LOW TEMPERATURE VOLUME 31, NUMBERS 8–9 AUGUST–SEPTEMBER 2005

On seven decades of antiferromagnetism ͓DOI: 10.1063/1.2008126

In tracing the history of the discovery of the cooperative distance between magnetic ions and assumed that the inter- ordering of spins in an antiparallel orientation of action between magnetic ions belonging to the same layer is their moments in the substances and the use of the term positive ͑ferromagnetic͒ while that between ions of different ‘‘antiferromagnetism’’ to denote this new magnetic state of layers is negative, i.e., promoting an antiparallel orientation in physical discourse, one would invariably start from of the magnetic moments. He introduced two magnetic sub- the experiments that raised the necessity of altering the mo- systems formed by the ions of neighboring layers with op- lecular field concept introduced by Weiss in 1906 for the positely directed magnetic moments, and the vectors now description of . These were experiments in- known as the ferromagnetic and antiferromagnetic vectors. vestigating the temperature dependence of the magnetic sus- Landau also took into account the interaction of the magnetic ceptibilities of the metals Mn, Cr, the alloys Co–Fe, Fe–Mn, moments of the ions with the lattice, which causes magnetic Fe–Cr, and the chlorides FeCl2 , CrCl3 , and NiCl2 .Asit anisotropy. The minimum of the energy in the model corre- would turn out, these substances are by no means the sim- sponds to an antiparallel orientation of the magnetic mo- plest of antiferromagnets, and it is possibly for that reason ments of the layers. The model explained the magnetic that the path to antiferromagnetism was not the easiest one. anomalies observed and predicted spontaneous cooperative For some time this new magnetic state was called anomalous magnetic ordering with zero resultant of ferromagnetism. the whole crystal but with nonzero magnetic moments of its In 1932 L. Ne´el, striking off from Heisenberg’s notions subsystems; such ordering should be accompanied by char- of the , deepened Weiss’s concept of the acteristic features on the depend dependence of the heat ca- molecular field by introducing the concept of a ‘‘local mo- pacity and . Landau obtained expres- lecular field.’’ He allowed for the possibility of this field sions for the temperature dependence of the magnetic being oriented in the opposite direction to that of the mag- susceptibility in all crystallographic directions ͓L. Landau, netic moment created by the neighboring ion. That made it Phys. Zs. Sowjet. 4, 675 ͑1933͔͒. possible to explain the temperature behavior of the magnetic In a parallel effort an experiment designed by L. Shub- susceptibility of PtCo alloys ͓L. Ne´el, Ann. Phys. ͑Paris͒ 17, nikov was carried out to detect the anomalies of the tempera- 5 ͑1932͔͒ and the anomalously large temperature- ture dependence of the heat capacity that should accompany ͓ independent susceptibility of L. Ne´el, J. Phys. et cooperative magnetic ordering in FeCl3 at low temperatures. Radium ͑Paris͒ 3, 160 ͑1932͔͒. Subsequently Ne´el intro- The experiment was brilliantly completed in 1934. The ob- duced the concept of interpenetrating magnetic sublattices served peak ͓O. N. Trapeznikowa and L. W. Shubnikov, Na- into the theory of the ‘‘generalized ferromagnet’’ ͑later called ture 134, 378 ͑1934͔͒ attested to the cooperative character of ‘‘antiferromagnet’’͒. The thermodynamic theory developed the spontaneous antiparallel ordering of the magnetic mo- by Ne´el made it possible to perceive the cooperative charac- ments and confirmed the conclusions of the theory that had ter of the antiferromagnetic ordering and to predict a peak on been developed at that time. Subsequently these same the temperature dependence of the susceptibility upon the anomalies of the heat capacity were found in all of the chlo- spontaneous onset of antiferromagnetic order ͓L. Ne´el, Ann. rides displaying anomalous magnetic behavior ͓O. N. Tra- Phys. ͑Paris͒ 5, 232 ͑1936͔͒. This peak was soon observed, peznikova, L. V. Shubnikov, G. A. Milyutin, Zh. E´ ksp. Teor. and the temperature of the transition to the new magnetic Fiz. 6, 421 ͑1936͔͒. Soon after, the expected features on the state has come to be called the Ne´el temperature. temperature dependence of their magnetic susceptibilities During this same period of years, L. V. Shubnikov and were also observed ͓L. W. Schubnikow and S. S. Schalyt, L. D. Landau in Kharkov were intrigued by the seemingly Phys. Z. Sowjet. 11, 566 ͑1937͒; Zh. E´ ksp. Teor. Fiz. 8, 518 contradictory magnetic properties of layered anhydrous crys- ͑1938͔͒. tals of the chlorides of iron, , nickel, and chromium. In The years 1932–1936 can be considered to be the date of the paramagnetic temperature region the magnetic suscepti- birth of antiferromagnetism. That is the period in which the bility of these salts increased with decreasing temperature as observation of the magnetic and thermal properties expected in ordinary ferromagnets—faster than required by the Curie- to accompany the formation of a magnetic Weiss law—but no ferromagnetism arose on further cooling with an antiparallel orientation of the magnetic moments was below the paramagnetic , and the suscepti- achieved. This state did not yet have a name. This was con- bility itself became dependent on the magnetic field strength ferred later, in 1937. In his Nobel Lecture, Ne´el would credit ͓H. R. Woltjer, Leiden Comm. N173b ͑1926͒; H. R. Woltjer F. Bitter with the introduction of the term ‘‘antiferromag- and Kamerlingh Onnes, Leiden Comm. N173c ͑1926͒;H.R. netism’’ ͓F. Bitter, Phys. Rep. 54,79–86͑1938͔͒. However, Woltjer and E. C. Wiersma, Leiden Comm. N201a ͑1930͔͒. the term can also be found in an earlier paper ͓J. H. van In 1933 Landau developed a phenomenological model of Vleck, Phys. Rev. 52, 1178–1198 ͑1937͔͒. a layered wherein he took into account the possibility Attempts at a quantum description of the phenomenon of of a strong dependence of the exchange interaction on the antiferromagnetism had not been so successful, since a strict

1063-777X/2005/31(8–9)/2/$26.00633 © 2005 American Institute of Physics 634 Low Temp. Phys. 31 (8–9), August–September 2005 N. F. Kharchenko

FIG. 1. Number of papers per year containing the term ‘‘antiferromagnetic’’ FIG. 2. Time dependence of the ratio of the number of papers per year in the title or abstract. The search was done using the search engine Google containing the term ‘‘antiferromagnetic’’ to the number of those containing Scholar. The dotted line shows the analogous result for the term ‘‘ferromag- the term ‘‘ferromagnetic.’’ netic.’’ The time interval over which the number of papers increases by a factor of e ͑the straight line͒ is 11.2 yr for ‘‘antiferromagnetic’’ and 12.5 yr for ‘‘ferromagnetic.’’ structures, frustrated antiferromagnetic structures, molecular cluster antiferromagnets, nuclear antiferromagnets, and, fi- antiparallel orientation of the spins of magnetic sublattices nally, artificial heteromagnetic nanostructures. We hope that did not correspond to the ground state of the quantum sys- the papers published in this issue will give an idea of the tems studied. This situation led to some skepticism as to different fields of antiferromagnetics research being pursued whether that picture was correct. Only the neutron- at various research centers of Europe, Asia, and America. diffraction experiments in 1949 ͓C. G. Shull and J. S. Smart, The pace of research on antiferromagnetism since the Phys. Rev. 76, 1256 ͑1949͔͒, in which additional diffraction time of its discovery is demonstrated in Figs. 1 and 2. Al- peaks were observed to appear upon the transition of the though the data presented are far from complete, they do MnO crystal to the magnetically ordered state, made it pos- reflect more accurately the evolutionary trends of such re- sible to assert that the antiferromagnetic ordering of the search. One might think that this temperature dependence moments on the macroscopic scale actually occurs. The mag- could be altered noticeably by recent developments such as netic structure of MnO was described beautifully by the the wide application of artificial antiferro/ferromagnetic model of two interpenetrating magnetic sublattices. The sta- structures for data storage and readout devices in information bility of such structures in time was later confirmed indi- technology and the growing interest in the role of antiferro- rectly and then directly, by the observation of stable 180- in high-temperature , but in re- degree ͑time-reversed͒ antiferromagnetic domains. ality the exponent characterizing the growth of interest varies Subsequent experimental and theoretical research on an- in direct proportion to time and fluctuates little. It is also tiferromagnets revealed many new structures and properties. curious that over the entire history of antiferromagnetism the It soon became clear that collinear antiferromagnets with ratio of the number of papers on antiferromagnets to the completely compensated magnetic moments represent only number of papers on ferromagnets is steadily growing. one of many groups of antiferromagnets. The great diversity In closing, we note that many famous names in physics of states united by the property of nearly complete compen- are linked with research on antiferromagnets, but the discov- sation of all their elementary magnetic moments is evident ery of antiferromagnetism as a physical phenomenon is as- merely from a list of the terms that have been introduced to sociated with the names of those who were first: L. Ne´el, L. identify the different classes of antiferromagnetic substances: D. Landau, and L. V. Shubnikov. There is no doubt that the collinear or noncollinear antiferromagnets, many-lattice non- physical concepts introduced by them, which have since collinear antiferromagnets, metamagnets, antiferromagnets come to be called antiferromagnetism, Ne´el temperature, with weak ferromagnetism, the set of modulated antiferro- magnetic sublattice, and the antiferromagnetic vector, will magnetic structures commensurate and incommensurate with endure for a long time to come. the period of the unit cell of the crystal, antiferromagnetic glasses, dimeric spin structures, antiferromagnetic chains and ladders, gap antiferromagnets, antiferromagnetic vortex N. F. Kharchenko LOW TEMPERATURE PHYSICS VOLUME 31, NUMBERS 8–9 AUGUST–SEPTEMBER 2005

Mesoscopic antiferromagnets: statics, dynamics, and quantum tunneling „Review… B. A. Ivanov*

Institute of Magnetism of the National Academy of Sciences of Ukraine, pr. Vernadskogo 36b, Kiev 03142, Ukraine; Taras Shevchenko National University, pr. Glushkova 2, Kiev 03127, Ukraine ͑Submitted March 4, 2005͒ Fiz. Nizk. Temp. 31, 841–884 ͑August–September 2005͒ The static and dynamic, quantum and classical properties of antiferromagnets ͑AFMs͒ are discussed from a unified point of view. Attention is directed mainly toward mesoscopic , i.e., materials with characteristic scales of nonuniformities of the order of atomic dimensions. The creation of such materials and their study and application have largely shaped the face of the physics of our day. This class includes small magnetic particles and their arrays, magnetic superlattices and clusters, and high-spin . The traditional problems of the physics of antiferromagnetism are also discussed ͑symmetry analysis of AFMs, reorientation transitions, equations of spin dynamics͒, but they are represented only to the extent that it is useful to do so for subsequent consideration of the quantum and classical properties of mesoscopic AFMs. For description of the spin dynamics of AFMs, a magnetic Lagrangian of a form matched with the quantum-mechanical Hamiltonian is constructed. The lowering of the symmetry of the AFMs, both that due to conventional causes such as an external magnetic field and that due to the partial decompensation of the sublattice spins, is taken into account. The latter effect is especially important for mesoscopic particles of AFMs of the ferrite type. The influence of defects and of the surface on the reorientation transitions in AFMs is discussed in detail. These effects, which are of fundamental importance for the description of small particles of AFMs, are observed for magnetic superlattices with an antiferromagnetic interaction of the elements of the superlattices. The description of macroscopic quantum effects in mesoscopic AFMs plays a prominent role. The spin Lagrangian obtained describes new tunneling effects such as an oscillatory dependence of the tunneling probability on magnetic field. Quantum effects in magnetic systems with a nonuniform ground state are investigated. These effects can be described as the change due to processes of tunneling of the topological charges of various natures that characterize these states. © 2005 American Institute of Physics. ͓DOI: 10.1063/1.2008127͔

I. INTRODUCTION physics,3–9 have played a large role in the study of the sub- ject of this review: the manifestations of antiferromagnetism The study and practical application of mesoscopic mate- in the properties of mesoscopic magnets. rials with characteristic length scales of the nonuniformities In the past seven decades the study of antiferromag- from micron to atomic dimensions started at the boundary of netism has comprised a significant part of the fundamental the twentieth and twenty-first centuries. The class of mesos- physics of magnetism.10,11 This is due not only to the fea- copic magnets includes materials with different spatial tures of ‘‘purely magnetic’’ properties of these materials, scales, such as small magnetic particles of micron and sub- among which one might mention the broad spectrum of micron dimensions and ordered arrays of such particles, magnetic superlattices, magnetic clusters, and, finally, high- spontaneous and field-induced order-disorder transi- ͑ 5,12͒ spin molecules, including tens of spins with a strong ex- tions see monographs and interesting resonance proper- 1,2 ties of AFMs, primarily the ‘‘exchange enhancement’’ of the change interaction. Mesoscopic magnets can be divided, in 7 terms of the character of their ordering ͑we are talking about resonance frequencies. No less important are the manifesta- the focus of the spins both in an individual small particle and tions of antiferromagnetism in such ‘‘nonmagnetic’’ proper- 8,13 13 the ordering of the magnetic moments of individual elements ties of materials such as optical, galvanomagnetic, and 13,14 of a superlattice͒, into the same classes as bulk materials acoustic. We know of no examples of the application of ͑ferromagnets, AFMs, ferrimagnets͒. These materials often conventional crystalline AFMs as functional materials for manifest unique physical properties that are absent in bulk solid-state electronics, although their many unique param- samples. It suffices to mention the macroscopic quantum ef- eters, e.g., the presence of magnetic resonance frequencies in 7 fects manifested in the coherent quantum behavior of tens, the submillimeter region and the enormous domain-wall ve- hundreds, and even thousands of spins. locities ͑tens of kilometers per second͒,6,15 look very prom- The physics of mesoscopic magnets owes its progress to ising for applications in modern functional electronics. In the the achievements of the traditional physics of magnetically past decade, however, novel objects that might be classified ordered crystals. Undoubtedly the significant results from the as ‘‘artificial AFMs’’ have arrived in the physics of magne- study of ‘‘ordinary’’ AFMs, elaborated and explained in the tism. Among these are one-dimensional ͑1D͒ superstructures many monographs devoted wholly or largely to this field of of the multilayer film type and two-dimensional ͑2D͒ super-

1063-777X/2005/31(8–9)/33/$26.00635 © 2005 American Institute of Physics 636 Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov structures comprising lattices of magnetic elements of sub- for the order parameter, our understanding of which has been micron scale on a nonmagnetic substrate. Such materials are furthered substantially in recent years, are represented to the promising for the development of new devices for high- extent that it is useful to do so for the subsequent discussion density magnetic recording,16–18 magnetic field sensors and of mesoscopic AFMs. This review does not claim to give a magnetic heads,19,20 and logic elements for computers.21 The complete description of all the phenomena mentioned above. interaction of the structural elements can be antiferromag- However, the author hopes that the discussion of the static netic in both 1D19,22 and 2D23 superlattices.23 Antiferromag- and dynamic ͑including nonlinear͒ quantum and classical netic order of the structural elements of a superlattice and the properties of AFMs, though brief, will be useful to the change of this order under the influence of magnetic field reader. ͑analogous to spin-flop or spin-flip transitions in crystalline 24–26 AFMs; see Sec. 4͒ determine its giant II. ANTIFERROMAGNETICS AT THE TURN OF THE magnetoresistance,19,20 a property of practical importance. CENTURY The main feature specific to the physics of mesoscopic At the present time many different types of spin ordering magnets is the presence in them of macroscopic coherent of condensed media are known, both crystalline and quantum tunneling effects, which were observed more than amorphous.10,11 From time immemorial man has made prac- 10 years ago.1,27–30 These effects are of interest because of tical use of magnets having nonzero mean values of the spin the importance, from a general physical standpoint, of mani- density ͗S͘ or magnetization M; in the simplest case festations of quantum effects in the macro world. Mesos- MϭϪg␮ ͗S͘, where g is the Lande factor (gϷ2 for ions copic objects, which have quantum-mechanical properties, B in the s state͒, and ␮ is the modulus of the Bohr magneton. are of interest for their potential application as elements of B 31 In the language of the theory of symmetry this means that quantum computers. Furthermore, a number of subtle and such materials ͑including not only pure ferromagnets but beautiful effects, e.g., the suppression of tunneling transi- also a large class of ferrimagnets, some spiral magnets, etc.͒ tions on account of interference of the instanton 32,33 are characterized by spontaneous breaking of the symmetry trajectories. The possibility of controlling tunneling ef- with respect to time inversion t→Ϫt, whereupon both the fects ͑turning the tunneling on or off͒ is important for the use 31 mean value of the spin of an individual ion and also the of mesoscopic magnets as elements of quantum computers. value of the magnetization changes sign. Antiferromagnets Analysis of macroscopic quantum tunneling effects re- are a fundamentally different class of materials. In them the quires the use of a number of complicated methods of mod- symmetry with respect to time inversion is spontaneously ern quantum physics of magnets and quantum field theory. broken, but the spontaneous magnetization is equal to zero. However, the exposition becomes much more compact and The phenomenon of antiferromagnetism was discovered over simple if results from the classical physics of magnets are 70 years ago, and a theoretical explanation for it was given at used. Comparison of the classical and quantum pictures of that time in the papers by Ne´el39,40 and Landau.41 magnets is made in Sec. 3. On the basis of that discussion, in The situation with complete compensation of the mag- Sec. 4 we investigate reorientation transitions in magnets, netization is presented most simply by assuming that the both in the framework of the conventional approach, which crystal lattice of the AFMs contains a finite number n of is adequate for describing bulk AFMs, and for mesoscopic magnetic sublattices, each of which has a nonzero magneti- magnets. In Sec. 5 we turn to an analysis of the semiclassical zation M␣ , but these magnetizations compensate each other dynamics of AFMs in a formulation that is closest to the so that the total magnetization of the AFMs in the ground conventional theory of the dynamical properties of magnets ϭ͚n ϭ state is equal to zero, M ␣ϭ1M␣ 0. The simplest ex- but which permits a consistent analysis of purely quantum ample, to which we shall limit discussion, is that of a two- effects. A detailed discussion of the problems of tunneling in sublattice AFM in the ground state of which the magnetiza- small particles of AFMs on this basis is presented in Sec. 6. tions M1 and M2 are equal in magnitude and antiparallel. In Finally, in Sec. 7 we discuss the question of the topological such a definition it is understood that the sublattices must spin nonuniformities in mesoscopic AFMs and, on the basis necessarily be crystallographically equivalent, i.e., there ex- of non-one-dimensional instantons, describe the tunneling ef- ists an element of the crystal symmetry group ͑the symmetry fects for such nonuniform states of AFMs. group of the paraphase͒ that takes one sublattice into the For reasons of space an author must always select those other. One of the most important contains reached after many areas that are of greatest interest at the given time. Naturally, years of study of the phenomenon of antiferromagnetism and even though one tries to be objective, such a selection cannot the ‘‘nonmagnetic’’ effects associated with it lies in the ex- avoid subjectivity. Here we do not touch upon questions of treme importance of symmetry analysis of AFMs. It is be- the dynamics of nonlinear excitations ͑solitons͒ in low- cause of symmetry restrictions that there can be a strictly dimensional crystalline AFMs ͑see reviews34–38͒. In this pa- zero value of the magnetization Mϭ0 in the face of a non- per attention is devoted mainly to the properties specific to zero mean value of the spins, i.e., exact compensation of the mesoscopic AFMs, both quantum ͑macroscopic quantum magnetizations of the individual spins in a finite region of a tunneling, destructive interference effects͒ and purely classi- medium. A symmetry description of magnets is based on the cal, deriving from the special role of the surface and of de- use of either magnetic4,9,42 or exchange43 symmetry groups. fects, and also to those physical phenomena in which the A symmetry operation that does not permute the magnetic aforementioned features of mesoscopic AFMs are manifested sublattices will be called even, following Turov,4 while one jointly. The traditional problems, e.g., the character of reori- that does permute will be called odd, and their signs will be entation transitions and the form of the dynamical equations denoted as ͑ϩ͒ or ͑Ϫ͒, respectively. A criterion of antiferro- Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov 637 magnetism is the presence of at least one odd symmetry el- Fe3ϩ with spin Sϭ5/2, coupled by an antiferromagnetic in- ement of the paraphase. This condition distinguishes AFMs teraction and ordered in an almost ideal of ferrimagnets ␣ from , the sublattices of which are inequivalent, a typical crystalline AFM— -Fe2O3 . However, the although their total magnetization can go to zero at a com- uncompensated magnetic moment of the ferritin particle is pensation point. This definition of antiferromagnetism ͑and ␮ not small: it has a value of around 200 B , i.e., of the order not the condition that the spontaneous magnetization equals of 1% of the maximum possible value. Thus, from the stand- ͒ zero is now standard. point of analysis of the magnetization, ferritin is a ferrimag- The magnetization of AFMs can be nonzero in the pres- net close to the compensation point, and that fundamentally ence of an external field H 0 without eliminating the spon- alters its dynamic properties, both classical and ͑especially͒ taneous symmetry breaking. Here it is appropriate to com- quantum; see Secs. 5 and 6. On the other hand, for the ma- pare the behavior of the different Heisenberg magnets, the jority of the magnetic , lying in the central part of such Hamiltonian of which contains only a purely isotropic ex- a particle, the standard description in the language of crystal change interaction of the form magnetic groups is applicable. In recent years, new objects that can be classified as Hˆ ͑Heis͒ϭJSˆ Sˆ , ͑1͒ 1• 2 mesoscopic magnets have appeared. Their characteristic Ͼ ˆ scales are still smaller than for the ferritin particles men- where J is the exchange integral, J 0 for AFMs, and S1,2 are spin operators. In a Heisenberg ferromagnet ͑FM͒,ifone tioned above but are larger than atomic scales. These are the disregards the anisotropic interactions of a relativistic nature, so-called high-spin molecules, containing tens of spins with the magnetization M is parallel to the field H, i.e., there is no an interaction close to the isotropic Heisenberg interaction ͑ ͒ ͑ spontaneous symmetry breaking for H 0. For an isotropic 1 and rather strong the exchange integral J has a value up ͒ 1,46 AFM the magnetization caused by the field is also nonzero, to 100 K ; see reviews. Such molecules are also known to Mϭ␹H, where ␹ is the susceptibility of the AFM. However, exhibit manifestations of the Dzyaloshinskii-Moriya interac- contrary to the FM case, a difference of the sublattice mag- tion. High-spin molecules typically have a regular distribu- netizations and, hence, antiferromagnetic order exists over a tion of magnetic ions, forming a quite definite magnetic ͉ ͉Ͻ structure. Essentially, the synthesis and study of high-spin wide region of values of the field H Hex , where Hex is the exchange field. These same relationships on the whole re- molecules, and especially the creation of high-quality single main valid when the relativistic interactions are taken into crystals of such molecules with good orientation of the mag- account ͑if they are not too strong͒: magnetic order exists in netic axes of the molecules in the crystal, have opened up a a ferromagnet only when the field H is oriented along certain new era in the study of the macroscopic quantum effects in ͑ selected symmetry axes, while in an AFM the ordering in the magnets. An important class of such molecules and a fun- ͒ presence of anisotropy remains unbroken as before at fields damentally new object in the physics of antiferromagnetism Ͻ is that of the so-called spin rings ͑the term ferric wheels is H Hex . The criterion Mϭ0 for defining antiferromagnetism is also used͒. In such molecules an odd number of magnetic not exhaustive, because in certain crystalline AFMs with the ions ͑the best-studied are molecules with 6, 8, and 10 iron 3ϩ relativistic interactions taken into account the compensation ions Fe with spin Sϭ5/2) form a closed ring with an of the magnetizations is not complete, even in the absence of antiferromagnetic interaction of the nearest neighbors (J 48–52 external magnetic field, and a weak spontaneous moment ϳ15– 30 K) and a completely compensated total spin. appears. For a two-sublattice AFMs the incomplete compen- In high-spin molecules proper ͑molecular magnets͒ the inter- sation of the magnetizations ͑the physics of weak ferromag- action is mainly ferromagnetic, and their total magnetic mo- ␮ 53 netism͒ can be described as a manifestation of noncollinear- ment can reach a value of 26 B . Some high-spin mol- ity, or canting, of the magnetizations M1 and M2 . Such ecules should be classified as molecular ferrimagnets. They AFMs are called weak ferromagnets or canted AFMs. The consist of two inequivalent groups of spins with a ferromag- theory of weak ferromagnetism was constructed by netic interaction within a group and an antiparallel orienta- Dzyaloshinskii44 on the basis of a phenomenological ap- tion of the spins of the two groups ͑analogous to sublattices͒. proach ͑see the recent monograph by Turov et al.9 for de- This class of objects includes the best-studied molecular 3ϩ tails͒, and a microscopic mechanism was proposed by magnet Mn12-ac, in which 8 manganese ions Mn with S 45 ϭ ϭ 48 Moriya. 3/2 give a total spin Stot 10. High-spin molecules with Application of the usual concepts of the physics of crys- half-integer spin are also known; among them are Mn4 com- 54,55 talline AFMs for the description of mesoscopic AFMs can plexes with spin 9/2. For Mn12-ac the decompensation is require a certain modification. For atomic clusters containing not small, but for the V16 with 15 vanadium atoms hundreds or thousands of spins with an antiferromagnetic with spin Sϭ1/2 the total spin is equal to 1/2.56 From the interaction, the role of the surface becomes extremely impor- point of view of classical symmetry analysis the high-spin tant. Even for an atomically smooth surface the number of molecules should be described by finite classes of magnetic particles in the sublattices can differ. As a consequence, an symmetry, but, unlike the known crystalline magnetic uncompensated total spin of the AFM particle arises. As an classes, here such symmetry elements as Cn , Dn ,Sn axes example, let us consider ferritin particles, which play an im- with nϭ5ornϾ6 can also be present. An example is the 3ϩ portant role in the vital activity of warm-blooded animals. Mn72Fe30 molecule, in which the 30 iron ions Fe , coupled These particles are used as a model object for experimental by an antiferromagnetic exchange interaction, form an icosa- study of the properties of AFM particles.27 The magnetic part dodecahedral structure with 5- and 3-fold axes.57 of a ferritin particle consists of approximately 4500 iron ions Finally, let mention the ‘‘antiferromagnetic’’ aspects of 638 Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov the physics of purely man-made magnetic materials ͑mag- Here the first three terms, in which the summation is over netic superstructures͒ created with the use of modern nano- nearest-neighbor pairs ͗␣␤͘ and each pair is counted once, technologies. This class includes multilayer systems consist- describe bilinear interactions of the spins. They determine ing of layers of ferromagnetic metals a few atoms thick, three types of interactions: the standard isotropic exchange separated by spacer layers of nonmagnetic or antiferromag- ͓the Heisenberg exchange of Eq. ͑1͔͒ with an exchange inte- netic metals.19 For such 1D superstructures the interaction of gral J; the antisymmetric Dzyaloshinskii–Moriya the magnetic elements ͑the individual layers͒ is determined interaction,44,45 which is written in terms of the dual vector d by the exchange across the spacers and can be ferromagnetic associated to the antisymmetric part of Jij ; and, the symmet- or antiferromagnetic; for these structures biquadratic ex- ric anisotropic interaction. This last is often called the inter- 19,22 change effects are also known. ion anisotropy. The Hamiltonian Hˆ also includes the single- In recent years, 2D superstructures composed of lattices ion anisotropy with the constant K, which can exist only for of submicron magnetic particles on a nonmagnetic substrate site spins Sу1. The last term describes the Zeeman interac- have also been widely studied. Depending on their shape, tion of the spins with the external magnetic field. Here and these particles are called magnetic dots, strips, rings, or below it is assumed for convenience and transparency that wires. Most often they are made of magnetically soft mate- the spin vector operator S␣ is parallel to the magnetization. 17 rials such as Co, Fe, and Permalloy. The interaction of We shall mainly consider the case of weakly anisotropic individual particles in such a system is determined by the magnets, for which ␬,KӶJ, and the value of the exchange- magnetic dipole interaction of their magnetic moments, i.e., relativistic constant dϳͱ(␬,K)JӶJ. 58 they are a pure implementation of dipolar AFMs. Thus the Let us discuss the transition from a quantum spin Hamil- orientation of the magnetic moments of the elements of the tonian of the type ͑2͒ to the customary phenomenological superstructures ͑macroscopic spins͒ is often description of magnets. Although in reality the macroscopic 23 antiferromagnetic, and such systems can be regarded as description is often used as a starting point, it is of interest to ‘‘artificial AFMs.’’ They exhibit effects inherent to ordinary analyze this transition, primarily for investigating macro- crystalline AFMs, such as spin-flop or spin-flip scopic quantum effects. No less important is the fact that a 24–26 transitions, with the feature that the boundary elements consistent microscopic analysis can shed light on the ad- ͑surface͒ play a significant role. These and other novel ef- equacy of the macroscopic approach for description of a fects, such as a transition to a specific incommensurate given experiment. phase, are discussed in detail in subsection 4.3. For transition from the spin operators to the correspond- We do not know of any effects in which the scatter in the ing field variables one usually uses the spin coherent states values of the magnetic moments of individual structural ele- ͑SCS͒ ͉␴͘, which were introduced by Radcliffe60 ͑see also ments leads to ‘‘decompensation of the macroscopic spins’’ the review61͒ are parametrized by a unit vector ␴, ␴2ϭ1, or of artificial AFMs. However, another important effect of non- the angular variables ␪, ␸: ideality of the structure on the atomic level is well known for ␴ ϭ ␪ ␸ ␴ ϭ ␪ ␸ ␴ ϭ ␪ ͑ ͒ them: topological defects of the atomic structure of AFMs or x sin cos , y sin sin , z cos . 3 FM-AFM interfaces give rise to mesoscopic nonuniformities on account of frustration effects. The properties of such non- The spin coherent states can be constructed by acting with uniform states, including the problem of quantum tunneling the rotation operator Oˆ on the spin state Sˆ with a specified ϭ for them, will be considered in Sec. 7. maximum projection Sz S on a certain direction n. These ␴"ˆ ␴"ˆ͉␴ ϭ ͦ␴ states are eigenstates of the operator S, S ͘ S ‹, and for them the mean value of the spin operator corresponds III. CLASSICAL AND QUANTUM DESCRIPTION OF to its classical value MAGNETS For magnetic insulators and for a number of materials ͗␴͉Sˆ ͉␴͘ϭS␴. ͑4͒ with metallic conductivity59 an adequate description can be given on the basis of a Hamiltonian written in terms of spin This property is handy for making the transition from the operators S␣ localized at crystal lattice sites ␣.Asanex- quantum physics of spins to a phenomenological theory of magnetism. The SCS are a complete ͑more precisely, over- ample, let us consider a Hamiltonian Hˆ containing an inter- (i) ( j) complete͒ set of quantum states of the spin operator Sˆ,a action bilinear in the operators, of the form JijS1 S2 . Here ͑ ͒ property reflected in the existence of the ‘‘decomposition of the tensor Jij specifies both the isotropic exchange and the anisotropic pair interaction of the spins, and it can also con- unity’’ relations tain an antisymmetric part due to the Dzyaloshinskii–Moriya ͑2Sϩ1͒ interaction.45 The discussion below will be based on a ͵ D␴͉␴͗͘␴͉ϭ ˆ ͵ D␴ϭ ͵ ␪ ␪ ␸ 1, ␲ sin d d , Hamiltonian with only a nearest-neighbor interaction. We 4 ͑ ͒ shall use a simple version of it with purely uniaxial symme- 5 try ͑the selected axis z is a Cϱ axis͒: where 1ˆ is the unit operator and D␴ is the measure, which is ␴2ϭ z z defined in terms of an area element on the sphere 1. Hˆ ϭ ͚ ͕JS␣ S␤ϩ͑d ͓S␣ϫS␤͔͒ϩ␬S␣S␤͖ ͗␣␤͘ • • For applications to the theory of Heisenberg magnets it is important that all the mean values of powers of the spin Ϫ Ϫ z 2Ϫ ␮ " ͑ ͒ components can be expressed in terms of the corresponding d K͚ ͑S␣͒ g B͚ H S␣ . 2 ␣ ␣ powers of the means of the spin components, e.g.,62,63 Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov 639

␺ 1 S 2 ͗␴͉͑ ˆ" ͒2͉␴͘ϭͩ Ϫ ͪ ͗␴͉ ˆ" ͉␴͘2ϩ ͑ ͒ ϵ͗␺ ͉ ͕ ˆ ប͖͉␺ ϭ ͵ D␺ ͕ A͓␺͔ ប͖ ͑ ͒ S e 1 S e . 6 P12 1 exp iHt/ 2 exp i / , 7 ␺ 2S 2 1 This relation is valid for any value of the spin; for spin S where A͓␺͔ is the action functional known from classical ϭ1/2 this corresponds to a well-known property of the Pauli physics, A͓␺͔ϭ͐dtL͓␺͔, L is the Lagrangian of the sys- matrices64 and rules out the existence of single-ion anisot- tem, and the integration is over all values of the variable ␺ ropy. Here it is important to note that the SCS are con- describing the GCS. We shall not discuss here the problem of structed without the use of any approximations of the large- consistent definition of the measure D␺. This approach is in spin type. However, there do exist a number of properties of essence based on the idea of the Feynman path integrals. It the SCS that permit their constructive use only in that semi- has the advantage that it can be applied to quantum field classical limit. problems when one is speaking about space-time configura- An important distinction of the SCS ͑as for other coher- tions ␺ϵ␺(x,t) of the field.67,68 ent states͒ from the states usually adopted in quantum me- Using the properties of the GCS ͉␺͘, we can write the chanics is the lack of orthogonality: Lagrangian L͓␺͔ in the form t͉␺͘ϪW͑␺͒, W͑␺͒ϵ͗␺͉Hˆ ͉␺͘. ͑8͒ץ/ץS LϭϪប͗␺͉ 1 ͉͗␴ ͉␴ ͉͘ϭͫ ͑ ϩ␴ ␴ ͒ͬ 1 2 1 1• 2 . 2 The Lagrangian has the standard form TϪW, where the form of the kinetic term T can easily be found for coherent In the limit S→ϱ the value of ͗␴ ͉␴ ͘ is exponentially 1 2 states of any Lie group. The ‘‘potential energy’’ W in this small; in certain calculations this formally permits one to formula is simply the quantum Hamiltonian of the system assume that these states are orthogonal in the large spin limit. ␺͉ ˆ ͉␺ Spin coherent states, like the ‘‘ordinary’’ coherent states averaged over the coherent state ͗ H ͘. ͉␣͘ for bosons, where ␣ is a complex number ͑introduced by Let us return to the description of magnets in terms of ͉␴͘ Glauber,65 these states are defined as eigenfunctions of the the SCS . The Lagrangian of an individual spin S in the ͑ ͒ 34–38,67 boson annihilation operator aˆ , aˆ ͉␣͘ϭ␣͉␣͘), are particular angular variables 3 is usually written in the form ͑ ͒ t͒͑1Ϫcos ␪͒ϪW͑␸,␪͒, ͑9͒ץ/␸ץexamples of the generalized coherent states GCS intro- LϭបS͑ duced by Perelomov.66 They can be constructed for any Lie where W(␲,␪) has the meaning of the classical energy group Gˆ . For spin coherent states, Gˆ coincides with the ro- ␴ ϭ͗␴͉ ˆ ͉␴͘ tation group of three-dimensional space SO͑3͒, and for boson W( ) H , expressed in terms of the angular vari- 61 ables for the vector ␴. Variation of this Lagrangian gives the states this is the so-called Heisenberg–Weyl group. All 70 GCS share the fundamental property that they minimize the well-known Landau–Lifshitz equation in angular variables ͑ 34–37,71͒ indeterminacy of the values of the variables determined by see the monograph and review . the quantum uncertainty relation. This means that the GCS The kinetic term can be represented in a more general ͑ ͒ ␴ ͑in particular, the SCS͒ are the quantum states whose prop- form than 9 directly for the vector dynamical variable : ␴ ␴ϫnץ erties are most nearly classical. The use of coherent states is LϭបSA͑␴͒ ϪW͑␴͒, A͑␴͒ϭ , ͑10͒ t ␴͑␴ϩ␴n͒ץ • most efficient when the Hamiltonian of the system is linear with respect to the generators of the Lie group Gˆ . Then if an initial state is coherent and described by a certain GCS, its where n is the quantization axis of the coherent state. For quantum evolution will reduce to a variation of the param- transparency of interpretation it is convenient to assume that ␴ ␴2ϭ eters of the GCS, which are described by classical equations the variable is not restricted by the condition 1 but ͉␴͉ϭ␴ of motion.61 This permits a simple and elegant construction has a length 1. Then this Lagrangian corresponds of exact solutions of a number of problems of the quantum formally to the dynamics of a charged particle with a radius ␴ evolution of an arbitrary spin in an external variable mag- vector , moving in a magnetic field with vector potential A. ͑ ͒ netic field.61 Exact solution of more-complex problems of The vector potential A in formula 10 is the field of a mag- ␴" ϭϪ␴ the interaction of several spins, with a Hamiltonian bilinear netic monopole, having a singularity at n , i.e., on a ␴ in the operators of the spin components, is also possible. semi-infinite line segment in the space . One usually ץ ␴ץ ϭ ␴ For our problem it is very important that the use of the chooses n ez , and here the quantity A( )•( / t) takes ͑ ͒ GCS permits one to obtain in a consistent manner the semi- the well-known form 9 . For constructing the Lagrangian of ͑ ͒ classical Lagrangian of the system, which in principle cannot an AFM it is more convenient to use the general form 10 . be recovered solely from the classical equations of motion. Let us discuss the question of the description of the static Lagrangians describing the same classical equations of mo- characteristics of an AFM — its energy first and foremost. ͑ ͒ tion of the system can differ by terms that are total deriva- The coherent state for a spin Hamiltonian of the type 2 is ͉␴͘ϭ͟ ͉␴ ͘ tives with respect to time. Such a term cannot influence the chosen in the form a direct product of SCS, ␣ ␣ , classical dynamics of the system but is responsible for inter- where ͉␴␣͘ are the SCS for the spin operator Sˆ ␣ located at ference of the instanton trajectories67,68 ͑see Sec. 6͒; it is also site ␣. In this case the transition from the spin Hamiltonian, important for the definition of the soliton momentum.69 which does not contain products of operators of the spin The most elegant way of obtaining the Lagrangian for components at a given site, to the classical energy is done by any coherent states is based on the possibility of representing replacing the operators Sˆ ␣ by classical vectors S␴␣ of length ͑ the quantum propagator P12 the probability amplitude of a S. transition from one quantum state to another͒ in terms of a This way of writing the energy corresponds in the high- path integral over trajectories:67 est degree with our intuitive ideas about how to make the 640 Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov

transition from the quantum spin Hamiltonian to the classical with the study of Bose–Einstein of neutral energy of a magnet. We note that the possibility of such a atoms with nonzero spin.78–80 transition ͑rather than writing, say, the spin S␣ in the form For many magnets the non-Heisenberg interactions are 2 ␴␣ͱS(Sϩ1), naively employing the condition Sˆ ϭS(S rather weak, and the SCS are the most convenient formalism ϩ1), which is incorrect͒, does not rely on the approximation for their semiclassical description. Let us return to an analy- Sӷ1. sis of almost-Heisenberg AFMs in the framework of the Let us now discuss the specifics of magnets whose SCS. We consider a system of localized spins in which the Hamiltonian contains products of the spin component opera- nearest neighbors are coupled by an antiferromagnetic inter- ͑ ͒ tors at a single site. Such terms are present for magnets with action of the type in Eq. 1 . We shall assume that the sites at single-ion anisotropy ͑2͒ or a biquadratic exchange interac- which the spins are located can be divided into two groups in 2 such a way that all the spins belonging to nearest-neighbor tion of the form JЈ(S␣ S␤) . Magnets with non-small inter- • pairs are in different groups. ͑This condition is violated, e.g., action of this type are often called non-Heisenberg.72 For for a frustrated triangular lattice.͒ For an ideal AFM these them the ‘‘classical’’ limit corresponds to the substitution two groups correspond to two magnetic sublattices. The in- ͗␴ ͉ ˆ 2͉␴ ͘→ Ϫ "␴ 2ϩ ␣ (S␣•e) ␣ (1 1/2S)(e ␣) const; see formula troduction of sublattices can also be thought of as a division ͑ ͒ 6 . Actually, however, the description of Heisenberg mag- of the lattice into pairs of spins ͑dimers͒. The Lagrangian of nets on the basis of the SCS ͉␴͘␣ encounters more-serious ␴ an AFM can be written in terms of two unit vectors 1 and problems than the renormalization of the single-ion anisot- ␴ for each such dimer. In the case of an extended magnet → Ϫ 2 ropy constants K K(1 1/2S). one can also introduce a more general description of slightly Let us discuss the specifics of Heisenberg magnets for ␴ nonuniform states on the basis of the spin densities 1(x,t) the example of a magnet with strong easy-plane single-ion and ␴ (x,t) ͓or their irreducible combinations l(x,t) and 2 Ͼ 72 2 anisotropy of the form KSz , K 0. The ground state of m(x,t); see Eq. ͑14͒; x is the coordinate of the dimer͔ as ϭ such an ion is a quantum state with fixed projection Sz 0, continuous functions of the coordinates. The energy of the ˆ ϭ ˆ ϭ ˆ 2 ϭ ˆ 2 ϭ ϩ for which ͗Sx͘ 0, ͗Sy͘ 0, but ͗Sx ͘ ͗Sy͘ S(S 1)/2. nonuniformity into account is taken into account by expand- The properties of this state are far from those of the corre- ing in the gradients; see formula ͑40͒ below. ␴ ϭ ␴ sponding SCS with projection z 0, in which the vector In many problems, e.g., the quantum tunneling in AFM has a definite direction in the (x,y) plane. Therefore it can molecules, the nonuniformities can be neglected, and all the ␴ hardly be expected that the ground state of such a Hamil- vectors 1,2 assumed parallel. Then it is convenient to intro- duce the sublattice spins S ϭ͚S␣ and S ϭ͚S␣ and the tonian, even in the presence of the exchange interaction, will 1 1 2 2 ϭ ␴ be well described in terms of SCS. Of course, this state can coherent states for the spins of each sublattice, S1 sN1 1 ϭ ␴ be described in the form of a superposition of spin coherent and S2 sN2 2 . Here and below, s denotes the value of the ͑ ͒ states completeness of the set of SCS! , but this approach is spin at the site, N1,2 is the number of sites in each sublattice, extremely awkward. and we shall use the notation For description of non-Heisenberg magnets Ostrovski and Loktev73,74 proposed to use the GCS for Lie group SU(2Sϩ1), which for spin SϾ1/2 is more general than the NϭN ϩN , N ϭNϭN ϪN , ͑11͒ SCS. ͓Recall that the SCS are constructed with the use of the 1 2 ex 1 2 rotation group SO͑3͒, which is equivalent to the group SU͑2͒.͔ This approach permits one to construct a consistent ͑ ͒ field theory of the unusual states of such magnets, including where Nex gives the excess uncompensated spin Sex ϭ not only the magnetically ordered phases but also the so- sNex , and N is the total number of sites in the AFM. Thus called nematic or quadrupolar phases, in which magnetic we arrive at a description of the uniform dynamics of the ␴ ordering is absent. These phases are invariant with respect to AFM sample in terms of two vectors of unit length, 1 and ␴ time inversion but can be characterized by spontaneous sym- 2 : metry breaking due to anisotropy of the quadrupole variables ˆ 2 ˆ 2 of the type ͗Sx ͘,͗Sy͘. The geometric image of the spin state ␴ץ ␴ץ of the nematic phase is an ellipsoid. In the nematic phase of 1 2 LϭបsN A ͑␴ ͒ ϩបsN A ͑␴ ͒ tץ t 2 2 2ץ spin systems with an isotropic Hamiltonian the low- 1 1 1 frequency dynamics is described in terms of the sigma- Ϫ ͑␴ ␴ ͒ ͑ ͒ model for the vector director n that specifies the orientation W 1 , 2 . 12 of this ellipsoid.75 Even in the quadrupolar-ferromagnetic phase with M 0 for such magnets the dynamics of the mag- netization M includes the variation of M in length and differs The kinetic terms are represented in terms of the vector po- substantially from that which is obtained with the use of the tential ͑10͒ of the field of a magnetic monopole, and 76 ␴ ␴ Landau–Lifshitz equation. The static state of this type has W( 1 , 2) is the classical energy of the AFM, which corre- long been studied; as an example we cite the analysis of a sponds to the Hamiltonian ͑2͒. As we are interested in the ͉ ͉ Ӷ domain wall in which M goes to zero at the center of the case of small decompensation, Nex N, in the expression for 77 wall. In the last few years the problem of the nematic this energy we keep the term with Nex 0 only in the Zee- ϭ ϭ phases and other nontrivial types of nonmagnetic order of man term and set N1 N2 N/2 everywhere else; we then spins systems has drawn increasing interest in connection get Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov 641

1 plicity we shall assume that dϭ0 and the magnetic field is W(␴ ,␴ )ϭ s2ZN͓J␴ ␴ ϩ͑d ͓␴ ϫ␴ ͔͔͒ 1 2 2 1• 2 • 1 2 parallel to the easy axis of the AFM ͑a more general case will be considered below͒. Analyzing the energy of the AFM 1 ͑16͒, one can show that in this case the vectors ␴ and ␴ lie ϩ s2N͓␬Z␴ ␴ ϪK͑␴2 ϩ␴2 ͔͒ 1 2 2 z,1 z,2 z,1 z,2 in a single plane ͑say, the (x,z) plane͒ at all values of the field. Then the irreducible vectors m and l are determined 1 ␪ Ϫ g␮ sH ͓N͑␴ ϩ␴ ͒ϩN ͑␴ Ϫ␴ ͔͒, solely by the two variables , m: 2 B • 1 2 ex 1 2 l ϭl cos ␪, l ϭl sin ␪, lϭͱ1Ϫm2, ͑13͒ z x ϭ ␪ ϭϪ ␪ ͑ ͒ where Z is the coordination number, the second group of mz m sin , mx m cos . 17 terms describes the anisotropy energy, and the third the Zee- Minimization of the energy with respect to the variable m ␴ ␴ man energy. In addition to the unit vectors 1 and 2 ,itis leads to the simple relation convenient to introduce their irreducible combinations ␮ ␪ g BH sin ϭ͑␴ Ϫ␴ ͒ ϭ͑␴ ϩ␴ ͒ ͑ ͒ mϭ . ͑18͒ l 1 2 /2, m 1 2 /2, 14 s͑2JZϩ2K cos 2␪ϩZ␬͒ which are related by Formula ͑18͒ reflects the anisotropy of the susceptibility of l2ϩm2ϭ1 and m"lϭ0. ͑15͒ the AFM with respect to the mutual orientation of the mag- netic field H and the vector l. In particular, for lʈH (␪ϭ0) These new variables describe the characteristic symmetry of the magnetization m is equal to zero even for H 0. It will AFMs with respect to permutation of the sublattices and are be convenient in what follows to eliminate the total spin m convenient for writing the phenomenological energy of an and to write the energy of the AFM in terms of the angular AFM. We write the energy ͑13͒ in terms of the vectors l and variable ␪ alone: m: ͑ ␮ ͒2 2 ␪ ␬ ͒ϭ 2 2ϩ 2 ϫ ͒ g BH N sin Z W͑l,m Js zNm s zN͑d•͓m l͔ WϭϪ ϩͩ Kϩ ͪ s2N sin2 ␪. 2͑2ZJϩZ␬ϩ2K cos 2␪͒ 2 1 ͑19͒ ϩ 2 ͓␬ ͑ 2Ϫ 2͒Ϫ ͑ 2ϩ 2͔͒ s N Z mz lz K lz mz 2 In the absence of magnetic field we get the standard result: Ϫ ␮ " Ϫ ␮ " ͑ ͒ both the exchange and the single-ion anisotropy give an ad- g BsNm H g BsNexl H. 16 ditive contribution to the effective : We note an important relationship that holds for any AFM: the anisotropy energy always contains only even powers of 1 W ͑␪͒ϭ s2K N sin2 ␪, K ϭ2Kϩ␬Z, ͑20͒ the components of the vector l or m, and the Dzyaloshinskii a 2 eff eff interaction is linear in m. In the general case instead of (d ϩ ␬Ͼ ϫ which is of the easy-axis type for 2K Z 0. For this ͓m l͔) the expression Dik(l)milk appears, where the ten- • model the stable phase at HϽH is the collinear phase ⌽ʈ , sor D (l) contains even power of the components of l and is 1 ik in which ␪ϭ0or␲, and for HϾH the stable phase is the determined by the symmetry of the AFM. 2 spin-flop phase ⌽Ќ , in which ␪ϭ␲/2. For model ͑19͒ the

characteristic fields for loss of stability of the collinear ⌽ʈ ⌽ IV. FIELD-INDUCED PHASE TRANSITIONS IN and spin-flop Ќ phases, H1 and H2 , are given by the for- ANTIFERROMAGNETS mulae The ground state of an AFM is determined by the mini- s H2 ϭ ͱ͑ ϩ␬ ͒͑ ϩ␬ ϩ ͒ ϭ 1 mum of the energy ͑16͒ with respect to the vectors l and m. H1 ␮ 2K Z 2ZJ Z 2K , H2 , g B Hsf Ne´el himself called attention to the fact that this state can change in a jump upon a continuous change of magnetic s ϭ ͱ͑ ϩ␬ ͒͑ ϩ␬ Ϫ ͒ ͑ ͒ field.40 In other words, phase transitions induced by a strong Hsf ␮ 2K Z 2ZJ Z 2k . 21 g B magnetic field can occur in AFMs. An example is the spin- flop transition ͑SFT͒,40 the experimental observation of In the spin-flop phase mz increases linearly with increasing which was a long time in coming because of the necessity of field as 81 using high fields. The SFT has been studied for over 50 H s͑2JZϩ␬ZϪ2K͒ ͑ m ϭ , H ϭ , HрH , ͑22͒ years and is still attracting great research interest see the z H ex g␮ ex monograph and reviews5,9,12,82͒. As we have said, in recent ex B ͑ ͒ years the study of ‘‘novel’’ AFMs ͑spin pairs and clusters, where Hex is the exchange field of the AFM in model 19 , Ӷ magnetic superlattices͒ has revealed a number of phenomena and H1 ,H2 Hex . If the value of the field H exceeds Hex , similar to the SFT. then lϭ0 for mϭ1, and the system reaches saturation. The transition to the saturated state ⌽ with lϭ0 is accompa- A. Classical bulk transitions PM nied by the destruction of the antiferromagnetic order and is In a boundaryless magnet the SFT occurs simultaneously called a spin-flip transition. It is a typical second-order tran- throughout the volume of the material; we shall call this a sition of the order-disorder type, but it is brought on not by bulk SFT. The origin of such a transition is easily explained an increase in temperature but by an increase of the external by analysis of a purely uniaxial AFM model ͑16͒. For sim- field. 642 Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov

FIG. 1. The ground state of a classical AFM with energy ͑19͒ for different K and different values of the field ͑schematic͒. The arrows indicate the directions of the sublattice spins in different states of the AFM; see text. ͑a͒ KϾ0; first-order transition, the continuations of the curves into the region of metastability of the phases are indicated by dotted lines, which terminate at the arrowheads at the instability points. ͑b͒ KϽ0, second-order transition, the dashed curve corresponds to the variation of the magnetization in a canted field.

It is important to take both types of anisotropy into ac- model in Eq. ͑16͒. For ␺ 0 the vector l is not directed along count simultaneously for the description of all possible symmetry directions, ␪ 0,␲/2. The phases that are trans- phases of an AFM and transitions between them. If the an- ⌽ ⌽ ␺ϭ ⌽ ⌽ formed into ʈ and Ќ at 0 will be called 1 and 2 , isotropy is of a purely exchange character, i.e., Kϭ0, then a respectively. The stability fields H1 and H2 of these phases degenerate SFT, for which critical fields H1 and H2 coincide, behave differently with increase of the angle ␺, H decreas- Ͼ 1 occurs. If K 0, then the SFT occurs as a first-order transi- ing and H increasing according to the laws83 ͑ ͒ 2 tion at the field Hsf introduced above in Eq. 21 , with a ⌬ ͑␺͒ϭ ͓ Ϫ ␺2/3͔ ͑␺͒ϭ ͓ ϩ ␺2/3͔ ͑ ͒ jump Sz of the total spin Sz of the system: H1 H1 1 a , H2 H2 1 a , 24 ␬Zϩ2K where aϭ(3/2)(K/JZ)1/3 is small, with a degree of small- ⌬ ϭ ͱ ͑ ͒ Sz 2sN . 23 ␺ϭ␺ 2JZϩ␬ZϪ2K ness K/J. It follows that at a certain value c all the characteristic values of the field coincide, H (␺ )ϭH (␺ ) Ͻ 1 c 2 c There is another possible case, when H1 H2 , and the sta- ϭH ϷH , and the ⌽ ↔⌽ exists only at ⌽ ⌽ c sf 1 2 bility regions of the phases ʈ and Ќ do not overlap. Then a sufficiently small value of the angle ␺р␺ , Ͻ Ͻ ⌽ c in the field interval H1 H H2 a canted phase Ͻ with 0 Ͻ␪Ͻ␲/2 exists, and the value of S changes from zero to 2 K 2/3 z ␺ ϭ ͩ ͪ ͑ ͒ ⌬ c . 25 Sz by a linear law but with a larger slope than in the region 3 JZ of the spin-flop phase ͑see Fig. 1͒. In this case the transition ͑ ͒ ␺ ⌽ʈ↔⌽Ќ occurs via two second-order transitions through the Thus in the model 16 on the (H, ) plane the line of first- canted phase. In model ͑19͒ this requires the presence of order phase transitions terminates in a critical point ␺ ␪ ␺ ϭ␪ ␺ ϭ␲ 83 competing anisotropies—single-ion anisotropy with KϽ0 ( c ,Hc), at which 1( c ,Hc) 2( c ,Hc) /4. ͑which for ␬ϭ0 corresponds to easy-plane anisotropy͒ and a For KϽ0, when two second-order transitions are real- rather large exchange anisotropy, with ␬ZϾ2͉K͉, so that the ized in the symmetric case ␺ϭ0, the tilt of the field is of effective magnetic anisotropy is easy-axis. fundamental importance: for ␺ 0 the second-order phase ⌽ ↔⌽ ␺ transitions 1,2 Ͻ vanish. Indeed, for 0 the vector l deviates from the symmetric directions ␪ϭ0,␲/2 at any B. Lowering of the symmetry of the problem value of the field, and therefore the symmetry differences of ͑ ⌽ ⌽ ⌽ We have discussed phase transitions the spin-flop and the phases 1 , 2 from the canted phase Ͻ vanish. At very spin-flip transitions͒ in the simplest and most symmetric case small ␺ 0 the angle ␪ varies rather sharply with field at of the model ͑2͒, assuming in addition that dϭ0 and that the values near where the second-order transitions would occur magnetic field is strictly parallel to the easy axis of the AFM. for ␺ϭ0, and the field dependence approximately reiterates For possible applications of the theory developed it is impor- ␪(H) for ␺ϭ0 ͑see Fig. 1͒ but is analytic and has no kinks. tant to know whether these transitions are preserved when In the analysis of phase transitions in a concrete model, one goes beyond the framework of such a model. As we have e.g., the model of Eq. ͑2͒, the question of whether the results said, in the case of a ferromagnet the Curie point vanishes in might change when some other invariants are taken into ac- the presence of magnetic field. For an AFM the situation is count. An example of such a situation was presented above: ⌽ ↔⌽ more favorable, and both the first-order spin-flop and the the second-order transitions 1,2 Ͻ are very sensitive to spin-flip transition at which the antiferromagnetic order is the symmetry of the problem and vanish at an arbitrarily destroyed ‘‘survive’’ in a quite general case. small deviation of the field from the axis, while the first- We start with an analysis of the question of what hap- order transition ‘‘survives’’ a rather strong perturbation of the pens when the magnetic field H deviates from the easy axis problem. A general ͑model-free͒ analysis of the existence of of the AFM by some small angle ␺ in the framework of the the spin-flip transition can be given on the basis of symmetry Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov 643

84 ⌽ arguments. In the saturated phase PM the magnetic mo- ments of the sublattices are parallel to each other and to the magnetic field, i.e., all the components of l are equal to zero. ⌽ The phase PM has a higher symmetry than any antiferro- magnetic phase, in which l 0. Its symmetry group coincides the symmetry group of the paramagnetic phase in the pres- H ence of an external magnetic field, G PM . The order param- ⌽ eter describing the transition to the paramagnetic phase PM is the vector l. This transition will be realized only in the case when none of the components of the vector l is an invariant with respect to elements of the symmetry group H G PM . It turns out that if the symmetry elements of the para- magnetic phase include odd translations or inversion, then the spin-flip transition will occur at arbitrary orientation of 84 the magnetic field H. If such elements are not present, then FIG. 2. Two possibilities for the arrangement of the vectors of the magnetic the spin flip transition will either be absent altogether or will field H, total spin S, sublattice spins S1 and S2 , and also the antiferromag- occur only for an orientation of the field H along preferred netic vector l in the noncollinear state of a particle of an uncompensated AFM in an external magnetic field. directions. This conclusion agrees with the results of an analysis of the effective energy of an AFM with the Dzy- aloshinskii interaction; see formula ͑38͒ below. sation N ϪN ӶN, the value of H is small, H ӶH , and Odd symmetry elements of the translation or inversion 1 2 c c ex type are characteristic for a number of crystalline AFMs, and therefore the ground state of the AFM is degenerate over a their presence forbids the existence of any type of Dzy- wide range of fields. aloshinskii interaction. Such discrete symmetry elements as The calculation presented above was done in neglect of odd inversion can be important not only for crystalline AFMs the magnetic anisotropy. Taking anisotropy into account is ͑AFMs with such an element have unique properties; see the somewhat awkward, and the result depends on the ratio of ͑ ͒ monograph by Turov et al.9͒ but also for finite spin clusters the characteristic fields Hc 26 and the field Hsf introduced with an antiferromagnetic interaction. above. Importantly, even when certain forms of anisotropy ͑ ͒ The foregoing discussion pertains only to the case of a are taken into account the degeneracy discrete of the completely compensated AFM. It is of interest to determine ground state is preserved upon decompensation of the AFM whether reorientation transitions are possible for small AFM in a strong magnetic field. Between these states, coherent particles, in which the spins of the sublattices are not equal, quantum tunneling effects are possible.

S1 S2 , and in the exchange approximation a nonzero mag- netic moment is present in the system. Assuming for defi- C. Spin-flop transition and incommensurate phases near Ͼ niteness that S1 S2 , we find that this moment will tend to the surface and defects S orient along the field in such a way that the spin 1 is parallel For large enough samples the demagnetizing effect of S to the field and 2 is antiparallel. However, besides the ‘‘fer- the surface of the a magnet will lead to the existence of a romagnetic’’ spin-orientation effect, which is linear in the thermodynamic-equilibrium domain structure. For an AFM field, there also exists another orientation effect which is ͑ ͒ ϭ such a structure intermediate state of the AFM consists of quadratic in the field, which also arises for S1 S2 and leads domains of the phases ⌽ʈ and ⌽Ќ ͑Ref. 86; see also the to a spin-flop transition. The combined effect of these two reviews12,87͒. For mesoscopic AFMs it becomes important factors in the case of sufficiently small decompensation gives that the spins at the boundary have a coordination number a certain analog of the spin-flop transition, which for sim- ˜ZϽZ. Quite some time ago Mills88 and also Keffer and plicity we shall discuss in the exchange approximation. Chow89 noted that in such a case the value of the spin-flop Suppose that the magnetizations of the sublattices make transition field ͑21͒ is lower at the boundary than in the bulk. angles ␪ and ␪ with the magnetic field. The collinear state 1 2 The calculation in those papers concerned the SFT in semi- described above, with antiparallel spins ␪ ϭ0, ␪ ϭ␲, is re- 1 2 infinite systems, in which the boundary of the crystal con- alized only at sufficiently low fields, HрH , where c tains spins of only one sublattice. This model is essentially equivalent to a spin chain, ˜ZϭZ/2 ͑see Fig. 3͒, for which the ϭ ͑ Ϫ ͒ ͑ ϩ ͒ ͑ ͒ Hc Hex S1 S2 / S1 S2 . 26 Х & 88–90 field of the surface SFT is Hc Hsf / . A surface SFT is accompanied by the appearance of a soft mode of surface у ␪ ϭ␪ ϭ 90 For H Hex the spins are parallel, 1 2 0. It is clear that . Granted, the possibility of realizing this interest- in terms of symmetry these states are equivalent to the para- ing phenomenon is largely determined by the quality of the ⌽ magnetic state PM and are not degenerate. However, in an surface, and for a real ͑non atomically smooth͒ surface of a Ͻ Ͻ intermediate field region Hc H Hex the sublattice spins S1 crystal its observation is difficult. That is possibly why at- ␪ ␪ ␲ 85 and S2 are noncollinear, 1 , 2 0, . Here the total spin is tempts at experimental observation of a surface SFT have not parallel to H, and the value of its transverse ͑perpendicular to met with success.91 ͒ ␪ ϩ ␪ ϭ H projection S1 sin 1 S2 sin 2 0, but the state of the sys- It turns out that the surface SFT is realized for one- tem is degenerate with respect to direction of the vector l dimensional magnetic superstructures of the type represented ͑see Fig. 2͒. For the case of interest, that of small decompen- by Fe/Cr multilayer films with an antiferromagnetic interac- 644 Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov

FIG. 4. Collinear ͑Ne´el͒ state of a two-dimensional AFM for several de- fects, bulk and surface: a—a single- defect formed by a nonmagnetic impurity; b and c—two-atom defects, compensated and uncompensated, re- FIG. 3. Collinear antiferromagnetic state for several systems discussed in spectively; d—a corner of the sample; e—two types of boundaries ͑see text͒. the text. Here and in the subsequent diagrams of spin structures in this The easy axis is perpendicular to the plane of the figure, and the magnetic Section the easy axis of the AFM is perpendicular to the plane of the figure, field is directed upward. The light and dark circles denote the spin-up and the magnetic field is assumed to be directed upward, the light and dark spin-down states, respectively, and the solid lines represent exchange bonds. circles denote the spin-up and spin-down states, respectively, and the solid line represent exchange bonds. Part ͑a͒ shows the uncompensated AFM surface considered in Refs. 88–90, and the arrow indicates the direction of the normal to the surface, n. The other parts show model low-dimensional and a finite density of nonmagnetic impurities of substitution systems: the spin chain ͑b͒ and spin ladder ͑c͒. creates an ensemble of identical ‘‘surfaces’’ on which obser- vation of the localized SFT effect is possible. Let us discuss the nature of the surface states and tran- tion of the layers, grown on an anisotropic MgO͑110͒ sub- sitions to these states for more complex systems. For all strate. These materials have growth anisotropy with an easy AFMs, excluding the simple chain, two types of surfaces can axis in the plane of the layer.92 The lower value of the ex- be distinguished from the standpoint of spin states. The sur- change field in the boundary layer can cause surface face can be compensated, containing an equal number of nonuniformity.93 The localized surface SFT in such materials spins of the different sublattices and having zero surface has been observed and investigated by various methods.24–26 magnetization in the Ne´el state, or uncompensated, with non- Initially, in the spirit of Refs. 88–90, it was interpreted94 as zero surface magnetization, as in Fig. 3a. It is intuitively the nucleation of the spin-flop phase near the surface, i.e., as clear that in these cases the nonuniform states that arise are a rotation of the magnetic moments of the surface layers by very different. As an example let us consider the layered an angle close to 90°. Later, on the basis of calculations in ͑2D͒ AFMs of the manganese halide type, which have been the model of a semi-infinite spin chain and analysis of the studied for many years.97 Substitution of the Mn2ϩ ions with experimental data, it was shown that the situation is more spin Sϭ5/2 by nonmagnetic impurities such as Zn or Cd complicated.26,95,96 It turns out that the boundary spin anti- ions98,99 isolated or combined into clusters gives rich possi- parallel to the field in the initial state reverses direction under bilities for realization of various magnetic defects ͑bulk and the influence of the field. Because of this, the state arising surface, compensated and uncompensated͒ with a locally near the surface is not the spin-flop but an antiferromagnetic weakened exchange interaction ͑see Fig. 4͒. state with zero magnetization, but with opposite directions of The basic regularities of these phenomena can be under- the magnetic moments in comparison with the bulk. In other stood from an analysis of a model object—the so-called spin words, the antiferromagnetic state near the surface differs ladder ͑SL͒.100,101 A spin ladder contains two spin chains from the bulk state by the direction of the vector l, and they coupled by an AFM interaction ͑see Fig. 3͒. Such spin sys- are separated by a 180° domain wall. Such a state is appro- tems occur as structural elements in many low-dimensional priately called an incommensurate phase.95 magnets and have interesting quantum properties.102,103 A References 24, 26 and 95 aroused interest not only in the spin ladder can be represented as a chain of spin dimers with problem of the SFT near the surface and the incommensurate AFM coupling of the spins both within each dimer and be- phases in AFMs. Keffer and Chow89 had noted the possibil- tween neighboring dimers. It is natural to introduce vectors ity of existence of localized transitions of a similar type in mn and ln for the structural dimers of a SL, where n is the bulk AFMs containing defects of the dislocation and grain- number of the dimer. Importantly, a finite SL can contain an boundary types. For dislocations the situation is not so analog of the compensated boundary illustrated in Fig. 3c simple ͑see Sec. 7 below͒, but it quite likely that analogs of and an uncompensated boundary with an incomplete dimer the surface SFT should be observed near defects of ‘‘ordi- ͑uncompensated spin͒ at the end ͑see Fig. 7 below͒ and also nary’’ crystalline AFMs. various bulk defects that do not reduce to breaking of the SL. Let us consider the simple case when magnetic ions in Let us give the results of an analysis for SLs with these low-dimensional magnets are replaced by nonmagnetic ions, types of boundaries. For an SL with an uncompensated spin and the corresponding exchange bonds are broken. The result the value of the critical field Hc for the transition to a non- is simplest for spin chains, in which the substitution of a uniform state differs significantly more strongly from the Ӎ magnetic ion by a nonmagnetic impurity within the chain bulk field H1 Hsf than for an SL with a regular dimer at the leads simply to the breaking of the chain and the formation end.100,101 Numerical minimization of the discrete Hamil- of two ‘‘surfaces’’ of the type considered in the early tonian ͑2͒ has shown that the structures of the phases are also 88,89 papers. In this case each chain ‘‘breaks’’ independently, different, e.g., the form of the distribution of mz and lz as Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov 645

␪ ϭ␪ Ϫnpϩ Ϫ n␪ Ϫnq n pe ͑ 1͒ qe ,

ϭ Ϫnpϩ Ϫ n Ϫnq ͑ ͒ mn m pe ͑ 1͒ mqe , 28 where

2 cosh ␳ϭ3˜J/JϪ1; 2 cosh qϭ3˜J/Jϩ1,

˜ϭͱ 2ϩ͑ 2Ϫ 2͒Ӎ 3J 9J h1 h 3J. ␪ The relations between the amplitudes p,q and m p,q can be ␪ ϭ ␪ ϭ Ϸ written in the form p Amp , q mq /A, where A h/6J Ӷ1. Therefore for a slowly damped exponential with pӶ1it Ӷ␪ is found that m p p , and for an oscillatory exponential with fast damping (qϳ1) the situation is the opposite. ͑ ͒ ͑ ͒ FIG. 5. Distributions of lz light symbols and mz dark symbols as func- The structure of the equations for the boundary dimer is tions of the number of the site for a semi-infinite spin ladder with single-ion different from ͑27͒. A stability analysis can be carried out anisotropy and different boundaries for values of the field above the critical.101 The circles are for a spin ladder with a regular end, and the using the equations for variables describing the spins that do diamonds are for a spin ladder with an uncompensated spin at the end. For not belong to regular dimers with nϾ0. These include the clarity the values of m have been multiplied by 4 for the SL with the ␪ z variables 0 , m0 describing the last dimer, and the angular uncompensated spin at the end and by 20 for the SL with the regular end. variables ␪, ␸ for the ‘‘additional’’ spins present in an un- The irregular behavior of mz(n) is clearly seen for the SL with the uncom- pensated spin; for the regular SL the details of its behavior near the bound- compensated SL. The solvability conditions of the linear ho- ary are shown in the inset. mogeneous equations for all these variables are satisfied only for a certain value of the magnetic field, which is what de- termines the stability field Hc and also the relation between ␪ ␪ functions of the distance n from the end of the ladder ͑see the variables p and q . Analysis of SL models with the use 101 Fig. 5͒.101 For a regular SL, which can be regarded as a of this example has led to the following results. model AFM with a compensated surface, the data on the For a regular SL the critical field, to a first approxima- tion in the small parameters (␬,K)/J, has the form distribution of lz show the existence of a surface SFT. In such systems the vector l rotates by approximately 90° in 2ϭ 2Ϫ ͑ ␮ ͒2͑ ϩ␬)͒2 ͑ ͒ going from the surface to the bulk; here the magnetization is Hc H1 2 s/g B K . 29 nonzero in the entire spin-flop region near the surface and This value is smaller than the bulk critical field H , but the falls to zero in the bulk. For an uncompensated SL, as can be 1 difference between H and H is proportional to the small seen in Fig. 5, an incommensurate state arises in which the c 1 parameter (␬,K)/J. A calculation showed that ␪ Ӷ␪ , and magnetization is close to zero both on the surface and in the q p the behavior of the vector l is regular, while the values of m bulk, while the nonuniformity of the spins is characterized p and m are comparable, m Ϸ2m /(1ϩ))Ӎ0.73m . This by the presence of a 180° domain wall. In both cases the q q p p explains the presence of irregular behavior of the magnetiza- dependence of the vector l on the number of the dimer n is tion in the immediate vicinity of the boundary of the SL ͑see rather smooth and well described by formulae obtained for Fig. 5͒. the components of l in antiferromagnetic domain walls: ␪ ϭ Ϫ Ϫ ⌬ ␪ In the analysis of an uncompensated SL the result for the tan n exp͓ (n n0)/ ␲/2͔ for a 90° wall, or tan( n/2) critical field turns out to be qualitatively different, Hc ϭexp͓Ϫ(nϪn )/⌬␲͔ for a 180° wall. However, the numerical 0 ϭͱ2/5H Ϸ0.63H . As in the case of the spin chain,88,89 the data show that, unlike that of l, the behavior of m near the 1 1 z value of H is substantially lower than H ; the difference boundary is substantially irregular. c 1 between H and H is of the order of H and does not con- To explain these features, let us consider the problem of c 1 c tain small factors. In this case ␪ Ӷ␪ , but again irregular stability of a collinear phase having the values ␪ ϭ0 and q p n behavior of the magnetization appears, and the ratio of m m ϭ0.101 The stability of this state can be investigated using q n and m has a different sign than for a spin ladder with a the Hamiltonian ͑2͒ written in the quadratic approximation in p regular boundary: m ϷϪm /(1ϩ))ӍϪ0.37m . This the small variables ␪ and m ; for the end dimer the index q p p n n property also corresponds well to the numerical data ͑see the nϭ0. The states of dimers with nϾ0 are described by the inset in Fig. 5͒. For a wide class of quasi-one-dimensional same equations as for a boundryless spin ladder: systems with an uncompensated boundary or bulk defect, the ͑ ␪ Ϫ␪ Ϫ␪ ͒ϩ ␪ Ϫ ϭ J 2 n nϩ1 nϪ1 Keff n hmn 0, critical field Hc is expressed by the simple formula

͑ ϩ ͒ ϩ ͑ ϩ ͒Ϫ ␪ ϭ ͑ ͒ 2 4J Keff mn J mnϩ1 mnϪ1 h n 0, 27 2˜Z H2ϭ H2, ͑30͒ ϭ ϩ ␬ ϭ ␮ c 2 1 where Keff 2K 3 , and h g BH/s is a dimensionless 2˜Z ϩZ field. The solutions of these equations are a complex analog ␪ ϰ Ϫ ˜Z Z of the Bloch state in an ideal lattice, n ,mn exp( Qn). The where and are the coordination numbers for the bound- corresponding ‘‘dispersion relation’’ has two solutions for the ary spin and the regular part of the AFM, respectively. For a ‘‘quasi-momentum’’ Q, and therefore two exponentials arise spin chain (˜Zϭ1 and Zϭ2) this gives the known result 2ϭ 2 88,89 in the solution: Hc H1/2. For a single nonmagnetic defect inside the 646 Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov

FIG. 7. Dependence of the quantities characteristic for analysis of the order (crit) Ϫ1 FIG. 6. Dependence of the z projection of the total spin Sz on magnetic field of a phase transition, Sz and (dSz /dH) , on the parameter k ϭ ϭ for a spin ladder with a regular dimer at the end and an effective anisotropy 2K/Keff , found numerically at the transition point H Hc for a spin lad- ϭ 101 constant Keff 0.06J. The data shown are for pure exchange and single- der with uncompensated spin at the boundary at a value of the effective ϭ 101 ion anisotropy and also for two types of combined anisotropy with different anisotropy constant Keff 0.02J. The structure of the SL in the collinear ␬ ϭ ratios of the contributions of the constants K and (k K/Keff). state is illustrated in the inset.

SL, which breaks one of the spin chains making up the spin quadratic in the variables of the linear problem, ␪ , m , and ladder, ˜Zϭ2 and Zϭ3, one obtains the answer H n n c that S ϭ0 in the Ne´el state, we arrive at the conclusion that ϭͱ8/11H Ϸ0.853H , which is in good agreement with the z 1 1 S should be considered to be the square of the order param- numerical results. z eter. For description of the phase diagram with a tricritical The majority of studies of the surface SFT have consid- point one must expand the energy of the system to sixth ered only modes with single-ion anisotropy and discussed order in the order parameter: only first-order transitions. A numerical analysis shows that two types of behavior of the magnetic-field dependence of 1 1 FϭϪ͑HϪH ͒S ϩ ␤S2ϩ ␥S3. ͑31͒ the z component of the total spin Sz are present. For a system c z 2 z 3 z with single-ion anisotropy for the nonuniform phase there is ͑ ͒ a finite value of Sz at the transition point, where the energies The form of the first term in formula 31 is obtained ana- of the collinear and nonuniform phases coincide. However, lytically from the expression for the energy of the SL and the for the case of exchange anisotropy the value of Sz tends equations of the linear approximation. Determination of the toward zero at the transition point, as is shown in Fig. 6. coefficients ␤ and ␥ requires solution of a nonlinear problem, Such behavior can be interpreted as signs of a first-order and but in the analysis they can be treated as phenomenological second-order transition, respectively. As in the case of the parameters and their values determined from the numerical bulk SFT, the kind of localized SFT depends largely on the data.101 If the coefficients are known, then the further analy- microscopic origins of the anisotropy. For all of the defect sis is simple: for ␤Ͼ0 the transition is second-order with a ϭ Ϫ ␤ ␤Ͻ models in a SL with pure single-ion anisotropy there is a linear dependence of Sz(H): Sz (H Hc)/ . For 0 and first-order transition. For the case of exchange anisotropy the ␥Ͼ0 the transition will be first-order with a nonzero value of (crit) (crit) transition to the nonuniform phase is a second-order transi- the total spin total spin Sz at the transition point: Sz ϭ ͉␤͉ ␥ Ϫ1ϭ␤ ␤ tion, and the amplitude of the nonuniform distribution goes 3 /4 . Thus both quantities, (dSz /dH) for Ͼ (crit) ␤Ͻ ͉␤͉ to zero at the transition point. In the vicinity of this point the 0 and Sz for 0, are proportional to the value of . state cannot be classified as a surface spin-flop phase or an Near the tricritical point ␤→0, and extrapolation of the incommensurate phase; it is more aptly characterized as a curves of these two quantities as functions of 2K/Keff to zero slightly distorted Ne´el state with a distortion amplitude that at a constant value of the effective uniaxial anisotropy Keff falls off in an oscillatory manner with distance from the sur- ϭ2Kϩ3␬ gives the value of the parameters of the tricritical face into the interior of the crystal. point. Such a graphical analysis of the SL model with an The presence of first- and second-order transitions for a uncompensated spin at the end is presented in Fig. 7. It gives given system indicates the presence of a tricritical point. The the position of the tricritical point for an anisotropy ratio Ϸ square-root dependence characteristic of a tricritical point is 2K/Keff 0.97. For a simple spin chain the tricritical point ϭ Ӎ Ϸ clearly seen in Fig. 6 for k K/Keff 0.78. Numerical mini- corresponds to the value 2K/Keff 0.67. For a regular SL the Ϸ mization of the energy of the system near second-order tran- tricritical point is found at 2K/Keff 0.78. Recall that the sitions and especially near a tricritical point is greatly com- bulk SFT occurs as a first-order transition for all values K plicated, but the analysis can be simplified by using Landau’s Ͼ0. One can therefore state that for localized transitions the phenomenological theory of phase transitions. Taking into case of a second-order phase transition is more typical than ϭ␪ account that the z projection of the dimer spin mz,n nmn is for a bulk transition. Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov 647

For a 2D AFM the situation remains qualitatively the Heisϭ ␮ Hex 2JS/(g B), and the value of the field Hex coincides same as for the SL: single-ion anisotropy gives a first-order with that given by the classical formula ͑22͒ for Kϭ0, ␬ transition, and anisotropy of the interaction gives a second- ϭ0, and Zϭ1. Thus the process of magnetizing a spin pair is order transition. For all the uncompensated surface or bulk reminiscent of the magnetizing of macroscopic isotropic defects the value of the critical field ͑30͒ is substantially AFMs, with the difference that for pairs of quantum spins the H H2 Ͻ lower than the instability field 1 , and the difference c linear dependence Sz(H) for H Hex becomes steplike, with Ϫ 2ϳ 2 H1 H1 does not contain any small parameters of the type a change in spin by unity. (K,␬)/J. In a 2D AFM, as for the SL, for all the compen- The exact solution of the quantum problems of finite 2Ϫ 2ϳ ␬ 2 Ӷ 2 sated defects Hc H1 (K, )H1/J H1 . There is only one spin clusters with an antiferromagnetic interaction for arbi- difference of a quantitative character, viz., the localization of trary value of the spin S is known only for certain simple the states in the 2D case is weaker, while the difference of models.105 An interesting model is that of a closed spin chain the fields of the bulk and localized SFTs is less than for the with four spins, SL. Since the SL is in essence a one-dimensional object, this 3 regularity corresponds to the known tendency for localiza- ˆ ϭ ˆ ˆ ϩ ˆ ˆ H J͚ Si•Siϩ1 JS4•S1 . tion effects to weaken as the dimensionality of the system iϭ1 increases.34,35 However, for such defects as an uncompen- sated surface and especially a corner of the sample ͑see Fig. The Hamiltonian of this model is easily rewritten in terms of ͒ ˆ slϭ ˆ ϩ ˆ ˆ slϭ ˆ ϩ ˆ 4d, e , the effects of localization of the SFT in a 2D AFM are the ‘‘sublattice spin’’ operators S1 S1 S3 , S2 S2 S4 , ˆ ϭˆ slϩ ˆ sl manifested more strongly than for a spin ladder. and the total spin operator Stot S1 S2 ; with accuracy up The possibility of realization of 2D localized transitions to an additive constant, it takes the form is due to some novel objects—arrays of magnetic dots of J different shapes, made from magnetically soft ferromagnets, Hˆ ϭ ͓Sˆ 2 Ϫ͑Sˆ sl͒2Ϫ͑Sˆ sl͒2͔. which can be arranged in square, rectangular, or staircase 2 tot 1 2 structures. The interaction between dots is of a magnetostatic The eigenvalues of the Hamiltonian are expressed in terms of nature, and in a weak field it leads to antiferromagnetic order the value of the total spin S of the system and the values of of the magnetic moments of the dots.23 An important circum- tot ˆ sl 2ϭ sl sl ϩ stance derives from the character of the anisotropy in these the sublattice spins, (S1,2) S1,2(S1,2 1). As in the classi- artificial systems. The effective anisotropy inherent to a cal Ne´el state, in the ground state of this system the sublat- sl ϭ single dot is determined by its shape, while the anisotropy of tice spins have the maximum possible value S1,2 2S. For the interaction energy of the dots is related to the shape of systems with a large number of spins such a simple treatment the lattice. This makes it possible the control independently cannot be done, and the conclusion that the sublattices are the parameters analogous to the single-ion or exchange an- ‘‘saturated’’ is apparently invalid. When the anisotropy is isotropy constants by modifying the geometry of the dot and taken into account the exact solution is known only for the ϭ ͑ ͒ lattice, respectively. , which corresponds to J,K 0in 2 . In this ϭ Ͻ case, in essence classical, we have Sz 0atH Hex and ϭ Ͼ saturation of the state, Sz 2S,atH Hex . At a value of the ϭ␬ ␮ D. Quantum analog of the spin-flop transition exchange field Hex S/(g B) the state of an Ising system is degenerate, and S can take any value from 0 to 2S. We have actually been discussing the spin-flop transition z In the classical theory even weak anisotropy can alter the in a model of two large spins. In recent years many experi- purely linear trend of S (H), giving rise to a spin-flop tran- ments have been done that can be described directly in terms z sition of either first or second order ͑see Fig. 3͒. Suppose the of this model. We are talking about processes of magnetiza- classical SFT field H ХSͱ2JK/(g␮ ) exceeds the field of tion of spin pairs with an antiferromagnetic interaction,104 sf B the first quantum jump, H for nϭ1, H ϭ ϭJ/(g␮ ), as including pairs of the high-spin Mn molecules with a maxi- n (n 1) B 4 can occur for moderate anisotropy KϾJ/S2. Then in the mum value of the total spin 2Sϭ9.54,55 For an ensemble of behavior of the spin pair one expects the appearance of the such pairs ͑and also more complex clusters like spin triplets, quantum analog of a spin-flop transition.106 The problem for quartets, etc.͒ a stepped magnetization is observed, i.e., a a spin pair in the presence of magnetic anisotropy of a gen- macroscopic manifestation of the spin quantization effect.104 eral form does not have an exact solution even in the most The process of magnetization of spin pairs when only symmetric case ͑2͒. A numerical analysis by Kireev and the isotropic exchange and the external field are taken into ac- author for a pair of spins in the simple model ͑2͒ with allow- count can be described on the basis of the exact solution of ance for general types of anisotropy showed that even for not the quantum problem ͑2͒ with Kϭ0, ␬ϭ0. Taking into ac- very large spins such as Sϭ3 – 5 one observes fair agreement ˆ ˆ ϭ ˆ 2 Ϫ ϩ count the identity 2S1•S2 Stot 2S(S 1), we can write this of the quantum and classical results. The behavior of the Hamiltonian solely in terms of the total spin operator of the quantum jumplike magnetization curve is determined, as for ˆ ϭˆ ϩ ˆ system, Stot S1 S2 , and its projection in the field direction classical AFMs, by the microscopic character of the mag- ˆ ϭ Sz . Corresponding to the eigenstates are fixed values of the netic anisotropy. In particular, the value Sz 0 is maintained 2 ϭ ϩ ӷ total spin Stot , Stot Stot(Stot 1), and Sz , and their energy is up to field values H H(nϭ1) . Further, for single-ion anisot- ϭ ϩ determined by the expression E(S,Sz) JStot(Stot 1)/2 ropy there exists a microscopic analog of the first-order spin- Ϫ ␮ g BHSz . Comparing these values of the energy, it is easy flop transition, viz., the jump of Sz from zero to values ϭ Ϫ ϭ ⌬ Ͼ to find that the value of Sz jumps from Sz n 1toSz n at Sz 1. For the case when in the macroscopic problem has ϭ ␮ ϭ a field Hn Jn/(g B). Saturation Stot 2S is reached at second-order transitions, for a quantum system the jumps 648 Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov

