s46/s52/s41\n/s32/s53/s48/s32/s120/s32/s49/s48/s48/s32/s32/s32/s40/s48/s46/s54/s53/s32/s120/s32/s49/s46/s51/s41\n/s32/s50/s53/s32/s120/s32/s50/s48/s48/s32/s32/s32/s40/s48/s46/s51/s50/s32/s120/s32/s50/s46/s54/s41\n/s86\n/s99/s61 /s49 /s32 /s86\n/s32/s109/s97/s103/s110/s101/s116/s105/s99/s32/s102/s105/s101/s108/s100/s32 /s72\n/s121/s32/s40 /s109/s84 /s41\n/s32/s32/s99/s114/s105/s116/s105/s99/s97/s108/s32/s99/s117/s114/s114/s101/s110/s116/s32 /s73\n/s99/s32/s40 /s65 /s41\n/s45/s48/s46/s53 /s48/s46/s48 /s48/s46/s53/s48/s53/s49/s48/s49/s53/s50/s48\n/s72\n/s121\n/s72\n/s120/s100\n/s70/s61/s51/s46/s57/s32 /s110/s109\n/s51/s48/s32/s120/s32/s51/s48/s32 /s109/s50\n/s32\n/s86\n/s99/s61 /s49 /s32 /s86/s53/s48/s32/s120/s32/s49/s48/s48/s32 /s109/s50\n/s86\n/s99/s61 /s48/s46/s50 /s32 /s86\n/s50/s53/s32/s120/s32/s50/s48/s48/s32 /s109/s50\n/s86\n/s99/s61 /s48/s46/s53 /s32 /s86/s50/s53/s32/s120/s32/s50/s48/s48/s32 /s109/s50\n/s32\n/s86\n/s99/s61 /s48/s46/s53 /s32 /s86/s50/s53/s32/s120/s32/s50/s48/s48/s32 /s109/s50\n/s86\n/s99/s61 /s48/s46/s50 /s32 /s86/s45/s48/s46/s53 /s48/s46/s48 /s48/s46/s53/s48/s53/s48/s49/s48/s48\n/s45/s48/s46/s53 /s48/s46/s48 /s48/s46/s53/s48/s50/s48/s52/s48/s54/s48/s56/s48\n/s53/s48/s32/s120/s32/s49/s48/s48/s32 /s109/s50\n/s86\n/s99/s61 /s49 /s32 /s86\n/s45/s49/s46/s48 /s45/s48/s46/s53 /s48/s46/s48 /s48/s46/s53 /s49/s46/s48/s48/s53/s49/s48/s49/s53/s50/s48/s50/s53/s51/s48\n/s32/s109/s97/s103/s110/s101/s116/s105/s99/s32/s102/s105/s101/s108/s100/s32 /s72\n/s120/s44/s121/s32/s40 /s109/s84 /s41/s99/s114/s105/s116/s105/s99/s97/s108/s32/s99/s117/s114/s114/s101/s110/s116/s32 /s73\n/s99/s32/s40 /s65 /s41\n/s45/s49/s46/s48 /s45/s48/s46/s53 /s48/s46/s48 /s48/s46/s53 /s49/s46/s48/s48/s53/s49/s48/s49/s53/s50/s48/s50/s53/s51/s48\n/s45/s49/s46/s48 /s45/s48/s46/s53 /s48/s46/s48 /s48/s46/s53 /s49/s46/s48/s48/s49/s48/s50/s48/s51/s48/s52/s48\nFIG. 3: (Color online) Measured Ic(Hy) pattern of SIFS JJs\nfor thin (top) and thick (bottom) Ni layer and different ge-\nometries. Ic(Hx) pattern of thick Ni layer is plotted in gray.\nOnset of magnetic anisotropy effects in Ni layer is between\n2.2 and 3.9nm.\nnetization estimated above. But as shown below even a\nremanent 2 dmagnetization of 1% of the maximal value\nmay notable change the Ic(H) pattern.\nA qualitativemodel for Ic(H) in presenceofa uniform,fixed 2dmagnetization /vectorMis derived. The short JJ model\nIc(H) =I0\nc/vextendsingle/vextendsingle/vextendsingle/vextendsingle/vextendsinglesin(πΦ\nΦ0)\nπΦ\nΦ0/vextendsingle/vextendsingle/vextendsingle/vextendsingle/vextendsingleand Φ = Φ H±ΦM\nis modified by 2 ddistributions of applied field flux Φ H\nand magnetization Φ Mwith\nΦ =/vextendsingle/vextendsingle/vextendsingle/vextendsingle/bracketleftbigg\n(2λL+dF)/parenleftbigg\n0Hx\nHy0/parenrightbigg\n±/parenleftbigg\nα0\n0α/parenrightbigg/bracketrightbigg\n×/parenleftbigg\nLx\nLy/parenrightbigg/vextendsingle/vextendsingle/vextendsingle/vextendsingle.\n/vectorM=α(Lx,Ly) is assumed to be orientated in-plane\n(Meissner-screening of S-electrodes) along the diagonal\n(i.e. longest) axis of sample (rough approximation for\nmagnetic shape anisotropy). The magnetization |/vectorM|is\nseveral orders of magnitude smaller than the upper limit\ngiven by a fully saturated magnetic layer. Note that the\neasy axis of ferromagnetic film can be determined by the\nmagnetic field during deposition, too. The model for Φ M\nis just exemplary for the effect of 2 din-plane magnetiza-\ntion inIc(H).\nIn Fig.1the surface plot of Ic(Hx,Hy) is depicted for\nvariousαand geometries. The position of I0\ncis shifted\nfrom the center ( Hx=Hy= 0mT) in opposite direction\nof/vectorM. Fig.2depictsIc(Hx,0) andIc(0,Hy) pattern.\nThese graphs resemble standard 1 d Ic(H) (H=Hx,Hy)\nmeasurements. For some |/vectorM| /negationslash= 0 T a single-peaked\nIc(Hx),Ic(Hy) pattern is calculated, which -on first\nglance- resembles a shifted |sin(H)/H|Fraunhofer pat-\ntern. However, the maximum Icis smaller than the\nrealI0\ncand the height of side-maxima do not obey the\nexpected value. For example the Ic(Hx,0) pattern of\n30×30µm2sample and α= 0.071 T/m is shifted by\nless than Φ 0and its maximum Icis already reduced to\n∼0.8I0\nc. This simple Ic(Hx,Hy) modelmayqualitatively\nexplainsomeexperimentalobservations(Fig. 3)onSIFS-\ntype JJs with Ni interlayer.\nForexperiment JJs with similar areas,i.e. 30 ×30,50×\n100 and 25 ×200µm2, were fabricated. The deposition\nand structuring of JJs is described in Ref. 29. The\nSIFS multilayer was magnetron sputter deposited with\nNi thickness dFranging from 1 −6nm. The tunnel bar-\nrier was formed for 30min at a partial oxygen pressure\nof 0.1 mbar. After oxidation a 2 nm Cu film was in-\nserted. All JJs were deposited in a single run by shift-\ning the substrate and target to obtain a wedge-shaped\nNi-layer. Normal state and subgap resistance indicate a\nsmall junction to junction variation. The IV and Ic(Hx),\nIc(Hy) characteristics were measured at 4 .2 K for two\nsets of samples ( dF= 2.2,3.9nm). Cooldown was done\nin zero field and thermal cycling up to ≈15 and 300K\nto check reproducibility. Transport measurements were\nmadein aliquid Hedip probeusinglow-noisehomemade\nelectronics and room-temperature voltage amplifier. The\nmagnetic fields ( Hx,0), (0,Hy) were applied in-plane of\nthe sample and parallel the junctions axis (Fig. 1). The\nvoltagecriteria VcforIc(H)determinationwas0 .2−1µV.\nA lower subgap resistance for dF= 3.9nm sample leads3\nto larger offset currents. Positive and negative current\nbranch of IVC had similar magnetic field dependence\n+Ic(Hy)≃ |−Ic(Hy)|. Magnetic field wassweptbetween\n±1.5mT for all samples. All junctions had their lateral\nsizes comparable or smaller than the Josephson penetra-\ntion length λJ, except the longest sample ( dF= 2.2nm,\n25×200µm2), whose Lyis not strictly inside the short\nJJ limit.\nThedF= 2.2 nm samples showed very regular Ic(Hy)\npattern. All maximum Ic’s were nearly centered and\nthe spread of jcwas∼1.7%, as determined from the\nmaximum Ic’s. The Ic(Hx) pattern is symmetric, too\n(not shown). The oscillation period of Ic(Hy) were de-\ntermined by magnetic cross-section ∼1/Lx, and nearly\nindependent ofaspect ratio. No indication fora distorted\nsupercurrent transport due to alloying at the Nb /Ni\ninterface [ 30] can be found. Effects due to magnetic\nanisotropy were not detectable, either because the sam-\nples were still inside the dead magnetic layer, or the\nanisotropy was absent or totally out-of-plane.\nThedF= 3.9nmsampleshadcompletelydifferent Ic(Hy)\npattern showing in-plane magnetic anisotropy with some\ncharacteristic features. All maximum Ic’s were shifted\nfrom the center, and the amplitude of shift increased\nwithLy, i.e.≈0.24mT for 30 ×30µm2,≈0.5mT for\n50×100µm2, and≈0.8mTfor25 ×200µm2samples. The\ndirectionofshift variesbetween sampleseveniftheywere\ncooled and measured at the same time (random polarity\nof magnetic configuration). The position of main peak of\nIc(Hy) was reproducible after thermal cycling to 300 K.\nThe width of main maxima (measured at offset line) was\nnot strictly ∼1/Lx, and varies from sample to sample.\nThe pattern were asymmetric, i.e. the height of same-\norder side maxima differed, probably due to non-uniform\nflux guidance in F-layer and re-orientation of domains.\nBy rotating the magnetic field by 90◦, i.e. measuring in\nIc(Hx) mode, low Ic’s, being nearly independent of Hx,\nwere detected. Even the squared shaped 30 ×30µm2\nJJ had an almost flat Ic(Hx) pattern. This indicates\nsome magnetic crystallographic anisotropy along y-axis,\nprobably caused by magnetron sputter deposition. Small\ndeviations of Ic(0,0) forIc(Hx,0) andIc(0,Hy) measure-ments can be related to variations of magnetic configu-\nration by the unshielded sample handling at 300 K. A\nconsiderable spread of maximum Ic(Fig.3) can be al-\nready seen for JJs with same geometry, which is even\nincreased by considering the maximum jcfor different\ngeometries. Simulations (Fig. 2) show that already a\nmoderate magnetization /vectorM(α <0.1 T/m) yields very\ndifferent maximum Ic’s ofIc(Hx) andIc(Hy). A sample\nto sample variation of direction and amplitude of /vectorMex-\nplains the data spread of Fig. 3. Obviously, this leads to\nvery large variations in the Ic(dF) dependence.\nIn literature, the Ic(H) pattern of SFS/SIFS JJs with\ncomparable strong magnets were either shown for sam-\nples with thin dF[19], or had deviations from the ideal\n|sin(H)/H|form. For example, the maximum Icin Ref.\n20, Fig. 5 (inset) is too small compared to the first side-\nmaxima. These samples were small in area ( /lessorequalslant1µm2),\nand the F-layer could have been in single domain state.\nAsIcvaried smoothly with H, either /vectorMrotated softly,\nor a multi-domain structure with averaged 2 dmagneti-\nzation existed. For both cases the 2 dnature of remanent\nmagnetization may have suppressed the maximum Ic-\ndetermined from Ic(H) pattern- below the real I0\nc.\nIn summary, the Ic(H) pattern along both field axis\nof SIFS JJs with Ni interlayer were measured. Assum-\ning magnetic anisotropy the characteristic features, i.e.\nshift or absence of central peak, can be qualitatively re-\nproduced by simulations. As conclusion, the 1 d Ic(H)\npattern in presence of magnetic anisotropy can not yield\ntherealI0\nc. Future experiments on SFS/SIFS JJs should\nbe done by 2-axis Ic(Hx,Hy) scan of JJs with well-\ncontrolled magnetic interlayer domain configuration. Su-\nperconducting spintronic devices such as FSF spin valves\nandπ, 0–πJJs are promising candidates for future clas-\nsical and quantum computers. As shown by this Letter\ncontrolofmagneticanisotropyandmagneticdomaincon-\nfigurationis essentialforphase-switchableFSF/SFIFS or\nSFS/SIFS devices.\nM.W. thanks D. Sprungmann for proof-reading, A.\nBannykh for help with fabrication and DFG for support\n(project WE 4359/1-1).\n[1] A. I. Buzdin, Rev. Mod. Phys. 77, 935 (2005).\n[2] E. A. Demler, G. B. Arnold, and M. R. Beasley, Phys.\nRev. B55, 15174 (1997).\n[3] V. V. Ryazanov, V. A. Oboznov, A. Y. Rusanov, A. V.\nVeretennikov, A. A. Golubov, and J. Aarts, Phys. Rev.\nLett.86, 2427 (2001).