⌬S ϭ1 are concentrated in a narrow region of field values ប ͑␴ ͒ ␴ ϩ ប ͑␴ ͒ ␴ z sN1 A1 1 • ˙ 1 sN2 A2 2 • ˙ 2 that correspond approximately to the characteristic fields lץ H1 ,H2 of the classical theory. ϭsប ͓N A ͑l͒ϪN A ͑Ϫl͔͒ t • 1 1 2 2ץ mץ ϩ ប ͓ ͑ ͒ϩ ͑Ϫ ͔͒ N1A1 l N2A2 l • ץ s V. LAGRANGIAN FOR THE DESCRIPTION OF THE t ͒ Ϫ ץ ͒ ץ ץ DYNAMICS OF REAL AFMs l A1͑l A2͑ l ϩsបm ͫN ϩN ͬ. ͑32͒ ץ 2 ץ 1 • ץ i 108 107 Kaganov and Tsukernik and Turov and Irkhin first t li li proposed to describe the dynamics of AFMs on the basis of a Up to this point we have not specified a concrete choice of system of two Landau–Lifshitz equations for the sublattice axes of the singularities n1 and n2 for the potentials A1 , A2 . magnetizations, which can be written in terms of the unit We choose them so as to eliminate the singular term, at least ␴ ␴ ͑ ͒ vectors 1 and 2 10 , as the field variables. An equivalent in the case of a compensated AFM.117,118 The corresponding but more convenient description is based on a system of ϭ Ϫ condition is A1(l) A2( l), which is possible when n1 equations for the irreducible vectors l, m ͑see the detailed ϭϪ n2 . For Nex 0 one cannot get rid of the singular term, analysis of this approach and further generalizations of the but it is proportional to the small quantity N ӶN. ͒ ex such equations in the monograph by Turov et al.9 . It later The dynamic term in ͑32͒ for the case of a ‘‘pure’’ AFM became clear that by using the natural small parameters of begins with the term linear in m, which can be represented in ␮ ϳͱ Ӷ ͓ ͑ ͔͒ the AFM theory, g BH/S KeffJ J see Eq. 2 , one can the form118 satisfactorily describe the behavior of an AFM using only the Aץ Aץ lץ equation for the antiferromagnetic vector l alone. In this case ប ͫ ϫ ͬ ϭ␧ ͩ jϪ k ͪ ͑ ͒ 33 . ץ ץ Bi ijk , ץ Ӷ m• B͉ ͉ the vector m is small, m 1, it is a slave variable and is t lk l j t.Inץ/lץ determined by the vector l and its time derivative this approximation the vector l, which for m 0 is three- Thus, in contrast to the FM case, the Lagrangian of a com- component with a variable length ͓see Eq. ͑15͔͒, can be pensated AFM contains only the gauge-invariant quantity B ϭ treated as a unit vector. A, which has the meaning of a magnetic monopole field, ϭ ͉ ͉3 " ϭ The dynamical equations for the antiferromagnetic unit B l/ l . Applying the condition m l 0, we easily obtain vector l can be constructed on the basis of symmetry for the slave variable m ץ arguments43 and can be obtained directly from the Landau– ប ͒ ͒ l Lifshitz equations for the sublattice magnetizations109–112 mϭ ͭ ␥͑H͑eff Ϫl͑H͑eff l͒͒ϩͫ ϫlͬͮ , ͑34͒ tץ • 2sJZ ͑see reviews12,34,36,38͒. In both approaches one obtains the ␥ϭ ␮ ប (eff) same classical equations of motion for the unit vector l, where g B / is the gyromagnetic ratio, and H is the which are called the ␴-model equations. The simplest ver- sum of the external field H and the Dzyaloshinskii field sion of the ␴-model has high dynamical symmetry, and the H(D). For the model ͑10͒, in which the Dzyaloshinskii– time derivatives of l appear in the Lagrangian in the trivial Moriya interaction has the purely antisymmetric form d ␴ ϫ␴ (D)ϭ ϫ ␮ 2 ץ ץ form ( l/ t) . The use of such equations as these have made •( 1 2), the Dzyaloshinskii field H sZ(d l)/g B . it possible to describe the Lorentz-invariant domain-wall dy- For the usual form of this interaction, proportional to ͓ ͑ ͔͒ namics observed experimentally for a number of AFMs, Dik(l)milk see the discussion following formula 16 , and (D)ϭ where it was found that the domain walls move with a char- the Dzyaloshinskii field Hi Dik(k)lk . Thus for a general acteristic velocity c equal to the velocity in the AFM model the expression for the effective field becomes AFM.6,15,37,38 Both the magnetic field and certain forms of effϭ ϩ ͑ ͒ ͑ ͒ the Dzyaloshinskii–Moriya interaction lower the dynamical Hi Hi Dik l lk . 35 ␴ symmetry of the classical -model, and its Lagrangian ac- We note that the vector m determines the value of the total .tץ/lץ quires terms linear in spin Stot of the system only in the case of a compensated A key development in the use of the equations of the ␴- ϭ sample, when Nex 0. If Nex 0, the total spin vector will model in the quantum physics of AFMs was to obtain the contain both a ‘‘transverse’’ term, proportional to the compo- ␴ Lagrangian of the -model on the basis of the spin coherent nents of m, and a ‘‘longitudinal’’ term, which is due to the 67,113–116 ␴ states. We begin by considering the -model for the excess spin and is parallel to l: uniform dynamics of the vector lϭl(t) with the magnetic ϭ ϩ ͑ ͒ field and the Dzyaloshinskii interaction taken into account, Stot sNexl sNm. 36 with the possibility of a small decompensation of the sublat- The derivation of the ␴-model equations can be carried fur- tice spins. The nonuniform dynamics will be briefly dis- ther without specifying a concrete form of Dik(l). Substitut- cussed at the end of this Section. In Secs. 6 and 7 we shall ing m into formula ͑10͒, we obtain an effective Lagrangian present this Lagrangian for investigating tunneling effects. for the vector l of an AFM sample containing N spins: The derivation of the ␴-model is done under the assump- lץ l 2ץ tion that the magnetic moment is small, ͉m͉Ӷ͉l͉. Using the ប2N 1 Lϭ ͫ ͩ ͪ Ϫ␥ͩ H͑eff͒ ͫlϫ ͬͪͬ tץ • tץ smallness of m, one can write the Lagrangian ͑10͒ in the 2JZ 2 form an expansion in powers of that vector. The kinetic term lץ -in the Lagrangian for the sublattice spins ͑12͒ can be de ϩsបN A͑l͒ ͩ ͪ ϪW˜ ͑l͒. ͑37͒ tץ • scribed in the approximation linear in m as ex Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov 649 t. They describeץ/lץ Here the first term governs the characteristic dynamics for a of additional gyroscopic terms linear in ‘‘pure’’ AFM, the second term is ‘‘ferromagnetic’’ in form a lowering of the real dynamical symmetry on an AFM on and is due to decompensation of the spins, and the last term account of the magnetic field and/or the Dzyaloshinskii in- W˜ (l) has the meaning of an effective ‘‘potential energy’’ of teraction. We note that in the original papers43,112,119 these the system: terms of different origin were written in the correct form of total derivatives. This pertains to terms resulting from the ͑g␮ ͒2N ˜ ͒ϭ ͒Ϫ ␮ " ͒ϩ B " ͒2Ϫ 2 magnetic field ͑cf. Refs. 43 and 51͒, the Dzyaloshinskii in- W͑l Wa͑l g BsNex͑H l ͓͑H l H ͔ 4JZ teraction ͑Refs. 119 and 118͒, and also from the form of the ␮ 2 Lagrangian for an AFM with incomplete compensation of ͑g B͒ N ϩ ͑Hϫ͓l͑H͑D͒ l͒ϪH͑D͔͒͒. ͑38͒ the spin ͑cf. Refs. 112 and 120͒. Therefore, in particular, the JZ • 2 soliton momentum P, which is not invariant with respect to In writing W˜ here we have written out separately all the the addition of total derivatives, and the dispersion relation ϭ terms that depend on the magnetic field. The expression for E E(P) for a soliton of the kink type were found correctly W(l) has the meaning of the effective anisotropy energy of in Refs. 119, 121, and 122. ͑ ͒ the AFM. This energy is made up of the ‘‘bare’’ anisotropy Let us discuss the form of the Lagrangian 37 in various ␴ energy mentioned above ͑the second group of terms in for- limiting cases. The standard -model of a pure AFM corre- ͑ ͒ ͑ ͒ ␴ ϭϪ␴ ϭ sponds to S ϭS ,orN ϭ0. For the transition to the La- mulae 13 and 16 , which are taken for 1 2 l or 1 2 ex mϭ0, respectively͒, and also the additional term (H(D) l)2 grangian of a ‘‘pure’’ ferromagnet the term proportional to 2 ץ ץ • Ϫ(H(D))2, which is acquired in the elimination of m.Itis ( l/ t) must be set to zero. Taking this term into account clear that there is no sense in separating out these contribu- goes beyond the accuracy of the calculation for sufficiently Ͼͱ 111,120 tions, and W(l) is the real anisotropy energy determined small Nex /N K/J. However, in doing so we alter the from static measurements. In essence, in writing W(l) one Hamiltonian structure of the system; specifically, the number ͑ ͒ must simply use the phenomenological expression whose of degrees of freedom the dimension of the phase space form is determined by the crystal symmetry of the magnet decreases by half, which is of fundamental importance for and, as a rule, is well known. Below we shall often consider the analysis of instantons. ͑ ͒ a model AFM of rhombic symmetry and use the expression In the derivation of the Lagrangian 37 the possibility of nonuniform states was neglected for simplicity. They can be ͑␪ ␸͒ϭ 2 ͓ ͑ 2ϩ 2͒ϩ 2͔ ͑ ͒ W , s N Ku lx ly K ply , 39 taken into account without difficulty by allowing for the de- pendence of the vector l on the spatial coordinates. For the where the constants K and K describe the uniaxial anisot- u p model of an AFM with a D-dimensional Heisenberg lattice ropy and the anisotropy in the basal plane (x,y), respec- with lattice constant a, this dependence is put in by changing tively. from the dynamical variable l(t) to the field variable l(x,t) The energy W˜ (l) of the system also contains terms linear and making the substitution and quadratic in the magnetic field. The meaning of the first term in ͑38͒ is simple: it describes the Zeeman energy of the dxD 1 N͑L/N͒→ ͵ ͫ͑L/N͒Ϫ Ja2s2͑ "I͒2ͬ. ͑40͒ excess spin in a magnetic field. A second term, quadratic in aD 2 “ the field components, is present for any AFM. The meaning of this contribution is very clear: it is due to the anisotropy of The value of L/N has the meaning of the value of the La- ͑ ͒ the susceptibility of the AFM. The role of this term can be grangian per spin for the uniform case 37 , and the second represented as the appearance of a field-induced uniaxial an- term in the integrand is taken with the opposite sign to the ͑ ͒ isotropy with a hard axis along the magnetic field. The com- energy of nonuniformity of the spins see, e.g., Ref. 71 .We effϭ petition of this induced anisotropy with the initial easy-axis note that for H 0 the derivatives with respect to the coor- crystalline anisotropy is responsible for the existence of the dinate and time appear in the Lagrangian in the Lorentz- spin-flop transition ͑see Sec. 4͒. Finally, the last term, which invariant combination t2͒, cϭsaJͱ2Z/ប, ͑41͒ץ/2ץis bilinear in the components of the external field H and the ٌ2Ϫc2͑ Dzyaloshinskii field H(D), describes the energy of the weak ferromagnetic moment caused by the Dzyaloshinskii interac- where c has the meaning of the magnon phase velocity in the tion with the external magnetic field.12 It contains terms lin- AFM. ear in the components of the vector l, making it possible to A special case is the one-dimensional AFM, e.g., a spin explain easily the conclusion of a symmetry analysis that the chain, when the vector l depends on one spatial coordinate, ϭ spin-flip transition exists for any orientation of the external l l(x,t). In the case of a simple chain the action acquires an 67,116 magnetic field only in the absence of the Dzyaloshinskii in- additional topological term lץ lץ teraction. Analyzing this term, one can easily indicate those 1 selected orientations of the field for which the spin-flip tran- Atopolϭ2␲បSQ, Qϭ ͵ dxdtͩ l ͫ ϫ ͬͪ . ͑42͒ tץ xץ • 4␲ sition can occur for nonzero H(D). Indeed, the terms linear in l, which destroy that transition, go to zero under the condi- The quantity Q is a topological invariant of the mapping of ϭ 2ϭ tion DikHi 0, which determines the desired direction of H. the (x,t) plane onto the sphere l 1. For a continuous func- Thus the Lagrangian for the vector l, with the external tion lϭl(x,t) having a unique value for x,t→ϱ, this quan- magnetic field and the general form of the Dzyaloshinskii tity takes on only integer values, Qϭ0,Ϯ1,Ϯ2,... . It is clear interaction taken into account, differs from the Lagrangian of that the variation of this quantity is equal to zero, and its the ␴-model of an ideal AFM67 by the presence of a number presence is not manifested in the equations of motion for the 650 Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov vector l. However, the topological term does influence the Without fear of contradiction it can be said the synthesis behavior of the non-small quantum fluctuations; in particular, and study of high-quality single crystals of high-spin mol- it leads to differences in the character of the ground state of ecules with precise orientation of the molecular axes ͑their the AFM chain with integer and half-integer spin.67,116 The properties were discussed in Sec. 2͒ have opened up a new topological term also determines interference effects for non- era of research on macroscopic quantum effects.1 Both ef- one-dimensional instantons for the problem of tunneling in fects of tunneling between excited levels, for which the split- disclinations and domain walls123,124 ͑see Sec. 7 below͒. The ting is larger,133,134 and tunneling in the ground state135–139 presence of the topological term depends fundamentally on have been observed for such systems. Parallel orientation of the magnetic symmetry of the AFM; in particular, it is the anisotropy axes of the molecules in single crystals has present for a simple spin chain but absent for a spin ladder made it possible to really address the question of observing ͑which we discussed in Sec. 4 in connection with the prob- the oscillations predicted by Garg131 in the tunneling splitting lem of the spin-flop transition͒; this circumstance leads to as a function of the external magnetic field, and only for a fundamentally different behavior of quantum fluctuations in field parallel to the hard axis of the magnetic particle. Such these systems.102 The question of whether topological invari- oscillations have been observed by the method of relaxation ants exist in AFMs of higher dimension, e.g., the Hopf in- time measurements at low temperatures139 for high-spin mol- variant, which describes the mapping of the three- ecules both with integer spin Sϭ10 ͑Refs. 137 and 138͒ and dimensional manifold (x,y,t) onto the sphere l2ϭ1, remains for systems with half-integer spin Sϭ9/2; for these cases the open to discussion. phase of the oscillations is shifted by ␲/2. Recently the di- rect observation of the tunneling level-splitting for antiferro- VI. TUNNELING IN SMALL AFM PARTICLES AND magnetic Fe8 molecules by the method of inelastic neutron 140 ANTIFERROMAGNETIC MOLECULES scattering was reported. This led to new theoretical studies of the tunneling problem for an ensemble of oriented In the study of the macroscopic quantum effects, the particles.141,142 phenomenon of coherent macroscopic quantum tunneling be- tween energetically equivalent but physically different states A. Instanton approach to the description of tunneling in systems with discrete degeneracy of the ground state is of The probability amplitude of a transition from state ͉2͘ to 1,28,29 ͉ ͑͘ particular interest. A typical effect of this type is the a different state 1 the propagator P12) can be written in mixing of two equivalent classical states and the tunneling terms of a path integral ͑7͒. If one is talking about tunneling, splitting of the energy levels corresponding to them, which i.e., motion in the classically forbidden region, then the value are degenerate in the classical case. The original papers on of the mechanical action in the ordinary mechanics of the the theory of quantum tunneling were done for small par- particles is purely imaginary. Then the Feynman exponent A ប ϪA ប ticles of ferromagnets under the assumption that all the spins exp(i / ) will be a purely real expression exp( Eu / ), ͑ A ϭϪ A in the particle are parallel to each other the large-spin where Eu i is the Euclidean action. For magnets 125–128 model͒. Then it turned out that from an experimental whose Lagrangian contains terms linear in the derivatives, standpoint AFMs are a more convenient class of objects for the Euclidean action can have both real and imaginary parts. investigation of quantum tunneling. According to In the classically forbidden region the real Feynman expo- 129,130 A A calculations, in AFMs the level splitting is stronger nent appears automatically, but the transition from i to Eu than in ferromagnets, and the effects can be observed at is conveniently done with the aid of a transition to imaginary higher temperature. It is not surprising that the first observa- time, t→i␶. Then for estimation of effects of macroscopic tion of quantum tunneling effects was done for ferritin par- quantum tunneling we immediately employ the Euclidean ticles of biological origin, which have an antiferromagnetic action functional, which can be written in the form structure27 ͑see Sec. 2͒. l 2ץ Tunneling in magnets entails subtle and beautiful effects ϱ ϱ ប2N 1 A ͓ ͔ϵ ͵ ␶ϭ ͵ ␶ͭ ͫ ͩ ͪ Eu l LEud d ␶ץ of interference of instanton trajectories. For ferromagnetic Ϫϱ Ϫϱ 2JZ 2 particles in the absence of magnetic field these effects lead to lץ lץ suppression of tunneling for half-integer total spin of the ͑eff͒ 32,33 ϩ ␥ͩ ͫ ϫ ͬͪͬϩ ប ͑ ͒ ͩ ͪ ␶ץ • ␶ is NexA lץ system, and this can be considered an example of the i H • l working of Kramers’ theorem on the ground-state degen- eracy. In the presence of magnetic field H, Kramers’ theorem ϩW˜ ͑l͒ͮ , ͑43͒ is inapplicable, and the result cannot be predicted before- hand. For ferromagnets, tunneling in the case H 0 was studied theoretically by Garg with the use of the instanton where the energy W˜ (l), the effective field H(eff), and the method.131 The analysis led to an extremely nontrivial result vector potential A(l) are determined by the same formulae as that a partial ͑non-Kramers͒ interference can arise. If the in the ‘‘ordinary’’ action ͑37͒. field is parallel to the hard axis of a particle with rhombic To analyze the tunneling in an AFM between different anisotropy, then the tunneling level-splitting will have an os- classical degenerate ground states having different values of cillatory dependence on the magnetic field. In pure AFMs, the vector l, labeled l(ϩ) and l(Ϫ), and to find the value of the i.e., for complete compensation of the sublattice spins, inter- tunneling splitting ⌬ of the energy of these states, one must 51,132 A ប ference effects can appear in an applied magnetic field estimate the tunneling exponent Re Eu / . The main object and in the presence of certain types of Dzyaloshinskii– in that formalism consists in the so-called Moriya interaction.118 instantons,28–30,68,143 which are solutions of the Euler– Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov 651

Lagrange equations for the Euclidean version of the AFM ␴-model the Lagrangian is real, all the ⌽(␣) are equal to Lagrangian lϭl(␶). The instantons correspond to the mini- zero, and therefore the combinatorial factor is trivial, ͱK A ␶→Ϯϱ mum of Re Eu͓l͔ under the condition that for the ϭn, i.e., for n equivalent paths the value of the level split- vector l tends toward two different ground states, l(␶) ting is equal to the contribution of one instanton, multiplied →l(Ϯ). In the semiclassical approximation the value of by the number of paths. A ប Re( Eu / ) is not small, and the contribution to the splitting is given only by Euclidian trajectories having the minimum B. Instantons as solutions of the dynamical problem value of the real part of the Euclidean action A ͓l͔.An Eu For analysis of the instanton solutions one can exploit important feature of magnets is the presence of several the extremely useful analogy with the problem of finding the equivalent instanton paths. In the leading approximation in separatrix solutions of finite-dimensional dynamical systems. the parameter exp(ϪA /ប) for physically observable split- Eu The vector l is conveniently parametrized by the usual angu- ting ⌬ of the degenerate energy levels of the ground state, lar variables: one can write143 ϭ ␪ ␸ ϭ ␪ ␸ ϭ ␪ ͑ ͒ l1 sin cos , l2 sin sin , l3 cos . 47 ⌬ϭប␻ ͚ͯ D ͑ϪA͑␣͒ ប͒ͯ ͑ ͒ 0 ␣ exp Eu / , 44 ͑ ͒ ␣ In the case of the general form of the Lagrangian 43 the structure of the instanton is determined by a general system where for convenience we have immediately separated out of two second-order Euler–Lagrange equations for the vari- ប␻ ␻ the factor 0 , where 0 is the frequency of small oscilla- ables ␪(␶),␸(␶), which are in general complex. Let us dis- tions of the vector l, and the index ␣ enumerates the equiva- cuss the possibility in principle of constructing the instanton lent instanton paths corresponding to the same ͑smallest͒ solutions of such a system. The boundary conditions for the A(␣) (0) →␶→0 for ␶ץ/lץ , value of Re Eu . The coefficient D␣ is the pre-exponential given problem have the form l→lϮ factor for a given instanton path, which in standard quantum Ϯϱ. This problem is similar to the problem of finding a mechanics is determined by the first correction to the classi- moving domain wall, which is described by solutions of the 64 cal solution. In the instanton approach this factor ͑the fluc- form lϭl(␰), where ␰ϭxϪvt, v being the velocity of the tuation determinant͒ is due to small fluctuations around the wall, with the same boundary conditions for ␰→Ϯϱ. Con- instanton. Its calculation is a complex mathematical problem sequently, the instanton ͑like a soliton of the wall type͒ cor- that will be discussed below. responds to a heteroclinic separatrix trajectory of the dy- Importantly, in magnets the corresponding partial contri- namical system ͑43͒, which joins the pair of equivalent (␣) butions D␣exp(ϪA /ប) can be ͑and often are͒ complex, (0) Eu minima lϮ of the potential W˜ (l). For the problem with more ⌽(␣) i.e., they contain a phase factor exp(i ), which leads to than one degree of freedom it is by no means always possible interference of the instanton paths. Often such interference to construct such a solution. The situation in the case of arises because of a difference of the values of the imaginary instantons is still worse than for solitons, since the solutions, A(␣) part of Eu . For example, in a rhombic ferromagnet the generally speaking, are not real and, in essence, the number A(␣) ␣ϭ quantities Eu for two instanton paths ( 1,2) contain of degrees of freedom is doubled. ␲ ប Ϫ ␲ ប ⌬ terms i Stot and i Stot . Therefore, the splitting An arbitrary solution of the equations of motion of a ϭ ͉ ϪA(1) ប ϩ ϪA(2) ប ͉ D exp( Eu / ) exp( Eu / ) acquires an interfer- given dynamical system can be constructed if the system is ͉ ␲ ͉ 32,33 ence factor cos Stot , which describes a suppression of integrable. If even the vector l is real, we will have a system 120 tunneling for half-integer spin. Chiolero and Loss gave an with two degrees of freedom. For it to be integrable there example of a tunneling problem in an AFM for which the must exist two independent integrals of motion. One of them pre-exponential factor D␣ is complex and gives a substantial is easily written: it is the Euclidean analog of the total en- contribution to the phase factor. To take interference effects ergy: into account explicitly for a system with several instantons ␸ 2ץ ␪ 2ץ ប2N ͑ for a magnet with easy axis Cn there exist n instanton tra- E ϭ ͫͩ ͪ ϩsin2 ␪ͩ ͪ ͬϪW͑␪,␸͒, ͑48͒ ␶ץ ␶ץ jectories͒ it is convenient to write118 Eu 4JZ ⌬ϭ ប␻ ͱ ͑Ϫ A͑␣͒ ប͒ ͑ ͒ but the second integral of motion by no means always exists. 2 0D K exp Re Eu / , 45 It can be constructed trivially for the case when the anisot- where D is now the modulus of the fluctuation determinant, ropy is purely uniaxial, i.e., when LEu does not depend on and we have explicitly separated out the phase ͑combinato- ␸. Below we shall use this integral of motion ͑68͒ to con- rial͒ factor K: struct instanton solutions for a purely uniaxial AFM in mag- ⌬⌽͑␣,␤͒ netic field. However, in the case of any nonzero decompen- Kϭ͚ͯ cosͩ ͪͯ, ⌬⌽͑␣,␤͒ϭ⌽͑␣͒Ϫ⌽͑␤͒, sation of the sublattices for a purely uniaxial model the ␣,␤ 2 12,34,35 144 ͑ ͒ moving domain wall and instanton are absent; their 46 existence is contingent on anisotropy in the basal plane. where the summation is over all pairs of instantons, and Analysis of dynamical systems with more than one de- ⌬⌽(␣,␤) is the difference of the phase factors for the ␣ and ␤ gree of freedom, in particular, the construction of integrable instantons. systems, has remained one of the most important problems Ϫ A ប In this formula exp( Re Eu / ) coincides exactly with of analytical dynamics for over a hundred years and is still the exponential factor that appears in the coefficient of trans- under intensive study.145 Among the integrable systems parency of a barrier in the semiclassical approximation of known are models of the dynamics of a unit vector, e.g., the ordinary quantum mechanics. For the simplest version of the classical Neumann problem146 of the motion of a material 652 Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov point along a sphere in the field of a potential that is a qua- ables pi ,qi , is equal to zero. Consequently, any ferromagnet dratic function of the Cartesian coordinates, and its generali- model whose energy W(cos ␪,␸) is analytic with respect to zation for a number of other polynomial potentials.147 We cos ␪ and ␸ in the sense indicated above reduces to an ex- ϭ note especially the integrable model with W Kiklilk for H actly integrable dynamical system. The efficiency of this way ϭ ϭ 0, Dij 0 and with a gyrotropic term of the monopole field of constructing the instanton trajectories for a number of fer- type.147,148 Integrability of such a model also follows from romagnet models was demonstrated in Ref. 150. The same the exact integrability of the Landau–Lifshitz equations for arguments can be applied to the instantons in an AFM, which the magnetization of a two-dimensional ferromagnet, which in the case of complex variables ␪, ␸ are described by a 149 was established by Sklyanin. Diagonalizing the matrix Kik dynamical system with four degrees of freedom, but the con- and taking into account the condition l2ϭ1, we arrive at the struction of one additional first integral does not solve the anisotropy energy ͑39͒ for a rhombic AFM. This model de- problem of integrability of the problem in general form. scribes the exact instanton solutions for the AFM model ͑77͒ For the often-used model of strong easy-plane anisot- at arbitrary decompensation.144 ropy the equations of motion give the approximate relation ␪ 36 ץ ␸ץ ␪Ӎ On going from an AFM to the Lagrangian of a ferromag- cos const•( / t), and cos is a dependent variable. In net, which is a nontrivial operation, the problem simplifies this approximation the Lagrangian takes on the standard Ϫ ␸ 2 ץ ␸ץ ␸ substantially. Formally, one can obtain the Lagrangian of a form for mechanics, M eff( )( / t) /2 W( ), where ͑ ͒ ␸ ␸ ferromagnet from 43 if the term quadratic in the derivative, M eff( ) and W( ) are the effective mass and the potential, ␶)2, is set to zero. However, in so doing one would alter which are determined by the parameters of the system. Inץ/lץ) the Hamiltonian structure of the system, and the dimension this case the problem reduces to a mechanical system with of the phase space would decrease by half. While for an one degree of freedom. However, such a reduction of the AFM the angles ␪ and ␸ that parametrize l can be treated as phase space of a Hamiltonian system with two degrees of canonical coordinates, and the momenta corresponding to freedom can lead to a loss of important physical effects in 150 them contain ␪˙ and ␸˙ , for a ferromagnet the Hamiltonian tunneling. variables are expressed only in terms of l. As the canonical pair one can choose the momentum Pϭcos ␪ and the coor- ϭ␸ dinate Q . A ferromagnet corresponds to an integrable C. Fluctuation determinant Hamiltonian system with one degree of freedom, and the role of Hamilton’s function is played by the energy W(␪,␸)of Let us discuss the evaluation of the pre-exponential fac- ͑ ͒ the ferromagnet. For instantons the corresponding Hamil- tor D in formula 45 for the tunneling level-splitting. In the tonian is complex, instanton approach this factor is due to the contribution of small deviations of the trajectory in the total Feynman inte- -W gral from the classical paths ␪ ,␸ corresponding to the inץ Qץ Wץ Pץ i ϭϪ , i ϭ , ͑49͒ 0 0 P stanton. This contribution can be estimated in the Gaussianץ ␶ץ Qץ ␶ץ approximation ͑the semiclassical approach!͒ with the aid of and the solution in a ferromagnet cannot be real, and there- an expansion of the Euclidean action to second order in the fore the number of degrees of freedom is actually equal to small deviations: two. However, in a recent paper by Kulagin and the author it ␪ ␶͒ϭ␪ ␶͒ϩ␽ ␶͒ ␸ ␶͒ϭ␸ ␶͒ϩ␦␸ ␶͒ ͑ ͒ was shown that for any ferromagnet model, integrability is ͑ 0͑ ͑ , ͑ 0͑ ͑ . 51 150 ϭ ϩ present even in this case. We write W H1 iH2 , where ␪ ϭ␪ ␶ ␸ ϭ H1 and H2 are real-valued functions. Assuming P and Q are For instantons of the form 0 0( ), 0 const, which of- complex, PϭPЈϩiPЉ, QϭQЈϩiQЉ, we separate the real ten arise in tunneling problems in AFMs, the variables ␽ and and imaginary parts of equations ͑49͒. As a result, we obtain ␦␸ describe the longitudinal and transverse deviations from a system of four real equations, the right-hand sides of which the instanton trajectory. Introducing the variable ␮ ϭ␦␸ ␪ will contain the derivatives of H1 and H2 with respect to the sin 0 , we obtain the factor D in the form variables PЈ,PЉ,QЈ,QЉ. However, if it is required that W be an analytic function of the complex variables P and Q and 1 Dϭ ͵ D͓␽͔D͓␮͔expͩ Ϫ ͵ dx͑⌿,Aˆ ⌿͒ͪ , ͑52͒ satisfy the Cauchy–Riemann conditions independently, then 2 the right-hand sides can be rewritten in terms of the deriva- ⌿ϭ ␽ ␮ ˆ tives of H1 only or of H2 only. As a result, the system of where ( , ) is a two-component vector, and A isa2 equations for the real variables becomes a Hamiltonian sys- ϫ2 matrix differential operator, with operators of the Schro¨- tem. In particular, by choosing pairs of canonical variables in dinger type with respect to the dimensionless variable ␰ ϭ␻ ␶ the following way: 0 along the diagonal. An example of such an operator that arises in a particular problem is given below; see for- p ϭPЈ, q ϭQЉ; p ϭPЉ, q ϭQЈ, ͑50͒ 1 1 2 2 mula ͑71͒. The off-diagonal structure of Aˆ corresponds to the system can be written in the form of the standard ͑real͒ two interacting field degrees of freedom, which substantially Hamiltonian system with two degrees of freedom, p˙ i complicates the calculation of D as compared to the standard ϭ ץ ץϭ ץ ץϭϪ H1 / qi , q˙ i H1 / pi , i 1,2, with the Hamiltonian case of instantons in a system with one degree of freedom. ϭ function H1 H1(p1 ,q1 ,p2 ,q2) and the additional integral Sometimes—for example, for instantons in the simple model 144 of motion H2 . Using only the Cauchy–Riemann condition of an uncompensated AFM—this factor can be calculated. for W, one can show that the Poisson bracket of the func- However, in a number of cases the operator Aˆ is diagonal, in tions H1 and H2 , calculated in terms of the canonical vari- which case the fluctuation determinant decomposes into a Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov 653

product of two independent factors, longitudinal Dʈ and external field, the Dzyaloshinskii interaction, and possible transverse DЌ . We shall limit discussion to this simpler ex- decompensation. As we shall see, the role of these interac- ample. tions is extremely diverse, and even if one could write down

Calculation of the longitudinal determinant Dʈ is a rather a general formula for the Euclidean action with all these nontrivial mathematical problem that has nonetheless be- factors included, it would be difficult to analyze. We there- come a standard element of the theory of instantons.68,143 Its fore consider the role of individual interactions that lower the value can be formally written in terms of an infinite product dynamical symmetry of an AFM, moving from the simple to of nonzero eigenvalues of the corresponding Schro¨dinger op- the complex. The presence of uncompensated spins compli- ␭ ϳ ͟ ␭ 1/2 erator k , Dʈ ͓ k(1/ k)͔ . Besides them, the Schro¨dinger cates the analysis significantly, and we shall therefore begin operator for the longitudinal deviations always contains a with the case of a completely compensated AFM, for which ␭ϭ ϭ zero mode with 0. The presence of the zero mode is due Nex 0. It is convenient to begin discussion with the tunnel- to the invariance of the Euclidean action with respect to the ing in the simplest case of a model with the highest dynami- ␶ → ␶Ϫ␶ ␶ ⌫ϭ substitution l0( ) l0( 0), where 0 is an arbitrary pa- cal symmetry, which corresponds to 0. The point is that rameter. The contribution of this mode cannot be described for certain AFM models with the Dzyaloshinskii–Moriya in- in the Gaussian approximation; its inclusion causes a large teraction and nonzero ⌫(␪,␺) 0 the results turn out to be A ប 1/2 ͑ ⌫ ␪ ␺ ϭ ͑ factor (Re Eu / ) to appear in Dʈ see details in the exactly the same as for ( , ) 0 this will not happen in reviews68,143͒. As a result, the longitudinal fluctuation deter- the case of magnetic field͒. minant can be written in the form If ⌫(␪,␸)ϭ0 the problem can be analyzed without dif- A 1/2 ficulty. We limit discussion to the case of a particle of an Re Eu DʈϭCͩ ͪ , ͑53͒ easy-axis AFM with an n-fold axis, in the ground state of 2␲ប which the vector l is parallel to the principal axis Cn . For ␻ ϭ␥ where C is a numerical factor of the order of unity. such a magnet 0 Hsf , and the anisotropy energy wa can The structure of the transverse fluctuation determinant be written in the form DЌ is not so universal. For a purely uniaxial AFM model ␻2 ͑␪ ␸͒ϭ␻2 2 ␪ϩ␻2 n ␪ 2͑ ␸ ͒ ͑ ͒ 0wa , 0 sin p sin sin n /2 , 55 with a symmetry axis Cϱ its form is exactly the same as for

Dʈ . Then the square of the standard large parameter where the second term corresponds to the simplest expres- A ប 1/2 151 ͓ ͑ ͒ (Re Eu / ) appears in the splitting see formula 72 sion for the anisotropy in the basal plane, and usually for n ͔ Ͼ ␻ Ӷ␻ below . This result is an example of the implementation of 2 one has p 0 . For such a choice of energy the z and the general rule of working with instantons in systems with x axes are always the easy and medium axes of the AFM, more than one zero mode, according to which each zero respectively. It is easy to see that the instanton solution cor- 143 ␸ϭ␸ ϭ ␪ϭ␪ ␶ mode gives such a factor in the level splitting. If there is a responds to a solution of the form 0 const, ( ). Ӷ ␸ small but finite anisotropy K p Ku in the basal plane x,y, The admissible values of 0 are determined by the relation then instead of this large parameter one will have ͱK /K . w ͑␪,␸͒ץ u p a ͯ ϰ ␸ ϭ ͑ ͒ ␸ sin n 0 0. 56ץ ␸ϭ␸ D. Instantons in uniaxial AFMs 0 ␸(0)ϭ␲ The structure of the instanton is determined by a general There exist 2n solutions of this equation, k k/n, k ϭ Ϫ ␸(0) system of second-order Euler-Lagrange equations for the an- 0,1,...,2n 1, of which n solutions with even k, k,min ␪ ␸ gular variables; this system can be written schematically in give a minimum of wa( , ), and the remaining n with ␸(0) ␪ ␲ the form k,max give a maximum of this function for all 0, . In- ␸(0) -␪ ␸ stantons with k,min correspond to lower values of the Eu ץ wa͑ , ͒ Ϫ␪¨ ϩ␸˙ 2 sin ␪ cos ␪ϩ␻2 ϩi␸˙ ⌫͑␪,␸͒ϭ0, clidean action. The function ␪͑␶͒ can be found from the ␪ץ2 0 second-order equations, for which the first integral ͑48͒ is ␪ ␸ ␪ ץ wa͑ , ͒ known. Taking into account the boundary conditions Ϫ͑␸˙ sin2 ␪͒2ϩ␻2 Ϫi␪˙ ⌫͑␪,␸͒ϭ0, ͑54͒ → ␲ ␶→Ϯϱ ␸ 0, for , we obtainץ2 0 ␪ 2 where a dot denotes a derivative with respect to ␶, and the d 2 ͑0͒ ͩ ͪ ϭ␻ w ͑␪,␸ ͒. ͑57͒ terms with ⌫(␪,␸) are determined by variation of the gyro- d␶ 0 a k scopic terms in the Lagrangian ͑43͒. The dimensionless func- The Euclidean action is real-valued on the trajectories for all tion w (␪,␸) is proportional to the anisotropy energy a values of ␸ and is given by the formula W(␪,␸). The coefficient of proportionality is chosen so as to ␻ ប2␻ ␲ make 0 coincide with the frequency of small oscillations of 0N Ku ␻2 ϭ ប2 A ϭ ͵ d␪ͱw ͑␪,␸͒Ӎ2បsNͱ . ͑58͒ the vector l, which corresponds to 0wa (4JZ/ N)W. Eu 2JZ a JZ ␪ ␺ ⌫ ␪ ␸ 0 The form of wa( , ) and the function ( , ), which is generated by the Dzyaloshinskii interaction or a magnetic Of course, if ⌫(␪,␸) 0 but does vanish at a certain value of ␸(0) field, is discussed below for specific examples. In the case of k,min , corresponding to a minimum of the anisotropy en- ␪ϭ␪ ␶ ␸ϭ␸(0) an AFM with a decompensation of the sublattice spins one ergy, then the ‘‘planar’’ real solution ( ), k ⌫ϭ␥ ␪ ϭ has Hex(Nex /N)sin , i.e., its form is the same as for const is possible. For these planar solutions the contribu- A the Landau–Lifshitz equation. tion of the gyroscopic term to the Euclidean action Eu can The general equations ͑54͒ contain a large number of be purely imaginary ͑examples will be given below͒.Asit independent parameters of various natures that are due to the turns out, however, if the function ⌫(␪,␸) goes to zero for 654 Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov

␸ ͑eff͒ ͑eff͒ץ ␪ ␸ ץ ␸ Hץ Hץ all of the same values of for which wa( , )/ does, r ␸(0) ␸(0) ˜ ϭ B͑␪ ␸͒ Bϭ ͑ ͑eff͒ ͒Ϫ k ϩ i viz., the minima k,min and the maxima k,max , i.e., B 2 , , 2 H •l lil j . ⌫ ␪ ␸ ϰ ␸ r dlk dlj ( , ) sin n , then the contribution of the gyroscopic term ͑60͒ to A is identically zero. In such a case both the instanton Eu For Hϭ0 the formula for B is conveniently written directly solution and the Euclidean action are exactly the same as in terms of the Dzyaloshinskii tensor D : when one is considering only the simplest model, with ij Dץ Dץ ␪ ␸ ϭ ⌫ ( , ) 0. The situation described is not realized for an Bϭ Ϫ ϩ ij Ϫ ij ͑ ͒ l j . 61 ץ l jlk ץ 3Dijlil j Dii AFM in a field, but for those AFMs for which the gyroscopic lk li terms are determined solely by the Dzyaloshinskii interaction The values of B(␪,␸) for different AFMs with the Dzy- it is in no way exotic. It occurs for all AFMs with an even aloshinskii interaction are given in Ref. 118. The imaginary ͑ ͒ according to Turov principal axis, as is confirmed by a part of the Euclidean action Im A for the ␣th instanton is symmetry analysis.118 Eu written in the form of a contour integral ͐A˜ dr. To calculate Analysis of the various AFM models with an external the value of the tunneling splitting one needs to know only magnetic field along the symmetry axis of the crystal or with the phase difference ⌬⌽(␣,␤)ϭ⌽(␣)Ϫ⌽(␤) for all instanton a Dzyaloshinskii interaction in an easy-axis AFM leads to the pairs ͑46͒. The ⌬⌽(␣,␤) are expressed in terms of the integral conclusion that besides the trivial case described above, two ͛ ˜ more situations can be realized.118,151 Thus for an AFM with A•dr over a closed contour consisting of these two trajec- sufficiently high symmetry, three types of behavior of the tories. Using Stokes’ theorem, one can write this quantity as instanton solutions have been established. The trivial case is the flux of the formal magnetic field B˜ introduced above, realized in an AFM with Hϭ0 and an even principal axis. through the part of the unit sphere bounded by that contour. For the other two nontrivial cases the behavior is substan- This leads to the simple expression tially different. ␲ ␸␤ ⌬⌽͑␣,␤͒ϭ ͵ sin ␪d␪ ͵ d␸B͑␪,␸͒, ͑62͒ 1. One nontrivial type of behavior corresponds to the ␸ ⌫ ␪␸(0) ϭ 0 ␣ case when ( k,min) 0 and a planar real instanton is real- ⌫ ␪ ␸(0) where ␸␣ and ␸␤ are the values of the angle ␸ for the given ized but ( , k,max) 0. For this case the value of the real A instantons. Importantly, for all possible forms of the Dzy- part of the Euclidean action Eu does not depend on the gyrotropy parameter ͉H(eff)͉, but a nonzero imaginary part of aloshinskii interaction and for all orientations of the external A ͉ (eff)͉ ˜ Eu appears which is proportional to H . magnetic field H, the total flux of the formal field B through ⌫ ␪ ␸(0) the sphere is equal to zero. We note that here the situation is 2. In the second nontrivial case, when ( , k,min) 0, the planar solution ␸ϭconst is absent. Then one must find a fundamentally different from the case of an uncompensated general solution of system ͑54͒ in the form ␪ϭ␪(␶), ␸ total spin Sex , when A describes the field of a magnetic ϭ␸(␶). This solution is complex and cannot always be con- monopole and the flux through some region of the sphere is ⌽ ϭ ␲ structed exactly. It can be shown, however, that for such proportional to its area, and the total flux tot 4 S.Of complex instantons the imaginary part of the Euclidean ac- course, in the case of an uncompensated magnet, too, the A ϭ A tion does not appear, Im Eu 0, and the real part Re Eu individual phases are determined by A, i.e., they depend on (␣,␤) contains only contributions quadratic in (H(eff))2. the gauge, but the phase difference ⌬⌽ is gauge invari- The latter case will be discussed in detail below for the ant. However, this phase difference is not equal to zero for example of an AFM in magnetic field. Now let us discuss the any trajectories. In particular, for two diametrically opposed first case, when the instanton trajectories are planar and real. trajectories ⌬⌽ϭ2␲S, so that in the spin tunneling problem In the case of real trajectories for calculation of the imagi- one obtains the phase factor cos(␲S) discussed above, and nary part of the Euclidean action one can use the following the tunneling is suppressed for a half-integer uncompensated simple method.117,118 We introduce a vector rϭrl that is not spin. We note that, in accordance with Dirac’s discussion of restricted by the condition r2ϭ1, and we write the term with the single-valuedness of the wave function of an electron in 67,152 ⌽ ϭ the first derivative from ͑43͒ in the form a monopole field, this condition yields cos( tot/2) 0, which leads to a half-integer quantization of the spin. r rϫH͑eff͒ Returning to the case of a compensated AFM, we noteץ Ϫi␥A˜ , where A˜ ϭ . ͑59͒ ͑ ͒ ␶ r2 that formula 62 permits easy calculation of the phase factorץ • K ͑46͒ in formula ͑45͒ for the level splitting. The overall This expression has the same structure as the term in the form of B(␪,␸) for an AFM with an n-fold axis is deter- nonrelativistic Euclidean Lagrangian describing the interac- mined by its symmetry.118 For an AFM with an even prin- (ϩ) tion of a charged particle having coordinate r and velocity ciple axis Cn , i.e., one that does not permute the permu- n vϭdr/d␶ with the formal magnetic field B˜ ϭcurlA˜ . The tation of the sublattices, the function B(␪,␸)ϰsin ␪ sin n␸, magnetic field enters the Lagrangian of a charged particle all the phase factors are equal to zero and there is no inter- ference. through the vector potential A˜ ͑we have introduced the no- Ϫ If the principle axis is odd, C( ) , however, the situation tation A˜ to prevent confusion of this quantity with the n is different for different n.Ifnϭ4k, i.e., nϭ4,8,... ͑for monopole-field vector potential A introduced above͒ at the crystalline AFMs only nϭ4 is meaningful, but for magnetic point r, which is determined only to within a certain gauge, clusters and high-spin molecules values greater than nϭ6 ˜ while at the same time the field B is gauge invariant. In the are also possible͒ the function B(␪,␺) is proportional to (eff) Lagrangian of an AFM for any H this formal magnetic sinn/2␪ sin(n␸/2) or to sinn/2␪ cos(n␸/2), depending on the field B˜ is directed radially and can be written as details of the position of the twofold axes in the basal plane. Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov 655

In this case the integral ͑62͒ can be nonzero, and destructive We recall that if the magnetic field is directed along the 118 Ͼ interference is possible. In the case of an odd axis with easy axis z of the AFM and H Hsf , then a spin-flop tran- ϭ ϩ ϭ ʈ ʈ ͑ n 2(2k 1), i.e., for n 2,6,10,..., in particular for rhom- sition occurs from the state with l ez to a state with l ey see bic and hexagonal AFMs, the function B(␪,␸) is also pro- Sec. 4͒. In this case the influence of the field on the static portional to sin(n␸/2) or cos(n␸/2). However, it contains a properties of the AFM can be described by renormalization n/2␪ ␪ ϭ Ϫ 2 2 factor sin cos , by virtue of which the phase difference of the constant Ku : Ku(H) Ku(1 H /Hsf). For the case ͑ ͒ 62 is identically equal to zero on account of the integral of H parallel to the medium axis ex , the constant Ku in- ␪ ϭ ϩ ␮ 2 over , and destructive interference is absent. creases with increasing field, Ku(H) Ku (g BH) / 2 ϭ Ϫ ␮ 2 Thus the Dzyaloshinskii interaction leads to destructive (4s JZ), while K p decreases, K p(H) K p (g BH) / 2 ͑ ͒ interference in tunneling only for AFMs with an odd princi- (4s JZ). At the characteristic field H p introduced in 64 the (Ϫ) ϭ ϭ ʈ Ͼ pal axis Cn for n 4k, i.e., n 4,8,... . A more important quantity K p(H) changes sign, and for H ex , H H p there is role in this phenomenon is played by the external magnetic a change of character of the axes in the basal plane: the y field. At low field one can assume that the trajectory is planar axis become the medium axis and the x axis the hard axis. and real, and the phase factor can then be written as Finally, if the magnetic field is directed along the hard axis e , then it effectively increases both K and K : K (H) ␲sg␮ HN y u p u,p ⌽ϭ ͩ B ␺ͪ ͑ ͒ ϭK ϩ(g␮ H)2/(4s2JZ), and the field does not influence cos cos cos , 63 u,p B 2JZ the character of the anisotropy axes at all. where ␺ is the angle between the plane in which the instan- If one naively uses the renormalized anisotropy con- ton trajectories lie and the external magnetic field H. The stants to describe the instantons in the framework of Eqs. more interest case of a non-small field is discussed in the ͑65͒͑which leads to incorrect results!͒, one finds that for the A next subsection. orientation of the field along the easy axis the quantity Eu → goes to zero for H Hsf , and for the field direction along A the medium axis the value of Eu on two classes of paths, E. Tunneling for an AFM in magnetic field ␸ϭ ␲ ϩ ␸ϭ␲ ϭ ( /2)(2k 1) and k, become equal at H H p . Let us consider the important case of an AFM in the However, as we shall show below, such a description is ab- presence of a strong magnetic field.51,151,153 For concrete cal- solutely incorrect. Besides the aforementioned renormaliza- culations we shall consider an AFM with rhombic anisotropy tion of the anisotropy, the field alters the character of the ͑ ͒ Ͼ Ͼ of the form 39 , where the constants Ku 0 and K p 0 de- dynamics of the vector l, and that also influences the struc- scribe uniaxial anisotropy and anisotropy in the basal plane, ture of the instantons and the value of the Euclidean action and the z and y axes are the easy and hard axes, respectively. for them. For subsequent applications it is convenient to rewrite this We begin by considering the case of the field parallel to energy in the form used in Eqs. ͑54͒, in terms of the charac- the easy axis z, which does not break the symmetry in the teristic values of the frequency: (x,y) plane. As a consequence, the gyroscopic term in Eqs. ͑ ͒ ␸ ⌫ϰ ␪ ␪ ␻2w ͑␪,␸͒ϭ␻2 sin2 ␪ϩ␻2 sin2 ␪ sin2 ␸, ͑64͒ 54 is independent of the angle : sin cos . In this case, 0 a u p ϭ under the condition K p 0 a purely uniaxial AFM model is ␻ ϭ␥ where u Hsf , Hsf is the field of the spin-flop transition, realized. Analysis of such a model leads to instructive physi- ␻ ϭ␥ ϭ ͱ ␮ 151 p H p , and the characteristic field H p 2s JZKp/g B cal results, and we shall therefore discuss this analysis. In describes the influence of the field on the anisotropy in the the purely uniaxial case the Lagrangian ͑43͒ is independent ␸ץ ץ ␸ ⍀ϭ ͒ ͑ basal plane see below . For this energy in the absence of of and is the Noetherian integral of motion LEu / ˙ , field the instantons corresponding to the condition ␪(ϩϱ) the expression for which can be written in the form ϭ0, ␪(Ϫϱ)ϭ␲ correspond to solutions with a planar rota- 2 tion of l͑␶͒ in the symmetry planes of the system for which ⍀ϭ͑␸˙ Ϫi␥H͒sin ␪ϭconst, ͑68͒ ␸ϭ(␲/2)k, with k an integer, and the problem is integrable. The instanton solutions of the cos ␪ϭtanh ␻¯ ␶, sin ␪ϭ1/cosh ␻¯ ␶, ͑65͒ type ␪→0,␲ for ␶→Ϯϱ correspond to ⍀ϭ0, i.e., ␸ ϭi␥H(␶Ϫ␶ ), where ␶ is an arbitrary constant. When this where ␻¯ is defined by the expression ␻¯ ϭ␻ for ␸ϭ0,␲ and 1 1 u formula is taken into account, the equation for ␪͑␶͒ assumes ␻¯ ϭͱ␻2ϩ␻2 for ␸ϭ(␲/2)(2kϩ1). The values of the Eu- u p the form ␪˙ 2ϭ␻2 sin2 ␪. Hence it is easy to obtain a solution clidean action for all these instantons can be written in a u of the form ͑65͒ with ␻¯ ϭ␻ . For the components of the unified way: u vector l in the instanton the following formulae are obtained: A ϭបN͑ប␻¯ /ZJ͒ for Hϭ0, ͑66͒ Eu ϭ ␻ ␶Ϫ␶ ͒ lz tanh͓ u͑ 0 ͔, and for zero field the minimum of the Euclidean action cor- responds to only one pair of instantons, with ␸ϭ0,␲, i.e., cosh͓␥H͑␶Ϫ␶ ͔͒ sinh͓␥H͑␶Ϫ␶ ͔͒ ϭ 1 ϭ 1 ͑ ͒ with a rotation of the vector l in the easier plane (z,x). lx ␻ ␶Ϫ␶ ͒ , ly i ␻ ␶Ϫ␶ ͒ . 69 cosh͓ u͑ 0 ͔ cosh͓ u͑ 0 ͔ When taking the magnetic field into account, we restrict → discussion to the case when it is directed along one of the This solution has the correct asymptotic behavior (lx,y for ␶→Ϯϱ ␥ Ͻ␻ symmetry axes: z, x,ory. The function ⌫(␪,␸) appearing ) for H u , i.e., at all values of the field smaller 118 in Eq. ͑54͒ for ␪, ␸ can be written in the compact form than the SFT field Hsf . Consequently, the instanton solution found is valid in the entire stability region of the state with ⌫͑␪,␸͒ϭ2␥͑H"l͒sin ␪, ͑67͒ lϭϮzˆ. Despite the fact that in this region the effective an- and it does not equal zero for any orientation of the field. isotropy constant Ku(H) varies from Ku to zero for H 656 Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov

→ ͑ ͒ Hsf , the Euclidean action for solution 69 is independent of H. Its imaginary part is equal to zero, and the real part is the same as for Hϭ0: A͑0͒ϭ ប ͱ ϭ Ͻ ͑ ͒ Eu 2s N Ku /JZ for K p 0, H Hsf . 70 This simple example has shown that the gyroscopic term can substantially suppress the static renormalization of the an- isotropy energy. For this simple example is is possible to do an exact calculation of the pre-exponential factor D in the form ͑52͒.151 The contribution of the two types of deviations, lon- gitudinal and transverse, are determined in this case by iden- ប2 tical operators Mˆ . These operators arise for many tunneling FIG. 8. Real part of the Euclidean action in units of N/2JZ as a function of magnetic field ͑schematic͒. The curves for the field along the easy, me- problems in magnets, and we shall therefore give their form dium, and hard axes are labeled EA, MA, and HA, respectively. The solid 151 A and the complete set of their eigenfunctions: lines represent a consistent calculation of the value of Eu , and the dotted line is the result obtained by a ‘‘naive’’ approach taking into account only d2 2 1 the static renormalization of the anisotropy constants; see text. ˆ ϭϪ Ϫ ϭ ϭ M 2 2 ; Mf0 0, f 0 ; d␰ cosh ␰ & cosh ␰

Ϫik␰ ͑tanh ␰Ϫik͒e Thus for magnetic field directed along the hard axis of the Mˆ f ϭ͑1ϩk2͒f , f ϭ . ͑71͒ k k k 2 A(HA) ͱL͑1ϩk ͒ AFM the real part of the Euclidean action Eu (H) is inde- pendent of the field strength ͑see Fig. 8͒, and the signs of its The total fluctuation factor in this problem is the square of imaginary part for the two instantons differ.51,132 In this case, ͓ ͑ ͔͒ the longitudinal fluctuation determinant Dʈ see Eq. 53 , however, the interference effects and the oscillatory depen- and as a result, the level splitting due to tunneling has the dence of the splitting on field are due not only to the imagi- form nary part of the Euclidean action. In this geometry the most A͑0͒ interesting effects appear on account of the fluctuation deter- ⌬ϭ ប␻ ͩ Eu ͪ ͑ϪA͑0͒ ប͒ ͑ ͒ minant. In this case the equations for small deviations of ␽ C u ␲ប exp Eu / , 72 2 and ␮ from the instanton solution are uncoupled, and D ϭ where C is a numerical factor of the order of unity. The DʈDЌ . The transverse determinant contains the Schro¨- ͱA ប dinger operator with a complex potential. The value of DЌ is additional large parameter Eu / appears in the problem in connection with the fact that the instanton solution ͑69͒ in complex, with different phases for the two different equiva- ␻ ϭ lent instantons, and that modifies the phase shift. Even for the model with p 0 contains two continuous parameters, ␶ ␶ ͑ ͒ moderate values of the field the presence of DЌ leads to a 1 and 0 see Sec. 6.3 . We note also that the level splitting ⌬ is independent of magnetic field, even when the fluctuation phase shift of the oscillatory function ⌬(H)by␲/2 in com- determinant is taken into account. This fact can be justified parison with that caused by the imaginary part of the Euclid- ␻ ean action itself.51,153 on the basis of an exact quantum treatment. Indeed, for p ϭ0 the Hamiltonian of the system commutes with the z In the presence of field the planar solution corresponds tot to rotation of the vector l in the plane perpendicular to the projection of the total spin, Sz . Consequently, the states of totϭ Ϯ direction of the field H. When H is oriented along the hard the system are characterized by a definite value Sz 0, 1, Ϯ2,... . The tunneling splitting of the doublet corresponds to axis the planar instanton clearly has the lowest value of the Stotϭ0, and the magnetic field cannot influence the energies Euclidean action for all values of the field. A planar rotation z ͑ of these states. We note that in these terms the spin-flop in other symmetry planes and the more so for a rotation in a Ͼ plane that is not a symmetry plane of the AFM͒ would cor- transition is due to the circumstance that for H Hsf a level totϾ respond to a nonzero function ⌫͑␪,␸͒ and is therefore impos- with Sz 0 becomes the lowest level. This in essence brings about the quantum analog of the spin-flop transition, which sible. Investigation of the distribution of l requires analysis was discussed in Sec. 4. of a system with two degrees of freedom; although it cannot Let us consider another simple case, when the magnetic be done exactly, it can be done using various approximate 151 field is directed along the hard axis of the AFM ͑the y axis͒. methods. Analysis showed a large diversity in the behav- In this case ⌫(␪,␸)ϭ2␥H sin2␪ sin ␸, and for instanton pairs ior of nonplanar instantons for different orientations of the with ␸ϭ0,␲ it has the value ⌫ϭ0. Consequently, in this field. Let us briefly discuss the results. case there are two planar instanton solutions of the form ␸ As was shown above on the basis of an exact solution, ϭ ϭ0,␲, ␪ϭ␪(␶). These solutions determine the instantons for a field along the easy axis at K p 0 the level splitting due with a rotation of the vector in the easier plane (z,x) and at to tunneling is independent of field up to the field of the the same time have the minimum value of the real part of spin-flop transition. In the presence of anisotropy in the basal A plane, K 0, the value of A (H) falls off with increasing Er . The Euclidean action for them is independent of field p Eu and is given by formula ͑66͒: field: 2 K ␲បg␮ HN ͑ ͒ Ku H A͑HA͒͑ ͒ϭ ប ͱ uϮ B ͑ ͒ A EA ͑H͒ϭ2sបNͱ ͱ1Ϫ␩ , ͑74͒ Eu H 2s N i . 73 Eu 2 JZ 2JZ JZ Hsf Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov 657

͑ ͒ ϭ ϭ FIG. 9. Tunneling splitting of the lower level of an AFM described by the Hamiltonian 39 with spin S 5 and uniaxial anisotropy Ku /J 0.1, for three ϭ ͑ ͒ ͑᭹ ᭺͒ ͑᭡ ᭝͒ different values of the planar anisotropy: K p /Ku 0 – , 0.2 , , and 0.4 , . a—Field parallel to the easy axis; the kinks correspond to the spin-flop transition. b – Field parallel to the hard axis ͑᭺, ᭝͒ and medium axis ͑᭹, ᭡͒.

␩ ␩ Ͻ Ͼ where the coefficient is small at small K p /Ku , field H Hc . For H Hc the tunneling is due to planar in- Ϸ␲ ␩ϭ (K p /Ku); the value 1 that is characteristic for the stantons and goes according to the scenario that is character- case when the influence of the field is taken into account in a istic for the above-described case of the field parallel to the →ϱ͑ ‘‘naive’’ manner is reached only in the limit K p /Ku see hard axis, with an oscillatory dependence of the level split- Fig. 8͒. Thus in a rhombic AFM for field parallel to the easy ting on the field at an almost constant oscillation amplitude. ⌬ axis the tunneling level-splitting (H) increases with in- Calculations in the framework of the instanton approach creasing field, and this effect is more pronounced for larger are in good agreement with the results of a numerical analy- values of K p . sis done by direct diagonalization of the Hamiltonian of a For field direction along the medium axis the situation is spin pair ͑see Fig. 9͒.151 For field along the easy axis in the somewhat more complicated: there can be competition be- purely uniaxial case K ϭ0 the energy difference ⌬ϭE tween different instantons. Here there exists a planar solution p 1 ϪE of the two lowest levels is independent of field up to with a rotation of l in the (z,y) plane, containing the hard 0

A the field of the spin-flop transition. For K p 0 the splitting axis y. For this solution the value of the real part of Eu is larger than in the case of the field along the hard axis ͑73͒, increases with increasing field, and this effect is more pro- where l rotates in the easier plane (z,x), containing the me- nounced for larger values of K p /Ku . In the case of magnetic dium axis x: field parallel to the hard axis and also for a sufficiently large Ͼ field H Hc directed along the medium axis, oscillations of K ϩK ␲បg␮ HN A͑MA,phanar͒͑ ͒ϭ ប ͱ u pϮ B the splitting with variation of the field are observed, with an Eu H 2s N i . JZ 2JZ amplitude that depends weakly on K .51,153 If the field is ͑ ͒ p 75 parallel to the medium axis, then at low fields one observes This value is also larger than for a planar instanton in the decay of the splitting, which then gives way to oscillatory case Hϭ0. The plane solution with rotation of l in the (z,x) behavior, the characteristic value of the transition field in- plane for a field H 0 parallel to the medium axis does not creasing and the amplitude of the oscillations decreasing exist, and at nonzero field it is transformed to a nonplanar with increase of K p /Ku . Granted, there is some quantitative solution. For this nonplanar solution the Euclidean action is disagreement. For example, a numerical analysis demon- A(MA,nonplanar) real, and its value Eu (H) grows with increasing strates that a characteristic field that can be interpreted as the field, and the growth is substantially faster than for the ‘‘na- field of the spin-flop transition exhibits dependence on the ive’’ way of taking the field into account ͑see Fig. 8͒.Asa value of K /K . A slight falloff of the oscillation amplitude result, for a certain critical value H the value of the action p u c with increasing field is also observed for tunneling through for the nonplanar solution becomes equal to the value of the planar instantons at a non-small value of the field H, when real part of the action for the planar solution ͑75͒, which is the pre-exponential factor should be independent of independent of field. At low anisotropy in the basal plane the 51,153 ͑ ͒ field. These effects are absent in the semiclassical de- value of Hc is substantially lower than the value of H p 64 : scription and can be attributed to quantum fluctuations in a ϭ ͑ ␲͒2͑ ͒Ӷ ͑ ͒ Hc H p 2/ K p /Ku H p . 76 system with finite spin. A more detailed comparison of the р The value of Hc remains lower than H p even in the limiting numerical data for different values of the spin S 20 with the →ϱ → case K p /Ku , at which Hc 0.616H p . Thus the nonpla- calculations based on the instanton approach can be found in nar instantons govern the tunneling only at sufficiently low Ref. 151. 658 Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov

␻ → ͑ ͒ ␻ ␻ 6.6. Influence of sublattice decompensation on the For n 0 this gives a finite value 70 , and for p / n → A tunneling 0 a divergence of Re Eu appears, and tunneling is sup- Let us turn to the case of an uncompensated AFM, as- pressed. suming the absence of external field and Dzyaloshinskii in- Analysis of the pre-exponential factor D in this problem ␽ ␦␸ teraction. In the general case this problem is much more reduces to a coupled system of equations for and , i.e., complicated for analysis than the exactly integrable case of a the differential operator Aˆ in ͑52͒ is not diagonal. The cal- pure ferromagnet or the symmetric models of compensated culation showed that the fluctuation determinant is real and ␻ AFMs discussed above when an external field or Dzy- does not influence the interference, and its dependence on n ϰ ͱ␻ 144 aloshinskii interaction is taken into account. However, there is nonanalytic: D 1/ n. is a case in which the system is integrable; for anisotropy of The question of the influence of external magnetic field the form ͑39͒, quadratic in the components of the vector l. on the coherent tunneling has been discussed repeatedly in The Euclidean version of the Lagrangian for this model is connection with the problem of tunneling in ferritin 154 conveniently written in the form particles. For an AFM with decompensated spins a weak field completely lifts the degeneracy of the system, and these ប2N 1 effects are thereby ruled out. In Sec. 4.2 however, in a dis- L ϭ ͫ ͑␪˙ 2ϩ␸˙ 2 sin2 ␪͒Ϫi␻ ␸˙ ͑1Ϫcos ␪͒ Eu 2JZ 2 n cussion of the states of an AFM with uncompensated spins, it was noted that for sufficiently high fields it is possible to 1 ϩ sin2 ␪͑␻2ϩ␻2 sin2 ␸͒ͬ, ͑77͒ have ground-state degeneracy due to the noncollinearity of 2 u p the sublattices. This leads to a new type of macroscopic tun- neling effects when a freely oriented AFM particle with in- by introducing three parameters with dimensions of fre- complete compensation of the spins is found in a strong ex- ␻ ϭ␥ ϭ͉ Ϫ ͉ quency. The parameter n Hex(Sex /Stot), Sex S1 S2 is ternal field.85 In this case the height of the tunnel barrier and ␻ ␻ the excess spin, and the quantities p , u determine the form the phase factor depend substantially on the field strength, of the energy of magnetic anisotropy with easy axis z and making it possible to control the tunneling probability by ͓ ͑ ͔͒ hard axis y see Eq. 64 . For constructing the instanton varying the field. solution it is helpful to recall the form of the Walker solution in real time, which describes a moving domain wall ͑see, e.g., Refs. 12, 34, and 35͒. In this solution the magnetization ␸ϭ␸ ϭ VII. NONUNIFORM STATES IN AFMs: STRUCTURE, is rotated in the plane 0 const, passing through the TOPOLOGICAL ANALYSIS, AND QUANTUM TUNNELING ␸ easy axis, and the value of 0 does not correspond to a plane EFFECTS of symmetry and is determined by the velocity of the wall. The instantons of Lagrangian ͑77͒ also have this property.144 It is a common belief that mesoscopic magnets should be ␸ ␸ ϭ␲ ϩ ␣ In that case the angle 0 is complex, 0 k/2 2i , where typical single-domain particles. However, it has recently k is an integer, been shown that even for extremely small particles of ferro- magnets the distribution of the spins in the ground state can ͱ4␻2␻¯ 2ϩ␻4ϩ␻4Ϫ␻2 ␻2 be nonuniform. For submicron particles of nonellipsoidal ␣ϭ n n p n ␻2ϭ␻2ϩ p cosh 2 , ¯ u . Ͻ ͑ ␻ 2 shape and sizes smaller than the critical, R Rc for Permal- p Շ ͑78͒ loy Rc 100 nm), certain specific weakly nonuniform mag- netic states are realized: the so-called of ‘‘flower’’ or ‘‘leaf’’ ␪͑␶͒ ͑ ͒ 155 Ͼ 156,157 For one has the standard formula of the type 65 , states. For R Rc , topologically nontrivial states ␪ϭ ͓␻ ␸ ␶͔ ␻2 ␸ ϭ␻2 158 cos tanh ¯ ( 0) , with a real parameter ¯ ( 0) u with a magnetic vortex appear which have recently been ϩ␻2 2 ␸ 159 p sin 0 . Here one needs to consider only one pair of studied experimentally. There is now a peak of interest in equivalent instantons, which have even kϭ0,2 and a lower vortex states in magnetic dots1,2 and magnetic rings160 ͑see A 161–164 value of the real part of the Euclidean action Eu . The the recent papers devoted to analysis of the dynamics A imaginary parts of Eu for instantons of this pair have dif- of vortices in such systems͒. ferent signs, and that gives an interference effect with the The dipole interaction is important for FMs and negli- ͉ ␲ ͉ usual phase factor for ferromagnets, cos Sex . gible for AFMs, but it is not the only source of nonunifor- A fundamental feature of uncompensated AFMs is the mities. Dimitrov and Wysin showed165 that the surface ␻ presence of nonanalyticity with respect to the parameter n single-ion anisotropy with local easy axis perpendicular to ϳ ␻ ϭ ͑ ͒ Sex . For a compensated AFM with n 0, equation 78 the surface of the sample at each point of the surface also ␸ ϭ␲ gives the real value 0 k/2, and tunneling is possible leads to a nonuniform nontopological ‘‘flower’’ or ‘‘leaf’’ even in a purely uniaxial model ͓see the exact solution ͑65͔͒. states. The origin of this nonuniformity is clear: in this case ␻ ␻ ϭ A For n 0 and p 0, solutions with finite Eu do not exist, the conditions of minimum energy in the bulk ͑uniform state͒ and for arbitrarily small decompensation it is necessary that and at the surface ͑the spin must be perpendicular to the there be anisotropy in the basal plane. This circumstance is surface͒ are contrary, i.e., frustration occurs. The strong sur- clearly manifested in the nonanalytic dependence of the Eu- face anisotropy limit gives a geometric problem for the dis- ␻ ␻ clidean action on the parameter n / p . In the most charac- tribution of a unit vector perpendicular to the surface of the ␻ ␻ Ӷ␻ teristic case p, n ¯ this dependence has the form object at each point of the surface. The solution of this prob- lem for any simply connected object has a singularity; such ប2 ␻ N n 166 Re A ϭ ͩ 2␻¯ ϩ␻ ln ͪ . ͑79͒ singularities actually arise for crystals and the A Eu n ␻ 3 167 2JZ p phase of superfluid He. Therefore the strong surface an- Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov 659

vector l locally in any region that does not contain the center of the dislocation. This is reminiscent of the procedure used in the theory of elasticity when introducing a strain field for describing a crystal with a dislocation.176,177 The character of the behavior around the singular line can be illustrated by the example of a screw dislocation in a layered AFM of the CoCl2 type, in which the exchange interaction of the spins in the planes is ferromagnetic and stronger than the antiferro- magnetic interaction of neighboring planes.178 The smallness of the latter makes it possible to describe a disclination con- sistently for vectors m, l in the framework of continuum theory.179 Near the dislocation line the ␴-model approxima- tion breaks down, and ͉m͉ϳ͉l͉, and at the very center of the disclination mϭ1 and lϭ0. Recently a general macroscopic approach was proposed for the description of topological de- fects in layered AFMs.180,181 FIG. 10. Spin chain with 9 spins; the thin solid lines are exchange bonds. Thus a disclination in an AFM with a dislocation can be The solid arrows show the direction of the spins for the solution ͑80͒ with treated in the framework of the ␴-model as a singular line of mϭ1; the dotted arrows show the direction of the vector l for the dimers. ͑ The regions of consistent numbering of the sublattices ͑the numbers 1 and 2 the field of the vector l the direction of l is not defined on next to the spins͒ are enclosed by the dotted-and-dashed lines; on the polar the line͒ with a jump of l at the surface, which rests on the (x) axis the consistency is disrupted, and the vector l has a jump. dislocation line and extends to the boundary of the crystal. The distribution of the vector l in a rectilinear disclination sufficiently far from its central part ͑core͒, for rӷa, is de- isotropy can give rise to vortices of the magnetization in both scribed by the formula FMs and AFMs.168 For AFMs specific nonuniform states can be caused by ␪Ӎ␲ ␸ϭ ͒␹ϩ␸ ͑ ͒ /2, ͑m/2 0 , 81 frustration of the exchange interaction. The concept of frus- tration is usually associated with spin glasses, but it is en- where r,␹ are polar coordinates on the plane perpendicular countered for crystalline AFMs if the division of the lattice to the disclination line. This formula, like that for the spin into a finite number of magnetic sublattices cannot satisfy the chain ͑80͒, contains the odd integer mϭϮ1,Ϯ3,... . The dis- condition of antiparallel orientation of neighboring spins; an tribution ͑81͒ is also characteristic for the director vector n of example is an AFM with the triangular lattice.169 The divi- a nematic with a disclination,166 and the num- sion of the AFM into magnetic sublattices is sensitive to the ber m is called the Frank index. The same distribution arises presence of atomic dislocations of the crystal lattice, which for a disclination in the state of a spin nematic, which is disrupt the topology of the atoms planes. As was shown by realized for magnets with a strong biquadratic exchange Dzyaloshinskii170 and also by Kovalev and Kosevich,171 this interaction.75 The simple distribution ͑81͒ is characteristic gives rise to macroscopic magnetic defects—domain walls both for a purely isotropic medium and in the presence of and disclinations; such defects have been observed in thin planar anisotropy, say, for an AFM with an isotropic easy films of chromium.172 For magnetic layered systems, atomic plane (x,y). In the latter case the polar axis for the vector l steps on the FM/AFM interface are also a source of forma- must be chosen along the hard axis z of the AFM. Formula 173,174 ͑ ͒ ӷ ͓ ͑ ͒ ͔ tion of such defects. 81 for large r r0 see Eq. 83 below is also valid for an The appearance of disclinations in a crystal with dislo- antiferromagnetic vortex, but the value of the Frank index cations can be explained by the circumstance that any con- for it is even, mϭϮ2,Ϯ4... .6,34,157 tour drawn along the exchange bonds and encompassing the For the description of the nonuniform states in an AFM axis of the dislocation contains an odd number of sites. and tunneling effects for them it is convenient to use the Therefore the problems of the ground state of an AFM with method of homotopic topology.156 The possibility that singu- dislocations and of a closed spin chain with antiferromag- lar lines of the disclination type, with nonuniformity far from netic interaction and an odd number of sites have a number the line, is due to the continuous degeneracy of the energy of of traits in common. The latter model in the exchange ap- the system with respect to a change of the order parameter in 175 M proximation has an exact solution, and it is helpful to dis- the degeneracy space D of the system. Let us consider the cuss it before considering the disclination. For such an odd behavior of the order parameter ͑the vector l for an AFM or ␪ ϭ␲ chain, all the spins lie in one plane, k /2, in which the n for a nematic͒ in traversing a closed contour ␥ around the ␴ ␸ direction of the kth spin k is determined by the angle k : singular line in coordinate space. Here the order parameter ⌫ M ␲ ␲k describes a closed contour in the degeneracy space D . ␸ ϭ ͓1Ϫ͑Ϫ1͒k͔ϩm , ͑80͒ This contour is a map of the singular line from the standpoint k 2 N of the behavior of the order parameter far from the line. Two ϭϮ Ϯ ⌫ ⌫ where m 1, 3,... is an odd integer, and the minimum of contours 1 and 2 are called homotopic if they can be the energy corresponds to mϭϮ1 ͑see Fig. 10͒. For an odd transformed into each other by a continuous deformation. M chain, as for a crystal with a dislocation, there is no possible The classes of all homotopic contours for a given space D division into sublattices that is consistent for the whole sys- form a group called the fundamental homotopy group of this ␲ M ⌫ ⌫ tem. However, one can introduce sublattices and define the space and is denoted 1( D). The product of 2 and 1 660 Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov

⌫ ⌫ M ϭ 2 corresponds to a contour 2• 1 obtained in traversing first in writing as D S /Z2 . Here Z2 is a group of two ele- ⌫ ⌫ 1 and then 2 . A contour that can be continuously de- ments, 0 and 1, 1  1ϭ0. This condition means that any con- formed into an infinitely small one ͑or, as we say, ‘‘con- tour on the sphere which connects points n and Ϫn can be tracted to a point’’͒ corresponds to the unit element of the assumed closed physically. All such contours can be de- ␲ M ⌫ group. The element of 1( D) that is the inverse of cor- formed into one another, but cannot be contracted to a point; ⌫Ϫ1 ␲ responds to a contour with the opposite direction of they correspond to a nontrivial element of the group 1 , traversal. while all the geometrically closed contours correspond to the Thus from the standpoint of the behavior at infinity, the unit element of this group. Consequently, the homotopy ␲ 2 singular lines of the order parameter are classified by group group 1(S /Z2) consists of two elements and is equivalent elements of the fundamental homotopy group of the degen- to that group Z2 described above. For an isotropic nematic all M ␲ M eracy space D of the system, 1( D), which are called the disclinations with odd Frank indices m can be trans- topological charges. Removable singularities correspond to formed into energetically more favorable ones with m the unit element of the group. When singular lines merge, a ϭϮ1, and the latter are topologically equivalent, while all kind of conservation law for topological charge is obeyed. the disclinations with even Frank indices are topologically When lines corresponding to the contours ⌫ and ⌫Ϫ1 merge, removable. In an isotropic nematic, disclinations with m they give a removable singularity. The mergine of singular ϭ2k can be brought to the ground state ͑formally the state ⌫ ⌫ ϭ lines with charges 1 and 2 give a line with a charge equal with m 0) by a nonplanar deformation, which is descrip- ⌫ ⌫ tively called ‘‘escaping into the third dimension.’’ 156 to the product of the elements 1• 2 . We note that the fun- damental homotopy group can be non-Abelian, but we shall The state of a planar nematic in which n lies in the plane not consider examples of that kind. of the sample is degenerate with respect to a change of n 1͕ 2ϩ 2 It is important to note that the classification of singular along a circle, or the one-dimensional sphere S lx ly ␲ M ϭ1,l ϭ0͖, with allowance for the factorization of the de- lines according to elements of the group 1( D) can be z M ϭ 1 incomplete; singular lines that are equivalent in the sense of generacy space D S /Z2 . In this case all of the different ␲ M values of m correspond to topologically different states of 1( D) cannot always be transformed into each other by a continuous deformation of the field of a vector l2ϭ1. For a the disclination of the nematic, and m plays the role of a ␲ complete description of the singular lines in magnets one 1-topological charge belonging to the group of integers must also take other topological charges into account. We with respect to summation Z. discuss this question below; right now let us turn to the clas- It is now understood what will happen in the case of an sification of vortices or disclinations according the topologi- AFM. For an AFM with a dislocation only those singular ␲ M lines such that the vector l changes sign on going around cal charges of the group 1( D). In the description of spin disclinations in an AFM their them are allowed. This means that in that case there are only similarity to disclinations in nematics is usually emphasized. disclinations with odd values of m. The minimum of the This similarity is undoubtedly productive for the topological energy of an isotropic AFM corresponds two disclinations ϭϩ ϭϪ classification of disclinations. However, there are essential with m 1 and m 1, illustrated in Fig. 11a. They can differences in the properties of disclinations for these ordered be transformed into each other by a rotation in spin space media. The point is that for the same nematic sample the about the x axis, which corresponds to a departure of l from Frank index m in the distribution ͑81͒ takes on any, even or the plane and is a typical example of ‘‘escaping into the third odd, integer values, mϭ0,Ϯ1,Ϯ2... ; mϭ0 corresponds to a dimension.’’ Thus for an isotropic AFM with a dislocation there is a nonuniform ground state, which cannot be charac- uniform distribution. For an AFM the parity of the Frank terized by any sort of discrete degeneracy. For an isotropic index m is wholly determined by the atomic structure of the AFM without a dislocation only even m are allowed, and the sample, which may be assumed to remain unchanged in the ground state corresponds to a uniform distribution of l with course of any changes in the spin structure. For AFMs with a mϭ0. dislocation only odd values of m are admissible in formula For an easy-plane AFM with an isotropic easy plane ͑81͒, while for an AFM with an ideal topology of the atomic (x,y) in the presence of a dislocation the minimum of energy planes either a uniform distribution with mϭ0 or a vortex also corresponds to two disclinations with mϭϮ1, which, with even mϭ2q is possible. unlike the isotropic case, are topologically different. In this In a purely isotropic magnet or a nematic, the state is case the ground state is twofold degenerate, and macroscopic degenerate with respect to changes of the unit vector on the quantum tunneling effects are possible ͑see below͒.Inthe sphere S2͕l2ϭ1͖. Any geometrically closed contour on the easy-plane AFM without a dislocation the only admissible sphere can be contracted to a point, and the fundamental distributions are of the form ͑81͒ with even value of the homotopy group of the sphere ␲ ( 2) consists of only the 1 S Frank index mϭ2q; these correspond to an antiferromag- unit element. Therefore no topologically nontrivial singular netic vortex. The integer qϭm/2, qϭϮ1,Ϯ2,... is called the lines for an order parameter of the spin ␴ type exist in a vorticity. In other words, the vorticity is the ␲ -topological ferromagnet. 1 charge of the vortex, which belongs to the group of integers However, for the description of a nematic it is important Z. One can write an expression for it in terms of an integral that the states with n and Ϫn are physically indistinguish- over a closed contour ␥: able; n is not a physical unit vector of the spin ␴ type but a director vector with an indeterminate sign, as is emphasized ␸ץ 2␲ 1 1 ͒ ͑ by its name. Therefore, for a nematic the degeneracy space is ϭ ͑ ͓ ϫٌ ͔͒ ϭ ␹ q Ͷ ez• l ␣l dx␣ ͵ d . 82 ␹ץ additionally factorized by the sign of n, and that is indicated 2␲ ␥ 2␲ 0 Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov 661

FIG. 11. Distribution of the vector l in an antiferromagnetic disclination associated with an atomic dislocation and the domain wall generated by the disclination; disclinations with values of the Frank index mϭ1 and Ϫ1 and the associated DWs with different chiralities are labeled by ͑ϩ͒ and ͑Ϫ͒ in parts a and b. a—exchange approximation; short distances. The line of dots shows a line of formal discontinuity of the vector l; the shaded area shows the central part, of the order of atomic size, in which the vector l cannot be uniquely introduced. b—The influence of anisotropy at distances greater than the width of the wall; the shaded region shows a central part corresponding approximately to the ‘‘exchange’’ region, where the distribution of l is as shown in part a.

Far from the vortex one has ␪ϭ␲/2, but at the center of the enough away from the core of the disclination, a domain wall vortex, unlike a disclination, there is no singularity. This is ͑DW͒ is formed; see Fig. 11b. The static two-dimensional ϭ possible when l is collinear with the unit vector along the DW corresponds to the solution l l0(x) with a rotation of ␪ϭ ␲ hard axis, ez , i.e., 0, . Thus at a given vorticity q we the vector l in the easy plane (x,y): have two different states of the vortex, the structures of which are described by the solutions ␵ ϭ␵ ͑ ͒ ϭ ϭ ͑ ͒ lx 0 tanh x/x0 , ly ͑ ͒ , lz 0, 84 ␸ϭ ␹ϩ␸ ␪ϭ␪͑Ϯ͒͑ ͒ ␪Ϯ͑ ͒ϭ ϭϮ ͑ ͒ cosh x/x0 q 0 , 0 r , cos 0 0 p 1. 83 (Ϯ) Ӎ 1/2 ӷ The functions ␪ (r) differ from ␲/2 in a vortex-core region where x0 (J/K p) a a is the DW thickness, and the num- ␵ ␵ϭϮ ␵ of diameter r0 , which, for a slightly anisotropic AFM, has bers 0 , 1. The value of 0 determines the values of lx ϭ 1/2ӷ the value r0 a(J/Ku) a. The two distributions of l with at two points far from the center of the DW. These two points ␲ the same values of the 1 topological charge q can differ by can be called a zero-dimensional sphere, by virtue of which ␲ ␴ ␲ the values of the other topological invariant, the so-called 2 the value of 0 has the meaning of the 0 topological charge ␵ topological charge, which corresponds to mapping not to of the DW. The change of 0 is due to the overcoming of a contours but to two-dimensional manifolds. In the case of a potential barrier proportional to the size of the system ͑for- 2D magnet we are talking about a mapping of the (x,y) mally, an infinite barrier͒, and tunneling cannot occur. The 2 2 ␲ plane onto the sphere S ͕l ϭ1͖. This charge is determined second topological charge of the type 1 is due to the map- ͓ ͕ 2 by the integral topological invariant Q introduced above Eq. ping of the line perpendicular to the DW onto the circle lx ͑ ͔͒ → ϩ 2ϭ ϭ ͖ ͑ ͒ 42 with the substitution t y. For a vortex the value of Q ly 1, lz 0 . This charge q is defined by the integral 82 ϱ is half-integer, QϭϪqp/2, where pϭϮ1 determines the in which ͛␥dx␣ is replaced by ͐Ϫϱdx; for a DW it has the ͑ ͒ ϭϪ␵ ␵ ␨ϭϮ sign of lz in the core of the vortex 83 . Vortex states with value q 0 /2. Consequently, the DW chirality 1, ϭϮ ␨ Q 1/2 differ topologically and cannot be transformed into which at fixed 0 determines the direction of rotation of l in ␲ each other by a continuous deformation. Consequently, the going along the chain, is the 1 topological charge of the ␲ ϭϮ vortex is characterized by two topological charges: the 1 DW. On different sides of a vortex with q 1, two such charge—the vorticity q—and the p2 charge, the polarization walls, with different signs of the chirality, appear. Such DWs pϭϮ1. This is manifested in the analysis of the merging of with a vortex are observed in thin films of ferromagnets and two vortices with vorticities q and Ϫq: if the polarizations are called walls with a vertical Bloch line.6,182 of the vortices are different, one obtains a localized ͑trivial Tunneling effects presuppose the presence of several dif- ␲ from the standpoint of 1) soliton with an integer value of ferent but energetically equivalent states of the system sepa- ␲ ϭϮ the 2 topological charge Q q, while if the polarizations rated by a finite potential barrier. For topological nonunifor- are of the same sign p, one obtains a topologically trivial mities the question arises of tunneling-induced change of the state. signs of the topological charges. A change of the ‘‘main’’ Let us study the transformation of the structure of vorti- charge such as the vorticity q for a vortex in an infinite ces and disclinations in the presence of weak anisotropy in system involves the overcoming of an infinite barrier. In that the easy plane (x,y). Let us consider a rhombic AFM with case tunneling-induced change of ‘‘auxiliary’’ topological ͑ ͒ Ͻ anisotropy energy of the form 39 , assuming that Ku 0 and charges—the chirality for a DW or the polarization for a Ͼ K p 0, i.e., z is the hard axis and x is the easy axis in the vortex—is possible. In small particles tunneling of a ‘‘main’’ easy plane. When the anisotropy in the basal plane is taken charge such as the Frank index for a disclination is also into account, the distribution of the vector l will no longer possible. The problems of the tunneling-induced change of have the simple form ͑81͒ or ͑83͒. For regions found far the chirality of a DW and tunneling of the topological charge 662 Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov

soliton of the vortex type, or, more precisely, its anisotropic analog—a vertical Bloch line in a two-dimensional DW, for which lϭl(x,y).6,182 Often the similarity of instantons and the static nonuniformities is extremely constructive and per- mits one to predict the character of the behavior of l in non- one-dimensional instantons and to estimate the tunneling am- plitude. An example is the simple case of non-one-dimensional instantons in an AFM with Heffϭ0. In this case the vector l in the solution will be real. For a Dϩ1-dimensional instanton describing tunneling for a D-dimensional nonuniform distri- bution l͑x͒ the problem of finding the real part of the Euclid- A ϭបW ean action Re Eu ͓l͔ reduces to minimization of the real dimensionless functional W͓l͔: dxDϩ1 1 FIG. 12. Structure of the instanton describing the tunneling transition be- W͓l͔ϭ ͵ ͫ Ja2s2ٌ͑l͒2ϩW͑l͒ͬ, ͑85͒ tween states of the DW with opposite chiralities ͑the top and bottom in the aDប 2 figure͒. The thick arrows denote the projection of the vector l on the easy Dϩ1 D plane; the solid lines denote the curves on which l makes a 45° angle to the where dx ϭdx d␶,ٌ denotes the gradient operator in the hard axis ͑it is perpendicular to the plane of the figure͒. These lines show the ϩ ϭ ␶ D 1-dimensional Euclidean space (x,x0), x0 c , and the DW boundaries and also the core region at the center of the vortex, which is function W͑l͒ is given by the same expression as the usual characteristic for an out-of-plane magnetic vortex. The vortex shown has ͑ ͒ polarization pϭ1, i.e., the direction of l͑0,0͒ is upward. static energy per spin for a magnet 37 . The imaginary part of the Euclidean action for real l can only be due to the topological term and is manifested only for Dϭ1. It is important no note that in the FM case such a situa- of a disclination in an easy-plane AFM are extremely similar, tion does not arise: instanton solutions are never real, and the and it is convenient to discuss them together. analysis of non-one-dimensional instantons in a ferromagnet We begin with an analysis of the tunneling of chirality is much more complicated. The problem of tunneling of the for a DW of the form ͑84͒ in quasi-one-dimensional AFMs DW chirality in a ferromagnet has been discussed by many of the spin-chain type or systems of coupled spin chains authors.63,183–187 It differs from the AFM case not only in forming a mesoscopic wire or ribbon.124 For the description that the corresponding instanton solution ␴ϭ␴(x,␶) is com- of the tunneling dynamics we use the Lagrangian ͑40͒ for a plex. It is also important that the DW chirality in a ferromag- nonuniform distribution of the vector l ͓more precisely, its net is directly related to the DW momentum. For example, Euclidean version for lϭl(x,␶)] with the topological term nonmoving DWs with different values of the chirality have ͑42͒ and anisotropy ͑39͒ included. The state of the wall ͑84͒ different values of the momentum ͑see the discussion of this is degenerate with respect to the value of the DW chirality, unusual property in Ref. 69͒. Therefore, for a free ͑with no and a tunneling transition between the two DW states with pinning͒ DW in a ferromagnet, the momentum of the DW is ␨ϭϮ1 can occur. For its description we must construct the conserved, and tunneling of the chirality is altogether two-dimensional instanton solution of the equations for l impossible.186 ϭl(x,␶). Its solution should give l→Ϯe for x→Ϯϱ, and x Now let us return to the AFM case. The construction of go over to two DWs with different chiralities for ␶→Ϯϱ. a two-dimensional vortex instanton solution describing the These conditions lead to the situation that in going along the tunneling of chirality in a DW can be done by minimization boundary of a sufficiently large large part of the Euclidean of the functional ͑85͒, which upon the substitution x→x, plane (x,␶) ͑see Fig. 12͒, the vector l rotates by an angle c␶→y corresponds formally to the well-studied energy of a 2␲qϭϮ2␲ in the easier plane (x,y) of the AFM. In this vertical Bloch line in a two-dimensional DW on the (x,y) case, for continuity of the solution it is necessary that at plane for a ferromagnet with rhombic anisotropy.6,182 The some point the vector l be perpendicular to the easy plane. value of the instanton action can easily be found by analogy That point is naturally called the center of the instanton and with the energy of this state. In the limiting cases of weak placed at the origin of coordinates in the Euclidean plane, and strong basal-plane anisotropy of W͓l͔, an estimate for xϭ0, ␶ϭ0, and at it l (0,0)ϭϮ1. Thus the instanton con- z the instanton solution W͓l ͔ϵW is given by the figuration has the properties of an out-of-plane magnetic vor- 0 0 formula124,188 tex for which the vorticity qϭ1 and the polarization pϭϮ1 ͑see Fig. 12͒. Thus for tunneling of the DW chirality 2␲ ln͑2.2ͱ␳͒, ␳ӷ1, W ϭs ͭ ͑86͒ there are two equivalent instanton paths having different 0 • ␳1/2 ␳Ӷ ϵ ␪ ϭ ϭϮ 8 , 1, signs of the polarization lz(0,0) cos (0,0) p 1. The instanton corresponds to the separatrix solution of a where we have introduced the rhombicity parameter ␳ ϭ ␳Ӷ partial differential equation with a finite value of the Euclid- K p /Ku . The limiting case of small rhombicity 1 can ean action ͑43͒. It is by no means always possible to con- be investigated in the effective one-dimensional instanton struct even one-dimensional instanton solutions. However, a approximation.188 We emphasize that, although the tunneling sufficiently complete analysis of the problem can be carried process involves a macroscopically large number of spins, ϭ ӷ out on the basis of the formal similarity between the two- N0 x0 /a 1, this number does not appear in the tunneling dimensional instanton of interest to us, with lϭl(x,␶), and a exponent. In the case ␳Ӷ1 the value of the tunneling expo- Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov 663

FIG. 13. Structure of two instanton paths determining the tunneling transi- tion between states of a disclination with opposite values of the Frank index. These two paths have the opposite signs of the topological invariant Q and of the imaginary part of the Euclidean action.

nent for one spin chain Sϳ1 is small, and the dynamics of the internal degrees of freedom of the domain wall is not 123,189 ␳Ӷ FIG. 14. Structure of the instanton determining the tunneling transition be- semiclassical but essentially quantum. For 1, one tween states of a magnetic vortex with opposite polarizations. The sphere can speak of macroscopic tunneling, which requires satisfac- around the origin of coordinates, on which a singular structure of the W у ‘‘hedgehog’’ type is shown, represents the region where the description of tion of the condition 0 1, only for a mesoscopic sample containing a large number of chains. In the case of strong the AFM in terms of a unit vector l is impossible. rhombic anisotropy the tunneling of the chirality is semiclas- sical even for a DW in a quasi-one-dimensional AFM.124 W (DW)ϭ ͑ ͒ For a two-dimensional instanton of the vortex type there (R/a)W0 , where W0 is given by formula 86 , in- is also an imaginary part of the Euclidean action, which is creases linearly as a function of the DW length R. For two determined by the topological term ͑42͒ in the latter. For the different paths the imaginary part of the Euclidean action is A ϭ ␲ ϭ ␲ two instanton paths Im Eu 2 sQ s qp, which for determined by the topological term ͑42͒; it is equal to ϭϮ A ϭϮ␲ (chain) (chain) p 1 gives Im Eu s. For half-integer spin of an Ϯ␲sN , where N is the number of atomic chains atom of the chain the interference of these two instanton that enclose the center of the disclination. For AFM displace- paths leads to suppression of tunneling. This can be ex- ments with half-integer spin s and an odd number of chains, plained by the fact that a DW in an antiferromagnetic spin tunneling is suppressed in the absence of magnetic field. This chain has a spin equal to the value of the atomic spin result can be explained by starting from Kramers’ theorem s.123,189,190 For a sample consisting of several coupled chains and noticing that the total number of spins in a sample with with half-integer spin, tunneling is forbidden for an odd an odd number of chains and an odd number of atoms in a number n of such chains. chain is necessarily odd. The behavior of the imaginary part of the instanton ac- For vortices in two-dimensional AFMs there arises the ␲ tion and the overall picture of the tunneling in an AFM is question of whether tunneling can affect the 2 topological fundamentally altered in the presence of even a weak exter- charge of the vortex: the polarization p.192 The vortex core ӷ ͒ nal magnetic field H. The magnetic field gives an additional region of diameter r0 a (a is the interatomic distance con- ͓ 2ϳ contribution to the imaginary part of the Euclidean action, tains a large of the order of (r0 /a) J/Ku] number of and the value of the tunneling splitting is an oscillatory func- spins, the direction of which changes in the quantum transi- tion of magnetic field.124 tion between vortex states with pϭϩ1 and Ϫ1. The corre- Let us consider the change of the Frank index of a dis- sponding three-dimensional instanton solution lϭl(x,y,␶)is clination as a result of tunneling in a small 2D AFM particle real; it inevitably contains a singular point of the hedgehog of radius R ͑an island of a quasi-monoatomic film͒.30,191 For type at the center of the instanton ͑see Fig. 14͒. Analysis of ␳ϭ ϳ ϳ simplicity we shall assume that K p /Ku 1 and x0 r0 . the behavior of the AFM near this singularity must go be- Ӷ For a small parameter R x0 a DW does not form, and there yond the framework of the standard ␴-model. It can be done is a smooth distribution of spins of the type ␸ϭϮ␹/2. In qualitatively in the same way as for a Bloch point in a this case there are also two instanton paths, the structure of ferromagnet.182 Estimates shown that the main contribution which is illustrated in Fig. 13. An estimate of the exponential to the Euclidean action is given not by the singularity but by factor gives a peripheral region with size of the order of that of the vortex ͑part͒ϭ␲ ͑ 2 ͒ ͑ ͒ core, r0 . The value of the factor in the tunneling exponent is W0 s R /ax0 . 87 192 W (vort)ϳ obtained in the form 0 (r0 /a). As in the case of This value is much smaller than the total number of spins in tunneling in a DW, this value is smaller than the number of (part)ϭ␲ 2 (vort)ϳ 2ϳ the particle, N (R/a) . For a disclination in a particle spins taking part in the process, N s(r0 /a) J/Ku , (part) of size R the factor W0 increases with increasing R up to but it nevertheless contains the large factor r0 /a ϳ ϳ ӷ ϳ 1/2 the value R x0 r0 at which a DW forms. For R x0 tun- (J/Ku) . Therefore, tunneling of the vortex polarization neling occurs within a DW, and the factor in the exponent, is possible in a quasi-two-dimensional AFM with sufficiently 664 Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov large anisotropy. It is important to emphasize that the vortex does not admit generalization to systems with large spin, and (vort) structure and, consequently, the value of W0 is deter- for a long time it remained unclear to what extent this result mined solely by the parameters of the AFM, and for an en- is applicable for AFMs with Sϭ1/2. At first glance it can be semble of vortices ͑as for high-spin molecules͒ there is no shown that for large spins or systems of the spin-ladder type spread of the parameters characterizing the tunneling. the quantum effects should be weaker, but reality turned out to be much more interesting. Haldane showed that for Sϭ1 VIII. CONCLUSION and all integer spins the long-range order is completely de- stroyed, the correlators fall off exponentially, and the spec- When an article is finished, one can better see what trum of excitations contains a finite gap ͑the so-called questions, for objective or subjective reasons, have been dis- Haldane gap ⌬ϳបc/␭).114,115 Zvyagin obtained a similar re- cussed in less than adequate detail or left out completely. sult for an exactly solvable model of coupled spins chains ͑in Let us list the questions that have been left out of this modern parlance, a spin ladder͒ with spin Sϭ1/2: the long- review but which would benefit from a brief discussion in 197 order to get a better idea of the range of problems that we range order is destroyed for an even number of chains. In have discussed. In this review we have discussed both purely the case of half-integer spin, for a chain or an odd number of ϭ classical aspects of AFM physics and problems of quantum coupled chains with spin S 1/2 there is quasi-long-range tunneling for small particles or clusters of AFMs. In the latter order with a gapless dispersion relation. These results for spin ladders have been confirmed by other methods,102,199–201 case it turned out that quantum tunneling effects in essence 198 violate the initial premise of the ‘‘classical’’ part of this re- including direct numerical simulation. For this review it is view. Indeed, the classical description is based on the Nee´l important that such results can be obtained on the basis of ␴ 113,202,203 picture of an AFM, which includes the concepts of magnetic the -model both for a spin chain and for a spin 200,201 sublattices and, consequently, the existence of an antiferro- ladder. The analysis scheme here is essentially the magnetic vector l. On the other hand, tunneling in an AFM same as that used for describing quantum tunneling in AFM essentially amounts to a change in the sign of l as a result of particles ͑see Sec. 6͒. The Ne´el state and the ␴-model for a tunneling, i.e., destruction of the Ne´el order. For small AFM 1D AFM are taken as starting points of the analysis, and then particles the intensity of these effects is proportional to the the quantum fluctuations in this state are investigated. The number of spins in the system, while tunneling is possible consistent incorporation of fluctuations, including non-small only for extremely small samples. ones, ‘‘rectifies’’ the problem that coherent states are not ex- One of the important results in AFM physics in the twen- act quantum states of the AFM Hamiltonian. The destruction tieth century is the establishment of the fact that non-small of the long-range order in this approach can be described on quantum fluctuations can break the Ne´el order even in infi- the basis of the instanton formalism. The corresponding in- nite systems, in particular, in quasi-one-dimensional AFMs. stantons are determined by two-dimensional solutions of the An example is the spin chain and spin ladders with two or type discussed in Sec. 7 for the problem of nonuniform tun- more legs. This striking and significant area of AFM physics neling. A remarkable property of these effects is that the is often called ‘‘quantum magnetism.’’ The author certainly difference of integer and half-integer spins enters the prob- had not planned any detailed discussion of this topic in this lem through the phenomenon of interference of instanton review. However, it is useful to at least outline the range of paths, and the imaginary part of the Euclidean action is de- topics and to cite references to the main publications. Fur- termined by the topological term ͑42͒͑see the review by thermore, since we have used models of the spin-ladder type Affleck116 for more details͒. in this review, it is necessary to discuss at least briefly the This brief discussion shows that not only is the use of degree to which the classical approach is applicable to the the classical ␴-model consistent with the presence of non- description of these materials, which are commonly associ- small quantum fluctuations, it is actually the most universal ated with the quantum theory of AFMs. and convenient formalism for investigating them. Simple in- Effects of non-small quantum fluctuations involving the tuitive arguments about the possible instanton paths and es- destruction of Ne´el order were first revealed in the analysis timates of the Euclidean action for them make it possible to of the exact solution of the quantum problem of the states of understand directly the role of the quantum destruction of the an isotropic spin chain with spin Sϭ1/2, constructed by Be- Ne´el long-range order. In particular, it is quite difficult to the in 1931193 and then generalized by Baxter for a chain picture the three-dimensional lϭl(x,y,␶) instanton of this with the same spin Sϭ1/2 but with an anisotropic nearest- type and the realization of similar effects in the 2D case. The neighbor interaction.194 For Sϭ1/2 chains there is no Ne´el instanton approach also lets one draw conclusions about the order even at zero temperature, and the correlators F(x) exponential suppression of tunneling effects for non-small ϵ͗l(0)l(x)͘ fall off in a power-law manner with increasing spins, the Haldane gap ⌬ϰexp(Ϫ2␲S). Recent numerical distance x between spins. Such a state is customarily called a simulations have established that even for spin Sϭ2, which critical state or a state with quasi-long-range order. ͑It cannot is next to the minimum Haldane spin Sϭ1, a disordered be called purely disordered, since that would require that the phase with a finite gap exists only for extremely small an- correlators have exponential behavior, Fˆ (x)ϰexp(Ϫx/␭), isotropy and is suppressed by a weak magnetic field.204 where ␭ is the correlation length.͒ Quasi-long-range order is Therefore it can be expected that the application of the clas- also characteristic for the Berezinskii phase, which exists in sical version of the ␴-model to two-dimensional magnets of two-dimensional easy-plane magnets at nonzero temperature the manganese halide type (Sϭ5/2) or even to spin ladders Ͻ T TBKT , where TBKT is the Berezinskii–Kosterlitz– with ions of this type in connection with the spin-flop tran- Thouless transition temperature.195,196 The Bethe method sition in Sec. 4 is entirely adequate. Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov 665

The spin order in materials with an antiferromagnetic 20R. P. Cowburn, D. K. Koltsov, A. O. Adeyeye, and M. E. Welland, J. interaction is not limited to the simplest Nee´l structure. The Appl. Phys. 87, 7082 ͑2000͒. 21 ͑ ͒ author’s only regret is that it was not possible to discuss the R. P. Cowburn and M. E. Welland, Science 287, 1466 2000 . 22S. O. Demokritov, J. Phys. D 31,925͑1998͒. nematic phases of non-Heisenberg magnets, frustrated AFMs 23K. Yu. Guslienko, Appl. Phys. Lett. 75, 394 ͑1999͒; K. Yu. Guslienko, S. or AFMs with noncollinear spin order, especially since such Choe, and S. Shin, ibid. 76, 3609 ͑2000͒. spin structures are known not only for crystalline AFMs but 24R. W. Wang, D. L. Mills, E. E. Fullerton, J. E. Mattson, and S. D. Bader, ͑ ͒ also for the unique magnetic molecules ͕Mo Fe ͖.57 How- Phys. Rev. Lett. 72, 920 1994 . 70 30 25R. W. Wang, D. L. Mills, E. E. Fullerton, S. Kumar, and M. Grimsditch, ever, the author hopes that this review, in which a wide class Phys. Rev. B 53, 2627 ͑1996͒. of problems, both static and dynamic, classical and quantum, 26S. Rakhmanova, D. L. Mills, and E. E. Fullerton, Phys. Rev. B 57,476 are discussed from a unified point of view for the simplest ͑1998͒. 27 two-sublattice AFMs will be useful to the reader. D. D. Awschalom, J. F. Smyth, G. Grinstein, D. P. DiVincenzo, and D. Loss, Phys. Rev. Lett. 68, 3092 ͑1992͒. This review could never have been written without the 28E. M. Chudnovsky and J. Tejada, Macroscopic Quantum Tunneling of the collaboration of many colleagues and discussions with them Magnetic Moment, Cambridge University Press ͑1998͒. of the questions addressed in it. In closing I express deep 29L. Gunter and B. Barbara ͑eds.͒, Quantum Tunneling of Magnetization, ͑ ͒ gratitude to V. G. Bar’yakhtar, V. E. Kireev, A. K. Kolezhuk, Vol. 301 of NATO ASI Series E, Kluwer, Dordrecht 1995 . 30B. A. Ivanov and A. K. Kolezhuk, in Frontiers in Magnetism of Reduced A. M. Kosevich, A. S. Kovalev, N. E. Kulagin, and C. E. Dimension Systems, V. G. Bar’yakhtar, P. E. Wigen, and N. A. Lesnik Zaspel for fruitful collaboration and helpful discussions. I am ͑eds.͒, Vol. 49 of NATO ASI Series 3. High Technology, Kluwer, Dor- also indebted to E. G. Galkina for a long collaboration and drecht ͑1998͒. 31 ͑ ͒ ͑ ͒ invaluable assistance on this review article. L. N. Leuenberger and D. Loss, Nature London 410,789 2001 ;W. Wernsdorfer, N. Aliaga-Alcalde, D. N. Hendrickson, and G. Christou, ibid. 416, 406 ͑2002͒. 32D. Loss, D. P. DiVincenzo, and G. Grinstein, Phys. Rev. Lett. 69, 3232 ͑1992͒. *E-mail: [email protected] 33J. von Delft and C. L. Henley, Phys. Rev. Lett. 69, 3236 ͑1992͒. 34A. M. Kosevich, B. A. Ivanov, and A. S. Kovalev, Nonlinear Waves of Magnetization. Dynamical and Topological Solitons ͓in Russian͔, 1W. Wernsdorfer, Adv. Chem. Phys. 118,99͑2001͒. Naukova Dumka, Kiev ͑1983͒. 2R. Skomski, J. Phys.: Condens. Matter 15, R841 ͑2003͒. 35A. M. Kosevich, B. A. Ivanov, and A. S. Kovalev, Phys. Rep. 194,117 3A. S. Borovik-Romanov, Antiferromagnetism ͓in Russian͔, Vol. 4 of Itogi ͑1990͒. Nauki, Fiz.-Mat. Nauki, Ya. G. Dorfman, ed., Izd-vo AN SSSR, Moscow 36H. J. Mikeska and M. Steiner, Adv. Phys. 40,191͑1991͒. ͑1962͒. 37V. G. Bar’yakhtar and B. A. Ivanov, ‘‘Solitons and Thermodynamics of 4E. A. Turov, Physical Properties of Magnetically Ordered Crystals, Aca- Low-Dimensional Magnets,’’ Soviet Scientific Reviews, Section A. Phys- demic Press, New York ͑1965͒, Izd-vo AN SSSR, Moscow ͑1963͒. ics ͑I. M. Khalatnikov, ed.͒, Vol. 16 ͑1992͒. 5K. P. Belov, A. K. Zvezdin, A. M. Kadomtseva, and R. Z. Levitin, Ori- 38B. A. Ivanov and A. K. Kolezhuk, Fiz. Nizk. Temp. 21, 355 ͑1995͓͒Low entational Transitions in Rare-Earth Magnets ͓in Russian͔, Nauka, Mos- Temp. Phys. 21, 275 ͑1995͔͒. cow ͑1979͒. 39L. Ne´el, Ann. Phys. ͑Paris͒ 17,61͑1932͒. 6V. G. Baryakhtar, M. V. Chetkin, B. A. Ivanov, and S. N. Gadetskii, 40L. Ne´el, Ann. Phys. ͑Paris͒ 5, 232 ͑1936͒. Dynamics of Topological Magnetic Solitons. Experiment and Theory, Vol. 41L. D. Landau, Phys. Zs. Sowjetunion 4,675͑1933͒. 139 of Springer Tracts in Modern Physics, Springer-Verlag, Berlin 42L. D. Landau and E. M. Lifshitz, Electrodynamics of Continuous Media, ͑1994͒. 1st ed., Pergamon Press, Oxford ͑1960͒, Gostekhizdat, Moscow ͑1957͒. 7A. G. Gurevich, Magnetic Resonance in Ferrites and Antiferromagnets 43A. F. Andreev and V. I. Marchenko, Usp. Fiz. Nauk 130,39͑1980͓͒Sov. ͓in Russian͔, Nauka, Moscow ͑1973͒; A. G. Gurevich and G. A. Melkov, Phys. Usp. 23,21͑1980͔͒. 44 ´ ͑ ͓͒ Magnetic Oscillations and Waves, CRC Press, Boca Raton, Florida I. E. Dzyaloshinski, Zh. Eksp. Teor. Fiz. 32, 1547 1957 Sov. Phys. ͑1996͒, Nauka, Moscow ͑1994͒. JETP 5, 1259 ͑1957͔͒. 8V. V. Eremenko, N. F. Kharchenko, Yu. G. Litvinenko, and V. M. Nau- 45T. Moriya, Phys. Rev. 120,91͑1960͒. menko, Magneto-Optics and Spectroscopy of Antiferromagnets, Springer- 46Jurgen Schnack, ‘‘Quantum theory of molecular magnetism,’’ in Quantum Verlag, New York ͑1992͒, Naukova Dumka, Kiev ͑1989͒. Magnetism, Vol. 645 of Lecture Notes in Physics, Springer-Verlag, 9E. A. Turov, A. V. Kolchanov, V. V. Men’shenin, I. F. Mirsaev, and V. V. Heidelberg ͑2004͒. Nikolaev, Symmetry and the Physical Properties of Antiferromagnets ͓in 47H. De Raedt, S. Miyashita, K. Michielsen, and M. Mashida, Phys. Rev. B Russian͔, Fizmatlit, Moscow ͑2001͒. 70, 064401 ͑2004͒. 10 ͑ ͒ 48 S. V. Vonsovski, Magnetism, Vols. 1 and 2, Wiley, New York 1974 , D. Gatteschi, A. Caneschi, L. Pardi, and R. Sessoli, Science 265, 1054 Nauka, Moscow ͑1971͒. ͑1994͒. 11G. Rado and H. Suhl ͑eds.͒, Magnetism, in four volumes, Academic Press, 49A. Cornia, A. G. M. Jansen, and M. Affronte, Phys. Rev. B 60, 12177 New York ͑1963͒. ͑1999͒. 12V. G. Bar’yakhtar and B. A. Ivanov, Intermediate State and the Dynamic 50O. Waldmann, R. Koch, S. Schromm, P. Mu¨ller, I. Bernt, and R. W. and Static Properties of Domain Walls in Two-Sublattice Magnets, Vol. 6 Saalfrank, Phys. Rev. Lett. 89, 246401 ͑2002͒. of Sov. Sci. Rev. Sec. A–Phys., I. M. Khalatnikov ͑ed.͒, Harwood, Am- 51A. Chiolero and D. Loss, Phys. Rev. Lett. 80, 169 ͑1998͒. sterdam ͑1985͒,p.404. 52F. Meier and D. Loss, Phys. Rev. Lett. 86, 5373 ͑2001͒; Phys. Rev. B 64, 13E. A. Turov, Kinetic, Optical, and Acoustic Properties of Antiferromag- 224411 ͑2001͒. nets ͓in Russian͔, Izd-vo UrO RAN, Sverdlovsk ͑1990͒. 53G. Christou, D. Gatteschi, D. N. Hendrickson, and R. Sessoli, MRS Bull. 14E. A. Turov and V. G. Shavrov, Usp. Fiz. Nauk 140,429͑1983͓͒Sov. 25,66͑2000͒. Phys. Usp. 26,593͑1983͔͒. 54W. Wernsdorfer, S. Bhaduri, C. Boskovic, G. Christou, and D. N. Hen- 15V. G. Bar’yakhtar, B. A. Ivanov, and M. V. Chetkin, Usp. Fiz. Nauk 146, drickson, Phys. Rev. B 65, 180403 ͑2002͒. 417 ͑1985͓͒Sov. Phys. Usp. 28,563͑1985͔͒. 55R. S. Edvards, S. Hill, S. Bhaduri, N. Aliaga-Alcalde, E. Bolin, S. Mac- 16V. Novosad, Y. Otani, A. Ohsawa, S. G. Kim, K. Fukamichi, J. Koike, K. cagnano, G. Christou, and D. N. Hendrickson, Polyhedron 22,1911 Maruyama, O. Kitakami, and Y. Shimada, J. Appl. Phys. 87, 6400 ͑2000͒. ͑2003͒. 17S. O. Demokritov, B. Hillebrands, and A. N. Slavin, Phys. Rep. 348, 441 56D. Caneschi, L. Pardi, A. L. Barra, A. Mu¨ller, and J. Do¨ring, Nature ͑2001͒. ͑London͒ 354, 463 ͑1991͒. 18S. O. Demokritov, J. Phys.: Condens. Matter 15, 2575 ͑2003͒. 57A. Mu¨ller, S. Sarkar, S. Q. N. Shah, H. Bo¨gge, M. Schmidtmann, Sh. 19M. N. Baibich, J. M. Broto, A. Fert, F. Nguyen Van Dau, F. Petroff, P. Sarkar, P. Ko´gerler, B. Hauptfleisch, A. X. Trautwein, and V. Schu¨n- Eitenne, G. Greuzet, A. Friederich, and J. Chazeles, Phys. Rev. Lett. 61, emann, Angew. Chem., Int. Ed. Engl. 38, 3238 ͑1999͒. 2472 ͑1988͒. 58J. M. Luttinger and L. Tisza, Phys. Rev. 70,954͑1946͒; P. I. Beloboro- 666 Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov

dov, R. C. Gekht, and V. A. Ignatchenko, Sov. Phys. JETP 57, 636 ͑1983͒; 95C. Micheletti, R. B. Griffiths, and J. M. Yeomans, J. Phys. A 30, L233 J. G. Brankov and D. M. Danchev, Physica A 144, 128 ͑1987͒; S. Prakash ͑1997͒. and C. L. Henley, Phys. Rev. B 42, 6574 ͑1990͒. 96N. Papanicolaou, J. Phys.: Condens. Matter 10, L131 ͑1998͒. 59T. Moriya, Spin Fluctuations in Itinerant Electron Magnetism, Springer- 97J. de Jongh and A. R. Miedema, Adv. Phys. 23,1͑1974͒. Verlag, Berlin ͑1985͒, Mir, Moscow ͑1988͒. 98K. Subbaraman, C. E. Zaspel, and J. E. Drumheller, Phys. Rev. Lett. 80, 60J. M. Radcliffe, J. Phys. A 4,313͑1971͒. 2201 ͑1998͒. 61A. M. Perelomov, Usp. Fiz. Nauk 123,23͑1977͓͒Sov. Phys. Usp. 20, 703 99C. E. Zaspel and J. E. Drumheller, Int. J. Modern Phys. 10, 3649 ͑1996͒. ͑1977͔͒; A. Perelomov, Generalized Coherent States and Their Applica- 100B. A. Ivanov, C. E. Zaspel, and A. Yu. Merkulov, J. Appl. Phys. 89, 7198 tions, Springer-Verlag, Berlin ͑1986͒. ͑2001͒. 62H. Kuratsuji and T. Suzuki, J. Math. Phys. 21, 472 ͑1980͒. 101B. A. Ivanov, C. E. Zaspel, and A. Yu. Merkulov, Phys. Rev. B 68, 212403 63S. Takagi and G. Tatara, Phys. Rev. B 54,9920͑1996͒. ͑2003͒. 64L. D. Landau and E. M. Lifshitz (Lifshits), Quantum Mechanics: Non- 102A. K. Kolezhuk and H. J. Mikeska, Phys. Rev. B 53, R8848 ͑1996͒. Relativistic Theory, transl. of 3rd ed., Pergamon Press, Oxford ͑1977͒, 103A. K. Kolezhuk and H. J. Mikeska, Int. J. Mod. Phys. B 12,2325͑1998͒. Nauka, Moscow ͑1989͒. 104Y. Shapira and V. Bindilatti, J. Appl. Phys. 92, 4155 ͑2002͒. 65R. J. Glauber, Phys. Rev. 130, 2529 ͑1963͒; 131, 2766 ͑1963͒. 105O. Waldmann, Phys. Rev. B 61, 6138 ͑2000͒. 106 66A. M. Perelomov, Commun. Math. Phys. 26,222͑1972͒. B. A. Ivanov and V. E. Kireev, JETP Lett. 81, 321 ͑2005͒. 107 ´ 67E. Fradkin, Field Theories of Condensed Matter Systems, Vol. 82 of Fron- M. I. Kaganov and V. M. Tsukernik, Zh. Eksp. Teor. Fiz. 34,106͑1958͒ tiers in Physics, Addison-Wesley ͑1991͒. ͓Sov. Phys. JETP 7,73͑1958͔͒. 108 68R. Rajaraman, Solitons and Instantons: An Inroduction to Solitons and E. A. Turov and Yu. P. Irkhin, Izv. AN SSSR Ser. Fiz. 22, 1168 ͑1958͒. 109 Instantons in Quantum Field Theory, North-Holland, Amsterdam ͑1982͒. I. V. Bar’yakhtar and B. A. Ivanov, Fiz. Nizk. Temp. 5, 759 ͑1979͓͒Sov. 69E. G. Galkina and B. A. Ivanov, JETP Lett. 71, 259 ͑2000͒. J. Low Temp. Phys. 5, 361 ͑1979͔͒. 110 70L. D. Landau and E. M. Lifshitz, ‘‘On the theory of the magnetic perme- H.-J. Mikeska, J. Phys. C 13, 2913 ͑1980͒. 111 ´ ͑ ͒ ability of ferromagnetic ,’’ see L. D. Landau, Collected Works ͓in B. A. Ivanov and A. L. Sukstanski, Zh. Eksp. Teor. Fiz. 84,370 1983 ͓ ͑ ͔͒ Russian͔, Nauka, Moscow ͑1969͒, Vol. 1, p. 128. Sov. Phys. JETP 57, 214 1983 . 112 ͑ ͒ 71A. I. Akhiezer, V. G. Bar’yakhtar, and S. V. Peletminskii, Spin Waves, B. A. Ivanov and A. L. Sukstanskii, Solid State Commun. 50,523 1984 . 113 ͑ ͒ ͑ ͒ North-Holland, Amsterdam ͑1968͒, Nauka, Moscow ͑1967͒. I. Affleck and I. Affleck, Nucl. Phys. B 257, 397 1985 ; 265,409 1986 ; ͑ ͒ 72E´ . L. Nagaev, Magnets with Nonsimple Exchange Interactions ͓in Rus- Phys. Rev. Lett. 57, 1048 1986 . 114 ͑ ͒ sian͔, Nauka, Moscow ͑1988͒. F. D. M. Haldane, Phys. Lett. A 93, 464 1983 . 115 ͑ ͒ 73V. S. Ostrovski , Zh. E´ ksp. Teor. Fiz. 91, 1690 ͑1986͓͒Sov. Phys. JETP F. D. M. Haldane, Phys. Rev. Lett. 50, 1153 1983 . 116 ͑ ͒ 64,999͑1986͔͒. I. Affleck, J. Phys.: Condens. Matter 1, 3047 1989 ; I. Affleck, in Fields, ͑ ͒ 74V. M. Loktev and V. S. Ostrovski , Fiz. Nizk. Temp. 20, 983 ͑1994͓͒Low Strings and Critical Phenomena,E.Bre´zin and J. Zinn-Justin eds. , ͑ ͒ 20 ͑ ͔͒ North-Holland, Amsterdam 1990 , p. 567. Temp. Phys. ,775 1994 . 117 75B. A. Ivanov and A. K. Kolezhuk, Phys. Rev. B 68, 052401 ͑2003͒. B. A. Ivanov, in Path Integrals from peV to TeV: 50 Years after Feynman’s 76B. A. Ivanov, A. N. Kichizhiev, and Yu. N. Mitsa , Zh. E´ ksp. Teor. Fiz. Paper, R. Casalbuoni, R. Giachetti, V. Tognetti, R. Vaia, and P. Verrucchi ͑ ͒ ͑ ͒ 102, 618 ͑1992͓͒JETP 75,329͑1992͔͒. eds. , World Scientific, Dordrecht 1999 , p. 410. 118 ´ ͑ ͓͒ 77L. N. Bulaevski and V. L. Ginzburg, Zh. E´ ksp. Teor. Fiz. 45, 772 ͑1963͒ B. A. Ivanov and V. E. Kireev, Zh. Eksp. Teor. Fiz. 121, 320 2002 JETP ͑ ͔͒ ͓Sov. Phys. JETP 18, 530 ͑1964͔͒. 94, 270 2002 . 119 ´ 78M. R. Mathews, B. P. Anderson, P. C. Halljan et al., Phys. Rev. Lett. 83, E. V. Gomona, B. A. Ivanov, V. A. L’vov, and G. K. Oksyuk, Zh. Eksp. ͑ ͓͒ ͑ ͔͒ 2498 ͑1999͒. Teor. Fiz. 97, 307 1990 Sov. Phys. JETP 70, 174 1990 . 120 ͑ ͒ 79K. W. Madison, F. Chevy, W. Wohlleben, and J. Dalibard, Phys. Rev. Lett. A. Chiolero and D. Loss, Phys. Rev. B 56,738 1997 . 121 14 84,806͑2000͒. B. A. Ivanov, A. K. Kolezhuk, and G. K. Oksyuk, Europhys. Lett. ,151 ͑1991͒. 80Fei Zhou, ‘‘Quantum spin nematic states in Bose-Einstein condensates,’’ 122I. V. Bar’yakhtar and B. A. Ivanov, Zh. E´ ksp. Teor. Fiz. 85,328͑1983͒ Electronic preprint ArXiv:cond-mat/0108473 ͑2002͒. ͓Sov. Phys. JETP 58, 190 ͑1983͔͒. 81N. J. Poulis and G. E. G. Hardeman, Physica ͑Utrecht͒ 18, 201 ͑1952͒; 123B. A. Ivanov and A. K. Kolezhuk, Fiz. Nizk. Temp. 21, 986 ͑1995͓͒Low 18,315͑1952͒. Temp. Phys. 21, 760 ͑1995͔͒. 82S. Foner, in Magnetism, Vol. 1, G. Rado and H. Suhl ͑eds.͒, Academic 124B. A. Ivanov, A. K. Kolezhuk, and V. E. Kireev, Phys. Rev. B 58, 11514 Press, New York ͑1963͒. ͑1998͒. 83V. A. Popov and V. I. Skidanenko, ‘‘,’’ Tr. FT- 125E. M. Chudnovski , Zh. E´ ksp. Teor. Fiz. 77, 2158 ͑1979͓͒Sov. Phys. INT AN USSR, Kharkov, vyp. 7, 49 ͑1970͒; G. K. Chepurnykh, Fiz. JETP 50, 1035 ͑1979͔͒. Tverd. Tela ͑Leningrad͒ 10,1917͑1968͓͒Sov. Phys. Solid State 10, 1517 126E. M. Chudnovsky and L. Gunter, Phys. Rev. B 37, 9455 ͑1988͒. ͑1968͔͒. 127M. Enz and R. Schilling, J. Phys. C 19, L711 ͑1986͒. 84 ͑ ͓͒ V. A. L’vov and D. A. Yablonski, Fiz. Nizk. Temp. 8,951 1982 Sov. J. 128J. L. van Hemmen and A. Su¨to¨, Physica B 141,37͑1986͒. Low Temp. Phys. 8, 479 ͑1982͔͒. 129 ͑ ͒ 85 B. Barbara and E. M. Chudnovsky, Phys. Lett. A 145,205 1990 . B. A. Ivanov and V. E. Kireev, JETP Lett. 69, 398 ͑1999͒. 130 86 I. V. Krive and O. B. Zaslavskii, J. Phys.: Condens. Matter 2, 9457 V. G. Bar’yakhtar, A. E. Borovik, and V. A. Popov, JETP Lett. 9,391 ͑1990͒. ͑ ͔͒ 1969 ; V. G. Bar’yakhtar, A. A. Galkin, S. M. Kovner, and V. A. Popov, 131A. Garg, Europhys. Lett. 22, 205 ͑1993͒. ´ ͑ ͓͒ ͑ ͒ Zh. Eksp. Teor. Fiz. 58,494 1970 Sov. Phys. JETP 31, 264 1970 ;K. 132V. Y. Golyshev and A. F. Popkov, Europhys. Lett. 29,327͑1995͒;V.Yu. ´ L. Dudko, V. V. Eremenko, and V. I. Fridman, Zh. Eksp. Teor. Fiz. 61, Galyshev and A. F. Popkov, Zh. E´ ksp. Teor. Fiz. 108, 1755 ͑1995͓͒JETP ͑ ͓͒ ͑ ͒ 1553 1971 ibid. 34,828 1971 ; A. King and D. Paquete, Phys. Rev. 81, 962 ͑1995͔͒. Lett. 30,662͑1972͒. 133J. R. Friedman, M. P. Sarachik, J. Tejada, and R. Ziolo, Phys. Rev. Lett. 87 V. G. Bar’yakhtar, A. N. Bogdanov, and D. A. Yablonski, Usp. Fiz. Nauk 76, 3830 ͑1996͒. 156,47͑1988͓͒Sov. Phys. Usp. 31, 810 ͑1988͔͒. 134L. Thomas, F. Lionti, R. Ballou, D. Gatteschi, R. Sessoli, and B. Barbara, 88 D. L. Mills, Phys. Rev. Lett. 20,18͑1968͒. Nature ͑London͒ 383,145͑1996͒. 89F. Keffer and H. Chow, Phys. Rev. Lett. 31, 1061 ͑1973͒. 135C. Sangregorio, T. Ohm, C. Paulsen, R. Sessoli, and D. Gatteschi, Phys. 90D. L. Mills and W. Saslow, Phys. Rev. B 171, 488 ͑1968͒. Rev. Lett. 78, 4645 ͑1997͒. 91W. E. Tennant, R. B. Bailey, and P. L. Richards, in Proceedings of the 136S. M. J. Aubin, N. R. Dilley, M. W. Wemple, M. B. Maple, G. Christou, Conference on Magnetism and Magnetic Materials, San Francisco, 1974, and D. N. Hendrickson, J. Am. Chem. Soc. 120, 839 ͑1998͒. C. D. Graham, G. H. Lander, and J. J. Rhyne ͑eds.͒, AIP Conf. Proc. 137W. Wernsdorfer and R. Sessoli, Science 284, 133 ͑1999͒. Number 24, American Institute of Physics, New York ͑1975͒. 138W. Wernsdorfer, M. Soler, G. Christou, and D. N. Hendrickson, J. Appl. 92E. E. Fullerton, M. J. Conover, J. E. Matson, C. H. Sowers, and S. D. Phys. 91, 7164 ͑2002͒. Bader, Phys. Rev. B 48, 15755 ͑1993͒. 139W. Wernsdorfer, S. Bhaduri, C. Boskovic, G. Christou, and D. N. Hen- 93F. C. Nortemann, R. L. Stamps, A. S. Carrico, and R. E. Camley, Phys. drickson, Phys. Rev. B 65, 180403R ͑2002͒. Rev. B 46, 10847 ͑1992͒. 140O. Waldmann, C. Dobe, H. Mutka, A. Furrer, and H. U. Gu¨del, ‘‘Meso- 94L. Trallori, P. Politi, A. Rettoti, M. G. Pini, and J. Willain, Phys. Rev. Lett. scopic quantum coherence in antiferromagnetic molecular clusters,’’ Elec- 72, 1925 ͑1994͒; J. Phys.: Condens. Matter 7, L451 ͑1995͒. tronic Preprint cond-mat/0410447 ͑2004͒. Low Temp. Phys. 31 (8–9), August–September 2005 B. A. Ivanov 667

141E. Kececioglu and A. Garg, Phys. Rev. Lett. 88, 237205 ͑2002͒; Phys. 168V. E. Kireev and B. A. Ivanov, Phys. Rev. B 68, 104428 ͑2003͒. Rev. B 67, 054406 ͑2003͒. 169R. S. Gekht, Usp. Fiz. Nauk 59, 261 ͑1989͒. 142 ͑ ͒ 170 ͑ ͒ E. M. Chudnovsky and X. M. Hidalgo, Europhys. Lett. 50, 395 2000 . I. E. Dzyaloshinski, JETP Lett. 25,414 1977 . 143 171 A. I. Vanshten, V. I. Zakharov, V. A. Novikov, and M. A. Shifman, Usp. A. S. Kovalev and A. M. Kosevich, Fiz. Nizk. Temp. 3,259͑1977͓͒Sov. Fiz. Nauk 136, 553 ͑1982͓͒Sov. Phys. Usp. 25, 195 ͑1982͔͒. J. Low Temp. Phys. 3, 125 ͑1977͔͒. 144B. A. Ivanov and V. E. Kireev, Fiz. Nizk. Temp. 25, 1287 ͑1999͓͒Low 172M. Kleiber, M. Bode, R. Ravlic´, and R. Wiesendanger, Phys. Rev. Lett. Temp. Phys. 25,966͑1999͔͒. 85, 4606 ͑2000͒. 145A. I. Perelomov, Integrable systems of classical mechanics and Lie alge- 173R. L. Stamps and R. E. Camley, Phys. Rev. B 54, 15200 ͑1996͒;R.L. bras, Birkha¨user, Basel ͑1990͒, Nauka, Moscow ͑1990͒. Stamps, R. E. Camley, and R. J. Hicken, ibid. 54, 4159 ͑1996͒. 146 C. Neumann, J. Reine Angew. Math. 56,46͑1859͒. 174P. Bodeker, A. Hucht, J. Borchers, F. Guthoff, A. Schreyer, and H. Zabel, 147 ´ ͑ ͒ V. M. Eleonski and N. E. Kulagin, Zh. Eksp. Teor. Fiz. 84, 616 1983 Phys. Rev. Lett. 81, 914 ͑1998͒; H. Zabel, J. Phys.: Condens. Matter 11, ͓Sov. Phys. JETP 57, 356 ͑1983͔͒. 9303 ͑1999͒. 148 ´ V. M. Eleonski, N. N. Kirova, and N. E. Kulagin, Zh. Eksp. Teor. Fiz. 77, 175B. A. Ivanov, V. E. Kireev, and V. P. Voronov, Fiz. Nizk. Temp. 23,845 ͑ ͓͒ ͑ ͔͒ 409 1979 Sov. Phys. JETP 50,209 1979 . ͑1997͓͒Low Temp. Phys. 23, 635 ͑1997͔͒. 149 E. K. Sklyanin, ‘‘On complete integrability of the Landau-Lifshitz equa- 176L. D. Landau and E. M. Lifshitz, Theory of Elasticity, 2nd Eng. ed., ͑ ͒ tion,’’ Preprint LOMI E-3-79 1979 . Pergamon Press, Oxford ͑1970͒, Nauka, Moscow ͑1965͒. 150 ´ ͑ ͒ B. A. Ivanov and N. E. Kulagin, Zh. Eksp. Teor. Fiz. 126, 1479 2004 177A. M. Kosevich, Dislocations in the Theory of Elasticity, Naukova ͓ ͑ ͔͒ JETP 99, 1291 2004 . Dumka, Kiev ͑1978͒. 151 ͑ ͒ B. A. Ivanov and V. E. Kireev, Phys. Rev. B 70, 214430 2004 . 178D. C. Wiesler, H. Zabel, and S. M. Shapiro, Z. Phys. B: Condens. Matter 152 ͑ ͒ R. Jackiw, Commun. Nucl. Part. Phys. 13, 141 1984 ; Usp. Fiz. Nauk 93, 277 ͑1994͒. 149 ͑ ͒ , 137 1986 . 179B. A. Ivanov and V. E. Kireev, Pis’ma Zh. E´ ksp. Teor. Fiz. 73,210͑2001͒ 153H. Hu, R. Lu¨, J.-L. Zhu, and J.-J. Xiong, Commun. Theor. Phys. 36, 245 ͓JETP 46, 188 ͑2001͔͒. ͑2001͒; Electronic Preprint cond-mat/0006292 ͑2000͒. 180A. S. Kovalev, Fiz. Nizk. Temp. 20, 1034 ͑1994͓͒Low Temp. Phys. 20, 154A. Garg, Phys. Rev. Lett. 71, 4249 ͑1993͒; 74, 1458 ͑1995͒. 815 ͑1994͔͒. 155N. A. Usov and S. E. Peschany, J. Magn. Magn. Mater. 130, 275 ͑1994͒; 181O. K. Dudko and A. S. Kovalev, Fiz. Nizk. Temp. 24,559͑1998͓͒Low M. Grimsditch, Y. Jaccard, and I. K. Shuller, Phys. Rev. B 58, 11539 Temp. Phys. 24,422͑1998͔͒; Fiz. Nizk. Temp. 25,25͑1999͓͒Low Temp. ͑1998͒; R. P. Cowburn and M. E. Welland, ibid. 58, 9217 ͑1998͒;R.P. Phys. 25,18͑1999͒; Fiz. Nizk. Temp. 26, 821 ͑2000͓͒Low Temp. Phys. Cowburn, A. O. Adeyeye, and M. E. Welland, Phys. Rev. Lett. 81, 5415 26, 603 ͑2000͔͒. ͑1998͒, E. V. Tartakovskaya, J. W. Tucker, and B. A. Ivanov, J. Appl. 182 Walls in Phys. 89, 8348 ͑2001͒; B. A. Ivanov and E. V. Tartakovskaya, Zh. E´ ksp. A. P. Malozemoff and J. C. Slonczewski, ͑ ͒ Teor. Fiz. 125, 1159 ͑2004͓͒JETP 98, 1015 ͑2004͔͒. Bubble Materials, Academic Press, New York 1979 , Mir, Moscow ͑ ͒ 156G. E. Volovik and V. P. Mineev, Zh. E´ ksp. Teor. Fiz. 72, 2256 ͑1977͒ 1982 . 183 ͑ ͒ ͓Sov. Phys. JETP 45,1186͑1977͔͒. H. B. Braun and D. Loss, Phys. Rev. B 53, 3237 1996 . 184 ͑ ͒ 157B. A. Ivanov, H. J. Schnitzer, F. G. Mertens, and G. M. Wysin, Phys. Rev. H. B. Braun and D. Loss, Int. J. Mod. Phys. B 10, 219 1996 . 185 ͑ ͒ B 58,8464͑1998͒. H. B. Braun, J. Kyriakidis, and D. Loss, Phys. Rev. B 56, 8129 1997 . 186 ͑ ͒ 158N. A. Usov and S. E. Peschany, J. Magn. Magn. Mater. 118, L290 ͑1993͒. J. Shibata and S. Takagi, Phys. Rev. B 62,9920 2000 . 187 ͑ ͒ 159K. Runge, T. Nozaki, U. Okami et al., J. Appl. Phys. 79, 5075 ͑1996͒;R. J. A. Freire, Phys. Rev. B 65, 104436 2002 . 188 ͑ ͒ P. Cowburn, D. K. Koltsov, A. O. Adeyeye et al., Phys. Rev. Lett. 83, B. A. Ivanov and A. K. Kolezhuk, JETP Lett. 60, 805 1994 . 189 1042 ͑1999͒; A. Ferna´ndez and C. J. Cerjan, J. Appl. Phys. 87, 1395 B. A. Ivanov and A. K. Kolezhuk, Phys. Rev. Lett. 74, 1859 ͑1995͒. 190 ͑2000͒; Jing Shi, S. Tehrani, and M. R. Scheinfein, Appl. Phys. Lett. 76, N. Papanicolaou, Phys. Rev. B 51, 15062 ͑1995͒. 191 2588 ͑2000͒; T. Pokhil, D. Song, and J. Nowak, J. Appl. Phys. 87, 6319 B. A. Ivanov, V. E. Kireev, and A. Yu. Merkulov, Mater. Sci. Forum ͑2000͒. 373–376, 807 ͑2001͒. 160M. Kla¨ui, C. A. F. Vaz, L. Lopez-Diaz, and J. A. C. Bland, J. Phys.: 192E. G. Galkina and B. A. Ivanov, JETP Lett. 61,511͑1995͒. Condens. Matter 15, R985 ͑2003͒. 193H. J. Bethe, Z. Phys. 71, 205 ͑1931͒. 161P. Park, P. Eames, D. M. Engebretson, J. Berezovsky, and P. A. Crowell, 194R. J. Baxter, Ann. Phys. 70, 323 ͑1972͒. Phys. Rev. B 67, 020403R ͑2003͒. 195V. L. Berezinskii, Sov. Phys. JETP 34,610͑1972͒. 162M. Buess, R. Ho¨llinger, T. Haug, K. Perzlmaier, U. Krey, D. Pescia, M. R. 196J. M. Kosterlitz and D. J. Thouless, J. Phys. C 6, 1181 ͑1973͒. Scheinfein, D. Weiss, and C. H. Back, Phys. Rev. Lett. 93, 077207 197A. A. Zvyagin, Sov. Phys. Solid State 32, 181 ͑1990͒;Sov.J.LowTemp. ͑2004͒. Phys. 18, 558 ͑1992͒. 163S. B. Choe, Y. Acremann, A. Scholl, A. Bauer, A. Doran, J. Sto¨hr, and H. 198E. Dagotto, J. Riera, and D. J. Scalapino, Phys. Rev. B 45, 5744 ͑1992͒. A. Padmore, Science 304,420͑2004͒. 199D. V. Khveshchenko, Phys. Rev. B 50,380͑1994͒. 164B. A. Ivanov and C. E. Zaspel, Phys. Rev. Lett. 94, 027205 ͑2005͒. 200D. Senechal, Phys. Rev. B 52, 15319 ͑1995͒. 165D. A. Dimitrov and G. M. Wysin, Phys. Rev. B 50, 3077 ͑1994͒; 51, 201G. Sierra, J. Phys. A 29, 3299 ͑1996͒. 11947 ͑1995͒. 202H. J. Schulz, Phys. Rev. B 34, 6372 ͑1986͒. 166P.-G. de Gennes, The Physics of Liquid Crystals, Clarendon Press, Oxford 203M. den Nijs and K. Rommelse, Phys. Rev. B 40, 4709 ͑1989͒. ͑1974͒, Mir, Moscow ͑1974͒. 204U. Schollwock and Th. Jolicoeur, Europhys. Lett. 30,493͑1995͒;U. 167N. D. Mermin, in Quantum Fluids and Solids, S. B. Trickey, E. D. Ad- Schollwock, O. Golinelli, and T. Jolicoeur, Phys. Rev. B 54, 4038 ͑1996͒. ams, and J. W. Duffey ͑eds.͒, Plenum Press, New York ͑1977͒,p.3;P.W. Anderson and R. G. Palmer, ibid.,p.23. Translated by Steve Torstveit LOW TEMPERATURE PHYSICS VOLUME 31, NUMBERS 8–9 AUGUST–SEPTEMBER 2005

Quantum Heisenberg antiferromagnets: a survey of the activity in Florence „Review… U. Balucani

Istituto dei Sistemi Complessi, Consiglio Nazionale delle Ricerche, via Madonna del Piano 10, I-50019 Sesto Fiorentino (FI), Italy L. Capriotti

Valuation Risk Group, Credit Suisse First Boston (Europe) Ltd., One Cabot Square, London, E14 4QJ, United Kingdom and Kavli Institute for Theoretical Physics, University of California, Santa Barbara, California 93106, USA A. Cuccoli, A. Fubini, and V. Tognetti

Dipartimento di Fisica dell’Universita` di Firenze, via G. Sansone 1, I-50019 Sesto Fiorentino (FI), Italy and Istituto Nazionale per la Fisica della Materia, Unita` di Ricerca di Firenze, via G. Sansone 1, I-50019 Sesto Fiorentino (FI), Italy T. Roscilde

Department of Physics and Astronomy, University of Southern California, Los Angeles, CA 90089-0484, USA R. Vaia

Istituto dei Sistemi Complessi, Consiglio Nazionale delle Ricerche, via Madonna del Piano 10, I-50019 Sesto Fiorentino (FI), Italy and Istituto Nazionale per la Fisica della Materia, Unita` di Ricerca di Firenze, via G. Sansone 1, I-50019 Sesto Fiorentino (FI), Italy P. Verrucchi

Istituto Nazionale per la Fisica della Materia, Unita` di Ricerca di Firenze, via G. Sansone 1, I-50019 Sesto Fiorentino (FI), Italy and Istituto dei Sistemi Complessi, Consiglio Nazionale delle Ricerche, via Madonna del Piano 10, I-50019 Sesto Fiorentino (FI), Italy ͑Submitted January 20, 2005͒ Fiz. Nizk. Temp. 31, 885–906 ͑August–September 2005͒ Over the years the research group in Florence ͑Firenze͒ has produced a number of theoretical results concerning the statistical mechanics of quantum antiferromagnetic models, which range from the theory of two-magnon Raman scattering to the characterization of the phase transitions in quantum low-dimensional antiferromagnetic models. Our research activity was steadily aimed to the understanding of experimental observations. © 2005 American Institute of Physics. ͓DOI: 10.1063/1.2008128͔

I. INTRODUCTION While these peculiar features of antiferromagnetism al- The Heisenberg model may well be considered the cor- ready occur in three-dimensional ͑3D͒ compounds, they are nerstone of the modern theory of magnetic systems; the rea- more pronounced in the low-dimensional ones, where other son for such an important role is the simple structure of the effects caused by the enhanced role of classical and quantum Hamiltonian, whose symmetries underlie its peculiar fea- fluctuations are present, and exotic spin configurations asso- tures. The basic forces which determine the alignment of the ciated with field theory models can appear. Indeed, the last spins are represented by the exchange integrals J. At vari- two decades have seen a renewed interest both in the case of ance with the ferromagnet, where the parallel alignment is the one-dimensional ͑1D͒ quantum Heisenberg antiferromag- promoted, in the antiferromagnet a lot of peculiar arrange- net ͑QHAF͒, for which a peculiar behavior of the ground ments of the spins can occur, with strong differences between state versus spin value was predicted,1 and of the two- classical and quantum systems. As a matter of fact, even for dimensional ͑2D͒ QHAF, because of its theoretically chal- nearest-neighbor antiferromagnetic interactions only the lenging properties and of the fact that it models the magnetic ground state of the Hamiltonian is different from the Ne´el behavior of the parent compounds of some high-Tc state with antialigned spins, and the ͑staggered͒ magnetiza- superconductors.2,3 The experimental activity on 2D antifer- tion shows the so-called spin reduction with respect to the romagnets stems from the existence of several real com- saturation value also at Tϭ0. The linear excitations of an pounds whose crystal structure is such that the magnetic ions antiferromagnet can be roughly associated in two families, form parallel planes and interact strongly only if belonging and pair excitations with vanishing total magnetization are to the same plane. As a consequence of such structure, their possible: the fact that the total momentum of them can be magnetic behavior is indeed 2D down to such low tempera- close to zero allows for their investigation by light scattering. tures that the weak interplane interaction becomes relevant,

1063-777X/2005/31(8–9)/18/$26.00668 © 2005 American Institute of Physics Low Temp. Phys. 31 (8–9), August–September 2005 Balucani et al. 669 driving the system towards a 3D ordered phase. scattering is expected to be spread over a band of frequencies In addition, the 2D Heisenberg model can be enriched in antiferromagnets. However, the strongly through symmetry-breaking terms—we considered easy-axis enhances the contribution of zone-boundary ͑ZB͒ ͑ ͒ ͑ ͒ 5 ϳ EA and easy-plane EP anisotropy, as well as an external excitations, i.e., at k kZB . uniform magnetic field—which are useful for reproducing While in ferromagnets the two-spin process is only due the experimental behavior of many layered compounds. In to a second-order mechanism, orders of magnitude smaller the EA case one is left with a chain with discrete reflection than the first order one, in antiferromagnets a different inde- symmetry, and the system undergoes an Ising-like phase pendent process is permitted, stronger than the correspond- transition. In the EP case or when a magnetic field is applied ing one for single-spin spectra.6 Specifically, an exchange the residual O͑2͒ symmetry prevents finite-temperature mechanism does not change the total z component of the ordering,4 but vortex excitations are possible and determine a spins: exciting two magnons in the two different sublattices Berezinskii-Kosterlitz-Thouless ͑BKT͒ transition between a (⌬Mϭ0)7 is the dominant scattering process. paramagnetic and a quasi-ordered phase. In spite of the tiny The one-spin Raman scattering peak disappears at the anisotropies of real systems ͑usually Շ0.01J), it can be Ne´el temperature because it probes the smallest wave vec- shown that they dramatically change the behavior of the spin tors, related with the long-range correlations. In contrast, array already at temperatures of the order of J. two-magnon Raman scattering essentially probes the highest In this paper we report about the progresses in the theory wave vectors, related to short-range correlations. Therefore of Heisenberg antiferromagnets that have been obtained by two-spin Raman scattering features persist also in the para- our group in Florence. The early work on the theory of two- magnetic phase7 where short-range order is still present. Let magnon Raman scattering is summarized in Sec. 2, while the us consider the following antiferromagnetic Hamiltonian subsequent Sections report the recent activity on low- with exchange integral JϾ0, z nearest neighbors with dis- dimensional antiferromagnetism. placements labeled by d, and two (a,b) sublattices:8 Section 3 is devoted to 1D models, and concerns the J study of the effect of soliton-like excitations in the com- Hϭ ͑ ͒ ͚ Si,a•Siϩd,b . 1 pound TMMC, as well as the anisotropic spin-1 model, for 2 id which a reduced description of the ground state allows one to 9 investigate the quantum phase transition in a unitarily trans- The scattering cross section S(␻) turns out to be propor- formed representation and to obtain quantitative results for tional to the Fourier transform of ͗M(0)M(t)͘, where the phase diagram. Section 4 concerns the theory of the iso- tropic 2D QHAF, for which we reproduced the experimental ϭ M ͑ ͒ M ͚ kSk•SϪk , 2 correlation length by means of a semiclassical approach, also k deriving the connection with ͑and the limitations of͒ famous is the effective Raman scattering operator. quantum field theory results. In Sec. 5 we summarize several Many antiferromagnetic compounds can be mapped onto recent results concerning the anisotropic 2D QHAF, with this model, even though a small next-nearest-neighbor ex- emphasis onto the different phase diagrams and the experi- change interaction without competitive effects, as well as mentally measurable signatures of XY or Ising behavior. anisotropy terms could be present. For instance, there are 3D Eventually, in Sec. 6 results on the 2D frustrated J1 – J2 iso- ͑ perovskite and rutile structures e.g., KNiF3 , NiF2) and 2D tropic model are described. ͑ layered structures e.g., K2NiF4 , LaCuO2). Let us remember that the exact ground state is not ex- II. TWO-MAGNON RAMAN SCATTERING IN HEISENBERG actly known, except in 1D models with Sϭ1/2 or Sϭϱ͑i.e., ANTIFERROMAGNETS the classical case͒: in the latter case it coincides with the The scattering of radiation is a very powerful tool for ‘‘Ne´el state’’ with antialigned sublattices. studying elementary excitations in condensed matter physics. In the ordered phase the theory can be developed in ␣ ␤ Any complete experiment gives rise to a quasi-elastic com- terms of two families of magnon operators ( k , k), through ponent due to nonpropagating or diffusive modes and to the Dyson-Maleev spin-boson transformation and a Bogoliu- symmetrically shifted spectra corresponding to the states of bov transformation: the system under investigation with an amplitude ratio gov- HϭE ϩH ϩVϩ..., ͑3͒ erned by the detailed balance principle. 0 0 The most sensitive probes for this investigation are un- where E0 is the ground-state energy in the interacting spin- doubtedly thermal neutrons, because the characteristic ener- wave approximation, and gies and wave vectors fit very well with those of the mag- netic elementary excitations. However, light-scattering H ϭ ␻ ͑␣†␣ ϩ␤†␤ ͒ ͑ ͒ 0 ͚ k k k k k 4 experiments can require a simpler apparatus and offer a bet- k ter accuracy, although the transfer wave vector k is much smaller than the size of the Brillouin zone, so that usually is the quadratic part of the Hamiltonian of a magnon only the center of this zone can be directly probed. In spite of whose frequencies, renormalized by zero-T quantum fluctua- this, two-spin Raman scattering involving the creation and tions, are destruction of a pair of elementary excitations can be per- C 1 formed, with the contribution of two magnons having equal ␻ ϭ ͩ ϩ ͪͱ Ϫ␥2 ␥ ϭ Ϫik"d ͑ ͒ k JSz 1 k; k ͚ e . 5 frequencies and opposite wave vectors. This two-magnon 2S z d 670 Low Temp. Phys. 31 (8–9), August–September 2005 Balucani et al.

FIG. 1. Theoretical two-magnon spectra in KNiF3 at different temperatures.10 FIG. 2. Experimental two-magnon spectra in KNiF3 at different temperatures.5

The last term in the Hamiltonian, V, represents the four- boring spins, thus giving a measure of the short-range anti- magnon interaction, whose most significant term refers to ferromagnetic order that is present at all finite temperatures. two magnons of each family and turns out to be In the disordered phase conventional many-body meth- 2Jz ods are of little use for a quantitative interpretation of the ϭ ␦ ␣␤ ␣†␣ ␤†␤ ͑ ͒ V ͚ qϩp,qЈϩpЈIqqЈ,ppЈ q qЈ p pЈ , 6 observed, largely spread spectra. The concept of quasi- N Ј Ј qq pp particle loses its meaning because of the overdamped char- ␣␤ ␥ S ␻ where the coefficients IqqЈppЈ are known functions of k . acter of the ‘‘excitations.’’ The calculation of ( ) can be In the Hartree-Fock approximation10 the temperature- instead approached by other more general theoretical meth- dependent Raman scattering operator ͑2͒ can be written as ods devoted to the representation of the dynamical correla- tion functions based on the linear response theory.11 Let us ϭ␣͑ ͒ ⌽ ͑␣ ␤ ϩ␣†␤†͒ ͑ ͒ consider the ‘‘Kubo relaxation function’’ associated with our M T S͚ k k k k k 7 k scattering process: with ␻ (T)ϭ␣(T)␻ . The two magnons created or de- 1 ␤ k k ͑ ͒ϵ ␭͗ ␭H ͑ ͒ Ϫ␭H ͑ ͒͘ ͑ ͒ ͑ ͒ f 0 t ͵ d e M 0 e M t . 8 stroyed by the operator 7 interact through V as given by ͗M͑0͒M͑0͒͘ 0 ͑6͒, so that the peak of the cross section S(␻) appears at ␻ Its Laplace transform f (z) is related to the scattering cross values smaller than 2 ZB by an amount of the order of J. 0 The explicit S(␻)atTϭ0 was calculated in the ‘‘ladder section: approximation’’ by Elliott and Thorpe and found in very ␻ good agreement with experiments.9 S͑␻͒ϰ Rf ͑zϭi␻͒. ͑9͒ 1ϪeϪ␤␻ 0 The finite-temperature calculation of the two-magnon Raman scattering cross section in the ordered region, up to Mori14 has given the following continued-fraction represen- ϳ 10 11 T 0.95TN was performed by Balucani and Tognetti, who tation of the relaxation function: calculated the two-magnon propagator in the ‘‘ladder ap- proximation,’’ also taking into account the damping and the temperature renormalization of the magnons at the boundary of the Brillouin zone.11 The calculated spectra S(␻), at in- creasing temperatures, were found in very good agreement with the experimental ones,8 and their characteristic param- eters ͑peak and linewidth͒ permitted determination of the temperature behavior of the frequency and damping of the ZB magnons.12 In Figs. 1 and 2 we show the excellent agree- ment of our theoretical approach with the experiments in the ordered phase.13 The validity of light scattering in probing the characteristic of ZB magnons has been confirmed both from the theoretical and the experimental point of view.12 In Fig. 3 our theoretical ZB magnon damping calculations are compared with experimental data from different techniques. In the paramagnetic phase all the experimental spectra show the persistence of a broad inelastic peak up to T ⌫ ϳ ӷ FIG. 3. Zone-boundary damping ZB versus temperature. The symbols refer 1.4TN . Only at T TN do the spectra have a structureless ␻ϭ to different experimental techniques: in particular, the unfilled circles are shape centered around 0. As matter of fact, the highest light scattering data.12 The dashed line is an improvement12 to a previous wave vectors sample only the behavior of clusters of neigh- ͑solid͒ theoretical curve.10 Low Temp. Phys. 31 (8–9), August–September 2005 Balucani et al. 671

Ӎϭ FIG. 4. Two-spin Stokes spectrum in KNiF3 at T 1.02TN . The line reports the theoretical shape,15 compared with experimental data.

FIG. 5. Experimental contribution of nonlinear excitations to the specific ⌬ ϭ Ϫ Ϫ⌬ heat of TMMC, C C(H) C(0) CSW . The field values are H ϭ5.39 T ͑᭹͒ and Hϭ2.5 T ͑᭺͒. The dotted-and-dashed line reports the 1 1 ͑ ͑ ͒ϭ ͑ ͒ϭ ͑ ͒ result of the free soliton gas phenomenology. The planar model interpolated f 0 z ϩ⌬ ͒ ; f n z ϩ⌬ ͒ , 10 crosses͒ appears to explain quantitatively the behavior of TMMC. z 1 f 1͑z z nϪ1 f nϩ1͑z which is formally exact, but allows one to perform some approximations in respect to the level of the termination f ϩ (z). The quantities ⌬ can be expressed in terms of n 1 n Hϭ ͑ Ϫ␦ z z ͒ ͑ ͒ J͚ Si•Siϩ1 Si SiϪ1 13 frequency moments: i ϱ ͗␻2n͘ϭ ͵ ␻␻2n ͑␻͒ ͑ ͒ with a very small easy-plane anisotropy (␦ϭ0.0086). d f 0 . 11 Ϫϱ A magnetic field of the order of 1–10 T can be applied perpendicularly (y axis͒ or along the chain. In the first case, In our calculations of f (z) in the entire paramagnetic 0 with approximations the more valid the lower the magnetic phase,15,16 the coefficients ⌬ and ⌬ have been evaluated 1 2 field (HϽ5 T), in the continuum limit TMMC can be repre- approximately by means of a decoupling procedure. More- sented by the classical sine-Gordon Hamiltonian: over, the third stage of the continued fraction ͑10͒ is evalu- ated assuming that A Hϭ ͵ dx͓⌽˙ 2ϩc2⌽22␻2͑1Ϫcos ⌽͔͒, ͑14͒ ⌬ ͑ ͒ϳ⌬ ͓ ͑ ͒ϩ Ј͑ ͔͒ ͑ ͒ 2 0 x 0 3 f 3 z 3 f 3 0 zf3 0 . 12 The parameters involved in ͑12͒ can be estimated from whose parameters are related with the magnetic Hamiltonian ͑ ͒ ϭ ␮ 13 , the reduced magnetic field h g BH, and the lattice knowledge of the short-time behavior of f 0(t) determined by the first moments, ͗␻2͘ and ͗␻4͘. spacing a as follows: The results of our approach in the paramagnetic region 1 ␦ ␦ are compared with the experiment in Fig. 4, showing the Aϭ , c ϭaJSͱ1Ϫ , ␻ ϭhͱ1Ϫ . ͑15͒ 8Ja 0 2 0 2 persistence of the peak of the ZB magnetic excitations above the critical temperature. The energy of a kink-soliton turns out to be ϭ ␻ Ӎ ͑ ͒ Es 8A 0c0 hS, 16

III. THE ONE-DIMENSIONAL ANTIFERROMAGNET and depends on the applied field. At variance with the ferro- magnetic solitons, these solitons can be easily excited at low- A. Solitons in the antiferromagnet TMMC est temperatures and can give a significant contribution to the 19 Interest in low-dimensional systems is motivated by the thermodynamics. When the field is applied longitudinally simpler calculations as compared with the 3D ones, accom- along the z axis only spin waves are present: therefore, the panied by interesting peculiar behavior. The powerful math- specific-heat measurements were performed in the two con- ematical approach based on the inverse-scattering and Bethe figurations. The contribution from the nonlinear excitations ⌬ ansatz techniques permits one to solve exactly some 1D was obtained as the difference C between the two experi- models, calculating thermodynamic and sometimes transport ments. quantities both in the classical and quantum cases.17 The The thermodynamic quantities were calculated by the 20 most celebrated realizations of these models occur in 1D classical transfer-matrix method for the sine-Gordon model ͑ ͒ 21 magnets. An original suggestion by Mikeska18 was that the 14 . We then used a classical discrete planar model: antiferromagnetic chain TMMC ͓(CH3)4NMnCl3͔ can be Hϭ 2 ⌽ Ϫ⌽ ͒ϩ Ϫ ⌽ ͒ ͑ ͒ mapped onto a classical 1D sine-Gordon model. The elemen- ͚ ͓2JS cos͑ i iϩ1 hS͑1 cos i ͔, 17 tary excitations of the sine-Gordon field are given in terms of i linear small-amplitude spin waves and nonlinear breathers verifying that it is qualitatively similar to the sine-Gordon. and kink-solitons. The nonlinear elementary excitations give The comparison21 is shown in Fig. 5, where the linear spin- a detectable contribution to the magnetic specific heat. wave specific heat was subtracted to emphasize the nonlinear TMMC is composed of Heisenberg (Sϭ5/2) antiferro- contribution, together with the prediction of the ‘‘classical magnetic chains along the z axis: soliton gas phenomenology.’’ 19 672 Low Temp. Phys. 31 (8–9), August–September 2005 Balucani et al.

This proved the presence of nonlinear excitations similar Upon its ground state, the antiferromagnetically ordered to sine-Gordon solitons, but the peak of the specific heat Ne´el state, one may construct three types of excitations: a occurs at temperatures where solitons cannot be considered single deviation, a direct soliton, an indirect soliton, where to be noninteracting and the ‘‘classical soliton-gas phenom- direct (indirect) refers to the fact that the excitation be gen- enology’’ breaks down. When the magnetic field is increased erated by flipping all the spins on the right of a given site to 9.98 T the model ͑17͒ is no longer able to describe the while keeping the z component of the spin on such site un- experiments. A quasi-uniaxial model21 was proposed and changed ͑setting it to zero͒. All of the above configurations found to be in good agreement. For general reference on the have energy ϩ2␭ with respect to that of the ground state, subject see Ref. 22. and, when properly combined, generate all of the excited states; among them, we concentrate upon those containing a couple of adjacent indirect solitons and notice that their en- ergy is ϩ3␭, while excited states containing two separate B. The SÄ1 quantum antiferromagnet indirect solitons have energy ϩ4␭. Therefore, indirect soli- We deal here with quantum antiferromagnetic spin tons are characterized by a binding energy ␭; moreover, one chains, focusing our attention on the class of models defined may easily see that isolated solitons may effectively intro- by the Hamiltonian duce disorder in the global configuration of the system, while coupled solitons only reduce the magnetization of each of the H 40 ϭ ͓͑ x x ϩ y y ͒ϩ␭ z z ϩ ͑ z͒2͔ ͑ ͒ two antiferromagnetic sublattices. In fact, strings contain- ͚ Si Siϩ1 Si Siϩ1 Si Siϩ1 d Si 18 J i ing any odd ͑even͒ number of adjacent solitons act on the order of the global configuration as if they were isolated Ͼ with exchange integral J 0 and single-ion anisotropy d. ͑coupled͒ solitons. One of the most surprising evidences of the difference As we move from the Ising limit, the transverse interac- between ferro- and antiferromagnetic systems is related with tion ͚ (SxSx ϩSySy ) comes into play, and it is seen41 to ϭ i i iϩ1 i iϩ1 the so-called Haldane conjecture, i.e., with T 0 properties lower the energy of the system more efficiently by delocal- of integer-spin antiferromagnetic chains. In general, we ex- izing indirect solitons rather than single deviations or direct pect three possible situations for the ground state of a mag- solitons, thus indicating that configurations which uniquely ͑ netic system: either it is ordered with finitely constant cor- contain indirect solitons are crucial in understanding how the ͒ ͑ relation functions , or quasi-ordered with power-law system evolves from the Ising limit ͑Ne´el phase͒ to the iso- ͒ decaying correlation functions , or completely disordered tropic case ͑Haldane phase͒. ͑ ͒ with exponentially decaying correlation functions . One From the above ideas we may draw a simple but sugges- could intuitively expect the third option to be possibly dis- tive scheme for such evolution: missed, based on the idea that, when thermal fluctuations are —in the Ising limit (␭→ϱ) the ground state is the anti- completely suppressed, the system is in an ordered or at least ferromagnetically ordered Ne´el state; quasi-ordered ground state. This idea is in fact proved correct —as ␭ decreases, indirect solitons appear along the for half-integer spin systems, thanks to the so-called Lieb- chain in pairs, thus keeping the antiferromagnetic order; 23 Schultz-Mattis theorem. Despite the impossibility of gen- —as ␭ is further lowered, indirect soliton pairs dissoci- eralizing such a theorem to integer-spin systems, its general ate due to the transverse interaction which, by spreading soli- 1 validity has been taken for granted till 1983, when Haldane tons along the chain, can cause the ground state to be disor- suggested, for the integer-spin Heisenberg chain, an unex- dered. ϭ pected T 0 behavior: a unique and genuinely disordered Due to the privileged role of indirect solitons in the ground state, meaning exponentially decaying correlation above scheme, we concentrate on configurations which con- functions and a finite gap in the excitation spectrum. After tain only indirect solitons. Such configurations generate a more than two decades Haldane’s idea that integer-spin sys- subspace for the Hilbert space of the system, which is re- tems can have a genuinely disordered ground state still ferred to in the literature as the reduced Hilbert space.42 24–27 stands as a conjecture. However, theoretical, States belonging to the reduced Hilbert space are strongly 28–32 33–39 experimental, and numerical works have definitely characterized by the fact that if one eliminates all the sites confirmed its validity. with Szϭ0, a perfectly antiferromagnetically ordered chain is ͑ ͒ ␭ Let us consider Eq. 18 for integer spin: in the (d, ) left. Remarkably, this type of order, which is called hidden plane one may identify different quantum phases, corre- order in the literature, is not destroyed by dissociation of sponding to models whose ground states share common fea- soliton pairs, and it actually characterizes the disordered ␭Ͼ tures. For 0 three phases are singled out: the Ne´el phase ground state of a Haldane system, as discussed below. ␭ӷ ( d), where the ground state has a Ne´el-like structure, the In 1992 Kennedy and Tasaki ͑KT͒ defined a nonlocal ӷ␭ so-called large-d phase (d ), where the ground state is unitary transformation43 which makes the hidden order vis- zϭ characterized by a large majority of sites where S 0, and ible, meanwhile clarifying its meaning. The transformation is the Haldane phase, which extends around the isotropic point defined by (dϭ0,␭ϭ1), and is characterized by disordered ground states. ϩ͓ ͔ ϭ Ϫ N0 N/2 We first deal with the Ising limit, U ͑ 1͒ ͟ Uk k H ϭ␭ z z /J ͚ Si Siϩ1 . i with Low Temp. Phys. 31 (8–9), August–September 2005 Balucani et al. 673

kϪ1 1 z x ͉⌿ ϵU͉⌽ ϭ ͉ ͑ ͒ ϭ ͫ ͩ ␲ ͪ Ϫ ͬ ͑ ␲ ͒ 0͘ 0͘ ͚ c͕s͖ s1s2 ...sN͘ 22 Uk exp i ͚ S p 1 exp i Sk 2 pϭ1 ͕s͖ z kϪ1 with ͕s͖ϵ(s ,s ,s ,...s ) and s ϵ͗⌿ ͉S ͉⌿ ͘ϭϩ1,0. 1 1 2 3 N i 0 i 0 ϩ ͫ ͩ ␲ z ͪ ϩ ͬ The simplest trial ground state allowing the description exp i ͚ S p 1 , 2 pϭ1 of soliton pair dissociation is that defined by Eq. ͑22͒ with c͕ ͖ϭt t ...t . The variational parameters where N is the number of sites of the chain, ͓N/2͔ is the s s1s2s3 s2s3s4 sNϪ2sNϪ1sN ϭ integer part of N/2, and N0 is the number of odd sites where are the six amplitudes tϩϩϩ , tϩϪ0 t0ϩϩ , tϩ0Ϫ , t0ϩ0 , 2 zϭ ͉⌿͘ t ϩϭtϩ , t , where ͉t ͉ represents the probabil- S 0. If the pure state has hidden order, meaning that it 00 00 000 siϪ1sisiϪ1 ͉⌿͘ z z z only contains indirect solitons, then U has spins with ity for (SiϪ1 ,Si ,Siϩ1) to be equal to (siϪ1 ,si ,siϩ1); a com- z ͉⌿ S 0 all parallel to each other. This point is made transpar- mon arbitrary factor may be used for normalizing 0͘.We 44 ent by the introduction of the string order parameter notice that the chosen form for c͕s͖ is such that the probabil- ͉⌿ jϪ1 ity for 0͘ to contain coupled solitons is finite indepen- ␣ ͑H͒ϵ ͳ ␣ ͫ ␲ ͚ ␣ͬ ␣ ʹ ͑ ͒ dently of that relative to the occurrence of isolated solitons, Ostring lim Si exp i Sl S j , 19 ϭ  ͉iϪ j͉→ϱ l i H whose presence is unambiguously marked by tϩ0ϩ 0. Without going into the details of the variational calcula- ␣ϭ where x, y, z, and ͗...͘H indicates the expectation value tions, reported in Ref. 46, we discuss here our final results. H over the ground state of the Hamiltonian . It may be shown Due to the normalization condition, the number of varia- z  that Ostring 0 if and only if the ground state belongs to the tional parameters is reduced from six to five; moreover, the reduced Hilbert space. In other terms, while ferromagnetic Ϫ energy ͗⌿ ͉UH˜ U 1͉⌿ ͘ is found to depend just on four order is revealed by the ferromagnetic order parameter 0 0 precise combinations of the original parameters, ␣ ␣ ␣ O ϭ lim ͗S S ͘H , 2 2 ferro i j ␥ϵ͉tϩϩ ͉ ͉t ϩ͉ ⇒͑ ϩϩ00ϩϩ ͒, ͉iϪ j͉→ϱ 0 00 ¯ ¯ ␲ϵ͉ ͉2͉ ͉⇒͑ ϩϩ ϩϩ ͒ the hidden order is revealed by the string order parameter tϩϩ0 tϩ0 ¯ 0 ¯ , ͑ ͒ Eq. 19 . In fact, the nonlocal transformation U relates the 2 ␹ϵ͉t ϩ͉ ͉t ϩ ͉⇒͑ 00ϩ00 ͒, above order parameters through the relation 00 0 0 ¯ ¯ ␩ ␩ Ϫ ␳ϵ͉ ͉⇒͑ ͒ ͑ ͒ ϭ ͑ H 1͒ ͑ ͒ t000 000 , 23 Ostring Oferro U U , 20 ¯ ¯ for ␩ϭx,z, meaning that the analysis of the hidden order in whose square moduli are related to the probabilities that the ͑⇒͒ ͉⌿ ͘ a system described by H may be developed by studying the corresponding strings be contained in 0 ; in particu- ␥2 ␲2 ferromagnetic order in the system described by the trans- lar, and refer to the probabilities for coupled and iso- Ϫ lated solitons, respectively, to occur in the ground state. formed Hamiltonian H˜ ϵUHU 1, which reads, for H de- ͑ ͒ Both the analytical expression for the energy and the fined by Eq. 18 , ␭ numerical minimization show that a critical value c ϭ␭ Ͼ ␭Ͼ␭ H˜ c(d) d exists such that for c the minimal energy is x x y i␲͑SzϩSx ͒ y z z ϭ ͓Ϫ ϩ i iϪ1 Ϫ␭ ␲ϭ␳ϭ ␭ϭ␭ ͚ Si Siϩ1 Si e Siϩ1 Si Siϩ1 attained for 0; the condition c(d) can, hence, J i define a curve of phase separation. We therefore single out ϩ ͑ z͒2͔ ͑ ͒ d Si . 21 three different phases, characterized by

Our work developed as follows: one first assumes that ͑a͒␲ϭ␳ϭ0,͑b͒all parameters0,͑c͒␹ϭtϩϩϩϭ0 the relevant configurations, as far as the Ne´el-Haldane tran- ͑24͒ sition is concerned, belong to the reduced Hilbert space; this in the ground state. The corresponding phase diagram is permits, by the KT transformation, to restrict the analysis to shown in Fig. 6, together with that obtained with a factorized zϭ zϭ ᭙ the subspace of states with either Si 1orSi 0, i. Then trial ground state43 and by numerical simulations.33 the expectation value of the transformed Hamiltonian H˜ in The (a)–(b) transition is seen to quite precisely de- Eq. ͑21͒ is minimized on a trial ground state whose structure scribe the Ne´el-Haldane one, and this leads us to define the takes into account at least short-range correlations between condition (a), meaning the occurrence of exclusively spins. By this procedure, we aim at following the effective coupled solitons, as typical of the Ne´el phase. As to the dissociation of soliton pairs, in order to clarify the connec- (c)–(b) transition, one should note that the use of the re- tion between the occurrence of isolated solitons in the duced Hilbert space is not fully justified in the ␭Ͻd region, ground state, and the transition towards the completely dis- where we in fact do not expect quantitatively precise results. 39,41,42,45 ordered Haldane phase. For a comparison between our results and the exact nu- In the framework of a standard variational approach, we merical data available, we have considered, along the dϭ0 ͗⌽ ͉H͉⌽ ͘ ␭ should minimize 0 0 with respect to a certain num- axis, two specific quantities: the critical anisotropy c(d), ber of variational parameters entering the expression of the where the Ne´el phase becomes unstable with respect to the ͉⌽ ͘ ␭ normalized trial ground-state 0 . By applying the nonlocal Haldane phase, and the ground-state energy E0(d, ) at the unitary transformation U we instead minimize isotropic point ␭ϭ1. For the critical anisotropy we find ͗⌿ ͉UHUϪ1͉⌿ ͘ ͉⌿ ͘ϵU͉⌽ ͘ ␭ ϭ 0 0 with 0 0 and the transformed c(0) 1.2044(5) as compared to the value obtained by ex- H˜ ϵUHUϪ1 ͑ ͒ ͉⌽ 37 ␭ Ϸ Hamiltonian defined by Eq. 21 ;if 0͘ be- act diagonalization, c(0) 1.19; for the energy we find ϭϪ ϭ longs to the reduced Hilbert space, it is E0(0,1) 1.3663(5) as compared to E0(0,1) 674 Low Temp. Phys. 31 (8–9), August–September 2005 Balucani et al.

x ͑ ͒ z ͑ ͒ FIG. 8. String order parameters Ostring squares , Ostring circles , and soliton ␭Ͼ ͑ ͒ ͑ ͒ ϭ FIG. 6. Phase diagram in the 0 half-plane: our results squares are density n0 triangles for d 0. shown together with those of Ref. 43 ͑dotted lines͒; the Haldane phase should correspond to the shaded area, according to the best available nu- merical data33 ͑solid lines͒. Given their essential role, we have also studied the x and z components of the string order parameter, as well as the soliton density n ϭ1Ϫ͗(Sz)2͘. After KT we expect Ϫ1.4014(5), again from an exact diagonalization 0 Oz (H)0 in both the Ne´el and the Haldane phase, and technique;47 the value obtained with the factorized trial string Ox (H)0 just in the Haldane phase. In fact, analytical ground state is E (0,1)ϭϪ4/3.43 string 0 expressions for Ox and Oz may be written46 in terms of four In Fig. 7 we show the variational parameters as ␭ is of the five variational parameters ͑25͒, and show that varied for dϭ0, i.e., along the y axis of the phase diagram; x O (H)ϭ0if␲ϭ␳ϭ0or␹ϭtϩϩϩϭ0, i.e., in phases (a) in fact, rather than the parameters with respect to which we string and (c); Oz (H)Ͼ0 in all phases, asymptotically vanish- have actually minimized the energy, the following combina- string ing as ␳→1, i.e., in the far large-d phase. tions are considered: x ϭ In more detail, we note that Ostring 0 whenever the ϭ␲2 2 ϭ␥ w͑1͒ /tϩϩϩ; w͑2͒ ; ground state does not contain strings made of an odd number of adjacent spins; as soon as the shortest string of such type, ϭ͑␹␥ ͒1/2 ϭ 2 ͑␳␥͒2/3 ͑ ͒ w͑2,2͒ tϩϩϩ ; w͑3͒ tϩϩϩ . 25 namely the isolated soliton, appears along the chain, then x z Ͼ The above quantities have a straightforward physical Ostring becomes finite. The unphysical result Ostring 0 in the meaning, as they are directly related with the probabilities (c) phase, vanishing only as d→ϱ rather than everywhere in for a soliton to appear along the chain as an isolated excita- the large-d phase, is due to our assuming the ground state to 2 belong to the reduced Hilbert space, which actually is licit tion (w(1)), as a part of a soliton pair (2w(2)), as a part of a string made of three adjacent solitons (3w3 ), and finally as only in the ␭Ͼd region. (3) x z ␭ a part of a string made of two soliton pairs separated by one In Fig. 8 we show Ostring , Ostring , and n0 as varies 4 with dϭ0. We underline that Ox becomes finite continu- site (4w(2,2)). From Fig. 7, it becomes evident that the string Haldane phase is featured by the occurrence of isolated soli- ously but with discontinuous derivative at the transition ͑re-  flecting the behavior of w and w shown in Fig. 7͒,so tons (w(1) 0), as well as of strings made of three adjacent (1) (3)  that the Ne´el-Haldane quantum phase transition is recog- solitons (w(3) 0). This result confirms that, as elicited by the analysis of nized as a second-order one. In Fig. 9 we zoom the order x the phase diagram, the Haldane phase is characterized by our parameter Ostring around the critical point: its behavior is x ϳ ␭ Ϫ␭ ␤ condition (b). seen to be described by a power law Ostring ( c ) ,as

x ϭ FIG. 9. Critical behavior of Ostring for d 0: squares are our results; curves x ϳ ␭Ϫ␭ ␤ ␤ are obtained by best-fit procedure from Ostring ( c) with fixed to ͑ ͒ ͑ ͒ ͑ ͒ ͑ ͒ ␤ϭ ͑ FIG. 7. Parameters w(1) squares , w(2) circles , w(3) upward triangles , 0.125 dashed curve and as fitting parameter, resulting in 0.217 solid ͑ ͒ ϭ ͒ ␭ ϭ and w(2,2) downward triangles for d 0. curve . Both procedures give c 1.2044(5), marked by a circle in figure. Low Temp. Phys. 31 (8–9), August–September 2005 Balucani et al. 675

expected for a continuous phase transition; our estimated system towards a 3D ordered phase: an antiferromagnetic value for the critical exponent is ␤ϭ0.217(5) as compared Heisenberg interaction and the small spin value make these to ␤ϭ0.125, corresponding to the Ising model in a trans- compounds behave as 2D QHAFs. Even the onset of 3D verse field, to whose universality class the Haldane transition magnetic long-range order is however strongly affected by is suggested to belong.42 At the isotropic point (dϭ0,␭ the 2D properties of the system: indeed, the observed 3D ϭ x ϭ 1) we find Ostring 0.3700(5), in full agreement with the magnetic transition temperature is comparable with the intra- value obtained by exact diagonalization.48 plane interaction energy, i.e., orders of magnitude larger than The overall good agreement between our results and the one would expect solely on the basis of the value of the numerical data available, allows us to conclude that the interplane coupling. Such apparently odd behavior can be Ne´el-Haldane transition is a second-order one, and that the easily understood by observing that the establishing of in- x ␰ string order parameter Ostring , revealing hidden order along plane correlations over a characteristic distance effectively the x direction, is the appropriate order parameter for the enhances the interplane coupling by a factor (␰/a)2, a being Haldane phase. The disordered ground state featuring the the lattice constant. The latter consideration is one of the Haldane phase is seen to originate by dissociation of soliton reasons explaining why most of the attention, both from the pairs, according to this path. Solitons occur just in pairs in experimental and theoretical point of view, has been paid to the antiferromagnetically ordered Ne´el phase; at the Ne´el- the low-temperature behavior of the correlation length ␰ of Haldane transition soliton pairs dissociate and the byproducts the 2D QHAF ͑in the following ␰ will always be given in rearrange into strings consisting of an odd number of soli- units of the lattice constant a). tons. These strings are ultimately responsible for the disorder The 2D QHAF is described by the Hamiltonian of the ground state. J Hϭ ͚ S S ϩ , ͑26͒ 2 i• i d IV. TWO-DIMENSIONAL ISOTROPIC HEISENBERG MODEL i,d The 2D isotropic QHAF on the square lattice is one of where J is positive and the quantum spin operators Si satisfy ͉ ͉2ϭ ϩ ϵ the magnetic models most intensively investigated in the last Si S(S 1). The index i (i1 ,i2) runs over the sites of a two decades. This is due both to its theoretically challenging square lattice, and d represents the displacements of the 4 Ϯ Ϯ properties and to it being considered the best candidate for nearest-neighbors of each site, ( 1,0) and (0, 1). modeling the magnetic behavior of the parent compounds of In addition to the first approximations usually employed some high-T superconductors.2,3 to investigate the low-temperature properties of magnetic c ͑ ͒ From a theoretical point of view the fully isotropic systems as, e.g., mean-field and modified spin-wave theory, Heisenberg model in d dimensions, thanks to the simple the critical behavior of the 2D QHAF was commonly inter- structure of its Hamiltonian ͑whose high symmetry is re- preted on the basis of the results obtained by field theory ␴ sponsible for most of its peculiar features͒, may well be con- starting from the so-called 2D quantum nonlinear model ͑ ␴ ͒ 58 sidered a cornerstone of the modern theory of critical phe- QNL M , whose action is given by nomena, with its relevance extending well beyond the only 1 u 2 2 2 ␶n͉ ͒; ͉n͉ ϭ1. ͑27͒ץmagnetic systems. The dϭ2 case earned additional interest, Sϭ ͵ dx͵ d␶ٌ͉͑n͉ ϩ͉ 2g 0 representing the boundary dimension separating systems with and without long-range order at finite temperature.4 The In the last equation n͑x͒ is a unitary 3D vector field, g antiferromagnetic coupling adds further appeal, as the ϭc⌳/␳ and uϭc⌳/T are the coupling and the imaginary- classical-like Ne´el state is made unstable by quantum fluc- time cut-off, respectively, and the two parameters ␳ and c are tuations, and the ground state of the system is not exactly usually referred to as spin stiffness and spin-wave velocity. known. It can be rigorously proven19 to be ordered for S Despite their names, however, the two parameters ␳ and c у1; for Sϭ1/2 there is no rigorous proof, although evidence are just phenomenological fitting constants which can be rig- for an ordered ground state can be drawn from many differ- orously related to the proper parameters J and S of the origi- ent studies ͑for a review, see, for instance, Ref. 50͒. nal magnetic Hamiltonian ͑26͒ only in the large-S limit.1,59 On the experimental side the attention to the properties The source of nonlinearity in the model of Eq. ͑27͒, which is of 2D QHAF was mainly triggered by the fact that among seemingly quadratic in the field variables, is the constraint the best experimental realizations of this model we find sev- imposed on the length of the field n. ␴ eral parent compounds of high-Tc superconductors, as, e.g., The relation between the 2D QHAF and the QNL M ͑ ͒ La2CuO4 or Sr2CuO2Cl2 Refs. 51–53 , both having spin S was first exploited to interpret the experimental data on cu- ϭ1/2. In such materials, as well as in other magnetic com- prous oxides by Chakravarty, Halperin, and Nelson ͑CHN͒,58 pounds with a layered crystal structure, such as La2NiO4 who used symmetry arguments to show that the long- ͑ ͒ ͑ ͒ ϭ ͑ Ref. 54 and K2NiF4 Refs. 52, 53 (S 1), Rb2MnF4 Ref. wavelength physics of the QHAF is the same of that of the ͒ ͑ ͒ ϭ ␴ 55 and KFeF4 Ref. 56 (S 5/2), or copper formate tetra- QNL M; in other words, the physical observables of the two deuterate ͑CFTD, Sϭ1/2) ͑Ref. 57͒, the magnetic ions form models show the same functional dependence upon T, if the parallel planes and interact strongly only if belonging to the long-wavelength excitations are assumed to be the only rel- same plane. The interplane interaction in these compounds is evant ones, as one expects to be at low temperature. orders of magnitude smaller than the intraplane one, thus The analysis carried out by CHN on the QNL␴M leads offering a large temperature region where their magnetic be- one to single out three different regimes, called quantum havior is indeed 2D down to those low temperatures where disordered, quantum critical ͑QCR͒, and renormalized clas- the weak interplane interaction becomes relevant, driving the sical ͑RCR͒, the most striking difference amongst them be- 676 Low Temp. Phys. 31 (8–9), August–September 2005 Balucani et al. ing the temperature dependence of the spin correlations. If g potential qˆគ -dependent terms, and its application may be ex- is such as to guarantee LRO at Tϭ0, the QNL␴Misinthe tended also to magnetic systems, according to the following RCR at very low temperature, and the correlation length ␰ scheme.65 The spin Hamiltonian H(Sᠪ) is mapped to H(pˆគ ,qˆគ ) 60 behaves as by a suitable spin-boson transformation; once the corre- H គ គ គ ϵ គ គ sponding Weyl symbol (p,q) with p (p1 ,...pN) and q e c 2␲␳ T ϵ ␰ ϭ ͩ ͪ ͩ ͪͫ Ϫ ͬ ͑ ͒ (q1 ,...qN) classical phase-space variables, has been deter- 3l exp 1 . 28 8 2␲␳ T 4␲␳ mined, the PQSCHA renormalizations may be evaluated, and H គ គ the effective classical Hamiltonian eff(p,q) and effective CHN also found that by raising the temperature, any 2D O pគ qគ ␴ classical function eff( , ) corresponding to the observable QNL M with an ordered ground state crosses over from the O H sᠪ ␰ of interest follow. Finally, the effective functions eff( ) RCR to the QCR, characterized by a correlation length O sᠪ s ϰ␣ ϭ and eff( ), both depending on classical spin variables with (T) c/T. ͉s͉ϭ1 and containing temperature- and spin-dependent quan- The first direct comparison between experimental data ␴ tum renormalized parameters, are reconstructed by the in- on spin-1/2 compounds and the prediction of the QNL M verse of the classical analog of the spin-boson transformation field theory in the RCR gave surprisingly good agreement used at the beginning. and caused an intense activity, both theoretical and experi- In order to successfully carry out such a renormalization mental, in the subsequent years. However, with the accumu- scheme, the Weyl symbol of the bosonic Hamiltonian must lation of new experimental data on higher-spin compounds it be a well-behaved function in the whole phase space. Spin- clearly emerged that the experimentally observed behavior of boson transformations, on the other hand, can introduce sin- ␰(T) for larger spin could be reproduced neither by the origi- ͑ ͒ ͑ ͒ gularities as a consequence of the topological impossibility nal simplified two-loop form of Eq. 28 given by CHN of a global mapping of a spherical phase space into a flat ͑which does not contain the term in square brackets͒ nor by ͑ ͒ one. The choice of the transformation must then be such that the three-loop result 28 derived by Hasenfratz and Nieder- the singularities occur for configurations which are not ther- ͑ ͒ 60 mayer HN ; moreover, no trace of QCR behavior was modynamically relevant, and whose contribution may be found in pure compounds. The discrepancies observed could hence be approximated. Most of the methods for studying be due to the fact that the real compounds do not behave like magnetic systems do in fact share this problem with the 2D QHAF or to an actual inadequacy of the theory. In par- PQSCHA; what makes the difference is that by using the ticular the CHN-HN scheme introduces two possible reasons PQSCHA one separates the classical contribution to the ther- for such inadequacy to occur: the physics of the 2D QHAF is mal fluctuations from the pure quantum one, and the ap- not properly described by that of the 2D QNL␴M and/or the ͑ ͒ proximation only regards the latter, being the former exactly two three -loop renormalization-group expressions derived taken into account when the effective Hamiltonian is recast by CHN-HN do hold at temperatures lower than those ex- in the form of a classical spin Hamiltonian. perimentally accessible. After an almost ten-year debate, the The spin-boson transformation which constitutes the first latter possibility has finally emerged as the correct one, being step of the magnetic PQSCHA is chosen according to the strongly supported not only by our own work but also by 61 symmetry properties of the original Hamiltonian and of its other independent theoretical approaches, joined with the ground state. In the case of the 2D QHAF both Dyson- analysis of the experimental and the most recent quantum ͑ ͒ Maleev and Holstein-Primakoff transformation can be em- Monte Carlo QMC data for the 2D QHAF. ployed, finally yielding:63 The theoretical approach we employed to investigate the 2D QHAF is the effective Hamiltonian method,62–64 devel- H ␪4 eff ϭ ϩ G ͑ ͒ oped within the framework of the pure-quantum self- ͚ si siϩd N ͑t͒, 29 2 • consistent harmonic approximation ͑PQSCHA͒ we intro- JS˜ 2 i,d duced at the beginning of the 1990s.65,66 The PQSCHA starts t sinh f from the Hamiltonian path-integral formulation of statistical G͑ ͒ϭ kϪ ␬2D ͑ ͒ t ͚ ln ␪2 2 , 30 mechanics, which allows one to separate in a natural way N k f k classical and quantum fluctuations; only the latter are then with the temperature- and spin-dependent parameters treated in a self-consistent harmonic approximation, finally yielding an effective classical Hamiltonian whose properties D ␪2ϭ1Ϫ , ͑31͒ can thereafter be investigated by all the techniques available 2 for classical systems. The idea of separating classical and quantum fluctuations turned out to be fruitful not only in 1 Dϭ ͑ Ϫ␥2͒1/2L ͑ ͒ view of the implementation of the PQSCHA, but also in the ͚ 1 k k , 32 ˜ k final understanding of the connection between semiclassical SN approaches and quantum field theories,67 which could be 68 ␻ 1 possible also thanks to the paper by Hasenfratz about cor- ϭ k L ϭ Ϫ ͑ ͒ f k , k coth f k . 33 rections to the field-theoretical results due to cutoff effects. 2˜St f k The PQSCHA naturally applies to bosonic systems, ␥ ϭ ϩ whose Hamiltonian is written in terms of conjugate operators In the previous equations k (cosk1 cosk2)/2, N is the គ ϵ គ ϵ ϭ ␦ ϵ qˆ (qˆ 1 ,...qˆ N), pˆ (pˆ 1 ,...pˆ N) such that ͓qˆ m ,pˆ n͔ i mn ; number of sites of the lattice, and k (k1 ,k2) is the wave the method, however, does not require H(pˆគ ,qˆគ ) to be stan- vector in the first Brillouin zone; ˜SϵSϩ1/2 is the effective dard, i.e., with separate quadratic kinetic pˆគ -dependent and classical spin length, which naturally follows from the renor- Low Temp. Phys. 31 (8–9), August–September 2005 Balucani et al. 677