\n[4] H. Sellier, C. Baraduc, F. Lefloch, and R. Calemczuk,\nPhys. Rev. B 68, 054531 (2003).\n[5] M. Weides, M. Kemmler, E. Goldobin, D. Koelle,\nR. Kleiner, H. Kohlstedt, and A. Buzdin, Appl. Phys.\nLett.89, 122511 (2006).\n[6] Y. Blum, A. Tsukernik, M. Karpovski, and A. Palevski,\nPhys. Rev. Lett. 89, 187004 (2002).\n[7] T. Kontos, M. Aprili, J. Lesueur, and X. Grison, Phys.Rev. Lett. 89, 137007 (2002).\n[8] V. A. Oboznov, V. V. Bol’ginov, A. K. Feofanov, V. V.\nRyazanov, and A. I. Buzdin, Phys. Rev. Lett. 96, 197003\n(2006).\n[9] A.Bauer, J.Bentner, M.Aprili, M.L.D.Rocca, M.Rein-\nwald, W. Wegscheider, and C. Strunk, Phys. Rev. Lett.\n92, 217001 (2004).\n[10] V. V.Ryazanov, V.A. Oboznov, A.V. Veretennikov,and\nA. Y. Rusanov, Phys. Rev. B 65, 20501 (2001).\n[11] W. Guichard, M. Aprili, O. Bourgeois, T. Kontos,\nJ. Lesueur, and P. Gandit, Phys. Rev. Lett. 90, 167001\n(2003).\n[12] M. Weides, M. Kemmler, H. Kohlstedt, R. Waser,\nD. Koelle, R. Kleiner, and E. Goldobin, Phys. Rev. Lett.4\n97, 247001 (2006).\n[13] M. Weides, C. Schindler, and H. Kohlstedt, J. Appl.\nPhys.101, 063902 (2007).\n[14] A. V. Ustinov and V. K. Kaplunenko, J. Appl. Phys. 94,\n5405 (2003).\n[15] T. Kato, A. A. Golubov, and Y. Nakamura, Phys. Rev.\nB76, 172502 (2007).\n[16] R. W. Houghton, M. P. Sarachik, and J. S. Kouvel, Phys.\nRev. Lett. 25, 238 (1970).\n[17] V. Shelukhin, A. Tsukernik, M. Karpovski, Y. Blum,\nK. B. Efetov, A. F. Volkov, T. Champel, M. Eschrig,\nT. L¨ ofwander, G. Sch¨ on, and A. Palevski, Phys. Rev. B\n73, 174506 (2006).\n[18] J. W. A. Robinson, S. Piano, G. Burnell, C. Bell, and\nM. G. Blamire, Phys. Rev. Lett. 97, 177003 (2006).\n[19] C. Bell, R. Loloee, G. Burnell, and M. G. Blamire, Phys.\nRev. B71, 180501 (2005).\n[20] J. W. A. Robinson, S. Piano, G. Burnell, C. Bell, and\nM. G. Blamire, Phys. Rev. B 76(2007).\n[21] S. A. Ahern, M. J. C. Martin, and W. Sucksmith, Proc.\nRoy. Soc. 248, 145 (1958).[22] L. Liebermann, J. Clinton, D. M. Edwards, and\nJ. Mathon, Phys. Rev. Lett. 25, 232 (1970).\n[23] G. Gubbiotti, G. Carlotti, M. G. Pini, P. Politi, A. Ret-\ntori, P. Vavassori, M. Ciria, , and R. C. OHandley, Phys.\nRev. B65, 214420 (2002).\n[24] C. Monton, F. de la Cruz, and J. Guimpel, Phys. Rev. B\n77, 104521 (2008).\n[25] A. G. Joshi, S. A. Kryukov, L. E. D. Long, E. M. Gonza-\nlez, E. Navarro, J. E. Villegas, and J. L. Vicent, J. Appl.\nPhys.101, 09G117 (2007).\n[26] S. V. Dubonos, A. K. Geim, K. S. Novoselov, and I. V.\nGrigorieva, Phys. Rev. B 65(2002).\n[27] H. Wu, J. Ni, J. Cai, Z. Cheng, and Y. Sun, Phys. Rev.\nB76, 024416 (2007).\n[28] F. J. Jedema, B. J. van Wees, B. H. Hoving, A. T. Filip,\nand T. M. Klapwijk, Phys. Rev. B 60, 16 549 (1999).\n[29] M. Weides, K. Tillmann, and H. Kohlstedt, Physica C\n437-438 , 349 (2006).\n[30] H. Chen, Y. Du, H. Xu, Y. Liu, and J. C. Schuster, J.\nMater. Sci. 40, 6019 (2005)." }, { "title": "1208.5629v1.Triplet_supercurrent_in_ferromagnetic_Josephson_junctions_by_spin_injection.pdf", "content": "arXiv:1208.5629v1 [cond-mat.supr-con] 28 Aug 2012Triplet supercurrent in ferromagnetic Josephson junction s by spin injection\nA. G. Mal’shukov1and Arne Brataas2\n1Institute of Spectroscopy, Russian Academy of Sciences, 14 2190, Troitsk, Moscow oblast, Russia\n2Department of Physics, Norwegian University of Science and Technology, NO-7491 Trondheim, Norway\nWe show that injecting nonequilibrium spins into the superc onducting leads strongly enhances\nthe stationary Josephson current through a superconductor -ferromagnet-superconductor junction.\nThe resulting long-range super-current through a ferromag net is carried by triplet Cooper pairs\nthat are formed in s-wave superconductors by the combined eff ects of spin injection and exchange\ninteraction. We quantify the exchange interaction in terms of Landau Fermi-liquid factors. The\nmagnitude and direction of the long-range Josephson curren t can be manipulated by varying the\nangles of the injected polarizations with respect to the mag netization in the ferromagnet.\nPACS numbers: 72.25.Dc, 71.70.Ej, 73.40.Lq\nI. INTRODUCTION\nStudies of hybrid structures combining supercon-\nducting and ferromagnetic components attract much\nattention due to their unique, rich, and complex\nphysical properties that are promising in a number\nof potential applications1. The interface of an s-\nwave superconductor with a ferromagnet is charac-\nterized by an unusual proximity effect that is spa-\ntially oscillating and can lead to a sign reversal of\nthe criticalcurrentthroughsuperconductor-ferromagnet-\nsuperconductor(SFS) Josephsonjunctions. Sucharever-\nsal is equivalent to a π-shift in the current-phase relation\nfor the Josephson current. This interesting property is\na motivation for using the so-called π-junctions as el-\nements of superconducting quantum circuits for poten-\ntial application in quantum computing2. However, the\nproximity effect in ferromagnets does not reach far. Two\ncritical tasks are to extend its range and to find a way to\nmanipulate the π-junction in order to switch the device\nbetween its various phase states. In contrast, Cooper-\npairs can be transferred over relatively long distances\neven in ferromagnets, if they are in a triplet state with\n±1 projections of their total spin onto the spin quantiza-\ntion axis. Various mechanisms have been proposed that\nconvert a singlet pair into a triplet pair, such as a spa-\ntially dependent magnetization3, spin-flip scattering at\nFS interfaces4, and precessing magnetization5. A num-\nber of works in this direction has been reviewed in Ref.1.\nIn this work, we will show that these tasks can be ful-\nfilledviatheproductionandmanipulationofalong-range\nproximity effect by injecting spins into superconducting\nleads. The novelty of our idea is based on the impor-\ntant, and so far unaddressed, role played by the electron-\nelectron interaction in SFS. Our insight is that the com-\nbined effects of spin-injection and electron-electron inter-\naction generate a long-range proximity effect despite the\nstrong exchange field in the ferromagnet. The conven-\ntional wisdom is that spin polarized electrons can only\nexist as excitations in s-wave superconductors, since the\nCooper pairs do not carry a spin. However, we will\ndemonstrate that this simple picture, which is based onthe neglect of electron-electron interactions beyond su-\nperconducting pairing correlations, misses qualitatively\nimportant effects. Quantitatively, in simple metals, the\nexchange interaction of itinerant carriers is noticeable\nandcanbedescribedintermsofLandauFermi-liquidfac-\ntors. Although the exchange interaction does not cause\nferromagnetism in s-wave superconductors, it causes a\ntransfer of spin polarization from the quasi-particle exci-\ntations to the condensate, in the form of polarized triplet\nCooper pairs. When such a triplet pairing is generated\nby the combined effects of spin-injection and exchange\ninteraction, these pairs subsequently tunnel through the\nferromagneticlayer via the long-rangeproximity effect, if\nthe spin polarizations in the leads and the layer are not\ncollinear. Only at this stage, which includes the so far\nunaddressed important electron-electron interaction, the\nsituation becomes similar to proposals of Ref.1,3 where\nan inhomogeneous magnetization gives rise to the long-\nrange effect provided by ±1 triplets. The relative angles\nbetween the spin polarizations in the superconducting\nleads and in the ferromagnet can be varied by control-\nling the injected spin polarizations, making it possible to\nvary the magnitude and sign of the Josephson current.\nThis enables manipulations of π-junctions. In addition\nto the Josephson supercurrent, which is driven by the\ndifference in the condensate phases, there is also a dissi-\npative DC current. The latter is induced by the spin po-\nlarization flow through the ferromagnetic layer with spin\ndependent conductivity. This dissipativecurrentalsocan\nbe manipulated by varying the injected polarization an-\ngles. As it will be shown, at some angles it vanishes, so\nthat the dissipative and supercurrents can be measured\nindependently.\nVariouseffectsofaninjected spinpolarizationandspin\ncurrent on the electric transport in SFS junctions6,7and\nother superconducting systems8–11have been recently\nconsidered. Despite this interest, the fact that the ex-\nchange interaction transfers the spin polarization from\nthe quasiparticles to the condensate has not been ad-\ndressed so far.\nThe article is organized by the following way. In Sec.II\nan expression is derived connecting the triplet compo-\nnents of the anomalous Green function to the nonequi-2\nFIG. 1: (Color online) A sketch of the system. The electric\ncurrent flows in normal leads N through contacts with ferro-\nmagnetic leads F Land F R. Spin density is injected from F L\nand F Rinto N and further penetrates across tunneling barri-\ners into superconductors S Land S R. The Josephson current\nflows between these leads through a ferromagnetic layer F.\nArrows show possible magnetizations of the ferromagnets.\nlibrium spin polarization in superconducting leads. In\nSec.III the Josephson and dissipative currents are calcu-\nlated. Finally, our results are discussed in Sec.IV.\nII. TRIPLET ELECTRON PAIRING FUNCTION\nINDUCED BY SPIN INJECTION\nHow to efficiently inject a spin polarization into para-\nmagnetic metals is well known12. A nonequilibrium spin\naccumulation is induced by the electric current through\na paramagnetic-ferromagneticinterface. We consider the\nscenario that the spin polarization further diffuses from\na paramagnetthrough a resistive barrierinto a supercon-\nducting lead, so that the electric circuit where the spin\ninjection takes place is effectively separated from the su-\nperconducting circuit. We assume that the steady state\nspin polarizations are generated in both superconduct-\ning leads, in the vicinity of the F-layer. The sketch of\nthe system is shown in Fig.1. For clarity, we simplify\nthe problem by assuming that the FS contacts contain\na barrier, so that the proximity effect is weak. We also\nassume that the spin relaxation time τspinin the leads is\nlong, so that the spin diffusion length lspinis large com-\npared to the SN contact sizes and the coherence length.