FIG. 10. Correlation length ␰ versus t, for Sϭ1/2 ͑leftmost͒ and Sϭ1. The FIG. 11. The function y(t)ϭtln␰ versus t, for ͑from the rightmost curve͒ 71 72 symbols are experimental data; for Sϭ1/2: 63Cu NQR data70 ͑circles͒ and Sϭϱ, 5/2, 1, and 1/2; the up-triangles and the diamonds are quantum ͑ ͒51 MC data for Sϭ1/2; also reported are neutron scattering data for La NiO neutron scattering data for La2CuO4 squares and for Sr2CuO2Cl2 2 4 ͑ ͒ 52,53 ϭ ͑unfilled circles͒,54 K NiF ͑squares͒,52,53 KFeF ͑filled circles͒,56 and up-triangles ; for S 1: neutron scattering data for La2NiO4 2 4 4 ͑ ͒54 ͑ ͒ 52,53 Rb MnF ͑down-triangles͒.55 The abrupt rising of the experimental data for down-triangles and for K2NiF4 diamonds . The classical result 2 4 ͑dotted-and-dashed line͒ is also reported. the Sϭ5/2 compounds at tӍ0.65 is due to the effect of the small, but finite, anisotropies and will be discussed in more detail in Sec. 5.

malization scheme, and tϵT/JS˜ 2 is the reduced temperature defined in terms of the energy scale JS˜ 2. The renormaliza- curves and experimental data for Sу1 ͑including the exact tion scheme is closed by the self-consistent solution of the Sϭϱ classical result͒ display a change of slope at interme- ␻ ϭ ␬2 Ϫ␥2 1/2 ␬2ϭ␪2 two coupled equations k 4 (1 k) and diate temperature, followed by a curvature inversion at lower Ϫt/(2␬2). The pure-quantum renormalization coefficient D t. On the other hand, by looking at the Sϭ1/2 case it be- ϭD(S,t) takes the main contribution from the high- comes clear why the QNL␴M approach gave such a good frequency part ͑short-wavelength͒ of the spin-wave spec- agreement when first used to fit the experimental data. The trum, because of the appearance of the Langevin function change in both the slope and the curvature of t ln ␰ is less L D k . measures the strength of the pure-quantum fluctua- pronounced and possibly occurs at lower temperatures, the tions, whose contribution to the thermodynamics of the sys- lower the spin: in the Sϭ1/2 case, it is difficult to say tem is the only approximated one in the PQSCHA scheme. whether these features are still present or not, but, if yes, they The theory is hence quantitatively meaningful as far as D is occur in a temperature region where the extremely high small enough to justify the self-consistent harmonic treat- value of ␰ (Ϸ104) makes both the experimental and the ment of the pure-quantum effects. In particular, the simple simulation data more difficult to obtain. criterion DϽ0.5 is a reasonable one to assess the validity of After having realized that the field theoretical prediction the final results. by CHN could not be applied to the Sу1 2D QHAF in the The most relevant information we get from Eq. ͑29͒ is temperature range probed by the experiments, the following that the symmetry of the Hamiltonian is left unchanged, so questions were waiting for a satisfactory answer: ͑i͒ the real that from a macroscopic point of view the quantum system range of applicability of the asymptotic three-loop expres- essentially behaves, at an actual temperature t, as its classi- sion ͑28͒ at different S, and ͑ii͒ the possible extension of the ͑ ͒ cal counterpart does at an effective temperature teff PQSCHA results to lower temperature, both in view of iii a ϭt/␪4(S,t). This allows us to deduce the behavior of many comprehensive description of the behavior of the correlation observables ͑but not all!—see Refs. 62–64 for details͒ di- length of the 2D QHAF in the entire range of temperature rectly from the behavior of the corresponding classical quan- and spin values. tities. This is the case of the correlation length, which turns A substantial contribution to settle this conundrum came out to be given simply by only from QMC simulations for higher spin values able to probe the very large correlation length region:73 indeed, ␰͑t͒ϭ␰ ͑t ͒ ͑34͒ cl eff high-precision Monte Carlo data for Sϭ1 and moderate cor- so that once ␪4(S,t) has been evaluated, the only additional relation length could still be very well interpreted by ␰ information we need is the classical cl(t), which is available PQSCHA and did not display the RCR asymptotic behavior, from classical Monte Carlo ͑MC͒ simulation and analytical as shown64 in Fig. 12, where we compared our curves for ␰ asymptotic expressions69 as t→0. and the staggered susceptibility ␹* with QMC data obtained Sample results obtained by PQSCHA are shown in the by Harada et al.74 figures. In Fig. 10 the correlation length for Sϭ1/2 and S QMC results for ␰ by Beard and coworkers73 showed ϭ1 is compared with experimental data; a similar compari- unambiguously that the three-loop Eq. ͑28͒ holds only for son, including MC data for Sϭ1/2 and experimental data on temperatures low enough to ensure an extremely large corre- ϭ ␰տ 5 ϭ ␰տ 12 ϭ S 5/2 compounds KFeF4 and Rb2MnF4 is made in Fig. 11, lation length, e.g., 10 for S 1, 10 for S 3/2, and but along the vertical axis the quantity t ln ␰ is plotted in generally cosmological correlation lengths for SϾ3/2, thus order to better appreciate the deviation from the predicted definitely excluding any possibility of employing QNL␴M RCR behavior, that would correspond to a straight line at low results to interpret available experimental data. t. From the last picture one can easily see that both PQSCHA In Ref. 68, Hasenfratz showed why cutoff effects, which 678 Low Temp. Phys. 31 (8–9), August–September 2005 Balucani et al.

prediction suggested that we replace the perturbative expres- ␰cl ␰ sion 3l with the exactly known classical cl , thus getting the results presented in Fig. 13. It is thus made clear that the main features of the quan- tum correlation length at intermediate temperatures are due to essentially classical nonlinear effects, which cannot be taken into account by perturbative approaches. Moreover, the effective exchange constant that defines the effective tem- H perature teff is seen to depend on the same pure-quantum renormalization coefficients defined by the PQSCHA, ac- cording to an expression which is very similar ͑equal͒ to that found by the latter approach in its standard ͑low-T) version: the behavior of ␰ in the full temperature and spin-value FIG. 12. Correlation length ␰ and staggered susceptibility ␹*ϵ␹/˜S2 versus ranges can thus be quantitatively described by Eq. ͑34͒ with- t for Sϭ1. Symbols are QMC data from Ref. 74. out any adjustable fitting parameter. The results obtained by the PQSCHA about the correla- tion length and other static quantities can also represent the needed information to be inserted within other frameworks, are so devious for Sϭ1/2, significantly modify the correla- like mode-coupling theory, to interpret experiments probing tion length for Sу1. Reaching such goal was possible only dynamic quantities, like nuclear magnetic resonance ͑NMR͒: going back to a direct mapping between the QHAF and the an example of the successful combination of PQSCHA and QNL␴M, and the resulting cutoff-corrected field-theoretical mode-coupling theory is given in Ref. 57. outcome is68

␰ ͒ϭ␰ ͒ ϪC͑T,S͒ ͑ ͒ H͑T,S 3l͑T,S e , 35 V. TWO-DIMENSIONAL ANISOTROPIC HEISENBERG MODEL where C(T,S), defined in Eq. ͑14͒ of Ref. 68, is an integral of familiar spin-wave quantities over the first Brillouin zone. While the theoretical debate mentioned in the previous With this correction, which is the leading order in the Section has been mainly dedicated to the isotropic 2D spin-wave expansion for the cutoff correction, it is possible QHAF, real compounds are not actually well described by the isotropic model when the temperature is low: indeed, the to obtain numerically accurate agreement with QMC data 4 down to ␰տ103 for all S. Mermin-Wagner theorem states that a finite-temperature In our most recent paper67 about 2D QHAF we showed transition cannot occur in the 2D isotropic QHAF, while the that by employing the explicit expression for C(T,S), and by experimental evidence of a transition suggests that 3D corre- substituting in Eq. ͑28͒ the spin stiffness ␳ and the spin-wave lations and anisotropy effects, as well as a combination of ͑ ͒ ͑ ͒ velocity c given by the mapping of QHAF onto the QNL␴M, both, must be considered. Easy-axis EA or easy-plane EP the leading terms of the result ͑35͒ not only can be cast in the anisotropies turn out to be fundamental in the analysis of the ␰ ϭ␰cl H critical behavior. form H(T,S) 3l(teff), in strict analogy with the PQSCHA ͑ ͒ H expression 34 , but the effective temperature teff is defined A. 2D antiferromagnet with easy-axis anisotropy through a renormalization constant which is again a function Several works ͑see Ref. 75 for a review͒ have shown of the pure-quantum fluctuations only! Such a remarkable that many additional interaction mechanisms may be taken and unexpected feature of the cutoff-corrected field-theory into account by inserting proper anisotropy terms in the mag- netic Hamiltonian; in particular, the transition observed in ͑ ͒ ϭ ͑ ͒ ϭ K2NiF4 Ref. 76 (S 1), Rb2FeF4 Ref. 76 (S 2), ͑ ͒ ͑ ͒ ϭ K2MnF4 Ref. 77 and Rb2MnF4 Ref. 76 (S 5/2), and others is seen to be possibly due to an easy-axis anisotropy. Such anisotropy has been often described in the literature through an external staggered magnetic field in order to al- low for a qualitative description of the experimental data. However, this choice lacks the fundamental property of de- scribing a genuine phase transition, as the field explicitly breaks the symmetry and makes the model ordered at all temperatures. To preserve the symmetry under inversion along the easy axis, it is actually appropriate to insert an exchange anisotropy term in the spin Hamiltonian. Sponta- neous symmetry breaking then manifests itself as a phase transition between ordered and disordered states. The EA- QHAF Hamiltonian then reads ␰ ␰ T JS2 Sϭ ␰ tH FIG. 13. Ratio / 3l versus / for 5/2. Solid line: cl( eff); dotted- J and-dashed line: ␰ Ӎ␰cl (tH ); dashed line: standard PQSCHA result; sym- Hϭ ͓␮͑ x x ϩ y y ͒ϩ z z ͔ ͑ ͒ H 3l eff ͚ Si Siϩd Si Siϩd Si Siϩd , 36 bols are QMC data.67 2 i,d Low Temp. Phys. 31 (8–9), August–September 2005 Balucani et al. 679

ϭ where i (i1 ,i2) runs over the sites of a square lattice, d connects each site to its four nearest neighbors, JϾ0 is the exchange integral, and ␮ is the EA anisotropy parameter (0 р␮Ͻ1). Again, JS˜ 2ϵJ(Sϩ1/2)2 sets the overall energy scale, and tϭT/JS˜ 2 is the reduced temperature. The model reduces to the isotropic QHAF when ␮ϭ1. Note that a ca- nonical transformation reversing the x and y spin compo- nents of one sublattice is equivalent to setting ␮→Ϫ␮,so that the physical properties of the model are even functions of ␮. The ␮ϭ0 case is called Ising limit, not to be confused with the genuine Ising model,78 reproduced by Eq. ͑36͒ with ␮ϭ0 and Sϭ1/2. Despite being a very particular case of Eq. ͑36͒, the 2D Ising model on the square lattice is a fundamen- tal point of reference for the study of the thermodynamic properties of the EA-QHAF. A renormalization-group 79 analysis of the classical model predicted the occurrence of ␮ cl FIG. 14. Critical temperature versus anisotropy for the 2D easy-axis an Ising-like transition at a finite temperature t (␮) of the cl ␮ c antiferromagnet. Dotted line: the fit of tc ( ) built up from classical MC order of unity for any value of ␮, no matter how near to the data82,83,87 ͑unfilled triangles͒. Solid lines: PQSCHA result for Sϭ1 and 5/2. isotropic value ␮ϭ1; this analysis has received the support Quantum MC data ͑unfilled circles90 and diamonds84͒ for Sϭ1/2, asymp- ␮ ϭ ͓ Ϫ␮ ͔ ͑ ͒ of several Monte Carlo simulations.80–83 totically described by tc( ) 2.49/ln 70/(1 ) dashed line , and exact result for the Ising model, ␮ϭ0 ͑unfilled square͒. Filled symbols are ex- As for the quantum case, up to a few years ago no in- ͑ ͒ 91 perimental data for the compounds YBa2Cu3O6 down triangle , K2NiF4 ͑ ͒ 92,93 ͑ ͒ 94 ͑ ͒ 95 formation was available about the value of the critical tem- cross , Rb2NiF4 up triangle , Rb2MnCl4 circle , Rb2MnF4 ␮ ͑ ͒ 92,96 perature tc( ,S) as a function of anisotropy and spin, save diamond . In the inset the region of weak anisotropy is enlarged. ϭ ͑ ͒78 the fact that tc(0,1/2) 0.567J Onsager solution and ϭ ͑ ͒ tc(1,S) 0 isotropic limit . As a consequence, it was also uncertain whether or not the small anisotropy (1Ϫ␮ temperature decreases very slowly ͑logarithmically͒ towards Ϫ2 Ӎ10 ) observed in real compounds could be responsible its vanishing value in the isotropic limit, so that tc remains for transitions occurring at critical temperatures of the order substantially of the order of unity. of J, also accounting for the fact that quantum fluctuations As a sample of the various results that were obtained, in are expected to lower the critical temperature with respect to Fig. 15 we report the comparison of the results of Ref. 85 97 the classical case. with the experimental data for the correlation length of the ϭ Over the last few years our group developed a quantita- S 5/2 magnet Rb2MnF4 , which results are quite well de- tive analysis of several thermodynamic properties of the scribed by the anisotropic model with Jϭ7.42 K and ␮ 65,66 ϭ model, by means of the effective Hamiltonian method for 0.9942. Rb2MnF4 is known to behave as a 2D magnet 76 spin Sу1 and by means of quantum Monte Carlo both above and below the observed transition, so that the simulations84 in the case Sϭ1/2. critical behavior is not contaminated by the onset of 3D or- The effective Hamiltonian85–87 for the EA-QHAF is ex- der and a clean characterization of the transition is possible. pressed as In Ref. 85, we have compared our theoretical results also with the neutron-scattering experimental data for the stag- H j gered magnetization, staggered susceptibility, and correlation eff ϭ eff ͓␮ ͑ x x ϩ y y ͒ϩ z z ͔ ͑ ͒ ͚ eff si siϩd si siϩd si siϩd , 37 length of Rb MnF and found excellent agreement both for ˜ 2 2 i,d 2 4 JS the overall temperature behavior and for the value of the ␮ Ͻ and shows a weaker renormalized exchange jeff(t, ) 1 and ␮ ␮ Ͼ␮ easy-axis anisotropy eff(t, ) , besides an additional free- H energy term that is not reported. By means of eff a series of thermodynamic quantities was studied:85,88 internal energy, specific heat, staggered magnetization, staggered correlation function, staggered correlation length, staggered susceptibil- ity. This required extensive classical Monte Carlo simula- tions, as varying the temperature gives an effective system ␮ ␮ with different effective anisotropy eff(t, ). The quantum phase diagram reported in Fig. 14 could be built up87,89 in a simpler way starting from the knowledge of the classical phase diagram and using a relation that follows from the form of Eq. ͑37͒, ͑␮ ͒ϭ ͑ ␮ ͒ cl͑␮ ͑ ␮ ͒͒ ͑ ͒ tc ,S jeff t, ,S tc eff t, ,S . 38 ϭ ␮ϭ ͑ ͒ In the region of very weak anisotropy, which is the most FIG. 15. Correlation length versus t for S 5/2, 0.9942 full curve and ␮ϭ1 ͑isotropic, dotted-and-dashed curve͒; the symbols are neutron scatter- important in view of the characterization of experimentally 97 98 ing data for Rb2MnF4 . The triangles are quantum Monte Carlo data for accessible materials, we verified that the Ising-like transition the isotropic model. 680 Low Temp. Phys. 31 (8–9), August–September 2005 Balucani et al.

quantity, undoubtedly shows a characteristic anisotropic be- havior with a different temperature dependence of the trans- verse and longitudinal branches: the former displays a mini- mum and the latter monotonically goes to zero, as expected for an EA antiferromagnet. This behavior results from the anisotropy-induced spin ordering, which makes the system more sensitive to the ap- plication of a transverse magnetic field than a longitudinal one. Both the minimum of the in-plane component and the decrease of the longitudinal component are close to the tran- sition, a feature also peculiar to the Ising model. Results for the critical temperature at Sϭ1/2 are already included in Fig. 14.

B. 2D antiferromagnet with easy-plane anisotropy FIG. 16. Specific heat versus tϭT/JS˜ 2 for Sϭ5/2. Mn-f-2U experiments 100 87,88 ͑squares͒, EA-QHAF with ␮ϭ0.9942 ͑solid line͒, isotropic QHAF In the case of an easy-plane anisotropy the Mermin- ͑dashed line͒. Note that the correct anisotropy for this compound is esti- mated to be ␮ϭ0.981. Error bars are due to the experimental uncertainly on Wagner theorem still holds, so that no finite-temperature J for Mn-f-2U. transition to a phase with a finite order parameter may occur. However, a BKT transition,102,103 related to the existence of vortex-like topological excitations, is known to characterize critical temperature, which coincides perfectly with that de- the class of the easy-plane models. The reference system for ϭ ͑ rived from the experimental analysis, Tc 38.4 K i.e., tc the easy-plane class is the planar rotator model,orXY ϭ0.575). model, defined in terms of two-component spins: above the Another quantity that shows a signature of the anisot- critical temperature tc the system is disordered, with expo- ropy and of the Ising transition is the specific heat: in Fig. nentially decaying correlation functions; in the region 0Ͻt Ͻ 16, a comparison with experimental data is shown in the case tc the system is in a critical phase with vanishing magne- of the Sϭ5/2 compound Mn-formate di-urea ͑Mn–f–2U͒,88 tization and power-law decaying correlators ͑quasi-long- whose anisotropy can be estimated, solely from knowledge range order͒;attϭ0 the magnetization acquires a finite of the exchange integral and the measured transition tem- value and the system is ordered. perature, to be ␮ϭ0.981. The comparison reveals the exis- The observation of clear signatures of BKT critical be- tence of a crossover from a high-temperature 2D-Heisenberg havior in real magnets is a controversial issue. However, it regime to a critical 2D-Ising regime that triggers the can explain the properties of several layered 99 ϭ 53,104–106 observed 3D phase transition at TN 3.77 K. compounds whose high-temperature phase can be Finally, for the strongest quantum case, Sϭ1/2, we have described by a purely 2D Heisenberg Hamiltonian, with an used the continuous-time quantum Monte Carlo method exchange interaction often displaying weak EP anisotropies, based on the loop algorithm.84 The general outcome of the on the order of 10Ϫ2 –10Ϫ4 times the dominant isotropic numerical simulations is that the thermodynamics of 2D coupling.105,106 Symmetry and universality arguments sug- quantum antiferromagnets is extremely sensitive to the pres- gest that the EP anisotropy drives the system towards a BKT ence of weak easy-axis anisotropies of the order of those of behavior at finite temperature, and the enhanced intraplane real compounds. For instance, in Fig. 17 it is shown that for correlations trigger the transition to the observed 3D ordered ␮ϭ0.99 the uniform susceptibility, which is a noncritical state. As a consequence, 2D critical behavior in close prox- imity of the would-be BKT transition is masked by these 3D effects. The EP-QHAF Hamiltonian reads J Hϭ ͓ x x ϩ y y ϩ␭ z z ͔ ͑ ͒ ͚ Si Siϩd Si Siϩd Si Siϩd , 39 2 i,d where ␭ is the easy-plane anisotropy parameter (0р␭Ͻ1). Again, JS˜ 2ϵJ(Sϩ1/2)2 sets the overall energy scale and t ϭT/JS˜ 2 is the reduced temperature. When ␭ϭ1 the model reduces to the isotropic QHAF. Note that a canonical trans- formation reversing the x and y spin components of one sublattice is equivalent to setting (J,␭)→(ϪJ,Ϫ␭), so that negative values of ␭ (Ϫ1Ͻ␭р0) correspond to the EP fer- romagnet. The ␭ϭ0 case is called the XY model or XX0 FIG. 17. Uniform susceptibility of the EA model for ␮ϭ0.99 from quantum model. However, at variance with the planar rotator model, Monte Carlo simulations.84 Filled diamonds: longitudinal branch; unfilled 101 out-of-plane fluctuations are present both in the classical and diamonds: transverse branch; stars: data for the isotropic model. The lines ͉␭͉Ͻ are a guide to the eye. The arrow indicates the estimated critical tempera- in the quantum EP models. Nevertheless, if 1, the clas- ture. sical EP model still undergoes a BKT phase transition.107 Low Temp. Phys. 31 (8–9), August–September 2005 Balucani et al. 681

␹ FIG. 19. Out-of-plane (zz) and in-plane (xx) uniform susceptibility u for ␭ϭ0.999. Diamonds are our QMC results; crosses are experimental 119,120 data for Sr2CuO2Cl2 . The circles report the result for the isotropic ϭ QHAF. The vertical lines mark the 3D transition temperature tc 0.176 and ϭ FIG. 18. BKT critical temperature versus anisotropy ␭ for the easy-plane the crossover temperature tco 0.227. model with Sϭ1/2,1,5/2,ϱ. The triangles report the classical (Sϭϱ)MC 110 cl ␭ data used to construct the curve for tc ( ) and hence the renormalized curves for Sϭ1 and 5/2, the curves obtained through the Holstein-Primakoff (␭Շ0.5) and the Villain (␭տ0.5) spin-boson transformation, are seen to be the BKT transition temperature stays large ͑i.e., comparable 84 connected in a fairly smooth way. The diamonds are our QMC data for to the exchange constant͒ also for very weak EP anisotropy. Sϭ1/2, while the circles are earlier QMC results.113 The inset expands the ␭ However, as explained above, the problem of detecting nearly isotropic region, in which the expected logarithmic behavior tc( ) ϳ(aϪln͉1Ϫ␭͉)Ϫ1 is fitted by the dashed curve. the incipient BKT transition requires looking for signatures of XY behavior in a region above the transition. We have shown115 that a suitable quantity is the uniform susceptibility ␹␣␣ ␣ϭ ␣ϭ Monte Carlo simulations of the classical systems108–111 con- u , which has in-plane ( x,y) and out-of-plane ( z) firm that the planar and the XXZ model share the same quali- components and is noncritical, i.e., it does not show singu- ␹zz tative behavior, but the value of the transition temperature of larities at tc . Figure 19 shows indeed that u deviates from ␹ 112 110,111 the isotropic u and displays a minimum. A similar feature is the planar model shrinks by 22% in the XX0 model 84,115 ͑i.e., with ␭ϭ0), as a consequence of out-of-plane fluctua- also present in other quantities and occurs around a tem- ␭ tions. A renormalization group calculation107 predicts that the perature tco( ) that can be generally defined as the minimum ␹zz ␭ ␹zz transition temperature vanishes logarithmically as the isotro- of u (t, ). The pronounced deviation of u from the iso- pic limit ͉␭͉→1 is approached, and this was also verified in tropic behavior is due to a simple statistical reason. The uni- classical MC simulations.110 Experiments53,54 and quantum form susceptibility arises from spin canting: two antiferro- MC simulations113 indicated that the qualitative behavior of magnetically coupled spins in an infinitesimal magnetic field the BKT transition is preserved in the quantum system, with h minimize their energy when they are located almost or- only quantitative modifications of the critical parameters due thogonally to h and slightly canted in its direction, thus giv- to the quantum fluctuations. ing a linear response; if they are locally parallel to h the տ We applied the effective Hamiltonian formalism65,66 to response is instead negligible. When for t tco the anisot- the EP-QHAF, finding that it was necessary87,114 to resort to ropy becomes effective, the fraction of spins aligned in the the Villain or to the Holstein-Primakoff transformation, de- EP rapidly increases compared to that of the isotropic case ϳ ͑ ␹zz pending on whether the anisotropy is strong or weak, respec- ( 2/3), and the response to a field along z i.e., u )is ͑ proportionally larger. tively. While the above approach gives reliable results with 53 a smooth enough connection at intermediate anisotropy͒ for The layered cuprate Sr2CuO2Cl2 is a good realization ϭ spin Sу1, we adopted quantum Monte Carlo of the EP-QHAF model, with J 1450 K; considering the simulations84,88,115–117 in the case Sϭ1/2. The effective spin-wave gap renormalization, its bare anisotropy is esti- 87,114,118 mated to be 1Ϫ␭Ӎ0.0014. Experimental data for the uni- Hamiltonian for the EP-QHAF, in terms of classical 119,120 spins, takes the form form susceptibility of this compound are reported in Fig. 19. They agree excellently with our results for ␭ H j ϭ0.999: the position of the minimum of ␹zz gives t eff ϭ eff ͓ x x ϩ y y ϩ␭ z z ͔ ͑ ͒ u co ͚ si siϩd si siϩd effsi siϩd , 40 ϭ0.227(15). Close to the critical region the experimental JS˜ 2 2 i,d data are affected by the 3D nature of the ordering of the real ␭ Ͻ ϭ and displays a weaker renormalized exchange jeff(t, ) 1 and magnet: the Ne´el transition is observed at tN 0.176(10) and ␭ ␭ Ͼ␭͑ Ϫ␭ easy-plane anisotropy eff(t, ) an additional free-energy compares well with the theoretical estimate tc(1 term is not reported͒. In analogy to the EA case, the BKT ϭ0.0014)ϭ0.179(10), confirming that 3D ordering is in- transition temperature can be obtained by renormalization of duced by the incipient intralayer BKT transition. the classical one using the self-consistent relation It is worthwhile to mention that in the triangular lattice the easy-plane antiferromagnet has very peculiar behavior, t ͑␭,S͒ϭ j ͑t,␭,S͒tcl͑␭ ͑t,␭,S͒͒. ͑41͒ c eff c eff already at the classical level, due to the frustration effect of In Fig. 18, the phase diagram of the EP-QHAF is re- accommodating three antiferromagnetic spins on a plaquette. ported, including the QMC results for Sϭ1/2. It is seen that Indeed, the minimum energy corresponds to a configuration 682 Low Temp. Phys. 31 (8–9), August–September 2005 Balucani et al. with the three sublattices aligned in the easy plane at equal angles 2␲/3. As clockwise and counterclockwise plaquette vorticities are possible, this configuration is twofold degen- erate and there is chiral symmetry which corresponds to an Ising-like order parameter. Therefore both a BKT and an Ising transition coexist in the system. We have studied the triangular antiferromagnet both in the classical121 and in the quantum122 case, constructing the phase diagram for varying anisotropy and showing that the transitions occur at slightly different temperatures.

C. 2D antiferromagnet in an applied Zeeman field FIG. 20. Phase diagram of the Sϭ1/2 2D QHAF in a magnetic field. Un- filled symbols refer to the classical limit of the model;123 the triangle126 and Interesting behavior is shown by the 2D Heisenberg an- the squares124 are QMC results. tiferromagnet when a magnetic field is applied. Indeed, a frustration phenomenon occurs, as antiferromagnetism tends to antialign spins while the field tends to align them with VI. FRUSTRATION IN THE 2D QUANTUM J1 – J2 123 itself: in the classical system it appears that the minimum- HEISENBERG MODEL energy configuration is the one with the spins almost or- thogonal to the field and canted in its direction. Therefore, The study of frustrated quantum spin systems is one of provided the field is not strong enough to overcome the ex- the most challenging and exciting topics in theoretical mag- change and to saturate the magnetization, it acts as an effec- netism because of the possible existence of a nonmagnetic tive easy-plane anisotropy, and one expects to observe BKT zero-temperature phase. A very extensively investigated, yet behavior. Remarkably, as this can be induced in a real largely debated model is the so-called J1 – J2 Heisenberg quasi-2D antiferromagnetic system by means of an applied model with competing antiferromagnetic couplings (J1 ,J2 Ͼ ͑ ͒ field, the strength of the effective anisotropy is in this case 0) between nearest neighbors nn and next-nearest neigh- ͑ ͒ tunable. Even though 2D criticality just acts as a trigger for bors nnn 3D ordering, by observing that the critical temperature fol- Hϭ ϩ ͑ ͒ lows the predicted behavior with field, an experiment could J1͚ Si•Sj J2͚ Si•Sj , 43 nn nnn realize an objective observation of genuine 2D behavior. The 2D QHAF in a uniform magnetic field is described where the spin operators are defined on a periodic lattice ϭ ϫ ␣ϭ by the Hamiltonian with N L L sites; hereafter J2 /J1 defines the frustra- tion ratio. In the classical limit (S→ϱ) the minimum energy con- J Hϭ Ϫ ␮ z ͑ ͒ figuration has conventional Ne´el order with magnetic wave ͚ Si•Siϩd g BH•͚ Si , 42 2 i,d i vector Qϭ(␲,␲) for ␣Ͻ0.5. Instead, for ␣Ͼ0.5, the anti- ferromagnetic order is established independently in the two H ␮ where is the applied Zeeman field, B is the Bohr mag- sublattices, with the two staggered magnetizations free to neton, and g is the gyromagnetic ratio. rotate with respect to each other. One of the two families of We have studied124,125 the Sϭ1/2 2D QHAF in an uni- collinear states, with pitch vectors Qϭ(␲,0) or ͑0,␲͒, are form magnetic field by means of the QMC method based on selected by an order-by-disorder mechanism as soon as ther- the worm algorithm. Our results confirmed that an arbitrarily mal or quantum fluctuations are taken into account. As a small field is able to induce a BKT transition and an ex- result, for ␣Ͼ0.5 the classical ground state breaks not only tended XY phase above it, as in the case of an easy-plane the spin rotational and translational invariance of the exchange anisotropy. The field-induced XY behavior be- Hamiltonian—as the conventional Ne´el phase—but also its comes more and more marked for increasing fields, while for invariance under ␲/2 lattice rotations, the resulting degen- ϫ strong fields the antiferromagnetic behavior along the field eracy corresponding to the group O(3) Z2 . Remarkably, axis is nearly washed out, so that the system behaves as a the additional discrete Z2 symmetry can, in principle, be bro- planar rotator model with antiferromagnetism surviving in ken at finite temperatures without violating the Mermin- the orthogonal plane only; the BKT critical temperature, as Wagner theorem. On this basis, in a seminal paper,127 Chan- ͑ ϵ ϵ ␮ ͑ ͒ reported in Fig. 20 where t T/J and h 2g BH/J), van- dra, Coleman, and Larkin CCL proposed that the 2D J1 – J2 ishes as the field reaches the saturation value hc and the model could sustain an Ising phase transition at finite tem- effective rotator length goes to zero. We have therefore perature, with an order parameter directly related to the Z2 shown that the model in a moderately strong field represents degree of freedom induced by frustration. They also pro- an ideal realization of the XY model and that XY behavior vided quantitative estimates of the critical temperatures in can be detected by measuring standard noncritical quantities, the large-␣ limit for both the classical and the quantum such as the specific heat or the induced magnetization. An cases. experimental realization of the XY model in purely magnetic This transition in the classical model has been estab- systems and a systematic investigation of the dynamics of lished by an extensive Monte Carlo investigation.128 In the vortex/antivortex excitations is therefore possible. quantum case, the occurrence of a low-temperature phase Low Temp. Phys. 31 (8–9), August–September 2005 Balucani et al. 683 with a discrete broken symmetry has been the subject of debates in connection with the discovery of three vanadate compounds (Li2VOSiO4 ,Li2VOGeO4 , and VOMoO4) whose relevant magnetic interactions involve nearest and next-nearest spin-1/2 V4ϩ ions on weakly coupled stacked planes. In particular, NMR and ␮SR measurements on ͑ ͒ Li2VOSiO4 Ref. 129 indicate the occurrence of a transition to a low-temperature phase with collinear order at TN Ӎ2.8 K. However, in the experiments with vanadate com- pounds, structural distortions and interlayer and anisotropy effects are likely to come into play,130 and on the other hand the theoretical investigation cannot rely on the insight pro- vided by quantum Monte Carlo methods, as their reliability in the presence of frustration is strongly limited ͑see the re- view articles in Refs. 131 and 132͒. FIG. 21. Renormalized critical temperature of the CCL transition for various A complete study of the thermodynamic properties of the values of the spin. Classical data ͑*͒ are taken from Ref. 128. The solid lines on the right are the CCL and the CCL* prediction for the classical and S quantum J1 – J2 model in its collinear phase has been pur- ϭ ͑ ͒ 65 1/2 case see text . The ellipse on the horizontal axis marks the nonmag- sued within the PQSCHA scheme described in Sec. 4, by ͑ ͒ ␣ ϭϱ ϭ ␣ ϭ Ӎ netic spin-liquid phase between c(S ) 0.5 and c(S 1/2) 0.6. which the thermodynamics is rephrased in terms of a classi- cal effective Hamiltonian with renormalized parameters de- ϭ ␣ ϭ pending on the spin value, temperature, and frustration. It is tration. In particular for S 1/2, c 0.6, in agreement with possible to show that, to O(1/S), the effective Hamiltonian the previous estimates of the zero-temperature quantum criti- can be recast in a form preserving all the symmetries of the cal point.132 original model, and that reads ͑except for uniform terms͒: VII. CONCLUSIONS Heffϭ eff˜ 2 ϩ eff˜ 2 ͑ ͒ J1 S ͚ si•sj J2 S ͚ si•sj , 44 The activity in magnetism of the Condensed Matter nn nnn Theory group in Florence134 stems from the early work on ˜ ϭ ϩ two-magnon Raman scattering in the seventies, and has where si are classical vectors of length 1, S S 1/2 is the effϭ ␪2ϩ␪2 ␪2 effϭ␪4 grown up over the years with the collaboration of several effective spin length, and J1 ( x y) 2J1/2 and J2 2J2 are the quantum-renormalized exchange integrals, with spin-, scientists. In this paper we summarize the relevant theoreti- temperature-, and frustration-dependent renormalization pa- cal work that concerns antiferromagnetic models. This activ- ␪ ␪ ␪ ity has been mainly concentrated on low-dimensional sys- rameters x , y , and 2 . The occurrence of the transition127 in the quantum case tems and has found one of its main motivations in the intent can be directly addressed within our approach by calculating of interpreting the data collected in experiments on real ma- the critical temperatures as functions of the spin and of the terials. Among the prominent subjects, we have reported frustration ratio.133 Using a simple scaling argument, one can about soliton-excitation effects in one-dimensional systems ␣ and critical and near-critical behavior and phase transitions relate the critical temperatures in the quantum case Tc(S, ) cl ␣ in two-dimensional systems. Besides that, we have also ad- to those of the classical model Tc ( ) through the following self-consistent relation:87,122 dressed some intriguing theoretical problems where funda- mental aspects of quantum mchanics come into play, as, e.g., ͑ ␣͒ϭ eff͑ ␣͒ cl͑␣eff͑ ␣͒͒ ͑ ͒ the ground state of antiferromagnetic chains with integer Tc S, j1 T,S, Tc Tc ,S, , 45 spin or the possible quantum critical regime predicted from effϭ eff˜2 ␣effϭ eff eff the field-theory treatment of the two-dimensional antiferro- where j1 J1 S /J1 and J2 /J1 . The classical transition cl ␣ magnetic Heisenberg model. temperature, Tc ( ) is accurately known through extensive MC simulations for ␣р2; it vanishes for ␣→2 and grows more or less linearly for ␣Ͼ1. *E-mail: [email protected] The behavior of the transition temperature versus frus- tration ratio is plotted in Fig. 21 for different values of the spin length. In order to represent the whole interval of ␣ 1F. D. M. Haldane, Phys. Rev. Lett. 50, 1153 ͑1983͒; Phys. Rev. A 93, 464 ෈͓1/2,ϱ͔, in Fig. 21 we have plotted both the MC and the ͑1983͒. 2A. Sokol and D. Pines, Phys. Rev. Lett. 71,2813͑1993͒. CCL estimates of the classical critical temperatures as a 3S. Chakravarty, in High Temperature Superconductivity, K. S. Bedell, function of ␣/(1ϩ␣). The mismatch between the MC and D. Coffey, D. E. Meltzer, D. Pines, and J. R. Schrieffer ͑eds.͒, Addison CCL predictions is a minor flaw that can be corrected by Wesley, Redwood City ͑1990͒. 4 ͑ ͒ slightly modifying CCL’s criterion for the determination of N. D. Mermin and H. Wagner, Phys. Rev. Lett. 17,1133 1966 . 5P. A. Fleury, Int. J. Magnetism 1,75͑1970͒. the transition temperature, as explained in Ref. 128. Remark- 6R. Loudon, Adv. Phys. 17, 243 ͑1968͒; M. G. Cottam and D. J. Lock- ably, while for large ␣ the transition temperature vanishes for wood, Light Scattering in Magnetic Solids, Wiley, New York ͑1986͒. ␣→ϱ for any value of the spin, in the opposite limit the 7P. A. Fleury, Phys. Rev. 180, 591 ͑1969͒. 8U. Balucani and V. Tognetti, Riv. Nuovo Cimento 6,39͑1976͒. critical temperatures vanish on approaching a critical value 9 ͑ ͒ ␣ Ͼ R. J. Elliott and M. Thorpe, J. Phys. C 2, 1630 1969 . c 0.5 that increases as S decreases, thus confirming the 10U. Balucani and V. Tognetti, Phys. Rev. B 8, 4247 ͑1973͒. existence of a nonmagnetic phase in the regime of high frus- 11U. Balucani, M. H. Lee, and V. Tognetti, Phys. Rep. 373,409͑2003͒. 684 Low Temp. Phys. 31 (8–9), August–September 2005 Balucani et al.

12F. Barocchi, P. Mazzinghi, V. Tognetti, and M. Zoppi, Solid State Com- 60P. Hasenfratz and F. Niedermayer, Phys. Lett. B 268, 231 ͑1991͒. mun. 25,241͑1978͒. 61N. Elstner, A. Sokol, R. R. P. Singh, M. Greven, and R. J. Birgeneau, 13S. R. Chinn, R. W. Davies, and H. J. Zeiger, AIP Conf. Proc. 5, 317 Phys. Rev. Lett. 75, 938 ͑1995͒; N. Elstner, Int. J. Mod. Phys. B 11, 1753 ͑1972͒. ͑1997͒. 14H. Mori, Prog. Theor. Phys. 33, 423 ͑1965͒; 34, 399 ͑1965͒. 62A. Cuccoli, V. Tognetti, P. Verrucchi, and R. Vaia, Phys. Rev. Lett. 77, 15U. Balucani and V. Tognetti, Phys. Rev. B 16,271͑1977͒. 3439 ͑1996͒. 16U. Balucani, M. G. Pini, and V. Tognetti, Nuovo Cimento B 46,81 63A. Cuccoli, V. Tognetti, P. Verrucchi, and R. Vaia, Phys. Rev. B 56, 14456 ͑1978͒. ͑1997͒. 17V. E. Korepin, N. M. Bogoliubov, and A. G. Izergin, Quantum Inverse 64A. Cuccoli, V. Tognetti, P. Verrucchi, and R. Vaia, Phys. Rev. B 58, 14151 Scattering Method and Correlation Functions, Cambridge Univ. Press ͑1998͒. ͑1993͒. 65A. Cuccoli, R. Giachetti, V. Tognetti, R. Vaia, and P. Verrucchi, J. Phys.: 18H. J. Mikeska, J. Phys. C 11, L29 ͑1978͒; 13, 2913 ͑1980͒. Condens. Matter 7, 7891 ͑1995͒. 19F. Currie, J. A. Krumhansl, A. R. Bishop, and S. E. Trullinger, Phys. Rev. 66A. Cuccoli, V. Tognetti, P. Verrucchi, and R. Vaia, Phys. Rev. A 45,8418 B 22,477͑1980͒. ͑1992͒. 20T. Schneider and E. Stoll, Phys. Rev. B 22, 5317 ͑1980͒. 67B. B. Beard, A. Cuccoli, R. Vaia, and P. Verrucchi, Phys. Rev. B 68, 21 F. Borsa, M. G. Pini, A. Rettori, and V. Tognetti, Phys. Rev. B 28, 5173 104406 ͑2003͒. ͑1983͒. 68P. Hasenfratz, Eur. Phys. J. B 13,11͑2000͒. 22 S. W. Lovesey, U. Balucani, F. Borsa, and V. Tognetti ͑eds.͒, Magnetic 69M. Falcioni and A. Treves, Nucl. Phys. B 265,671͑1986͒; S. Caracciolo Excitations and Fluctuations, Vol. 54 of Springer Series in Solid State, and A. Pelissetto, Nucl. Phys. B 420,141͑1994͒. Springer Verlag, Berlin ͑1984͒. 70P. Carretta, A. Rigamonti, and R. Sala, Phys. Rev. B 55, 3734 ͑1997͒. 23 E. H. Lieb, T. D. Schultz, and D. C. Mattis, Ann. Phys. ͑N.Y.͒ 16, 407 71J.-K. Kim, D. P. Landau, and M. Troyer, Phys. Rev. Lett. 79, 1583 ͑1997͒; ͑1961͒. private communication. 24 I. Affleck, T. Kennedy, E. H. Lieb, and H. Tasaki, Phys. Rev. Lett. 59, 799 72M. S. Makivic´ and H.-Q. Ding, Phys. Rev. B 43, 3562 ͑1991͒. ͑ ͒ ͑ ͒ 1987 ; Commun. Math. Phys. 115,477 1988 . 73B. B. Beard, R. J. Birgeneau, M. Greven, and U.-J. Wiese, Phys. Rev. 25 D. P. Arovas, A. Auerbach, and F. D. M. Haldane, Phys. Rev. Lett. 60, 531 Lett. 80, 1742 ͑1998͒; B. B. Beard, Nucl. Phys. Proc. Suppl. 83–84,682 ͑ ͒ 1988 . ͑2000͒. 26 ͑ ͒ I. Affleck, J. Phys.: Condens. Matter 1, 3407 1989 . 74K. Harada, M. Troyer, and N. Kawashima, J. Phys. Soc. Jpn. 67, 1130 27 ͑ ͒ M. Takahashi, Phys. Rev. Lett. 62, 2313 1989 . ͑1998͒. 28 W. J. L. Buyers, R. M. Morra, R. L. Armstrong, M. J. Hogan, P. Gerlach, 75J. Kanamori, in Magnetism, Vol. 1, G. T. Rado and H. Suhl ͑eds.͒, Aca- ͑ ͒ and K. Hirakawa, Phys. Rev. Lett. 56, 371 1986 . demic Press, New York ͑1963͒. 29 M. Steiner, K. Kakurai, J. K. Kjems, D. Petitgrand, and R. Pynn, J. Appl. 76R. J. Birgeneau, H. J. Guggenheim, and G. Shirane, Phys. Rev. B 1, 2211 Phys. 61, 3953 ͑1987͒. ͑ ͒ 30 1970 . M. Date and K. Kindo, Phys. Rev. Lett. 63, 1424 ͑1990͒. 77 31 R. J. Birgeneau, H. J. Guggenheim, and G. Shirane, Phys. Rev. B 8,304 P. Gaveau, J. P. Boucher, L. P. Regnault, and J. P. Renard, Europhys. Lett. ͑1973͒. 12,647͑1990͒. 78 ͑ ͒ 32 L. Onsager, Phys. Rev. 65,117 1944 . L. C. Brunel, T. Brill, I. Zaliznyak, J. P. Boucher, and J. P. Renard, Phys. 79M. Bander and D. L. Mills, Phys. Rev. B 38, 12015 ͑1988͒. ͑ ͒ Rev. Lett. 69, 1699 1992 . 80J. D. Patterson and G. L. Jones, Phys. Rev. B 3,131͑1971͒. 33 ͑ ͒ R. Botet, R. Jullien, and M. Kolb, Phys. Rev. B 28, 3914 1983 . 81K. Binder and L. P. Landau, Phys. Rev. B 13, 1140 ͑1976͒. 34 ͑ ͒ J. B. Parkinson and J. C. Bonner, Phys. Rev. B 32, 4703 1985 . 82P. A. Serena, N. Garcı´a, and A. Levanyuk, Phys. Rev. B 47, 5027 ͑1993͒. 35 ¨ ͑ ͒ M. P. Nightingale and H. W. J. Blote, Phys. Rev. B 33, 659 1986 . 83M. E. Gouveˆa, G. M. Wysin, S. A. Leonel, A. S. T. Pires, T. Kamppeter, 36 2 ͑ ͒ T. Kennedy, J. Phys.: Condens. Matter , 5737 1990 . and F. G. Mertens, Phys. Rev. B 59, 6229 ͑1999͒. 37T. Sakai and M. Takahashi, Phys. Rev. B 42, 1090 ͑1990͒. 84A. Cuccoli, T. Roscilde, V. Tognetti, P. Verrucchi, and R. Vaia, Phys. Rev. 38T. Delica, K. Kopinga, H. Leschke, and K. K. Mon, Europhys. Lett. 15, B 67, 104414 ͑2003͒. 55 ͑1991͒. 85 39 ͑ ͒ A. Cuccoli, V. Tognetti, T. Roscilde, P. Verrucchi, and R. Vaia, Phys. Rev. H. Koehler and R. Shilling, J. Phys.: Condens. Matter 4,7899 1992 . ͑ ͒ 40 B 62, 3771 2000 . H.-J. Mikeska, Chaos, Solitons, Fractals 5, 2585 ͑1995͒. 86 41 ͑ ͒ A. Cuccoli, T. Roscilde, V. Tognetti, R. Vaia, and P. Verrucchi, J. Magn. H.-J. Mikeska, Europhys. Lett. 19,39 1992 . ͑ ͒ 42 ͑ ͒ Magn. Mater. 226–230, 562 2001 . G. Gomez-Santos, Phys. Rev. Lett. 63, 790 1989 . 87 43 ͑ ͒ A. Cuccoli, T. Roscilde, V. Tognetti, R. Vaia, and P. Verrucchi, Eur. Phys. T. Kennedy and H. Tasaki, Phys. Rev. B 45,304 1992 ; Commun. Math. ͑ ͒ ͑ ͒ J. B 20,55 2001 . Phys. 147, 431 1992 . 88 44 ͑ ͒ A. Cuccoli, T. Roscilde, V. Tognetti, R. Vaia, and P. Verrucchi, J. Appl. M. den Nijs and K. Rommelse, Phys. Rev. B 50, 3907 1994 . ͑ ͒ 45 ͑ ͒ Phys. 93, 7637 2003 . H. Tasaki, Phys. Rev. Lett. 66,798 1991 . 89 46H.-J. Mikeska and P. Verrucchi, Phys. Rev. B 52, 3571 ͑1995͒. A. Cuccoli, V. Tognetti, T. Roscilde, P. Verrucchi, and R. Vaia, Braz. 47 ͑ ͒ J. Phys. 30, 697 ͑2000͒. O. Golinelli, Th. Joliceur, and R. Lacaze, Phys. Rev. B 50, 3037 1994 . 90 48 ͑ ͒ H.-Q. Ding, J. Phys.: Condens. Matter 2, 7979 ͑1990͒. N. Elstner and H.-J. Mikeska, Phys. Rev. B 50, 3907 1994 . 91 49E. J. Neves and J. F. Perez, Phys. Lett. 114A,331͑1986͒. J. Rossat-Mignod, L. P. Regnault, C. Vettier, P. Burlet, J. Y. Henry, and ͑ ͒ 50E. Manousakis, Rev. Mod. Phys. 63,1͑1991͒. G. Lapertot, Physica B 169,58 1991 . 92 51M. A. Kastner, R. J. Birgeneau, G. Shirane, and Y. Endoh, Rev. Mod. H. W. de Wijn, L. R. Walker, and R. E. Walstedt, Phys. Rev. B 8,285 ͑ ͒ Phys. 70,897͑1998͒. 1973 . 93 52M. Greven, R. J. Birgeneau, Y. Endoh, M. A. Kastner, B. Keimer, J. Skalyo, Jr., G. Shirane, R. J. Birgeneau, and H. J. Guggenheim, Phys. M. Matsuda, G. Shirane, and T. R. Thurston, Phys. Rev. Lett. 72, 1096 Rev. Lett. 23, 1394 ͑1969͒. 94 ͑1994͒. K. Nagata and Y. Tomono, J. Phys. Soc. Jpn. 36,78͑1974͒. 95 53M. Greven, R. J. Birgeneau, Y. Endoh, M. A. Kastner, M. Matsuda, and B. Schro¨der, V. Wagner, N. Lehner, K. A. M. Kesharwani, and R. Geick, G. Shirane, Z. Phys. B: Condens. Matter 96, 465 ͑1995͒. Phys. Status Solidi B 97, 501 ͑1980͒. 54K. Nakajima, K. Yamada, S. Hosoya, Y. Endoh, M. Greven, and R. J. 96R. A. Cowley, G. Shirane, R. J. Birgeneau, and H. J. Guggenheim, Phys. Birgeneau, Z. Phys. B: Condens. Matter 96,479͑1995͒. Rev. B 15, 4292 ͑1977͒. 55Y. S. Lee, M. Greven, B. O. Wells, R. J. Birgeneau, and G. Shirane, Eur. 97Y. S. Lee, M. Greven, B. O. Wells, R. J. Birgeneau, and G. Shirane, Eur. Phys. J. B 5,15͑1998͒. Phys. J. B 5,15͑1998͒. 56S. Fulton, S. E. Nagler, L. M. N. Needham, and B. M. Wanklyn, J. Phys.: 98B. B. Beard, Nucl. Phys. B, Proc. Suppl. 83–84, 682 ͑2000͒. Condens. Matter 6, 6667 ͑1994͒; S. Fulton, R. A. Cowley, A. Desert, and 99K. Takeda, H. Deguchi, T. Hoshiko, and K. Yamagata, J. Phys. Soc. Jpn. T. Mason, J. Phys.: Condens. Matter 6, 6679 ͑1994͒. 58, 3489 ͑1989͒. 57P. Carretta, T. Ciabattoni, A. Cuccoli, E. Mognaschi, A. Rigamonti, 100K. Takeda, M. Mito, K. Nakajima, K. Kakurai, and K. Yamagata, Phys. V. Tognetti, and P. Verrucchi, Phys. Rev. Lett. 84,366͑2000͒. Rev. B 63, 024425 ͑2001͒. 58S. Chakravarty, B. I. Halperin, and D. R. Nelson, Phys. Rev. B 39, 2344 101J.-K. Kim and M. Troyer, Phys. Rev. Lett. 80, 2705 ͑1998͒. ͑1989͒. 102V. L. Berezinskii, Zh. E´ ksp. Teor. Fiz. 59, 907 ͑1970͓͒Sov. Phys. JETP 59I. Affleck, Nucl. Phys. B 257,397͑1985͒. 32, 493 ͑1971͔͒. Low Temp. Phys. 31 (8–9), August–September 2005 Balucani et al. 685

103J. M. Kosterlitz and D. J. Thouless, J. Phys. C 6, 1181 ͑1973͒;J.M. 120D. Vaknin, S. K. Sinha, C. Stassis, L. L. Miller, and D. C. Johnston, Phys. Kosterlitz, J. Phys. C 7, 1046 ͑1974͒. Rev. B 41, 1926 ͑1990͒. 104B. J. Suh, F. Borsa, L. L. Miller, M. Corti, D. C. Johnston, and D. R. 121L. Capriotti, R. Vaia, A. Cuccoli, and V. Tognetti, Phys. Rev. B 58,273 Torgeson, Phys. Rev. Lett. 75,2212͑1995͒. ͑1998͒. 105See, e.g., L. J. de Jongh ͑ed.͒, Magnetic Properties of Layered Transition 122L. Capriotti, A. Cuccoli, V. Tognetti, P. Verrucchi, and R. Vaia, Phys. Rev. Metal Compounds, Kluwer, Dordrecht ͑1990͒. B 60, 7299 ͑1999͒. 106D. C. Johnston, in Handbook of Magnetic Materials, Vol. 10, K. H. J. 123D. P. Landau and K. Binder, Phys. Rev. B 24, 1391 ͑1981͒. Buschow ͑ed.͒, Elsevier Science, New York ͑1997͒. 124A. Cuccoli, T. Roscilde, R. Vaia, and P. Verrucchi, Phys. Rev. B 68, 107S. B. Khokhlachev, Zh. E´ ksp. Teor. Fiz. 70, 265 ͑1976͒. 060402 ͑2003͒. 108C. Kawabata and A. R. Bishop, Solid State Commun. 42,595͑1982͒. 125A. Cuccoli, T. Roscilde, R. Vaia, and P. Verrucchi, J. Magn. Magn. Mater. 109R. W. Gerling and D. P. Landau, in Magnetic Excitations and Fluctua- 272–276, 884 ͑2004͒. tions, S. W. Lovesey, U. Balucani, F. Borsa, and V. Tognetti ͑eds.͒, 126O. F. Syljua˚sen, Phys. Rev. B 61, 846 ͑2000͒. Spinger Verlag, Berlin ͑1984͒. 127P. Chandra, P. Coleman, and A. I. Larkin, Phys. Rev. Lett. 64,88͑1990͒. 110A. Cuccoli, V. Tognetti, and R. Vaia, Phys. Rev. B 52, 10221 ͑1995͒. 128C. Weber, L. Capriotti, G. Misguich, F. Becca, M. Elhajal, and F. Mila, 111 H. G. Evertz and D. P. Landau, Phys. Rev. B 54, 12302 ͑1996͒. Phys. Rev. Lett. 91, 177202 ͑2003͒. 112 R. Gupta and C. Baillie, Phys. Rev. B 45, 2883 ͑1992͒. 129R. Melzi, P. Carretta, A. Lascialfari, M. Mambrini, M. Troyer, P. Millet, 113 H. Q. Ding and M. S. Makivic´, Phys. Rev. B 42,6827͑1990͒; 45, 491 and F. Mila, Phys. Rev. Lett. 85, 1318 ͑2000͒; R. Melzi, S. Aldrovandi, ͑1992͒; H.-Q. Ding, Phys. Rev. Lett. 68, 1927 ͑1992͒; Phys. Rev. B 45, F. Tedoldi, P. Carretta, P. Millet, and F. Mila, Phys. Rev. B 64, 24409 230 ͑1992͒. ͑2001͒. 114 L. Capriotti, A. Cuccoli, V. Tognetti, R. Vaia, and P. Verrucchi, Physica D 130P. Carretta, R. Melzi, N. Papinutto, and P. Millet, Phys. Rev. Lett. 88, 119,68͑1998͒ 47601 ͑2002͒; P. Carretta, N. Papinutto, C. B. Azzoni, M. C. Mozzati, 115 A. Cuccoli, T. Roscilde, R. Vaia, and P. Verrucchi, Phys. Rev. Lett. 90, E. Pavarini, S. Gonthier, and P. Millet, Phys. Rev. B 66, 094420 ͑2002͒. ͑ ͒ 167205 2003 . 131V. N. Kotov, J. Oitmaa, O. P. Sushkov, and W. H. Zheng, Philos. Mag. B 116 A. Cuccoli, T. Roscilde, V. Tognetti, R. Vaia, and P. Verrucchi, J. Appl. 80, 1483 ͑2000͒. ͑ ͒ Phys. 93, 7640 2003 . 132L. Capriotti, Int. J. Mod. Phys. B 15, 1799 ͑2001͒. 117 A. Cuccoli, T. Roscilde, V. Tognetti, R. Vaia, and P. Verrucchi, J. Magn. 133L. Capriotti, A. Fubini, T. Roscilde, and V. Tognetti, Phys. Rev. Lett. 92, ͑ ͒ Magn. Mater. 272–276, e651 2004 . 157202 ͑2004͒. 118 A. Cuccoli, V. Tognetti, P. Verrucchi, and R. Vaia, Phys. Rev. B 51, 12840 134See the group homepage http://www.fi.infn.it/cmtg/. ͑1995͒. 119D. Vaknin, L. L. Miller, J. L. Zaretsky, and D. C. Johnston, Physica C This article was published in English in the original Russian journal. Repro- 274,331͑1997͒. duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 31, NUMBERS 8–9 AUGUST–SEPTEMBER 2005

Quantum information storage and state transfer based on spin systems Z. Song*

Department of Physics, Nankai University, Tianjin 300071, China C. P. Sun†

Department of Physics, Nankai University, Tianjin 300071, China; Institute of Theoretical Physics, Chinese Academy of Sciences, Beijing, 100080, China ͑Submitted December 23, 2004͒ Fiz. Nizk. Temp. 31, 907–917 ͑August–September 2005͒ The idea of quantum state storage is generalized to describe the coherent transfer of quantum information through a coherent data bus. In this universal framework, we comprehensively review our recent systematical investigations to explore the possibility of implementing the physical processes of quantum information storage and state transfer by using quantum spin systems, which may be an isotropic antiferromagnetic spin ladder system or a ferromagnetic Heisenberg spin chain. Our studies emphasize the physical mechanisms and the fundamental problems behind the various protocols for the storage and transfer of quantum information in solid state systems. © 2005 American Institute of Physics. ͓DOI: 10.1063/1.2008129͔

I. INTRODUCTION quantum memory. A protocol was presented to implement a quantum storage element for the electronic spin state in a The current development of quantum information sci- ring array of interacting nuclei. Under appropriate control of ence and technology demands optimal systems serving as both the electron and the external magnetic field, an arbitrary long-lived quantum memories, through which the quantum quantum state of the electronic spin qubit, either a pure or a information carried by a quantum system with short decoher- ence time can be coherently transferred.1 In this sense a mixed state, can be coherently stored in the nuclear spin quantum channel or a quantum data bus is needed for perfect wave and then read out in the reverse process. transmission of quantum states. In this article, we will dem- On the other hand, designed for a more realistic quantum onstrate that both the quantum information storage and the computing, a scalable architecture of quantum network 11,12 quantum state transfer can be uniquely described in a univer- should be based on the solid state system. However, the sal framework. intrinsic feature of solid state based channels, such as the There exist some schemes2–5 concerning the quantum finiteness of the correlation13,14 and the environment induced storage of photon states, and there are also some efforts de- noise ͑especially the low-frequency noise͒ may block this voted to the universal quantum storage for a qubit ͑a basic scalability. Fortunately, analytical study shows that a spin two-level system͒ state, which is necessary in quantum com- system possessing a commensurate structure of energy spec- putation. For example, most recently an interesting trum matched with the corresponding parity can ensure the protocol6–8 was presented to reversibly map the electronic perfect state transfer.15–17 Based on this fact, an isotropic spin state onto the collective spin state of the surrounding antiferromagnetic spin ladder system can be pre-engineered nuclei. Because of the long decoherence time of the nuclear as a novel robust kind of quantum data bus.18 Because the spins, the information stored in them can be robustly pre- effective coupling strength between the two spins connected 9 served. It was found that, only under two homogeneous con- to a spin ladder is inversely proportional to the distance of ditions with low excitations, such many-nuclei system ap- the two spins, the quantum information can be transferred proximately behaves as a single-mode boson to serve as an between the two spins separated by a longer distance. An- efficient quantum memory. other example of the near-perfect transfer of quantum infor- The low-excitation condition requires a ground state mation was given to illustrate an application of the theorem. with all spins oriented, which can be prepared by applying a The proposed protocol of such near-perfect quantum state magnetic field polarizing all spins along the same direction. With the concept of spontaneous symmetry breaking ͑SSB͒, transfer utilizes a ferromagnetic Heisenberg chain with uni- one can recognize that a ferromagnetic Heisenberg spin form coupling constant but in an external parabolic magnetic 17 chain usually has a spontaneous magnetization, which natu- field. rally offers a ground state of this kind. In event of SSB, the The present paper will give a broad overview of the intrinsic interaction between spins will strongly correlate present situation of the our investigations mentioned above with the nuclei to form the magnon, a collective mode of on quantum state storage and quantum information coherent spin wave, even without any external magnetic field. With transfer based on quantum spin systems. We will understand these considerations, Wang, Li, Song, and Sun10 explored the the physical mechanisms and the fundamental problems be- possibility of using a ferromagnetic quantum spin system, hind these protocols in the view of a unified conception, the instead of the free nuclear ensemble, to serve as a robust generalized quantum information storage.

1063-777X/2005/31(8–9)/9/$26.00686 © 2005 American Institute of Physics Low Temp. Phys. 31 (8–9), August–September 2005 Z. Song and C. P. Sun 687

II. GENERALIZED QUANTUM STORAGE AS A DYNAMIC PROCESS For the dynamic process recording and reading quantum information carried by quantum states, we first describe the idea of generalized quantum storage, which was also intro- duced in association with the Berry’s phase factor.19 Let M be a quantum memory possessing a subspace spanned by ͉ ϭ ͉ ϭ␦ M n͘ (n 1,2,...,d, ͗M n M m͘ nm), which can store the ͉ quantum information of a system S with basis vectors Sn͘, nϭ1,2,...,d. If there exists a controlled time evolution in- ͉  ͉ terpolating between the initial state Sn͘ M͘ and the final ͉  ͉ state S͘ M n͘ for each index n and arbitrarily given states ͉S͘ and ͉M͘, we define the usual quantum storage by using a FIG. 1. Demonstration of quantum state transfer as a process of generalized factorized evolution of time Tm quantum information storage by grouping the data bus D and the target subsystem SB as a generalized quantum memory. ͉⌽ ϭ ͉⌽ ϭ͉  ͉ ͑ ͒ ͑Tm͒͘ U͑Tm͒ ͑0͒͘ S͘ M n͘, 1 starting from the initial state ͉⌽(0)͘ϭ͉S ͘  ͉M͘. The cor- n ϭ͉  ͉ B ͉ D͘ UB Sn͘. Here D͘ is a robust state of the data bus and responding readout process is an inverse evolution of time T f (ϾT ) UB represents some local unitary transformations with re- m spect to B, which are independent of the initial state. With ͉⌽ ͒ ϭ ͉͒⌽ ͒ ϭ͉  ͉ ͑ ͒ ͑T f ͘ U͑T f ͑0 ͘ Sn͘ M͘. 2 this notation, the quantum state transfer indeed can be re- In this sense, writing an arbitrary state ͉S(0)͘ϭ͚ c ͉S ͘ of garded as a generalized QDP. n n n ͉ A͘ϭ͚ ͉ A͘ S into M with the initial state ͉M͘ of quantum memory can In fact, if one inputs a state of S ncn Sn localized ϭ be realized as a controlled evolution from time tϭ0to at A at t 0, the initial state of whole system can be written ϭ as t Tm

͉␺͑ ͒͘ϭ ͉ A͘  ͉ ͘ ͑ ͒ ͉ ͘  ͉ ͘→͉ ͘  ͉ ͘ ͑ ͒ 0 ͚ cn Sn M , 7 ͚ cn Sn M S ͚ cn M n . 3 n n n ͉ ͘ϭ͉ ͘  ͉ B͘ The readout process from M is another controlled evolution where M D UB S . The quantum state transfer can from time tϭT to tϭT : be usually described as a factorized time evolution at time m f ϭ t T f ͉  ͉ → ͉  ͉ ͑ ͒ S͘ ͚ cn M n͘ ͚ cn Sn͘ M͘. 4 ͉␺͑ ͒ ϭ͉  ͉  ͉ B ϭ͉  ͉ n n T f ͘ S͘ D͘ ͚ cnUB Sn ͘ S͘ ͚ cn M n͘ n n Obviously, the combination of these two processes forms a ͑8͒ cyclic evolution in which a state totally returns to the initial with ͉M ͘ϭ͉D͘  U ͉SB͘. The above equations just demon- one. n B n strate that the quantum state transfer is essentially a general- However, in the view of the decoding approach, one ized quantum memory with W ϭ(1 U ). In this sense the does not need ‘‘totally return’’ to revive the information of M B revisable quantum state transfer can be regarded as a general initial state, and a difference is allowed by the n-independent readout process. unitary transformation WϭW  1, namely, S Now we would like to remark on the differences be- tween generalized quantum state storage and other two types ͉  ͉ → ͉  ͉ ͑ ͒ S͘ WM ͚ cn M n͘ (WS͚ cn Sn͘) M͘. 5 n n of quantum processes, quantum teleportation and quantum copy. In fact, quantum teleportation is theoretically perfect, ͑ ͒ This is a quantum dynamic process QDP for recording and yielding an output state which revives the input with a fidel- reading, which defines a quantum storage. Because the factor ity Fϭ1. Actually one of necessary procedure in teleporta- WS is known to be independent of the initial state, it can be tion is to measure the Bell state at location A, which will ͚ ͉ ͘ easily decoded from WS ncn Sn by the inverse transforma- induce wave packet collapse. On the other way around, the tion of WS . We notice that the quantum storage usually re- quantum state storage process is always on time evolution lates to two quantum subsystems. without any measurement. As for quantum copy the initial We will show as follows that the quantum state transfer state remains unchanged during its copy and can be gener- can be understood as a generalized quantum storage with ated in a dynamic process. three subsystems, the input one with the Hilbert space SA, the data bus with D, and the output one with SB. As illus- III. QUANTUM STATE TRANSFER IN SPIN SYSTEMS trated in Fig. 1, the two subsystems SA and SB are located at two distant locations A and B, respectively. Then the Hilbert A robust quantum information processing based on solid space of the total system can be written as state system is usually implemented in a working space spanned by the lowest states, which are well separated from S ϭSA  D  SBϵSA  M, ͑6͒ T other dense spectra of high excitations. In this sense the en- where MϭD  SB can be regarded as the generalized quan- ergy gap of the solid state system is an important factor we ͉ tum memory with the memory space spanned by M n͘ should take into account. The decoherence induced by the 688 Low Temp. Phys. 31 (8–9), August–September 2005 Z. Song and C. P. Sun

environmental noise can also destroy the robustness of quan- identify the above SRS as the middle inversion of spins with tum information processing, such as the low-frequency ͑e.g., respect to the center of the quantum spin chain. As the dis- 1/f ) noise prevalent in solid state devices. cussion in Ref. 20, we write spin inversion operation People believe that the gap of the data bus can suppress P⌿͑s ,s ,...,s Ϫ ,s ͒ϭ⌿͑s ,s Ϫ ,...,s ,s ͒ ͑13͒ the stay of transferred state in the middle way in order to 1 2 N 1 N N N 1 2 1 ⌿ enhance the fidelity, but the large gap may result in a shorter for the wave function (s1 ,s2 ,...,sNϪ1 ,sN) of the spin ϭ correlation length. The relationship between correlation chain. Here, sn 0,1 denotes the spin values of the nth qubit. length and the energy gap is usually established in the system with translational symmetry. So we need to consider some modulated-coupling systems or artificially engineered irregu- B. Perfect state transfer in modulated coupling system lar quantum spin systems where the strong correlation be- Based on the above analysis, in principle, perfect quan- tween two distant site can be realized. tum state transfer is possible in the framework of quantum mechanics. According to SPMC, many spin systems can be A. Theorem for the perfect quantum state transfer pre-engineered for perfect quantum states transfer. For in- Quantum mechanics shows that perfect state transfer is stance, two-site spin-1/2 Heisenberg system is the simplest possible. To sketch our central idea, let us first consider a example which meets the SPMC. Recently, Christandl 15,16 single-particle system with the usual spatial reflection sym- et al. proposed an N-site XY chain with an elaborately metry ͑SRS͒ in the Hamiltonian H. Let P be the spatial designed modulated coupling constants between two nearest- reflection operator. The SRS is implied by ͓H,P͔ϭ0. Now neighbor sites, which ensures a perfect state transfer. It is ␲ ␺͑ ͒ we prove that at time /E0 any state r can evolve into the easy to find that this model corresponds the SPMC for the Ϯ␺ Ϫ ␧ ϭ reflected state ( r) if the eigenvalues n match the pari- simplest case Nn n. A natural extension of the application  ties pn in the following way: of the theorem leads to discover other models with Nn n. Following this idea, a new class of different models whose ␧ ϭ ϭϮ͑Ϫ ͒Nn ͑ ͒ n NnE0 ,pn 1 9 spectrum structures obey the SPMC exactly were proposed

for arbitrary positive integer Nn and for perfect state transfer. Consider an N-site spin-1/2 XY ␸ ϭ␧ ␸ ␸ ϭ ␸ ͑ ͒ chain with the Hamiltonian H n͑r͒ n n͑r͒, P n͑r͒ pn n͑r͒. 10 NϪ1 ␸ Here n(r) is the common eigen wave function of H and P, ϭ ͓ x x ϩ y y ͔ ͑ ͒ H 2 ͚ Ji Si Siϩ1 Si Siϩ1 , 14 r is the position of the particle. We call Eq. ͑9͒ the spectrum- iϭ1 parity matching condition ͑SPMC͒. The proof of the above x y z where S , S , and S are Pauli matrices for the ith site, and rigorous conclusion is a simple but heuristic exercise in basic i i i J is the coupling strength for the nearest-neighbor interac- quantum mechanics. In fact, for the spatial reflection opera- i tion. For the open boundary condition, this model is equiva- tor, P␺(r)ϭϮ␺(Ϫr). For an arbitrarily given state at t lent to the spinless fermion model. The equivalent Hamil- ϭ0, ␺(r,t)͉ ϭ , it evolves to t 0 tonian can be written as NϪ1 ␺͑r,t͒ϭexp͑ϪiHt͒␺͑r͒ϭ͚ C exp͑ϪiN E t͒␸ ͑r͒ n n 0 n ϭ ͓k͔ † ϩ ͑ ͒ n H ͚ Ji ai aiϩ1 h.c., 15 ͑11͒ iϭ1 ϭ ␸ ͉␺ ϭ␲ where a† , a are the fermion operators. This describes a at time t, where Cn ͗ n ͘. Then at time t /E0 ,we i i have simple hopping process in the lattice. According to the SPMC, we can present different models ͑labeled by different ␲ N k෈ ␺ͩ r, ͪ ϭ͚ C ͑Ϫ1͒ n␸ ͑r͒ϭϮP␺͑r͒ ͑12͒ positive integer 0,1,2,...) through pre-engineering of the n n ϭ [k]ϭͱ Ϫ E0 n coupling strength as Ji Ji i(N i) for even i and Ji ϭ [k]ϭͱ ϩ Ϫ ϩ ␺ r ␲ E ϭϮ␺ Ϫr 20 Ji (i 2k)(N i 2k) for odd i. By a straightfor- that is ( , / 0) ( ). This is just the central result ␧ ward calculation, one can find the k-dependent spectrum n discovered for quantum spin system that the evolution opera- ϭϪ ϩ Ϫ Ϫ ϭ ␧ ϭϪ ϩ tor becomes a parity operator ϮP at some instant tϭ(2n N 2(n k) 1 for n 1,2,...,N/2, and n N 2(n ϩ ␲ Ϫ ␲ ϭϮ ϩk)Ϫ1 for nϭN/2ϩ1,...,N. The corresponding 1) /E0 , that is exp( iH /E0) P. From the above ar- guments we have the consequence that if the eigenvalues k-dependent eigenstates are ␧ ϭ N N n NnE0 of a 1D Hamiltonian H with spatial reflection ͉␸ ϭ ͉ ϭ †͉ ͑ ͒ symmetry are odd-number spaced, i.e., N ϪN Ϫ are always n͘ ͚ cni i͘ ͚ cniai 0͘, 16 n n 1 ϭ ϭ odd, any initial state ␺(x) can evolve into Ϯ␺(Ϫx) at time i 1 i 1 ϭ␲ t /E0 . In fact, for such 1D systems, the discrete states where the coefficients cni can be explicitly determined by the alternate between even and odd parities. Consider the odd- recurrence relation presented in Ref. 18. ␧ ϭ number-spaced eigenvalues n NnE0 . The next-nearest It is obvious that the model proposed in Ref. 15 is just level must be even-number spaced; then the SPMC is satis- the special case of our general model with kϭ0. For arbi- fied. Obviously, the 1D SPMC is more realizable for the trary k, one can easily check that it meets the our SPMC by construction of the model Hamiltonian to perform perfect a straightforward calculation. Thus we can conclude that state transfer. these spin systems with a set of pre-engineered couplings [k] Now, we can directly generalize the above analysis to Ji can serve as the perfect quantum channels that allow the many particle systems. For the quantum spin chain, one can qubit information transfer. Low Temp. Phys. 31 (8–9), August–September 2005 Z. Song and C. P. Sun 689

FIG. 3. Schematic illustration of the energy levels of the system. When the connections between two qubits and the medium are switched off (J ϫ 0 FIG. 2. Two qubits A and B connect to a 2 N-site spin ladder. The ground ϭ0), the ground states are degenerate ͑a͒. When J switches on, the ground ͑ ͒ ͑ ͒ 0 state of H with a-type connection a is singlet triplet when N is even state͑s͒ and the first excited state͑s͒ are either singlet or triplet. This is ͑ ͒ ͑ ͒ odd , while for the b-type connection b , one should have the opposite approximately equivalent to that of two coupled spins ͑b͒, ͑c͒. result.

same rung, and ͗ij͘ʈ denotes nearest neighbors on either leg C. Near-perfect state transfer of the ladder. In the term Hq , L and R denote the sites In real many-body systems, the dimension of the Hilbert connecting to the qubits A and B at the ends of the ladder. space increases exponentially with the size N. For example, There are two types of connection between SA (SB) and the for an N-site spin-1/2 system the dimension is Dϭ2N, and ladder, which are illustrated in Fig. 2. According to Lieb’s the symmetry of the Hamiltonian cannot help so much. So it theorem,22 the spin of the ground state of H with the con- is almost impossible to obtain a model to be exactly engi- nection of type a is zero ͑one͒ when N is even ͑odd͒, while neered. In the above arguments we just show the possibility for the connection of type b, one should have the opposite to implement the perfect state transfer of any quantum state result. For the two-leg spin ladder HM , analytical analysis over arbitrary long distances in a quantum spin chain. It and numerical results have shown that the ground state and sheds light into the investigation of near-perfect quantum the first excited state of the spin ladder have spin 0 and 1, state transfer. There is a naive way that one selects some respectively.13,14 It is also shown that there exists a finite spin ⌬ϭ MϪ Mϳ special states to be transported, which is a coherent superpo- gap E1 Eg J/2 between the ground state and the first sition of commensurate part of the whole set of eigenstates. excited state ͑see Fig. 3͒. This fact has been verified by For example, we consider a truncated Gaussian wave experiments13 and is very crucial for our present investiga- packet for an anharmonic oscillator with lower eigenstates to tion. be harmonic. It is obvious that such system allows some Thus, it can be concluded that the medium can be ro- special states to transfer with high fidelity. We can imple- bustly frozen to its ground state to induce the effective ϭ ment such approximate harmonic system in a natural spin Hamiltonian Heff JeffSA•SB between the two end qubits. chain without the pre-engineering of couplings but in the With the effective coupling constant Jeff to be calculated in presence of a modulated external field. Another way to real- the following, this Hamiltonian depicts the direct exchange ize near perfect state transfer is to achieve the entangled coupling between two separated qubits. As the famous Bell ͉ states and fast quantum states transfer of two spin qubits by states, Heff has singlet and triplet eigenstates j,m͘AB : ͉ ϭ & ͉↑ ͉↓ Ϫ͉↓ ͉↑ ͉ ϭ͉↑ ͉↑ ͉ connecting two spins to a medium which possesses a spin 0,0͘ 1/ ( ͘A ͘B ͘A ͘B) and 1,1͘ ͘A ͘B , 1, Ϫ ϭ͉↓ ͉↓ ͉ ϭ & ͉↑ ͉↓ ϩ͉↓ ͉↑ gap. A perturbation method, the Fro¨hlich transformation, 1͘ ͘A ͘B , 1,0͘ 1/ ( ͘A ͘B ͘A ͘B), which shows that the interaction between the two spins can be can be used as a channel to share entanglement for a perfect mapped to the Heisenberg type coupling. quantum communication in a longer distance. Spin ladder. We sketch our idea with the model illus- The above central conclusion can be proved with both trated in Fig. 2. The whole quantum system we consider here analytical and numerical methods as follows. To deduce the ϫ ͉␺ ͉␺ consists of two qubits (A and B)anda2 N-site two-leg above effective Hamiltonian we use g͘M ( ␣͘M) and Eg ͑ ͒ spin ladder. In practice, this system can be realized by an (E␣) to denote ground excited states of HM and the corre- engineered array of quantum dots.21 The total Hamiltonian sponding eigenvalues. The zero-order eigenstates ͉m͘ can ϭ ϩ H Hm Hq contains two parts, the medium Hamiltonian then be written in a joint way as ͉ ͘ ϭ͉ ͘  ͉␺ ͘ ϭ ϩ ͑ ͒ j,m g j,m AB g M , HM J ͚ Si•Sj J ͚ Si•Sj 17 ͗ij͘Ќ ͗ij͘ʈ ͉␺ jm z ϭ͉  ͉␺ ͑ ͒ ␣ ͑s ͒͘ j,m͘AB ␣͘M . 19 describing the spin-1/2 Heisenberg spin ladder consisting of zϭ z two coupled chains, and the coupling Hamiltonian Here we have considered that the z component S SM ϩSz ϩSz ϭ ϩ ͑ ͒ A B of the total spin is conserved with respect to the Hq J0S⌳•SL J0SB•SR 18 z 2 connection Hamiltonian Hq . Since SM and SM commute ͉␺ ͉␺ z describing the connections between qubits A, B and the lad- with HM , we can label g͘M as g(sM ,sM)͘M , and then zϭ ϩ z der. In the term HM , i denotes a lattice site on which one s m sM can characterize the non-coupling spin state jm z electron sits, ͗ij͘Ќ denotes nearest-neighbor sites on the ͉␺␣ (s )͘. 690 Low Temp. Phys. 31 (8–9), August–September 2005 Z. Song and C. P. Sun

When the connections between the two qubits and the ϭ medium are switched off, i.e., J0 0, the degenerate ground ͉ states of H are just j,m͘g with the degenerate energy Eg and spin 0, 1, respectively, which is illustrated in Fig. 3a. When the connections between the two qubits and the medium are switched on, the degenerate states with spin 0,1 ͑Ref. 23͒ should split as illustrated in Fig. 3b, c. In the case with J0 ӶJ at lower temperature kTϽJ/2, the medium can be frozen to its ground state, and then we have the effective Hamil- tonian

Ј Ј ͉g͗ j,m͉H ͉␺ j m ͑sz͉͒͘2 ϳ q ␣ ͉ ͘ ͗ ͉ Heff ͚ Ϫ j,m gg j,m jЈ,mЈ, j,m,sz Eg E␣ 1 1 1 3 ϭJ .Diag.ͩ , , ,Ϫ ͪ ϩ␧͑20͒ eff 4 4 4 4 FIG. 4. The spin gaps obtained by numerical method for the systems L ϭ ϭ ϭ 4,5,6,7,8,10 with J 10,20,40 and J0 1 are plotted, corresponding to the where ϳ magnitude of Jeff . It indicates that Jeff 1/(LJ). J2͓L͑␣͒R*͑␣͒ϩR͑␣͒L*͑␣͔͒ ϭ 0 Jeff ͚ Ϫ , ␣ Eg E␣ ͉␺ ϭ ͉  ͉␤ ͑ ͒ ͘ ͚ c jm j,m͘AB jm͘M 22 3J2͓͉L͑␣͉͒2ϩ͉R͑␣͉͒2͔ jm ␧ϭ͚ 0 ͑ ͒ ͑ Ϫ ͒ . 21 ͉␤ ␣ 4 Eg E␣ where ͕ jm͘M͖ is a set of vectors of the data bus, which is not necessarily orthogonal. Then we have the condition ͚ ͉ ͉2 ␤ ͉␤ ϭ ͉␺͘ This just proves the above effective Heisenberg Hamiltonian jm c jm M͗ jm jm͘M 1 for normalization of . In this ͑5͒. Here the matrix elements of interaction K(␣) sense the practical description of the A – B subsystem of two ϭ ␺ ͉ z ͉␺ ϭ M͗ g SK ␣(1,0)͘M (K S,L) can be calculated only for quits can only be given by the reduced density matrix the variables of the data bus medium. We also remark that, because Sz and S2 are conserved for H , the off-diagonal ␳ ϭ ͉␺ ␺͉ ϭ ͉ ͉2͉ ͉ q AB TrM͑ ͗͘ ͒ ͚ c jm j,m͘AB͗ j,m elements in the above effective Hamiltonian vanish. jm To sum up so far, we have shown that at lower tempera- Ͻ ϩ * ͗␤ ͉␤ ͘ ͉ ͘ ͗ Ј Ј͉ ture kT J/2, H can be mapped to the effective Hamiltonian ͚ c jЈmЈc jmM jЈmЈ j,m M j,m AB j ,m jЈmЈ jm Heff , which seemingly depicts the direct exchange coupling between two separated qubits. Notice that the coupling ͑23͒ strength has the form J ϳg(L)J2/J, where g(L) is a function eff 0 where Tr means the trace over the variables of the medium. of LϭNϩ1, the distance between the two qubits concerned. M By a straightforward calculation we have Here we take the Nϭ2 case as an example. According to Eq. ͑21͒ one can get J ϭϪ(1/4)J2/J and (1/3)J2/J when A and 1 eff 0 0 ͉ ͉2ϭ͉ ͉2ϭͳ ␺ͯͩ ϩ z z ͪͯ␺ʹ c c Ϫ S S , B connect the plaquette diagonally and adjacently, respec- 11 1 1 4 A• B tively. This result is in agreement with the theorem22 about the ground state and the numerical result when J ӶJ.In 1 0 ͉c ͉2ϭͳ ␺ͯͩ ϪS S ͪͯ␺ʹ , ͑24͒ general cases, the behavior of g(L) versus L is very crucial 00 4 A• B for quantum information since L/͉J ͉ determines the char- eff ͉c ͉2ϭ1Ϫ2͉c ͉2Ϫ͉c ͉2. acteristic time of quantum state transfer between the two 10 11 00 qubits A and B. In order to investigate the profile of g(L), a Now we need a criterion to judge how close the practical numerical calculation is performed for the systems L reduced eigenstate is to the pure state for the effective two- ϭ ϭ ϭ ͑ ͒ 4,5,6,7,8,10, with J 10,20,40 and J0 1. The spin gap s site coupling Heff . As we noticed, it has the singlet and trip- ͉ ͉ between the ground state͑s͒ and first excited state͑s͒ is ͑are͒ let eigenstates j,m͘AB in the subspace spanned by 0,0͘AB zϭ z ϩ z ϭ ͉ ͉2ϭ͉ ͉2ϭ͉ ͉2ϭ calculated, corresponding to the magnitude of Jeff . The nu- with S SA SB 0, we have c11 c10 c1Ϫ1 0, ͉ ͉2ϭ ͉ ͉ ͉2 merical result is plotted in Fig. 4, which indicates that Jeff c00 1; for the triplet eigenstate 1,0͘AB we have c11 ϳ ϭ͉ ͉2ϭ͉ ͉2ϭ ͉ ͉2ϭ 1/(LJ). It implies that the characteristic time of quantum c1Ϫ1 c00 0, c10 1. With the practical Hamil- ͉ ͉2 ϭ state transfer linearly depends on the distance and then guar- tonian H, the values of c jm , i 1,2,3,4 are calculated nu- ͉␺ antees the possibility of realizing the entanglement of two merically for the ground state g͘ and first excited state ͉␺ ͑ ͒ ϭ ϭ separated qubits in practice. 1͘ of finite system s L 4,5,6,7,8,10 with J 10,20,40 ϭ zϭ In order to verify the validity of the effective Hamil- (J0 1) in the S 0 subspace, which are listed in the Table tonian Heff , we need to compare the eigenstates of Heff with 1a, b, c of Ref. 17. It shows that, at lower temperature, the ͉ ͉2 those reduced states from the eigenstates of the whole sys- realistic interaction leads to results for c jm which are very tem. In general the eigenstates of H can be written formally close to that described by Heff , even if J is not so large in as comparison with J0 . Low Temp. Phys. 31 (8–9), August–September 2005 Z. Song and C. P. Sun 691

We assert that the above tables reflect all the facts dis- tinguishing the difference between the results about the en- tanglement of two end qubits generated by Heff and H. Though we have ignored the off-diagonal terms in the re- duced density matrix, the calculation of the fidelity ͉ ϵ ͉␳ ͉ ϭ͉ ͉2 F( j,m͘) M͗ j,m AB j,m͘M c jm further confirms our observation that the effective Heisenberg type interaction of two end qubits can approximate the realistic Hamiltonian very well. Then the quantum information can be transferred between the two ends of the 2ϫN-site two-leg spin ladder, that can be regarded as the channel to share entanglement with separated Alice and Bob. Physically, this is just due to a large spin gap existing in such a perfect medium, whose ground state can induce a maximal entanglement of the two end qubits. We also pointed out that our analysis is appli- cable for other types of medium systems as data buses, ͉ ͉ which possess a finite spin gap. Since L/ Jeff determines the characteristic time of quantum state transfer between the two FIG. 5. Schematic illustration of the time evolution of a Gaussian wave qubits, the dependence of J upon L becomes important and packet. It shows that the near-perfect state transfer over a long distance is eff possible in the quasi-harmonic system. relies on the appropriate choice of the medium. In conclusion, we have presented and studied in detail a protocol for quantum state transfer. Numerical results show ϭ ϭ We consider a Gaussian wave packet at t 0, x NA as that the isotropic antiferromagnetic spin ladder system is a the initial state perfect medium through which the interaction between two 2Nϩ1 separated spins is very close to the Heisenberg type of cou- 1 ͉␺͑ ͒͘ϭ ͩ Ϫ ␣2͑ Ϫ Ϫ ͒2 ͪ ͉ ͘ ͑ ͒ NA,0 C ͚ exp i NA 1 i 28 pling, with a coupling constant inversely proportional to the iϭ1 2 distance, even if the spin gap is not so large compared to the ͉ ͘ couplings between the input and output spins and the me- where i denotes the state with 2N spins in the down state dium. and only the ith spin in the up state, and C is the normaliza- ␣2ϭ ⌬2 Spin chain in modulated external magnetic field. Let us tion factor. The coefficient 4ln2/ is determined by ⌬ ͉␺ ͘ consider the Hamiltonian of a (2Nϩ1)-site spin-1/2 ferro- the width of the Gaussian wave packet. The state (0) ͉␺ ͘ϭ Ϫ ͉␺ ͘ magnetic Heisenberg chain evolves to (t) exp( iHt) (NA,0) at time t, and the ͉␺ fidelity for the state (0)͘ transferring to the position NB is 2N 2Nϩ1 defined as ϭϪ ϩ ͑ ͒ z ͑ ͒ H J͚ Si•Siϩ1 ͚ B i Si 25 ϭ͉ ␺ ͉ Ϫ ͉␺ ͉ ͑ ͒ iϭ1 iϭ1 F͑t͒ ͗ ͑NB,0͒ exp͑ iHt͒ ͑NA,0͒͘ . 29 ͉␺ ͘ with the uniform coupling strength ϪJϽ0, but in the para- In Fig. 5 the evolution of the state (0) is illustrated sche- matically. From the investigation of Ref. 25, we know that bolic magnetic field ϭϪ ϭϪ for small NA NB x0 , where NB is the mirror counter- ͒ϭ Ϫ Ϫ ͒2 ͑ ͒ part of N , but in the large ⌬ limit, if we take B B͑i 2B0͑i N 1 26 A 0 ϭ8(ln2/⌬2)2, F(t) has the form where B is a constant. In single-excitation invariant sub- 0 1 2t space with the fixed z component of the total spin, SzϭN ͑ ͒ϭ ͫϪ ␣2 2 ͩ ϩ ͪͬ ͑ ͒ F t exp NA 1 cos ␣2 30 Ϫ1/2, this model is equivalent to the spinless fermion hop- 2 ping model with the Hamiltonian which is a periodic function of t with the period Tϭ␣2␲ and ϩ has maximum of 1. This is in agreement with our above J 2N 1 2N 1 ϭϪ ͑ † ϩ ͒ϩ ͑ ͒ † ͑ ͒ analysis. However, in quantum communication, what we are H ͚ ai aiϩ1 h.c. ͚ B i ai ai 27 2 iϭ1 2 iϭ1 concerned with is the behavior of F(t) in the case of the ӷ⌬ ϭ ͉ ͉ϭ ͉ ͉ transfer distance L , where L 2 NA 2 NB . For this where for simplicity we have neglected a constant in the purpose the numerical method is performed for the case L ϭ ⌬ϭ ϭ ⌬2 2␭ ␭ Hamiltonian. For the single-particle case with the basis set 500, 2,4,6 and B0 8(ln2 ) . The factor deter- ͕͉n͘ϭ͉0,0,...,1nЈth,0,...͘,nϭ1,2,...͖, which is just the same mines the maximum fidelity, and then the optimal field dis- as that of the Hamiltonian of a Josephson junction in the tribution can be obtained numerically. In Fig. 2a, b, c of Ref. 24 ϭ ϭ ␭ Cooper-pair number basis for EJ J, Ec 2B0 , analytical 18 the functions F(t) are plotted for different values of .It analysis and numerical results have shown that the lower shows that for the given wave packets with ⌬ϭ2, 4, and 6 ␭ energy spectrum is indeed quasi-harmonic in the case EJ there exists a range of within which the fidelities F(t) are ӷ ͑ ͒ Ec Ref. 24 . Although the eigenstates of the Hamiltonian up to 0.748, 0.958, and 0.992, respectively. For finite dis- ͑25͒ do not satisfy the SPMC precisely, especially in the tance, the maximum fidelity decreases as the width of Gauss- high-energy range, there must exist some Gaussian wave ian wave packet increases. On the other hand, the strength of packet states expanded by the lower eigenstates. This kind of the external field also determines the value of the optimal state can be transferred with high fidelity. fidelity for a given wave packet. There exists an optimal 692 Low Temp. Phys. 31 (8–9), August–September 2005 Z. Song and C. P. Sun

FIG. 6. The electronic spin state onto the collective spin state of the sur- rounding nuclei ensemble in a quantum well.

external field to obtain maximal fidelity, while the period of 2 F(t) is close to Tϭ␣ ␲. This shows a difference from the FIG. 7. The configuration geometry of the nuclei-electron system. The nu- ideal system, i.e., continuous harmonic systems, in which the clei are arranged in a circle within a quantum to form a ring array. To turn on fidelity is independent of the strength of the external field. the interaction one can push a single electron towards the center of the circle along the axis perpendicular to the plane. Numerical results indicate that it is possible to realize near- perfect quantum state transfer over a longer distance in a practical ferromagnetic spin chain system. In summary, we have shown that a perfect quantum and its conjugate Bϩ are introduced to depict the collective transmission can be realized through a universal quantum excitations in ensemble of nuclei with spin I0 from its polar- channel provided by a quantum spin system with spectrum ized initial state structure, in which each eigenenergy is commensurate and matches with the corresponding parity. According to this N 20 ͉ ϭ͉Ϫ ϭ ͉Ϫ SPMC for the a mirror inversion symmetry, we can imple- G͘ NI0͘ ͟ I0͘i ϭ ment the perfect quantum information transmission with sev- i 1 eral novel pre-engineered quantum spin chains. For more practical purpose, we prove that an approximately commen- which denotes the saturated ferromagnetic state of ensemble surate spin system can also realize near-perfect quantum of nuclei. There is an intuitive argument that if the gi have different values, while the distribution is ‘‘quasi- state transfer in a ferromagnetic Heisenberg chain with uni- † form coupling constant in an external field. The fidelity for homogeneous,’’ B and B can also be considered as boson ͓ ϩ͔→ the system in a parabolic magnetic field has been studied by operators satisfying B,B 1 approximately. Song, Zhang, and Sun analyzed the universal applicabil- a numerical method. The external field plays a crucial role in 9 the scheme. It induces a lower quasi-harmonic spectrum, ity of this protocol in practice. It was found that only under which can drive a Gaussian wave packet from the initial two homogeneous conditions with low excitations does the position to its mirror counterpart. The fidelity depends on the many-nuclei system behave approximately as a single-mode initial position ͑or distance L), the width ⌬ of the wave boson and can its excitation serve as an efficient quantum packet, and the magnetic field distribution B(i) via the factor memory. The low-excitation condition requires a ground ␭. Thus for given L and ⌬, proper selection of the factor ␭ state with all spins oriented, which can be prepared by ap- can achieve the optimal fidelity. Finally, we conclude that it plying a magnetic field polarizing all spins along a single is possible to implement near-perfect Gaussian wave packet direction. With consideration of the spontaneous symmetry transmission over a longer distance in many-body system. breaking for all spins oriented, a protocol for a quantum storage element was proposed utilizing a ferromagnetic quantum spin system, instead of the free nuclear ensemble, to serve as a robust quantum memory. IV. QUANTUM STORAGE BASED ON THE SPIN CHAIN The configuration of the quantum storage element is il- Recently a universal quantum storage protocol6–8 was lustrated in Fig. 7. The nuclei are arranged in a circle within presented to reversibly map the electronic spin state onto the a quantum to form a spin ring array. A single electron is just collective spin state of the surrounding ensemble of nuclei in localized at the center of the ring array, surrounded by the a quantum well ͑see Fig. 6͒. Because of the long decoherence nuclei. The interaction of the nuclear spins is assumed to time of the nuclear spins, the information stored in them can exist only between the nearest neighbors while the external be robustly preserved. magnetic field B0 threads through the spin array. Then the ͑ (i) (i) (i) electron-nuclei system can be modeled by a Hamiltonian H When all nuclei with spin operators Ix , Iy , Iz )of ϭH ϩH ϩH . It contains the electronic spin Hamiltonian spin I0 are coupled with a single electron spin with strength e n en 9 H ϭg ␮ B ␴z, the nuclear spin Hamiltonian gi , a pair of collective operators e e B 0 ͑ ͒ ͚N i N N iϭ1giIϪ Bϭ ͑31͒ ϭ ␮ zϪ ͑ ͒ ͱ ͚ 2 Hn gn nB0 ͚ Sl J͚ Sl•Slϩ1 32 2I0 g j lϭ1 lϭ1 Low Temp. Phys. 31 (8–9), August–September 2005 Z. Song and C. P. Sun 693

␳ ϭ␳   ͉Ϫ Ϫ͉ ͑ ͒ with the Zeeman splitting and the ferromagnetic interaction ͑T͒ b͑0͒ wF ͗͘ , 38 JϾ0, and the interaction between the nuclear spins and the ϭ͚ ͉ ͗͘ ͉ electronic spin where wF n,mϭ0,1wnm nN mN is the storing state of the Nth magnon with ␭ N ϭ ␴ϩ Ϫϩ ͑ ͒ Hen ͚ Sl h.c. 33 i N ϭ ϭ␳ ͫ ͑ Ϫ ͒␲ͬ ͑ ͒ 2 l 1 w exp m n . 39 nm nm 2 ͑ ͒ Here ge (gn) is the Lande g factor of the electron nuclei , ␳ ␮ ␮ ͑ ͒ Here, to simplify our expression, we have denoted ϩϩ and B ( N) is the Bohr magneton nuclear magneton . The Ϫ ϩ ϵ␳ , ␳Ϯϵ␳ , ␳Ϫϩϵ␳ , ␳Ϫϵ␳ . The difference be- Pauli matrices S and ␴ represent the nuclear spin of the 00 01 10 11 l tween w and ␳ (0) is only an unitary transformation inde- lth site and the electronic spin, respectively. The denomina- F e pendent of the stored initial state ␳ (0). tor N in Eq. ͑33͒ originates from the envelope normalization e So far we have discussed the ideal case with homoge- of the localized electron wave function.6–8 The hyperfine in- neous coupling between the electron and the nuclei, that is, teractions between nuclei and electron are proportional to the the coupling coefficients are the same constant ␭ for all the envelope function of localized electron. The electronic wave nuclear spins. However, the inhomogeneous effect of cou- function is supposed to be cylindrically symmetric, e.g., the pling coefficients has to be taken into account if what our s-wave component. Thus the coupling coefficient ␭ concern extends beyond the s-wave component, where the ϰ͉␺(r)͉2 is homogeneous for all the N nuclei in the ring wave function is not strictly cylindrically symmetric. In this array. case, the quantum decoherence induced by the so-called To consider the low spin-wave excitations, the discrete quantum leakage has been extensively investigated for the Fourier transformation defines the bosonic operators atomic ensemble based quantum memory.26 We now discuss 1 N 2␲kl the similar problems for the magnon-based quantum ϭ ͩ ͪ Ϫl ͑ ͒ bk ͚ exp i S , 34 memory. ͱ ϭ N N l 1 ␭ ϰ͉␺ ͉2 ␺ In the general case the l (rl) , where (rl) is the in the large N limit. Then one can approximately diagonalize envelope function of the electron at site rl , vary with the the Hamiltonian ͑32͒ as positions of the nuclear spins. In this case, the Hamiltonian NϪ1 contains terms other than the interaction between the spin ϭ ϩ ␻ † and Nth mode boson, that is, the inhomogeneity induced HT HN ͚ kbk bk kϭ1 interaction ͑ ͒ NϪ1 where HN is a Jaynes-Cummings JC type Hamiltonian s ϭ␭ͱ ͩ ␴ ␹ ϩ ͪ ͑ ͒ V ϩ ͚ kbk h.c. 40 ⍀ s 2N kϭ1 H ϭ␻ b† b ϩ ␴zϩ␭ͱ ͑␴ϩb ϩ␴Ϫb† ͒. ͑35͒ N N N N 2 2N N N should be added in our model Hamiltonian HT , where Then we obtain the dispersion relation for the magnon or N ␭ spin-wave excitation ␹ ϭ l ␲ k ͚ exp͑i2 kl/N͒. lϭ1 ␭N 2␲k ␻ ϭ ␮ ϩ Ϫ ͑ ͒ k gn nB0 2Js 2Js cos . 36 ␭ ␭ ϭ ␭ ͱ ␲␴ N For a Gaussian distribution of l , e.g., l ( / 2 )exp ϫ Ϫ Ϫ 2 ␴2 ␴ ␭ ϭ␭ ( (l 1) /(2 )) with width and 1 , the correspond- The above results show that HT contains only the interaction ing inhomogeneous coupling is described by of the Nth magnon with the electronic spin, and the other Ϫ NϪ1 magnons decouple with it. Here the frequency of the 1 N 1 1 Ϫ͑lϪ1͒2 2␲kl ␹ ϭ ͩ ϩ ͪ ͑ ͒ ␻ ϭ ␮ ⍀ k ͚ exp 2 i . 41 boson N gn nB0 and the two-level spacing N Ϫ ϭ ͱ ␲␴ 2␴ N ϭ ␮ l 1 0 2 2g* BB0 can be modulated by the external field B0 si- ␹ multaneously. Figure 8 shows the magnitude of k for different Gauss- ␭ ␴ The process of quantum information storage can be ian distributions of l with different widths . It indicates implemented in the invariant subspace of the electronic spin that the modes near 1 and NϪ1 have a stronger coupling and the Nth magnon. Now we can describe the quantum with the electron. When the interaction gets more homog- ͑ ␴͒ ␹ storage protocol based on the above spin-boson model. Sup- enous with larger the coupling coefficients k for all the pose the initial state of the total system is prepared so that modes from 1 to NϪ1 become smaller. When the distribu- there is no excitation in the N nuclei at all, while the electron tion is completely homogeneous, all the couplings with the ␳ ϭ͚ ␳ ͉ ͉ ͉ϩ͘ Ϫ is in an arbitrary state e(0) n,mϭϮ nm n͗͘m , where N 1 magnon modes vanish, and then we obtain the Hamil- ͉͑Ϫ͒͘ denotes the electronic spin up ͑down͒ state. The initial tonian HT . state of the total system can then be written as In the following we will adopt a rather direct method to analyze the decoherence problem of our protocol resulting ␳͑0͒ϭ␳ ͑0͒  ͉0 ͗͘0 ͉  ␳ ͑0͒ ͑37͒ b N N e from dissipation. If N is so large that the spectrum of the ␳ ϭ͉ ͉ ͉ in terms of b(0) ͕0͖͘NϪ1͕͗0͖ , where n1 ,n2 ,...,nNϪ1͘ quantum memory is quasicontinuous, this model is similar to ϵ͉ ϭ Ϫ ͕nk͖͘NϪ1 (k 1,2,...,N 1) denotes the Fock state of the the ‘‘standard model’’ of quantum dissipation for the vacuum Ϫ ϭ ϭ 27 other N 1 magnons. If we set B0 0, at t T induced spontaneous emission. The NϪ1 magnons will in- ϵ(␲/␭)ͱN/2s, the time evolution from ␳͑0͒ is just de- duce the quantum dissipation of the electronic spin with a scribed as a factorized state decay rate 694 Low Temp. Phys. 31 (8–9), August–September 2005 Z. Song and C. P. Sun

or dephasing can result from the mixing of different magnon modes. We acknowledge the collaborations with P. Zhang, Yong Li, Y. D. Wang, B. Chen, X. F. Qian, T. Shi, Ying Li and R.Xin, which have resulted in our systematic research on quantum spin-based quantum information processing. We acknowledge the support of the CNSF ͑Grants Nos. 90203018, 10474104, 10447133͒, the Knowledge Innovation Program ͑KIP͒ of the Chinese Academy of Sciences, and the National Fundamental Research Program of China ͑No. 2001GB309310͒.

*E-mail: [email protected] †E-mail: [email protected] FIG. 8. The fidelity F(t) in the large N limit. The vertical line indicates the instant ␲/2g at which the quantum storage is just implemented. Here ␥/g ϭ1/50. The inset shows the decaying oscillation with details of F(t)ina small region near the instant ␲/2g. 1 D. P. DiVincenzo and C. Bennet, Nature ͑London͒ 404,247͑2000͒ and references therein. 2 M. D. Lukin, Rev. Mod. Phys. 75, 457 ͑2003͒. 3 M. Fleischhauer and M. D. Lukin, Phys. Rev. Lett. 84, 5094 ͑2000͒; Phys. ͑ ͒ N Rev. A 65, 022314 2002 . ␭2s͉␹ ͉2 s 4 ͑ ͒ ␥ϭ ␲ k ␦ ␻ Ϫ ␭ͱ ͑ ͒ C. P. Sun, Y. Li, and X. F. Liu, Phys. Rev. Lett. 91, 147903 2003 . ͩ ͪ 5 2 ͚ k 2 . 42 E. Pazy, I. DAmico, P. Zanardi, and F. Rossi, Phys. Rev. B 64, 195320 kϭ1 2N 2N ͑2001͒. 6 Let ͉⌿͘ be the ideal evolution governed by the expected J. M. Taylor, C. M. Marcus, and M. D. Lukin, Phys. Rev. Lett. 90, 206803 ͉⌿Ј͘ ͑2003͒. Hamiltonian HT without dissipation and be the realistic 7 A. Imamoglu, E. Knill, L. Tian, and P. Zoller, Phys. Rev. Lett. 91, 017402 evolution governed by the Hamiltonian with dissipation. ͑2003͒. Supposeing that the initial state of the electron is (͉ϩ͘ϩ͉ 8 M. Poggio et al., Phys. Rev. Lett. 91, 207602 ͑2003͒. 9 ͑ ͒ Ϫ͘)/&, we can analytically calculate the fidelity Z. Song, P. Zhang, T. Shi, and C. P. Sun, Phys. Rev. B 71, 205314 2005 . 10 Y. D. Wang, Y. Li, Z. Song, and C. P. Sun, cond-mat/0409120, submitted F͑t͒ϭ͉͗⌿͉⌿Ј͉͘ to Phys. Rev. A ͑2004͒. 11 D. P. DiVincenzo, D. Bacon, J. Kempe, G. Burkard, and K. B. Whaley, 1 ␥ Nature ͑London͒ 408,339͑2000͒. ϭ ͫ ϩ ͩ Ϫ ͪͬ ␾͑ ͑⌬Ј ϩ␾͒ 12 S. Bose, Phys. Rev. Lett. 91, 207901 ͑2003͒. 1 exp t sec cos gt cos 1t 2 2 13 E. Dagotto and T. M. Rice, Science 271, 618 ͑1996͒. 14 S. White, R. Noack, and D. Scalapino, Phys. Rev. Lett. 73,886͑1994͒;R. ϩ ⌬Ј ͒ ͑ ͒ sin gt sin 1t , 43 Noack, S. White, and D. Scalapino, Phys. Rev. Lett. 73,882͑1994͒. 15 M. Christandl, N. Datta, and J. Landahl, Phys. Rev. Lett. 92, 187902 where ͑2004͒. 16 ␾ϭarcsin ͱ2N2 /␭2s, gϭ␭ͱs/2N and ⌬Јϭͱg2Ϫ␥2. M. Christandl, N. Datta, T. C. Dorlas, A. Ekert, A. Kay, and A. J. Landahl, ␥ 1 Phys. Rev. A 71, 032312 ͑2005͒; quant-ph/0411020. Figure 8 shows the curve of the fidelity F(t) changing 17 T. Shi, Ying Li, Z. Song, and C. P. Sun, Phys. Rev. A 71, 032309 ͑2005͒. 18 with time t. We can see that the fidelity exhibits a exponen- Ying Li, T. Shi, B. Chen, Z. Song, and C. P. Sun, Phys. Rev. A 71, 022301 ͑2005͒. tial decay behavior with a sinusoidal oscillation. At the in- 19 C. P. Sun, P. Zhang, and Y. Li, quant-ph/0311052; Y. Li, P. Zhang, stance when we have just implemented the quantum storage P. Zanardi, and C. P. Sun, Phys. Rev. A 70, 032330 ͑2004͒. process, the fidelity is about 1Ϫ␲␥/8. Therefore, the devia- 20 C. Albanese, M. Christandl, N. Datta, and A. Ekert, Phys. Rev. Lett. 93, ͑ ͒ tion from the ideal case with homogeneous couplings is very 230502 2004 ; quant-ph/0405029. 21 D. Loss and D. P. DiVincenzo, Phys. Rev. A 57,120͑1998͒;B.E.Kane, small for ␥/gӶ1. Since the ring-shape spin array with inho- Nature ͑London͒ 393, 133 ͑1998͒. mogeneous coupling is just equivalent to an arbitrary 22 E. Lieb, Phys. Rev. Lett. 62, 1201 ͑1989͒; E. Lieb and D. Mattis, J. Math. Heisenberg spin chain in the large N limit, the above argu- Phys. 3, 749 ͑1962͒. 23 Z. Song, Phys. Lett. A 231, 135 ͑1997͒; Z. Song, Phys. Lett. A 233,135 ments means that an arbitrary Heisenberg chain can be used ͑1997͒. for quantum storage following the same strategy addressed 24 Yu. Makhlin, G. Schon, and A. Shnirman, Rev. Mod. Phys. 73, 357 above if ␥/g is small, i.e., the inhomogeneous effect is not ͑2001͒. 25 very strong. T. Shi, B. Chen, Z. Song, and C. P. Sun, Commun. Theor. Phys. 43,795 ͑2005͒. On the other hand, if N is small, the spectrum of the 26 C. P. Sun, S. Yi, and L. You, Phys. Rev. A 67, 063815 ͑2003͒. quantum memory is discrete enough to guarantee the adia- 27 W. H. Louisell, Quantum Statistical Properties of Radiation, Wiley, New batic elimination of the NϪ1 magnon modes, i.e., York ͑1990͒. ␭ͱ ␹ ͉␻ ͉Ӷ Ϫ s/2N k / k 1 for the N 1 magnon modes. As a con- This article was published in English in the original Russian journal. Repro- sequence of this adiabatic elimination, quantum decoherence duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 31, NUMBERS 8–9 AUGUST–SEPTEMBER 2005

Localized-magnon states in strongly frustrated quantum spin lattices J. Richter*

Institut fu¨r Theoretische Physik, Universita¨t Magdeburg, P.O. Box 4120, D-39016 Magdeburg, Germany ͑Submitted January 19, 2005͒ Fiz. Nizk. Temp. 31, 918–928 ͑August–September 2005͒ Recent developments concerning localized-magnon eigenstates in strongly frustrated spin lattices and their effect on the low-temperature physics of these systems in high magnetic fields are reviewed. After illustrating the construction and the properties of localized-magnon states we describe the plateau and the jump in the magnetization process caused by these states. Considering appropriate lattice deformations fitting to the localized magnons we discuss a spin- Peierls instability in high magnetic fields related to these states. Last but not least we consider the degeneracy of the localized-magnon eigenstates and the related thermodynamics in high magnetic fields. In particular, we discuss the low-temperature maximum in the isothermal entropy versus field curve and the resulting enhanced magnetocaloric effect, which allows efficient magnetic cooling from quite large temperatures down to very low ones. © 2005 American Institute of Physics. ͓DOI: 10.1063/1.2008130͔

I. INTRODUCTION theoretical study of frustrated quantum spin systems is chal- lenging and is often faced with particular problems. While The interest in quantum spin antiferromagnetism has a for unfrustrated systems a wide class of well developed very long tradition; see, e.g., Ref. 1. Stimulated by the recent many-body methods are available, at least some of them, progress in synthesizing magnetic materials with strong e.g., the powerful quantum Monte Carlo method, fail for quantum fluctuations,2 particular attention has been paid to frustrated systems. Furthermore several important exact low-dimensional quantum magnets showing novel quantum statements like the Marshall-Peierls sign rule12 and the Lieb- phenomena like spin-liquid phases, quantum phase transi- 13 tions or plateaus and jumps in the magnetization process. Mattis theorem are not generally valid if frustration is ͑ ͒ However, quantum spin systems are of interest in their own present see, e.g., Refs. 14, 15 . right as examples of strongly interacting quantum many- On the other hand, the investigation of strongly frus- body systems. trated magnetic systems surprisingly led to the discovery of We know from the Mermin-Wagner theorem3 that ther- several new exact eigenstates. To find exact eigenstates of mal fluctuations are strong enough to destroy magnetic long- quantum many-body systems is, in general, a rare exception. range order ͑LRO͒ for Heisenberg spin systems in dimension For spin systems one has only a few examples. The simplest DϽ3 at any finite temperature T. For Tϭ0, where only example for an exact eigenstate is the fully polarized ferro- quantum fluctuations are present, the situation seems to be magnetic state, which becomes the ground state of an anti- more complicated. While for one-dimensional ͑1D͒ antifer- ferromagnet in a strong magnetic field. Furthermore the one- romagnets, in general, the quantum fluctuations are strong and two-magnon excitations above the fully polarized ferro- enough to prevent magnetic LRO, the competition between magnetic state also can be calculated exactly ͑see, e.g., Ref. interactions and fluctuations is well balanced in two dimen- 1͒. An example for non-trivial eigenstates is Bethe’s famous 16 sions, and one meets magnetic LRO as well as magnetic solution for the 1D Heisenberg antiferromagnet ͑HAFM͒. disorder at Tϭ0, depending on the details of the lattice.4–8 It Some of the eigenstates found for frustrated quantum mag- was pointed out many years ago by Anderson and Fazekas9,10 nets are of quite simple nature and their physical properties, that competition of magnetic bonds for instances due to tri- e.g., the spin correlation functions, can be calculated analyti- angular configurations of antiferromagnetically interacting cally. Note that such states are often eigenstates of the un- spins may influence this balance and can lead to disordered frustrated system, too, but they are irrelevant for the physics ground-state phases in two-dimensional ͑2D͒ quantum anti- of the unfrustrated system if they are lying somewhere in the ferromagnets. In the context of spin glasses this competition spectrum. However, the interest in these eigenstates comes of bonds, later called frustration, was discussed in great de- from the fact that they may become ground states for par- tail. These studies on spin glasses have shown that frustration ticular values of frustration. Therefore these exact eigenstates may have an enormous influence on the ground state and play an important role either as ground states of real quan- thermodynamic properties11 of spin systems. tum magnets or at least as ground states of idealized models The investigation of frustration effects in spin systems, which can be used as reference states for more complex especially in combination with strong quantum fluctuations, quantum spin systems. is currently a hot topic in solid state physics. We mention Two well-known examples for simple eigenstates of some interesting features like quantum disorder, incommen- strongly frustrated quantum spin systems are the Majumdar- 17 surate spiral phases, ‘‘order by disorder’’ phenomena to Gosh state of the 1D J1 – J2 spin-half HAFM and the or- name a few, which might appear in frustrated systems. The thogonal dimer state of the Shastry-Sutherland model.18 Both

1063-777X/2005/31(8–9)/9/$26.00695 © 2005 American Institute of Physics 696 Low Temp. Phys. 31 (8–9), August–September 2005 J. Richter

eigenstates are product states built by dimer singlets. They become ground states only for strong frustration. These eigenstates indeed play a role in realistic materials. The Majumdar-Ghosh state has some relevance in quasi-1D spin- ͑ ͒ Peierls materials like CuGeO3 see, e.g., Ref. 19 . The or- thogonal dimer state of the Shastry-Sutherland model is the magnetic ground state of the quasi-two-dimensional 20 SrCu2(BO3)2 compound. Other frustrated spin models in one, two or three dimensions are known which have also dimer-singlet product states as ground states ͑see, e.g., Refs. 21–24͒. Note that these dimer-singlet product ground states have gapped magnetic excitations and lead therefore to a plateau in the magnetization m at mϭ0. Recently it has been demonstrated for the 1D counterpart of the Shastry- 22,25–27 Sutherland model that more general product eigen- FIG. 1. Excitation energies wi(k) of the one-magnon states for the isotropic spin-half Heisenberg antiferromagnet ͓i.e., Jϭ1, ⌬ϭ1, sϭ1/2, and hϭ0in states containing chain fragments of finite length can lead to ͑ ͔͒ ͑ ͒ 26 Eq. 1 on the kagome´ lattice right inset . The left inset shows the path in an infinite series of magnetization plateaus. Finally, we the Brillouin zone corresponding to k values used as x coordinate. This mention the so-called frustrated Heisenberg star, where also figure was provided by J. Schulenburg. exact statements on the ground state are known.28 In this paper we review recent results concerning a new class of exact eigenstates appearing in strongly frustrated an- 1 N ͉ ϭ ˆ Ϫ͉ ϭ͉⌿ ͉⌿ tiferromagnets, namely the so-called localized-magnon 1͘L ͚ aiSi 0͘ L͘ R͘, c i states. These states have been detected as ground states of certain frustrated antiferromagnets29–31 in a magnetic field, 0 ᭙i෈L ͑local͒ a ͭ ͑3͒ and their relevance for physical properties of a wide class of i ϭ0 ᭙i෈R ͑remainder͒, frustrated magnets has been discussed in Refs. 7 and 29–37. ͉⌿ where L͘ belongs to the localized excitation living on the ͉⌿ local region L and R͘ describes the fully polarized ferro- magnetic remainder. We split the Hamiltonian into three II. LOCALIZED MAGNON STATES ˆ ϭ ˆ ϩ ˆ ϩ ˆ ˆ parts H HL HLϪR HR , where HL contains all bonds Jil ෈ ˆ ෈ We consider a Heisenberg XXZ antiferromagnet for gen- with i,l L, HR contains all bonds Jkj with k, j R, and ˆ ෈ ෈ eral spin quantum number s in a magnetic field h HLϪR contains all bonds Jlk with l L and k R. The re- quirement that the localized-magnon state ͉1͘ is simulta- 1 L ˆ ϭ ͭ ⌬ ˆ z ˆ zϩ ͑ ˆ ϩ ˆ Ϫϩ ˆ Ϫ ˆ ϩ͒ͮ Ϫ ˆ z neously an eigenstate of all three parts of the Hamiltonian, H ͚ Jij Si S j Si S j Si S j hS , ͗ij͘ 2 ˆ ͉ ϭ ͉ ˆ ͉ ϭ ͉ ˆ ͉ i.e., HL 1͘L eL 1͘L , HR 1͘L eR 1͘L and HLϪR 1͘L ϭ ͉ ͘ у ͑ ͒ eLϪR 1 L , leads to two criteria for the exchange bonds Jij 0. 1 31 Jij , namely ϭ ˆ zϭ͚ ˆ z The magnetization M S iS commutes with the Hamil- i ϭ ᭙ ෈ ͑ ͒ tonian and is used as a relevant quantum number to charac- ͚ Jrlal 0 r R 4 l෈L terize the eigenstates of Hˆ . Let us consider strong magnetic and fields exceeding the saturation field hsat . Then the system is in the fully polarized ferromagnetic eigenstate ͉FM͘ϭ͉ϩs, ϭ ᭙ ෈ ͑ ͒ ϩs,ϩs,ϩs,ϩs ...͘, which will be considered as the mag- ͚ Jrl const l L. 5 ෈ non vacuum state, i.e. ͉0͘ϭ͉FM͘. The one-magnon states r R above this vacuum are given by Equation ͑4͒ represents a condition on the bond geometry, whereas Eq. ͑5͒ is a condition for the bond strengths and is 1 N ͉ ϭ Ϫ͉ ͑ ˆ Ϫ ͉͒ ϭ ͑ ͉͒ ͑ ͒ automatically fulfilled in uniform lattices with equivalent 1͘ ͚ aiSi 0͘; H EFM 1͘ wi k 1͘, 2 c i sites. Note, however, that the second condition is not a nec- essary one, i.e., one can find models with eigenstates of form where EFM is the energy of the fully polarized ferromagnetic ͑3͒ violating ͑5͒; see Ref. 30. This more general case appears state ͉FM͘ and c is a normalization constant. For several strongly frustrated lattices one observes flat dispersion ϭ modes wi(k) const of the lowest branch. One example is shown in Fig. 1. Here we mention that there is some relation to the flat-band ferromagnetism in electronic systems dis- cussed by Mielke and Tasaki; see Refs. 38–40. Consequently, one can construct localized states by an appropriate superposition of extended states with different k vectors. The general form of these localized states can be FIG. 2. Typical lattice geometries supporting the localized-magnon states 30,31 written as ͑3͒. The magnon lives on the restricted area indicated by a thick line. Low Temp. Phys. 31 (8–9), August–September 2005 J. Richter 697

Due to the mutual independence of the localized mag- р ϰ nons for n nmax N the energy Eloc of the localized-magnon states is proportional to the number n of localized magnons, ϭ ϭ Ϫ ␧ i.e., at h 0 one has Eloc EFM n 1 , where EFM is the ͑ ͒ ␧ energy of the fully polarized vacuum state and 1 is the energy gain by one magnon. As a results there is a simple linear relation between Eloc and the total magnetization M valid for all systems hosting localized magnons: FIG. 3. Localized magnons on the kagome´ lattice. One-magnon state indi- E ͑M,hϭ0͒ϭϪaNϩbsM, ͑7͒ cated by a circle, the ϩ and Ϫ signs correspond to the sign of the coeffi- loc ϭϮ ͓ ͑ ͔͒ ͑ ͒ cients ai 1 see Eq. 6 a . Magnon crystal state corresponding to the where the parameters a and b depend on details of the sys- ϭ ´ maximum filling nmax N/9 of the kagome lattice with localized magnons tem like the exchange constant J , the anisotropy parameter ͑circles͒͑b͒. The figures were taken from Ref. 30. ij ⌬, the coordination number z of the lattice, etc. For example, for the isotropic spin-half HAFM with NN exchange J on the kagome´ lattice one finds30,35 if ͉⌿ ͉͘⌿ ͘ is not an individual eigenstate of both Hˆ and L R L ͑ ϭ ͒ϭ 2 Ϫ ϭϪ 2 ϩ ͑ ͒ ˆ ˆ ϩ ˆ Eloc M,h 0 2s NJ 6snJ 4s NJ 6sJM. 8 HLϪR but of (HLϪR HL). Hence, the geometry condition ͑4͒ is the criterion of major importance. A typical geometry Note that also the spin-spin correlation functions of such fulfilling condition ͑4͒ is realized by an even polygon sur- states can be easily found.31 rounded by isosceles triangles; see Fig. 2. The lowest one- For the physical relevance of these eigenstates it is cru- magnon state living on an even polygon has coefficients ai in cial that they have lowest energy in the corresponding sector ͉ 1͘L alternating in sign. But also finite strings of two or three of M. Indeed this can be proved for quite general antiferro- sites attached by appropriate triangles can fulfill the criterion magnetic spin models.29,41 ͑4͒; see Fig. 2. Since the condition ͑4͒ for the existence of localized- As an example we consider the HAFM ͑1͒ on the magnon states is quite general, one can find a lot of magnetic 30 ϭ kagome´ lattice, i.e., we have Jij 1 for nearest-neighbor systems in one, two, and three dimensions having localized- ͑ ͒ ϭ 29–32,35–37,43 NN bonds and Jij 0 otherwise. Its one-magnon dispersion magnon eigenstates. To illustrate that we show in shown in Fig. 1 exhibits one flat mode. The resulting local- Fig. 4 some 1D and in Fig. 5 some 2D and 3D antiferromag- ized magnon lives on a hexagon; see Fig. 3a. Its wave func- netic spin lattices. But localized-magnon states are observed tion is for frustrated magnetic molecules also.29 1 6 ͉ kagome´ϭ ˆ Ϫ͉ ϭ͑Ϫ ͒i ͑ ͒ 1͘L ͚ aiSi 0͘, ai 1 . 6 III. PLATEAUS AND JUMPS IN THE MAGNETIZATION ͱ ϭ 12S i 1 ͑i෈hexagon͒ CURVE Note that the number of hexagons N/3 corresponds to the As discussed in the previous Section, the localized- number of states in the flat branch of w(k). Because of the magnon states are the lowest states in the corresponding sec- localized nature of the magnon we can put further such mag- tor of magnetization M. Hence they become ground states in nons on the lattice in such a way that there is no interaction appropriate magnetic fields h. Furthermore we stated that between them. The maximum filling nmax of the kagome´ lat- there is a linear relation between the energy of these states ͑ ͒ tice with localized magnons is shown in Fig. 3b. The result- Eloc and M; cf. Eq. 7 . When a magnetic field h is applied ϭ ϭϪ ϩ Ϫ ing eigenstate is a magnon ‘‘crystal’’ state with nmax N/9 the energy reads Eloc(M,h) aN bsM hM, and one magnons and a magnetic unit cell three times as large as the has complete degeneracy of all localized-magnon states at ϭ ϭ geometric one. Therefore the states with n the saturation field h hsat bs. As a result of this degen- ϭ͕0,1,2,3,...,N/9͖ localized magnons represent the class eracy the zero-temperature magnetization mϭM/(Ns) of exactly known product eigenstates with quantum numbers jumps between the saturation value mϭ1 and the value 1 ϭ ϭ Ϫ ϭ Ϫ Ϫ Ϫ Ϫ M Sz Ns n ͕Ns,Ns 1,Ns 2,...,Ns N/9͖. nmax /(Ns) corresponding to the maximum number nmax of

FIG. 4. One-dimensional systems with localized- magnon states. The magnons live on the restricted area indicated by a thick line: sawtooth chain42,43 ͑a͒, kagome´ chain I44 ͑b͒, kagome´ chain II45 ͑c͒, diamond chain46 ͑d͒, dimer-plaquette chain22 ͑e͒, frustrated ladder47 ͑f͒. Note that there are special restrictions for the exchange integrals to have localized-magnon states as lowest eigenstates in the corresponding sector of M. 698 Low Temp. Phys. 31 (8–9), August–September 2005 J. Richter

FIG. 5. Two- and three-dimensional antiferromagnets with localized- magnon states. The magnons live on the restricted area indicated by a thick line: planar pyrochlore ͑check- erboard͒ lattice31,48 ͑a͒, square- kagome´ lattice37,49 ͑b͒, star lattice7,36 ͑c͒, pyrochlore lattice50 ͑d͒.

IV. MAGNETIC-FIELD INDUCED SPIN-PEIERLS INSTABILITY independent localized magnons. Since nmax is proportional to N but independent of the spin quantum number s, the height IN QUANTUM SPIN LATTICES WITH LOCALIZED- ␦ ϭ ϭϱ MAGNON STATES of the jump m nmax /(Ns) goes to zero for s , i.e., the magnetization jump due to the localized-magnon states be- comes irrelevant if the spins become classical. The influence of a magnetoelastic coupling in frustrated In Fig. 6 we present two examples of the magnetization antiferromagnets on their low-temperature properties is cur- curves; further examples can be found in Refs. 7, 27, 31, 36, rently widely discussed. Lattice instabilities breaking the 37. The jumps of height ␦mϭ1/2 ͑sawtooth chain͒ and ␦m translational symmetry are reported for 1D and 2D as well ϭ2/7 ͑kagome´ lattice͒ are well pronounced. Furthermore we 3D quantum spin systems; see, e.g., Ref. 53. In all those see a wide plateau at the foot of the jump for the sawtooth studies the lattice instability was considered at zero field. As 54 chain. Note that there are general arguments in favor of a discussed already in an early paper by Gross, a magnetic plateau just below the jump.7,51,52 Therefore we might expect field usually acts against the spin-Peierls transition and might a plateau preceeding the jump in all magnetic systems with favor a uniform or incommensurate phase. In contrast to localized magnon states. Though from Fig. 6 it remains un- those findings, in this Section we discuss a lattice instability clear whether there is a plateau for the kagome´ lattice, too, a in frustrated spin lattices hosting localized magnons for more detailed analysis32 indeed yields evidence for a finite which the magnetic field is essential.32 plateau width of about ␦hϳ0.07J. First we point out that due to the localized nature of the

ϭ FIG. 6. Panel a: Magnetization versus magnetic field h of the isotropic spin-half Heisenberg antiferromagnet. The sawtooth chain with J1 1andJ2 ϭ ϭ ͑ ͒ ͑ ͒ 2J1 ; cf. Fig. 4. The figure was taken from Ref. 31. Panel b: The kagome´ lattice with N 27 dashed line and 36 sites solid line . The inset illustrates the ‘‘magnon crystal’’ state corresponding to maximum filling with localized magnons, where the location of the magnons is indicated by the circles in certain hexagons. The figure was taken from Ref. 8. Low Temp. Phys. 31 (8–9), August–September 2005 J. Richter 699

states are the ground states of the systems then we have a favorable spin-Peierls distortion with ␦*ϭJ/(12␥). As dis- cussed in the previous Section, we expect a plateau at the foot of the magnetization jump. The spin-Peierls distortion due to localized-magnon states then takes place for the val- ues of the magnetic field belonging to this plateau. Now the question arises whether the lattice distortion under consideration is stable below and above this plateau, Ͻ Ϫ ϭ i.e., for M N/2 nmax and M N/2. It is easy to check that the lattice distortion illustrated in Fig. 7 is not favorable for the fully polarized vacuum state, i.e. for MϭN/2. For mag- netizations M below this plateau, we are not able to give a rigorous answer, but numerical results for finite kagome´ lat- ϭ FIG. 7. Kagome´ lattice with one distorted hexagon which can host localized tices of size N 18,27,36,45,54 indicate that there is no spin- magnons. The parts of the lattices before distortions are shown by dashed Peierls deformation adopting the lattice distortion shown in lines. All bonds in the lattices before distortions have the same length. The Ͻ Ϫ ͑ ͒ Fig. 7 for M N/2 nmax see Ref. 32 for more details . figure was taken from Ref. 32. We mention that the scenario discussed above basically remains unchanged for the anisotropic Hamiltonian ͑1͒ with ⌬ 1 and also for spin quantum numbers sϾ1/2.32 magnons we have an inhomogeneous distribution of NN From the experimental point of view the discussed effect ͗ ˆ ˆ ͘ 31 should manifest itself most spectacularly as a hysteresis in spin-spin correlations Si•Sj . In the case that one magnon is distributed uniformly over the lattice, the deviation of the the magnetization and the deformation of kagome´-lattice an- NN correlation from the ferromagnetic value, i.e. the quan- tiferromagnets in the vicinity of the saturation field. We em- ͗ ˆ ˆ ͘Ϫ phasize that the discussed spin-Peierls instability in high tity Si•Sj 1/4, is of order 1/N. On the other hand, for a localized magnon ͑3͒ we have along the polygon/line hosting magnetic fields may appear in the whole class of frustrated the localized magnon in general actually a negative NN cor- quantum magnets in one, two and three dimensions hosting independent localized magnons provided it is possible to relation ͗Sˆ Sˆ ͘, and all other NN correlations are positive. i• j construct a lattice distortion preserving the symmetry of the For instance for the spin-half HAFM on the kagome´ lattice ͑ ͒ ͗ ˆ ˆ ͘ϭϪ localized-magnon cell. the localized-magnon state 6 leads to Si•Sj 1/12 for neighboring sites i and j on a hexagon hosting a magnon and to ͗Sˆ Sˆ ͘ϭϩ1/6 for neighboring sites i and j on an attach- i• j V. FINITE LOW-TEMPERATURE ENTROPY AND ENHANCED ing triangle. Hence a deformation with optimal gain in mag- MAGNETOCALORIC EFFECT IN THE VICINITY OF THE netic energy will lead to an increase of antiferromagnetic SATURATION FIELD bonds on the polygon/line hosting the localized magnon ͑i.e., hexagon for the kagome´ lattice͒ and to a decrease of the It is well-known that strongly frustrated Ising or classical bonds on the attaching triangles. Heisenberg spin systems may exhibit a huge ground-state Let us discuss the situation for the isotropic spin-half degeneracy; see e.g., Refs. 5–7, 55–58. For quantum sys- HAFM on the kagome´ lattice in more detail. A correspond- tems the degeneracy found for classical systems is often ing deformation which preserves the symmetry of the cell lifted due to fluctuations, and the quantum ground state is which hosts the localized magnon is shown in Fig. 7. To unique ͑the so-called ‘‘order from disorder’’ check the stability of the kagome´ lattice with respect to a phenomenon60,61͒. However, highly frustrated antiferromag- spin-Peierls mechanism we must compare the magnetic and netic quantum spin lattices hosting localized magnons are an the elastic energies. For the kagome´ lattice the deformation example where one finds a huge ground-state degeneracy in shown in Fig. 7 leads to the following changes in the ex- a quantum system. As discussed in Sec. 3 the localized- change interactions: J→(1ϩ␦)J ͑along the edges of the magnon states become degenerate at saturation field. As hexagon͒ and J→(1Ϫ␦/2)J ͑along the two edges of the tri- pointed out first in Ref. 7 the degeneracy grows exponen- angles attached to the hexagon͒, where the quantity ␦ is pro- tially with system size N. This huge degeneracy and its con- portional to the displacement of the atoms and the change in sequence for the low-temperature physics are discussed in the exchange integrals due to lattice distortions is taken into more detail in Refs. 33–35. account in first order in ␦. The magnetic energy ͑8͒ is low- For some of the spin systems hosting localized magnons ered by distortions and becomes for one magnon and one the ground-state degeneracy at saturation and therefore the ϭ ϭ ␦ ϭ Ϫ corresponding distortion Eloc(n 1,h 0, ) NJ/2 3J residual entropy at zero temperature can be calculated ex- Ϫ ␦ р 3 J/2. Considering n nmax independent localized mag- actly by mapping the localized-magnon problem onto a re- nons and corresponding distortions the energy gain is then lated lattice gas model of hard-core objects.33–35 These lat- ϭϪ ␦ emag 3n J/2. On the other hand, the elastic energy in the tice gas models of hard-core objects have been studied over ϰ␦2 ͑ ͒ harmonic approximation increases according to eelast . the last decades in great detail see, e.g., Ref. 59 . Let us Therefore a minimal total energy is obtained for a finite ␦ illustrate this for the sawtooth chain. Here the local areas ϭ␦*Ͼ0. For the kagome´ lattice the elastic energy for one where the magnons can live are the valleys of the sawtooth ϭ ␥␦2 ͑ ␥ distorted cell is eelast 9 the parameter is proportional chain; see Fig. 4a. Because a certain local area of the lattice to the elastic constant of the lattice͒. If the localized-magnon can be occupied by a magnon or not, the degeneracy of the 700 Low Temp. Phys. 31 (8–9), August–September 2005 J. Richter

monomers or by rigid dimers occupying two neighboring sites. The limiting behavior of W for N→ϱ was found by Fisher many years ago:59 1ϩͱ5 Wϭexpͩ log Nͪ Ϸexp͑0.240606N͒ 2 S FIG. 8. The sawtooth chain which hosts three localized magnons at fat V → ϭ0.240606. ͑10͒ parts ͑a͒ and the auxiliary lattice used for the exact calculation of the N ground-state degeneracy at saturation ͑b͒. The localized magnons are eigen- ϭͱ ϩ⌬ ͑ ͒ The relevance of this result for experimental studies emerges states for J2 2(1 )J1 Ref. 30 . The figure was taken from Ref. 35. at low but finite temperatures. In Ref. 33 this mapping was used to obtain a quantitative description of the low- ground state at saturation, W, grows exponentially with N, temperature magnetothermodynamics in the vicinity of the giving rise to a finite zero-temperature entropy per site at saturation field of the quantum antiferromagnet on the ´ saturation: kagome lattice and on the sawtooth chain. In what follows we report on the extension of the ana- S 1 lytical findings for the zero-temperature entropy at the satu- ϭ lim log WϾ0. ͑9͒ N N→ϱN ration field to finite temperatures and to arbitrary magnetic fields using exact diagonalization data for the sawtooth chain Now we map the original lattice which hosts localized mag- of up to Nϭ20 sites.34,35 In Fig. 9a we show the isothermal nons onto an auxiliary lattice which is occupied by hard-core entropy versus magnetic field for several temperatures. The objects, which are rigid monomers and dimers in the case of presented results show that for several magnetic fields below the sawtooth chain; see Fig. 8. The auxiliary chain ͑Fig. 8b͒ Ͻ saturation h hsat one has a twofold or even threefold degen- consists of NϭN/2 sites which may be filled either by rigid eracy of the energy levels, leading in a finite system to a

FIG. 9. Field dependence of the isothermal entropy per site for the Heisenberg antiferromagnet on the sawtooth chain of different length (sϭ1/2, ⌬ϭ1). ϭ ϭ ͓ ͑ ͔͒ ͑ ͒ J1 1, J2 2, i.e., the condition for bond strengths Eq. 4 is fulfilled and the localized-magnon states are exact eigenstates a . Influence of a deviation from the perfect condition for bond strengths ͑b͒. The figures were taken from Ref. 35. Low Temp. Phys. 31 (8–9), August–September 2005 J. Richter 701

finite zero-temperature entropy per site. Correspondingly one finds in Fig. 9a ͑upper panel͒ a peaked structure and more- over a plateau-like area just below hsat . However, it is clearly seen in Fig. 9a ͑upper panel͒ that the height of the peaks and of the plateau decreases with system size N, lead- ϭ ϭ Ͻ Ͼ ing to limN→ϱS/N 0atT 0 for h hsat and h hsat . Only ϭ ϭ the peak at h hsat 4 is independent of N and remains finite for N→ϱ. At finite but low temperatures this peak survives as a well-pronounced maximum, and it only disappears if the temperature is of the order of the exchange constant; see Fig. 9a ͑lower panel͒. Note that the value of the entropy at satu- ration, which agrees with the analytical prediction ͑10͒,is almost temperature independent up to about TϷ0.2. Thus, the effect of the independent localized magnons leading to a finite residual zero-temperature entropy is present at finite FIG. 10. Lines of constant entropy versus field ͓i.e., adiabatic ͑de͒magneti- Շ zation curves͔ for the Heisenberg antiferromagnet on the sawtooth chain of temperatures T 0.2, producing a noticeable maximum in ϭ ⌬ϭ ϭ ϭ different length (s 1/2, 1, J1 J/2, J2 J). The figure was taken from the isothermal entropy curve at the saturation field. We men- Ref. 34 with the kind permission of A. Honecker. tion that the numerical results for higher spin quantum num- bers s suggest that the enhancement of the entropy at satu- ration for finite temperatures becomes less pronounced with tive of the isothermal entropy with respect to the magnetic increasing s. field With respect to the experimental observation of the ץ ץ ץ T ͑ S/ h͒T maximum in the low-temperature entropy at the saturation ͩ ͪ ϭϪT , ͑11͒ h Cץ field in real compounds we are faced with the situation that S the condition on bond strengths ͓see Eq. ͑4͔͒ under which the where C is the specific heat. Again one can calculate the field localized-magnon states become the exact eigenstates are dependence of the temperature for an adiabatic ͑de͒magneti- certainly not strictly fulfilled. For the considered isotropic zation process for finite systems by exact diagonalization. spin-half HAFM on the sawtooth chain this condition is ful- Some results for the isotropic spin-half HAFM on the saw- ϭ filled for J2 2J1 ; see Fig. 9a. Based on the numerical cal- tooth chain are shown in Fig. 10. The lowest curves in Fig. culations one is able to discuss the ‘‘stability’’ of the maxi- 10 belongs to S/Nϭ0.05 and Nϭ12,16,20, respectively. The mum in the entropy against deviation from the perfect other curves correspond to S/Nϭ0.1,0.15,...,0.4,0.45. The condition for bond strengths. In Fig. 9b we show the field magnetocaloric effect is largest in the vicinity of the satura- dependence of the entropy at low temperatures for the saw- tion field. In particular, a demagnetization coming from mag- tooth chain of Nϭ16 sites with J ϭ1 and J ϭ1.9 and J 1 2 2 netic fields larger than hsat is very efficient. If one starts for ϭ Ͼ 2.1. Since the degeneracy of the ground state at saturation h hsat with an entropy lower than the residual entropy at ϭ ͓ ͑ ͔͒ is lifted when J2 2J1 , the entropy at saturation at very low hsat , S/N 0.240606 see Eq. 10 , then one observes even ͑ → → temperatures long-dashed and short-dashed curves in the a cooling to T 0ash hsat . Thus frustrated magnetic sys- upper panel of Fig. 9b͒ is not enhanced at the saturation field. tems hosting localized magnons allow magnetic cooling However, the initially degenerate energy levels remain close from quite large T down to very low temperatures. We men- to each other if J2 deviates only slightly from the perfect tion, that the results shown in Fig. 10 also clearly demon- ϭ value 2J1 . Therefore with increasing temperature those lev- strate that finite-size effects are very small for h hsat at any els become accessible for the spin system, and one again temperature. Therefore the above discussion is valid also for obtains a maximum in the entropy in the vicinity of satura- large systems N→ϱ. tion at low but nonzero temperatures; see the lower panel in Fig. 9b ͑long-dashed and short-dashed peaks in the vicinity of saturation͒. We emphasize that the low-temperature maxi- VI. SUMMARY mum of S/N at saturation is a generic effect for strongly We have reviewed recent results29–37 on exact eigen- frustrated quantum spin lattices which may host independent states constructed from localized magnons which appear in a localized magnons. class of frustrated spin lattices. For these eigenstates several Let us remark that the ground-state degeneracy problem quantities like the energy and the spin-spin correlation can be of antiferromagnetic Ising lattices in the critical magnetic calculated analytically. The physical relevance of the field ͑i.e., at the spin-flop transition point͒, which obviously localized-magnon eigenstates emerges at high magnetic do not contain quantum fluctuations, has been discussed in fields where they can become ground states of the spin sys- the literature; see, e.g., Ref. 62. tem. It has been pointed out very recently by Zhitomirsky and For frustrated magnetic systems having localized- Honecker34,63 that the most spectacular effect accompanying magnon ground states several interesting physical effects as- a maximum in the isothermal entropy S(h) is an enhanced sociated with this states may occur. First one finds a macro- magnetocaloric effect. Indeed, the cooling rate for an adia- scopic jump in the zero-temperature magnetization curve at ͑ ͒ batic de magnetization process is proportional to the deriva- the saturation field hsat . This jump is a true quantum effect 702 Low Temp. Phys. 31 (8–9), August–September 2005 J. Richter which vanishes if the spins become classical (s→ϱ). At the 7 J. Richter, J. Schulenburg, and A. Honecker, ‘‘Quantum magnetism in two foot of the jump one can expect a plateau in the magnetiza- dimensions: from semi-classical Ne´el order to magnetic disorder,’’ in tion curve. Quantum Magnetism, Vol. 645 of Lecture Notes in Physics, U. Scholl- wo¨ck, J. Richter, D. J. J. Farnell, and R. F. Bishop ͑eds.͒, Springer-Verlag, Since all localized-magnon states have the same energy Berlin ͑2004͒, p. 85; see also cond-mat/0412662. ϭ 8 at h hsat a huge degeneracy of the ground state at saturation A. Honecker, J. Schulenburg, and J. Richter, J. Phys.: Condens. Matter 16, is observed which increases exponentially with system size S749 ͑2004͒. 9 ͑ ͒ N, thus leading to a nonzero residual zero-temperature en- P. W. Anderson, Mater. Res. Bull. 8, 153 1973 . 10 P. Fazekas and P. W. Anderson, Philos. Mag. 30, 423 ͑1974͒. tropy. For some of the frustrated spin models hosting local- 11 K. Binder and A. P. Young, Rev. Mod. Phys. 58,801͑1986͒. ized magnons the residual entropy at saturation field can be 12 W. Marshall, Proc. R. Soc. London, Ser. A 232,48͑1955͒. estimated exactly. At finite temperatures T the localized- 13 E. Lieb and D. Mattis, ‘‘Ordering energy levels of interacting spin sys- ͑ ͒ magnon states produce a maximum in the isothermal entropy tems,’’ J. Math. Phys. 3,749 1962 . 14 J. Richter, N. B. Ivanov, and K. Retzlaff, Europhys. Lett. 25,545͑1994͒. versus field curve in the vicinity of the saturation field for not 15 J. Richter, N. B. Ivanov, K. Retzlaff, and A. Voigt, J. Magn. Magn. Mater. ͑ ͒ too large T. This maximum in the isothermal entropy at hsat 140–144, 1611 1995 . leads to an enhanced magnetocaloric effect. If one starts for 16 H. A. Bethe, Z. Phys. 71,205͑1931͒. Ͼ 17 C. K. Majumdar and D. K. Ghosh, J. Math. Phys. 10, 1399 ͑1969͒. h hsat with an entropy lower than the residual entropy at 18 ͑ ͒ → → B. S. Shastry and B. Sutherland, Physica B 108, 1069 1981 . hsat then one observes even a cooling to T 0ash hsat . 19 D. Poilblanc, J. Riera, C. A. Hayward, C. Berthier, and M. Horvatic, Phys. This may allow cooling from quite large T down to very low Rev. B 55, R11941 ͑1997͒. temperatures. 20 S. Miyahara and K. Ueda, Phys. Rev. Lett. 82, 3701 ͑1999͒. 21 ͑ ͒ Last but not least, the localized-magnon states may lead A. Pimpinelli, J. Phys.: Condens. Matter 3, 445 1991 . 22 N. B. Ivanov and J. Richter, Phys. Rev. A 232,308͑1997͒; J. Richter, N. to a spin-Peierls instability in strong magnetic fields, for in- B. Ivanov, and J. Schulenburg, J. Phys.: Condens. Matter 10,3635͑1998͒. stance for the antiferromagnetic kagome´ spin lattice. For this 23 K. Ueda and S. Miyahara, J. Phys.: Condens. Matter 11, L175 ͑1999͒. system the magnetic-field driven spin-Peierls instability 24 H.-J. Schmidt, J. Phys. A 38, 2133 ͑2005͒. 25 ͑ ͒ breaks spontaneously the translational symmetry of the A. Koga, K. Okunishi, and N. Kawakami, Phys. Rev. B 62,5558 2002 ; A. Koga and N. Kawakami, Phys. Rev. B 65, 214415 ͑2002͒. kagome´ lattice and appears only in a certain region of the 26 J. Schulenburg and J. Richter, Phys. Rev. B 65, 054420 ͑2002͒. magnetic field near saturation. 27 J. Schulenburg and J. Richter, Phys. Rev. B 66, 134419 ͑2002͒. We emphasize that the reported effects are generic in 28 J. Richter and A. Voigt, J. Phys. A 27, 1139 ͑1994͒; J. Richter, A. Voigt, ¨ ͑ ͒ highly frustrated magnets. To observe them in experiments and S. Kruger, J. Magn. Magn. Mater. 140–144,1497 1995 ; J. Richter, A. Voigt, and S. Kru¨ger, J. Phys. A 29, 825 ͑1996͒. one needs frustrated magnets with small spin quantum num- 29 J. Schnack, H.-J. Schmidt, J. Richter, and J. Schulenburg, Eur. Phys. J. B ber s and sufficiently small exchange coupling strength J to 24,475͑2001͒. reach the saturation field. There is an increasing number of 30 J. Schulenburg, A. Honecker, J. Schnack, J. Richter, and H.-J. Schmidt, Phys. Rev. Lett. 88, 167207 ͑2002͒. synthesized quantum frustrated spin lattices, e.g., quantum 31 64–67 J. Richter, J. Schulenburg, A. Honecker, J. Schnack, and H.-J. Schmidt, J. antiferromagnets with a kagome´-like structure. Though Phys.: Condens. Matter 16, S779 ͑2004͒. such materials often do not fit perfectly to the lattice geom- 32 J. Richter, O. Derzhko, and J. Schulenburg, Phys. Rev. Lett. 93, 107206 etry having localized-magnon ground states, the physical ef- ͑2004͒. 33 M. E. Zhitomirsky and H. Tsunetsugu, Phys. Rev. B 70, 100403͑R͒ fects based on localized magnons states may survive in non- ͑ ͒ ͑ ͒ 2004 . ideal geometries see Sec. 5 , which may open the window to 34 M. E. Zhitomirsky and A. Honecker, J. Stat. Mech.: Theor. Exp. P07012 the experimental observation of the theoretically predicted ͑2004͒. 35 O. Derzhko and J. Richter, Phys. Rev. B 70, 104415 ͑2004͒. effects. 36 The author is indebted to O. Derzhko, A. Honecker, H.-J. J. Richter, J. Schulenburg, A. Honecker, and D. Schmalfuss, Phys. Rev. B 70, 174454 ͑2004͒. Schmidt, J. Schnack, and J. Schulenburg for the fruitful col- 37 J. Richter, J. Schulenburg, P. Tomczak, and D. Schmalfuss, cond-mat/ laboration in the field of localized-magnon states in frus- 0411673 ͑2004͒, submitted to Phys. Rev. B . 38 trated antiferromagnets. Most of the results discussed in the A. Mielke, J. Phys. A 24, 3311 ͑1991͒. 39 H. Tasaki, Phys. Rev. Lett. 69, 1608 ͑1992͒. paper are based on common publications with these col- 40 A. Mielke and H. Tasaki, Commun. Math. Phys. 158, 341 ͑1993͒. leagues. In particular, I thank J. Schulenburg for many valu- 41 H.-J. Schmidt, J. Phys. A 35, 6545 ͑2002͒. able discussions and hints. Furthermore I thank H. Frahm for 42 T. Nakamura and K. Kubo, Phys. Rev. B 53, 6393 ͑1996͒;D.Sen,B.S. directing my attention to the flat-band ferromagnetism. I also Shastry, R. E. Walstedt, and R. Cava, Phys. Rev. B 53, 6401 ͑1996͒. 43 V. Ravi Chandra, D. Sen, N. B. Ivanov, and J. Richter, Phys. Rev. B 69, acknowledge the support from the Deutsche Forschungsge- 214406 ͑2004͒. meinschaft ͑Project No. Ri615/12-1͒. 44 Ch. Waldtmann, H. Kreutzmann, U. Schollwo¨ck, K. Maisinger, and H.-U. Everts, Phys. Rev. B 62, 9472 ͑2000͒. 45 P. Azaria, C. Hooley, P. Lecheminant, and A. M. Tsvelik, Phys. Rev. Lett. ͑ ͒ *E-mail: [email protected] 81, 1694 1998 . 46 H. Niggemann, G. Uimin, and J. Zittartz, J. Phys.: Condens. Matter 9, 9031 ͑1997͒. 47 F. Mila, Eur. Phys. J. B 6, 201 ͑1998͒; A. Honecker, F. Mila, and M. Troyer, Eur. Phys. J. B 15, 227 ͑2000͒. 1 D. C. Mattis, The Theory of Magnetism I, Springer-Verlag, Berlin ͑1991͒. 48 S. E. Palmer and J. T. Chalker, Phys. Rev. B 64, 094412 ͑2001͒. 2 P. Lemmens and P. Millet, in Quantum Magnetism, Vol. 645 of Lecture 49 R. Siddharthan and A. Georges, Phys. Rev. B 65, 014417 ͑2002͒;P.Tom- Notes in Physics, U. Schollwo¨ck, J. Richter, D. J. J. Farnell, and R. F. czak and J. Richter, J. Phys. A 36,5399͑2003͒. Bishop ͑eds.͒, Springer-Verlag, Berlin ͑2004͒,p.433. 50 B. Canals and C. Lacroix, Phys. Rev. Lett. 80, 2933 ͑1998͒. 3 N. D. Mermin and H. Wagner, Phys. Rev. Lett. 17, 1133 ͑1966͒. 51 T. Momoi and K. Totsuka, Phys. Rev. B 61, 3231 ͑2000͒. 4 C. Lhuillier, P. Sindzingre, and J.-B. Fouet, Can. J. Phys. 79, 1525 ͑2001͒. 52 M. Oshikawa, Phys. Rev. Lett. 84,1535͑2000͒. 5 R. Moessner, Can. J. Phys. 79, 1283 ͑2001͒. 53 S.-H. Lee, C. Broholm, T. H. Kim, W. Ratcliff II, and S-W. Cheong, Phys. 6 G. Misguich and C. Lhuillier, in Frustrated Spin Systems,H.T.Diep͑ed.͒, Rev. Lett. 84,3718͑2000͒; Y. Yamashita and K. Ueda, Phys. Rev. Lett. 85, World-Scientific ͑2004͒. 4960 ͑2000͒; A. Keren and J. S. Gardner, Phys. Rev. Lett. 87, 177201 Low Temp. Phys. 31 (8–9), August–September 2005 J. Richter 703

͑2001͒; F. Becca and F. Mila, Phys. Rev. Lett. 89, 037204 ͑2002͒;P. 61 E. F. Shender, Zh. E´ ksp. Teor. Fiz. 83, 326 ͑1982͒͑Sov. Phys. JETP 56, Carretta, N. Papinutto, C. B. Azzoni, M. C. Mozzati, E. Pavarini, S. 178 ͑1982͒͒. Gonthier, and P. Millet, Phys. Rev. B 66, 094420 ͑2002͒; O. Tcherny- 62 B. D. Metcalf and C. P. Yang, Phys. Rev. B 18, 2304 ͑1978͒. shyov, R. Moessner, and S. L. Sondhi, Phys. Rev. Lett. 88, 067203 ͑2002͒; 63 M. E. Zhitomirsky, Phys. Rev. B 67, 104421 ͑2003͒. F. Becca, F. Mila, and D. Poilblanc, Phys. Rev. Lett. 91, 067202 ͑2003͒. 64 A. P. Ramirez, B. Hessen, and M. Winklemann, Phys. Rev. Lett. 84, 2957 54 M. C. Gross, Phys. Rev. B 20, 4606 ͑1979͒. ͑2000͒. 55 G. H. Wannier, Phys. Rev. 79, 357 ͑1950͒. 65 Zenji Hiroi, Masafumi Hanawa, Naoya Kobayashi, Minoru Nohara, Hide- 56 R. Moessner and S. L. Sondhi, Phys. Rev. B 63, 224401 ͑2001͒. nori Takagi, Yoshitomo Kato, and Masashi Takigawa, J. Phys. Soc. Jpn. 57 M. E. Zhitomirsky, Phys. Rev. Lett. 88, 057204 ͑2002͒. 70, 3377 ͑2001͒. 58 M. Udagawa, M. Ogata, and Z. Hiroi, J. Phys. Soc. Jpn. 71, 2365 ͑2002͒. 66 A. Fukaya, Y. Fudamoto, I. M. Gat, T. Ito, M. I. Larkin, A. T. Savici, Y. J. 59 M. E. Fisher, Phys. Rev. 124, 1664 ͑1961͒; R. J. Baxter, I. G. Enting, and Uemura, P. P. Kyriakou, G. M. Luke, M. T. Rovers, K. M. Kojima, A. S. K. Tsang, J. Stat. Phys. 22, 465 ͑1980͒; R. J. Baxter and S. K. Tsang, J. Keren, M. Hanawa, and Z. Hiroi, Phys. Rev. Lett. 91, 207603 ͑2003͒. ͑ ͒ ͑ ͒ Phys. A 13,1023 1980 ; R. J. Baxter, J. Phys. A 13,L61 1980 ;R.J. 67 D. Bono, P. Mendels, G. Collin, and N. Blanchard, Phys. Rev. Lett. 92, Baxter, Exactly Solved Models in Statistical Mechanics, Academic Press, 217202 ͑2004͒. London ͑1982͒; R. J. Baxter, Ann. Combinatorics 3, 191 ͑1999͒. 60 J. Villain, R. Bidaux, J. P. Carton, and R. Conte, J. Phys. A 41, 1263 This article was published in English in the original Russian journal: Repro- ͑1980͒. duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 31, NUMBERS 8–9 AUGUST–SEPTEMBER 2005

Correlations, spin dynamics, defects: the highly frustrated kagome´ bilayer D. Bono,* L. Limot,† and P. Mendels‡

Laboratoire de Physique des Solides, UMR 8502, Universite´ Paris-Sud, 91405 Orsay, France G. Collin

Laboratoire Le´on Brillouin, CE Saclay, CEA-CNRS, 91191 Gif-sur-Yvette, France N. Blanchard

Laboratoire de Physique des Solides, UMR 8502, Universite´ Paris-Sud, 91405 Orsay, France ͑Submitted March 21, 2005͒ Fiz. Nizk. Temp. 31, 929–951 ͑August–September 2005͒

The compounds SrCr9pGa12Ϫ9pO19 and Ba2Sn2ZnGa10Ϫ7pCr7pO22 are two highly frustrated magnets possessing a quasi-two-dimensional kagome´ bilayer of spin-3/2 chromium ions with antiferromagnetic interactions. Their magnetic susceptibility was measured by local nuclear magnetic resonance and nonlocal ͑SQUID͒ techniques, and their low-temperature spin dynamics by muon spin resonance. Consistent with the theoretical picture drawn for geometrically frustrated systems, the kagome´ bilayer is shown here to exhibit: ͑i͒ short range spin-spin correlations down to a temperature much lower than the Curie–Weiss temperature, no conventional long-range transition occurring; ͑ii͒ a Curie contribution to the susceptibility from paramagnetic defects generated by spin vacancies; ͑iii͒ low-temperature spin fluctuations, at least down to 30 mK, which are a trademark of a dynamical ground state. These properties point to a spin-liquid ground state, possibly built on resonating valence bonds with unconfined spinons as the magnetic excitations. © 2005 American Institute of Physics. ͓DOI: 10.1063/1.2008131͔

I. INTRODUCTION as highly frustrated magnet ͑HFM͒, is therefore an ideal can- A. Highly frustrated magnets didate to possess a RVB ground state. Moreover, since HFMs are insulators, the experimental investigation of their ground Anderson’s initial proposal of a resonating valence bond state, and the related theoretical modeling, is greatly simpli- ͑RVB͒ state was intended as a possible alternative for the fied by the absence of phenomena such as superconductivity, Ne´el ground state of a triangular network with Heisenberg charge ordering or itinerant magnetism. 1 spins coupled by an antiferromagnetic interaction. The RVB The original features of the ground state of the Heisen- state, known also as a ‘‘spin-liquid’’ state because of the berg two-dimensional ͑2D͒ kagome´ network, which is the short-range magnetic correlations and spin fluctuations down central issue of this paper, is already apparent through a clas- ϭ 8–11 to T 0, was also proposed to explain the high-Tc behavior sical description of their magnetism, yielding a dynami- 2 3 of cuprates and, more recently, of the superconducting cal ground state with infinite degeneracy. More precisely, the compound NaxCoO2•yH2O. Although Anderson’s conjec- ‘‘order by disorder’’ mechanism selects a coplanar spin ar- 4 ture was proven to be wrong for the triangular network, rangement in the ground state. However, the system can there is a growing consensus that the RVB state is the ground swap from one coplanar spin configuration to the other by a state of the so-called highly frustrated networks.5,6 Like the rotation of a finite number of spins, with no cost in energy. triangular network, the magnetic frustration of these systems These zero-energy excitation modes, or soft modes, prevent a is exclusively driven by the triangular geometry of their lat- selection of one of the coplanar configurations1͒ and consti- tice provided that spins are coupled through an antiferromag- tute a low-energy reservoir of magnetic excitations. It results netic ͑AFM͒ interaction. This is different from spin glasses, in a finite entropy per spin in the ground state, without, still where frustration arises from the randomness of the magnetic up to now, any definitive conclusions on a possible long- interactions.7 However, compared to a triangular network, range order, even at Tϭ0. these networks have a corner-sharing geometry, as in the In a quantum description of the magnetism for Heisen- two-dimensional kagome´ ͑corner sharing triangles, Fig. 1͒, berg spins 1/2 on the kagome´ network, the magnetic ground the three-dimensional pyrochlore ͑corner sharing tetrahedra, state is disordered.11 It is predicted to be RVB-like, built on a Fig. 1͒, and the quasi-two-dimensional kagome´ bilayer ͑cor- macroscopic number of singlet states.13,14 If any, a gap be- ner sharing triangles and tetrahedra, Fig. 2͒ lattices. The tween the ground state and the first triplet state is expected to corner-sharing geometry introduces a ‘‘magnetic flexibility,’’ be fairly small, of the order of J/20,15 where J is the AFM thereby enhancing the frustration of these networks ͑thus the interaction between nearest neighbor spins.2͒ Most strikingly term ‘‘highly’’͒ and destabilizing the Ne´el order. A highly and similar to the classical case, the corner-sharing geometry frustrated network, and the experimental counterpart known and the half-integer spins lead to an exponential density of

1063-777X/2005/31(8–9)/18/$26.00704 © 2005 American Institute of Physics Low Temp. Phys. 31 (8–9), August–September 2005 Bono et al. 705

further than nearest neighbor yield long-range magnetic or- ͑ ͒ der at 1 K in the pyrochlore compound Gd2Ti2O7 Ref. 29 . In Y2Mo2O7 , lattice distortions have been proposed to re- lieve the frustration in the ground state.30 However, some ͑ ͒ samples, like the pyrochlore Tb2ϪpYpTi2O7 Refs. 31, 32 or ͑ ͒ the spinel ZnCr2O4 Ref. 33 display spin dynamics down to very low temperature. Anyway, the high spin values and/or the 3D character of the frustrated geometry of these com- pounds remain far from the ideal Heisenberg kagome´ case we are dealing with. Along with the spin-liquid ground state FIG. 1. The kagome´ ͑left͒ and the pyrochlore ͑right͒12 corner-sharing lat- tices. The coordinance of each site is 4 and 6, respectively. of the ideal systems, the plethoric number of possible ground states in the presence of disorder ͑including the case of fer- romagnetic interactions in the pyrochlore lattice, yielding the 34,35 low-lying singlet excitations,16 with energies smaller than the frustrated so-called ‘‘’’ ground state, named so as gap, and to a nonzero entropy per spin at Tϭ0.17 The same it is equivalent to the order observed for hydrogen atoms in conclusions seem to emerge for the AFM pyrochlore ice H2O), constitutes one of the rich aspects of the HFMs. lattice.18 The nature of the magnetic excitations, possibly un- 17,19 confined spinons, is still an open question. B. SCGO and BSZCGO As a major drawback, every additional contribution to ´ the nearest-neighbor AFM interaction, such as next-nearest- Among all HFMs, the chromium-based spin-3/2 kagome ͓ ͔ neighbor interactions, dipolar interactions, anisotropy, bilayer compounds SrCr9pGa12Ϫ9pO19 SCGO(p); Ref. 36 ͓ ͔ Dzyaloshinskii-Moriya interactions, lattice distortions or and Ba2Sn2ZnGa10Ϫ7pCr7pO22 BSZCGO(p); Ref. 37 are magnetic defects generated by the presence of magnetic im- the most likely candidates to possess a spin-liquid ground purities or spin vacancies, all of which we designate by ‘‘dis- state. This is not surprising in view of the low spin value, 2D order,’’ may release the frustration and stabilize an original character, and low level of disorder of these compounds ground state.20–25 Experimentally, disorder is present in all compared to other HFMs. They are in fact an experimental ´ HFMs and indeed causes cooperative phase transitions at realization of a quasi-2D kagome bilayer of spin-3/2 chro- ͑ ͒ low temperature, reminiscent of spin glasses or AFM sys- mium ions with antiferromagnetic interactions Fig. 2 . tems. This is typically observed in the jarosites kagome´ fam- Moreover, the AFM interaction between nearest neighboring ily, where the disorder is governed by spin anisotropy and by spins is dominated by a direct overlap exchange resulting in ϳ ͑ ͒ dipolar interactions ͑Ref. 26͒.3͒ In the spin-1/2 kagome´ com- a coupling of J 40 K see Sec. 2 . This coupling is almost ͓ ͔ two orders of magnitude larger than the typical disorder- pound Cu3(titmb)2(OCOCH3)6 •H2O, the competition be- tween first- and second-neighbor interactions, respectively related interactions, like single-ion anisotropy, estimated to 38,39 ϳ ferromagnetic and antiferromagnetic, yields an original 0.08 K, dipolar interactions 0.1 K, or next-nearest- Ͻ ground state, with a double peak in the specific heat and neighbor interactions 1 K. These compounds, however, al- plateaus in the magnetization under particular conditions.27 ways have a small amount of substitutional disorder, since a ϭ ϭ In the Sϭ1 kagome´ staircase Ni V O , the competition be- Cr coverage higher than p 0.95 and p 0.97 cannot be 3 2 8 ͑ ͒ ͑ ͒ tween first and second neighbor interactions, combined to reached in SCGO(p) Ref. 40 and BSZCGO(p) Ref. 41 , respectively. Hence there are always at least some percent of Dzyaloshinskii-Moriya interactions and spin anisotropy, pro- 3ϩ 28 the magnetic Cr ions which are substituted with nonmag- duces a very rich field and temperature phase diagram. In 3ϩ three dimensions ͑3D͒, dipolar couplings and interactions netic Ga ions. These low amounts of spin vacancies, and the related magnetic defects that they produce, turn out to be insufficient to destroy the spin-liquid behavior of the ground state. We will show that local techniques like nuclear mag- netic resonance ͑NMR͒ and muon spin relaxation ͑␮SR͒ are the most suited to probe frustration-related properties in these systems, since they can bypass the magnetic contribu- tion of these defects. Figure 3 presents a simplified chemical structure for the ideal pϭ1 unit cell of SCGO and of BSZCGO, highlighting the chromium kagome´ bilayers.4͒ As can be seen, BSZCGO is a pure kagome´ bilayer, contrary to SCGO, where there are two additional chromium sites, labeled Cr͑c͒, between the bilayers. These pairs of Cr spins are coupled by an AFM exchange constant of 216 K with a weak coupling to the kagome´ bilayers (ϳ1K).42 They form an isolated singlet at FIG. 2. Kagome´ bilayers in SCGO ͑left͒ and in BSZCGO ͑right͒. There are low temperatures, and the resulting susceptibility is negli- two different sites in the kagome´ bilayer lattice, one of coordinance 5, the gible at TՇ50 K compared to the susceptibility of the other of coordinance 6. The magnetism of both compounds arises from the ´ ´ kagome bilayer in a pure compound. Moreover, the Ga/Cr stacking of magnetically decoupled kagome bilayers, ensured in SCGO by ͑ the presence of the nonmagnetic singlets and in BZSCGO by a large inter- substitutions are substoechiometric on these sites see Sec. bilayer distance. 2.3.3͒. The full crystal structure is obtained by the stacking 706 Low Temp. Phys. 31 (8–9), August–September 2005 Bono et al.

FIG. 5. Collection of spin-glass-like transition temperatures Tg for SCGO and for BSZCGO versus Cr concentration p ͑our data and from Refs. 37, 38, 50, 51͒. The line is a guide to the eye.

␹ ␹Ϫ1 macro in these compounds. The linearity of macro at high ␹Ϫ1 ϭ temperature and the extrapolation of this line to macro 0to a negative temperature is typical of AFM interactions and gives an order of magnitude of the Curie-Weiss temperature ␪ ϭ ϩ ͑ CW zS(S 1)J/3 600 K and 350 K for SCGO and BSZCGO͒. However, contrary to the case of ‘‘conventional’’ 43 ␹ antiferromagnets, no kink is evidenced in macro around ␪ CW . This is actually one of the common signatures of the frustration in HFMs:5 the magnetic correlation length cannot increase because of frustration, and mean field theory re- FIG. 3. Cr3ϩ network of SCGO and BSZCGO with their oxygen ͑᭺͒ envi- ϳ␪ ϩ mains valid for T CW . At low temperature, a spin-glass- ronments, along with the two Ga3 sites. The dotted lines represent some of the Ga–O–Cr hyperfine coupling paths. like transition occurs in both systems, around a freezing tem- ϳ perature Tg 3 K in SCGO(p) and 1.5 K in BSZCGO(p) ͑Fig. 5͒. of the unit cells presented in Fig. 3. The magnetism of both However, the spin-glass transition is unconventional. In compounds arises by magnetically decoupled kagome´ bilay- fact, in conventional 3D spin glasses the nonlinear suscepti- ers, ensured in SCGO by the presence of the nonmagnetic bility diverges at Tg , the specific heat is proportional to T at singlets at low temperature ͑only weakly affected by the low temperature, and Tg is proportional to the number of ͑ ͒ Ga/Cr substitution, see Sec. 2.3.3͒, and in BSZCGO by the defects. In kagome´ bilayers, however, i the nonlinear sus- ͑ ͒ large inter-bilayer distance of 9.4 Å ͑6.4 Å in SCGO͒. These ceptibility diverges at Tg Ref. 44 , but Tg increases as the ͑ ͒ compounds can therefore be considered as ideal two- number of spin vacancies decreases Fig. 5 , indicating that dimensional systems. the freezing is related to an intrinsic property of the frus- ͑ ͒ 2 Figure 4 shows the typical macroscopic susceptibility trated bilayer. ii The specific heat is proportional to T ,as in AFM 2D long-range ordered systems. Its unusual large value at low temperature unveils a high density of low-lying excitations,37,44 and its insensitivity to an external magnetic field suggests a large contribution from singlet excitations, which is one of the most striking features predicted for the Sϭ1/2 kagome´ lattice.45–47 ͑iii͒ Magnetic fluctuations and short-range correlations, consistent with a 2D magnetic net- work, are encountered at low temperature.48,49 The kagome´ bilayers therefore combine properties of 3D spin glasses, 2D AFM ordering, 2D fluctuating magnetic states and original properties expected in Sϭ1/2 kagome´ systems. Given these two temperature scales of the macroscopic ␪ susceptibility, CW and Tg , Ramı´rez proposed the definition ϭ␪ Ͼ of a ‘‘frustration ratio,’’ f CW /Tg , with f 10 in all HFMs.52 The kagome´ bilayers SCGO and BSZCGO, respec- tively, display f ϳ150 and 230,41 the largest ratios reported ␹ ͑ ͒ FIG. 4. Macroscopic static susceptibility macro(T) of BSZCGO 0.97 mea- ͑ ͒ ␹Ϫ1 ͑ ͒ so far in compounds where the frustration is driven by corner sured with a SQUID under 100 G left scale and macro(T) right scale . 5 Inset: ac susceptibility ͑78 Hz͒ below 2 K, focusing on the spin-glass-like sharing equilateral triangles. The large frustration, the transition (Tg). Heisenberg spins along with the low disorder, make them the Low Temp. Phys. 31 (8–9), August–September 2005 Bono et al. 707

archetypes of HFM and maybe the best candidates for a RVB TABLE I. Quadrupole parameters for the 71Ga sites in SCGO40,53 and in 41 69␯ ϭ 69 71 71␯ Ϸ 71␯ spin-liquid ground state. BSZCGO. Notice that Q ( Q/ Q) Q 1.59 Q . Quadrupole ef- fects are therefore more pronounced for the 69Ga isotope than for the 71Ga. We present a review of the magnetic properties of SCGO(p) and BSZCGO(p) powder samples, which covers a large range of Cr concentration p. A comparison between samples of different concentration allows us to identify the frustrated and disorder-related magnetic properties of the kagome´ bilayer. Through local probes it is then possible to determine the true nature of the bilayer’s static susceptibility ͑by NMR͒, as well as the bilayer’s low-temperature dynam- ics and ground-state magnetic excitations ͑by ␮SR͒. The out- line of the paper is as follows. Section 2 is dedicated to the *Determined using point charge simulation. susceptibility of the kagome´ bilayer. It is shown that the susceptibility of the bilayer of SCGO and of BSZCGO can be accessed directly through gallium NMR 40,41,53–56 Before going any further, we briefly recall some relation- experiments. Both susceptibilities exhibit a maxi- ship between the contribution of each gallium nucleus to the ϳ ϳ mum in temperature at T 40 K J and decrease below this Ga͑1͒ NMR line and its local magnetic environment in order maximum down to at least J/4. This behavior indicates that to underline what can be exactly probed through Ga͑1͒ short range magnetic correlations persist at least down to 10 NMR. We suppose first that the susceptibility in the kagome´ K, being consistent with the existence of a small spin gap. bilayer varies from chromium site to chromium site and label ´ The comparison of the kagome bilayer’s susceptibility with it, in a generic manner, by ␹.AGa͑1͒ at site i will contribute ␹ macro shows that the spin vacancies of the bilayer generate to the NMR spectrum at a position depending upon the num- paramagnetic defects responsible for the low-temperature ber of the nearest neighbors ͑NN͒ occupied chromium sites Curie upturn observed in the macroscopic susceptibility and their susceptibility ␹. This corresponds to the shift Ki in ͑ both systems have also extra Curie contributions, coming the NMR spectrum for a gallium at site i: from broken Cr͑c͒ spin pairs in SCGO and from bond length distribution in BSZCGO͒. Section 3 is dedicated to the ␮SR Kiϭ ͚ A␹, ͑1͒ study of the low-temperature spin dynamics of these occupied NN Cr͑a,b͒ 49,56–58 HFMs. It is shown that magnetic excitations persist where A is the hyperfine constant ͑the chemical shift is ne- ͑ down to at least 30 mK the lowest temperature that could be glected here͒. The average shift, K, of the NMR line, is accessed͒. This temperature sets an upper limit for the value simply related to the average Ki over all the gallium sites. of a spin gap. A qualitative and quantitative analysis of the Hence, it is practically proportional to the average suscepti- data shows that these excitations are possibly coherent un- bility ␹ of the kagome´ bilayer. On the other hand, the confined spinons of a RVB ground state. The energy scale T kag g distribution of Ki around K defines the magnetic width of the would then correspond, in this phenomenological approach, Ga NMR line and reflects the existence of a distribution of ␹. to the signature of this coherent resonating state. A summary As we show in the following sections, the NMR width and concluding remarks can be found in Sec. 4. probes a susceptibility related to disorder in BSZCGO and in SCGO. II. FRUSTRATED VERSUS DISORDER-RELATED Along with this hyperfine interaction, the 69,71Ga Hamil- SUSCEPTIBILITY tonian in SCGO and in BSZCGO has a quadrupolar contri- A. Gallium NMR in SCGO and BSZCGO bution. The quadrupole interaction of gallium nuclei in both The NMR experiments were performed on 69Ga SCGO and BSZCGO is a consequence of the coupling of the Ϫ ͑ ͒ (69␥/2␲ϭ10.219 MHz/T, 69Qϭ0.178ϫ10 24 cm2) and nucleus to the electric field gradient EFG produced by the 71Ga (71␥/2␲ϭ12.983 MHz/T, 71Qϭ0.112ϫ10Ϫ24 cm2) surrounding electronic charges. Following the usual nota- nuclei in SCGO and BSZCGO using a ␲/2– ␶ – ␲ spin echo tions, the nuclear Hamiltonian may be expressed as ␥ Q ␯ sequence, where and are, respectively, the gyromagnetic h Q HϭϪ␥បI͑¯1ϩK¯ ͒H ϩ ͓3I2ϪI2ϩ␩͑I2ϪI2͔͒, ͑2͒ ratio and the quadrupolar moment of the nuclei. The gallium 0 6 z x y ions are located on two distinct crystallographic sites, labeled ␯ Ga͑1͒ and Ga͑2͒ in Fig. 3. As pointed out in Refs. 40, 41, 53, where H0 is the applied magnetic field, Q is the quadrupolar 56, the interest in Ga NMR lies in the coupling between frequency, and 0р␩р1 is the asymmetry parameter. The gallium nuclei with their neighboring magnetic Cr3ϩ ions principal axes of the shift tensor K¯ are collinear with the through a Ga-O-Cr hyperfine bridge ͑Fig. 3͒. In particular, direction of the nuclear spin operators Ix , Iy , and Iz . the 69,71Ga(1) nuclei are exclusively coupled to the Cr3ϩ As shown in Fig. 3, the gallium sites of both HFMs have ions of the kagome´ bilayer: to nine chromium ions of the different crystallographic environments. Some simple argu- upper and lower kagome´ layers ͓labeled Cr͑b͒ in Fig. 3͔, and ments can be used to anticipate differences concerning their to three chromium ions in the linking site in between the two spectra ͑Table I͒: kagome´ layers ͓labeled Cr͑a͒ in Fig. 3͔. Through the NMR of ͑i͒ The Ga͑1͒ sites of both HFMs have a nearly tetrahe- Ga͑1͒, it is possible to probe locally the magnetic properties dral oxygen environment. Given this nearly cubic symmetry, ␯ of the kagome´ bilayers of both HFMs. This is the central the EFG is small, i.e., Q is small. However, whereas the topic of the NMR study presented in this paper. Ga͑1͒ site of SCGO is occupied only by Ga3ϩ ions, the 708 Low Temp. Phys. 31 (8–9), August–September 2005 Bono et al.