\nConsequently, the spin densities sL(R)and the order pa-\nrameters ∆ L(R)only vary slowly in space near the left\n(L) and right ( R) contacts.\nThe electronic transport through an SFS system,\nwhose characteristic dimensions are larger than the elas-\ntic mean free path, can be described in terms of Us-\nadel equations for angular averaged Green’s functions\ng(for a review see13). These functions are matrices in\nthe Keldysh, spin, and Nambu spaces. We choose the\nspin and Nambu spaces so that the one-particle destruc-\ntion operators are c1k↑=ck↑,c1k↓=ck↓,c2k↑=c†\n−k↓,\nc2k↓=−c†\n−k↑, where the labels 1 and 2 denote theNambu spinor components, while ↑and↓are the spin\nindices. The Keldysh component gKof the Green func-\ntion can be represented as13\ngK=grh−hga, (1)\nwheregrandgaaretheretardedandadvancedfunctions,\nrespectively, and the distribution function his a diagonal\nmatrix in the Nambu space.\nIn order to determine the distribution hin the super-\nconducting leads, the interfaces between these leads and\nthespin-polarizednormalmetalsmustbeconsidered. We\nusestandardboundaryconditionsrelatingfluxesthrough\nS-N (S-F) interfaces to Green functions in superconduc-\ntors and normal metals (ferromagnets). It is assumed\nthat the spin relaxation rates in the superconducting\nleads are slow enough ( lspin≫rsnσs) and the leak-\nage of the spin polarization through the SF boundary is\nsufficiently slow rsn/Asn≪rsf/Asf, where 1 /rsnand\n1/rsfare the interface conductances (per unit square) of\nSN and SF interfaces, AsnandAsfare the SN and SF\ncontact areas, and σsis the normal-state conductivity of\nthe superconductor’s lead. With these assumptions, the\ndistribution functions in the superconductor, h(s), and\nnormal metal, h(n), are equal to each other, h(s)=h(n).\nWe further assume that nonequilibrium spins in N-leads\nare thermalized with chemical potentials µ↑andµ↓for\nthe two spin directions. Therefore, denoting by the sub-\nscripts 11 and 22 the corresponding matrix elements in\nthe Nambu space, we get for h↑(↓)≡h(s)\n11↑(↓)=h(n)\n11↑(↓)\nand¯h↑(↓)≡h(s)\n22↑(↓)=h(n)\n22↑(↓)\nh↑(↓)=¯h↑(↓)= tanhω−µ↑(↓)\n2kBT. (2)\nAt the same time, the retarded ( gr) and advanced ( ga)\nGreenfunctions havethe sameforms asin an equilibrium\nsuperconductor.\nOur calculation so far re-iterates the conventional wis-\ndom of spin-injection in superconductors: the effects are\nlimited to a spin-dependent statistical distribution func-\ntion, whiletheretardedandadvancedGreenfunctionsdo\nnotchange. Inthispicture, spininjectiondoesnotleadto\nthe appearance of triplet correlations in the condensate\nwave-function, which would cause long-range Josephson\ntunneling through a ferromagnetic layer. Fortunately,\nthere is a mechanism to generate triplet correlations in\nspin-polarized superconducting leads, which others have\nso far overlooked. The electron-electron exchange in-\nteraction provides a coupling between a spin accumu-\nlation and the spectral properties of superconductors,\nin that spin polarized quasiparticles produce an effec-\ntive Zeeman field. The latter, in its turn, gives rise to\ntripletcorrelationsthat aredescribed viathe correspond-\ning spin components of the anomalous functions gr\n12and\nga\n12. In Fermi-liquid theory, the effective Zeeman energy\nisǫxc(σN), where Nis a unit vector parallel to the in-\njected spin polarization S=NSand\nǫxc= GS/2NF. (3)3\nThe spin-accumulation magnitude is\nS=−NF\n4(1+G)/integraldisplay\ndωTr[(1+τ3)\n2σzgK],(4)\nwhereτ3andσzare the Pauli matrices acting in the\nNambu and spin spaces, respectively, and NFis the den-\nsity of states at the Fermi level. The renormalization\nfactor 1/(1+G), where G is the exchange Landau-Fermi\nliquid parameter, appears when the spin-density of Eq.\n4 is expressed in terms of a semiclassical Green func-\ntion integrated over energy14. This factor is not quali-\ntatively important in our case, since G is not too close\nto the paramagnet instability G = −1.15The exchange\nCoulomb interaction in metals gives rise to a negative G.\nFor example, the calculated value is -0.17 in Al16. The\nspindensity(4) stronglydependsontemperature, mostly\nvia the temperature dependence of the superconducting\ngap in the energy spectrum. In order to determine Sand\n∆ in both leads, Eq. (4) have to be solved together with\ntheS-depended selfconsistency equation for ∆.9\nVia the effective Zeeman energy of Eq.3, the retarded\nand advanced Green functions become spin-dependent.1\nIndeed, choosing the quantization axis along N, the\nanomalous functions fr\n↑↓=gr\n12↑↑andfr\n↓↑=−gr\n12↓↓be-\ncome\nfr\n↑↓(↓↑)=±|∆|exp(iφ)/radicalbig\n(ω∓ǫxc+iδ)2−|∆|2,(5)\nwhere the phase φof the order parameter ∆ equals φL\nandφRat the left and right contacts, respectively. The\ntriplet component of this function with 0-spin-projection\nonto the z-axis isfr\n0= (fr\n↑↓+fr\n↓↑)/√\n2, while the triplet\ncomponents with ±1-projections vanish, fr\n±1=fr\n↑↑(↓↓)=\n0. Theadvancedfunction, aswellastheconjugatedfunc-\ntionsf†, are determined from symmetry relations.\nIt ismoretransparentto discussthe Greenfunctions in\na basis where the spin quantization axis is parallel to the\nmagnetization in the ferromagnetic layer, which is along\nz, as shown in Fig. 1. We assume that the spin polariza-\ntionsintheleftandrightleadsarerotatedwithrespectto\nthis axis by the angles θLandθR, respectively. We follow\nthe convention that the three components of the triplet\nf0,f1,f−1arerelated to a 3Dvector a= (ax,ay,az) with\naz=f0,ax= (f−1−f1)/√\n2 anday=i(f−1+f1)/√\n217.\nHence, in the geometry shown in Fig. 1, after a rota-\ntion ofaaround the y-axis, we get in the new basis\nf′\n0=f0cosθandf′\n1=−f′\n−1=−f0sinθ/√\n2. So, by\nusing Eq. (5) the triplet components in the left and right\nsuperconducting leads are\nf±1R(L)=−sinθR(L)\n2(f↑↓+f↓↑), (6)\nwherethelabels randahavebeenomitted fromhereand\nthe same magnitudes of ǫxcare assumed in both leads.In the new basis, the distribution function (2) is\nhL(R)=h↑(1+σzcosθL(R))\n2+h↓(1−σzcosθL(R))\n2+\nσysinθL(R))h↑−h↓\n2. (7)\nIII. THE JOSEPHSON AND DISSIPATIVE\nCURRENTS\nWhat we have established is that the superconduct-\ning leads acquire triplet pairing correlations determined\nby non-equilibrium spin polarizations whose directions\nare tilted with respect to the ferromagnet’s magnetiza-\ntion in the SFS junction. We will show that the current\nthroughsuchatriplet pairing-ferromagnet-tripletpairing\nsystem consists of two parts, a dissipative contribution\ncontrolled by the non-equilibrium distribution of spins\nin the device, and a super-current driven by the phase\ndifference between superconductors and provided by the\ntriplet components of the superconducting condensates\nin the left and right leads.\nLet us first consider the dissipative current. It can be\nexpressed in terms of the distribution function hfinside\nthe ferromagnet. Due to precession in the exchange field\nBex, the spins that are not parallel to it decay quickly\non the length-scale/radicalbig\nDf/Bex, whereDfis the diffusion\nconstant. Therefore, only the components of hfthat are\nparallel and anti-parallel to z, denoted as hf↑andhf↓, re-\nmain finite inside the ferromagnet, if the junction length\nL≫/radicalbig\nDf/Bex. When the spin relaxation length is\nlarger than L, in the linear approximation these collinear\ncomponents satisfy the spin-conserving diffusion equa-\ntionDfσ∇2\nxhfσ= 0, where σ=↑,↓, that takes into ac-\ncount spin-dependent diffusion coefficient in a strong fer-\nromagnet. The solution of this equation is a linear func-\ntion ofxwhose slope is obtained from the boundary con-\nditions∓rsfσσfσ∇xhfσ|x=xL(R)=hL(R)σ−hfσ|x=xL(R),\nwherehL(R)σare given by the first two terms of Eq. (7).\nTaking into account that rsfσandσfσcan depend on the\nelectron spin and assuming equal barrier transmittances\nat L and R contacts we obtain\nhfσ=hRσ+hLσ\n2+hRσ−hLσ\n1+2γσx\nL, (8)\nwherexR(L)=±L/2 andγσ= (rsfσσfσ/L)≫1. Using\nEqs. (8) and (7) we compute the dissipative part of the\ncurrent through the junction:\njd=/summationdisplay\nσ/integraldisplay\ndωσfσ∇xhfσ=\nδµ\neL/parenleftbiggσf↑\n1+2γ↑−σf↓\n1+2γ↓/parenrightbigg\n(cosθR−cosθL).(9)\nThis current is proportional to the difference in the spin-\nup and spin-down conductances (2 rsfσ+L/σfσ)−1of the\ntotal ferromagnetic layer, including the interfaces; this4\nis the well known19connection between spin and elec-\ntric transport in ferromagnets. The electric current at-\ntains its maximum when cos θR=−cosθL=±1, and\nvanishes at θR=θL, as well as at θR,θL=±π/2.\nSuch an angular dependence has a simple physical ex-\nplanation. The electric current (9) is proportional to\nthe spin-current through the junction. The latter attains\nits maximum when the nonequilibrium spin polarizations\nin the ferromagnetic layers are oppositely directed and\nit vanishes if these polarizations are collinear and have\nequal magnitudes. The spin current obviously also van-\nishes if these polarizationsare perpendicular to the ferro-\nmagneticmagnetizationaxis, sinceperpendicularcompo-\nnents do not penetrate deep into ferromagnet. The spin\nflow through the junction is accompanied by energy dis-\nsipation. It is determined by the Ohmic losses in the\nferromagnet during transport of spin polarized electrons\nbetween the leads having spin dependent electrochemical\npotentials. The dissipative current of Eq. (9) is inde-\npendent of the superconducting phases φLandφR. We\nassume that the electric potentials of both contacts are\nequal. Iftheloadispresentinthecircuit, thespincurrent\nwill induce a voltage difference. The latter, in its turn,\ncan cause periodic oscillations of the Josephson current6.\nWhen/radicalbig\nDf/Bexis much shorter than the junction\nlengthLand the coherence length, the up and down-\nspin Fermi surfaces become decoupled. In this regime,\nthe supercurrent jsthrough the junction is determined\nby the decoupled tunneling of ±1 triplet Cooper pairs at\ntheir respective ferromagnet’s Fermi surfaces. Unlike the\ndissipative current, the spin-dependence of the electron\ndiffusion coefficients and conductivities is not so impor-\ntant, at least in the case when the Bex≪EF. Therefore,\nin the leading approximation we set Df↑=Df↓=Df,\nand a similar relation for the conductivities. Further-\nmore, in the linear approximation, only the first term of\nEq. (8) has to be taken into account. Moreover, since\nthe Josephson current is determined by part of the dis-\ntribution function that is odd in frequency, from Eqs.\n(8), (7) and (2) only the spin-independent part hf↑+hf↓\ncontributes to the current. It is given by\njs=σf\n16e/summationdisplay\nm=±1/integraldisplay\ndω/bracketleftbig\n(fr\nm∇xfr†\nm−∇xfr\nmfr†\nm)−\n(fr→fa)](hf↑+hf↓)\n2, (10)\nwherefr\nm(m=±1) are the retarded triplet compo-\nnents of the anomalous function in ferromagnet and\nfa\nm(ω) =fr\nm(−ω), while fr†\nm(ω) =fr∗\nm(−ω). When the\nspin-relaxation length is much larger than L, and within\nthe linearized approximation, fr\n±1obey1\nDf∇2\nxfm+2iωfm= 0, (11)\nwith the boundary conditions18rsfσf∇xfm|x=xR(L)=\n±fmR(L), where fmR(L)are given by Eqs. (5-6). Af-\nter transforming the integral in Eq. (10) into a sum overFIG. 2: (Color online) The Josephson current as a function\nof the spin potential δµmeasured in units of the unperturbed\nsuperconducting gap ∆ 0, atL=ζ, where ζ=/radicalbig\nDf/2|∆0|\nandI0= (2ζπkBT/er2\nsfσf)×10−2\nthe frequencies ωn=πkBT(2n+ 1),jscan be finally\nrepresented in the form\njs= sin(φL−φR)LsinθRsinθL\neσfR2\nsfK, (12)\nwhere\nK=|∆|2kBT/summationdisplay\nωn>0,ν=±11\nkνLsinh(kνL)×\n/bracketleftBigg\n1/radicalbig\n(ωn+iHν)2+|∆|2−1/radicalbig\n(ωn−iH−ν)2+|∆|2/bracketrightBigg2\n(13)\nandkν=/radicalbig\n2(ωn+iνδµ)/Df, withHν=ǫxc+νδµand\n2δµ=µ↑−µ↓.\nIV. DISCUSSION\nAs follows from Eq. (12), the Josephson current de-\npends on the directions of the nonequilibrium spin po-\nlarizations in the superconducting leads. The current\nreaches its maximum when the spin accumulations in the\nleads are perpendicular to the magnetization in the fer-\nromagnet, θR=θL=π/2. It reverses its sign when\nthe spin polarization in one of the leads flips its direc-\ntion. Therefore, in the setup shown in Fig. 1, the junc-\ntion can be switched into the π-state by simply revers-\ning the electric current through one of the FN contacts.\nIt should be noted that the dissipative current given by\nEq.(9) vanishes when the relative angles are such that5\nthe supercurrent reaches its maximum. Hence, the dissi-\npative transport can be turned off, a feature that can be\nimportant for practical purposes. Eqs. (12) and (13) also\nimply that the long-range proximity effect, described via\njs, vanishes when the exchange interaction ǫxc= 0. The\ndependence of jsonthe spin-potential δµisshownin Fig.\n2. A finite spin-potential causes variations of the order\nparameter ∆ and spin density entering in Eq. (12) which\nhave been found from a pair of self-consistent equations.\nIn our calculation of ∆ and S, we neglected the exchange\nfieldǫxc, assuming that ǫxc≪δµ. This is a realistic\nassumption, taking into account that S < δµN Fin Eq.\n(3) and|G|is considerably less than 1 in some supercon-\nducting metals (e.g. Al). In this limit, the dependence of\n∆ onδµis formally the same as in a thermally equilib-\nrium superconductor subject to a Zeeman splitting equal\ntoδµ. Such a scenario is well studied in the literature\n(see e.g.21). In Fig. 2, we see that the critical current\nchanges sign at some values of δµ. This is caused by\ninjection of nonequilibrium spins into the ferromagnetic\nlayer. As a result, the distribution function in Eq. 10 is\ndifferentfromtheequilibriumdistribution. Thereissome\nsimilarity of this effect with a current reversal observed\nin Josephson transistors20. At the lower temperature,\nthe supercurrent versus spin-potential is more peaked in\nthe range of higher δµ, since the spin density increases\nsharply together with ǫxcin this range. At even higher\nδµ, the superconductivity is destroyed by spin injection.\nThat causes a sudden drop of the current. We believethat this narrow range can be easily observed experi-\nmentally in the set up shown in Fig. 1, because δµcan\nbe fine tuned by varying the current through the normal\nleads.\nFig. 2 is calculated at L=/radicalbig\nDf/2|∆0|, that is the\ncharacteristic length of the ±1-triplet proximity effect\nin the range of temperatures considered. This length is\nobviously much larger than the s-wave Cooper pair pen-\netration depth/radicalbig\nDf/Bexand therefore clearly demon-\nstrates how the range of the proximity effect becomes\nmuch longer by spin injection into the superconducting\nleads.\nIn conclusion, spin injection into s-wave superconduc-\ntors can dramatically increase the stationary Josephson\ncurrent in SFS system. This enhancement is provided\nby±1 triplet components of the electron pairing func-\ntion. They are generated in superconducting leads by\nexchange fields that are noncollinear with the ferromag-\nnet magnetization. These fields, in turn, are induced by\nan injected spin polarization. Besides a strong effect on\nthe Josephson current, spin injection also gives rise to a\ndissipative current that at zero bias potential is induced\ndue to spin dependence of the ferromagnet conductivity.\nBoth Josephson and dissipative currents can be manipu-\nlated by varying the injected spin directions in the leads\nenabling control of π-junctions.\nAnatoly Mal’shukov gratefully acknowledges the hos-\npitality of NTNU.\n1F. S. Bergeret, A. F. Volkov, K. B. Efetov, Rev. Mod.\nPhys.77, 1321 (2005); A. I. Buzdin, Rev. Mod. Phys, 77,\n935 (2005)\n2A. K. Feofanov, et. al., Nature Phys. 6, 593 (2010); L.\nB. Ioffe, V. B. Geshkenbein, M. V. Feigel’man, A. L.\nFauch‘ere, and G. Blatter, Nature 398, 679 (1999); A. V.\nUstinov and V. K. Kaplunenko, J. Appl. Phys. 94, 5405\n(2003); T. Ortlepp et. al., Science 312, 1495 (2006)\n3F. S. Bergeret, A. F. Volkov, and K. B. Efetov, Phys. Rev.\nLett.86, 4096 (2001); A. Kadigrobov, R. I. Shekhter, and\nM. Jonson, Europhys. Lett. 54, 394 (2001).\n4M. Eschrig, J. Kopu, J. C. Cuevas, and G. Sch¨ on,\nPhys. Rev. Lett. 90, 137003 (2003); M. Eschrig and T.\nL¨ ofwander, Nature Phys. 4, 138 (2008).\n5M. Houzet, Phys. Rev. Lett. 101, 057009 (2008)\n6A. G. Mal’shukovand A.Brataas, Phys.Rev.B 82, 144511\n(2010);\n7A. M. Bobkov and I. V. Bobkova, Phys. Rev. B 84, 054533\n(2011); Phys. Rev. Lett. 108, 197002 (2012)\n8T. Yamashita, S.Takahashi, H.Imamura, andS.Maekawa,\nPhys. Rev. B 65, 172509 (2002); S. Takahashi and S.\nMaekawa, ibid. 67, 052409 (2003).\n9S. Takahashi,H. Imamura, and S. Maekawa, Phys. Rev.\nLett.82, 3911 (1999)\n10J. P. Morten, A. Brataas, and W. Belzig, Phys. Rev. B\n70, 212508 (2004); ibid72, 014510 (2005); N. Poli et. al.,Phys. Rev. Lett. 100, 136601 (2008).\n11M. Johnson, Appl. Phys. Lett. 65, 1460 (1994); P. S. Luo\net. al., Phys. Rev. B 79, 140508 (2009)\n12F. J. Jedema, H. B. Heersche, A. T. Filip, J. J. S. Basel-\nmans, and B. J. van Wees, Nature (London) 416, 713\n(2002); S. O. Valenzuela and M. Tinkham, Nature (Lon-\ndon)442, 176 (2006).\n13J. Rammer, H. Smith, Rev. Mod. Phys. 58, 323 (1985)\n14J. M. Serene, D. Rainer, Phys. Rep. 101, 221 (1983)\n15L.P. Pitaevskii, E.M. Lifshitz . Statistical Physics, Part 2.\nVol. 9. Butterworth-Heinemann (1980).\n16J. A. Leiro, Solid. State Comm., 93, 953 (1995)\n17L.D. Landau, E.M. Lifshitz . Quantum Mechanics: Non-\nRelativistic Theory. Vol. 3. Pergamon Press (1977)\n18M. Yu. Kupriyanov and V. F. Lukichev, Zh. Eksp. Teor.\nFiz.94, 139 (1988) [Sov. Phys. JETP 67, 1163 (1988)].\n19A. G. Aronov, Pis’ma Zh. Eksp. Teor. Fiz. 24, 37 (1976),\nJETP Lett. 24, 32 (1976); M. Johnson and R. H. Silsbee,\nPhys. Rev. Lett. 55, 1790 (1985).\n20L. N. Bulaevskii, V. V. Kuzii, and A. A. Sobyanin, Solid\nState Commun. 25, 1053 (1977); J. J. A. Baselmans, A.\nMorpurgo, B. J. van Wees, and T. M. Klapwijk, Nature\n397, 43 (1999).\n21Yu. A. Izyumov, Yu. N. Skryabin, Phys. Stat. Sol. (b) 61,\n9 (1974)" }, { "title": "0802.4009v1.Ferromagnetic_to_spin_glass_cross_over_in__La_Tb___2_3_Ca__1_3_MnO__3_.pdf", "content": "arXiv:0802.4009v1 [cond-mat.str-el] 27 Feb 2008Ferromagnetic to spin glass cross over in (La,Tb) 2/3Ca1/3MnO 3\nC. Raj Sankar, S. Vijayanand, and P. A. Joy∗\nPhysical and Materials Chemistry Division, National Chemi cal Laboratory, Pune 411008, India\nIn the series La 2/3−xTbxCa1/3MnO3, it is known that the compositions are ferromagnetic for\nsmaller values of xand show spin glass characteristics at larger values of x. Our studies on the\nmagnetic properties of variouscompositions intheLa 2/3−xTbxCa1/3MnO3series showthatthecross\nover from ferromagnetic to spin glass region takes place abo vex≈1/8. Also, a low temperature\nanomaly at 30 K, observed in the ac susceptibility curves, di sappears for compositions above this\ncritical value of x. A mixed phase region coexists in the narrow compositional r ange 0.1 ≤x≤\n0.125, indicating that the ferromagnetic to spin glass cros s over is not abrupt.