Ga͑1͒ site of BSZCGO is randomly occupied either by Zn2ϩ or by Ga3ϩ ions. As a consequence, point charge simulations of the EFG on the Ga͑1͒ site of BSZCGO show that there is ␩ ␯ ͑ ͒ a large distribution of and Q , contrary to the Ga 1 of SCGO. ͑ii͒ The Ga͑2͒ site of SCGO is surrounded by a bi- pyramid of 5 oxygen ions, whereas the Ga͑2͒ of BSZCGO lies in an oxygen-elongated tetrahedron. As both environ- ments lack cubic symmetry, the quadrupolar frequency is ␯ large. For this site, point charge simulations yield a Q that is twice as large in SCGO as in BSZCGO. ͑iii͒ The ratio Ga͑1͒:Ga͑2͒ is 2:1 in SCGO, whereas it is 1:2 in BSZCGO. Consequently, the relative spectral weight of the Ga͑1͒ and Ga͑2͒ sites will be opposite in the two compounds. An extensive study of the 69,71Ga NMR spectra in SCGO was presented in Refs. 40 and 53. It allowed identification of the NMR lines and extraction of the parameters of the nuclear Hamiltonian, among them the quadrupolar param- eters given in Table I. It was shown that the Ga NMR spec- trum of both isotopes is the sum of the Ga͑1͒ and Ga͑2͒ sites, plus a small extra contribution related to the presence of the FIG. 6. Comparison of the 69,71Ga NMR spectra in SCGO and BSZCGO ␯ Ϸ nonstoichiometric gallium substituted on the chromium sites obtained by sweeping the field with a constant frequency l 40 MHz, at ϭ ͑ ͒ ͓labeled Ga͑sub͒ in Fig. 8͔. A similar analysis allowed iden- T 40 K. The arrows point the quadrupolar singularities for the Ga 1 and Ga͑2͒, respectively, located inside and outside the bilayers ͑Fig. 3͒. ‘‘CL’’ tification of the NMR lines in BSZCGO and determination of indicates the central line (1/2↔Ϫ1/2 transition͒ singularities. The continu- 41 the related parameters of the nuclear Hamiltonian, also ous and dashed lines show the shift of the in-range singularities of the Ga listed in Table I. sites, from one system to the other. We cannot identify the first order satel- ͑ ͒ Aside from the Ga-substituted sites, the Ga NMR spec- lites of the Ga 1 in BSZCGO on this spectrum. trum displays four sets of lines corresponding to the two ͑ ͒ ͑ ͒ isotopes distributed on both Ga 1 and Ga 2 sites. In a pow- it is advantageous to work at a frequency higher than 40 der sample, as used here, for a given site and a given isotope, MHz, in order to isolate the Ga͑1͒ contribution.5͒ Field- the line shape results from the distribution of the angles be- sweep 71Ga NMR experiments were performed at an rf fre- tween the magnetic field and the EFG principal axes. This ␯ Ϸ ͑ ͒ quency l 84 MHz. At high temperature Fig. 9a a ͑ ͒ yields singularities rather than well defined peaks the so- 71Ga(1) shift is evidenced, similar to the one in SCGO.6 At ͒ called powder line shape , as in the Ga NMR spectra of low temperature, the 71Ga(1) and 71Ga(2) lines broaden and ͑ ͒ ͑ ͒ SCGO 0.95 and of BSZCGO 0.97 presented in Fig. 6. start to overlap. However, the two lines can be resolved by Powder-pattern simulations for two different rf frequen- exploiting the different transverse relaxation times T2 of the cies (␯ ϭ40 MHz and 84 MHz͒ are presented in Fig. 7, us- ϳ ␮ ϳ ␮ l two gallium sites (T2,Ga(1) 20 s, T2,Ga(2) 200 s). The ing the parameters of Table I. They agree perfectly with the first step consists in using a large time separation between experimental NMR spectra and clearly show that the the two rf pulses (␶ϭ200 ␮s) in order to eliminate the Ga͑1͒ BSZCGO spectrum is more intricate than the SCGO one, because of the more pronounced quadrupolar contribution. This is probably the reason why the Ga͑sub͒ line cannot be resolved in BSZCGO, contrary to SCGO. Most of the SCGO data were acquired in a field-sweep ␯ Ϸ spectrometer with an rf frequency l 40 MHz, because then the 69,71Ga(1) NMR lines can be isolated, as shown in Fig. 7.40 At high temperatures ͑Fig. 8a͒ the NMR line shape remains unaltered and shifts with decreasing temperature. At low temperature (TϽ50 K), the line broadens with decreas- ing temperature ͑Fig. 8b͒. The 69,71Ga(1) shift is still quali- tatively visible down to 5 K, where the Ga NMR signal is lost ͑see below͒. Fits of the low-temperature spectra ͑using a convolution of a high-temperature unbroadened spectrum with a Gaussian͒ allow one to extract quantitatively the shift and the width of the NMR lines for all the samples studied FIG. 7. Quadrupolar powder pattern simulations for both sites and isotopes ␯ ϭ ͑ ͒ (0.72рpр0.95). All display similar shifts with temperature, with the parameters of Table I, for frequencies l 40 MHz left and 84 ͑ ͒ but different broadening, the line shape increasing with MHz right in SCGO and BSZCGO. The area of each contribution is ob- 40 tained considering the stoichiometry of each site in the samples and the Ga/Cr substitution at a given temperature. natural abundance of each isotope. The full NMR spectra are obtained by In the case of BSZCGO, it can be seen from Fig. 7 that adding all the contributions ͑not shown͒. Low Temp. Phys. 31 (8–9), August–September 2005 Bono et al. 709

FIG. 9. 71Ga spectra in BSZCGO. a—High temperature. The 71Ga(2) first- 71 ϭ ␲␯ 71␥ FIG. 8. Ga spectra in SCGO (Hl 2 l / ). a—High temperature. The order quadrupolar contribution appears as a flat background in this field 71 Ga(2) contribution appears as a flat background in this field range. range at the 71Ga(1) line position. b—Low temperature, lines are broad- 71 b—Low temperature, lines are broadened. The Ga(2) contribution re- ened. ᭺ and ᮀ are the short and the rescaled long spectra, respectively. ᭝ is 71 mains flat at the position of the Ga(1) line, which shifts to higher fields in for the Ga͑1͒ contribution, given by their subtraction. The dotted arrows this temperature region when the temperature decreases. point at the center of the Ga͑1͒ line, which shifts to higher fields when T decreases. Continuous ͑dashed͒ lines are Gaussian broadened 71Ga(1) 71 ␯ ϭ ͑ ͓ Ga(2)͔ quadrupolar powder pattern simulations, with Q 3.5 MHz 12 MHz͒ and ␩ϭ0.6 (␩ϭ0.04). contribution. The isolated Ga͑2͒ line is extracted using a Gaussian convolution of the quadrupolar powder pattern. In the second step, a short time separation is employed between there is no evidence for such a recovery in SCGO, nor in ␶Ϸ ␮ ͑ ͒ the pulses ( 10 s) so that the Ga 1 contribution is re- BSZCGO. covered. The Ga͑1͒ line is finally isolated by subtracting the long-␶ spectrum corrected in intensity to account for the ͒ ´ shorter time separation employed.7 Such a procedure can be B. Susceptibility of the kagome bilayer ͑ ͒ ␹ employed down to 10 K. The width and the shift of the Ga 1 The kagome´ bilayer susceptibility kag is probed through 71␯ 69,71 40,41,54 line can then be extracted from a fit using a Q the shift K of the Ga(1) NMR lines, presented in ϭ3.5 MHz quadrupolar powder pattern convoluted by a Gaussian line shape, and are little affected by the uncertainty 71␯ ͑ ͒ on Q . The ratio between the intensities of the Ga 2 and Ga͑1͒ lines ͑integrated area below the lines͒ was found to stay close to 2 down to 10 K, in agreement with the 2:1 stoichiometric ratio for the Ga͑2͒ and Ga͑1͒ sites of BSZCGO. The temperature dependence of the intensity of the 71Ga(1) NMR line, which corresponds to the number of de- tected 71Ga(1) nuclei, is presented in Fig. 10 for SCGO. A similar trend is observed in BSZCGO, with larger error bars due to the preponderant Ga͑2͒ contribution. As temperature decreases the spin dynamics slows down, and results in a ͑ ͒ decreasing longitudinal relaxation time T1 see Sec. 3 , hence in a progressive loss of the NMR signal below 15 K. Such a wipeout is reminiscent of the one observed in spin 59 glasses at temperatures near Tg . However, contrary to spin FIG. 10. NMR signal intensity in several SCGO(p) samples. A loss of 69,71 Շ ϳ glasses, where ultimately the signal is recovered below Tg , Ga sites is observed for T 5Tg ( 16 K). The line is a guide to the eye. 710 Low Temp. Phys. 31 (8–9), August–September 2005 Bono et al.

͑ ͒ BSZCGO 0.97 . The ‘‘Curie–Weiss’’ constant CNMR found from the fit is 20% larger in SCGO than in BSZCGO, which indicates that hyperfine couplings are stronger in SCGO, likely because of the shorter Ga–O–Cr bonds ͑Fig. 3͒. Al- though very commonly employed to extract information from the HFMs susceptibility, this phenomenological law is a rather crude approximation, since the linear behavior of the inverse of the susceptibility is expected to hold, within a տ ␪ mean field approach, only when T 2 NMR , which is not the ␪ case here. From NMR , one can therefore only grossly esti- mate the coupling constants J of the kagome´ bilayers. In the absence of any prediction for Sϭ3/2, we used the high tem- ␹ ϭ perature series expansion of kag , derived for an S 1/2 ␪ 60 kagome´ lattice, to correct NMR by a factor of 1.5. The FIG. 11. Shift of the 71Ga(1) line in SCGO and BSZCGO. Lines are GCC coupling constant is finally extracted from the mean-field ͑ ͒ calculations see text . Inset: Full scale down to 0. The line is the 36 spin relation ␪ ϭ1.5zS(Sϩ1)J/3. The coordinance is z Sϭ1/2 cluster calculation mentioned in the text (␹ ). NMR ED ϭ5.14 and corresponds to the average number of nearest neighbors for a chromium ion of the bilayer. The couplings Fig. 11 for the purest samples of both systems in a tempera- then read Jϭ45 K and 40 K for SCGO and BSZCGO, re- ture range displaying no significant loss of intensity (T spectively. These values are consistent with the couplings տ ͑ ␹ 61,62 10 K). The shift and kag) increases when the tempera- observed in other chromium-based compounds, and they ture decreases to 45 K, where a maximum is reached, and indicate that the Cr–Cr AFM interaction stems from the di- then decreases again as the temperature is lowered. This be- rect overlap between orbitals of neighboring chromium ions, havior is common to both systems and is field- and rather than from an oxygen-mediated superexchange cou- dilution-independent.54 Most importantly, it is quite clear pling. A direct interaction results in a coupling J very sensi- from Fig. 12 that there is a discrepancy between the low- tive to the Cr–Cr distance ͑d͒, and follows a phenomenologi- ␹ ␹ 40 temperature behavior of kag and of macro . Instead of a cal law: maximum, the susceptibility probed by SQUID measure- ments follows a Curie-like law at low temperature. This es- ␦J/␦dϭ450 K/Å. ͑4͒ ␹ 8͒ tablishes that macro is a two-component susceptibility, re- sulting from the sum of the kagome´ bilayer’s susceptibility The slightly stronger coupling found in SCGO compared to and, as evidenced in Sec. 2.3, of a susceptibility related to BSZCGO results then, following this viewpoint, from the ´ ͑ ͒ magnetic defects (␹ ): shorter Cr–Cr bond lengths of its kagome bilayer Fig. 3 . def ␹ We now turn to the low-temperature behavior of kag , ␹ ϭ␹ ϩ␹ ͑ ͒ macro kag def . 3 and in particular we discuss the maximum around 45 K. The Concerning the high-temperature (Tу80 K) behavior of first possible interpretation of this maximum is the existence ϭ of a spin gap. Such a gap ⌬ is an important issue for the K, a phenomenological Curie–Weiss law, K CNMR /(T ϩ␪ ␪ ϭ Ϯ kagome´ bilayer, since it would be the signature of the exis- NMR), is an accurate fit, yielding NMR (440 5) K and (380Ϯ10) K, respectively, for SCGO͑0.95͒ and tence of a singlet ground state, found in the theoretical quan- tum description of an Sϭ1/2 kagome´ layer.15 Its value is predicted to be ⌬ϷJ/20, and should be of the same order for higher spins.9͒ Since Ga NMR cannot access temperatures lower than 10 K, which for SCGO and BSZCGO corre- sponds to J/4, we cannot conclude whether the maximum in temperature is related to a gap. From an experimental point of view, its observation would require temperatures TϽ⌬ in order to observe the exponential decrease predicted for the susceptibility.47 In the inset of Fig. 11, we compare our ex- perimental data to the susceptibility computed with exact diagonalization on a 36 spin Sϭ1/2 kagome´ cluster.19 We see that the position of the predicted maximum does not match the experimental one. Also, the sharp peak of the suscepti- ␹ bility of the calculation is not seen in kag . Clearly, larger cluster sizes and the kagome´ bilayer geometry are needed for a better comparison to the NMR data. For the time being, it may be safely concluded, that the behavior of the NMR shift, ␹ hence of kag , is consistent only with a spin gap smaller than J/10. ͑ ͒ ␹ ͑ ͒ ͑ ͒ FIG. 12. K left and macro right for SCGO 0.95 . The NMR shift and the Rather than the signature of a gap, the maximum in ␹ SQUID susceptibility do not follow exactly the same law at high tempera- kag ␹ ͑ ͒ ͑ ͒ was assigned in Ref. 54 to the signature of a moderate in- ture, since macro also probes the susceptibility of the Cr c –Cr c spin pairs of SCGO ͑a detailed analysis can be found in Fig. 13 of Ref. 40͒. crease of the spin-spin correlations of the kagome´ bilayer Low Temp. Phys. 31 (8–9), August–September 2005 Bono et al. 711

which, however, must remain short ranged given the absence of any phase transition. This conclusion was confirmed by neutron measurements on SCGO for temperatures ranging from 200 K down to 1.5 K,63 and by susceptibility calcula- tions performed on the kagome´ and the pyrochlore lattices with Heisenberg spins, in which the spin-spin correlation length is kept of the order of the lattice parameter.64 These calculations, using the so-called ‘‘generalized constant cou- pling’’ method ͑GCC͒,64 consist in computing the suscepti- bility of isolated spin-clusters ͑triangles for the kagome´ and tetrahedra for the pyrochlore͒ and in coupling these clusters following a mean field approach. In the case of the pyro- chlore lattice, an exponential decrease of the susceptibility is obtained at low temperature. In this case, the ground state of a cluster is non-magnetic and the spin gap is between the S ϭ0 ground state and the magnetic Sϭ1 excited states. In the ´ case of the kagome lattice, the ground state of the triangle is FIG. 13. Ga͑1͒ NMR width versus temperature for SCGO samples of dif- ϭ magnetic (Stotal 1/2) which gives rise to the nonphysical ferent Cr concentration. The width ⌬H is normalized by the reference field ϭ␯ 69,71␥ ͑ divergence in the susceptibility as T→0, due to the choice of Hl l / to superimpose the results from the two gallium isotopes as a particular cluster.64 The GCC simulations for the Sϭ3/2 explicitly shown for pϭ0.95). The solid lines are ϰ1/T Curie-like fits. kagome´ and pyrochlore lattices, in between which the kagome´ bilayer’s susceptibility is expected to lie, are pre- lattice dilution, similarly to the low-temperature behavior of sented in Fig. 11. As shown, they agree with the data down to ␹ ͑see Fig. 16͒, but differs from the NMR shift. The Tϭ40 KӶ␪ , yielding a maximum and a behavior macro NMR perfect scaling of the widths of the two 69,71Ga(1) isotopes around the maximum in agreement with the experimental ͑e.g., SCGO͑0.95͒ in Fig. 13͒ underlines the magnetic origin findings. From the simulations, the values of ␪ are 260 K CW of the low-temperature broadening. for SCGO͑0.95͒ and 235 K for BSZCGO͑0.97͒, which cor- Figure 14 presents the linewidth obtained for respond to Jϭ40 K and Jϭ37 K, close to the previous val- SCGO͑0.95͒ and for a comparable dilution in ues obtained through the high-temperature Curie–Weiss law. BSZCGO͑0.97͒. Surprisingly, the Curie upturn in In conclusion, the GCC computation fits quite well the BSZCGO͑0.97͒ is four times larger than in SCGO͑0.95͒. data down to TϳJӶ␪ , where JϷ40 K. It only shows CW This ratio cannot be explained by a higher hyperfine cou- that the strongest underlying assumption of this cluster pling constant for BSZCGO, since, as mentioned in Sec. 2.2, mean-field approach is relevant, i.e., the spin-spin correlation the constants C obtained from the high-temperature length is of the order of the lattice parameter, which prevents NMR Curie-Weiss analysis of the shift in SCGO and BSZCGO any magnetic transition to a long range ordered state. point to the opposite variation. Further insight concerning this mismatch between the two HFMs can be gained by plot- C. Defect-related susceptibility ting the widths as a function of dilution at a given tempera- ͑ ͒ In this section, we focus on the defects contribution to ture inset of Fig. 14 . In SCGO the width extrapolates to 0 ϭ the susceptibility. The NMR linewidth and the SQUID data when p 1, which shows that the width is only related to at low temperature show that the paramagnetic susceptibility dilution. On the contrary, in BSZCGO the width reaches an ␹ ␹ def observed in macro comes from the dilution of the kagome´ bilayer and from other kinds of defects, namely the Cr͑c͒ pairs in SCGO and bond defects in BSZCGO. The Ga͑1͒ NMR width not only enables to evidence that the va- cancy of a spin on the network, i.e. the dilution, generates a paramagnetic defect, but also allows to shed light on the more fundamental question concerning the extended nature of the defects. We first present the experimental facts con- cerning the magnetic defects, as measured quantitatively through the NMR linewidth and the SQUID susceptibility ␹ macro and then elaborate on their nature in both compounds.

1. NMR linewidth The Ga͑1͒ NMR linewidth of SCGO and of BSZCGO was measured for various Cr-concentrations p. Figure 13 ⌬ presents the typical p-dependence of the linewidth in both FIG. 14. Magnetic contribution to the NMR linewidth H/Hl .For samples—in the figure, only for SCGO. In agreement with BSZCGO͑0.97͒, the data for 71Ga(2) ͑ᮀ͒ are rescaled by a factor of 6, corresponding to the ratio of the coupling constants deduced from the high- the qualitative presentation of the raw spectra in Sec. 2.1, the ᭹ ⌬ 71 ϰ temperature shifts. are for H/Hl͓ Ga(1)͔. The lines are 1/T Curie- width increases as temperature drops and is well described like fits. Inset: p dependence of ⌬H71͓ 71Ga(1)͔ at 50 K for SCGO(p)and by 1/T Curie law ͑lines in Fig. 13͒. It is very sensitive to the BSZCGO(p). The lines are a guide to the eye. 712 Low Temp. Phys. 31 (8–9), August–September 2005 Bono et al.

71 ͑ ͒ ␯ Ϸ FIG. 15. Ga(1) spectrum at 150 K in SCGO 0.95 ( l 40 MHz). The lines are quadrupolar powder pattern simulations with the parameters of Table I, shifted with a value proportional to zЈ, the number of NN Cr3ϩ for ͑ ͒ a site. Their area is weighted by the probability P(zЈ) when pϭ0.95, pre- FIG. 16. Low-temperature macroscopic susceptibility SQUID for some ͑ ͒ ͑ ͒ sented in the inset. SCGO Cr concentrations. Inset: SCGO 0.95 data circles fitted with Eq. ͑5͒.

Hence, the explanation for the low-temperature broadening р Ͻ asymptotic nonzero value for 0.86 p 1. As a first assump- must be searched elsewhere ͑see Sec. 2.3.3͒. tion, one could wonder whether the dilution of the lattice is larger than the nominal one. This can be ruled out since ͑i͒ 2. SQUID the expected evolution of the line shape with p is found at 300 K ͑see the following͒͑ii͒ muon spin relaxation measure- The SQUID measurements were carried out over a wide ments indicate a regular evolution of the dynamical proper- range of Cr concentrations (0.29рpр0.97) for temperatures ties with p ͑Sec. 3͒. Therefore, the increased low-temperature down to 1.8 K ͑no difference was observed between the field upturn in BSZCGO comes from dilution-independent para- cooled ͑FC͒ and zero-field cooled ͑ZFC͒ susceptibility above ͒ magnetic defects, which are not present in SCGO. the freezing temperature Tg in a field of 100 G . Figure 16 We may already rule out possible scenarios which could presents the typical low-temperature macroscopic suscepti- be responsible for the low-temperature broadening. For in- bility of these HFMs. Like the NMR linewidth, the suscep- stance, the broadening generated by the suppression of chro- tibility exhibits a low-temperature upturn, increasing with mium ions in the nuclear environment. Such a distribution is growing dilution. In Sec. 2.2, by a comparison between the present in the Ga͑1͒ NMR spectra of SCGO and of NMR shift and the macroscopic susceptibility, we estab- ␹ BSZCGO, but is only responsible for a minor broadening of lished that macro must be a two-component susceptibility, the line as shown in Fig. 15 for SCGO͑0.95͒. This broaden- and, in particular, that it should possess, compared to the ␹ ing mechanism results from the fact that each Ga͑1͒ nucleus shift, a paramagnetic susceptibility def to explain its low- ␹ has a probability P(zЈ) ͑inset of Fig. 15͒ to have zЈ chro- temperature upturn. The dilution-dependent upturn of macro ␹ mium neighbors (0рzЈр12). Assuming that the Ga/Cr sub- in Fig. 16 therefore establishes that def is related to the stitution does not affect the Ga–O–Cr hyperfine couplings Ga/Cr substitution in SCGO(p). However, we also know and that all the chromium ions have the same susceptibility, from NMR that there also exist extra paramagnetic contribu- the shift of a Ga͑1͒ nucleus surrounded by zЈ neighboring tions, related to intrinsic defects in BSZCGO, for example. chromium ions is then proportional to zЈ, according to Eq. In order to quantify the low-temperature dependence of ␹ ͑1͒. To construct the spectrum, each gallium nucleus is asso- the susceptibility, we follow Ref. 65 and fit macro by a two ciated to a quadrupole line simulated with the parameters of component expression: Table I, with a shift reflecting its Cr environment and an C Cdef intensity weighted by the total number of gallium nuclei hav- ␹ ϭ ϩ . ͑5͒ macro Tϩ␪ Tϩ␪ ing the same environment, i.e., weighted by P(zЈ). The def simulated line perfectly matches the experimental Ga͑1͒ The first Curie–Weiss term roughly takes into account the NMR line at 150 K ͑Fig. 15͒. However, the broadening re- contribution from the kagome´ bilayer and governs the behav- ␹ sulting from this distribution scales with the susceptibility, ior of macro at high temperature. The second term, the more hence with the shift K. Since K decreases at low tempera- relevant for this section, also a priori Curie–Weiss, quanti- ␹ ␹ ture, this broadening mechanism cannot explain the low- fies the contribution of def . A more accurate fit of macro temperature upturn of the width. For the same reason, a spa- would consist in using a low-temperature susceptibility mim- tial distribution of the hyperfine constant, which yields also a icking the NMR shift. Such a fit, however, yields the same width ϰ K, cannot justify the broadening observed and also qualitative results as presented below. has to be ruled out.10͒ Even if a such refined analysis is not From these fits ͑e.g., inset of Fig. 16͒, an effective mo- 3ϩ ϭ Ϯ ␮ possible in BSZCGO because of a stronger broadening than ment per Cr is extracted, peff 4.1 0.2/ B in both 56 40 ␮ SCGO, similar conclusions are derived for this compound. SCGO and BSZCGO, close to the 3.87/ B expected for Low Temp. Phys. 31 (8–9), August–September 2005 Bono et al. 713

this compound, on the contrary of SCGO, a fine analysis of the spin-vacancy defects cannot be carried out, even by Ga͑1͒ NMR. A. Spin vacancies. In AFM systems such as the 1D spin chains66,67 the quasi-2D spin ladders68 and the 2D cuprates69 it is now well established that a vacancy ͑or a magnetic im- purity͒ generates a long range oscillating magnetic perturba- tion and creates a paramagnetic-like, i.e., not necessarily ϰ1/T, component in the macroscopic susceptibility. The symmetric broadening of the NMR line observed in these systems is related to the oscillating character of the pertur- bation. The dilution effects of the kagome´ bilayer would then be described by the same physics of these correlated systems. ͑ ͒ FIG. 17. Fits of the SQUID data ͑Eq. ͑5͒͒.a—᭝ come from Ref. 50. The In SCGO, the symmetric feature of the Ga 1 NMR line lines are a guide to the eye. b—The dashed line represents the kagome´ indicates that spin vacancies induce a perturbation extended bilayers defects contribution to Cdef(p) in SCGO(p), the other lines are a in space. This rules out descriptions of the defect in terms of ͑ ͒ 65,70 guide to the eye. The gray area represent the Cr c contribution in Cdef(p) uncorrelated paramagnetic centers, since the broadening ͑ ͒ ͑ ͒ for SCGO(p) see text . Inset: same data as b for a broader range of p. of the line would then be asymmetric. Hence, given the AFM interactions of the kagome´ bilayer, the perturbation was as- signed to a staggered polarization of the network.40 The ex- Sϭ3/2 spins. In BSZCGO, the Curie-Weiss temperature is act diagonalization of 36 spin 1/2 clusters on the kagome´ found to increase with p ͑Fig. 17a͒, in agreement with mean- lattice with spin vacancies by Dommange et al. confirms this field theory. From the linear variation, following the proce- interpretation.23 They show that within an RVB ground state, dure described in Sec. 2.2, we extract Jϭ40 K,41 consistent the spin-spin correlations around a vacancy are enhanced, with our NMR results. However, two linear regimes of ␪(p) and that the magnetization is staggered around it, a picture are observed for SCGO(p), below and above pϳ0.55. This was first reported in Ref. 50 and cannot be understood in that should also hold for higher spin values. However, it terms of simple mean field theory. Taking into account a seems that the image of a simple envelope of the staggered ͑ ͒ magnetization, like the Lorentzian now granted in high-Tc third term in Eq. 5 , corresponding to the susceptibility of 69 the Cr͑c͒ pairs in SCGO, derived in Ref. 40 and vanishingly cuprates , is not correct here since this localization of sin- small below 50 K, does not affect this result. glets around the vacancies shifts the staggered cloud from The second term in Eq. ͑5͒ is introduced to fit the low- the nonmagnetic impurity. Although the agreement is quali- temperature behavior of the defects. For all samples studied, tative at the moment, further work is required within this it is close to a pure Curie law, within error bars. The values framework for a quantitative modeling of the NMR line- shape. of the Curie constants Cdef extracted are presented in Fig. 17b. A p dependence qualitatively similar to the NMR width B. Bond disorder in BSZCGO. We previously saw that in ϭ BSZCGO(p), low-temperature macroscopic susceptibility is found and Cdef extrapolates again to 0 for p 1 in SCGO, whereas it displays a finite value in BSZCGO. When the Cr and NMR width data can be both satisfactorily accounted р for, only if a novel p-independent defect-like contribution is concentration is 0.8, the dependency of Cdef is qualita- tively the same in both systems ͑inset of Fig. 17b͒. taken into consideration, a result somehow surprising in view To give an order of magnitude, this term represents, for of the close similarity between the kagome´ bilayers of SCGO the purest samples, a number of Sϭ1/2 paramagnetic spins and BSZCGO. The only major change lies in the 1:1 random 3ϩ 2ϩ equal to the number of spin vacancies in SCGO(p) ͓40͔ and occupancy of the Ga͑1͒ site by Ga or Zn ions which to 15–20% of the number of Cr3ϩ spins in BSZCGO(p).41 induces different electrostatic interactions with the neighbor- ing ions. As an example, distances to O2Ϫ in a tetrahedral ϭ ϩ Ϫ environment vary from rGa3ϩ-O2Ϫ 0.47 Å to rZn2 -O2 3. Paramagnetic defects: discussion ϭ0.60 Å which is at the origin of a change of the average The NMR and the macroscopic susceptibility measure- Ga͑1͒-O bonds, from 1.871 Å for SCGO42 to 1.925 Å in ϩ ments indicate that there are paramagnetic defects in SCGO BSZCGO.42 Similarly, one expects that the Cr3 will be less ϩ ϩ and BSZCGO. Some of them are related to the spin vacan- repelled by Zn2 than by Ga3 , which certainly induces cies in the kagome´ bilayers and affect the neighboring chro- magnetic bond-disorder, i.e., a modulation of exchange inter- ϩ mium ions in such a way that their overall behavior is para- actions J between neighboring Cr3 . From Eq. ͑4͒, one can magnetic. These defects therefore yield information on how estimate that a modulation as low as 0.01 Å in the ϩ ϩ the correlated spin network of the kagome´ bilayer responds Cr3 -Cr3 distances would yield a 10% modulation of J. to the presence of the vacancy. In the following, we discuss The presence of nonperfect equilateral triangles classi- this point as well as the presence of other types of defects in cally induces a paramagnetic component in the susceptibility both compounds. of the frustrated units.70 The fact that we do not observe any The second type of defects results from lattice imperfec- extra Curie variation as compared to SCGO(p) in the aver- tions in BSZCGO, and from the dilution of the spin-pair sites age susceptibility, probed through the NMR shift K ͑Fig. 11͒, in SCGO. Unfortunately, in the case of BSZCGO the second indicates that such unbalanced exchange interactions only type of defects dominates the paramagnetic response, and in induce a staggered response in the same manner as spin va- 714 Low Temp. Phys. 31 (8–9), August–September 2005 Bono et al. cancies. Whether this could be connected with the localiza- the muon spin systems. However, while the out-of- tion of singlets in the vicinity of defects and a staggered equilibrium state is reached using rf pulses in NMR, the cloud around them, like for spin vacancies, should be more muon spin system is already in an excited state. Indeed, deeply explored. This might highlight the relevance of a muons (␮ϩ, Sϭ1/2) are implanted in the sample, 100% spin RVB approach to the physics of the kagome´ network. Further polarized along the z axis. Therefore, the muon spins always insight into the exact nature of the defects likely requires a depolarize to reach the Boltzmann distribution and ␮SR can better determination of the bond disorder through structural be performed in zero external field, contrary to conventional 71 studies at low T, to avoid the usual thermal fluctuations at NMR. The time dependence of their polarization Pz(t) room T. along the z axis is studied through their decay into a C. Broken spin pairs in SCGO. The Ga/Cr substitution positron72 and is directly linked to both the spin fluctuations onaCr͑c͒ site of SCGO could in principle break a Cr͑c͒- ͓the time correlation function of the local field H␮ is usually ͑ ͒ 2 ϭ Ϫ␯ Cr c spin pair and free a paramagnetic spin, which then exponential, i.e., ͗H␮(0)•H␮(t)͘/͗H␮(0) ͘ exp( t)] and ␹ ͑ ͒ could contribute to the paramagnetic upturn of macro . How- the random local field distribution, characterized by a width ever, this contribution turns out to be small. The fact that the ⌬/␥␮ (␥␮ is the muon gyromagnetic ratio͒. The longitudinal broken spin-pair susceptibility is small, can be deduced by relaxation time is related to the spin-correlation function comparing NMR and SQUID measurements for SCGO and through BSZCGO. The starting point is to notice that the crossing ϱ between the variation of the Curie constants C of both ϳ ͗ ͑ ͒ ͑ ͒͘ ͑␥ ͒ def 1/T1 ͵ S 0 •S t cos ␮HLFt dt. samples determined by the fits of Eq. ͑5͒ to the macroscopic 0 susceptibility ͑Fig. 17b͒, is not observed in the NMR width Zero-field and longitudinal-field ͑LF͒ ␮SR experiments, ͑inset of Fig. 14͒. Considering from our NMR analysis that where the external field HLF is applied along the z axis, there is a 20% difference between the hyperfine coupling, we allow for instance to distinguish magnetic ͑randomly or or- ⌬ 73,74 can evaluate that the scaling factor between H/Hl(p) and dered͒ frozen states from dynamical ones. Moreover, Cdef(p) should be 20% larger in BSZCGO(p) than in whereas specific heat is sensitive to all kinds of excitations, SCGO(p). An error of 10% comes from the possible differ- including low-lying singlets at low temperature in the ence between the bilayer susceptibility measured with NMR kagome´ bilayer samples,46 the muons probe magnetic exci- and SQUID, since the Ga͑1͒ nuclei do not probe stoichio- tations only, with a specific range of frequencies 11͒ metrically the magnetic sites Cr͑b͒ and Cr͑a͒. We first de- (ϳ109 Hz) sitting in a typical time window ͑10 ns–10 ␮s͒ in ͓⌬ ͔ termine the scaling factor H/Hl(p) /Cdef(p)in between NMR and neutron experiments. BSZCGO(p), where 100% of the magnetic ions belong to We present here our ␮SR study of the spin dynamics in ⌬ 56,58 the kagome´ bilayers. With the same ratio, the H/Hl(p) SCGO(p) and BSZCGO(p). We first show that while data for SCGO(p) therefore yield a lower Cdef(p) than the conventional ␮SR polarization functions can be used in the measured one, outside the error bars ͑dashed line in Fig. whole temperature range for the very diluted samples and in 17b͒. This difference is likely due to dilution of the Cr͑c͒ տ the high-temperature regime (T 3Tg) for the purest ones, pairs in SCGO(p). Their contribution, which is not probed they cannot be used satisfactorily in the low-temperature re- with the Ga͑1͒ NMR ͑gray area in Fig. 17b͒, corresponds to gime in the latter case (pտ0.7). Moreover, in this first step, 3ϩ a 0.04p Cr Sϭ3/2 paramagnetic term, whereas a 2p/9 we qualitatively show that the magnetic state of all the proportion would be expected in the case of a stoichiometri- samples is dynamical down to the experimental limit of 30 cally substituted sample. This substoichiometric contribution mK ͑Sec. 3.2͒. We further use a model independent analysis of the Cr͑c͒ site is consistent with of the data, simply taking the time necessary for the muon 40 measurements. spin polarization to decrease down to the value 1/e. The comparison between SCGO and BSZCGO clearly shows a correlation between the muon spin relaxation rate and Tg , III. SPIN DYNAMICS AND MAGNETIC EXCITATIONS suggesting that this freezing temperature is not an impurity A. Muon: an appropriate probe of dynamics phase. On the contrary, it seems to be closely linked to the slowing down of the spin dynamics in the bulk sample ͑Sec. We mentioned that an abrupt loss of the Ga͑1͒-NMR 3.3͒. Finally, a phenomenological model for the muon relax- intensity occurs below 10 K ͑Fig. 10͒. Hence, Ga͑1͒-NMR ation, based on sporadic dynamics due to spin excitations in cannot be employed to probe the low-temperature magnetic a singlet sea, proposed by Uemura et al.,49 is extended to all properties of SCGO and of BSZCGO. This is unfortunate, fields and temperature range. Its connection to the RVB pic- since the low-temperature spin dynamics reveal information ture is discussed, and we argue that such coherent states that, in particular, gives a clue as to the ground state of these might mediate the interactions between ‘‘impurities,’’ which systems. On the contrary, due to the smaller coupling of the induce the spin-glass freezing ͑Sec. 3.4͒. muon to the electronic moments and a shorter time window than the NMR one, the ␮SR technique has proven to be a B. Conventional approaches leading tool for the study of quantum dynamical states in a vast family of fluctuating systems. 1. High-temperature behavior The electronic dynamics is commonly probed, either for At high temperature, a conventional paramagnetic be- NMR or for ␮SR, through the measurements of the longitu- havior is found for all the samples, with a stretched exponen- ϭ Ϫ ␭ ␤ dinal relaxation time, necessary to recover the thermody- tial variation of Pz(t) exp͓ ( t) ͔. Figure 18 shows that as namic equilibrium after the excitation of the nuclear spin or expected, ␤→1 in the dense magnetic cases74 ͑in our case, Low Temp. Phys. 31 (8–9), August–September 2005 Bono et al. 715

FIG. 18. Exponent ␤(p) in BSZCGO(p) using the relaxation function ϭ Ϫ ␭ ␤ Pz(t) exp͓ ( t) ͔ appropriate for the fast fluctuations limit at high tem- perature (Tу10 K). FIG. 19. BSZCGO͑0.97͒ data at 0.03 K. The lines are fits with different models ͑see text͒. We used ⌬Ϸ2 ␮sϪ1,4␮sϪ1 (␯Ϸ17 ␮sϪ1), and ϩ with a high coverage of the kagome´ bilayer lattice by Cr3 ) 65 ␮sϪ1 (␯Ϸ600 ␮sϪ1), for the static, dynamic and ‘‘Uemura’’ cases, re- and ␤→0.5 in the dilute cases73 ͑i.e., with a low coverage͒. spectively. We can give here an estimate of the fluctuation fre- quency in this temperature range and of the NMR time win- and powder samples, the ‘‘1/3 tail’’ is still observed in zero dow which would be required to measure this spin dynamics. field experiments for the same reasons, but oscillations ap- ϳ ␭ϳ ␮ Ϫ1 At 50 K, the muon relaxation rate is 0.03 s and pear in P (t), which correspond to well defined local fields. ␮ Ϫ1 ͑ ͒ ͑ ͒ z 0.01 s in SCGO 0.95 and BSZCGO 0.97 . In this ap- One should notice that nuclear moments which appear propriate fast fluctuating paramagnetic limit, the fluctuation static on the ␮SR time window always contribute to the ␯ϳͱ បϳ 13 Ϫ1 ͑ rate can be estimated from zJS/kB 2•10 s Ref. time-evolution of the polarization. A small longitudinal field ͒ Ϸ 49 using the coupling J 40 K determined previously by of the order of a few 10s of G (⌬ ϳ0.1 ␮sϪ1) is com- ϭ ND NMR, and the average number z 5.14 of Cr nearest neigh- monly applied to ‘‘decouple’’ this contribution. In other ⌬ϭͱ␭␯ ϳ ␮ Ϫ1 ␮ Ϫ1 bors. We extract /2 600 s and 300 s , words, since the electronic and dipolar contributions are mul- which corresponds to an average field at the muon sites of tiplied, such a contribution, if any, ‘‘disappears,’’ and only 7000 G and 3500 G for SCGO and BSZCGO, respectively. the electronic contribution, not perturbed by such low fields, Since the relationship between the longitudinal relax- is measured. Hence, experiments performed at H ϳ100 G ␮ LF ation time T1 measured in NMR and SR is are roughly equivalent to a zero-field measurement without NMR ␮SRϳ͓⌬ ͑ 71␥71 ͔͒2 nuclear dipoles. T1 /T1 / A , Figure 19 shows P (t) measured in BSZCGO͑0.97͒ at 71 ϳ ␮ z where A 70 kOe/ B is the hyperfine coupling constant of ϭ ϭ 71 0.03 K for HLF 50 G and for HLF 5000 G. Neither oscil- the Ga nuclei obtained from NMR experiments, we can lations nor a 1/3 tail are observed when H ϭ50 G. Al- NMR ␮SRϳ LF estimate that T1 /T1 5. This is consistent with our data though not emphasized in the figure, one can note that the NMRϳ ␮ T1 100 s at high temperature and establishes that be- initial polarization is Gaussian, which is even clearer in NMRϳ ␮ low 5 K, the relaxation must be T1 1 s, which is non- SCGO(p).49 In the static case such a shape is observed in measurable as it lies out of the NMR time window. Such a dense magnetic randomly frozen systems, where the polar- small relaxation time is responsible for the wipeout of the ization is given by the Gaussian Kubo-Toyabe ͑KT͒ ͑ ͒ Ga–NMR intensity Fig. 10 . function.73 We fitted our data at short times with this function in zero field, and plot the expected polarization when HLF 2. Failure of conventional approaches at low temperature ϭ500 G in Fig. 19, which is found to be completely decou- for the purest samples pled. This contrasts with our experimental finding, i.e., the The derivation of zero field and longitudinal field ␮SR polarization at 5000 G is hardly decoupled. The lack of a 1/3 relaxation functions is presented in Refs. 73 and 74 for text- tail, the absence of oscillations and of a decoupling effect book cases. We first rule out the possibility of some static show that the magnetic state of BSZCGO(0.97) is fluctuating freezing as the source of the muon relaxation. Two options at 0.03 K. All these considerations remain valid in both can be considered. ͑i͒ In randomly frozen ͑static͒ magnetic SCGO(p) and BSZCGO(p) compounds when pտ0.6. states, the major results are the following: first, the muon The computation of the dynamical relaxation functions spin polarization displays a ‘‘1/3 tail’’ in zero external field, is based on a strong collision approximation.73 Using the у ⌬ → ͑ ͒ i.e., Pz(t 5/ ) 1/3. Indeed, 1/3 of the frozen magnetic corresponding dynamical Kubo-Toyabe DKT function, internal fields are statistically parallel to z and do not con- analytically derived by Keren75 for ␯у⌬, allows us to fit the ͑ ͒ tribute to the muon spin depolarization. Second, Pz(t)is zero field data on the whole time range Fig. 19 . Again, the ‘‘decoupled,’’ i.e., does not relax, for longitudinal fields experimental data is much less sensitive to the applied field տ ⌬ ␥ ͑ ͒ ϭ HLF 10 / ␮ . ii In the case of ordered magnetic states HLF 5000 G than the expected polarization. 716 Low Temp. Phys. 31 (8–9), August–September 2005 Bono et al.

FIG. 21. Comparison of the evolution of the muon spin relaxation rate ␭ versus T in BSZCGO͑0.97͒ and SCGO͑0.95͒,froma1/e analysis. Inset: recovery of a part of the asymmetry at low temperature and long times in ͑ ͒ ͑ ͒ SCGO 0.95 below Tg log-log scale .

teractions are still present. This is the so-called ‘‘cooperative .’’

FIG. 20. a,b—Temperature-dependence of the muon polarization Pz(t)in C. Model-independent basic analysis ͑ ϭ ϭ BSZCGO under weak HLF 10 G for p 0.43 and 100 G for p 0.97). Notice the different time scales for both samples. c,d—HLF dependence at As a first step to a quantitative analysis, we estimate the base temperature. The line for pϭ0.43, ZF, is a square-root exponential muon spin relaxation rate ␭ using the 1/e point of the polar- ⌬ ϳ ␮ Ϫ1 times the Kubo-Toyabe function in zero field ( ND 0.08 s ), account- ization P (t), as discussed in Ref. 57. This allows one to ing for the nuclear dipole contribution. The other lines are fits described in z the text. single out information about two important issues using a simple model-independent analysis: the role of the spin- glass-like transition and the impact of the spin vacancies on This so-called ‘‘undecouplable Gaussian line shape’’ was the dynamics. first reported in Ref. 49 for SCGO͑0.89͒ and further in other From high temperature, ␭ increases by more than two 57,76,77 kagome´ compounds or spin singlet compounds like the orders of magnitude down to T for both kagome´ bilayers at ͑ ͒ g doped Haldane chain Y2ϪxCaxBaNiO5 Ref. 78 . Uemura their highest level of purity, to reach a relaxation rate plateau et al.49 proposed a relaxation model, presented further in this ␭ Շ 49,57 T→0 for T Tg . Using a temperature-scale twice larger paper, which captures some of the facets of this relaxation. for SCGO͑0.95͒ than in BSZCGO͑0.97͒, the temperature de- For now, we just notice that this model cannot justify our pendence of ␭ of both samples perfectly scales on the tem- data for all fields ͑Fig. 19͒, even if it shows a lower decou- perature axis, as shown in Fig. 21.12͒ This ratio is very close pling effect, consistent with our ͑and previous͒ observations. ͑ ͒ to the 2.3 2 ratio between their freezing temperatures Tg , Before going further with more complex models ͑Sec. 3.4͒, which points to a link between the formation of the spin- we will first, in the following, give qualitative arguments to glass-like state and the presence of fluctuations. This seems characterize the low-temperature magnetism of our samples. to be a quite common feature of various systems with a sin- In order to illustrate the evolution of the properties with glet ground state.49,57,76–78 p, we present two typical low and high occupations p of the ␭ In Fig. 22 we report the variation of T→0(T) for various kagome´ bilayers in Fig. 20 which emphasize the qualitative BSZCGO(p) samples and compare them to SCGO(p).49,57 differences for various frustrated network coverage rates p. In addition to the high-temperature behavior already com- mented, the evolution of Pz(t) at low temperature is mark- edly different. For pϭ0.97 the relaxation rate increases by more than two orders of magnitude to reach a temperature- Շ Ϸ ͑ ͒ independent value for T Tg 1.5 K Fig. 20b , with the un- decouplable character presented above. For a low coverage rate, pϭ0.43, we also find a dynamical state but, at variance with the previous case, only a weak temperature-dependence is observed. Also, the polarization displays a square root ex- ponential decay for all temperatures ͑Fig. 20a͒ and for any ͑ ͒ longitudinal field HLF Fig. 20c . A weak plateau of the re- laxation rate is still observed below 0.5 K but is no more ϭ present for p 0.3, where Pz(t) is found to be weakly temperature- and HLF-dependent. This is typical of the pure paramagnetic fast fluctuations limit for dilute magnetic sys- ␭ ͑ ͒ FIG. 22. The p dependence of T→0 for BSZCGO(p) unfilled symbols tems, which contrasts with the results of SQUID measure- and SCGO(p) ͑filled symbols͒. Continuous ͑dashed͒ lines are ϳp3 ments showing that strongly frustrated antiferromagnetic in- ϩ1.9(3)p6 (ϰp3; Ref. 57͒ fits. Low Temp. Phys. 31 (8–9), August–September 2005 Bono et al. 717

Although additional p-independent defects are dominant in BSZCGO(p), as evidenced by SQUID and NMR measure- ments, it is very surprising to find, for the two systems, a very similar quantitative low-temperature relaxation rate ␭ T→0 with increasing Cr concentration over the entire p range studied. Therefore we conclude that only defects re- lated to dilution of the frustrated magnetic network influence the relaxation rate. In a classical framework, a coplanar arrangement with zero-energy excitation modes, involving spins on hexagons, was proposed in the literature to be selected at low temperature.79 Inelastic neutron scattering experiments on SCGO and the spinel compound ZnCr2O4 are indications in 33,48 favor of such excitations. Therefore, one expects the FIG. 23. Schematic representation of the creation and deconfinement of a muon spin relaxation to scale with the number of fully occu- spinon in a RVB ground state on a kagome´ lattice. pied hexagons, ϰ p6. In Fig. 22, the BSZCGO(p) data alto- gether with our SCGO(pу0.89) samples indicate that 6 ␭ → (p) is well accounted for by adding a dominant p term muon site by a paramagnetic neighboring released spin with T 0 ␯ to the p3 term proposed for SCGO(pр0.89) in Ref. 57. In a a fluctuation frequency and a corresponding exponential Ϫ␯ ⌬ case of purely dipolar couplings of the muon with the elec- time correlation function exp( t). is therefore related to tronic magnetic moments, the average of the local dipolar the muon location in the unit cell and its average is assumed ⌬ϰ ͚ 3 ϰ to be p-independent for a given system but is expected to field is given by ͗ i1/ri ͘ p, where the sum is made 3ϩ vary from BSZCGO to SCGO. The Gaussian at early times, over the Cr ions, ri is the distance between the muon and the ith position, and ͗ ͘ is the average over the muon posi- the weakness of the field dependence and a dynamical relax- • ␯ tions. One therefore obtains ␭(p)ϳ⌬(p)ϰp in the case of ation down to 0 are altogether related to the f factor and ϳ⌬ slow fluctuations, and . In addition to this sporadic relaxation, we found that a more conventional Markovian relaxation needs to be intro- ␭͑ ͒ϳ ⌬͑ ͒2␯ ͑␯2ϩ␥2 2 ͒ϰ 2 p 2 p / ␮HLF p duced to fit the long-times tail (tտ3 ␮s) for all fields and in the case of fast fluctuations, i.e., ␭(p)ϰp␩, with ␩р2, temperatures. We therefore write which is not consistent with our data. This indicates that the ͑ ͒ϭ K͑ ⌬ ␯͒ϩ͑ Ϫ ͒ Ϫ␭t ͑ ͒ Pz t xPz t, f , fHLF , f 1 x e , 6 relaxation is not induced by single spin excitations but rather collective excitation processes extending at least on triangles where x represents the weight of the short-time sporadic dy- and/or hexagons. However, these excitations still involve namical function, associated with the spinons dynamics. In only a finite number of spins since the increase of spin va- the following, we first detail how all these parameters can be cancies only affect smoothly the spin dynamics. On the con- reliably deduced from the data and sketch a physical picture trary, a more pronounced effect is observed in the Sϭ1/2 consistent with our results. Ӷ kagome´-like compound volborthite.77 We first present our analysis for T Tg in BSZCGO(p). In order to limit the number of free parameters, we make the minimal assumption that the external field does not influence D. RVB ground state with coherent unconfined spinons the dynamics of the coherent spinon term. The dynamics ͑␯͒ Finding a phenomenological model reproducing the po- and the average dipolar field created by a spinon on a muon ͑⌬͒ larization Pz(t) for all fields, all temperatures, and all dilu- site are shared for all p and HLF . For low fields, x is tions has been for long a pending challenge in the analysis of found close to 1 for the purest samples ͑Fig. 20a,b͒, making these ␮SR experiments in kagome´ frustrated antiferromag- ␭Ј a nonrelevant fitting parameter. It also enables us to de- ␯ ⌬ Ͻ nets. In the quantum framework of a singlet ground state, termine and when HLF 500 G, for various p. The pa- one needs unpaired spins, i.e., isolated magnetic moments, in rameter f is adjusted for each p and its variation accounts for ␭ order to generate magnetic excitations responsible for the the evolution of T→0(p). On the contrary, the high-field muon spin relaxation. Such excitations can be ascribed to data enable us to monitor the evolution of x with HLF and to unconfined spinons,19 for which the location of spin 1/2 var- determine a value for ␭. We could not extend the fits below ies in time with no loss of coherence of the excited state ͑Fig. pϭ0.71 since the weak field dependence prevents an unam- 23͒. Hence, a given muon spin couples to a spin only a short biguous determination of the parameters. fraction ft of the time t after implantation, and one can use a We find perfect fits of our data ͑Fig. 20d͒ with ␯ Ϫ1 Ϫ1 model of ‘‘sporadic’’ field fluctuations to describe Pz(t). ϳ1000 ␮s , ⌬ϳ350 ␮s ϳ␥␮ϫ4 kG and an average This was the initial guess by Uemura et al. to account value of f ϳ0.006. We find a nearly flat p-independent long- Ӷ ␭ ϳ ␮ Ϫ1 Ã for the weakness of the field dependence of Pz(t) at low time relaxation rate for T Tg ( Ј 0.05 s , in Fig. temperature in SCGO͑0.89͒,49 although the spins are S 24d͒. An important finding is that x is of the order of p at ϭ3/2. In the absence of any theoretical predictions, we will low fields ͑Fig. 24a,b͒. This may be related to the theoretical just assume that the mechanism is comparable. This polar- computations showing that the correlations are enhanced ization is given by a ‘‘sporadic’’ dynamical Kubo-Toyabe around spin vacancies in the kagome´ lattice,23 which would K ⌬ ␯ function, Pz ( ft, ,HLF , ), which rewrites simply as destroy locally the spin-liquid state. Besides, x decreases ap- K ⌬ ␯ ⌬ ␥ ϳ Pz (t, f ,HLF , ). / ␮ is the local field created on the preciably for HLF 10 kG whatever the value of p. We can 718 Low Temp. Phys. 31 (8–9), August–September 2005 Bono et al.

with 1Ϫp but rather decreases, opposite to the case of ca- nonical spin glasses. We could further confirm the existence of such a frozen component in SCGO(pϭ0.95– 0.89), since a clear recovery of a small part (ϳ4%) of the asymmetry is found at long times ͑e.g., Fig. 21, inset͒. It is observed for tϳ100␭Ϫ1, which is out of the experimental range for BSZCGO(p). To summarize, the picture based on a coherent RVB state, which magnetic excitations are mobile fluctuating spins 1/2 on the kagome´ lattice, explains well the data, pro- vided that ͑i͒ these excitations can be generated even for T →0. This underlines the smallness of the ‘‘magnetic’’ gap, if ⌬Ͻ ͑ ͒ any, typically J/1000; ii an energy scale related to Tg ͑fields of the order of 10 kG or, equivalently, temperatures of the order of Tg , which vary very little with p) is high enough to destroy the coherent RVB type state; ͑iii͒ the sub- FIG. 24. Fitting parameters of Eq. ͑6͒ for BSZCGO and SCGO ͑unfilled and stitution defects are accounted for by an additional classical filled symbols͒; f is common for both systems. relaxation process.

associate this decrease to the existence of an energy scale E. Spin-glass-like transition ϳ 1 K, which is of the order of Tg . Finally, we observe a linear variation of f with p ͑Fig. 24c͒, and f tends to vanish The ‘‘intrinsic’’ spin-glass-like transition at Tg is one of around pϳ0.5, a limit consistent with our classical approach the most puzzling observation in these frustrated magnets. ͑Fig. 22͒. Indeed, since we find ␯ϳ⌬ when T →0, the rela- Indeed, the origin of such a spin-glass state in a disorder-free system still awaits for a complete understanding, although tionship f (p)⌬ϳ␭ → (p) is expected. This indicates that T 0 recent theoretical approaches catch some of the facets of this even far from the substituted sites, the coherent state is 80 somehow affected, e.g., the density of spinons could be original ground state. The most realistic model, presented by Ferrero et al. in Ref. 81, uses the so-called ‘‘dimerized’’ smaller. 14,16 ␭ approach, i.e., considers a geometry of the kagome´ lat- As suggested by the similar p variation of T→0 in both BSZCGO(p) and SCGO(p), we assume that the excitation tice with two different kinds of equilateral triangles, corre- ͑ ͒ modes are identical in both systems, i.e., f and ␯ are kept the sponding to the Cr b layers in SCGO(p) and BSZCGO(p) ͑ ͒ ϭ same for comparable p. We find ⌬ϳ1200 ␮sϪ1 for our Fig. 3 . It predicts a spin-glass-like transition in the S 1/2 ´ SCGO(pу0.89) samples, in agreement with previous work kagome lattice, related to the freezing of the chirality, with ϳ Ј Ј ´ on SCGO͑0.89͒.49 It is quite rewarding to find that a common Tg 0.5J . Here J is the largest coupling in the kagome ͑ ͒ physical picture underlies all the sets of data at low tempera- planes corresponding to the thick lines in Fig. 3 . Although ture for both kagome´ bilayers. 0.05J is of the order of magnitude of the experimental Tg , We now extend the fits of the muon spin polarization the very similar values of the Cr–Cr bonds in both systems ͑ ͒ P (t) to the whole temperature range ͑Fig. 20b͒, fixing f to and Eq. 4 do not permit prediction of a factor 2 in Tg . z ␮ its low-temperature value in order to limit the number of free From our SR results, we propose an other interpreta- parameters. It would also, phenomenologically, mean that tion which still awaits for theoretical confirmation. We find once a spinon is created, its deconfinement process remains no energy gap for the spin 1/2 excitations. This could explain temperature independent, which is a quite reasonable as- quite well why the transition to the spin-glass state is fairly sumption from a quantum point of view. As expected from independent of p. Indeed, in this framework, spinons could ͑ ͒ mediate the interactions between magnetic defects localized the change of shape around Tg Refs. 49, 57 , the weight x of the sporadic term decreases to finally enter a high- around spin vacancies. This is corroborated by all the scaling temperature regime ͑Fig. 24b͒ with an exponential muon re- properties in T/Tg . The transition to the spin-glass state laxation (x→0). The similarity between the field and tem- would correspond then to the formation of the coherent sin- perature dependence of x indicates that the sporadic regime glet state rather than any interaction strength between de- is destroyed with an energy of the order of the freezing tem- fects. The bond defects inherent to BSZCGO(p) probably perature T . Figure 22b displays a sharp crossover from one decrease the value of this coherent state energy scale since it g is more favorable to create localized singlets in this geom- state to the other between Tg and 3Tg , corresponding to the steep decrease of ␭͑Fig. 21͒, for both BSZCGO(p) and etry. This is qualitatively consistent with the twice lower Ͼ ␯ ␭ ͑ ͒ spin-glass transition temperature in this family. SCGO(p). For T Tg , and Ј Fig. 24d decrease by one order of magnitude up to 10Tg . At high temperature, it is natural to think in terms of F. Extension to other compounds? paramagnetic fluctuating spins. At lower temperatures, ␭ seems to diverge at Tg and below, the weight of the expo- We already noticed that the ␮SR undecouplable Gauss- nential term at low fields (1Ϫxϳ1Ϫp) could correspond to ian line shape, altogether with a plateau of the relaxation localized frozen spins, reflecting the glassy component mea- rate, have been reported in other compounds, and now elabo- sured by SQUID. It is noteworthy that Tg does not increase rate about the possible links between these systems. Low Temp. Phys. 31 (8–9), August–September 2005 Bono et al. 719

The other kagome´ compounds, volborthite77,82,83 and Cr oretical predictions. This response generates a Curie-like up- jarosite76 display the same properties as the kagome´ bilayers turn in the macroscopic susceptibility at low temperature. and our phenomenological model seems correct. However, local measurements of the susceptibility, using Ga Kojima et al. already noticed the similarity between the NMR, show that the intrinsic susceptibility of the kagome´ ͑ ͒ 78,84 doped Haldane chain Y2ϪxCaxBaNiO5 and SCGO 0.89 . bilayers decreases below 45 K, a temperature of the order of However, contrary to the kagome´ bilayers, the relaxation J. This behavior was also predicted theoretically and is con- function becomes more ‘‘conventional’’ when the system be- sistent with short-ranged spin-spin correlations, on the order comes purer, i.e., the muon spin relaxation function is a of the lattice parameter. We cannot give a definitive conclu- square root exponential in the pure system at low tempera- sion about the existence of a spin gap with NMR experi- ture. It is not surprising since in the pure Haldane chain one ments because the gallium nuclei cannot be probed by NMR expects a valence bond crystal.78,84 Therefore no magnetic at low temperature. However a maximum value of ϳJ/10 is excitation are allowed at low temperature and no depolariza- found for such a gap. tion is expected in ␮SR. However, mobile excitations are We measured the spin dynamics using muon spin relax- added when the chains are doped,85 i.e., the muon spin re- ation experiments, down to 30 mK. We find that both sys- laxation becomes unconventional. This corresponds to the tems SCGO and BSZCGO display a dynamical magnetic data and is consistent with our unconfined spinon phenom- state down to this temperature, unconventional for the purest enological model. samples. Very simple considerations show that the spin- Ӷ␪ Finally, Fukaya et al. reported the same behavior in glass-like state, measured at Tg CW with macroscopic sus- ϭ 86 Sr(Cu1ϪxZnx)2(BO3)2 for x 0 and 0.02. This system dis- ceptibility techniques, is correlated to the slowing down of plays, theoretically and experimentally, an exact singlet the Cr spin dynamics in the whole sample. The data are well dimer state and its intrinsic susceptibility vanishes for TՇ3 accounted by a description based on unconfined spinons as K.87,88 Nonetheless, a CW term, corresponding to 0.72% of the magnetic excitations and we suggest that this phenom- Sϭ1/2 impurities with respect to the Cu sites, is observed enological approach remains valid for several spin singlet below 4 K,87 which is finally comparable, from the macro- compounds at low temperature. Indeed, they all display strik- scopic susceptibility aspect, to the former cases of pure ing similar aspects such as a spin-glass transition of the mag- volborthite and pure Y2CaBaNiO5 . However, according to netic defect channel in the macroscopic susceptibility and an us, there is today no clue about the possibility of mobile undecouplable Gaussian line shape with a relaxation rate pla- excitations in this magnetic network with a few percent of teau of the muon polarization below this spin-glass transi- defects. tion. We conclude that all these systems display ͑i͒ a theoret- In all these systems the fast depolarization of the muon ical and/or experimental singlet ground state; ͑ii͒ a defect down to the experimental limit of Tϳ30 mK underlines the term observed through a low temperature CW like upturn in existence of magnetic excitations at low temperature and the macroscopic susceptibility; ͑iii͒ a spin-glass-like transi- hence the weakness of an hypothetical unconfined spinon ͑ tion of this defect term at a temperature Tg but gap. An other spin dynamics study with neutron spin echo 13͒ ͑ ͒ Sr(Cu1ϪxZnx)2(BO3)2); iv a plateau of the muon spin technique is in progress and could give more details about → 89 relaxation rate below Tg . this T 0 spin-liquid state. It is important to notice that the clear experimental cor- We thank H. Alloul, F. Bert, J. Bobroff, R. J. Cava, D. relation between the muon spin relaxation rate plateau and Huber, A. Keren, C. Lhuillier, M. Mekata, G. Misguich, R. the spin-glass-like transition is strongly against a pure muon- Moessner, C. Mondelli, H. Mutka, B. Ouladdiaf, C. Payen, P. induced effect argument for this plateau. Even if the muon Schiffer and P. Sindzingre for fruitful discussions. We also has indeed an effect on the physics of these compounds, the wish to thank A. Amato, C. Baines, and A.D. Hillier, our ␭ ␮ plateau below Tg shows that a bulk transition occurs at this SR local contacts, whose outstanding efforts have made temperature, i.e., that this freezing cannot be due to isolated these experiments possible. impurities. The ␮SR experiments were performed at the Swiss Theoretical approaches are now required to link a pos- Muon Source, Paul Scherer Institute, Villigen, Switzerland, sible unconfined spinons state to a spin-glass-like behavior and financially supported by the Federal Office for Education and a relaxation plateau of the implanted muon spins. They and Science, Berne, Switzerland, as well as at the ISIS facil- could therefore corroborate our phenomenological approach. ity, Rutherford Appleton Laboratory, Didcot, United King- dom. IV. CONCLUSION Part of this work was financially supported by the EC IHP Program for large scale facilities and by the EC Frame- With Heisenberg spins and nearest neighbor couplings of work Programmes V and VI. the order of 40 K, the kagome´ bilayers of SCGO and We are grateful to the muon beamline groups and to the BSZCGO are today the archetypes of highly frustrated mag- technical staff of Laboratoire de Physique des Solides, Orsay. nets. The comparison between SQUID and NMR experi- ments on the two slightly different compounds SCGO and BSZCGO allows us to isolate the intrinsic properties of the *Present address: Kamerlingh Onnes Laboratory, Leiden University, P.O. frustrated kagome´ bilayer geometry and to understand better Box 9504, 2300 RA Leiden, The Netherlands. † the role of the ͑non͒magnetic defects. We find that defects Present address: Institut fu¨r Experimentelle und Angewandte Physik, Christian-Albrechts-Universita¨t zu Kiel, D-24098 Kiel, Germany. localized in the frustrated network induce a spatially ex- ‡E-mail: [email protected] tended response of the magnetic system, consistent with the- 1͒This is different from the situation encountered in spin glasses, where the 720 Low Temp. Phys. 31 (8–9), August–September 2005 Bono et al.

cooling history finishes by selecting one of the numerous local minima Lecheminant, and L. Pierre, Eur. Phys. J. B 2, 501 ͑1998͒. present in the free energy. 16 F. Mila, Phys. Rev. Lett. 81, 2356 ͑1998͒. 2͒ Hϭ ͚ 17 ͑ ͒ We define the Hamiltonian as J ͗i, j͘Si•Sj , where i, j are the nearest G. Misguich, D. Serban, and V. Pasquier, Phys. Rev. B 67, 214413 2003 . neighbors. 18 B. Canals and C. Lacroix, Phys. Rev. Lett. 80, 2933 ͑1998͒. 3͒The Sϭ3/2 Cr jarosite is the only exception.76 19 C. Lhuillier and P. Sindzingre, in Quantum Properties of Low- 4͒The official labeling is as follows for SCGO ͓BSZCGO͔:Ga͑4f͓͒Ga͑2d͔͒ Dimensional Antiferromagnets, Y. Ajiro and J. P. Boucher ͑eds.͒, Kyushu for Ga͑1͒,Ga͑4e͓͒Ga͑2c͔͒ for Ga͑2͒,Cr͑1a͓͒Cr͑2a͔͒ for Cr͑a͒,Cr͑12k͒ University Press, Fukuoka ͑2001͒, p. 111, cond-mat/0212351. ͓ ͑ ͔͒ ͑ ͒ ͑ ͑ ͒ 20 ͑ ͒ Cr 6i for Cr b and Cr 4fvi) for Cr c in SCGO. We used simplified S. E. Palmer and J. T. Chalker, Phys. Rev. B 62, 488 2000 . notations for clarity in the comparison of the two systems. 21 O. Tchernyshyov, R. Moessner, and S. L. Sondhi, Phys. Rev. Lett. 88, 5͒Since the Ga͑2͒ line of BSZCGO is governed by quadrupolar effects, its 067203 ͑2002͒. ϰ ␯2 ␯ 90 22 ͑ ͒ width is Q/ l . Therefore this line is broader at low field. On the M. Elhajal, B. Canals, and C. Lacroix, Phys. Rev. B 66, 014422 2002 . 23 contrary, the quadrupolar frequency of the Ga͑1͒ is smaller and the line- S. Dommange, M. Mambrini, B. Normand, and F. Mila, Phys. Rev. B 68, width is proportional to the applied external field. 224416 ͑2003͒. 24 6͒Notice that the high-temperature analysis of both 71Ga(1) and 71Ga(2) M. E. Zhitomirsky, A. Honecker, and O. A. Petrenko, Phys. Rev. Lett. 85, ͑ ͒ lines in BSZCGO allows us to compare their shifts, which yields an esti- 3269 2000 . 25 mate of the chemical shift of 0.15͑3͒%. J.-C. Domenge, P. Sindzingre, and C. Lhuillier, cond-mat/0502414. 26 7͒We checked that the transverse relaxation time T is homogeneous over the A. S. Wills, Can. J. Phys. 79, 1501 ͑2001͒. 2 27 whole 71Ga(2) line. Its shape is hence used for the subtraction whereas its Y. Narumi, K. Katsumata, Z. Honda, J. C. Domenge, P. Sindzingre, C. ϳ ͓ ␶ Ϫ␶ 71͔ Lhuillier, Y. Shimaoka, T. C. Kobayashi, and K. Kindo, Europhys. Lett. intensity is multiplied by exp 2( l s)/T2 . ͒ 65,705͑2004͒. 8 In the case of SCGO, ␹ is a three-component susceptibility, the third macro 28 G. Lawes, M. Kenzelmann, N. Rogado, K. H. Kim, G. A. Jorge, R. J. contribution coming from the Cr͑c͒-Cr͑c͒ spin pairs, which we drop here Cava, A. Aharony, O. Entin-Wohlman, A. B. Harris, T. Yildirim et al., for clarity. Phys. Rev. Lett. 93, 247201 ͑2004͒. 9͒C. Lhuillier, private communication. 29 N. P. Raju, M. Dion, M. J. P. Gingras, T. E. Mason, and J. E. Greedan, 10͒A dipolar broadening due to the diluted paramagnetic defects91 would be Phys. Rev. B 59, 14489 ͑1999͒. too small to justify the broadening of the 69,71Ga spectra. 30 A. Keren and J. Gardner, Phys. Rev. Lett. 17, 177201 ͑2001͒. 11͒ ͑ ͒ ͑ ͒ ͑ ͒ The Ga 1 nuclei are coupled to 9 Cr b and3Cra . The measured sus- 31 J. A. Hodges, P. Bonville, A. Forget, A. Yaouanc, P. Dalmas de Re´otier, G. ceptibility is hence ␹ ϰ9␹ ϩ3␹ , where ␹ and ␹ are the Cr͑b͒ kag,NMR b a b a Andre´, M. Rams, K. Kro´las, C. Ritter, P. C. M. Gubbens et al., Phys. Rev. ͑ ͒ and Cr a susceptibilities. On the other hand, the macroscopic susceptibil- Lett. 88, 077204 ͑2002͒. ity probes these susceptibilities with their stoichiometric ratio, i.e., 32 A. Keren, J. S. Gardner, G. Ehlers, A. Fukaya, E. Segal, and Y. J. Uemura, ␹ ϰ ␹ ϩ␹ ͑ ͒ kag,macro 6 b a Ref. 55 . A small difference between the ratios Phys. Rev. Lett. 92, 107204 ͑2004͒. ␹ ␹ ͓⌬ ͔ b / a would yield a slightly different ratio H/Hl(p) /Cdef(p). 33 S. H. Lee, C. Broholm, W. Ratcliff, G. Gasparovic, Q. Huang, T. H. Kim, 12͒ We attribute the factor of 8 between ␭ in the two systems to a different and S. W. Cheong, Nature ͑London͒ 418, 856 ͑2002͒. coupling of the muon spin to the Cr spins. Transverse field experiments 34 M. Harris, Nature ͑London͒ 399,311͑1999͒. ͑ ͒ show for instance a larger linewidth in SCGO 0.95 than in 35 A. P. Ramirez, A. Hayashi, R. J. Cava, R. Siddhartan, and B. S. Shastry, ͑ ͒ 41 ⌬ BSZCGO 0.91 despite a lower defect term. On the other hand, , Nature ͑London͒ 399, 333 ͑1999͒. ␮ ⌬ϰ͚ 3 which is mainly created by dipolar interactions in SR, is 1/r , 36 X. Obradors, A. Labarta, A. Isalgue´, J. Tejada, J. Rodrı´guez, and M. Per- where the sum is over all the spins in the lattice and r is the distance net, Solid State Commun. 65,189͑1988͒. between the muon site and these spins.73 Usually, the muon is located near 37 I. S. Hagemann, Q. Huang, X. P. A. Gao, A. P. Ramirez, and R. J. Cava, 2Ϫ 92 an O ion. In a very simplified approach, we computed ⌬, considering Phys. Rev. Lett. 86, 894 ͑2001͒. Ϫ that the muon would be located on the O2 sites, as an average, and find 38 A. P. Ramirez, G. P. Espinosa, and A. S. Cooper, Phys. Rev. B 45, 2505 ⌬Ϸ1100 ␮sϪ1 and 800 ␮sϪ1 in SCGO͑1͒ and BSZCGO͑1͒. This is in ͑1992͒. perfect agreement with the high-temperature evaluation of ⌬. Since ␭ 39 H. Ohta, M. Sumikawa, M. Motokawa, H. Kikuchi, and H. Nagasawa, J. ϰ⌬2 in the fast fluctuation limit and ␭ϰ⌬ in the static limit, we expect, Phys. Soc. Jpn. 65, 848 ͑1996͒. 40 with this simple approach, that ␭ will be 1.5–2 times larger in SCGO(p). L. Limot, P. Mendels, G. Collin, C. Mondelli, B. Ouladdiaf, H. Mutka, N. Blanchard, and M. Mekata, Phys. Rev. B 65, 144447 ͑2002͒. The missing factor of 4 may either come from a more complex muon sites 41 distribution, due to the O2Ϫ ions around the Cr͑c͒ sites, or to a different D. Bono, P. Mendels, G. Collin, and N. Blanchard, Phys. Rev. Lett. 92, 89 217202 ͑2004͒. dynamics range, as suggested by neutron spin echo. 42 13͒To our knowledge, no FC-ZFC magnetization data have been published S. H. Lee, C. Broholm, G. Aeppli, T. G. Perring, B. Hessen, and A. Taylor, Phys. Rev. Lett. 76, 4424 ͑1996͒. about Sr(Cu Ϫ Zn ) (BO ) . 1 x x 2 3 2 43 L. J. de Jongh and A. R. Miedema, Adv. Phys. 50,947͑2001͒. 44 A. P. Ramirez, G. P. Espinosa, and A. S. Cooper, Phys. Rev. Lett. 64, 2070 ͑1990͒. 45 1 P. W. Anderson, Mater. Res. Bull. 8, 153 ͑1973͒. P. Lecheminant, B. Bernu, C. Lhuillier, L. Pierre, and P. Sindzingre, Phys. ͑ ͒ 2 235 ͑ ͒ Rev. B 56, 2521 1997 . P. W. Anderson, Science , 1196 1987 . 46 3 G. Baskaran, Phys. Rev. Lett. 91, 097003 ͑2003͒. A. P. Ramirez, B. Hessen, and M. Winklemann, Phys. Rev. Lett. 84, 2957 ͑ ͒ 4 B. Bernu, P. Lecheminant, C. Lhuillier, and L. Pierre, Phys. Rev. B 50, 2000 . 47 10048 ͑1994͒. P. Sindzingre, G. Misguich, C. Lhuillier, B. Bernu, L. Pierre, C. Waldt- ͑ ͒ 5 A. P. Ramirez, in Handbook on Magnetic Materials, K. J. H. Busch ͑ed.͒, mann, and H. U. Everts, Phys. Rev. Lett. 84,2953 2000 . 48 Elsevier Science, Amsterdam ͑2001͒, Vol. 13, p. 423. C. Broholm, G. Aeppli, G. P. Espinosa, and A. S. Cooper, Phys. Rev. Lett. 6 R. Stewart ͑ed.͒, Proceedings of the Highly Frustrated Magnetism 2003 65, 3173 ͑1990͒. 49 Conference, Grenoble, France ͑2004͒, J. Phys.: Condens. Matter 16, Y. J. Uemura, A. Keren, K. Kojima, L. P. Le, G. M. Luke, W. D. Wu, Y. S553–922 ͑2004͒. Ajiro, T. Asano, Y. Kuriyama, M. Mekata et al., Phys. Rev. Lett. 73,3306 7 K. Binder and A. P. Young, Rev. Mod. Phys. 58, 801 ͑1986͒. ͑1994͒. 8 D. A. Huse and A. D. Rutenberg, Phys. Rev. B 45, 7536 ͑1992͒. 50 B. Martı´nez, F. Sandiumenge, A. Rouco, A. Labarta, J. Rodrı´guez- 9 J. T. Chalker, P. C. W. Holdsworth, and E. F. Shender, Phys. Rev. Lett. 68, Carvajal, M. Tovar, M. T. Causa, S. Galı´, and X. Obradors, Phys. Rev. B 855 ͑1992͒. 46, 10786 ͑1992͒. 10 I. Ritchey, P. Chandra, and P. Coleman, Phys. Rev. B 47, 15342 ͑1993͒. 51 C. Mondelli et al., private communication. 11 G. Misguich and C. Lhuillier, in Frustrated Spin Systems,H.T.Diep͑ed.͒, 52 A. P. Ramirez, Annu. Rev. Mater. Sci. 24, 453 ͑1994͒. World Scientific, Singapore ͑2003͒, cond-mat/0310405, and references 53 A. Keren, P. Mendels, M. Horvatic´, F. Ferrer, Y. J. Uemura, M. Mekata, therein. and T. Asano, Phys. Rev. B 57, 10745 ͑1998͒. 12 R. Moessner, Can. J. Phys. 79, 1283 ͑2001͒. 54 P. Mendels, A. Keren, L. Limot, M. Mekata, G. Collin, and M. Horvatic´, 13 C. Zeng and V. Elser, Phys. Rev. B 51, 8318 ͑1995͒. Phys. Rev. Lett. 85, 3496 ͑2000͒. 14 M. Mambrini and F. Mila, Eur. Phys. J. B 17, 651 ͑2000͒. 55 L. Limot, P. Mendels, G. Collin, C. Mondelli, H. Mutka, and N. Blan- 15 C. Waldtmann, H. U. Everts, B. Bernu, C. Lhuillier, P. Sindzingre, P. chard, Can. J. Phys. 79, 1393 ͑2001͒. Low Temp. Phys. 31 (8–9), August–September 2005 Bono et al. 721

56 D. Bono, P. Mendels, G. Collin, and N. Blanchard, J. Phys.: Condens. 76 A. Keren, K. Kojima, L. P. Le, G. M. Luke, W. D. Wu, Y. J. Uemura, M. Matter 16, S817 ͑2004͒. Takano, H. Dabkowska, and M. J. P. Gingras, Phys. Rev. B 53, 2451 57 A. Keren, Y. J. Uemura, G. Luke, P. Mendels, M. Mekata, and T. Asano, ͑1996͒. Phys. Rev. Lett. 84, 3450 ͑2000͒. 77 A. Fukaya, Y. Fudamoto, I. M. Gat, T. Ito, M. I. Larkin, A. T. Savici, Y. J. 58 D. Bono, P. Mendels, G. Collin, N. Blanchard, F. Bert, A. Amato, C. Uemura, P. P. Kyriakou, G. M. Luke, M. T. Rovers et al., Phys. Rev. Lett. Baines, and A. D. Hillier, Phys. Rev. Lett. 93, 187201 ͑2004͒. 91, 207603 ͑2003͒. 59 ͑ ͒ D. E. MacLaughlin and H. Alloul, Phys. Rev. Lett. 36, 1158 1976 . 78 K. Kojima, A. Keren, L. P. Le, G. M. Luke, B. Nachumi, W. D. Wu, Y. J. 60 ͑ ͒ A. B. Harris, C. Kallin, and A. J. Berlinsky, Phys. Rev. B 45, 2899 1992 . Uemura, K. Kiyono, S. Miyasaka, H. Takagi et al., Phys. Rev. Lett. 74, 61 ͑ ͒ K. Motida and S. Miyahara, J. Phys. Soc. Jpn. 28, 1188 1970 . 3471 ͑1995͒. 62 ͑ ͒ E. J. Samuelsen, M. T. Hutchings, and G. Shirane, Physica Amsterdam 79 A. Chubukov, Phys. Rev. Lett. 69, 832 ͑1992͒. 48 ͑ ͒ ,13 1970 . 80 F. Mila and D. Dean, Eur. Phys. J. B 26, 301 ͑2002͒. 63 C. Mondelli, K. Andersen, H. Mutka, C. Payen, and B. Frick, Physica B 81 M. Ferrero, F. Becca, and F. Mila, Phys. Rev. B 68, 214431 ͑2003͒. 267–268,139͑1999͒. 82 64 ͑ ͒ Z. Hiroi, M. Hanawa, N. Kobayashi, M. Nohara, H. Takagi, Y. Kato, and A. J. Garcia-Adeva and D. L. Huber, Phys. Rev. B 63, 174433 2001 . ͑ ͒ 65 ͑ ͒ M. Takigawa, J. Phys. Soc. Jpn. 70, 3377 2001 . P. Schiffer and I. Daruka, Phys. Rev. B 56, 13712 1997 . 83 66 M. Takigawa, N. Motoyama, H. Eisaki, and S. Uchida, Phys. Rev. B 55, F. Bert, D. Bono, P. Mendels, J. C. Trombe, P. Millet, A. Amato, C. ͑ ͒ 14129 ͑1999͒. Baines, and A. Hillier, J. Phys.: Condens. Matter 16, S829 2004 . 84 ͑ ͒ 67 F. Tedoldi, R. Santachiara, and M. Horvatic, Phys. Rev. Lett. 83,412 K. Kojima, Ph. D. thesis, The University of Tokyo 1995 . 85 ͑ ͒ ͑1999͒. B. Ammon and M. Imada, Phys. Rev. Lett. 85, 1056 2000 . 86 68 M. Azuma, Z. Hiroi, M. Takano, K. Ishida, and Y. Kitaoka, Phys. Rev. A. Fukaya, Y. Fudamoto, I. M. Gat, T. Ito, M. I. Larkin, A. T. Savici, Y. J. Lett. 73,3463͑1994͒. Uemura, P. P. Kyriakou, G. M. Luke, M. T. Rovers et al., Physica B 326, 69 S. Ouazi, J. Bobroff, H. Alloul, and W. A. MacFarlane, Phys. Rev. B 70, 446 ͑2003͒. 104515 ͑2004͒. 87 H. Kageyama, K. Yoshimura, R. Stern, N. V. Mushnikov, K. Onizuka, M. 70 R. Moessner and A. J. Berlinsky, Phys. Rev. Lett. 83,3293͑1999͒. Kato, K. Kosuge, C. P. Slichter, T. Goto, and Y. Ueda, Phys. Rev. Lett. 82, 71 A. Abragam, Principles of Nuclear Magnetism, Clarendon Press, Oxford 3168 ͑1999͒. ͑1961͒. 88 K. Kodama, M. Takigawa, M. Horvatic´, C. Berthier, H. Kageyama, Y. 72 S. L. Lee, S. H. Kilcoyne, and R. Cywinski ͑eds.͒, Muon Science: Muons Ueda, S. Miyahara, F. Becca, and F. Mila, Science 298,395͑2002͒. in Physics, Chemistry and Materials, Proceedings of the Fifty-First Scot- 89 H. Mutka et al., in preparation. tish Universities Summer School in Physics, August 1998, Scottish Uni- 90 M. H. Cohen and F. Reif, Solid State Phys. 5, 321 ͑1957͒. versities Summer School in Physics & Institute of Physics Publishing, 91 R. E. Walstedt and L. R. Walker, Phys. Rev. B 9, 4857 ͑1974͒. Bristol and Philadelphia ͑1999͒. 92 J. H. Brewer, R. F. Kiefl, J. F. Carolan, P. Dosanjh, W. N. Hardy, S. R. 73 R. S. Hayano, Y. J. Uemura, J. Imazato, N. Nishida, T. Yamazaki, and R. Kreitzman, Q. Li, T. M. Riseman, P. Schleger, H. Zhou et al., Hyperfine ͑ ͒ Kubo, Phys. Rev. B 20,850 1979 . Interact. 63, 177 ͑1990͒. 74 Y. J. Uemura, T. Yamazaki, D. R. Harshman, M. Senba, and E. J. Ansaldo, Phys. Rev. B 31,546͑1985͒. This article was published in English in the original Russian journal. Repro- 75 A. Keren, Phys. Rev. B 50, 10039 ͑1994͒. duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 31, NUMBERS 8–9 AUGUST–SEPTEMBER 2005

Ordering in two-dimensional Ising models with competing interactions G. Y. Chitov*

Department 7.1-Theoretical Physics, University of Saarland, Saarbru¨cken D-66041, Germany; Department of Physics and Astronomy, Laurentian University, Sudbury, ON, P3E 2C6 Canada C. Gros

Department 7.1-Theoretical Physics, University of Saarland, Saarbru¨cken D-66041, Germany; Institute for Theoretical Physics, Frankfurt University, Frankfurt 60438, Germany ͑Submitted February 28, 2005͒ Fiz. Nizk. Temp. 31, 952–967 ͑August–September 2005͒ We study the 2D Ising model on a square lattice with additional non-equal diagonal next-nearest neighbor interactions. The cases of classical and quantum ͑transverse͒ models are considered. Possible phases and their locations in the space of three Ising couplings are analyzed. In particular, incommensurate phases occurring only at non-equal diagonal couplings, are predicted. We also analyze a spin-pseudospin model comprised of the quantum Ising model coupled to XY spin chains in a particular region of interactions, corresponding to the Ising sector’s super- antiferromagnetic ͑SAF͒ ground state. The spin-SAF transition in the coupled Ising-XY model into a phase with co-existent SAF Ising ͑pseudospin͒ long-range order and a spin gap is considered. Along with destruction of the of the Ising sector, the phase diagram of the Ising-XY model can also demonstrate a re-entrance of the spin-SAF phase. A detailed study of the latter is presented. The mechanism of the re-entrance, due to interplay of interactions in the coupled model, and the conditions of its appearance are established. Applications of the spin-SAF theory for the transition in the quarter-filled ladder compound ͓ ͔ NaV2O5 are discussed. © 2005 American Institute of Physics. DOI: 10.1063/1.2008132

I. INTRODUCTION Our interest in the subject comes from the earlier work on a quantum Ising model coupled to the spin chains.6 This The role of competing interactions in ordering is a fas- kind of coupled so-called spin-pseudospin ͑or spin-orbital͒ cinating problem of condensed matter physics. One of the models appear in context of the phase transition in NaV2O5 , most canonical examples of such systems are frustrated Ising which has inspired a great experimental and theoretical effort models which demonstrate a plethora of critical properties, in recent years. ͑For a review see Ref. 7.͒ Going deeper into ͑ far from being exhaustively studied. For a review see Ref. analysis, we came to realize that the Ising sector of the prob- ͒ 1͒ 1. The frustrations can be either geometrical, like, e.g., in lem is in fact the 2D transverse NN and NNN Ising model on the Ising model on a triangular lattice, or they can be brought a square lattice. It turns out that even its classical counterpart about by the next-nearest neighbor ͑NNN͒ interactions. (⍀ϭ0) has been studied1 only for the case of equal NNN Competing interactions ͑frustrations͒ can, e.g., result in new couplings, J ϭJ . The elementary plaquette of this lattice phases, change the Ising universality class, or even destroy 1 2 with the notations for couplings is shown in Fig. 1. To the the order at all. Another interesting aspect of the criticality in best of our knowledge, this 2D NN and NNN Ising model in frustrated Ising models is an appearance of quantum critical point͑s͒͑QCP͒ at special frustration points of model’s high transverse field is a complete terra incognita even at J1 ϭ degeneracy, and related quantum phase transitions.2 J2 . Inclusion of a transverse field ͑⍀͒ brings an extra scale So, as the first step, we find the ground-state phase dia- ⍀ϭ into the game, giving raise to a new and complicated critical gram of the classical ( 0) Ising model at arbitrary cou- behavior. The Ising models with ⍀ϭ0 and ⍀0 are also plings Jᮀ ,J1 ,J2 . Along with the three ordered phases found 8 ϭ often called classical and quantum, respectively. For a review earlier by Fan and Wu for the case J1 J2—ferromagnetic ͑ ͒ ͑ ͒ on the Ising models in transverse field ͑IMTF͒ see Ref. 3. FM , antiferromagnetic AF , and super-antiferromagnetic Most studies of the frustrated quantum Ising models are re- ͑SAF͒—the model has a fourth phase which can occur if the ϭϪ stricted to their ground-state properties, when mapping of the sign of (J1 /J2) 1. We call it super-ferro- 9 d-dimensional quantum model at Tϭ0 onto its antiferromagnetic ͑SFAF͒. From mean-field-type arguments (dϩ1)-dimensional classical counterpart helps to analyze we predict also the existence of an incommensurate ͑IC͒ the ground-state phase diagram of the former. For the NNN phase at TϾ0 in this model. The 2D IC phase is also called 2D models we are interested in, there has been a consider- floating.10 Similar phase is known for the well-studied 2D able effort on the quantum anisotropic NNN Ising ͑ANNNI͒ ANNNI model.1,10,11 We present a qualitative temperature model, reviewed in Ref. 3. The studies of some other 2D phase diagram for the regions of the coupling space where frustrated transverse Ising models have appeared only the IC phase is located. Note that the three phases—SAF, recently.4,5 SFAF, IC—can occur only in the presence of competing in-