\nI. INTRODUCTION\nThere has been a wide interest in the study of the sub-\nstituted perovskite-type manganites, La 1−xAxMnO3, in\norder to understand the different aspects of the complex\nmagnetic behaviour exhibited by these compounds.1,2,3\nThe double exchange interactions involving Mn3+and\nMn4+ions give rise to ferromagnetism in the substituted\nmanganites. The changes in the Mn-O-Mn bond angle,\nfrom structural distortions, are very crucial in determin-\ning the strength of the ferromagnetic interactions. High-\nest Curie temperature in the La 1−xCaxMnO3series is\nobtainedfor x≈1/3andhencetherearemanystudiesre-\nported forthe compositionLa 2/3A1/3MnO3. Manyinter-\nestingnewmagneticbehaviorsareobservedwhen La3+is\npartially substituted by other trivalent rare-earth ions in\nLa2/3Ca1/3MnO3,4,5,6,7,8,9though there are no changes\nin the Mn3/Mn4+ratio after substitution. Hwang et al.4\nfound a direct correlationbetween the Curie temperature\nand the average ionic radius of the La-site ions, where\nthe Curie temperature decreases with decreasing aver-\nage ionic radius in La 0.7−xRxCa0.3MnO3(R = Pr, Y),\nindicating the role of lattice effects in determining the\nferromagnetic properties. Partial substitution of La by\nsmaller ions leads to a decrease in the Mn-O-Mn bond\nangle and this affects the long range ferromagnetic ex-\nchange interactions.6,7,8\nFor La 0.67−xTbxCa0.33MnO3(in the re-\nports, the chemical composition is used as\n(La1−xTbx)2/3Ca1/3MnO3), a gradual decrease in\nthe Curie temperature is observed with increasing\nconcentration of Tb.7,9This is also associated with a\nbroadening of the magnetic transition and ultimately a\nspin glass behaviour is observed at larger concentrations\nof Tb.10,11In La 0.67−xTbxCa0.33MnO3, the evolution\nof a spin glass (SG) or a cluster glass (CG) state is\nthought to be due to the competing interactions of\nferromagnetic (FM) and antiferromagnetic (AFM) types\nor the random distribution of Mn-O-Mn bond angle\nwhich suppresses the exchange strength between the\nMn ions significantly.7,8,11At sufficiently large values\nofx(x= 0.22), short range ordered magnetic clusters\nare formed with typical magnetic coherence length of\naround 18 ˚A, at low temperatures.10\nThe reported studies on La 0.67−xTbxCa0.33MnO3areperformed on ferromagnetic compositions ( x≤0.1) or\nspin glass compositions ( x >0.15) and therefore no in-\nformation is available on the changeover region from fer-\nromagnetic to spin glass characteristics. In order to\nexplore the critical concentration of the Tb ions which\nresults in the formation of magnetic clusters leading\nto spin glass characteristics, a series of compositions\nwith very close values of xbetween 0.1 and 0.15 in\nLa0.67−xTbxCa0.33MnO3have been studied. The criti-\ncal concentration is found to be x≈1/8, and interesting\nmagnetic properties are observed for compositions close\nto the cross over region.\nII. EXPERIMENTAL\nThe polycrystalline La 0.67−xTbxCa0.33MnO3compo-\nsitions were synthesized by the conventional solid state\nroute from La 2O3, Tb4O7, CaCO 3and MnO 2by mix-\ning these oxides in the required stoichiometry for x= 0,\n0.03, 0.07, 0.10, 0.11, 0.12, 0.125, 0.13, 0.15, 0.20, and\n0.25. The well-mixed powders were initially heated at\n1273 K for 48 h, and subsequently at 1473 K for 48h,\n1573 K for 48 h and finally at 1623 K for 24h, with in-\ntermediate grindings at every 24 h steps to ensure the\nsample homogeneity. Finally the powder samples were\npelletized and sintered at 1673 K for 24 h. The samples\nwere characterized by powder X-ray diffraction using Cu\nKαradiation and Ni filter. The Mn4+contents in the\ncompounds were estimated by the iodometric titration\nmethod.12The magnetization measurements were per-\nformed on a vibrating sample magnetometer. Temper-\nature variation of the ac magnetic susceptibility of the\nsamples was measured by the mutual inductance method\nin a field of 2 Oe and at a frequency of 210 Hz.\nIII. RESULTS AND DISCUSSION\nThe samples were characterized for their phase purity\nby X-raydiffraction. All samples showedsinglephase na-\nture and the diffraction patterns were indexed on the dis-\ntorted orthorhombic structure with space group Pbnm.6\nThe lattice parameters were obtained by least squares\nrefinement of the diffraction patterns and found to be2\n/s48 /s49/s48/s48 /s50/s48/s48 /s51/s48/s48/s48/s46/s48/s48/s46/s53/s49/s46/s48/s32/s48/s46/s48\n/s32/s48/s46/s48/s51\n/s32/s48/s46/s48/s55\n/s32/s48/s46/s49/s48\n/s32/s48/s46/s49/s50/s53\n/s32/s48/s46/s49/s51/s51/s48/s32/s75/s47\n/s109/s97/s120\n/s84/s32/s40/s75/s41\nFIG. 1: Temperature variation of the ac susceptibility of so me\ncompositions in La 0.67−xTbxCa0.33MnO3. The numbers in-\ndicate the values of x.\ncomparable for those of the corresponding compositions\nreported previously.6,11The Mn4+content was found to\nbe matching with the expected value of 33% for all the\ncompositions.\nThe ac susceptibility curves of the unsubstituted and\nsome Tb-substituted compositions are shown in Fig. 1.\nWith increasing concentration of Tb, the magnetic tran-\nsition temperature is decreased, and the susceptibility\ncurves show a cusp-like feature when the concentration\nis increased beyond x= 0.1. The shapes and features of\nthe ac susceptibility curves of the Tb-substituted com-\npositions (for x≤0.1 and x >0.15) are similar to\nthat reported earlier.6,9,10Although well-defined mag-\nnetic transitions are observed for the samples with low\nconcentrations of Tb, there is a distinct anomalous fea-\nture in the susceptibility curves of these compositions at\nlow-temperatures. A decrease in the susceptibility is ob-\nserved below ∼30 K, as soon as a small amount of Tb is\nincorporated ( x= 0.03).\nAs shown in Fig. 2, for x= 0.1, a small step in the\nsusceptibility curve is observed below the main magnetic\ntransition. For x= 0.11, there is a cusp like feature ob-\nserved at a higher temperature which is succeeded by the\nnormal flat curve, and finally a drop in the susceptibility\nbelow 30 K. This is a complex feature and is not com-\nmonly reported for the manganite compositions. As the\nconcentration of Tb ions is further increased, it can be\nseen that the cusp-like feature becomes more prominent,\nthe flat region gradually vanishes, but a dip in suscep-\ntibility is still observed at 30 K for the compositions up\ntox= 0.125. For x >0.125 there is only a cusp-like\nfeature in the ac susceptibility curve as observed for spin\nglasssystems. The composition with x=0.33is reported\nto show spin glass behavior.7Neutron diffraction studies\nhave indicated ferromagnetically ordered state for x=\n0.067 at low temperatures, whereas short range ordered\nclusters are found for x= 0.2 (the authors have used the/s48 /s51/s48 /s54/s48 /s57/s48 /s49/s50/s48 /s49/s53/s48/s48/s46/s48/s48/s46/s53/s49/s46/s48/s47\n/s109/s97/s120\n/s84/s32/s40/s75/s41/s32/s48/s46/s49\n/s32/s48/s46/s49/s49\n/s32/s48/s46/s49/s50\n/s32/s48/s46/s49/s50/s53\n/s32/s48/s46/s49/s51\n/s32/s48/s46/s49/s53\n/s32/s48/s46/s50/s48\nFIG. 2: Temperature dependence of the ac susceptibility of\nthe compositions in La 0.67−xTbxCa0.33MnO3, for 0.1 ≤x≤\n0.2. The numbers indicate the values of x.\nchemicalformulaas(La 1−xTbx)2/3Ca1/3MnO3, sothatx\n= 0.1 and 0.3, respectively, correspond to x= 0.067 and\n0.2 in La 0.67−xTbxCa0.33MnO3).11These clusters show\nspin glass-type of behavior due to competition between\nFM and AFM types of exchanges.6,9\nForthe compositions with x >0.125, thereis a gradual\nshift of the cusp towards lower temperatures as the value\nofxis increased. Also, it was found that the relative\nsusceptibility value alsodecreasedwith the increasingTb\nconcentration. These observations point out the evolu-\ntion of short range ordering due to the decreasing Mn-\nO-Mn bond angle and consequent formation of magnetic\nclusters as the concentration of Tb is increased. This\nclustering is likely to be due to the breaking or weaken-\ning of the double exchange closer to the Tb-sites caused\nby the local structural distortion and the spin glass like\nfeature originates from the formation of such clusters.\nThe decreasing temperature corresponding to the cusp,\nidentified as T ghenceforth, suggests the confinement of\nmagnetic clusters to shorter length scales as the number\nof Tb centers is increased (increasing values of x).\nAs shown in Fig. 3, the ferromagnetic transition tem-\nperature decreases almost linearly with increasing x, up\ntox= 0.125. For the compositions showing both fer-\nromagnetic and spin glass characteristics (0.1 ≤x≤\n0.125),TcandTgare determined as illustrated in the\ninset of Fig. 3 for x= 0.11. Here, the χ-T data for x=\n0.15 (spin glass composition) is shifted towards the right\nside and that of x= 0.07 (ferromagnetic composition) is\nshifted towards the left side along the x-axis, to match\nwith the observed data of the mixed phase composition\nand the curves are normalized with respect to the maxi-\nmum values. Tcis taken as the mid point ofthe magnetic\ntransition. T gchanges abruptly around x= 0.125 and\nvaries from 66 K for x= 0.13 to 43 K for x= 0.25.Tg\nhas been found to be almost independent (40-50 K) of x\nfor higher Tb concentrations.6Thus, the lowest value of3\n/s48/s46/s48 /s48/s46/s49 /s48/s46/s50 /s48/s46/s51/s48/s53/s48/s49/s48/s48/s49/s53/s48/s50/s48/s48/s50/s53/s48/s51/s48/s48\n/s53/s48 /s49/s48/s48 /s49/s53/s48\n/s32/s84\n/s67\n/s32/s84\n/s103/s84/s32/s40/s75/s41\n/s120 /s32/s105/s110/s32/s76/s97\n/s48/s46/s54/s55/s45/s120/s84/s98\n/s120/s67/s97\n/s48/s46/s51/s51/s77/s110/s79\n/s51/s32/s120 /s32/s61/s32/s48/s46/s49/s49\n/s32/s120 /s32/s61/s32/s48/s46/s48/s55\n/s32/s120 /s32/s61/s32/s48/s46/s49/s53/s84\n/s103\n/s84\n/s99/s32\n/s32/s84/s32/s40/s75/s41\nFIG. 3: Variation of TcandTgas a function of xin\nLa0.67−xTbxCa0.33MnO3. Inset: illustration of the method\nof extracting TcandTg, forx= 0.11, using the data for x=\n0.15 (for Tg) andx= 0.07 (for Tc).\npossible Tgis larger than the temperature (30 K) where\na decrease in the ac susceptibility is observed for 0 < x≤\n0.125.\nThe decrease in the susceptibility below 30 K as soon\nas a small amount of Tb is incorporated in the lat-\ntice of La 2/3Ca1/3MnO3has been ascribed to spin glass\nbehavior.9However, neutron diffraction studies showed\nferromagnetic ordering for x= 0.067 down to 7 K.11\nThus, the feature at 30 K for 0.03 ≤x≤0.125 is not\nlikely to be spin glass transition. It is possible that at\nsmall concentrations the Tb3+ions are randomly dis-\ntributed in the lattice, the double exchange is disturbed\naround the Tb centers and therefore tiny magnetic clus-\ntersareformedwithreducedMn-O-Mnangle. Thesetiny\nmagnetic clusters remain isolated until x= 0.125 (1/8)\nabove which larger magnetic clusters are formed due to\nthe breaking or considerable weakening of the three di-\nmensional long range ordering. Thus, the temperature\nat which a decrease in the susceptibility is observed, due\nto these small clusters, remains the same until x= 1/8.\nThe three dimensional ordering is affected when x >1/8.\nHowever, there is another possibility that, at small con-\ncentrations, the Tb ions form a TbMnO 3like local envi-\nronment in the lattice of La 0.67−xTbxCa0.33MnO3. For\nTbMnO 3, anincommensurate-commensuratephasetran-\nsition, which is accompanied by a ferroelectric transition,\nassociated with a lattice modulation, is observed close to\n∼30 K and large magnetic field controlled polarization\neffects are reported at this temperature.13,14\nA local phase separation exists in the x= 0.1 sample,\nas evidenced by a small step- like magnetic transition, in-\ndicating that ferromagnetic clusters are started forming\nat this value of x. It is possible that, above this value of\nx, some magnetic clusters with short range ordering are\nseparated whose size decreases with increasing x. Thus,\nat intermediate values of x, the lattice is consisting of/s45/s49/s53 /s45/s49/s48 /s45/s53 /s48 /s53 /s49/s48 /s49/s53/s45/s57/s48/s45/s54/s48/s45/s51/s48/s48/s51/s48/s54/s48/s57/s48\n/s48/s46/s48/s48 /s48/s46/s48/s54 /s48/s46/s49/s50 /s48/s46/s49/s56 /s48/s46/s50/s52/s52/s48/s54/s48/s56/s48/s49/s48/s48\n/s65/s116/s32/s49/s50/s32/s75/s77/s32/s40/s101/s109/s117/s47/s103/s41\n/s72/s32/s40/s107/s79/s101/s41/s32/s120 /s32/s61/s32/s48/s46/s48\n/s32/s120 /s32/s61/s48/s46/s49/s48\n/s32/s120 /s32/s61/s48/s46/s49/s51\n/s32/s120 /s32/s61/s32/s48/s46/s49/s53\n/s32/s120 /s32/s61/s32/s48/s46/s50/s48\n/s65/s116/s32/s49/s53/s32/s107/s79/s101/s32/s77/s32/s40/s101/s109/s117/s47/s103/s41\n/s120 /s32/s105/s110/s32/s76/s97\n/s50/s47/s51/s45/s120/s84/s98\n/s120/s67/s97\n/s49/s47/s51/s77/s110/s79\n/s51\nFIG. 4: The M-H curves of La 0.67−xTbxCa0.33MnO3for dif-\nferent values of x. Inset: variation of magnetization at 15 kOe\nas a function of x.\nlarger ferromagnetic clusters with sufficiently long range\nordering and smaller short range ordered clusters. Also,\nfor the largerlong range orderedpart, the magnetic tran-\nsition temperature decreases due to the decrease in the\nMn-O-Mn bond angle. Thus, the clustering may be seen\nto start when x= 0.11, where a cusp is also observed\nalong with the normal magnetic features.\nFig. 4 shows the dc magnetization curves of different\ncompositions, measured at 12 K, up to a maximum field\nof 15 kOe. The magnetization is saturated above 10\nkOe for x≤0.1. The variation of the magnetization\nat the maximum measured field of 15 kOe, as a function\nofxin La2/3−xTbxCa1/33MnO3is shown in the inset\nof Fig. 4. The magnetization remains almost the same\nup tox= 0.1 and then decreases above this value of\nx. Also, the magnetization is not saturated for x >\n0.1 in the measured field range. Recent studies sug-\ngest the existence of quantum critical point (QCP) ef-\nfect in La 0.67Ca0.33Mn1−xGaxO3.15QCP is defined as\na second order transition accompanied by the change of\na non-thermal parameter. The observation of QCP in\nLa0.67Ca0.33Mn1−xGaxO3system is as predicted by the\ntheoretical calculations and the QCP in this system was\nexpected for a value of 10-20% of Ga substitution.16The\nsubstitution of a nonmagnetic ion like Ga at the Mn-\nsublattice of the perovskite-type oxide causes the local-\nization of the electronic states suppressing the double ex-\nchange mechanism. For La 0.67Ca0.33Mn1−xGaxO3, the\nspontaneous magnetic moment calculated from the ex-\nperimental neutron diffraction patterns recorded at 1.5\nK decreased for x >0.1 and vanished for x >0.16, in-\ndicating electron localization which suppresses the dou-\nble exchange mechanism. The present dc magnetization\ndata on La 2/3−xTbxCa1/33MnO3shows almost a similar\ntrend, except for the finite value of the magnetization at\nhigher values of x, suggesting the possible existence of\nQCP in this system also. However, this needs to be ver-4\n/s48 /s49 /s50 /s51/s48/s50/s48/s52/s48/s54/s48\n/s120 /s32/s61/s32/s48/s46/s49/s50/s53/s77/s32/s40/s101/s109/s117/s47/s103/s41\n/s48 /s53 /s49/s48 /s49/s53/s48/s50/s48/s52/s48/s54/s48/s56/s48\n/s120 /s32/s61/s32/s48/s46/s49/s51\n/s32/s32\n/s48 /s53 /s49/s48 /s49/s53/s48/s50/s48/s52/s48/s54/s48\n/s120 /s32/s61/s32/s48/s46/s49/s53/s77/s32/s40/s101/s109/s117/s47/s103/s41\n/s72/s32/s40/s107/s79/s101/s41/s48 /s53 /s49/s48 /s49/s53/s48/s50/s48/s52/s48\n/s120 /s32/s61/s32/s48/s46/s50/s48/s32\n/s72/s32/s40/s107/s79/s101/s41\nFIG. 5: The virgin magnetization (red curves) and part of the\nhysteresis loop (blue curves) of x= 0.125, 0.13, 0.15, and 0.2\ninLa0.67−xTbxCa0.33MnO3. Notethatthe Hscale is different\nforx= 0.125.\nified with the help of neutron diffraction measurements,\nas made in the case of the Ga substituted system.\nAnother interesting behavior observed in the low tem-\nperature M−Hmeasurements for the compositions im-\nmediately above x= 1/8 is an irreversible jump to a fer-\nromagnetic state at higher magnetic fields, as shown in\nFig. 5. This is observedonly for the virgin magnetization\nmeasurements. Up to x= 0.125, a normal feature is ob-\nserved, where the virgin magnetization curve lies inside\nthe hysteresis loop. For x= 0.2, the behaviour is similar\nto that observed for some typical spin glass systems,12\nwhere the virgin magnetization curve initially lies out-\nside the loop and then merges with the loop at higher\nfields. On the other hand, for x= 0.13 and 0.15, the en-\ntire virgin magnetization curve lies outside the hysteresis\nloop above a certain small field (this small crossing field\nis observed for x= 0.2 and 0.25 also, and increases with\nx), and an anomalous step-like feature is observed in the\nvirginmagnetizationcurve,similartothat ofametamag-\nnetic transition. However, this transition is completelyirreversible. The field above which a broad step is ob-\nserved is larger for x= 0.15 compared to that for x=\n0.13. After the magnetic field is increased in the negative\ndirection to -15 kOe and when brought back to +15 kOe\nthrough H = 0, the transition is not observed. This is\nan irreversible ferromagnetic transition in the sense that\nthe step-like feature is never obtained when the measure-\nmentswererepeatedimmediatelyorevenafteratimegap\nof 30 minutes. In the subsequent measurements, the first\npart of the curve always lies inside the hysteresis loop,\nlike that for the compositions for x≤0.125. It appears\nthat an irreversible magnetic field induced phase transi-\ntion is occurred. Similar characteristicswerereported for\nsomePr-basedsubstituted manganitecompositions. Dho\nand Hur explained this behavior in terms of the reorien-\ntation ofthe Mn spins pinned by localizedPrmoments,17\nwhereas Woodward et al.18explained the observations in\nterms of an avalanche behavior. It may be noted that\nthe first part of the virgin curve is similar to that of the\nspin glass composition x= 0.2, in the present case, and\ntherefore, it is possible that the second jump is due to a\nfield induced growth of the larger ferromagnetic clusters.\nOnce the clusters are grown, it is not possible to revert\nback to the original state due to the unavailability of suf-\nficient thermal energy. The original state is found only\nwhen the temperature is raised above the peak tempera-\nture and then cooled back in zero field.\nIV. CONCLUSIONS\nThe present studies made on a series of close compo-\nsitions in La 0.67−xTbxCa0.33MnO3indicate that single\nphase ferromagnetic compositions are possible for x <\n0.1, mixed long range ordered ferromagnetic and short\nrange ordered magnetic clusters coexist for 0.1 ≤x≤\n0.125 and spin glass like phases are formed for larger val-\nues ofx. The ferromagnetic clusters present in the com-\npositions immediately above the cross over region show\nmagneticfield induced growthand givelargermagnetiza-\ntionathigherfields. Furtherdetailedstudiesarerequired\nto understand the complex magnetic behavior shown by\nthese Tb substituted manganite compositions at the in-\ntermediate and the cross over regions.\n∗Electronic address: pa.joy@ncl.res.in\n1Y. Tokura, Colossal Magnetoresistive Oxides (Gordon and\nBreach, Singapore, 2000).\n2E. Dagotto, Nanoscale Phase Separation and Colossal\nMagneto-Resistance (Springer Verlag, Berlin, 2003).\n3B. I. Belevtsev, Low Temp. Phys. 30, 421 (2004).\n4H. Y. Hwang, S-W Cheong, P. G. Radaelli, M. Marezio,\nand B. Batlogg, Phys. Rev. Lett. 75, 914 (1995).\n5A. Maignan, C. Simon, V. Caignaert, and B. Raveau, J.\nAppl. Phys. 79, 7891 (1996).6J. Blasco, J. Garcia, J. M. de Teresa, M. R. Ibarra, P. A.\nAlgarabel, and C. Marquina, J. Phys.: Condens. Matter\n8, 7427 (1996).\n7J. M. De Teresa, M. R. Ibarra, J. Garcia, J. Blasco, C.\nRitter, P. A. Algarabel, C. Marquina, and A. del Moral,\nPhys. Rev. Lett. 76, 3392 (1996).\n8T. Terai, T. Kakeshita, T. Fukuda, T. Saburi, N.\nTakamoto, K. Kindo, and M. Honda, Phys. Rev. B 58,\n14908 (1998).\n9J. C. Nie, J. H. Wang, and B. R. Zhao, J. Magn. Magn.5\nMater.192, L379 (1999).\n10J. M. De Teresa, C. Ritter, M. R. Ibarra, P. A. Algarabel,\nJ. L.Garcia-Munoz, J.Blasco, J. Garcia, andC. Marquina,\nPhys. Rev. B 56, 3317 (1997).\n11M. Watahiki, N. Metoki, J. Suzuki, K. Oikawa, J. Nie, M.\nTachiki, and Y. Yamada, J. Phys. Soc. Japan 70, 1090\n(2001).\n12C. R. Sankar and P. A. Joy, Phys. Rev. B 72, 132407\n(2005).\n13T. Kimura, T. Goto, H. Shintani, K. Ishizaka, T. Arima,\nand Y. Tokura, Nature 426, 55 (2003).\n14T. Goto, T. Kimura, G. Lawes, A. P. Ramirez, and Y.Tokura Phys. Rev. Lett. 92, 257201 (2004).\n15J. M. De Teresa, P. A. Algarabel, C. Ritter, J. Blasco, M.\nR. Ibarra, L. Morellon, J. I. Espeso, and J. C. Gomez-Sal,\nPhys. Rev. Lett 94, 207205 (2005).\n16J. L. Alonso, L. A. Fernandez, F. Guinea, V. Laliena, and\nV. Martin-Mayor, Phys. Rev. B 66, 104430 (2002).\n17J. Dho and N. H. Hur, Phys. Rev. B 67, 214414 (2003).\n18F. M. Woodward, J. W. Lynn, M. B. Stone, R. Mahendi-\nran, P. Schiffer, J. F. Mitchell, D. N. Argyriou, and L. C.\nChapon, Phys. Rev. B 70, 174433 (2004)." }, { "title": "2202.01611v1.On_the_Divergence_of_the_Ferromagnetic_Susceptibility_in_the_SU_N__Nagaoka_Thouless_Ferromagnet.pdf", "content": "arXiv:2202.01611v1 [cond-mat.str-el] 3 Feb 2022On the Divergence of the Ferromagnetic Susceptibility in th e SU(N)\nNagaoka-Thouless Ferromagnet\nRajiv R. P. Singh1and Jaan Oitmaa2\n1Department of Physics, University of California Davis, CA 9 5616, USA\n2School of Physics, The University of New South Wales, Sydney 2052, Australia\n(Dated: February 4, 2022)\nUsing finite temperature strong coupling expansions for the SU(N) Hubbard Model, we calculate\nthe thermodynamic properties of the model in the infinite- Ulimit for arbitrary density 0 ≤ρ≤1\nand allN. We express the ferromagnetic susceptibility of the model a s a Curie term plus a ∆ χ,\nan excess susceptibility above the Curie-behavior. We show that, on a bipartite lattice, graph by\ngraph the contributions to ∆ χare non-negative in the limit that the hole density δ= 1−ρgoes\nto zero. By summing the contributions from all graphs consis ting of closed loops we find that the\nlow hole-density ferromagnetic susceptibility diverges e xponentially as exp∆ /TasT→0 in two\nand higher dimensions. This demonstrates that Nagaoka-Tho uless ferromagnetic state exists as a\nthermodynamicstateofmatteratlowenoughdensityofholes andsufficientlylowtemperatures. The\nconstant ∆ scales with the SU(N) parameter Nas 1/Nimplying that ferromagnetism is gradually\nweakened with increasing Nas the characteristic temperature scale for ferromagnetic order goes\ndown.