1063-777X/2005/31(8–9)/13/$26.00722 © 2005 American Institute of Physics Low Temp. Phys. 31 (8–9), August–September 2005 G. Y. Chitov and C. Gros 723

Ising sector of this model is given by the Hamiltonian of the NN and NNN IMTF, and the Ising variables ͑called pseu- dospins for this case͒ represent physically the charge degrees of freedom. We model the spin sector by the array of the XY spin chains. The 2D long-range charge order in NaV2O5 is identified as the SAF phase of the Ising model, and we re- strict our analysis to the SAF region of the couplings (Jᮀ ,J1 ,J2). The coupled model is handled by combining the mean-field treatment of its Ising sector with the use of FIG. 1. Couplings on an elementary plaquette in the NN and NNN 2D Ising exact results available for the XY spin chains. Since the SAF model ͑1͒. Ordering patterns in the super-antiferromagnetic ͑SAF͒ and state is only fourfold degenerate, the mean-field predictions super-ferro-antiferromagnetic ͑SFAF͒ phases. for the Ising sector of the coupled model are expected to be at least qualitatively correct. The mean-field equations for the coupled model are, teractions in the Ising model, and the latter two occur only if with some minor modifications, the same as we have ob-  J1 J2 . tained earlier.6 A striking feature of the coupled model is that This analysis of the classical Ising model lays the it always orders from the charge-disordered spin-gapless groundwork for venturing into its study in presence of a state into the phase of co-existent SAF charge order and spin transverse field. The role of transverse field is subtle. A more gap. We call this the spin-SAF transition. By always we straightforward aspect is that its increase above certain criti- mean that the critical temperature of the spin-SAF transition cal value can eventually destroy the ordered state of the clas- is nonzero for all Ising couplings within the whole consid- sical model, and in the ground state the transverse field re- ered SAF region. In other words, the QCP of the IMTF is sults in appearance of a QCP. This is similar to the well- destroyed, and this is due to the spin-charge ͑-pseudospin͒ understood quantum NN Ising model. Another particularly coupling. interesting aspect in the role of transverse field is that it can This property of the spin-SAF transition and the param- lift degeneracy of the ground state and stabilize new phases eters of the spin-SAF phase were studied earlier,6 so in this at finite temperature in a highly frustrated model, like, e.g., work we only reinstate some points and stress the distinc- 4,5 the antiferromagnetic isotropic triangular Ising model, tions pertinent to the present model. Ͼ ⍀ϭ which is disordered at any T 0 when 0. The other remarkable feature of the coupled model’s The behavior of the systems with infinitely degenerate phase diagram is re-entrance, which was not well understood ͑ ground states with or without a finite ground-state entropy in our earlier work.6 Now we carry out an analytical study of ͒ per spin can be quite complicated in the presence of trans- the re-entrance and establish the conditions when it can oc- verse field. It lies beyond the scope of the present work, and cur. This analysis allows us to understand the detailed definitely cannot be understood from the mean-field analysis mechanism of this interesting phenomenon generated by we apply in this study. For the NN and NNN Ising model we competing interactions. ͑ ͒ only identify the lines planes in the space of couplings The rest of the paper is organized as follows. Section 2 (Jᮀ ,J1 ,J2) where the model is highly degenerate, and in contains our results on the ordering in the 2D NN and NNN their neighborhood we expect some new exotic phases gen- Ising model at ⍀ϭ0 and ⍀0. The results on the spin-SAF ⍀ erated by 0 to appear. transition in the coupled model are presented in Sec. 3. The ϭ From mapping of the NN and NNN IMTF at T 0 onto final Sec. 4 presents the summary and discussion. its classical 3D counterpart, we qualitatively predict the ͑mean-field͒ ground-state phase boundaries of the quantum II. 2D NEAREST- AND NEXT-NEAREST-NEIGHBOR ISING model in the coupling space (Jᮀ ,J1 ,J2). In particular, it MODEL follows from our analysis that in the presence of transverse field the IC ground-state phase can penetrate into some parts We consider the 2D Ising model on a square lattice with of the FM, AF, SAF regions of the classical model (⍀ the Hamiltonian ϭ 0). 1 1 ϭ T xT xϩ T xT x ͑ ͒ Finally, we consider the coupled spin-pseudospin model. H ͚ Jᮀ i j ͚ Jkl k l , 1 2 ͗i,j͘ 2 ͗͗ k,l͘͘ It is proposed to analyze the transition in NaV2O5 . This material provides a unique example of a correlated electron where the bold variables denote lattice vectors, the first ͑sec- system, where the interplay of charge and spin degrees of ond͒ sum includes only nearest neighbors ͑next-nearest freedom results in a phase transition into a phase with coex- neighbors͒ of the lattice, respectively. Spins along the sides istent spin gap and charge order. NaV2O5 is the only known of an elementary plaquette interact via the NN coupling Jᮀ , so far quarter-filled ladder compound. Each individual rung while spins along plaquette’s diagonals interact via the NNN ϭ ͑ ͒ of a ladder is occupied by single electron which is equally couplings Jkl J1,2 see Fig. 1 . The way we defined the distributed between its left/right sites in the disordered phase. Hamiltonian corresponds to antiferromagnetic couplings for ϭ At Tc 34 K this compound undergoes a phase transition JᮀϾ0 and ferromagnetic for JᮀϽ0. when a spin gap opens, accompanied by charge ordering.7 Ground-state phases The problem of the in NaV2O5 localized on the rungs of the 2D array of ladders is mapped onto the coupled There is no exact solution of the model ͑1͒. Its possible spin-pseudospin model on an effective square lattice. The ordered phases and critical properties have been studied 724 Low Temp. Phys. 31 (8–9), August–September 2005 G. Y. Chitov and C. Gros within various approaches for equal diagonal couplings J1 ϭ ͑ J2 see Ref. 1 for a review and references on the original literature͒. We will consider arbitrary Ising couplings ͑ ͒ (J1 ,J2 ,Jᮀ), so the model 1 can be either frustrated or not ͑see footnote 1͒. The ground-state phase diagram can be found from energy arguments, as was first done by Fan and 8 ϭ ͑ ͒ Wu for J1 J2 . Their phase diagram is shown in Fig. 3c. From direct counting of the ground-state energies of possible spin arrangements we construct the phase diagram for J1  J2 . Along with the phases found by Fan and Wu— ferromagnetic, antiferromagnetic, and super- antiferromagnetic—there is a fourth phase which can occur if J1 and J2 have opposite signs. The name of the SAF phase comes from viewing it as two superimposed antiferromag- netic lattices ͑one lattice of circled sites and another of squared sites in Fig. 1͒. In the SAF state there are two frus- trated bonds Jᮀ per plaquette and its energy is fourfold de- generate, since each of the superimposed lattices can be flipped independently. In addition to the two SAF states with alternating ferromagnetic order along the horizontal chains ͑one of these is shown in Fig. 1͒, there are two states with the vertical ferromagnetic order. The new fourth phase shown in Fig. 1 can be viewed as two superimposed lattices, each of which is ordered ferro- FIG. 2. Phase diagram of the ground states ͑GS͒ of the model ͑1͒. For ͑ Ͻ Ϫ ϫ Ϫ ϫ Ϫ magnetically along one side e.g., J2 0) and antiferromag- visualization purposes it is drawn within the ( 1,1) ( 1,1) ( 2,2) par- ϩ ϭ͉ ͉ netically along the other ͑e.g., J Ͼ0). So we will call it allelepiped. The SAF GS lies between the frustration planes J1 J2 Jᮀ . 1 ϭ͉ ͉ ͑ super-ferro-antiferromagnetic ͑SFAF͒.9 The SFAF state also The SFAF GS lies between the frustration planes J1 Jᮀ . The AF GS FM GS͒ lies above ͑beneath͒ the frustration planes and above ͑beneath͒ the basal has two frustrated plaquette’s bonds and fourfold degeneracy. plane in the third quadrant, respectively. The second quadrant ͑not shown͒ is ϭ The ordering pattern shown in Fig. 1 changes only by a obtained by a reflection in the J1 J2 plane. Different sectors shown by ϭ lattice spacing shift over flipping of the sublattices. The di- hatched and twiggy lines on the exactly solvable planes Jᮀ 0 are explained rection of the ferromagnetic order is determined by the fer- in the text. romagnetic diagonal. trated ͑diagonal͒ bonds per plaquette is two in the first quad- The ground-state phases in the space (J1 ,J2 ,Jᮀ) are shown in Fig. 2. In order to facilitate perception of this pic- rant, one in the second and the fourth, and zero in the third. ture, we also present in Figs. 3 and 4 several plane projec- Transitions at finite temperature should be absent on the Ͼ frustration planes FP/FPЈ where the model is highly degen- tions of the 3D Fig. 2. In the first quadrant (J1 ,J2 0) in the ϩ Ͼ͉ ͉ erate. We are not aware of studies of the ground state in these region J1 J2 Jᮀ lying between two frustration planes ͑FP͒ cases and cannot say at the moment whether the system pos- sesses some kind of a long-range order at zero temperature or ϩ ϭ͉ ͉ ͑ ͒ ϭ ϭ Ͻ Ј J1 J2 Jᮀ :FP 2 not, except a rather trivial line Jᮀ J1 0,J2 0 of the FP planes crossing where the model becomes a set of decoupled the ground state of the model is SAF. A continuous transition Ising chains, and four special lines on the FP planes where it from the paramagnetic ͑PM͒ to the SAF phase occurs at becomes the exactly-solvable isotropic triangular Ising ͑ITI͒ Ͼ some critical temperature Tc 0. From the arguments known model. The latter case will be discussed momentarily. ϭ ͑ for the case J1 J2 Ref. 12; see also Ref. 13 for a more general symmetry analysis of the Ginzburg-Landau func- Exactly-solvable limits tional͒ this transition is nonuniversal: the critical indices de- In the 3D space (J1 ,J2 ,Jᮀ) besides the frustration pend continuously on the couplings Jᮀ ,J1 ,J2 . planes FP and FPЈ, there are three special planes where one Ͼ Ͼ In the region J1 Jᮀ of the fourth quadrant (J1 0,J2 of the couplings is zero. On these planes the model ͑1͒ re- Ͻ 0) lying between the other pair of frustration planes duces to the exactly solvable cases. ϭ͉ ͉ ͑ ͒ J1 Jᮀ :FPЈ 3 the ground state is SFAF. The same arguments12,13 suggest nonuniversality of the PM→SFAF transition. In the regions lying above ͑beneath͒ the frustration planes FP and FPЈ, and above ͑beneath͒ the basal plane Jᮀ ϭ0 in the third quadrant, the ground state is a usual AF ͑FM͒, respectively. The transition PM→AF ͑FM͒ belongs to the 2D Ising universality class. The second quadrant J1 ͑ ͒ Ͻ Ͼ FIG. 3. Plane projections of the 3D diagram shown in Fig. 2. a : Upper part 0,J2 0, not shown in Fig. 2, is obtained by reflection in JᮀϾ0, view from the top. ͑b͒: Lower part JᮀϽ0, view from the top. ͑c͒: ϭ ͑ ͒ ϭ ϵ Ј 8 the plane J1 J2 . In the AF FM state the number of frus- Compactification of 3D Fig. 2 in the special case J1 J2 J . Low Temp. Phys. 31 (8–9), August–September 2005 G. Y. Chitov and C. Gros 725

ϭ ͑ ϭ solvable ‘‘triangulation’’ planes J1 0 or J2 0) is the same as the AF ground state in the interior in this region of the phase diagram. The situation on the ‘‘triangulation’’ planes in the lower part (JᮀϽ0) of the phase diagram is exactly analogous to the upper part, with an obvious replacement AF↔FM. Note that the ground states change on the lines where the triangulation and frustration planes cross. To put it differ- ently, these are the lines of quantum phase transitions. The ͑ ͒ →͉ ͉Ϫ AF FM phase disappears in the limit J1 Jᮀ 0(J2 ϭ0). Also, the zero-temperature AF in-chain order ͑ATI ͑ ͒͒ →͉ ͉ϩ QC described above disappears in the limit J1 Jᮀ 0 ϭ ϭ (J2 0) as well. The ITI model J1 Jᮀ is disordered at any nonzero temperature ͑indicated as ITI ͑QC͒ in Fig. 2͒. Its ground state, albeit having finite entropy per site, possesses periodic ͑with a period of three lattice spacings͒ long-range order.16 FIG. 4. Plane phase diagram for the ratios of the couplings, corresponding ϭ Ͼ ͑ ͒ The basal plane Jᮀ 0 in Fig. 2 corresponds to the case to the 3D Fig. 2 at Jᮀ 0. The phase boundaries thick solid lines corre- ͑ ͒ spond to the frustrations planes FP and FPЈ. The thick dashed lines ͑online͒ when Hamiltonian 1 represents two decoupled identical indicate the boundaries of the incommensurate global minima locus (y NN Ising models residing on two superimposed lattices Ͼ1/2, xϾ1/2, yϽ1/4x) discussed in the text. The case JᮀϽ0 can be ob- ͑shown by circles and squares in Fig. 1͒. Diagonal couplings .J /͉Jᮀ͉,AF→FM in this figureۋtained by the substitutions J /Jᮀ 1,2 1,2 J1,2 are the NN couplings of these Ising models. This is the only exactly-solvable limit ͑labeled by 2ϫ2DI in Fig. 2͒ within the SFAF ͑or SAF͒ region of the phase diagram. In this limit the PM→SAF ͑or SFAF͒ phase transition enters Let us start with the upper part (JᮀϾ0) of the SAF region into the 2D Ising universality class.

Ͻ ϩ ͑ ͒ Jᮀ J1 J2 . 4 Incommensurate „floating… phase ϭ On the SFAF–SAF boundary J2 0 the model is equivalent So far we have discussed the ground-state phases of the to the anisotropic Ising model on a triangular lattice ͑ATI͒, model and the critical behavior on the boundaries of these 14–17 for which exact results are available. The antiferromag- phases with the disordered phase, as well as on the special Ͼ Ͼ netic (Jᮀ ,J1 0) ATI model with one strong bond J1 Jᮀ is planes ͑lines͒. However, there is also a possibility that order- 17 disordered at any nonzero temperature. It orders only at T ing into the ground-state phases of Fig. 2 happens not nec- ϭ0, i.e., it is quantum critical ͑QC͒. The highly degenerate essarily from the PM phase, but from some other one͑s͒ oc- ground state ͑however with a vanishing zero-temperature en- curring at nonzero temperature. A very simple analysis tropy per site͒ can be viewed as a 2D array of antiferromag- indicates that this indeed can take place in our model. ͑ netically ordered along the strong bond J1) correlated Fourier-transforming the Hamiltonian ͑1͒ we obtain ͑we set chains. The oscillating ͑with a period of four lattice spacings͒ the lattice spacing to unity͒ power-law decay of the spin-spin correlation function along 17 Jᮀ directions indicate on the preference of the ferromag- Hϭ͚ J͑q͒Tx͑q͒Tx͑Ϫq͒, ͑5͒ netic order along the ‘‘missing’’ diagonal J2 . This resembles q the SFAF state, however any couple of adjacent J chains is 1 ͑ ͒ϭ ͑ ϩ ͒ϩ ͑ Ϫ ͒ uncorrelated. We label this state occurring on two sectors of J q Jᮀ cos qx cos qy J1 cos qx qy ϭ ͑ ϭ ͑ ͒ the J1 0 or J2 0) planes as ATI QC on the phase dia- ϩJ cos͑q ϩq ͒, gram ͑Fig. 2͒. Since the critical behavior of the ATI model 2 x y ͉ ͉Ͻ ͉ ͉р␲ with two equal weak ferromagnetic bonds Jᮀ J1 is where q runs within the first Brillouin zone qx,y .At equivalent to the totally antiferromagnetic ATI model,17 the mean-field level, a minimum of J(q)inq space defines the x ATI ͑QC͒ state smoothly continues into the lower (JᮀϽ0) wave vector q0 of the critical freezing mode T (q0), i.e., the ͗ x ͘ϰ ϩ␸ part of the SFAF–SAF boundary. order parameter Tm cos(q0•m ) below a certain criti- ϭ Ͼ ͑ ↔ The sectors J2 0, J1 0 and 1 2) above the FP J1 cal temperature Tc . In different regions (J1 ,J2 ,Jᮀ) of the ϩ ϭ ͑ ͒ J2 Jᮀ correspond to the antiferromagnetic ATI model ground-state phase diagram Fig. 2 we find a minimum at Ͻ 17 F/Aϭ ␲ ␲ with one weak bond J1 Jᮀ . It is known to have only two q (0,0)/( , ) giving the FM/AF order parameter and → SAFϭ ␲ ␲ phases and to order at finite temperature. Tc(PM AF) as a two minima q1,2 ( ,0)/(0, ) giving two components of ͑ function of couplings is also known exactly. The sectors J2 the SAF order parameter. The latter represent two possible ϭ Ͻ Ͼ ͑ ↔ 0, J1 0, Jᮀ 0 and 1 2) correspond to the antiferro- ordering patterns of the SAF phase and via some transforma- magnetic ATI model which is not even frustrated, and tion can be related to magnetizations of the superimposed → Ͼ Ͼ → 12 Tc(PM AF) 0 at any Jᮀ 0. The PM AF transition in sublattices. ͒ The important point is that the positions of the the ATI model belongs to the 2D Ising class. So, except for commensurate ͑C͒ extrema q# (#ϭF, A, SAF) of J(q)do the SFAF–SAF boundary, the ordered phase on the exactly- not depend on the couplings. 726 Low Temp. Phys. 31 (8–9), August–September 2005 G. Y. Chitov and C. Gros

FIG. 6. Qualitative temperature phase diagram. The IC floating ͑Fl͒ phase lies beneath the dashed lines. Crosses indicate the borders of the IC minima ͑ ͒ Ͼ Ͻ ϭ locus. a : For fixed J1 0, J2 0. On the plane Jᮀ 0 the floating phase must be absent. We assume that it disappears smoothly at Jᮀϭ0, resulting in ͑ ͒ ͑ ͒ Ͼ ͉ ͉Ͻ a Lifshitz point L . b : The same for fixed J1 0, Jᮀ J1 .

only an algebraic long-range order; it is called floating.10 The FIG. 5. Positions of extrema of J(q) ͑5͒ in the Brillouin zone. Open sym- origin of the IC floating phase in our model is frustration bols connected to their bold counterparts by a reciprocal lattice vector. In- ͑competing interactions͒. Such phase is well known from Ϯ a,s ͑ ͒ ͑ ͒͑ Ͼ Ͻ commensurate extrema q 7 , 8 shown for the case J1 0, J2 0) another example of frustrated Ising model, i.e., the ANNNI ͑ ͒ exist if conditions 6 are satisfied. model which was intensively studied in the past.1,10,11 In that model the floating phase locks into the antiphase which has the wave vector qϭ(0,␲/2). ͑The antiphase is the analog of There are however two other pairs of extrema Ϯqs,a our SFAF phase.͒ The ANNNI model also provides an ex- which exist if ample showing that the mean-field ͑minimization͒ analysis does not work well in defining boundaries between the float- ͉ ͉Ͻ ͉ ͉ ͉ ͉ ͑ ͒ Jᮀ 2 J1 and/or 2 J2 . 6 ing and commensurate phases in 2D, and the extend of the ͒ These extrema lie on the diagonal of the Brillouin zone. qs,a IC phase is less than the mean field suggests.2 are generically incommensurated ͑IC͒ and depend on the We will not attempt to locate exactly the phase bound- couplings as aries at finite temperature in this study. Following Domany et al.13 in classification of the ordered phases by commensu- J s ϭϪ s ϭ ͩ Ϫ ᮀ ͪ ͑ ͒ rability p, i.e., the ratio of superstructure’s period and lattice qx qy arccos , 7 2J2 spacing along a given direction, we can label the phases as ;(1ϫ2 ͑or 2ϫ1ۋ2ϫ2; SAFۋ1ϫ1; AFۋfollows: F Jᮀ ͒ -4ϫ4.3 From mapping of the 2D IC-C phase transiۋqaϭϪqaϭarccosͩ Ϫ ͪ . ͑8͒ SFAF x y J 2 1 tion to the Kosterlitz–Thouless problem, it is established10,18 In Fig. 5 we show the positions of all extrema of J(q) within that there is no such ͑continuous͒ transition for commensu- the Brillouin zone. In the SFAF ground-state region ͑e.g., in rate phases with small p2Ͻ8. From this result with the pro- Ͼ Ͻ ͒ ͑ ͒ the fourth quadrant J1 0, J2 0 shown in Fig. 2 the pair of viso of continuity of phase transition s in the model, we extrema Ϯqa ͑8͒ gives global minima of J(q) ͑the other pair conclude that the IC floating phase cannot ‘‘spill’’ beyond the of solutions ͑8͒ Ϯqs, when it exists, corresponds to SFAF (pϭ4) ground-state region of the phase diagram, even ͒ ͉ a͉ϭ͉ a͉Ͼ␲ ͑ ͒ maxima , and qx qy /2. As we see, two IC modes if a naive mean-field analysis suggests that within the F, Ϯqa could give components of the SFAF ground-state order AF, and SAF regions there are some parts where it could be parameter’s wave vectors ϮqSFAFϭϮ(␲/2,Ϫ␲/2) only if possible ͑see Fig. 4͒. The only high-temperature phase in Jᮀϭ0. ͑In the second quadrant of the SFAF region when which the latter three regions have a common border is the Ͻ Ͼ s a J1 0, J2 0, the vectors q and q exchange their roles. disordered PM. ↔ Because of the J1 J2 symmetry, in the following, for con- Combining this with the known exact critical properties creteness, we will always discuss the fourth quadrant.͒ of the model on the special planes discussed above, we end The locus of the IC global minima does not coincide up with the qualitative finite-temperature phase diagram with the SFAF ground-state region, but overlaps with the shown in Fig. 6. Since on the plane Jᮀϭ0 the model ͑1͒ is neighboring FM, AF, and SAF phases. The minimum J just two decoupled Ising lattices, the floating phase must be Ϯ a ͉ ͉ϭ ( q ) is located between two planes Jᮀ 2J1 in the fourth absent. Mean-field arguments suggest that the floating phase quadrant, and in two regions of the half of the first quadrant disappears exactly at Jᮀϭ0, giving rise to a Lifshitz point Ͻ ͑ ͒ ϭϩ ϭϩ ͱ ͑ ͒ ͑ ͒ (J2 J1): i between Jᮀ 2J1 and Jᮀ 2 J1J2; ii L . Note that the floating phase does not appear if the diag- ͑ Ϫۋϩ . The regions of the IC minima in the other half (J2 onal couplings are equal even at the mean-field level see Ͼ ͑ ͒ ͒ J1) of the first quadrant as in the second quadrant are Fig. 4 , which agrees with known, more sophisticated analy- ↔ a→ s 1 obtained from the described above by J1 J2 , q q .On ses of this case. the plane phase diagram shown in Fig. 4 this locus is re- stricted by the lines yϭ1/2, xϭ1/2, yϭ1/4x, shown by the The model in transverse field dashed lines. So we can conclude that at finite temperature the model Now we turn to the analysis of the NN and NNN Ising possesses an IC phase and there is an IC–C phase transition Hamiltonian H ͑1͒ in the presence of a transverse field. The where the IC wave vector qa locks into one of the ͑commen- total Hamiltonian of the Ising model in transverse field surate͒ ground-state phase vectors. As in 2D the IC phase has ͑IMTF͒ reads Low Temp. Phys. 31 (8–9), August–September 2005 G. Y. Chitov and C. Gros 727

According to Fig. 4, it means that we choose the couplings to ϭ Ϫ⍀ T z ͑ ͒ HIMTF H ͚ i . 9 lie above the hyperbola yϭ1/4x. The first condition in ͑11͒ i SAF ensures that the couplings lie in the region where q1,2 ϭ ␲ ␲ The Ising operators are normalized to satisfy the spin algebra (0, )/( ,0) provide a global minimum of model’s spec- trum, so the phase with the IC solutions qa/s does not inter- ͓T ␣ T ␤͔ϭ ␦ ␧ T ␥ ͑ ͒ vene. The second in the above conditions stipulates that even i , j i ij ␣␤␥ i . 10 if Jᮀϭ0 we stay away from the planes where our model becomes the triangular Ising. A transverse field can generate There is no exact solution of the transverse 2D Ising model exotic temperature phases in that model. Such phases have even for the case of NN couplings only (J ϭJ ϭ0). The 1 2 been found4,5 in the particular case of the isotropic ͑antifer- ground-state phase diagram of the Hamiltonian ͑9͒ can be ͒ romagnetic͒ transverse triangular Ising model.4 analyzed from the known mapping of the d-dimensional From mapping between the quantum and classical Ising IMTF at zero temperature onto the (dϩ1)-dimensional Ising models we conclude that at zero temperature our IMTF with model at a given ͑nonzero͒ ‘‘temperature.’’ 3 In our case the the couplings satisfying ͑11͒ possesses a single QCP which 2D NN and NNN IMTF maps onto the 3D Ising model com- separates the SAF and PM ground-state phases. The mean prised of the 2D layers ͑1͒ coupled ferromagnetically in the field predicts a two-phase PM/SAF diagram. The critical third ͑Trotter͒ direction with the coupling J ϰϪln coth ⍀ T temperature T of the second-order PM–SAF phase transi- Ͻ0. For such a (2ϩ1)-dimensional model a mean-field c tion evolves smoothly from the QCP T (J/⍀ )ϭ0 to the analysis gives a qualitatively correct diagram of the ground- c cr asymptotic limit T (ϱ) of the classical model ͑see dashed state phases of the 2D IMTF.3 The new coupling J does not c T curve in Fig. 8, where gϵ(J ϩJ )/⍀). It is also known that bring any additional frustration to the 2D NN and NNN 1 2 the mean field gives a qualitatively correct phase diagram for model. Analysis of J (q ,q ,q )ϭJ cos q ϩJ(q ,q ), 3 x y T T T x y the IMTF when dу2.3 Thus we argue that the mean-field where J(q ,q ) is given by ͑5͒, shows that J does not x y T result shown by the dashed curve in Fig. 8 does represent the modify the domains of the global minima in the (J ,J ,Jᮀ) 1 2 phase diagram of the IMTF ͑1͒, ͑9͒, ͑11͒, while its quantita- space, adding only a trivial q ϭ0 third component to the T tive aspects, e.g., the exact value of the QCP, should be cor- two-dimensional vectors q# discussed above ͑cf. Fig. 5.͒ The rected via more accurate treatments. temperature phase diagram of the 3D Ising model with the ͑ spectrum J3(qx ,qy ,qT) if we label the phases according to the in-plane ordering pattern defined by the 2D vectors q#) looks similar to that shown in Fig. 6, with one very important III. COUPLED SPIN-PSEUDOSPIN MODEL distinction: the above-mentioned argument related to phase’s Now we turn to the analysis of the IMTF ͑1͒, ͑9͒ coupled commensurability p does not apply in 3D, so the IC region is to the quantum spins (S) residing on the same sites of the not restricted to lie above the SFAF phase, but can spill into lattice as the Ising spins do. The latter we will call pseu- the neighboring regions of the (J1 ,J2 ,Jᮀ) space. From the dospins from now on. Such coupled spin-pseudospin ͑or mean-field arguments, the IC region is given by the locus of spin-orbital͒ models emerge in various contexts, most nota- Ϯ a/s the IC minima J( q ) defined in the previous Section. So bly the Jahn–Teller transition-metal compounds19 or many the IC phase, instead of being locked between the special kinds of low-dimensional quantum magnets. For a recent Ј ϭ planes FP and J2 0 as shown in Fig. 6a and 6b, respec- short overview and more references, see Ref. 6. tively, can spread up to the locus boundaries shown by the Models of the type we study in the present work appear crosses. From the equivalence between the zero-temperature in analysis of the quarter-filled ladder compound NaV O . ͑ ͒ 2 5 d-dimensional IMTF quantum Ising and the The Hamiltonian of this material can be mapped onto a spin- ϩ (d 1)-dimensional classical Ising model, we infer that the pseudospin model with spins and pseudospins residing on ground-state phase diagram of the former on the plane the same rung of a ladder.20–22 The ladders form a 2D lattice. ⍀ ( /J,J#) should have the same structure as the described x  The long-range pseudospin order ͗Ti ͘ 0 represents physi- above (T,J#) diagram of the latter. So we expect the trans- cally the charge disproportionation between left/right sites on verse field to generate the IC ground-state phase not only in a rung below Tc . the SFAF region of the Ising coupling space, but also in the This system was analyzed on the effective triangular lat- neighboring parts of the F, AF, and SAF regions. From mini- tice shown by solid lines in Fig. 7a.20,22 However, in the case mization arguments the latter are restricted by the dashed of NaV2O5 the Ising couplings generated by Coulomb repul- lines on the plane diagram in Fig. 4 and by the crosses in Fig. Ͼ sion are antiferromagnetic and J1 Jᮀ . Since the triangular 6. Ising model with one strong side is disordered,17 one needs From analogies with the ANNNI model, we rather ex- ͒ an extra coupling (J2-diagonal to stabilize the observed pect this ‘‘IC region’’ to be filled with infinitely many com- 6 SAF long-range order. In our earlier study we took J1 into mensurate phases with different p,10,11 but detailed analysis account explicitly, while the other diagonal J2 was effec- of this question, as well as the full finite-temperature phase tively generated via the spin-pseudospin coupling.5͒ ͑ ͒ diagram of the transverse model 9 , need a separate study. In this work we take into account the Ising couplings In the rest of the paper we will be particularly interested Jᮀ ,J1 ,J2 between neighboring sites of the effective lattice, in the SAF region of the coupling space, and restrict our- as shown in Fig. 7a. Then such effective lattice can be selves to the couplings mapped onto the square lattice shown in Fig. 7b with the NN and NNN Ising couplings. For NaV O all J Ͼ0, so the 2 Ͻ ͑ ͒Ͼ ͑ ͒ 2 5 # Jᮀ 4J1J2 , J1 ,J2 0. 11 model is frustrated. As follows from the geometry of the 728 Low Temp. Phys. 31 (8–9), August–September 2005 G. Y. Chitov and C. Gros

T ͑ ͒ FIG. 7. 2D effective lattice of coupled ladders. A vertical line and a represent dot a single ladder and its rung where pseudospin mn and spin Smn not shown reside. In the region encircled by the dashed line the Ising couplings between pseudospins are indicated. Two pseudospins from the (mϪ1)-th and (m ϩ ͑ ͒ ␧ T x 1)-th ladders are coupled not only by J2 bold dashed line but also via the dimerization constant .The mn-ordering pattern shown corresponds to the SAF phase ͑a͒. The same after mapping on a square lattice ͑b͒.

ϰ␧ original NaV2O5 lattice, J2 is the weak diagonal and J1 is interaction , and the in-ladder spin-pseudospin interaction the largest coupling: ϰ␭. The former, in terms of the equivalent square lattice, linear over difference of the charge displacement operators Ͻ Ͻ ͑ ͒ J2 Jᮀ J1 . 12 T x on the NNN sites along the weak J2 diagonal, is respon- We assume Jᮀ to be small enough not only to lie beneath the sible for the simultaneous appearance of the SAF order and frustration plane ͑2͒, but to satisfy a more stringent condition the spin gap, as well as for the destruction of the IMTF ͒ ͑11͒. Then according to the above analysis, the IMTF with QCP.7 The latter, quadratic in the NNN charge operators these couplings has a two-phase ͑PE-SAF͒ diagram with a along the J1-diagonal, is responsible for the re-entrance. 6͒ QCP. In the literature on NaV2O5 its charge order is called With all these terms the total effective Hamiltonian reads the ‘‘zig-zag phase,’’ what characterizes the antiferroelectric order in a single ladder only. In fact, the two-dimensional HϭH ϩ ͓ ϩ␭T z T z IMTF ͚ SmnSm,nϩ1 J m,n m,nϩ1 long-range charge order in NaV2O5 is SAF. For a detailed m,n explanation of this point, including interpretation of the ex- ϩ␧͑T x ϪT x ͔͒ ͑ ͒ perimental crystallographic data on the charge order23 in mϩ1,nϩ1 mϪ1,n . 14 terms of the Ising pseudospins; see Ref. 24. The dimensionless couplings J,␭,␧ in the Hamiltonian ͑14͒ In the following we will work with dimensionless quan- are positive, and the spin operators satisfy the same algebra Hϭ ⍀ → ⍀ tities: Hamiltonians H/ , temperature T T/ , and ͑10͒ as the pseudospins ͑while S and T commute͒. The sums ϵ ⍀ Ising couplings g# J# / . With the site labeling shown in above run through 1рmрM and 1рnрN. For brevity we Fig. 7, the IMTF Hamiltonian is: ϵ will use the notation Dmn Smn•Sm,nϩ1 . In this study we 1 consider the model with the XY spin sector: H ϭ ͭ ϪT z ϩ ͓ ͑T x T x IMTF ͚ mm gᮀ mn mϩ1,nϩ1 m,n 2 ϭ x x y y ͑ ͒ Dmn SmnSm,nϩ1SmnSm,nϩ1 . 15 ϩT x xT x ͒ϩg T x T x mn mϩ1,n 1 mn m,nϩ1 The range of couplings under consideration will be restricted to ϩ T x T x ͔ͮ ͑ ͒ g2 mn mϩ2,nϩ1 . 13 ␭͒Շ ␧Շ ␭ ͑ ͒ ͑J, g1 , max J, . 16 For the decoupled spin sector of the total Hamiltonian Spin-SAF phase transition we take into account only the strongest coupling between spins on the NN rungs of a ladder. In terms of the effective We treat the Hamiltonian ͑14͒ following conventional lattice ͑cf. Fig. 7͒ this translates into a set of decoupled wisdom of molecular-field approximations ͑MFA͒.26 In the Heisenberg chains with the usual antiferromagnetic spin ex- present version of MFA the pseudospins are decoupled and ␳Tϰ Ϫ␤ T change J. These parallel chains are oriented along the averaged with the density matrix exp( hmn• mn), ͑ ͒ J1-diagonals. where hmn is the Weiss molecular field, while the spin sec- As we infer from our previous work on a simpler version tor is treated exactly via a Jordan-Wigner transformation. of the IMTF Hamiltonian,6 there are two spin-pseudospin The details are presented in Ref. 6. Similar to the pure interaction terms resulting in two qualitatively distinct as- IMTF with couplings ͑11͒ we assume the possibility of the pects of model’s criticality: the inter-ladder spin-pseudospin SAF order in the coupled model ͑14͒. So we take the follow- Low Temp. Phys. 31 (8–9), August–September 2005 G. Y. Chitov and C. Gros 729

ing ansa¨tze for the Ising pseudospin averages ͑i.e., the charge ordering parameters in terms of the real physical quantities͒ T z ϭ ͑ ͒ ͗ mn͘ mz , 17 T x ϭ͑Ϫ ͒mϩn ͑ ͒ ͗ mn͘ 1 mx . 18 It is easy to see from the Hamiltonian ͑14͒ that ansatz ͑18͒ creates a dimerization in the spin sector, therefore a natural assumption for the dimerization operator average is ϭϪ ϩ Ϫ mϩn␦ ͑ ͒ ͗Dmn͘ ͓t ͑ 1͒ ͔. 19 With the new coupling ϵ ϩ ͑ ͒ g g1 g2 20 FIG. 8. Critical temperature of the PE-SAF phase transition as a function of the molecular-field equations and results derived in Ref. 6 for the Ising coupling g at different values of J,␭,␧ from the numerical solution ϭ ϭ ͑ ϭ the case g2 gᮀ 0 i.e., g g1) can be applied here. Some of Eqs. ͑21͒. The dashed line corresponds to the pure IMTF (Jϭ␭ϭ␧ of them we reproduce in this paper in order to make it more ϭ0). Two stars on the abscissa show the positions of critical couplings ϭ ͑ ͒ ␭ϭ ͑ ͒ self-contained, and for the use in what follows as well. g␭ 2.0627 2.6366 for 0.1 1.0 , respectively. The large unfilled circles on the curves with ␧0 indicate the right boundary of the exponential BCS The average quantities are determined by the system of ͑ ͒ ͑ ͒ regime 25 . At large values of g not shown all curves Tc(g) approach the four coupled equations ϭ asymptotic line Tc g/4. ϩ ␭ ␤ 1 1 2 tmz n m ϭ tanh , ͑21a͒ z 2 n 2 considered in our analysis of the coupled model͒ where the ϩ ␧␩ ␤ mx g 2 n ͑ ͒ m ϭ tanh , ͑21b͒ Ising sector of 14 can order into, e.g., FM, AF, or SFAF x 2 n 2 phase, the dimerization ͑gap͒ in the spin sector does not oc- cur. ␲/2 2 ␸ 1 cos 1 ͑ ͒ ϭ ͵ ␸ ˜␤␰͑␸͒ϵ ͑⌬ ˜␤͒ ͑ ͒ The behavior of Tc(g) in the coupled model 14 shows t ␲ d ␰͑␸͒ tanh ␲ tn , , 21c 0 two striking new features in comparison to the pure IMTF: re-entrance and destruction of the QCP.6 In the absence of ⌬ ␲/2 sin2 ␸ ⌬ ␩ϭ ͵ d␸ tanh ˜␤␰͑␸͒ϵ ␩ ͑⌬,˜␤͒, the spin-charge coupling ␧, the model ͑14͒ has a QCP at ␲m ␰͑␸͒ ␲m n x 0 x ␭ ͑21d͒ ϵ ϩ ͑ ͒ g␭ 2ͩ 1 ␲ͪ , 24 ␩ϭͱ 2ϩ 2 where hx hz is the absolute value of the Ising mo- ͑ ͒ ␭ lecular field where Tc vanishes see Fig. 8 . renormalizes the QCP in comparison with the pure IMTF value gϭ2. The coupling ␧ ϭ ϩ ␭ ͑ ͒ hz 1 2 tmz , 22a responsible for the spin gap generation also destroys the ϭ ϩ ␧␦ ͑ ͒ QCP, resulting in the exponential behavior of T in the re- hx gmx 2 . 22b c gion gрg␭ , where the model would have been disordered at The other auxiliary parameters are defined as follows: any temperature if ␧ϭ0. This constitutes an important feed- ␦ϵ ␩ ͑ ͒ back from the spins on the charge degrees of freedom, allow- mx , 23a ing the very possibility of the model to order at all. Approxi- ␰͑␸͒ϵͱ 2 ␸ϩ⌬2 2 ␸ ͑ ͒ cos sin , 23b mate analytical solutions for Tc(g) in the regime of strong Ising couplings and the BCS regime are:6 ␧ 2 mx ⌬ϵ , ͑23c͒ Jϩ␭m2 g z , gӷg␭ , 4 ␤ T Ϸ ͑25͒ ˜␤ϵ ͑ ϩ␭ 2͒ ͑ ͒ c ˜ ␲˜ J mz . 23d Ά AJ J 2 ͫ Ϫ ͑ Ϫ ͒ͬ exp g␭ g , BCS regime, 2 4␧2 At some critical temperature T the coupled model under- c ϵ ͓␲ Ϫ␥ ͔Ϸ ␥Ϸ goes a phase transition. It is a second-order transition, with where A 8/ exp(1 ) 1.6685, 0.5772 is Euler’s the thermodynamic behavior of the physical quantities pre- constant, and dicted by the Landau theory of phase transitions.6 With the ␭ ͑ ͒ ␧ ˜JϵJϩ . ͑26͒ spin-pseudospin -charge coupling present, the SAF 4  ⌬ ϭ ␧ charge order mx 0 and the spin gap SG 2 mx appear The boundary where the low-temperature BCS regime sets in simultaneously below Tc . By analogy with the spin-Peierls transition, when the Peierls instability ͑freezing͒ cre- and the related formulas are applicable, is given approxi- ates the spin gap, it is natural to call this type of transition mately by the condition: the spin-super-antiferroelectric ͑spin-SAF͒ transition. 4␧2 It is worth noting an important property of the Hamil- BCS regime: gϽg␭ϩ . ͑27͒ tonian ͑14͒: in the other domains of Ising couplings ͑not ␲˜J 730 Low Temp. Phys. 31 (8–9), August–September 2005 G. Y. Chitov and C. Gros

͑ The BCS regime has many analogies with the standard In all regimes of couplings the re-entrance occurs at Tc theory of superconductivity, apart from the exponential de- Ͻ1/2.) By carrying out the leading-term expansions of the ͑ ͒ ͑ ͒ pendence of Tc on couplings. In particular, several physical functions in Eqs. 28 , 29 we obtain quantities ͑order parameter, BCS ratio, specific heat jump͒ ␭˜J manifest certain ‘‘universal’’ behavior near Tc , similar to ϭ ϩ ͑Ϫ ͒ϩ ͑ ͒ 6 g 2 4 exp 1/Tc 33 that known from the BCS theory. 4Tc 6 Another peculiarity of Tc(g) found earlier from the nu- ͑ for a single-valued function g(Tc). The nonmonotonic i.e., merical solution of Eqs. ͑21͒ is re-entrance in the intermedi- ͒ re-entrant behavior of Tc(g) is related to the existence of an ate regimeg gϳg␭ . The re-entrance occurs in the coupled extremum of g(Tc). The coupling gmin which defines the model with the QCP (␧ϭ0), while when ␧0 the critical ͑ minimal value of g for the order in mx to be possible in the temperature can even manifest a double re-entrant behavior pure IMTF with ␭ϭ0 this was the QCP͒ and which at the ͑ ͒ before it reaches the BCS regime see Fig. 8 . A detailed same time is the left border of the re-entrant region analysis of the coupled model in the regime of re-entrance Ͻ Ͻ ͑ ͒ was not done previously. We will address this in the next gmin g g␭ , 34 subsection, mainly analytically, in order to get more insight ͑ ͒ is defined from the minimum of the function g(Tc) 33 . on the underlying physics and, in particular, to establish con- This point corresponds to the critical temperature ditions when the re-entrance can occur. 16 ϭ Ϫ1 ␬ ␬ ϵ ͑ ͒ T ln o , o 35 * ␭˜J Re-entrance for which Let us first reproduce some earlier formulas6 for the 4 reader’s convenience. At TрT one equation from the pair Ϸ ϩ ͑ ϩ ␬ ͒ ͑ ͒ c gmin 2 ␬ 1 ln o . 36 ͑21a,b͒ can be written in the form o The consistency of the solution ͑35͒ with ͑32͒ implies the 4␧2 2␭ mϪ1ϭgϩ ␩ Ϫ t TрT ͑ ͒ condition z ␲͑ ϩ␭ 2͒ n ␲ n , c . 28 J mz 2 ϭ ͑ ͒ Ͻ ␬ Ͻ ͑ ͒ At T Tc we have Eqs. 21a as 2 ln o . 37 ˜J ␤ ␭ 1 c 2 mz m ϭ tanh ͩ 1ϩ t ͪ , ͑29͒ The other regime corresponds to the case when z 2 2 ␲ n Ͻ ͕ ˜ ͖ ͑ ͒ ␩ ͑ ͒ ⌬ϭ Tc min 1/2, J/2 . 38 and parameters tn , n are given by Eqs. 21c,d with 0. The latter two functions have the following expansions:6 Proceeding in the same way as above, we obtain for g(T ) in this case: ␲ 1 c xͩ 1Ϫ x2 ͪ ϩO͑x5͒, xϽ1, 4 4 ␭␲ ϭ ϩ ͑Ϫ ͒Ϫ 2 ͑ ͒ t ͑0,x͒Ϸ ͑30͒ g g␭ 4 exp g␭/2Tc Tc . 39 n ͭ 2 2 ␲ 1 3˜J 1Ϫ ϩO͑1/x4͒, xϾ1, x2 24 Let us point out that the conditions ͑32͒, ͑38͒ determine two and different regimes (xϽ1orxϾ1) of the asymptotics ͑30͒, ͑ ͒ ͑ ͒ ͑ ͒ 31 which we apply to obtain g(Tc)as 33 or 39 .Soif ␲ 1 ˜Ͻ ͑ ͒ xͩ 1Ϫ x2 ͪ ϩO͑x5͒, xϽ1, J 1 there are regions of Tc where condition 32 is satisfied, 4 12 then the approximation ͑33͒ applies. However, at sufficiently ␩ ͑ x͒Ϸ ͑ ͒ n 0, ͭ 2 31 Ͻ˜ ␲ 1 low temperatures (Tc J/2) we inevitably enter the other ln Axϩ ϩO͑1/x4͒, xϾ1. 48 x2 regime ͑38͒ where the asymptotics ͑30͒, ͑31͒ change (x ͒ ͑ Ͼ ۋ Ͻ 1 x 1), and the function g(Tc) crosses over from 33 ͑ ͒ ˜ ͑ ͒ Case ␧Ä0; re-entrance with QCP to 39 . If, on the contrary, J is large, then condition 32 never applies, and the approximation ͑39͒ describes the ͑ ͒ ͑ ͒ As one can easily see from Eqs. 28 , 29 there is no whole region T Ͻ1/2. ␭ϭ c re-entrance when 0. This is a well-known fact for the Extrema T of the function g(T ) ͑39͒ are determined * c pure IMTF, both on the mean-field level and beyond the by the transcendental equation MFA.3,26 To study the re-entrance analytically and in particu- ␭ ␭␲ lar, to establish whether there is some minimal value of 3 exp͑Ϫg␭/2T ͒ϭ T . ͑40͒ when it appears, we should distinguish between two * 2 * 3g␭˜J asymptotic regimes of the mean-field equations. Let us first ͑ ˜Ͻ This equation always has a trivial solution TЈ ϭ0 corre- consider the regime it can occur only if J 1) when (Tc * ϵ ␤ sponding to the ͑local͒ maximum of g(T ). This is the QCP, 1/ c) c and the curve Tc(g) approaches the QCP normally to the 1 1 abscissa ͑see Fig. 8͒. Two nontrivial solutions of ͑40͒ exist if ˜JϽT Ͻ . ͑32͒ 2 c 2 the couplings satisfy the condition Low Temp. Phys. 31 (8–9), August–September 2005 G. Y. Chitov and C. Gros 731

˜Ͼ ␭1/2 ͑ ͒ Case ␧Å0; double re-entrance, no QCP J C1 g␭ , 41 The absence of re-entrance at ␭ϭ0 can be proven rigor- where ously. Indeed, combining Eqs. ͑28͒ and ͑29͒, we obtain the equation ␲ e 3 C ϵͱ ͩ ͪ Ϸ0.3121. ͑42͒ 4␧2 J 1ϩ2m 1 24 3 Ϫ1ϭ ϩ ␩ ͩ z ͪ ͑ ͒ mz g ␲ n 0, ln Ϫ 46 J 2 1 2mz There is only one solution within the validity region of the m ෈͓ ͔ ͑ ͒ which has one and only one solution z 0,1/2 for a given approximation 39 , and it corresponds to the minimum of g ͑ ͒ value of . This solution in its turn provides a unique value g(Tc). If the couplings satisfy 41 then ͑ ͒ of Tc via Eq. 29 , and thus no re-entrance. At ␭0 continuous evolution of T (g) between the re- 1 c ϵ ϩ Ͼ ͑ ͒ gime of strong Ising coupling and the BCS regime ͓cf. Eq. u ln 3 ln ao 1, 43 3 ͑25͔͒ can occur either with a double re-entrance ͓i.e., with one minimum and one maximum of g(Tc)] within the re- 24˜J2 ϵ entrant region ao ␲␭ 2 , g␭ Ͻ Ͻ ͑ ͒ gmin g gmax , 47 and the minimum can be found analytically as or without re-entrance. In the latter case the function Tc(g) ͓ ͑ ͑ ͒͒ or g(Tc)] has only an inflexion point see Fig. 8 . g␭ 1 ln u Following the analysis given in the previous subsection, T Ϸ ͩ Ϫ ͪ . ͑44͒ * 6 u u2 we obtain for the case of small ˜J in the regime ͑32͒

␭˜JϪ2␧2 For the left border of the re-entrant region we obtain ϭ ϩ Ϫ ϩ ͑ ͒ g 2 4 exp͑ 1/Tc͒ . 48 4Tc ␭␲ 4␭␲ ϭ Ϫ 2 ϩ 3 ͑ ͒ Again, the re-entrant behavior is conditional on the existence gmin g␭ T T . 45 2 * 2 * 3˜J 3g␭˜J of a minimum of g(Tc). It exists if, at least,

The above equations agree well with the numerical solutions 1 ␧Ͻ ͱ␭˜JϷ0.7071ͱ␭˜J. ͑49͒ of the MFA ͑21͒ at different values of couplings ͑from com- & parison of the asymptotics ͑33͒, ͑39͒ and the numerical ͑ ͒ curves at various couplings and temperatures we found the The unique minimum of g(Tc) 48 defines the left border of deviations ϳ5% at most͒. More importantly, the analytical the re-entrance region gmin and corresponds to the critical ␬, andۋ temperature T given by Eqs. ͑35͒, ͑36͒ with ␬ results of this subsection allows us to understand in detail the * 0 interplay of the scales provided by a model’s couplings and 16 the temperature, resulting in the re-entrance. Let us explain ␬ϵ . ͑50͒ this on the example of two characteristic numerical curves ␭˜JϪ2␧2 shown in Fig. 8. The consistency imposes a constraint analogous to ͑37͒, ϭ ␭ϭ ␧ϭ ˜ϭ The curve shown for J 0, 1, 0(J 0.25) corre- more stringent than the ‘‘minimal requirement’’ ͑49͒. ˜ տ˜ ϭ sponds to the case of small J.AtTc J/2 0.125 it is well The other regime ͑38͒ is described by the approximation described by the equations for the regime ͑32͒. Its re-entrant ␲ 2 behavior and, in particular, the minimum g is due to the Tc min ϭ ϩ ͑Ϫ ͒Ϫ ͑␭˜ϩ␧2͒ ͑ ͒ g g␭ 4 exp g␭/2Tc J last term on the right-hand side of 33 . At lower tempera- ˜ 3 Շ ͑ ͒ 3J tures Tc 0.125 the asymptotics 33 is not applicable, the 1/2 curve is described by ͑39͒. Note that since C ␭ g␭ 2 1 4␧ A˜J Ϸ0.8229, the condition ͑41͒ for the minimum is broken, and Ϫ ln . ͑51͒ the asymptotics ͑39͒ describes the featureless low- ␲˜J 2Tc temperature evolution of this curve towards the maximum at As we have explained in the previous subsection for the case the QCP. ␧ϭ ͑ ͒ ˜ The second curve in Fig. 8 with Jϭ0.75, ␭ϭ1, ␧ϭ0 0, the asymptotics 48 is applicable only for small J at the intermediate temperatures ͑32͒, while ͑51͒ can be applied (˜Jϭ1) corresponds to the case of large ˜J. The whole re- Ͻ at arbitrary low temperatures, including the BCS region. The entrant region (Tc 0.4), including the position of the mini- 1/2 latter, given by the exponential dependence in ͑25͒, can be mum g (C ␭ g␭Ϸ0.8229) is described by the interplay min 1 recovered if we retain only the first and last ͑leading͒ terms of the last two terms on the right-hand side of Eq. ͑39͒. on the right-hand side of Eq. ͑51͒. Comparison of Eqs. ͑36͒, ͑45͒ also allows one to understand In the regime ͑38͒ the re-entrance occurs if the equation the more pronounced re-entrant behavior for the case of for extrema of ͑51͒ smaller J. As we see from our analysis of Eqs. ͑33͒, ͑39͒ in the ␲T 6␧2˜J2 both regimes ͑32͒, ͑38͒, the re-entrant behavior on the phase ͑Ϫ ͒ϭ * ͫ ͑␭˜ϩ␧2͒ 2 Ϫ ͬ ͑ ͒ exp g␭/2T J T 52 * 3 * ␲2 diagram occurs at any ␭0. 3g␭˜J 732 Low Temp. Phys. 31 (8–9), August–September 2005 G. Y. Chitov and C. Gros has nontrivial solutions. Note in making comparison of Eq. As follows from the inequalities ͑49͒, ͑53͒, ͑56͒ an in- ͑52͒ to its counterpart ͑40͒ at ␧ϭ0, that coupling ␧ destroys crease of ␧ can suppress the re-entrance even at ␭0. The the QCP,6 as is immediately seen from ͑51͒. So the trivial necessary conditions ͑49͒, ͑53͒ for extrema of the two as- solution T ϭ0 of Eq. ͑52͒ corresponds to the unphysical ymptotics ͑48͒, ͑51͒ are close, albeit with a rather small mis- * singularity of g. As follows from ͑52͒, ͑38͒, nontrivial solu- match of the coefficient. Conditions for re-entrance are more tions are possible if, at least, stringent, since they require the consistency between the so- lutions for extrema and the validity ranges of the appropriate ␲ 1 ͑ ͒ ␧ ϳ ͱ␭˜ ␧Ͻ ͱ␭˜JϷ0.8357ͱ␭˜J. ͑53͒ asymptotics. The overestimated critical value 0 0.7 J 2ͱ6 ͱ1Ϫ␲2/24 gives a good simple estimate for the boundary where the re-entrance disappears from the whole curve g(T ), whether If this condition is satisfied, the transcendental equation ͑52͒ c ˜JϾ ˜JϽ T g has at least one solution, corresponding to maximum of 1or 1. An example of the curve c( ) without re- g(T ). To leading order in ␧, it occurs at the temperature entrance is shown in Fig. 8. c To summarize our analysis of the re-entrance for the 1/2 ͱ6 ˜J cases ␧ϭ0 and ␧0: it reveals the robustness of this phe- TЈ Ϸ ͩ ͪ ␧͑54͒ ͑ ͒ * ␲ ␭ nomenon in the coupled model 14 and its underlying mechanism, namely, competition between different scales and the coupling defined by the couplings J,␭,␧,g and the temperature. These competing scales ͑interactions͒ are not related to the Ising 2 ͱ 4␧ A␲ ␭˜J frustration which is present in the model as well Ϸ Ϫ ͑ ͒ gmax g␭ ln . 55 ͓(Jᮀ ,J ,J )Ͼ0͔, since the latter is not accounted for ex- ␲˜J ͱ ␧ 1 2 2 6 plicitly by our mean-field equations. This competing mecha- If the couplings meet both the conditions ͑41͒ and nism for the re-entrance appears to be robust and not being an artifact of the MFA. Re-entrant phases due to competing 1/2 ␭ interactions are known also, e.g., from exact solution of the ␧ϽC g␭ͩ ͪ , ͑56͒ 27 2 Ising model on the Union Jack lattice, or from analyses of ˜J decorated Ising models.28 where It is not clear for us at the moment how the proposed re-entrance can be observed. NaV2O5 is, up-to-date, the only ␲ 47Ϫ13ͱ13 known compound with the spin-SAF transition and does not C ϵ ͱ Ϸ0.0936, ͑57͒ 2 2 36 show re-entrance. This is in agreement with our estimates for the parameters for the effective Hamiltonian for this com- then a second solution of ͑52͒ (TЉ ), corresponding to mini- * pound. They give its g located on the disordered side of the mum of g(Tc), exists. This minimum, located between ͑destroyed͒ QCP, and the re-entrance on the whole curve T (g) would be only very weak ͑i.e., localized near gϳg␭), g␭ c TЈ ϽTЉ Ͻ ͑58͒ if any. It appears experimentally that, e.g., external pressure * * 6 cannot modify the in-plane parameters of NaV2O5 strongly is given approximately by Eq. ͑44͒, where u in ͑43͒ is modi- enough, such that re-entrance would be generated. The varia- ۋ fied by a0 a, and tions of the interlayer couplings under pressure, on the other hand, generate various types of order ͑including devil’s stair- 24˜J3 case͒ with regard to the plane stacking, while the in-plane aϵ . ͑59͒ SAF order remains unaffected.24,29 2 2 ␲g␭͑␭˜Jϩ␧ ͒ For validity of expansion ͑44͒ we assume uϾ1. IV. SUMMARY AND DISCUSSION The analytical results of this subsection allows us to de- scribe the behavior of the coupled model at ␧0, following We study the 2D Ising model on a square lattice with from the MFA equations ͑21͒, both qualitatively and quanti- nearest-neighbor (Jᮀ) and unequal next-nearest neighbor ␧ϭ tatively. Two counterparts ( 0.1) of the numerical curves (J1,2) interactions. The cases of classical and quantum mod- discussed in the previous subsection are shown in Fig. 8. For els are considered. this case of small ␧, re-entrance is possible, according to We find the ground-state phase diagram of the classical ͑ ͒ ͑ ͒ conditions 49 , 53 . The re-entrant behavior at the tempera- Ising model at arbitrary Jᮀ ,J1 ,J2 . Along with the three or- տ Ј tures Tc T is not modified essentially by the presence of dered phases—ferromagnetic, antiferromagnetic, and SAF— * ␧ ␧ϭ ϭ 8  new coupling as compared to the case 0, and is in fact known for J1 J2 , in a more general case J1 J2 there is a controlled by couplings J,␭,g. Since we have already dis- region of the coupling space with the super-ferro- cussed it in detail for the case ␧ϭ0, we will not dwell on it antiferromagnetic ͑or 4ϫ4) ground-state phase and an in- any more. In the low-temperature regime ͑51͒ coupling ␧ commensurate phase at finite temperature, not reported be- changes drastically the behavior of g(T )atT ՇTЈ , creat- fore. The three phases—SAF, SFAF, IC—can occur only in c c * ing a maximum at g(TЈ ) described well by the approxima- the presence of competing interactions ͑frustrations͒ on the ͑ ͒ * tion 54 and turning g(Tc) away from the QCP towards the Ising model’s plaquette. BCS region. For the BCS regime Eq. ͑25͒ provides a virtu- A particularly interesting conclusion from the analysis of ally exact solution. the quantum model’s phase boundaries is that transverse field Low Temp. Phys. 31 (8–9), August–September 2005 G. Y. Chitov and C. Gros 733

⍀ can stabilize the IC ground-state phase ͓located for ⍀ separate study,24 is the 3D nature of the transition in ϭ ϭϪ 0 in the region with sign (J1 /J2) 1] in some parts of NaV2O5 . According to the correlation lengths 30 ϭ the AF and SAF regions of the coupling space where measurements, upon approaching Tc 34 K the 2D cross- Ͼ  (J1 ,J2) 0 but J1 J2 . These regions, along with vicinities over of the pretransitional structural fluctuations occurs of the special planes of degeneracy ͑triangulation and frus- somewhere at Tϳ50 K. The model considered in the present tration͒ in coupling space, are good candidates for the quan- work deals with a single plane, leaving aside the question of tum model to demonstrate a very nontrivial critical behavior. charge ordering along the third ͑stacking͒ direction. The Leaving this for a future work, we hope that our findings will phase transition in NaV2O5 quadruples the unit cell in the inspire additional interest in this model. Taking into account stacking direction, and the recent x-ray experiments, carried 23,33 only one simple example of a mapping shown in Fig. 7, it is out deep in the ordered phase revealed peculiar stacking  clear the model with J1 J2 is not so exotic. ordering patterns of the super-antiferroelectrically charge- We analyze the IMTF coupled to the XY spin chains in ordered planes. In addition, the pressure can change these ͑ ͒ patterns and even generate a multitude of higher-order com- the restricted SAF region of (Jᮀ ,J1 ,J2) where the IMTF has a simple two-phase ͑disordered-SAF͒ diagram with a mensurate superstructures in the stacking direction ͑devil’s 29 QCP, similar to that of the transverse NN model. Our interest staircase͒. To explain these phenomena we proposed a 3D in this model is motivated by the problem of the phase tran- extension of the Ising sector with additional competing cou- plings between the nearest and next-nearest planes.24 In the sition in the quarter-filled ladder compound NaV2O5 . The predictions of the mean-field equations for the critical prop- limit Jᮀ the Ising sector reduces to two identical interpen- erties of the coupled spin-pseudospin model do not differ etrating decoupled 3D ANNNI models. Although inclusion essentially from our earlier results for a simpler of the competing interlayer couplings accommodates the ex- Hamiltonian.6 Due to the spin-pseudospin coupling ␧, the planation for the observed stacking charge order in the ͑ ͒ QCP of the NN and NNN IMTF is destroyed, and in the framework of the spin-SAF in-plane mechanism of the whole SAF region ͑11͒ of Ising couplings the spin- transition in NaV2O5 , a deeper understanding of the critical pseudospin model undergoes the spin-SAF transition. We properties of a very complicated model with the 3D Ising should point out, that albeit the exponential BCS regime ͑25͒ sector warrants further study. We are grateful to D. I. Khomskii, E. Orignac, and P. N. on the disordered side of the IMTF QCP gϽg␭ formally Timonin for helpful discussions. This work is supported by extends up to gϭ0, decreasing g, i.e., J ϩJ , will eventu- 1 2 the German Science Foundation. G.Y.C. also acknowledges ally remove us from the coupling region ͑11͒ where the SAF support from a LURF grant. pseudospin solution of the MFA equations is applicable. We perform a detailed analytical study of re-entrance in the coupled model. In particular, we establish the conditions *E-mail: [email protected] 1͒ 1 when it can occur. The analytical results not only agree well We use the term ‘‘frustration’’ in a broad sense, meaning only that there is no spin arrangement on an elementary plaquette which can satisfy all with the direct numerical calculations in various regimes, but bonds. allows us to understand the physical mechanism of re- 2͒The well-studied ANNNI phase diagram with the FM and antiphase entrance due to interplay of competing interactions in the ground states1,10,11 is an analog of the lower part of the fourth quadrant ͑FM-SFAF͒ of our diagram in Fig. 2. coupled model. ͒ 3 According to this notation, the antiphase of the 2D ANNNI model is (1 In this work we gain more insights on the transition in ϫ4). For the same reasons we give for our case, the floating phase of that the spin-pseudospin model, and we can sharpen our previous model exists only within the antiphase ground-state region. 6 4͒ statements concerning the applications to NaV2O5 . The By analogy with the triangular Ising case of Refs. 4 and 5, we expect the present analysis of Ising sector allows us to identify the 2D transverse field to bring about new exotic phases near the frustration planes FP, FPЈ, where the ground state of our model is also infinitely degenerate. long-range charge order in that compound as the SAF phase. 5͒ ϰ␸ T ϯT ␸ A linear coupling ( 1 2) with some Gaussian mode results in an As follows from known results on the ordering of the frus- effective anti-/ferrro-magnetic interaction between T and T . Such terms 1 1 2 trated 2D Ising model into the SAF phase, the spin-SAF can create or renormalize the couplings of the Ising effective Hamiltonian. transition has non-universal coupling-dependent critical indi- In his context the Ising model with two NNN couplings of different signs ces. Experiments indicate rather wide regions of the two- is less academic than it might appear. 6͒Since in applications to NaV O the Ising pseudospins T represent charge dimensional structural ͑charge-ordering͒ fluctuations charac- 2 5 30–32 displacements, the appropriate names for the phases are ‘‘paraelectric’’ terized by the critical index ␤Ϸ0.17– 0.19, close to ␤ ͑PE͒ and ‘‘super-antiferrolelectric.’’ We keep the same abbreviation SAF ϭ1/8 of the 2D Ising model. Due to known difficulties in for the latter. 7͒ extracting critical indices from experimental data, it seems This type of coupling is allowed by the symmetry of the original NaV2O5 lattice.6 Numerical estimates of ␧ from a microscopic Hamiltonian are problematic to diagnose the deviations from universality ␧ Ͻ Ͻ ͑ given in Ref. 25. Note also that effectively couples the spin chains. caused by 0 Jᮀ /J1 1. Note that in the limit Jᮀ the PM- SAF transition enters into the 2D Ising universality class, and for NaV2O5 , due to its geometry, the ratio Jᮀ /J1 should 1 be small.͒ R. Liebmann, Statistical Mechanics of Periodic Frustrated Ising Systems, Springer, Berlin ͑1986͒. On the theoretical side, the critical indices of the PM- 2 S. Sachdev, Quantum Phase Transitions, Cambridge University Press, SAF transition as functions of couplings in the NN and NNN Cambridge ͑1999͒. Ising model have been calculated by various methods only at 3 B. K. Chakrabarti, A. Dutta, and P. Sen, Quantum Ising Phases and Tran- ϭ 1 sitions in Transverse Ising Models, Springer, Berlin ͑1996͒. J1 J2 . The critical exponents are unknown for the case 4 ͑ ͒  R. Moessner and S. L. Sondhi, Phys. Rev. B 63, 224401 2001 . J1 J2 , and it appears to be an interesting problem to study. 5 M. V. Mostovoy, D. I. Khomskii, J. Knoester, and N. V. Prokof’ev, Phys. Another very interesting issue we addressed recently in a Rev. Lett. 90, 147203 ͑2003͒. 734 Low Temp. Phys. 31 (8–9), August–September 2005 G. Y. Chitov and C. Gros

6 G. Y. Chitov and C. Gros, Phys. Rev. B 69, 104423 ͑2004͒. 22 M. V. Mostovoy, D. I. Khomskii, and J. Knoester, Phys. Rev. B 65, 7 P. Lemmens, G. Gu¨ntherodt, and C. Gros, Phys. Rept. 375,1͑2003͒. 064412 ͑2002͒. 8 C. Fan and F. Y. Wu, Phys. Rev. 179, 560 ͑1969͒. 23 S. Grenier, A. Toader, J. E. Lorenzo, Y. Joly, B. Grenier, S. Ravy, L. P. 9 Such a state, with the name of (4ϫ4) superstructure, was described be- Regnault, H. Renevier, J. Y. Henry, J. Jegoudez, and A. Revcolevschi, fore, e.g., by D. P. Landau and K. Binder, Phys. Rev. B 31, 5946 ͑1985͒. Phys. Rev. B 65, 180101͑R͒͑2002͒. It can also occur in the 2D Ising model when, along with NN and NNN, 24 G. Y. Chitov and C. Gros, J. Phys.: Condens. Matter 16, L415 ͑2004͒. the third-neighbor interactions are included. 25 C. Gros and G. Y. Chitov, Europhys. Lett. 69,447͑2005͒. 10 P. Bak, Rep. Prog. Phys. 45, 587 ͑1982͒. 26 R. Blinc and B. Zˇ ekzˇ, Soft Modes in Ferroelectrics and Antiferroelectrics, 11 W. Selke, Phys. Rep. 170,213͑1988͒. North-Holland, Amsterdam ͑1974͒. 12 S. Krinsky and D. Mukamel, Phys. Rev. B 16, 2313 ͑1977͒. 27 V. G. Vaks, A. I. Larkin, and Y. N. Ovchinnikov, Zh. E´ ksp. Teor. Fiz. 49, 13 E. Domany, M. Schick, J. S. Walker, and R. B. Griffiths, Phys. Rev. B 18, 1180 ͑1965͓͒Sov. Phys. JETP 22, 820 ͑1966͔͒. 2209 ͑1978͒. 28 E. H. Fradkin and T. P. Eggarter, Phys. Rev. A 14, 495 ͑1976͒. 14 G. H. Wannier, Phys. Rev. 79, 357 ͑1950͒; Errata: Phys. Rev. B 7, 5017 29 K. Ohwada, Y. Fujii, N. Takesue, M. Isobe, Y. Ueda, H. Nakao, Y. Waka- ͑1973͒. bayashi, Y. Murakami, K. Ito, Y. Amemiya, H. Fujihisa, K. Aoki, 15 R. M. F. Houtappel, Physica ͑Amsterdam͒ 16, 425 ͑1950͒. T. Shobu, Y. Noda, and N. Ikeda, Phys. Rev. Lett. 87, 086402 ͑2001͒. 16 J. Stephenson, J. Math. Phys. 11,413͑1970͒. 30 S. Ravy, J. Jegoudez, and A. Revcolevschi, Phys. Rev. B 59,681͑1999͒. 17 J. Stephenson, J. Math. Phys. 11,420͑1970͒. 31 B. D. Gaulin, M. D. Lumsden, R. K. Kremer, M. A. Lumsden, and 18 S. N. Coppersmith, D. S. Fisher, B. I. Halperin, P. A. Lee, and W. F. H. Dabkowska, Phys. Rev. Lett. 84, 3446 ͑2000͒. ͑ ͒ ͑ ͒ Brinkman, Phys. Rev. B 25, 349 1982 ; Phys. Rev. Lett. 46, 549 1981 ; 32 Y. Fagot-Revurat, M. Mehring, and R. K. Kremer, Phys. Rev. Lett. 84, ͑ ͒ Erratum: Phys. Rev. Lett. 46,869 1981 . 4176 ͑2000͒. 19 ͑ ͓͒ K. I. Kugel and D. I. Khomskii, Usp. Fiz. Nauk 136,621 1982 Sov. 33 S. van Smaalen, P. Daniels, L. Palatinus, and R. K. Kremer, Phys. Rev. B ͑ ͔͒ Phys. Usp. 25,231 1982 . 65, 060101 ͑2002͒. 20 M. V. Mostovoy and D. I. Khomskii, Solid State Commun. 113, 159 ͑1999͒. This article was published in English in the original Russian journal. Repro- 21 D. Sa and C. Gros, Eur. Phys. J. B 18, 421 ͑2000͒. duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 31, NUMBERS 8–9 AUGUST–SEPTEMBER 2005

Frustrated vortex in a two-dimensional antiferromagnet M. M. Bogdan*

B. Verkin Institute for Low Temperature Physics and Engineering, National Academy of Sciences of Ukraine, pr. Lenina 47, Kharkov 61103, Ukraine ͑Submitted April 6, 2005͒ Fiz. Nizk. Temp. 31, 968–973 ͑August–September 2005͒ The interaction of a magnetic vortex with the frustration created by a magnetic defect is investigated in a discrete Heisenberg model of a two-dimensional antiferromagnet with easy- plane anisotropic exchange. Numerical solutions are obtained for the static Landau- Lifshitz equations describing the spin distribution in a system with magnetic frustration and a vortex. It is found that the energy of the magnet is minimum in the case when the center of the vortex coincides with the position of the magnetic impurity. It is shown that as a result of the attraction between the vortex and frustration, a two-dimensional solitonic bound state localized at the magnetic defect—a frustrated vortex—arises in the magnet. The energy of such a vortex is lower than that of the free vortex, and this effect can be manifested in features of the behavior of the EPR linewidth in two-dimensional magnets. © 2005 American Institute of Physics. ͓DOI: 10.1063/1.2008133͔

I. INTRODUCTION stant ␭Ͻ1 corresponds to easy-plane anisotropy, and the constant ␩ is introduced to take into account a weak anisot- The class of quasi-two-dimensional antiferromagnets ropy in the plane: 1Ϫ␩Ӷ1Ϫ␭Ӷ1. It will be understood that comprising metalorganic compounds can evidently be the results obtained below are applicable primarily to yttrium supplemented by high-T superconductors in the magneti- c compounds, for which ␭Ϸ0.99, whereas for lanthanum com- cally ordered phase.1 The CuO planes in high-T supercon- 2 c pounds it is important to take the Dzyaloshinski–Moriya in- ducting compounds such as YBa Cu O ϩ and 2 3 6 x teraction into account as well. For yttrium compounds, since La Ϫ Sr CuO Ϫ are two-dimensional Heisenberg layers of 2 x x 4 y the spins of two nearest-neighbor copper planes interact an- antiferromagnetically coupled copper spins with a value ϩ tiferromagnetically with an exchange JЌ , we shall assume close to 1/2.2 The superexchange J between Cu2 ions takes Ϫ when studying the static configurations that the nearest spins place through the oxygen ions O2 . These almost isotropic in different planes are pairwise antiparallel. two-dimensional layers are coupled by a weak exchange JЌ The static spin distributions are solutions of the follow- ӶJ of the same order as the easy-plane magnetic 3 4 ing Landau-Lifshitz equations: anisotropy. It is known that the electronic state of the CuO2 planes and, hence, the character of the magnetic exchange ␦H 2ϩ ϫ ϭ ϭϪ ͑ ͒ between Cu ions depend substantially on the oxygen con- Sr Fr 0, Fr ␦ , 2 Sr tent in high-Tc superconductors. In particular, it has been established for lanthanum and yttrium superconductors that where Fr is the effective field at site r, and H is the Hamil- variation of the oxygen concentration leads to the formation tonian of the magnetic system. In the presence of a bond of ‘‘holes’’ in the CuO2 plane. These ‘‘holes’’ are charge defect ͑the exchange constant between two spins has a dif- 5 carriers in the superconducting phase, and in the opinion of ferent sign from the interaction in the host matrix͒, as a result 2 Aharony and co-authors, their localization at the oxygen in of frustration, the ground state can be nonuniform ͑in the XY the magnetic phase leads to a change in character of the model this is a threshold effect in the parameter JЈ/J).7 Such 2ϩ superexchange between the Cu ions, from antiferromag- a state, with a nonuniform distribution of the antiferromag- 3 netic with JϷ1000 K to ferromagnetic with JЈϷϪ3000 K. netic vector field localized near the defect is called a Villain 6 The resulting frustration destroys the long-range magnetic ground state.6 On the other hand, such a localized spin struc- order. ture can also be interpreted naturally as a two-dimensional The quantum description of these phenomena is a very magnetic soliton localized at the defect. Therefore, the term 7 laborious task, and for that reason the analytical approach magnetic frustration has come to denote a solitonic spin dis- usually turns to the framework of classical Heisenberg tribution arising near an effective frustrated bond.2 Such 8 models. In the Heisenberg XYZ model the magnetic inter- frustrations can influence the magnetic characteristics of actions in the copper planes in YBa Cu O ϩ are described ͑ ͒ 2 3 6 x high-Tc superconducting HTSC compounds, in particular, by a Hamiltonian of the form the temperature of the Ne´el phase transition in ͑ ͒ YBa2Cu3O6ϩx Ref. 9 , and can contribute to the suscepti- ϭ ͓ x x ϩ␩ y y ϩ␭ z z ͔ ͑ ͒ H0 J͚ SrSrϩa SrSrϩa SrSrϩa , 1 bility of the crystals. Direct experimental measurement of the r,a susceptibility of the CuO2 planes is difficult because of the on a square lattice, where Sr is the classical spin, the sum- large values of the exchange constants J and JЈ. However, it mation is over lattice sites and nearest neighbors, the con- has become possible to estimate the susceptibility of the

1063-777X/2005/31(8–9)/5/$26.00735 © 2005 American Institute of Physics 736 Low Temp. Phys. 31 (8–9), August–September 2005 M. M. Bogdan

CuO layers in an indirect way thanks to the discovery of a ϭ ϩ ϭ˜ ͑ ϩ ͒ ͑ ͒ 2 H H0 Hfr , Hfr JSh• Sr Sr , 3 new class of HTSC compounds containing rare-earth ͑R͒ 1 2 10,11 ˜ ions: GaBa2Cu3O6ϩx ,Nd2ϪxCexCuO4 , etc., in which where Sh is a hole spin, and the exchange J between the hole the rare-earth ions neighbor the CuO2 planes. A soliton ap- spin and a copper spin is assumed, for the sake of definite- proach to the treatment of questions related to the structure ness, to be antiferromagnetic, i.e., ˜JϾ0. ͑We recall that all of of magnetic frustrations in the CuO2 layers of HTSC com- the spins have been replaced by classical unit vectors, and pounds and their rare-earth analogs in the presence of mag- their absolute value Sϭ1/2 is subsumed in the renormaliza- netic field was proposed in Refs. 12 and 13; the contribution tion of the exchange constants.͒ This model explicitly takes of frustrations to the magnetic susceptibility of magnets was into account the interaction of the magnetic impurity with the investigated, and the magnetic fields induced by the field of a spins of the host and admits a transition in terms of this frustration in the rare-earth ion layers were found. Since the parameter to the defectless case, ˜J→0. We note that the characteristic interactions in the rare-earth layers are substan- static Landau-Lifshits equation for the hole spin, tially smaller than the exchange J in the CuO2 plane, the S ϫ͑S ϩS ͒ϭ0 ͑4͒ magnetic properties of the planes containing the Ga, Nd, etc. h r1 r2 ions are studied by conventional techniques. Susceptibility has the obvious solution in accordance with the fact that the measurements in compounds of the class R CuO have 2 4 hole spin is antiparallel to the sum of the copper spin vectors: revealed11 weak ferromagnetism, one of the causes of which S ϩS may be that frustrations having a magnetic moment are r1 r2 S ϭϪ . ͑5͒ present in the CuO2 planes. Since frustrations in the CuO2 h ͉ ϩ ͉ Sr Sr planes influence the magnetic ordering in the adjacent rare- 1 2 earth layers,13 one expects that frustration contributions will Therefore, in the proposed model the frustrated contribution also be manifested in magnetic resonance experiments. of the hole to the interaction between the statically distrib- On the other hand, it is well known that magnetic vortex uted copper spins is equal to H ϭϪ˜J͉S ϩS ͉. We note fr r1 r2 solitons can exist in two-dimensional isotropic and easy- that the effective interaction of the copper spins via the hole 14,15 plane magnets. In the case of weak anisotropy these vor- spins turns out to be of a ferromagnetic character indepen- tices have a structure with spin components that come out of dently of the initial sign of the interaction of the hole spin 16 the plane. Such solitons can contribute to the EPR line and copper spin ͑see also Ref. 2͒. Here we note that, unlike 17,18 broadening in quasi-two-dimensional magnets. Their in- the models using the frustrated bond approximation,2,8 which teraction with magnetic and nonmagnetic defects of substi- lead to results that are equivalent for ferro- and antiferromag- 19–23 tution has become a topic of research only recently. nets to within a change of the signs of the exchange con- In this paper we investigate the character of the interac- stants, the present model pertains specifically to antiferro- tion of a magnetic frustration and a vortex in an easy-plane magnets, since it is due to the inevitable frustrating influence antiferromagnet and predict the formation of their bound of an interstitial magnetic impurity on the antiferromagnetic state—a frustrated vortex. Such nonlinear excitations can ordering of the host spins. contribute to the resonance and thermodynamic characteris- The solution ͑5͒ of equation ͑4͒ suggests a numerical tics of quasi-two-dimensional magnets and HTSC com- algorithm that permits effective solution of the static equa- pounds in the magnetically ordered phase. tion ͑2͒ in the general case. An iterative method of solving the static Landau–Lifshitz equations for arbitrary values of II. TWO-DIMENSIONAL HEISENBERG MAGNET WITH A the spins S and arbitrary interactions governing the equilib- FRUSTRATING IMPURITY rium state of the system ͑in particular, spin distributions like ͒ The use of a planar model for describing magnetic two-dimensional magnetic frustrations and vortices is based ͑ ͒ frustration,2,8,12 raises the question of the correctness of re- upon the following idea. It follows from Eq. 2 that the placing the almost isotropic Heisenberg Hamiltonian custom- vector Sr should always be parallel to the effective field Fr , arily used for HTSC compounds by the XY model. The im- and the elementary iteration step can therefore be written as portant circumstance that calls the applicability of such an iϩ1ϭ i i ͑ ͒ Sr S•Fr /Fr , 7 approximation into question is this: two-dimensional i i solitons—magnetic vortices—can exist in easy-plane where Fr is the length of the vector Fr , and the index i is the magnets.14,16 If the anisotropy is not small, then the vortex number of the iteration. If the initial spin distribution if suf- spin configuration is planar. In the almost isotropic case, ficiently close in form to a magnetic frustration or vortex, which corresponds to the CuO plane, the vortex configura- then the iterative calculation converges very rapidly and 2 ͑ ͒ tion acquires spin components that come out of the plane. leads to a stable solution of Eqs. 2 . This naturally raises the questions of how the magnetic frus- trations will behave in an almost isotropic Heisenberg model, III. INTERACTION OF A MAGNETIC FRUSTRATION AND A and what will be the nature of the interaction of a frustration VORTEX IN A TWO-DIMENSIONAL ANTIFERROMAGNET and a magnetic vortex in such a system? To answer these questions, we formulate the following In this paper we report a numerical investigation of the classical model describing the interaction of the spin of a equilibrium spin configurations in the framework of the dis- ͑ ͑ ͒ ϫ hole and the spins of the copper in the CuO2 plane see also crete Heisenberg model 3 ona41 40 spin matrix (n,m) Ref. 7͒. The Hamiltonian ͑1͒ is supplemented with a term ͑the lattice constant is taken equal to unity͒. First we ob- appropriate to such an interaction: tained solutions describing magnetic frustrations; they turned Low Temp. Phys. 31 (8–9), August–September 2005 M. M. Bogdan 737

FIG. 3. Energy of an antiferromagnet containing a vortex and a frustration, measured in units of J, as a function of the distance between their centers.