\nINTRODUCTION\nThe Hubbard model [1–3] is a central model for de-\nscribing the behavior of electrons in solid state systems\nand has had a huge impact in our understanding of con-\ndensed matter physics [4–6]. Strong correlations, arising\nfrom the on-site repulsion in the Hubbard model, can\nbe used to understand many basic solid state phenom-\nena including metal-insulator transitions, antiferromag-\nnetism, superconductivity, spin-liquids and itinerant fer-\nromagnetism.\nNagaoka-ThoulessFerromagnetismisaclassicproblem\nin itinerant magnetism [7]. Nagaoka [8] and Thouless\n[9] independently showed that when the Hubbard repul-\nsionUis large enough, a single hole introduced into a\nsystem with one-particle per site, polarizes the system\naround the hole. There have been several variational\nand numerical studies [10–14] of Nagaoka-Thoulessferro-\nmagnetism, especially in the ground state of the system.\nAtfinite temperatures, rigorousmathematicalarguments\nhave been made to show that magnetization in a field ex-\nceeds the pure paramagnetic value [15, 16] and Dynami-\ncal Mean-Field Theory (DMFT) [17] was used to obtain\na phase diagram in the density-temperature plane.\nThe cold atomic gases in optical lattices provide a new\nmotivation for study of the Hubbard model [18–23]. In\nthese systems, it is possible to build an ensemble that is\nwell described by the Hubbard model and where the mi-\ncroscopic parameters such as Uandtcan be controlled\nand a priori well understood. Furthermore, cold atomic\ngases allow one to change the number of fermion species\nfrom two to a larger Nand thus study the Hubbard\nmodel with SU(N) symmetry [24–29] for different values\nofN.\nFinite temperature strong-couplingexpansion is a nat-\nural way to address the magnetic behavior of the Hub-bard model at finite temperatures, at various hole densi-\nties, in the thermodynamic limit [30–32]. These expan-\nsionscanbedevelopedinthegrandcanonicalensembleat\nfixed fugacity ζ= expβµin powers of βt,w= exp−βU,\nand1/βU. After changingvariablesfrom fugacitytopar-\nticle density ρ, one can obtain temperature dependent\nthermodynamic properties at various densities. For Uof\norder or larger than the bandwidth they allow one to re-\nlate the thermodynamics of the Hubbard model at low\ntemperatures to a generalized Heisenberg or t-J model\n[33–36]. The expansions simplify in the limit U→ ∞,\nin which case many terms can be set to zero and can\nbe used to study the problem of Nagaoka-Thouless fer-\nromagnetism.\nThe first few terms of the expansion suffice to give\nan accurate numerical description of the thermodynamic\nproperties of the model at temperatures larger than the\nhopping parameter t. And, as shown previously for the\nSU(2) t-J model [37–39], series extrapolationmethods al-\nlowoneto goto muchlowertemperatures. But, anumer-\nical extrapolation is difficult to control reliably down to\nT= 0. Here, we are interested in the entire temperature\nrange 0< T <∞. We show that, close to half filling, i.e.\nin the limit δ= 1−ρgoing to zero, the thermodynamic\nuniform magnetic susceptibility can be computed all the\nway toT→0 by summing over the loop graphs in each\norder of perturbation theory. This calculation provides\na lower bound for the susceptibility and leads to a func-\ntion which diverges exponentially to infinity as exp∆ /T\nas the temperature goes to zero. This shows that, for\nlarge enough U, the Nagaoka-Thoulessferromagnetic be-\nhavior is a thermodynamic phenomena at low density of\nholes and low enough temperatures. These results are\ntrue for any N >1 of the SU(N) models [40] and in any\ndimension greater than one. However, the constant ∆\nscales as 1 /N, that is the temperature scale for the tran-2\nsition goes down as Nincreases. For the SU(2) case, our\nresults are in agreement with DMFT which also found\nthat the transition temperature goes to zero as δ→0\n[17].\nMODEL AND METHODS\nTheSU(N)HubbardmodelisdefinedbyaHamiltonian\nH=H0+V, where the unperturbed Hamiltonian H0is\nan on-site term:\nH0=U/summationdisplay\nini(ni−1)\n2−µ/summationdisplay\nini−h/summationdisplay\ni(n1i−ni\nN),(1)\nwithnithe total number operator for particles on site i\nandµis the chemicalpotential. The last term his a spin-\npolarizing field that lowers the energy when the particle\nis in the first spin state n1i= 1 and raises it for all other\nstates and has an overall zero trace. The perturbation V\nis the hopping term:\nV=−t/summationdisplay\nN/summationdisplay\nα=1(C†\ni,αCj,α+h.c.), (2)\nwhere the sum < i,j > runs over nearest-neighbor pairs\nof sites of a lattice and the sum over αruns over the\nNspecies of Fermions. The total number of fermions\nof each species is a constant of motion. Thus both the\nchemical potential and field terms commute with the rest\nof the Hamiltonian.\nUsing the formalism of thermodynamic perturbation\ntheory [30, 31],the logarithm of the grand partition func-\ntion, per site, can be expended as\nlnZ=lnz+∞/summationdisplay\nr=1/integraldisplayβ\n0dτ1/integraldisplayτ1\n0dτ2.../integraldisplayτr−1\n0dτr\n<˜V(τ1)...˜V(τr)>N(3)\nwherezis the single-site partition function,\n˜V=eτH0Ve−τH0, (4)\nand,\n< X >= Tre−βH0X/Tre−βH0. (5)\nIn each order, the terms in the expansion can be ex-\npressed in terms of various graphs on the lattice as:\nlnZ= lnz+/summationdisplay\nGLGz−Ns(βt)NbXG,(6)\nIntheexpression,thegraph GhasNssitesand Nbbonds.\nLGis the lattice constant of the graph defined as the\nextensive part of the graph count, per lattice site. The\nweight-factor XGisthereducedcontributionofthegraphobtained from an evaluation of the traces which depends\nonβt,βU, fugacity ζ, fieldhandN.\nIn theU→ ∞limit, no double occupancy is allowed\nand the weight-factor for a graph with Ns-sites reduces\nto\nXG=Ns−1/summationdisplay\nn=1xG\nnζn. (7)\nHerexG\nnis a polynomial in the SU(N) parameter Nof\nordern.\nFrom the partition function, the particle density (per\nsite) can be obtained via the relation\nρ=ζ∂\n∂ζlnZ. (8)\nThermodynamic quantities such as Internal energy per\nsite,e, and entropy per site, s, are obtained using the\nrelations\ne=−(∂\n∂βlnZ)ζ, (9)\nand\ns=−β(∂\n∂βlnZ)ζ−ρlnζ+lnZ. (10)\nThe ferromagnetic susceptibility per site is defined by\nthe second derivative of ln Zwith respect to the spin-\npolarizing field h. It is given by\nχ=1\nβ∂2\n∂h2lnZ (11)\nThe field term in the Hamiltonian is defined solely for\ncalculating the susceptibility. Otherwise, we will restrict\nall calculations to h= 0.\nSINGLE-SITE TERM AND SERIES EXPANSIONS\nIn thelimit of U→ ∞the single-sitepartitionfunction\nto order h2becomes z=z0+h2z1, where,\nz0= 1+Nζ (12)\nand\nz1=β2ζ\n2N−1\nN(13)\nFor all calculations other than the susceptibility, we can\nseth= 0. In zeroth order the particle density is given\nby\nρ0=Nζ\n1+Nζ. (14)3\nThe zeroth order susceptibility per site is given by\nχ=2\nβz1\nz0= (N−1\nN2)βρ. (15)\nThis is a Curie law as no double occupancy means we\nhave local moments at all temperatures.\nWe will define excess susceptibility, over and above the\nCurie-law as\nχ=Cρ\nT+∆χ, (16)\nwith Curie constant Cequal toN−1\nN2. Note that in this\nequation the density is the full density not the bare den-\nsity obtained in zeroth order. Our goal is to calculate\n∆χ.\nIn our studies, we will restrict ourselves to bipartite\nlattices. All the graphs that contribute to the zero-field\npartition function on a bipartite lattice to eighth order\ntogether with their weights for arbitrary Nare given in\nSupplementary materials.\nFIG. 1: Two classes of graphs that contribute to the suscepti -\nbility expansion. (a) Tree graphs, with noclosed loops. Eve ry\nbond must be doubled in order to have a non-zero trace con-\ntribution. In the single hole sector, each fermion moves bac k\nandforthastheholes movesaroundthegraph. Everyspincan\nbe independently of any spin species. Such graphs only con-\ntribute to the Curie-law and do not contribute to the excess\nsusceptibility at all. (b) Graphs consisting of closed loop s. In\nthe single hole sector, as the hole traverses the loop, each s pin\nmoves to its neighboring position. In order to have a non-zer o\ntrace contribution, all fermions must be of the same species .\nThus these graphs have a maximum relative contribution to\nthe excess susceptibility.\nNear one-particle per site all properties can be ex-\npanded in powers of the hole density δ0, given by\nδ0= 1−ρ0=1\n1+Nζ=1\nz0, (17)\nFor the Nagaoka-Thouless problem, we are interested in\nthelimit δ= 1−ρgoingtozero. Thus, wewillkeepterms\nlinear in δ0and drop all terms proportional to higher\npowers of δ0. These linear in δ0terms come from exactlyone hole in each cluster. We should note that this does\nnot mean we are looking at a single hole in the ther-\nmodynamic system. Our formalism implies that we are\nstudying the limit of low hole density as similar behavior\nwill be happening independently all over the system.\nIn the large Ulimit, the weight of a graph XGis a\npolynomial in ζof orderNs−1 , where Nsis number of\nsites in the cluster. The restriction to lowest power of δ0\nreduces the weight factor for a graph to:\nXG=xGζNs−1. (18)\nThe coefficients xG, which depend on Nand the field h,\nturn out to be always positive as can be seen from the\nexplicit calculations to eighth orderin the supplementary\nmaterials. These terms correspond either to a single hole\nmovingback and forth on a tree like graphwith no closed\nloops or a single hole moving in closed loops. In both\ncases they are positive. We will see that this will lead us\ntothe resultthat the contribution toexcesssusceptibility\nfrom every graph is non-negative. This means that even\na partial summation of graphs is a lower bound on the\nexcess susceptibility.\nIn this limit, the relation between the full density func-\ntion and fugacity becomes\nρ=ρ0+/summationdisplay\nGLG(βt)Nb(Ns−1−Nsρ0)xGζNs−1\nzNs\n0(19)\nThe excess susceptibility is given by\n∆χ=/summationdisplay\nGLG(βt)Nb(C1G−C2G),(20)\nwhere\nC1G=1\nβ∂2\n∂h2(XG\nzNs), (21)\nand\nC2G= (Ns−1−Nsρ0)C\nTXG\nzNs. (22)\nFor tree graphs, with no closed loops (see Fig. 1), the\ncoefficient xGin zerofieldisproportionalto NNs−1. This\nreflects the fact that in the absence of closed loops ev-\nery spin can independently be of any species. For these\ngraphs the contribution to excess susceptibility vanishes\nidentically. Physically this is a reflection of the fact that\nsusceptibility of independent spins is already contained\nin the Curie law. Thus, we only need to consider those\nweights where the power of Nin a graph with Nssites\nis less than Ns−1. In all these terms at least some of\nthe spins are constrained to be of the same species. Even\nthe smallest such constraint can be shown to lead to a\npositive contribution to the susceptibility.\nAt the other extreme are those terms where zero-field\nxGscales linearly with N. This implies that every spin4\nin the graph must be of the same species to contribute to\na non-zero trace. An example is a graph consisting of a\nsingleclosedloop(SeeFig.1). Itmusthavethisbehavior.\nAs a hole traversesaroundthe loop, everyFermion in the\nloop movesto its neighboringposition and hence must be\nof the same species as its neighbor to contribute to the\ntrace. These graphs contribute maximally to the excess\nsusceptibility. It can be shown that for a single loop of\nlengthlthe zero-field weight-factor is\nXG=2l\nl!N ζl−1(23)\nThe excess susceptibility contribution from this graph\ncan be shown to be\nC\nT(βt)l\nl!2l(l−1)(l−2)Nζl−1\nzl\n0. (24)\nExpressing this in terms of ρ0gives\nC\nT(βt)l\nl!2l(l−1)(l−2)ρl−1\n0δ0\nNNs−2. (25)\nIt is well known that for large l, the number of polygons\nof even length lembedded in a bipartite lattice scale as\n[41]\npl=Aµl\npla−3, (26)\nwhere the constant µpcalled the connectivity constant\nis known approximately for most lattices [42]. Ignoring\nthe weak dependence on the exponent awhich will only\naffect the prefactor, the contributions of polygons can be\nsummed to obtain an excess susceptibility of\n∆χ∝N2δ0\nρ0C\nTexp∆\nT, (27)\nwith ∆ =tρ0µp\nN.Ignoring the slowly varying prefactor,\nthis shows that the excess susceptibility diverges expo-\nnentially as T→0. We believe the primary role of the\nadditional terms not included in this summation is to\ndecorate these graphs and renormalize the bare density\nρ0to the full density ρ.\nThisresultprovidesalowerboundtothemagneticsus-\nceptibility and implies that the Nagaoka-Thouless ferro-\nmagnet is a thermodynamic phase of matter for low hole\ndensity and low enough temperatures. The characteristic\ntemperaturescaleatwhichthesusceptibilitybecomesex-\nponentially large is inversely proportional to the SU(N)\nparameter N. Thus, the tendency for ferromagnetism\ngradually weakens with increase in N. These results are\nin agreement with the earlier dynamical mean-field the-\nory results for the SU(2) case in that the ferromagnetic\nphase boundary was found to go to zero temperature\nas the hole density goes to zero [17]. They are also in\nagreement with mathematical arguments that show the\nexistence of finite magnetization in a field that exceeds\nthe paramagnetic value at any temperature [16].The extension of these results to finite U/tand finite\nhole doping can be done numerically as was done for\nthe SU(2) t-J models some time ago [37–39, 43]. Those\nstudies show that at small enough J/tand close to half\nfilling the peak in the magnetic susceptibility shifts to\nq= 0. However, from a small number of terms in the\nexpansion it is more difficult to rigorously establish the\ndivergence of the susceptibility.\nDISCUSSIONS AND CONCLUSIONS\nIn this paper we have used finite temperature strong\ncoupling expansions for the Hubbard model to revisit the\nproblem of Nagaoka-Thouless ferromagnetism at large U\nandsmall holedopingnearone-particleper site. We have\nshown that the ferromagnetic susceptibility of the sys-\ntem diverges exponentially as exp∆ /TasT→0. Thus\nat sufficiently low temperatures, the system must turn\nferromagnetic.\nWhile the Hubbard model is an approximatemodel for\nsolid state systems, it can be a well characterized and ac-\ncuratemodel in coldatomic gasesin opticallattices. Fur-\nthermore, in these systems, the Hubbard parameters can\nbe tuned by lasers and the number of fermion species can\nbe made larger than 2 and the system can have SU(N)\nsymmetry. We have shown that Nagaoka-Thouless fer-\nromagnetism is present for all Nand is only weakened\ngradually with increase in N. Fundamentally, this ferro-\nmagnetism arises from the fact that as a hole traverses\na closed loop, non-zero trace arises only if all the spins\nbelong to the same species. This combined with positive\ntrace on loops of bipartite lattices implies an exponen-\ntially divergent susceptibility as T→0.\nWe hope our work would stimulate further experi-\nmental search for Nagaoka-Thouless ferromagnetism and\nmeasurements of the temperature dependence of the sus-\nceptibility at low hole densities in solid state and cold\natom systems.\nAcknowledgement: This work is supported in part\nby the US National Science Foundation grant DMR-\n1855111.\n[1]J. Hubbard, Proc. R. Soc. A 276, 238–257 (1963).\n[2]J. Kanamori. Prog. Theor. Phys., 30:275, 1963.\n[3]M. C. Gutzwiller. Phy. Rev. Lett., 10:159, 1963.\n[4]D. C. Mattis, The Theory of Magnetism Made Simple\n(World Scientific, 2006).\n[5]A. Auerbach, Interacting Electrons and Quantum Mag-\nnetism(Springer, 1994).\n[6]E. H. Lieb, The Hubbard model: Some Rigorous Results\nand Open Problems. In Condensed Matter Physics and\nExactly Soluble Models , pages 59–77. Springer Berlin Hei-\ndelberg, (2004).5\n[7]For a historical review see H. Tasaki, J. Stat. Phys. 84,\n535-653 (1996).\n[8]Y. Nagaoka Phys. Rev. 147, 392 (1966).\n[9]D. J. Thouless, Planet. Space Sci. 86, 893–904 (1965).\n[10]B. S. Shastry, H. R. Krishnamurthy, and P. W. Anderson\nPhys. Rev. B 41, 2375 (1990).\n[11]F. Becca and S. Sorella Phys. Rev. Lett. 86, 3396 (2001).\n[12]M. Kollar, R. Strack, and D. Vollhardt Phys. Rev. B 53,\n9225 (1996).\n[13]J. A. Riera and A. P. Young Phys. Rev. B 40, 5285(R)\n(1989); A. Barbieri, J. A. Riera, and A. P. Young Phys.\nRev. B 41, 11697(R) (1990).\n[14]L. Liu, H. Yao, E. Berg, S. R. White, and S. A. Kivelson\nPhys. Rev. Lett. 108, 126406 (2012).\n[15]M. Aizenman and E. H. Lieb, Phys. Rev. Lett.,\n65(12):1470–1473 (1990).\n[16]T. Miyao, arXiv:2112.14475 (2021).\n[17]H. Park, K. Haule, C. A. Marianetti, and G. Kotliar,\nPhys. Rev. B 77, 035107 (2008).\n[18]A. Mazurenko, C. S. Chiu, G. Ji, M. F. Parsons, M.\nKan´ asz-Nagy, R. Schmidt, F. Grusdt, E. Demler, D.\nGreif and M. Greiner Nature 545, 462–466 (2017).\n[19]P. T. Brown, D. Mitra, E. G.-Sanchez, R. Nourafkan, A.\nReymbaut, C.-D. H´ ebert, S. Bergeron, A. -M. S. Trem-\nblay, J. Kokalj, D. A. Huse, P. Schauss, W. S. Bakr,\nScience, 363, 379-382 (2019).\n[20]D. Mitra, P. T. Brown, E. G.-Sanchez, S. S. Kondov,\nT. Devakul, D. A. Huse, P. Schauss, W. S. Bakr Nature\nPhysics 14, 173-177 (2017).\n[21]L. W. Cheuk, M. A. Nichols, K. R. Lawrence, M. Okan,\nH. Zhang, E. Khatami, N. Trivedi, T. Paiva, M. Rigol,\nM. W. Zwierlein\n[22]U. Schneider, L. Hackerm¨ uller, J. P. Ronzheimer, S. Will,\nS. Braun, T. Best, I. Bloch, E. Demler, S. Mandt, D.\nRasch and A. Rosch. Nature Physics 8, 213–218 (2012).\n[23]J. Koepsell, J. Vijayan, P. Sompet, F. Grusdt, T. A.\nHilker, E. Demler, G. Salomon, I. Bloch and C. Gross\nNature 572, 358–362 (2019).\n[24]C. Hofrichter, L. Riegger, F. Scazza, M. H¨ ofer, D. R. Fer-\nnandes, I. Bloch, and S. F¨ olling Phys. Rev. X 6, 021030\n(2016).\n[25]C. Honerkamp and W. Hofstetter Phys. Rev. Lett. 92,\n170403 (2004).\n[26]S. Taie, E. I.-G.-Padilla, N. Nishizawa, Y. Takasu, Y.Kuno, H.-T. Wei, R. T. Scalettar, K. R. A. Hazzard, Y.\nTakahashi, Cond-mat arXiv:2010.07730 .\n[27]D. Tusi, L. Franchi, L. F. Livi, K. Baumann, D. B.\nOrenes, L. Del Re, R. E. Barfknecht, T. Zhou, M. In-\nguscio, G. Cappellini, M. Capone, J. Catani, L. Fallani,\narXiv:2104.13338 (2021).\n[28]L. Del Re and M. Capone, Phys. Rev. A 98, 063628\n(2018);\n[29]E. I.-G.-Padilla, S. Dasgupta, H.-T. Wei, S. Taie, Y.\nTakahashi, R. T. Scalettar, and K. R. A. Hazzard Phys.\nRev. A 104, 043316 (2021).\n[30]J. Oitmaa, C. Hamer and W. Zheng, Series Expansion\nMethods for Strongly Interacting Lattice Models , Cam-\nbridge University Press, Cambridge, UK, (2006).\n[31]J. A. Henderson, J. Oitmaa, and M. C. B. Ashley Phys.\nRev. B 46, 6328 (1992).\n[32]D.F.B. ten Haaf and J.M.J. van Leeuwen, Phys. Rev. B\n46, 6313 (1992).\n[33]A. H. MacDonald, S. M. Girvin, and D. Yoshioka, Phys.\nRev. B 37, 9753 (1988).\n[34]J.-Y.P.Delannoy, M.J.P.Gingras, P.C.W.Holdsworth,\nand A.-M. S. Tremblay, Phys. Rev. B 72, 115114 (2005).\n[35]H.-Y. Yang, A. M. L¨ auchli, F. Mila, and K. P. Schmidt\nPhys. Rev. Lett. 105, 267204 (2010).\n[36]R. R. P. Singh and J. Oitmaa, Cond-mat\narXiv:2201.06677 (2022).\n[37]W. O. Putikka, M. U. Luchini, and T. M. Rice Phys.\nRev. Lett. 68, 538 (1992); W. O. Putikka, M. U. Luchini,\nand M. Ogata Phys. Rev. Lett. 69, 2288 (1992); W. O.\nPutikka, M. U. Luchini, and R. R. P. Singh Phys. Rev.\nLett. 81, 2966 (1998).\n[38]R. R.P. SinghandR.L. Glenister Phys.Rev.B 46, 11871\n(1992); R. R. P. Singh and R. L. Glenister Phys. Rev. B\n46, 14313(R) (1992).\n[39]L. P. Pryadko, S. A. Kivelson, and Oron Zachar Phys.\nRev. Lett. 92, 067002 (2004).\n[40]H. Katsura and A. Tanaka, Phys. Rev. A 87, 013617\n(2013).\n[41]J. A. Hammersley, Mathematical Proceedings of the\nCambridge Philosophical Society, 57, 516 (1961).\n[42]For a review see A. J. Guttmann, Cond-mat\narXiv:1212.3448 (2012).\n[43]J. Yedidia, Phys. Rev. B 41 9397 (1990)." }, { "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" }]