ϫ FIG. 1. A 13 12 fragment of the spin distribution at a magnetic frustration. ͑ The positions of the central spins are the sites ͑21,20͒ and ͑21,21͒, between for the energy of a free vortex i.e., in the system without a which is located the spin of a magnetic impurity. frustrating impurity, Wϭ0) for a 41ϫ40 spin matrix gives ⌬ ϭ Ϫ ϭ Ev Ev E0 9.51. The next step consists in the calculation of the interac- tion of a magnetic vortex and a frustration in the framework out to be practically the same as in the planar model,2,8,12,13 of the proposed model. For this the problem is stated for the i.e., localized in the easy plane at for arbitrarily weak anisot- following physical situation: a hole is introduced into a sys- ropy ͑see Fig. 1͒. This circumstance permits a complete jus- tem of spins with a vortex at the center and is moved to tification of the use of the planar model for calculating the different distances from the center of the vortex. It was as- structure of the frustrations in the presence of magnetic sumed in the calculations that the anisotropy constants ␩ field.12,13 We note that in the purely isotropic case (␩ϭ␭ ϭ1 and ␭ϭ0.99, and the ratio of exchange constants W ϭ1), all the spins of the frustration also lie in a single plane, ϭ3. For the stable configuration obtained, the energy of the but there is degeneracy with respect to rotation of that plane system was found. Figure 3 shows the dependence of that around an arbitrary axis. In the frustration model with inter- energy on the distance R between the hole spin and vortex action ͑3͒ a nonuniform ground state arises at arbitrarily center. It is seen that the interaction is of an attractive char- ϭ˜ ϭ small values of the parameter W J/J. For W 0 the energy acter. It follows from an analysis of the spin distributions ϫ of the uniform antiferromagnetic ground state of a 41 40 obtained that at large distances between the hole spin and ϭϪ matrix of spins in units of J has the value E0 3199. The vortex center the presence of frustration is practically unno- energy of the system with frustration, Efr , initially falls off ticeable ͑see Fig. 4͒, and outwardly the vortex differs little quadratically as a function of the parameter W but then, after from the free vortex.16 At the same time, the vortex turns out Ϸ ͑ ͒ W 3, its decline becomes practically linear Fig. 2 . Thus to be strongly deformed in the energetically most favorable the energy of the system with frustration is lowered by an state, when the vortex center lies on the axis between two ⌬ ϭ Ϫ amount Efr Efr E0 , which is naturally called the self- copper spins, coinciding precisely with the position of the ϭ energy of a frustration. For example, for W 3 the energy hole spin ͑Fig. 5͒. Such a bound state of the magnetic vortex ⌬ ϭϪ Efr 2.08. and an impurity spin is naturally called a frustrated vortex. The proposed numerical method is efficient for calculat- The strong frustrating influence of the magnetic impurity on ing the structure of a vortex with components that come out the deviation of the components of the vortex from the easy from the easy plane, since in that case the spin distributions plane is clearly seen in Fig. 6, which shows the modulus of in the course of the iterations converge rapidly to the stable z the Sr components of the host spins. solutions. As the initial distribution for the vortex in the cal- ⌬ To find the energy E f v of a frustrated vortex, one must culations we used approximate analytical expressions for subtract from the total energy E ϭϪ3192.31 the energy easy-plane vortices from Ref. 16. The numerical calculation f v

z FIG. 4. Modulus of the projection Sr on the coordinates in an antiferromag- FIG. 2. Energy of an antiferromagnet with frustration Efr , measured in units net containing a vortex and a magnetic frustration in the case when the of J, as a function of the frustration interaction parameter W. impurity spin is located between sites ͑11,20͒ and ͑11,21͒. 738 Low Temp. Phys. 31 (8–9), August–September 2005 M. M. Bogdan

for this is that this broadening of the resonance line includes a contribution from magnetic vortices.17 An estimate of the energy of the vortices can be obtained from the temperature dependence of the EPR linewidth.18 In HTSC compounds EPR directly on the copper ions is not observed; this remains one of the unexplained mysteries of the physics of high- temperature superconductivity.24,25 This may be because of the very strong broadening of the resonance line in these compounds, which may include a contribution from mag- netic vortices. The resonance is observed in the rare-earth analogs of HTSC compounds,11 and the contribution of frus- trated vortices might be observed indirectly through the fields induced in the planes of the rare-earth ions, as in the simpler case of a replica of a magnetic frustration.13 It should be noted, however, that for further generalization of the re- ϫ FIG. 5. A 13 12 fragment of the distribution of the projections of spins in sults to the lanthanum compounds La2ϪxSrxCuO4Ϫy it is a frustrated vortex. The arrows indicate the components of the vectors Sr in necessary to take the Dzyaloshinski–Moriya interaction into the XY plane. The impurity spin is located in the middle between sites ͑21,20͒ and ͑21,21͒. account in a consistent way, and also the possible motion of the holes, which requires consideration of more-complex spin Hamiltonians.26,27 ϭϪ Efr 3201.08 of a nonuniform ground state of the system ⌬ with an isolated frustration. The energy obtained, E f v ϭ ⌬ IV. CONCLUSIONS 8.77, is less than the energy Ev of a free vortex by the amount of the binding energy, ⌬E ϭϪ0.74, which amounts b We have formulated a discrete classical Heisenberg to around 8% of the vortex energy. As is seen in Fig. 3, these model of a two-dimensional antiferromagnet with easy-plane calculations are equivalent to finding the binding energy in a anisotropy of the exchange with an interstitial magnetic im- frustrated vortex by subtracting from E ͑the total energy at f v purity. Such a model can describe the behavior of magnetic Rϭ0) the energy of the system with the magnetic frustration copper layers of the HTSC compound YBa Cu O ϩ with removed to a large distance and the free vortex. 2 3 6 x holes in the CuO planes. We have investigated the interac- Returning to the initial physical statement of the prob- 2 tion of a magnetic frustration and a magnetic vortex in the lem, we emphasize that introducing a hole into a CuO plane 2 framework of this model and obtained the following results. containing a free vortex leads to the formation of their bound 1. We have proposed an efficient algorithm for numerical state—a frustrated vortex, and lowers the energy of the sys- solution of the static Landau–Lifshitz equations for calcula- tem as a whole by an amount ⌬EϭϪ2.82, which consists of tion of the equilibrium stable spin configurations of magnetic the energy of a frustration and the binding energy. systems with an arbitrary character of the spin interactions. Thus, for given dimensions of the lattice ͑or density of 2. The solutions of the Landau–Lifshitz equations corre- vortices and holes͒ and values of the exchange interactions, a sponding to magnetic frustration and a magnetic vortex with numerical estimate of the energy of a frustrated vortex shows spin components coming out of the easy plane have been that the decrease of the system energy due to the introduction found numerically, and the energy characteristics of these of a hole into it and the localization of the vortex at the hole solutions have been calculated as functions of the model pa- is of the order of 30% of the energy of a free vortex. rameters. Such a decrease might be detected in EPR experiments 3. We have shown that the energy of a magnet contain- in layered magnets—rare-earth analogs of HTSC com- ing a vortex and frustration is minimum in the case when the pounds. Upon localization of the magnetic impurity at a site center of the vortex coincides with the position of the hole. in a metalorganic antiferromagnet one observes temperature As a result of the attraction between the vortex and frustra- broadening of the EPR line. One of the possible explanations tion a two-dimensional bound state localized at the magnetic defect—a frustrated vortex—appears. The energy of such a vortex is lower than that of the free vortex. 4. This effect can be detected in EPR experiments, since an estimate of the energy of magnetic vortices can be ob- tained from the temperature dependence of the resonance linewidth, to which these nonlinear excitations give an expo- nential contribution. Thus in two-dimensional antiferromagnetic systems with interstitial impurities ͑such as holes in HTSC compounds͒ the magnetic frustrations and vortices can form bound states having a substantial influence on the thermodynamics and FIG. 6. Frustrated vortex with center of localization at a magnetic defect. resonance properties of these systems. z The author thanks N. F. Kharchenko and J. M. Tran- The dependence of the modulus of the projection Sr on the coordinates is shown. quada for helpful discussions and advice. Low Temp. Phys. 31 (8–9), August–September 2005 M. M. Bogdan 739

*E-mail: [email protected] 14 A. M. Kosevich, B. A. Ivanov, and A. S. Kovalev, Nonlinear Waves of Magnetization. Dynamical and Topological Solitons ͓in Russian͔, ͑ ͒ 1 L. J. de Jongh and A. R. Miedema, Adv. Phys. A 23,1͑1974͒. Naukova Dumka, Kiev 1983 . 15 ͑ ͒ 2 A. Aharony, R. J. Birgeneau, A. Coniglio, M. A. Kastner, and H. E. Stan- A. A. Belavin and A. M. Polyakov, JETP Lett. 22, 245 1975 . 16 ley, Phys. Rev. Lett. 60, 1330 ͑1988͒. M. E. Gouvea, G. M. Wysin, A. R. Bishop, and F. G. Mertens, Phys. Rev. ͑ ͒ 3 J. M. Tranquada, A. H. Moudden, A. I. Goldman, P. Zolliker, D. E. Cox, B 39, 11840 1989 . 17 G. Shirane, S. K. Sinha, D. Vaknin, D. C. Johnston, M. S. Alvarez, A. J. F. Waldner, J. Magn. Magn. Mater. 31–34, 1203 ͑1983͒. 18 Jacobson, J. T. Lewandowski, and J. M. Newsam, Phys. Rev. B 38, 2477 F. Waldner, J. Magn. Magn. Mater. 54–57, 873 ͑1986͒. 19 ͑1988͒. C. E. Zaspel, T. E. Grigereit, and J. E. Drumheller, Phys. Rev. Lett. 74, 4 J. M. Tranquada, S. M. Heald, A. R. Moodenbaugh, and Xu Youwen, 4539 ͑1995͒. Phys. Rev. B 38, 8893 ͑1988͒. 20 C. E. Zaspel, C. M. McKennan, and S. R. Snaric, Phys. Rev. B 53, 11317 5 V. J. Emery, Phys. Rev. Lett. 58, 2794 ͑1987͒. ͑1996͒. 6 J. Villain, J. Phys. C 10, 4793 ͑1977͒. 21 K. Subbaraman, C. E. Zaspel, and J. E. Drumheller, J. Appl. Phys. 79,101 7 R. J. Gooding and A. Mailhot, Phys. Rev. B 44, 11852 ͑1991͒. ͑1996͒. 8 J. Vannimenus, S. Kirkpatrick, F. D. M. Haldane, and C. Jayaprakash, 22 K. Subbaraman, C. E. Zaspel, and J. E. Drumheller, Phys. Rev. Lett. 80, Phys. Rev. B 39, 4634 ͑1989͒. 2201 ͑1998͒. 9 M. M. Bogdan and A. S. Kovalev, Fiz. Nizk. Temp. 16, 1576 ͑1990͓͒Sov. 23 M. M. Bogdan and C. E. Zaspel, Phys. Status Solidi 189,983͑2002͒. J. Low Temp. Phys. 16,887͑1990͔͒. 24 F. Mehran and P. W. Anderson, Solid State Commun. 71,29͑1989͒. 10 H. Takagi, S. Uchida, and Y. Tokura, Phys. Rev. Lett. 62,1197͑1989͒. 25 P. Simon, J. M. Bassat, Y. Henrion, B. Rives, J. P. Loup, P. Odier, and S. 11 R. D. Zysler, M. Tovar, C. Rettori, D. Rao, H. Shore, S. B. Oseroff, D. C. B. Oseroff, Physica C 235–240, 1647 ͑1994͒. Vier, S. Schultz, Z. Fisk, and S-W. Cheong, Phys. Rev. B 44, 9467 ͑1991͒. 26 R. J. Gooding, N. M. Salem, R. J. Birgeneau, and F. C. Chou, Phys. Rev. 12 M. M. Bogdan, A. S. Kovalev, and A. A. Stepanov, Fiz. Nizk. Temp. 18, B 55, 6360 ͑1997͒. 838 ͑1992͓͒Low Temp. Phys. 18, 591 ͑1992͔͒. 27 R. J. Gooding and A. Mailhot, Phys. Rev. B 48, 6132 ͑1993͒. 13 A. S. Kovalev and M. M. Bogdan, Fiz. Tverd. Tela ͑St. Petersburg͒ 35, 1773 ͑1993͓͒Phys. Solid State 35,886͑1993͔͒. Translated by Steve Torstveit LOW TEMPERATURE PHYSICS VOLUME 31, NUMBERS 8–9 AUGUST–SEPTEMBER 2005

Fermionic versus bosonic descriptions of one-dimensional spin-gapped antiferromagnets S. Yamamoto* and K. Funase

Division of Physics, Hokkaido University, Sapporo 060-0810, Japan ͑Submitted January 31, 2005͒ Fiz. Nizk. Temp. 31, 974–983 ͑August–September 2005͒ In terms of spinless fermions and spin waves, we describe the magnetic properties of a spin-1/2 ferromagnetic-antiferromagnetic bond-alternating chain which behaves as a Haldane-gap antiferromagnet. On the one hand, we employ the Jordan–Wigner transformation and treat the fermionic Hamiltonian within the Hartree–Fock approximation. On the other hand, we employ the Holstein–Primakoff transformation and modify the conventional spin-wave theory so as to restore the sublattice symmetry. We calculate the excitation gap, the specific heat, the magnetic susceptibility, magnetization curves, and the nuclear spin-lattice relaxation rate with varying bond alternation. These schemes are further applied to a bond-alternating tetramerized chain which behaves as a ferrimagnet. The fermionic language is particularly stressed as a useful tool for investigating one-dimensional spin-gapped antiferromagnets, while the bosonic one works better for ferrimagnets. © 2005 American Institute of Physics. ͓DOI: 10.1063/1.2008134͔

I. INTRODUCTION strategy is still indispensable to low-temperature thermody- namics especially of spin-gapped antiferromagnets, where Haldane1,2 sparked renewed interest in one-dimensional grand canonical sampling is hardly feasible numerically. Heisenberg antiferromagnets, predicting that their low- Then we are led to describe massive spin chains in terms of energy structures should vary qualitatively depending on conventional languages such as the Jordan–Wigner fermions whether the constituent spins are integral or fractional. A and the Holstein–Primakoff spin waves. magnetic excitation gap immediately above the ground state, The Jordan–Wigner transformation is an efficient ap- which is referred to as the Haldane gap, was indeed observed proach to low-dimensional quantum magnetism. Spin-1/2 ar- in quasi-one-dimensional spin-1 Heisenberg antiferromag- rays with uniform42 and alternating43–45 antiferromagnetic ͑ ͒ nets such as CsNiCl3 Ref. 3 and Ni(C2H8N2)2NO2(ClO4) exchange interactions between nearest neighbors were thus ͑Refs. 4, 5͒. A rigorous example of such a massive phase was investigated and their energy structures, magnetization also given theoretically.6,7 Significant numerical efforts8–12 curves, and thermodynamic properties were indeed revealed were dovoted to detecting the Haldane gap in the higher-spin well. Two-leg antiferromagnetic spin ladders were also dis- systems. Competition between massive and massless phases cussed within this scheme46,47 and the interchain-coupling in low-dimensional quantum magnets was extensively stud- effect on the lowest-lying excitation was elucidated. More ied especially by the nonlinear-sigma-model quantum field refined fermionization was further proposed for coupled spin theory,13–23 and a wide variety of spin gaps-energy gaps in chains. Ordering spinless fermions along a snake-like path, magnetic excitation spectra-were further predicted. There Dai and Su48 succeeded in interpreting massive and massless followed stimulative findings, including quantized plateaux excitations with varying number of the ladder legs. Their in zero-temperature magnetization curves,24–26 gap forma- idea was generalized to investigate zero-temperature magne- tion in coupled spin chains27,28 and the dramatic crossover tization curves49 and thermodynamic quantities.50 In such from one- to two-dimensional quantum antiferromagnets,29 circumstances, we consider fermionizing an spin-1/2 and an antiferromagnetic excitation gap with a ferromagnetic ferromagnetic-antiferromagnetic bond-alternating chain, background.30–33 which converges to the spin-1 antiferromagnetic Heisenberg Besides the sigma-model study, analytic approaches chain as the ferromagnetic coupling tends to infinity and played a crucial role in revealing the nature of Haldane-gap therefore reproduces many of observations common to antiferromagnets. The valence-bond-solid model13,14 stimu- Haldane-gap antiferromagnets. lated considerable interest in matrix-product Bosonic theory has significantly been developed for one- representation34–40 of the Haldane phase. The Lieb– dimensional quantum magnets in recent years. While the Schultz–Mattis theorem41 was generalized24 to clarify a Schwinger-boson mean-field theory is unable to distinguish mechanism for gap formation in a magnetic field. However, fractional-spin chains from integral-spin ones, it is still use- these arguments were essentially restricted to the ground- ful in predicting the asymptotic dependence of the Haldane state behavior and can hardly be extended to finite- gap on spin quantum number51 and explaining quantum temperature properties. Numerical tools such as quantum phase transitions of Haldane-gap antiferromagnets in a Monte Carlo and density-matrix renormalization-group tech- field.52 The Schwinger-boson representation was further ap- niques are indeed useful for such a purpose, but an analytic plied to ferrimagnetic spin chains53,54 and ladders.55 It was a

1063-777X/2005/31(8–9)/8/$26.00740 © 2005 American Institute of Physics Low Temp. Phys. 31 (8–9), August–September 2005 S. Yamamoto and Kei-ichi Funase 741 major breakthrough leading to the subsequent development A. Fermionic approach 56–62 of the spin-wave theory in low dimensions that In accordance with the bond dimerization, we introduce 63 gave a spin-wave description of the one- Takahashi two kinds of spinless fermions through the Jordan-Wigner dimensional ferromagnetic thermodynamics introducing an transformation additional constraint on the number of spin waves. This modified spin-wave scheme was further applied to spin- nϪ1 nϪ1 50,64,65 ϩ ϭ † ͫ ␲ͩ † ϩ † ͪͬ gapped antiferromagnets and qualitatively improved S2nϪ1 an exp i ͚ amam ͚ bmbm , ϭ ϭ for one-dimensional ferrimagnets.66,67 The antiferromagnetic m 1 m 1 modified spin-wave theory is less quantitative than the ferri- n nϪ1 magnetic version,54,68 but it enlighteningly interpreted novel ϩ ϭ † ͫ ␲ͩ † ϩ † ͪͬ S2n bn exp i ͚ amam ͚ bmbm , observations such as the temperature dependence of the mϭ1 mϭ1 Haldane gap64,65 and the field dependence of the nuclear 1 1 spin-lattice relaxation rate.69 Such spin-wave understanding z ϭ † Ϫ z ϭ † Ϫ ͑ ͒ S2nϪ1 anan , S2n bnbn . 2 is well supported by other analytic descriptions.70–72 As for 2 2 finite-temperature calculation of spin-gapped antiferromag- Decomposing the fermionic Hamiltonian at the Hartree-Fock nets, the Schwinger-boson mean-field theory is of no use, level, we obtain a mean-field Hamiltonian as while the modified spin-wave theory maintains its validity to 54 H ϭ ϩ Ϫ ͒ a certain extent. Thus, we apply the modified spin-wave HF E0 ͑JAF JF scheme to the spin-1/2 ferromagnetic-antiferromagnetic N 1 1 bond-alternating chain with particular emphasis on a com- ϫ ͫͩ Ϫ ͪ † ϩͩ Ϫ ͪ † ͬ ͚ db anan da bnbn parison between fermionic and bosonic descriptions of spin- nϭ1 2 2 gapped antiferromagnets. N 1 Our theoretical attempt is much motivated by existent ϩ ͫͩ Ϫ ͪ † ϩ ͬ JAF ͚ pAF anbn H.c. bond-alternating chain compounds such as IPACuCl3 ͓IPA nϭ1 2 ϭ ϭ ͑ ͒ isopropylammonium (CH3)2CHNH3͔ Ref. 73 and ͓ ϭ N 1 (4-BzpipdH)CuCl3 4-BzpipdH 4-benzylpiperidinium Ϫ ͫͩ Ϫ ͪ † ͬ ϭ ͔ ͑ ͒ JF ͚ pF bnanϩ1H.c. C12H18N Ref. 74 . nϭ1 2 These materials behave as spin-1 Haldane-gap antiferro- N magnets at low temperatures,75–78 while such spin-1 features Ϫ ␮ ͑ † ϩ † ͒ ͑ ͒ Ј g BH ͚ anan bnbn , 3 are broken up into paramagnetic spin 1/2 s with increasing nϭ1 temperature.79–81 Besides the thermal crossover from quan- ϭ † ϭ † ϭ † tum spin 1Јs to classical spin 1/2Јs, their enriched ground- where da ͗anan͘HF , db ͗bnbn͘HF , pAF ͗bnan͘HF , pF 82–86 87 ϭ͗ † ͘ state properties and novel edge states are of great in- anϩ1bn HF , and terest to both theoreticians and experimentalists. 1 1 E ϭͫJ ͩͯp ͯ2Ϫd d Ϫ ͪ ϪJ ͩͯp ͯ2Ϫd d ϩ ͪ 0 AF AF a b 4 F F a b 4 II. FORMALISM ϩ ␮ ͬ ͑ ͒ g BH N, 4 We consider the ferromagnetic-antiferromagnetic bond- dimeric spin-1/2 Heisenberg chain, whose Hamiltonian is with ͗...͘ denoting the thermal average over the Hartree- given by HF Fock eigenstates. Defining the Fourier transformation as N Hϭ Ϫ ͒ ͚ ͓͑JAFS2nϪ1 S2n JFS2n S2nϩ1 1 ϭ • • ϭ ik͑nϪ1/4͒ n 1 an ͚ e ak , ͱN k Ϫ ␮ ͑ z ϩ z ͔͒ ͑ ͒ g BH S2nϪ1 S2n . 1 88–91 92 1 The ground-state properties and low-lying excitations ϭ ik͑nϩ1/4͒ ͑ ͒ bn ͚ e bk 5 of this model were well investigated by numerical tools and ͱN k variational schemes. In particular, the string order parameter originally defined for spin-1 Heisenberg chains93 was gener- and then a unitary transformation as 88–90 ␪ alized to this system and the breakdown of a hidden i k a uk vke ␣ ϫ 84,85 ͩ k ͪ ϭͩ ͪͩ k ͪ ͑ ͒ Z2 Z2 symmetry was extensively argued. As the ferro- Ϫ , 6 b i0k Ϫ ␤ magnetic coupling tends to infinity, the string order remains k vke uk k finite and the Haldane gap converges to that originating in where decoupled singlet dimers. On the other hand, the thermodynamic properties have 1 ␩ 87,94 u ϭͱ ͩ 1Ϫ ͪ , much less been calculated so far and there is no guiding k 2 ͱ␩2Ϫ͉␥ ͉2 theory for extensive experimental findings. Employing two k different languages, we calculate various thermal quantities 1 ␩ and give rigorous information on their low-temperature be- v ϭͱ ͩ 1ϩ ͪ , k 2 ͱ␩2ϩ͉␥ ͉2 havior. k 742 Low Temp. Phys. 31 (8–9), August–September 2005 S. Yamamoto and Kei-ichi Funase

B. Bosonic approach ␪ 1 1 Ϫ ␥ ϵ͉␥ ͉ei kϭJ ͩ Ϫp ͪ eik/2ϪJ ͩ Ϫp*ͪ e ik/2, k k AF 2 AF F 2 F Next we consider a single-component bosonic represen- tation of each spin variable at the cost of the rotational sym- 1 metry. We start from the Holstein-Primakoff transformation ␰ϭ ͑J ϪJ ͒͑d ϩd Ϫ1͒, 2 AF F a b ϩ ϭͱ Ϫ † S4nϪ4ϩ␶ 2S a␶:na␶:na␶:n , 1 ␩ϭ Ϫ Ϫ ͑ ͒ ͑JAF JF͒͑da db͒, 7 z ϭ Ϫ † 2 S4nϪ4Ϫ␶ S a␶:na␶:n , ϩ ϭ † ͱ Ϫ † and twice the lattice constant is set equal to unity, we can S4nϪ2ϩ␶ b␶n 2S b␶:nb␶:n, diagonalize the Hamiltonian as z ϭϪ ϩ † ͑ ͒ S4nϪ2ϩ␶ S b␶:nb␶:n , 15 H ϭ ϩ ͑␧ϩ␣†␣ ϩ␧Ϫ␤†␤ ͒ ͑ ͒ HF E0 ͚ k k k k k k , 8 ␶ϭ k where 1,2; that is, we assume the chain to consist of four sublattices. Under the large-S treatment, the Hamiltonian can where the dispersion relations are given by be expanded as ␧Ϯϭ␰Ϯͱ␩2ϩ͉␥ ͉2Ϫ ␮ ͑ ͒ HϭϪ ͑ ϩ ͒ 2 ϩ ϩ ϩH ϩH ϩ ͑ Ϫ1͒ k k g BH. 9 2 JF JAF S N E1 E0 1 0 O S , ͑ ͒ Ϯ 16 In terms of the fermion distribution functions ¯nk Ϯ i ␧ k T Ϫ where E and H give the O(S ) quantum corrections to the ϭ͓e k / B ϩ1͔ 1, the internal energy, the total magnetiza- i i ground-state energy and the dispersion relations, respec- tion, and the magnetic susceptibility are expressed as tively. The naivest diagonalization of the Hamiltonian ͑16͒, whether up to O(S1)oruptoO(S0), results in diverging ϭ ϩ ␧␴ ␴ ͑ ͒ E E0 ͚ ͚ k ¯nk , 10 sublattice magnetizations even at zero temperature. In order k ␴ϭϮ to suppress the quantum as well as thermal divergence of the number of bosons, we consider minimizing the free energy ϭ ␴Ϫ ͑ ͒ 56–58 M ͚ ͚ ¯nk N, 11 constraining the sublattice magnetizations to be zero: k ␴ϭϮ N ͑g␮ ͒2 ͑ † ϩ † ͒ϭ ͑ ͒ ␹ϭ B ␴͑ Ϫ ␴͒ ͑ ͒ ͚ ͚ a␶:na␶:n b␶:nb␶:n 4SN. 17 ͚ ͚ ¯nk 1 ¯nk , 12 nϭ1 ␶ϭ1,2 kBT k ␴ϭϮ Within the conventional spin-wave theory, spins on one sub- respectively. Another quantity of wide interest is the nuclear lattice point predominantly up, while those on the other pre- spin-lattice relaxation rate 1/T . Considering the electronic- 1 dominantly down. The condition ͑17͒ restores the sublattice nuclear energy-conservation requirement, the Raman process symmetry. In order to enforce the constraint ͑17͒, we first usually plays a leading role in the relaxation, which is for- introduce a Lagrange multiplier and diagonalize mulated as N ␲͑ ␮ ប␥ ͒2 1 4 g B N Ϫ HϭHϩ ␯ ͑ † ϩ † ͒ ͑ ͒ ϭ ͚ Em /kBTͯ͗mЈ͉͚ ͑A Sz JAF S ͚ ͚ a␶:na␶:n b␶:nb␶:n . 18 Ϫ e Ϫ ប⌺ Em /kBT n 2n 1 nϭ1 ␶ϭ1,2 T1 me m,mЈ n 2 We define the Fourier transformation as ϩB Sz ͉͒mͯ͘ ␦͑E ϪE Ϫh␻ ͒ ͑ ͒ n 2n mЈ m N , 13 1 ϭ Ϫik͑nϪ5/8ϩ␶/4͒ a␶:n ͚ e a␶:k , where An and Bn are the dipolar coupling constants between ͱN k ␻ ϭ␥ the nuclear and electronic spins, N NH is the Larmor ␥ frequency of the nuclei, with N being the gyromagnetic 1 ϭ ik͑nϪ1/8ϩ␶/4͒ ͑ ͒ ratio, and the summation ͚ is taken over all the electronic b␶:n ͚ e b␶:k , 19 m ͱN k eigenstates m with energy Em . Assuming the Fourier com- ponents of the coupling constants to have little momentum and then the Bogoliubov transformation as ͚ ϵ Ӎ ͚ ϵ dependence as n exp(ikn)An Ak A and n exp(ikn)Bn Bk ӍB, we obtain the fermionic expression of the Raman relax- ation rate as

1 ϭ␲ ␮ ␥ 2 ប 2 , ͑g Bh N͒ / N T1 ͑20͒ ϫ 2ϩ 2ϩ ␪ Ϫ␪ ͒ ͚ ͓A B 2AB cos͑ kЈ k ͔ k,kЈ where four times the lattice constant is set equal to unity. We ␺Ϯ H determine the coefficients i:k so as to diagonalize up to ␴ ␴ ␴ ␴ H ϫ ͑ Ϫ ͒␦͑␧ Ϫ␧ Ϫ ␻ ͒ ͑ ͒ order O(S) and take 0 into account in the calculation ͚ ¯nk 1 ¯nkЈ kЈ k h N . 14 95 ␴ϭϮ perturbationally. Then the Hamiltonian ͑16͒ is written as Low Temp. Phys. 31 (8–9), August–September 2005 S. Yamamoto and Kei-ichi Funase 743

FIG. 1. The spinless-fermion ͑SF͒, modified-spin-wave ͑MSW͒, quantum Monte Carlo ͑QMC͒, and numerical-diagonalization ͑Exact͒ calculations of the ground-state energy ͑the left͒ and the excitation gap immediately above the ground state ͑the right͒ for the bond-alternating dimerized chain, where Lϭ2N is the number of spins.

ϵ Ϫ Ϯ ϵ Ϯ ␴ ¯n␶:k , respectively, which are expressed as ¯n␶:k ¯nk E ϭϪ2J ͑1ϩ␥ϩ␯͒SNϩJ ͚ ͚ ␻ , ␴ ␴ 1 AF AF k J ␻ ϩ␦␻ k T Ϫ k ␴ϭϮ ϭ͓e AF( k k )/ B Ϫ1͔ 1. Here the Lagrange multiplier ␯ ϭ ⌬ ⌳Ϫ␥⌫͒Ϫ ϩ␥͒⌬2Ϫ␥⌫2Ϫ⌳2 is determined through E0 2JAF͓2 ͑ ͑1 ͔, ␴␥ 1ϩ␥ϩ␯ ͩ ϩ ͪ ͑ ϩ ␴͒ϭ ͑ ϩ ͒ H ϭ ͑␻ϩ␣† ␣ ϩ␻Ϫ␤† ␤ ͒ ͚ ͚ 1 ␹ ␭␴ 1 2¯nk 2N 1 2S . 1 JAF͚ ͚ k ␶k ␶k k ␶,k ␶,k , k ␴ϭϮ k ␶ϭ k k 1,2 ͑23͒

H ϭ ͑␦␻ϩ␣† ␣ ϩ␦␻Ϫ␤† ␤ ͒ Then the internal energy and the magnetic susceptibility are 0 JAF͚ ͚ k ␶:k ␶:k k ␶:k ␶:k 96 k ␶ϭ1,2 given by ϩH ϩH ͑ ͒ irrel quart , 21 ϭ ϩ ␻␴ ␴ E Eg 2͚ ͚ ˜ k ¯nk , ␥ϭ k ␴ϭϮ where JF/JAF, ␴ ͑ ␮ ͒2 ␻ ϭ ͱ͑ ϩ␥ϩ␯͒2Ϫ ϩ␥2ϩ ␴␹ 2 g B ␴ ␴ k S 1 1 2 k, ␹ϭ ͑ ϩ ͒ ͑ ͒ ͚ ͚ ¯nk ¯nk 1 , 24 3k T k ␴ϭϮ ␴␥ ϩ␥ϩ␯ B ␴ 1 ␦␻ ϭ͓͑⌳Ϫ␥⌫͒Ϫ͑1ϩ␥͒⌬͒]ͩ 1ϩ ͪ ϭϪ ϩ 2 ϩ ϩ k ␹ ␭␴ where Eg 2(JF JAF)S N E1 E0 . k k k ␴␹ ϩ␥ 2 III. CALCULATIONS ␥ϩ␴␹ k sin k 2 Ϫ␥͑⌬ϩ⌫͒ ␴ ϩ͑⌬Ϫ⌳͒ ␴ , ␭ ␴␹ ␭ First we calculate the ground-state energy Eg and the k k k lowest excitation gap Egap and compare them with numerical 1 ␴␥ 1ϩ␥ϩ␯ findings in Fig. 1. The spinless fermions are much better than ⌫ϭ ͩ ϩ ͪ Ϫ ͚ ͚ ͫ 1 ␹ ␭␴ 1ͬ the modified spin waves at describing both quantities. As J 4N k ␴ϭϮ k k F goes to zero, the fermionic findings are refined and end up k ␴ k with the exact values E /NϭϪ3J /4 and E ϭJ . The ͑1ϩ␥ϩ␯͒␴␹ cos2 Ϫ␥͑␥ϩ␴␹ ͒␭ sin2 g AF gap AF k 2 k k 2 modified spin waves considerably underestimate the spin ϫ , k gap. They can not distinguish massive spin chains from ␹2ϩ␥͑␥ϩ2␴␹ ͒sin2 massless critical ones54 to begin with, but they are still useful k k 2 in qualitatively investigating dependences of the Haldane k gap on temperature and spin quantum number.65 ␴␹ ϩ␥ sin2 1 k 2 Secondly we calculate the thermodynamic properties. ⌳ϭ ͚ ͚ ␴ , Figure 2 shows the temperature dependences of the zero-field 4N ␴ϭϮ ␴␹ ␭ k k k specific heat and magnetic susceptibility. Due to the signifi- 1 ␴␥ 1ϩ␥ϩ␯ cant underestimate of the spin gap, the modified spin-wave ⌬ϭ ͚ ͚ ͫͩ ϩ ͪ Ϫ ͬ ͑ ͒ 1 ␴ 1 , 22 description is much less quantitative than the fermionic one 4N ␴ϭϮ ␹ ␭ k k k at low temperatures. Furthermore, the modified spin waves ␭␴ϭ␻␴ ␹ ϭ͓ ϩ␥ϩ␯ 2Ϫ 2 ͔1/2 H with k k /S and k (1 ) sin (k/2) . irrel completely fail to reproduce the antiferromagnetic Schottky- H H and quart in 0 contain off-diagonal one-body terms such as type peak of the specific heat. Because of the Lagrange mul- ␣ ␤ ␶:k ␶:k and residual two-body interactions, respectively, tiplier ␯, which turns out to be a monotonically increasing both of which are neglected in the perturbational treatment. function of temperature, the dispersion relations ͑22͒ lead to ␣† ␣ ␤† ␤ At finite temperatures we replace ␶:k ␶:k and ␶:k ␶:k endlessly increasing energy and thus nonvanishing specific ͗␣† ␣ ͘ϵ ϩ ͗␤† ␤ ͘ by their canonical averages ␶:k ␶:k ¯n␶:k and ␶:k ␶:k heat at high temperatures. The spinless fermions succeed in 744 Low Temp. Phys. 31 (8–9), August–September 2005 S. Yamamoto and Kei-ichi Funase

FIG. 2. The spinless-fermion ͑SF͒, modified-spin-wave ͑MSW͒, and quantum Monte Carlo ͑QMC͒ calculations of the specific heat ͑the upper three͒ and the magnetic susceptibility ͑the lower three͒ as functions of temperature for the bond-alternating dimerized chain, where Lϭ2N is the number of spins.

reproducing the overall thermal behavior. The present ap- Next we consider the total magnetization as a function of proaches have the advantage of giving the low-temperature an applied field and temperature. We compare the fermionic behavior analytically. Equation ͑9͒ shows that the dispersion description of magnetization curves with numerical findings relation of the low-lying excitations reads in Fig. 3. The spinless fermions again work very well. Quan- ␧ϮӍϮ ϩ 2͒Ϫ ␮ ͑ ͒ tum Monte Carlo sampling becomes less and less feasible ͑Egap JAFvk g BH, 25 k with decreasing temperature, while we have no difficulty in ␮ Ͻ provided g BH Egap , where calculating Eq. ͑11͒ even at zero temperature. The ground- ϰ Ϫ 1/2 E ϭͱJ2 ˜p2 ϩJ2˜p2 ϩ2J J ˜p ˜p , state magnetization turns out to behave as M (H Hc) gap AF AF F F F AF F AF ␮ ϭ ͑ ͒ near the critical field g BHc Egap Ref. 97 . Magnetization ϭ ͑ ͒ 49 98 2Egapv JF˜pF˜pAF , 26 plateaus of multi-leg spin ladders and mixed spin chains ϭ Ϫ ϭ Ϫ are also well interpreted in terms of the spinless fermions. with ˜pF Re pF 1/2 and ˜pAF Re pAF 1/2. Then the low- temperature properties are calculated as On the contrary, in the modified spin-wave theory, the num- ber of sublattice bosons are kept constant and therefore we 2 C kBT Ϫ Egap Egap 3 have no quantitative information on the uniform magnetiza- Ӎͱ Egap /kBT ϩ ϩ ␲ e ͫͩ ͪ ͬ, NkB vJAF kBT kBT 4 tion as well as the staggered one. Though the Schwinger- boson mean-field theory,51,54,99,100 which consists of a rota- ␹ JAF JAF Ϫ Ӎͱ Egap /kBT ͑ ͒ tionally invariant bosonic representation, still works with an ͑ ␮ ͒2 ␲ e . 27 g B N vkBT applied field and/or existent anisotropy52,101,102 to a certain These features are found in the antiferromagnetic extent but rapidly loses its validity with increasing 54 Heisenberg two-leg ladder as well50,72 and can be regarded as temperature. common to spin-gapped antiferromagnets. The power-law Thus we are fully convinced that the spinless fermions prefactor to the activation-type temperature dependence, are superior to the modified spin waves in investigating which can hardly be extracted from numerical findings, is quantum and thermal properties of spin-gapped antiferro- essential in estimating the spin gap experimentally. magnets. Lastly in this section, we calculate the nuclear spin-

FIG. 3. The spinless-fermion and quantum Monte Carlo calculations of magnetization curves for the bond-alternating dimerized chain, where Lϭ2N is the number of spins. Low Temp. Phys. 31 (8–9), August–September 2005 S. Yamamoto and Kei-ichi Funase 745

FIG. 4. The spinless-fermion calculations of the nuclear spin-lattice relaxation rate as a function of temperature ͑the left͒ and an applied magnetic field ͑the right͒ for the bond-alternating dimerized chain.

N lattice relaxation rate 1/T1 in terms of the spinless fermions Hϭ ϩ ͒ in an attempt to stimulate further experimental interest in this ͚ ͓JAF͑S4nϪ3 S4nϪ2 S4nϪ2 S4nϪ1 ϭ • • system. If we again employ the approximate dispersion ͑25͒ n 1 at moderate fields and temperatures, k TӶE Ϫg␮ H, Eq. Ϫ ͑ ϩ ͔͒ ͑ ͒ B gap B JF S4nϪ1•S4n S4n•S4nϩ1 . 29 ͑14͒ can be further calculated analytically as Ϫ ͓ ϭ ϭ ͔ ͑ ␮ ប␥ ͒2 Cu(3 Clpy)2(N3)2 3-Clpy 3-chloropyridine C5ClH4N 1 g B N Ϫ 106 Ӎ ͑ ϩ ͒2 Egap /kBT ͑Ref. 105͒ is well described by this Hamiltonian and be- ␲ប A B e T1 2 vJAF haves as if it is a ferrimagnet of alternating spins 3/2 and 68 g␮ H ប␻ 1/2. In the conventional spin-wave scheme, the spin devia- ϫ B ͩ N ͪ ͑ ͒ ͗ † ͘ ͗ † ͘ cosh K , 28 tions in each sublattice, a a␶ and b b␶ , diverge in k T 0 2k T ␶:n :n ␶:n :n B B the antiferromagnetic ground state but stay finite in the fer- where K0 is the modified Bessel function of the second kind rimagnetic one. Without quantum divergence of the sublat- Ӎ Ϫ␥Ϫ Ͻ Ӷ ␥ and behaves as K0(x) ln 2 ln x for 0 x 1 with be- tice magnetization, it is not necessary to diagonalize the ef- ing Euler’s constant. Considering the significant difference fective Hamiltonian ͑29͒. In an attempt to keep the ប␻ between the electronic and nuclear energy scales ( N dispersion relations free from temperature, we may simply Շ Ϫ5 ប␻ Ӷ 10 J), there usually holds the condition N kBT.At diagonalise the original Hamiltonian ͑18͒ and then introduce 54 low temperatures, 1/T1 also exhibits an increase of the acti- a Lagrange multiplier so as to minimize the free energy. vation type but with logarithmic correction, which is much For ferrimagnets such an idea is much superior to the origi- weaker than the power correction in the case of the suscep- nal antiferromagnetic modified spin-wave scheme.56–58 tibility. Such a pure spin-gap-activated temperature depen- Figure 5 shows the thus-modified spin-wave calculations dence of 1/T1 , which is shown in Fig. 4, should indeed be as well as the Hartree–Fock calculations in terms of the spin- observed experimentally, unless magnetic impurities mask less fermions in comparison with numerical findings. The the intrinsic properties. Equation ͑28͒ further reveals a ferrimagnetic modified spin waves work very well, contrast- unique field dependence of 1/T1 : With increasing field,1/T1 ing with the antiferromagnetic ones. They reproduce the first decreases logarithmically and then increases exponen- Schottky-type peak of the specific heat and the ferrimagnetic tially, which is visualized in Fig. 4. The initial logarithmic minimum of the susceptibility-temperature product at inter- behavior comes from the Van Hove singularity peculiar to mediate temperatures. Although the modified spin-wave de- one-dimensional energy spectra and may arise more gener- scription of the antiferromagnetic increase of ␹T is some- ally from a nonlinear dispersion relation at the band bottom. what moderate, it converges into the paramagnetic value Therefore, besides spin-gapped antiferromagnets, one- S(Sϩ1)/3 at high temperatures. The description is more and dimensional ferromagnets and ferrimagnets may exhibit a more refined with decreasing temperature and is expected to similar field dependence.69,72,103,104 Relaxation-time mea- be accurate at sufficiently low temperatures.67 The T1/2 initial surements on spin-gapped chain antiferromagnets such as behavior of C and the TϪ2-diverging behavior of ␹ are both 66 IPACuCl3 and (4-BzpipdH)CuCl3 are strongly encouraged. correctly reproduced. Besides static properties, T1 measurements107 on a ferrimagnetic chain compound NiCu(C H N O )(H O ) 2H O was elaborately interpreted IV. BOND-ALTERNATING FERRIMAGNETIC CHAIN 7 6 2 6 2 3 • 2 in terms of the modified spin waves.108 Before closing our comparative study, we briefly men- On the other hand, the spinless fermions misread the tion a bond-alternating but ferrimagnetic chain calculated low-temperature properties of ferrimagnetic chains. A fatally within the same schemes. We take another interest in weak point of their description is the onset of a Ne´el-ordered the ferromagnetic-ferromagnetic-antiferromagnetic-antiferro- state. With increasing JF , the transition temperature Tc goes magnetic bond-tetrameric spin-1/2 Heisenberg chain, whose up and the applicability of the Hartree-Fock fermions is re- Hamiltonian is given by duced. Indeed the fermionic description is not so bad away 746 Low Temp. Phys. 31 (8–9), August–September 2005 S. Yamamoto and Kei-ichi Funase

FIG. 5. The spinless-fermion, modified-spin-wave, and quantum Monte Carlo calculations of the specific heat ͑the left͒ and the magnetic susceptibility ͑the right͒ as functions of temperature for the bond-alternating tetramerized chain, where Lϵ4N is the number of spins. The Hartree-Fock fermions encounter a paramagnetic-to-Ne´el-ordered phase transition with decreasing temperature and the transition temperature is indicated by arrows.

99,110–112 upward from Tc , but it is much less complementary to nu- dom in the same footing. The present naive fermi- merical tools in ferrimagnetic systems. onic representation is highly successful for spin-1/2 gapped antiferromagnets, including various bond-alternating and/or V. SUMMARY coupled chains. It is complementary to numerical tools espe- cially at low temperatures and allows us to readily infer both We have comparatively discussed fermionic and bosonic static and dynamic properties of spin-gapped antiferromag- descriptions of the bond-dimeric Heisenberg chain as an ex- nets. ample of spin-gapped antiferromagnets. The fermionic lan- The authors are grateful to H. Hori for valuable com- guage is based on the Jordan–Wigner spinless fermions ments. This work was supported by the Ministry of Educa- within the Hartree–Fock approximation, while the bosonic tion, Culture, Sports, Science and Technology of Japan, and formulation consists of constraining the Holstein–Primakoff the Iketani Science and Technology Foundation. bosons to restore the sublattice symmetry. The spinless fer- mions well describe both ground-state and finite-temperature *E-mail: [email protected] properties. The zero-field specific heat and magnetic suscep- tibility behave as Ϫ Ϫ ϰ 3/2 Egap /kBT 1 ͑ ͒ C ͑kBT͒ e F. D. M. Haldane, Phys. Lett. A 93,464 1983 . 2F. D. M. Haldane, Phys. Rev. Lett. 50, 1153 ͑1983͒. and 3W. J. L. Buyers, R. M. Morra, R. L. Armstrong, M. J. Hogan, P. Gerlach, Ϫ Ϫ and K. Hirakawa, Phys. Rev. Lett. 56,371͑1986͒. ␹ϰ͑ ͒ 1/2 Egap /kBT kBT e , 4J. P. Renard, M. Verdaguer, L. P. Regnault, W. A. C. Erkelens, J. Rossat- Mignod, and W. G. Stirling, Europhys. Lett. 3, 945 ͑1987͒. respectively, at sufficiently low temperatures, while the re- 5K. Katsumata, H. Hori, T. Takeuchi, M. Date, A. Yamagishi, and J. P. laxation rate behaves as Renard, Phys. Rev. Lett. 63,86͑1989͒. 6 Ϫ T. Affleck, T. Kennedy, E. H. Lieb, and H. Tasaki, Phys. Rev. Lett. 59, ϰ Egap /kBT ͑ ␮ ͓͒ 1/T1 e cosh g BH/kBT 0.80908 799 ͑1987͒. 7 Ϫ ͑ប␻ ͔͒ T. Affleck, T. Kennedy, E. H. Lieb, and H. Tasaki, Commun. Math. Phys. ln N /kBT 115, 477 ͑1988͒. 8 ͑ ͒ at moderately low temperatures and fields. On the other S. Yamamoto, Phys. Rev. Lett. 75, 3348 1995 . 9U. Schollwoeck and Th. Jolicoeur, Europhys. Lett. 30,493͑1995͒. hand, the modified spin waves give much poorer findings. In 10S. Yamamoto, Phys. Lett. A 213, 102 ͑1996͒. particular, they significantly underestimate the spin gap and 11Y.-I. Wang, S. Qin, and L. Yu, Phys. Rev. B 60, 14529 ͑1999͒. fail to reproduce the Schottky-type peak of the specific heat. 12S. Todo and K. Kato, Phys. Rev. Lett. 87, 047203 ͑2001͒. 13I. Affleck, Nucl. Phys. B 257, 397 ͑1985͒. The same schemes have further been applied to the bond- 14I. Affleck, Nucl. Phys. B 265, 409 ͑1986͒. tetrameric ferrimagnetic chain, where the modified spin 15G. Sierra, J. Phys. A: Math. Gen. 29,3299͑1996͒. waves work very well and are superior to the spinless fermi- 16G. Sierra, Phys. Rev. Lett. 77, 3443 ͑1996͒. 17 ons both qualitatively and quantitatively. The fermionic lan- T. Fukui and N. Kawakami, Phys. Rev. B 55, R14709 ͑1997͒. 18T. Fukui, M. Sigrist, and N. Kawakami, Phys. Rev. B 56,2530͑1997͒. guage is useful in describing disordered ground states and 19T. Fukui and N. Kawakami, Phys. Rev. B 56, 8799 ͑1997͒. their excitations, whereas the bosonic one in depicting or- 20T. Fukui and N. Kawakami, Phys. Rev. B 57, 398 ͑1998͒. dered ground states and their fluctuations. 21A. Koga, S. Kumada, N. Kawakami, and T. Fukui, J. Phys. Soc. Jpn. 67, ͑ ͒ The modified spin-wave theory is fully applicable to 622 1998 . 22K. Takano, Phys. Rev. Lett. 82, 5124 ͑1999͒. higher-spin systems. The Jordan–Wigner transformation can 23K. Takano, Phys. Rev. B 61, 8863 ͑2000͒. 109 also be generalized to higher-spin systems, where spin-1 24M. Oshikawa, M. Yamanaka, and I. Affleck, Phys. Rev. Lett. 78,1984 chains, for instance, are mapped onto an extended t – J model ͑1997͒. 25K. Totsuka, Phys. Lett. A 228, 103 ͑1997͒. of strongly correlated electrons. However, the double-graded 26 ͑ ͒ ϵ Ϫ † D. C. Cabra, A. Honecker, and P. Pujol, Phys. Rev. Lett. 79, 5126 1997 . Hubbard operators such as ˜cn,↑ (1 cn,↓cn,↓)cn,↑ demand 27E. Dagotto, J. Riera, and D. Scalapino, Phys. Rev. B 45, R5744 ͑1992͒. that we should treat the fermion and boson degrees of free- 28S. Gopalan, T. M. Rice, and M. Sigrist, Phys. Rev. B 49, 8901 ͑1994͒. Low Temp. Phys. 31 (8–9), August–September 2005 S. Yamamoto and Kei-ichi Funase 747

29E. Dagotto and T. M. Rice, Science 271, 618 ͑1996͒. 74S. A. Roberts, D. R. Bloomquist, R. D. Willett, and H. W. Dodgen, J. Am. 30A. K. Kolezhuk, H.-J. Mikeska, and S. Yamamoto, Phys. Rev. B 55, Chem. Soc. 103, 2603 ͑1981͒. R3336 ͑1997͒. 75H. Manaka, I. Yamada, and K. Yamaguchi, J. Phys. Soc. Jpn. 66,564 31G.-S. Tian, Phys. Rev. B 56, 5355 ͑1997͒. ͑1997͒. 32S. Maslov and A. Zheldev, Phys. Rev. B 57,68͑1998͒. 76M. Hagiwara, Y. Narumi, K. Kindo, T. Kobayashi, H. Yamakage, K. 33Y. Takushima, A. Koga, and N. Kawakami, Phys. Rev. B 61, 15189 Amaya, and G. Schumauch, J. Phys. Soc. Jpn. 66, 1792 ͑1997͒. ͑2000͒. 77H. Manaka, I. Yamada, Z. Honda, H. A. Katori, and K. Katsumata, J. 34W.-D. Feritag and E. Mo¨ller-Hartmann, Z. Phys. B 83, 381 ͑1991͒. Phys. Soc. Jpn. 67, 3913 ͑1998͒. 35A. Klo¨mper, A. Schadschneider, and J. Zittartz, Z. Phys. B 87, 281 78H. Manaka, I. Yamada, N. V. Mushnikov, and T. Goto, J. Phys. Soc. Jpn. ͑1992͒. 69, 675 ͑2000͒. 36A. Klo¨mper, A. Schadschneider, and J. Zittartz, Europhys. Lett. 24, 293 79H. Manaka and I. Yamada, Phys. Rev. B 62, 14279 ͑2000͒. ͑1993͒. 80H. Manaka, I. Yamada, M. Hagiwara, and M. Tokunaga, Phys. Rev. B 63, 37K. Totsuka and M. Suzuki, J. Phys. A: Math. Gen. 27, 6443 ͑1994͒. 144428 ͑2001͒. 38K. Totsuka and M. Suzuki, J. Phys.: Condens. Matter 7, 1639 ͑1995͒. 81H. Manaka, I. Yamada, Y. Miyajima, and M. Hagiwara, J. Phys. Soc. Jpn. 39S. Yamamoto, Phys. Lett. A 225, 157 ͑1997͒. 72, 2694 ͑2003͒. 40 S. Yamamoto, Int. J. Mod. Phys. B 12, 1795 ͑1998͒. 82K. Hida, Phys. Rev. B 45, 2207 ͑1992͒. 41 E. H. Lieb, T. Schultz, and D. J. Mattis, Ann. Phys. ͑N.Y.͒ 16, 407 ͑1961͒. 83K. Hida, Phys. Rev. B 46, 8268 ͑1992͒. 42 L. N. Bulaevskii, Sov. Phys. JETP 16, 685 ͑1963͒. 84M. Kohmoto and H. Tasaki, Phys. Rev. B 46, 3486 ͑1992͒. 43 L. N. Bulaevskii, Sov. Phys. JETP 17, 684 ͑1963͒. 85M. Yamanaka, Y. Hatsugai, and M. Kohmoto, Phys. Rev. B 48, 9555 44 V. M. Kontorovich and V. M. Tsukernik, Sov. Phys. JETP 26,687͑1968͒. ͑1993͒. 45 A. A. Zvyagin, Sov. Phys. Solid State 32, 181 ͑1990͒. 86K. Okamoto, D. Nishino, and Y. Saika, J. Phys. Soc. Jpn. 62,2587͑1993͒. 46 M. Azzouz, Phys. Rev. B 48,6136͑1993͒. 87K. Funase and S. Yamamoto, Phys. Lett. A 334,220͑2005͒. 47 M. Azzouz, L. Chen, and S. Moukouri, Phys. Rev. B 50, 6233 ͑1994͒. 88S. Takada, J. Phys. Soc. Jpn. 61, 428 ͑1992͒. 48 ͑ ͒ X. Dai and Z. Su, Phys. Rev. B 57, 964 1998 . 89K. Hida and S. Takada, J. Phys. Soc. Jpn. 61, 1879 ͑1992͒. 49 ͑ ͒ H. Hori and S. Yamamoto, J. Phys. Soc. Jpn. 71, 1607 2002 . 90K. Hida, J. Phys. Soc. Jpn. 62,439͑1993͒. 50 ͑ ͒ H. Hori and S. Yamamoto, J. Phys. Soc. Jpn. 73, 549 2004 . 91T. Sakai, J. Phys. Soc. Jpn. 64, 251 ͑1995͒. 51 ͑ ͒ D. P. Arovas and A. Auerbach, Phys. Rev. B 38,316 1988 . 92K. Hida, J. Phys. Soc. Jpn. 63, 2514 ͑1994͒. 52 ͑ ͒ H. Xing, G. Su, S. Gao, and J. Chu, Phys. Rev. B 66, 054419 2002 . 93M. den Nijs and K. Rommelse, Phys. Rev. B 40, 4709 ͑1989͒. 53 C. Wu, B. Chen, X. Dai, Y. Yu, and Z.-B. Su, Phys. Rev. B 60, 1057 94K. Hida, J. Phys. Soc. Jpn. 67, 1416 ͑1998͒. ͑ ͒ 1999 . 95S. Yamamoto, T. Fukui, and T. Sakai, Eur. Phys. J. B 15,211͑2000͒. 54 ͑ ͒ S. Yamamoto, Phys. Rev. B 69, 064426 2004 . 96 87 ͑ ͒ 55 M. Takahashi, Prog. Theor. Phys. Suppl. ,233 1986 . X. Y. Chen, Q. Jiang, and W. Z. Shen, J. Phys.: Condens. Matter 15, 915 97T. Sakai and M. Takahashi, Phys. Rev. 57, R8091 ͑1998͒. ͑ ͒ 2003 . 98S. Yamamoto and T. Sakai, J. Phys.: Condens. Matter 11, 5175 ͑1999͒. 56 ͑ ͒ M. Takahashi, Phys. Rev. B 40, 2494 1989 . 99A. Auerbach and D. P. Arovas, Phys. Rev. Lett. 61,617͑1988͒. 57 ͑ ͒ J. E. Hirsch and S. Tang, Phys. Rev. B 40, 4769 1989 . 100S. Sarker, C. Jayaprakash, H. R. Krishnamurthy, and M. Ma, Phys. Rev. B 58 ͑ ͒ S. Tang, M. E. Lazzouni, and J. E. Hirsch, Phys. Rev. B 40, 5000 1989 . 40 ͑ ͒ 59 , 5028 1989 . V. Y. Irkhin, A. A. Katanin, and M. I. Katsnelson, Phys. Rev. B 60, 1082 101C. J. De Leone and G. T. Zimanyi, Phys. Rev. B 49,1131͑1994͒. ͑ ͒ 1999 . 102M.-R. Li, Y.-J. Wang, and C.-D. Gong, Z. Phys. B 102,129͑1997͒. 60 ͑ ͒ S. Yamamoto and T. Nakanishi, Phys. Rev. Lett. 89, 157603 2002 . 103S. Yamamoto, J. Phys. Soc. Jpn. 69, 2324 ͑2000͒. 61 ͑ ͒ M. Kollar, I. Spremo, and P. Kopietz, Phys. Rev. B 67, 104427 2003 . 104H. Hori and S. Yamamoto, Phys. Rev. B 68, 054409 ͑2003͒. 62 ͑ ͒ H. Hori and S. Yamamoto, J. Phys. Soc. Jpn. 73, 1453 2004 . 105A. Escuer, R. Vicente, M. S. E. Fallah, M. A. S. Goher, and F. A. Mautner, 63 ͑ ͒ M. Takahashi, Phys. Rev. Lett. 58,168 1987 . Inorg. Chem. 37, 4466 ͑1998͒. 64 ͑ ͒ S. M. Rezende, Phys. Rev. B 42, 2589 1990 . 106M. Hagiwara, Y. Narumi, K. Minami, K. Kindo, H. Kitazawa, H. Suzuki, 65 ͑ ͒ S. Yamamoto and H. Hori, J. Phys. Soc. Jpn. 72, 769 2003 . N. Tsujii, and H. Abe, J. Phys. Soc. Jpn. 72, 943 ͑2003͒. 66 ͑ ͒ S. Yamamoto and T. Fukui, Phys. Rev. B 57, R14008 1998 . 107 113 ͑ ͒ 67 N. Fujiwara and M. Hagiwara, Solid State Commun. ,433 2000 . S. Yamamoto, T. Fukui, K. Maisinger, and U. Schollwo¨ck, J. Phys.: Con- 108H. Hori and S. Yamamoto, J. Phys.: Condens. Matter 16,9023͑2004͒. ͑ ͒ dens. Matter 10, 11033 1998 . 109C. D. Batista and G. Ortiz, Phys. Rev. Lett. 86, 1082 ͑2001͒. 68 ͑ ͒ T. Nakanishi and S. Yamamoto, Phys. Rev. B 65, 214418 2002 . 110Y. Suzumura, Y. Hasegawa, and H. Fukuyama, J. Phys. Soc. Jpn. 57, 2768 69 ͑ ͒ S. Yamamoto and H. Hori, J. Phys. Soc. Jpn. 73, 822 2004 . ͑1987͒. 70 ͑ ͒ Th. Jolicoeur and O. Golinelli, Phys. Rev. B 50, 9265 1994 . 111G. Kotliar and J. Liu, Phys. Rev. B 38, 5142 ͑1988͒. 71 ͑ ͒ J. Sagi and I. Affleck, Phys. Rev. B 53, 9188 1996 . 112D. Yoshioka, J. Phys. Soc. Jpn. 58,32͑1989͒. 72M. Troyer, H. Tsunetsugu, and D. Wo¨rtz, Phys. Rev. B 50, 13515 ͑1994͒. 73L. P. Battaglia, A. B. Corradi, G. Marcotrigiano, L. Menabue, and G. C. This article was published in English in the original Russian journal. Repro- Pellacani, Inorg. Chem. 19, 125 ͑1980͒. duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 31, NUMBERS 8–9 AUGUST–SEPTEMBER 2005

Superconductivity and normal state properties of non-centrosymmetric CePt3Si: a status report E. Bauer*

Institut fu¨r Festko¨rperphysik, Technische Universita¨t Wien, A-1040 Wien, Austria I. Bonalde

Centro de Fı´sica, Instituto Venezolano de Investigaciones Cientı´ficas, Apartado 21874, Caracas, 1020-A, Venezuela M. Sigrist

Theoretische Physik, ETH-Ho¨nggerberg, 8093 Zu¨rich, Switzerland ͑Submitted March 1, 2005͒ Fiz. Nizk. Temp. 31, 984–994 ͑August–September 2005͒

Ternary CePt3Si crystallizes in the tetragonal P4mm structure which lacks a center of inversion. Ϸ ͑ ͒ Antiferromagnetic order sets in at TN 2.2 K followed by superconductivity SC below Tc Ϸ Ј ϷϪ Ϸ 0.75 K. Large values of Hc2 8.5 T/K and Hc2(0) 5 T were derived, referring to Cooper pairs formed out of heavy . The mass enhancement originates from Kondo Ϸ interactions with a characteristic temperature TK 8 K. CePt3Si follows the general features of correlated electron systems and can be arranged within the Kadowaki–Woods plot next ␮ to the unconventional SC UPt3 . NMR and SR results show that both magnetic order and SC coexist on a microscopic scale without having spatial segregation of both phenomena. The absence of an inversion symmetry gives rise to a lifting of the degeneracy of electronic bands by spin-orbit coupling. As a consequence, the SC order parameter may have uncommon features as indicated from a very unique NMR relaxation rate 1/T1 and a linear temperature dependence of the penetration depth ␭.©2005 American Institute of Physics. ͓DOI: 10.1063/1.2008135͔

I. INTRODUCTION dates some aspects of the scheme of symmetry classification Electron correlations in solids are a ‘‘magic door’’ to the and leads to a mixture of singlet and triplet pairing in gen- discovery of unexpected features and phases of metals, inter- eral. metallics and oxides at low temperatures. Of particular im- A recently discovered example in this respect is CePt3Si ͑ ͒ 1 portance are phase transitions at Tϭ0. Critical fluctuations with the tetragonal P4mm No. 99 , the first associated with such a phase transition can lead to strong heavy fermion SC without a center of inversion. The absence renormalization of normal metallic properties and novel ex- of inversion symmetry induces peculiar band splittings, as otic phases may emerge from these strongly fluctuating en- we will discuss below, which are detrimental to certain kinds of Cooper pairing channels. In particular, spin triplet pairing vironments. One of the most exciting features in this context 2 is the occurrence of superconductivity ͑SC͒. becomes unlikely under these circumstances. The appearance of SC in such a scenario can deviate The aim of the present paper is to provide a review of from the common BCS type in many essential aspects. results of current research dedicated to the various physical Strong correlation effects responsible for the heavy electron properties of CePt3Si in both the normal and superconduct- behavior from narrow f -electron bands, may hamper the pos- ing states and to locate the system in the standard generic sibility of conventional Cooper pairing, i.e., pairing in the phase diagram of heavy fermion compounds. most symmetric (s-wave͒ form as it is favored by electron- This paper is organized as follows. After a discussion of phonon interaction. In turn, magnetic fluctuation may pro- normal state properties of CePt3Si, the SC features of vide the necessary attractive interaction in a different angular CePt3Si are examined before theoretical considerations con- momentum channel. This means that Cooper pairs may have cerning the SC order parameter are made. either spin-singlet or spin-triplet configuration and the orbital II. RESULTS AND DISCUSSION angular momentum may lead to a highly anisotropic gap A. Normal state properties with zero nodes. Almost all previously studied SC exhibiting strong electron correlations in the normal state region are Physical properties of ternary CePt3Si are dominated by characterized by a center of inversion in its crystal structure. the onset of long range, presumably antiferromagnetic order Ϸ ϭ This allows us to distinguish spin-singlet and spin-triplet below TN 2.2 K, followed by SC below Tc 0.75 K. This components of the SC order parameter and consider them intriguing coincidence of two ordering phenomena in separately. Superconductivity in materials having no inver- CePt3Si have to be considered in the context of crystal elec- sion symmetry is rare. This lack of a inversion center invali- tric field ͑CEF͒ splitting and Kondo interaction which sub-

1063-777X/2005/31(8–9)/9/$26.00748 © 2005 American Institute of Physics Low Temp. Phys. 31 (8–9), August–September 2005 Bauer et al. 749

on the same sample upon lowering the temperature. To cor- roborate the nFl property deduced for CePt3Si from the spe- cific heat data and to exclude short range order effects and inhomogeneities above the magnetic phase transition, C p(T) was studied for diluted Ce0.2La0.8Pt3Si as well. Results are ⌬ ͑ ͒ shown in Fig. 1 as C p versus ln T squares . This diluted sample—without magnetic ordering—exhibits a similar logarithmic contribution to the specific heat, like the parent CePt3Si, and thus establishes this feature as an intrinsic property. In order to analyze in more detail the magnetically or- 3 dered region of CePt3Si, a model by Continentino is applied for the specific heat well below Tmag :

FIG. 1. Temperature-dependent magnetic contribution to the specific heat, 39 T ⌬ ⌬ ͑ 7/2 1/2 C p of CePt3Si plotted as C p /T on a logarithmic temperature scale filled ͑ ͒ϭ ⌬ ͑Ϫ⌬ ͒ͫ ϩ ͩ ͪ Cmag T g T exp SW /T 1 ͒ ⌬ SW ⌬ circles . The phonon contribution to define C p is taken from C p(T)of 20 SW ͑ ͒ LaPt3Si filled diamonds . The long-dashed line represents the magnetic 2 entropy ͑right axis͒. The short-dashed line is a guide to the eye and roughly 51 T ϩ ͩ ͪ ͬ. ͑1͒ indicates the non-Fermi liquid behavior. The solid line is a fit according to 32 ⌬ ͑ ͒ ⌬ ϳ 3 SW Eq. 1 and the dotted-and-dashed line is a fit according to C p T . This expression is based on antiferromagnetic magnons ␻ϭͱ⌬2 ϩ 2 2 ⌬ with a dispersion relation SW D k , where SW is stantially modify Hund’s jϭ5/2 ground state of the Ce ion. the spin-wave gap and D is the spin-wave velocity; g The response of the system associated with the mutual inter- ϰ1/D3ϰ1/⌫3, and ⌫ is an effective magnetic coupling be- play of these mechanisms will be highlighted below. tween the Ce ions. A least squares fit of Eq. ͑1͒ to the data ͑ ͒ ⌬ Ϸ Substantial information concerning the magnetic and the below TN solid line, Fig. 1 reveals SW 2.7 K, a reason- paramagnetic properties of CePt3Si can be deduced from the able gap value with respect to the ordering temperature. A temperature-dependent magnetic contribution to the specific recent neutron diffraction study carried out on CePt3Si con- Ϸ heat Cmag(T). The latter may be defined by the difference firmed antiferromagnetic ordering below TN 2.2 K with a ⌬ ϭ C p between CePt3Si and isostructural nonmagnetic refer- wave vector k (0,0,1/2), i.e., doubling of the magnetic unit ⌬ ϳ ⌬ 4 ence LaPt3Si, with C p Cmag . Plotted in Fig. 1 is C p /T cell along the c direction. A second model calculation with ϰ 3 versus T of CePt3Si together with the raw data of LaPt3Si. simple antiferromagnetic spin waves, yielding Cmag T , The low temperature behavior of the latter can be accounted again gives reasonable agreement with the heat capacity ␪ ϭ for in terms of the Debye model with D 255 K together data. Using both models, one can estimate a Sommerfeld ␥Ϸ 2 ⌬ 2 with a Sommerfeld value 9 mJ/mol•K . C p(T)of coefficient of 0.42 and 0.39 J/mol•K for the former and CePt3Si defines three regimes: the SC state of CePt3Si below latter model, respectively. Such figures are in excellent ϭ Tc 0.75 K; the magnetically ordered range below TN agreement with an extrapolation of high field specific heat Ϸ 2.2 K, and the paramagnetic region above TN . This region data where SC is already suppressed by the applied magnetic is characterized by a Schottky-like anomaly with a weak lo- field. These strongly enhanced Sommerfeld values character- cal maximum around 40 to 50 K. Such numbers imply a CEF ize CePt3Si as a typical heavy fermion compound. level approximately 100 K above the ground state doublet. Considering Kondo type interactions to be responsible However, since C p(T) of LaPt3Si slightly exceeds the spe- for the significant renormalization of electrons, the tempera- ⌬ ϳ cific heat data of CePt3Si, C p(T) Cmag(T) becomes nega- ture dependent magnetic entropy Smag(T) allows one to esti- tive, and hence a reliable evaluation of CEF level scheme is mate the Kondo temperature TK . Applying results derived not possible, at least, considering this quantity only. The in- from the renormalization group technique5 for effective spin- ͑ ͒ tegrated entropy Smag right axis, Fig. 1 reaches R ln 2 1/2 systems to Smag(T) derived for CePt3Si yields TK 2 Ϸ around 25 K, and the entropy of 8.7 J/mol•K integrated up 7.2 K. A second possible estimate for TK follows from the ϭ 2 to 100 K is slightly less than R ln 4 11.5 J/mol•K . These competition of the RKKY interaction and the , results confirm again that the ground state of Ce3ϩ ions is a which leads to a significant reduction of the specific heat ϭ doublet with the first excited level above about 100 K. The jump at T TN . Following the procedure developed in Ref. Ϸ twofold degeneracy of the ground state doublet, however, is 6 gives TK 9 K, in reasonable agreement with the previous lifted by magnetic order as well as by Kondo type interaction estimate. spreading entropy to higher temperatures. The temperature-dependent electrical resistivity ␳(T)of ␳ The second interesting aspect in the paramagnetic tem- CePt3Si is plotted in Fig. 2a together with (T) of isostruc- ␳ perature range of CePt3Si is an almost logarithmic tail of tural LaPt3Si. (T) of CePt3Si drops to zero from a residual ⌬ ␮⍀ midϭ C p /T above TN , extrapolating to about 12 to 13 K. The value of 5.8 •cm with Tc 0.75 K, thus indicating su- logarithmic temperature dependence observed just above the perconductivity. At high temperatures, ␳(T) is characterized magnetic transition may be considered as hint of non-Fermi by a negative logarithmic contribution, followed by pro- liquid ͑nFl͒ behavior. Therefore, it is a unique observation at nounced curvatures around 75 K and 15 K, which may re- ambient pressure that non-Fermi liquid behavior, magnetic flect crystal electric field effects in the presence of Kondo ordering and eventually a SC transition consecutively arises type interactions. Further evaluation of ␳(T) requires knowl- 750 Low Temp. Phys. 31 (8–9), August–September 2005 Bauer et al.

␳ FIG. 2. Temperature-dependent electrical resistivity of CePt3Si and LaPt3Si plotted on a logarithmic temperature scale. The magnetic contribu- ␳ ͑ ͒ ͑ ͒ tion mag(T) dashed line refers to the right axis a . Low-temperature ␳ details of (T) of CePt3Si. The solid and the dotted-and-dashed lines are least squares fits ͑see text͒ and the dashed-line shows d␳(T)/dT ͑b͒.

␳ edge of the phonon contribution ph which, in a first approxi- mation, may be derived from homologous and isotypic LaPt3Si. LaPt3Si is metallic in the temperature range mea- ␳ Laϭ␳Laϩ␳La ␳ sured and (T) 0 ph(T)( 0 is the residual resistiv- ity͒ can simply be accounted for in terms of the Bloch Gru¨n- ␪ Ϸ ͓ eisen model with the Debye temperature D 160 K solid ͔ ␳ ϭ ␮⍀ line, Fig. 2a and 0 10.7 •cm. According to Matthies- ␳ ␳ Ce sen’s rule, (T) of CePt3Si can be expressed as (T) ϭ␳Ceϩ␳Ce ϩ␳ 0 ph(T) mag(T). The temperature-dependent mag- ␳ netic contribution to the resistivity, mag(T), then follows simply from the difference of ␳(T)Ce and ␳(T)La, assuming ␳CeϷ␳La ␳La ␳ ph ph . Furthermore, 0 is subtracted. mag(T) exhibits a Ͼ distinct logarithmic contribution for T 100 K; the maxi- FIG. 3. A plot of the T2 coefficient of the electrical resistivity A versus the mum around 80 K may indicate the overall crystal field split- T-linear specific heat coefficient ␥. Solid and dashed lines represent 1 ␥2ϭ ϫ Ϫ5 ␮⍀ 2 2 Ϫ2 ␥2ϭ ϫ Ϫ6 ␮⍀ ting of the jϭ5/2 Ce 4 f state ͑dashed line, Fig. 2a, right A/ 1 10 •cm•mol K mJ and A/ 0.4 10 •cm 2 2 Ϫ2 ͑ ͒ axis͒. •mol K mJ figure taken from Ref. 8 . Figure 2b exhibits low-temperature features of the elec- trical resistivity of CePt3Si. Besides the onset of supercon- ␳ ductivity, there is a distinct change of the slope in (T) spin gap obtained from this fit, ⌬ ϭ2.77 K, agrees excel- ␳ SW around 2 K, which becomes more evident from a d /dT plot lently with that value obtained from specific-heat analysis. ͓ ͔ right axis, Fig. 2b . In the context of the specific heat study, The coefficient A derived is much larger than usually this anomaly is interpreted as a signature of the onset of observed for simple metals like K or Cu, and thus evidences long-range magnetic order. A least-squares fit according to again that the low temperature state of CePt Si is dominated ␳ϭ␳ ϩ 2 ␳ ϭ ␮⍀ 3 0 AT reveals the residual resistivity 0 5.2 by heavy quasiparticles. In fact, considering the electron- ϭ ␮⍀ •cm and a material-dependent constant A 2.35 2 electron interaction in terms of the Barber model indicates cm/K . ϰ 2 • A (N(EF)) , where N(EF) is the electronic density of To account for the magnetically ordered region of states at the Fermi energy. Since also the Sommerfeld value CePt3Si in more detail, the above indicated model of Conti- depends on the same quantity, it is natural to arrange certain nentino can be adopted for the temperature dependent elec- systems according their A and ␥ values. Such a classification trical resistivity, where conduction electrons are scattered on 7 3 was made for the first time by Kadowaki and Woods, and antiferromagnetic magnons many highly correlated electron systems have been shown to 8 2 T satisfy this scheme. A closer inspection yields for the ratio ␳ϭ␳ ϩA⌬3/2 T1/2 exp͑Ϫ⌬ /T͒ͫ1ϩ ͩ ͪ A/␥2ϭ1ϫ10Ϫ5 ␮⍀ mol2 K2 mJϪ2. Arranging CePt Si in 0 SW SW 3 ⌬ • • • 3 SW such a Kadowaki-Woods plot, see Fig. 3, obviously shows 2 T 2 that the present compound is found at the very same site as ϩ ͩ ͪ ͬ. ͑2͒ 15 ⌬ the unconventional superconductor UPt3 . SW Temperature-dependent magnetic susceptibility provides ⌬ ␮ Again, SW is the spin-wave gap. A least squares fit of Eq. information concerning the effective magnetic moments eff ͑ ͒ 2 to the experimental data above Tc and below the ordering involved in a particular system, the interaction strength be- temperature yields an equally good agreement ͑dot-and-dash tween these moments via the paramagnetic Curie tempera- ␪ line, Fig. 2b͒ as the earlier model ͑solid line, Fig. 2b͒. The ture p , and about phase transitions present in a certain Low Temp. Phys. 31 (8–9), August–September 2005 Bauer et al. 751

FIG. 4. Temperature-dependent susceptibility of CePt Si plotted as 1/␹ ver- FIG. 5. Magnetic scattering obtained at 6.4 K with the incident energy of 35 3 ϭ sus T. The solid line is a least-squares fit according to the modified Curie- meV. The dashed line is for the elastic component with FWHM 2.4 meV Weiss law. The inset shows the low temperature behavior deduced from an while the short-dashed line represents the quasi-elastic component with ϭ ac susceptibility study ͑Ref. 9͒. FWHM 0.8 meV. The dotted-and-dashed line is for the sum of two Lorentzian components centered at 13 and 20 meV with FWHM ϭ10.0 meV ͑Ref. 17͒.

sample. Results taken from a SQUID measurement per- formed at ␮ Hϭ1 T are shown in Fig. 4 together with an ac order to account for the excitations observed, the standard 0 ϩ susceptibility measurement in the inset of this figure. At el- CEF Hamiltonian for Ce3 with C4v point symmetry is con- ␹ evated temperatures, (T) of CePt3Si exhibits a Curie-Weiss sidered: ͑ ͒ behavior—and the anomaly around 2.2 K inset, Fig. 4 in- H ϭB0O0ϩB0O0ϩB4O4. ͑4͒ dicates a magnetic phase transition. To account qualitatively CEF 2 2 4 4 4 4 m m for the region above about 50 K, a least-squares fit according Bl and Ol are the CEF parameters and the Steven operators, to the modified Curie-Weiss law, i.e., respectively. Due to CEF effects, the 6-fold degenerate ground state of the Ce3ϩ ions is split in tetragonal symmetry C into 3 doublets. Applying Eq. ͑4͒ to the data shown in Fig. 5 ␹ϭ␹ ϩ ͑ ͒ 0 3 0ϭϪ 0 Tϩ␪ reveals CEF parameters B2 0.4972 meV, B4 p ϭ 4ϭ 0.0418 meV, and B4 0.2314 meV. These parameters are ␹ ⌬ ϭ ⌬ ϭ was applied. 0 represents a temperature independent Pauli- consistent with levels at 1 13 and 2 20 meV and per- like susceptibility and C is the Curie constant. mit a reasonably good description of the magnetic scattering Results of this procedure are shown by the solid line in ͑solid line, Fig. 5͒. Keeping the CEF parameters unchanged, Fig. 4. The effective magnetic moment deduced from the the data obtained at 94 K are equally well explained. More- Curie constant C matches the theoretical value associated over, the two CEF excitations centred at 13 and 20 meV are with the 3ϩ state of cerium, thereby implying a rather stable also consistent with the heat capacity data as discussed ␪ magnetic moment. The paramagnetic Curie temperature p above. We furthermore studied low energy excitations using ϷϪ46 K is large and negative, being indicative of strong lower incident energy to find a weak feature around 1.4 meV. antiferromagnetic interactions. In the Kondo picture, already The dispersion of that intensity at Tϭ5 K, particularly adopted to explain the large Sommerfeld value ␥, the value around Qϭ0.8 ÅϪ1 is a signature for the development of ␪ of p suggests a Kondo temperature of the order of 10 K short-ranged magnetic correlations and can be considered as Ϸ͉␪ ͉ (TK p /4). origin of the anomalous behavior of the specific heat above ␪ ʈ ʈ Ϸ Very large values of p (H ͓100͔ and H ͓001͔) were magnetic ordering. At higher temperatures (T 30 K) scat- also deduced from a previous investigation4 of single crys- tering becomes Q-independent. This feature is completely talline CePt3Si, confirming the substantial antiferromagnetic absent in non-magnetic LaPt3Si. Unlike our study performed interaction strength. at the instrument HET of ISIS ͑UK͒, the inelastic neutron For better understanding of the magnetic ground state scattering experiment reported in Ref. 4 indicated two CEF and expected localized character of Ce 4 f electrons, neutron peaks at 1.0 and 24 meV. Based on the latter, a set of CEF inelastic scattering experiments were performed. In order to parameters was deduced, sufficient to account for isothermal accurately and reliably determine magnetic scattering from magnetization and the temperature dependent magnetic sus- the magnetic moments of Ce, both CePt3Si and LaPt3Si were ceptibility. However, this latter CEF level scheme does not investigated in powder form under identical conditions. For a properly describe the temperature dependent magnetic con- proper phonon subtraction, two well-established methods tribution to the specific heat and magnetic entropy. Such a were used,10 yielding almost identical results. discrepancy is rather serious, since, different to magnetiza- The intensity shown in Fig. 5 for a sample temperature tion and susceptibility, not any theoretical model is necessary Tϭ6.4 K is expected to be solely of magnetic origin as the for the calculation, except basic thermodynamics. In order to phonon contribution has been subtracted from the data. In get more reliable data for the analysis, particularly at low 752 Low Temp. Phys. 31 (8–9), August–September 2005 Bauer et al.

FIG. 6. Temperature-dependent specific heat C p /T of CePt3Si; the dashed 3 FIG. 7. Temperature-dependent specific heat C p /T of CePt3Si for various line is a T extrapolation of C p(T)at0T. 3 values of applied fields; the dashed line is a T extrapolation of C p(T)at0 ͑ ͒ T a . Temperature dependence of the upper critical field Hc2 . The solid Ј ϷϪ straight line yields Hc2 8.5 T/K; the dashed line is a guide to the eye ͑Ref. 1͒. PL indicates the Pauli-Clogston limiting field ͑b͒. energy excitations, neutron inelastic measurements with high resolution and low incident energy neutrons are in progress. Even using the electronic specific heat coefficient in the SC ␥ Ϸ 2 ⌬ ␥ B. Superconducting properties of CePt3Si state, s 0.18(1) J/mol•K , we obtained C p /( sTc) Ϸ0.55 that is still below the BCS value. Signs of bulk SC of CePt Si below T ϭ0.75 K are nu- 3 c Again, two scenarios may explain the substantial reduc- merous: zero resistivity, diamagnetic signal in the suscepti- ⌬ ␥ tion of C p /( Tc) with respect to the BCS value; strongly bility, a jump in the specific heat and NMR relaxation rate at ⌬ ␥ anisotropic gaps yield a reduced magnitude of C p /( Tc) Tc . ͑Ref. 11͒, and not all electrons condense into Cooper pairs, Substantial information concerning the superconducting and the rest stay essentially normal. This may imply that the state is provided by heat capacity data taken at low tempera- electrons responsible for normal state features, such as anti- tures for CePt3Si. Results are shown in Fig. 6 as C p /T ver- ferromagnetic order, coexist with those forming the Cooper sus T. The phonon contribution is negligible in the tempera- ␥ Ϸ 2 pairs. In fact, the finite value of s 0.18 J/mol•K provides ture range shown. Besides the already mentioned magnetic evidence that even at Tϭ0 K a significant portion of the ϭ phase transition at TN 2.2 K and the logarithmic contribu- Fermi surface is still not involved in the SC condensate. tion above that temperature, the superconducting transition at Very recently, a specific-heat study was performed on a Ϸ Tc 0.75 K is the most prominent feature. well heat-treated polycrystalline CePt Si sample, revealing ␥ Ϸ 2 3 The Sommerfeld coefficient n 0.39 J/mol K of Ϸ • two consecutive phase transitions at Tc1 0.8 and Tc2 CePt3Si at zero field, obtained from an extrapolation of the Ϸ0.55 K. The extreme different field response for both tran- antiferromagnetically ordered region, evidences the large ef- sitions, however, are possibly indicative of two different su- fective masses of the charge carriers involved. The extrapo- perconducting states.12 lation shown in Fig. 6 ͑dashed line͒ satisfies the basic re- The upper critical field Hc2(0) and the slope dHc2 /dT quirement of superconductivity, the entropy balance between ϵ Ј Hc2 are indispensable quantities for the determination of the SC and normal state regions. Another careful extrapola- microscopic parameters describing the superconducting tion of the heat capacity of data within the superconducting state. The temperature- and field-dependent specific heat ␥ temperature range towards zero yields about n* C (T,H) of CePt Si is shown for the low-temperature range Ϸ 2 p 3 210 mJ/mol•K . This nonvanishing contribution within in Fig. 7a. the SC state may hint at two mechanisms: not all electrons The application of magnetic fields reduces Tc , resulting are involved in the SC condensate, rather, roughly half of in a rather large change of dH /dTϷϪ8.5 T/K, in good ␥ Ϸ␥ c2 them ( s n*) are forming normal state long range mag- agreement with the conclusion drawn from electrical resis- netic order, which then co-exists with superconductivity on a ͓ ͔ tivity see Fig. 7b . An extrapolation of Tc(H) towards zero microscopic scale; a somewhat gapless superconducting Ϸ yields Hc2(0) 5 T, well above the paramagnetic limiting state. This would cause finite values of the Sommerfeld con- Ϸ 1 field H p 1.1 T. Furthermore, an estimation of the Som- stant and power laws of the thermodynamic and transport merfeld coefficient from the high field data gives properties instead of an exponential behavior as in typical 2 0.36 J/mol•K , in fair agreement with the value obtained BCS systems. The coexistence of both states is evidenced from an extrapolation of the normal state in the zero field ␮ ͑ ͒ from SR spectroscopy and a superconducting state with data see Fig. 6 . The upturn of C p /T at the lowest tempera- nodes in the gap ͑gapless at certain sites͒ is supported by tures, that gets stronger with increasing magnetic field, de- magnetic penetration depth ␭(T) measurements ͑see below͒. rives from the nuclear contribution of 195Pt. The jump of the specific heat anomaly associated with In order to derive a set of parameters characterizing the superconductivity, ⌬C /T͉ Ϸ0.1 J/mol K2, leads to p Tc • superconducting state of CePt3Si the BCS theory is ⌬ ␥ Ϸ 13,14 C p /( nTc) 0.25, which is significantly smaller than the adopted. Although substantial deviations from a spherical ⌬ ␥ Ϸ figure expected from the BCS theory ( C p /( Tc) 1.43). Fermi surface are expected for tetragonal CePt3Si, reason- Low Temp. Phys. 31 (8–9), August–September 2005 Bauer et al. 753 able physical parameters can be expected ͑compare, e.g., Refs. 14 and 15͒. ␥ ϭ 2 ͑ The starting parameters are s 0.18(2) J/mol•K as a ͒ Ј ϭϪ ␳ ϭ ␮⍀ lower limit , Hc2 8.5 T/K and 0 5.2 •cm. ͓ ͑ ͒ The effective Fermi surface Ss is then computed Eq. 2 ͔ clӍ ϫ 20 Ϫ2 in Ref. 14 as Ss 3.7 10 m within the clean limit and dlӍ ϫ 20 Ϫ2 Ss 3.5 10 m for the dirty limit. Considering the dirty Ј ϷϪ limit only, one gets Hc2 0.77 T/K, a value rather low Ј Ϸ with respect to the experimentally derived slope Hc2 Ϫ 8.5 T/K. This indicates that CePt3Si is not a typical dirty- limit superconductor. Thus, the further calculations are based ␥ on clean-limit results. Combining the Fermi surface with s Ӎ gives the Fermi velocity vF 5300 m/s and in the context of ␳ Ӎ ␮⍀ the residual resistivity, 0 5.2 •cm, a mean free path Ӎ ϫ 8 ␰ ltr 8 10 m can be estimated. The coherence length 0 for → T 0 was obtained from two independent relations. One fol- FIG. 8. Zero field depolarization rate of CePt3Si at various temperatures ␰ ϭ Ӎ ͑Ref. 17͒. lows from the BCS equation, 0 0.18hvF /(kBTc) 9.7 ϫ10Ϫ9 m. A second expression stems from the well known ␮ ϭ⌽ ␲␰2 ␰ ϭ ϫ Ϫ9 formula 0Hc2 0 /(2 0), yielding 0 8.1 10 m, in reasonable agreement with the former. above and below Tc did not show any change of the mag- ␬ netic signal, supporting the view of a microscopic coexist- The evaluation of the Ginzburg-Landau parameter GL ϭ␭/␰ requires knowledge of the thermodynamic critical ence between magnetism and SC. ␮ ϭ This points to a novel state for SC Ce-based heavy- field Hc(0) 26(2) mT, which can be calculated from 0 fermion systems at ambient pressure, for which, to date, the free energy difference between the superconducting and 18 the normal state: magnetism was found to be either absent or strongly com- peting against SC.19 The observed coexistence is reminiscent of the situation observed in UPd Al ,20 where a model of two ⌬ ͑ ͒ϭ Ϫ ϭ␮ 2͑ ͒ 2 3 F T Fn Fs 0Hc T /2 independent electron subsets, localized or itinerant, was pro- 21 T Ϫ posed in view of similar microscopic data. TЈ ͑Cs Cn͒ ϭ ͵ ͵ dTЉdTЈ. Additional microscopic information about the SC state TЉ Tc Tc can be obtained from the temperature-dependent 195Pt 22 nuclear spin-relaxation rate 1/T1 . Results in the form of a Cs is obtained from the zero-field specific heat, and Cn is plot of (1/T1T)/(1/T1T)T versus T/Tc are shown in Fig. 9 3 c taken from the T extrapolation, as indicated by the dashed for 8.9 and 18.1 MHz. The relaxation behavior 1/T1T of Ϸ line in Fig. 6. With Hc2(0) 5 T one derives a value for CePt3Si is reminiscent of a kind of Hebel–Slichter ␬ ϭ ͱ Ӎ 23 GL Hc2(0)/ 2Hc) 140, which, in turn, determines the anomaly, indicating coherence effects as in conventional ␭ → Ӎ ϫ Ϫ6 London penetration depth L(T 0) 1.1 10 m. Since BCS SC. The peak height, however, is significantly smaller the evaluation of the various parameters is based on s-wave than that observed for conventional BCS SC and, addition- models, care has to be taken when using their absolute val- ally, shows no field dependence at the 8.9 MHz (Hϳ1T) ues; nevertheless the right order of magnitude can be antici- pated. ͑ ͒ ␳ Ϸ ␮⍀ Evaluating Eq. A.13 of Ref. 14 with max 100 Ӎ ϫ 21 Ϫ2 •cm yields ShT 3.1 10 m , the Fermi surface at el- evated temperatures. The discrepancy between Ss and ShT suggests that only a part of the Fermi surface is involved in forming Cooper pairs while the remaining one engages in normal state magnetic correlations. This finding seems to be convincingly supported from the lessened value of ⌬ ␥ C p /( Tc). In terms of the coexistence of both supercon- ϭ ductivity (Tc 0.75 K) and long-range magnetic order (TN ϭ 2.2 K), the downsized specific heat jump at Tc may ex- plain, at least partly, that the Fermi surface is likely to be ͑ ␥ subdivided into a superconducting part related to s) and a normal-state region. Microscopic evidence for the latter con- clusion can be found from zero-field ␮SR spectroscopy data obtained in the magnetic phase below and above Tc in the ͑ ͒ magnetic phase Fig. 8 . At temperatures much above TN , FIG. 9. A plot of (1/T T)/(1/T T) versus T/T at 8.9 MHz (Hϳ1T͒ and the ␮SR signal is a characteristic of a paramagnetic state 1 1 Tc c 18.1 MHz (Hϳ2 T). The dashed line is for the Balian-Werthamer model with a depolarization solely arising from nuclear moments. ͑ ͒ ⌬ ϭ BW isotropic triplet SC state with a value of 2 /kBTc 4. The dotted line Below T the ␮SR signal indicates that the full sample ⌬ ϭ N assumes a point-node model with 2 /kBTc 3.6, and the dotted-and-dashed 16 ⌬ ϭ ͑ ͒ volume orders magnetically. High statistic runs performed line represents a fit by a line-node gap model with 2 /kBTc 5.1 Ref. 17 . 754 Low Temp. Phys. 31 (8–9), August–September 2005 Bauer et al.

ϳ and 18.1 MHz (H 2 T) run. Notably, CePt3Si is the first HF SC that exhibits a peak in 1/T1T just below Tc . ϳ 1/T1T at H 2 T seems to saturate at low temperature, which can be assigned to the presence of vortex cores where the normal-state region is introduced. 1/T1T at 8.9 MHz (Hϳ1 T), however, continues to decrease down to T ϭ0.2 K, the lowest temperature yet measured. Neither an exponential law nor the usual T3 law, reported for most of the unconventional HF SC ͑see, e.g., Ref. 24 and references cited therein͒, is observed for the data down to Tϭ0.2 K Ϸ ( 0.3Tc). However, further studies at temperatures below 0.2Tc are required to make any definite conclusion on the specific behavior of 1/T1 . The nuclear spin-lattice relaxation rate 1/T1 in the super- conducting state can be expressed as 1 2␲A2 ϱ ϭ ͵ ͓ 2͑ ͒ϩ 2͑ ͔͒ ͑ ͓͒ Ϫ ͑ ͔͒ ប Ns E M s E f E 1 f E dE, T1 0 ͑5͒

Ns is the density of states and describes the distinct features of an isotropic, polar or axial SC state; M s is the anomalous density of states arising from coherence effects which can be calculated from

N͑0͒ 2␲ ␲ ⌬͑␪,␸͒ M ͑E͒ϭ ͵ ͵ sin ␪d␪d␸, ͑6͒ s ␲ 2 2 4 0 0 ͱE Ϫ⌬ ͑␪,␸͒ ⌬͑ ␸͒ with 0, the direction-dependent energy gap. M s is ex- pected to exist only in s-wave superconductors. To account for the relaxation behavior below Tc in non- centrosymmetric CePt3Si, three unconventional models were adopted for a description of the temperature dependence of 1/T at Hϳ1 T. The dashed line in Fig. 9 represents a fit ⌬␭ ⌬␭ 1 FIG. 10. Normalized variation (T)/ 0 versus T/Tc in a polycrystalline ͑ ͑ ͒ ⌬␭ according to the Balian–Werthamer model BW isotropic CePt3Si sample a . Low-temperature behavior of (T) in polycrystalline ͒ ⌬ ϭ 25 ␭ spin-triplet SC state with 2 /kBTc 3.9. Note that the CePt3Si and Cd samples. (T) is independent of temperature below 0.28Tc peak of the BW model in 1/T T originates from the presence in cadmium, as expected for an s-wave superconductor. The inset displays a 1 linear low-temperature behavior of ␭(T) in CePt Si ͑b͒. of an isotropic energy gap. The dotted-and-dashed line is a fit 3 ⌬ ϭ using a line-node model with 2 /kBTc 5.1. The dotted line ⌬ ϭ refers to a point-node model with 2 /kBTc 3.6. The mod- els used, however, failed to give satisfactory description of the observed temperature dependence of 1/T1 over the entire tained using a tunnel diode oscillator system running at 9.5 temperature range. MHz, for a polycrystalline CePt3Si sample at temperatures The data seem to start following the line-node model at down to 0.049 K.26 Figure 10a depicts the normalized varia- the lowest measured temperatures. However, the data are de- ⌬␭ ⌬␭ tion of (T)/ 0 versus T/Tc in the whole temperature scribed reasonably well by the BW ͑nodeless͒ model just ⌬␭ ϭ␭ Ϫ␭ region below Tc . Here (T) (T) (0.049 K) and below Tc . The experimentally observed peak in 1/T1T ⌬␭ 0 is the total penetration depth shift. Figure 10b displays would indicate the presence of an isotropic energy gap, even ⌬␭(T) versus T/T in the low-temperature range TϽ0.6T though a coherence effect—inherent for the isotropic spin- c c for polycrystalline CePt Si and Cd samples. Cadmium is a singlet s-wave pairing state—is absent. 3 classic s-wave superconductor, for which the low tempera- In almost all previous studies on either conventional and ture dependence of ␭(T) is exponential. The inset to Fig. 10b unconventional SC, it was assumed that the crystal has an ⌬␭ inversion center, which allows separate consideration of the is a close-up of (T) versus T/Tc for the CePt3Si sample Ͻ even ͑spin-singlet͒ and odd ͑spin-triplet͒ components of the for temperatures T 0.3Tc , where it can be clearly seen that the penetration depth data of the CePt3Si sample follow a SC order parameter. In CePt3Si, however, a center of sym- metry is absent. Therefore, the novel relaxation behavior linear temperature behavior below 0.17Tc . We remark here that the low-temperature behavior of ␭(T) is not affected by found below Tc hints at a possibly new class of a SC state being realized in noncentrosymmetric heavy fermion com- the type of sample ͑single crystal, powder, etc.͒ in which the pounds. measurement is performed, if the sample is of high An important probe of the structure of the superconduct- quality.27–29 ing energy gap is the temperature dependence of the mag- For a clean, local superconductor the penetration depth netic penetration depth ␭(T). Figure 10 shows results ob- is given by Low Temp. Phys. 31 (8–9), August–September 2005 Bauer et al. 755 f E ץ ␭2͑0͒ ϱ ϭͫ 1ϩ2 ͳ ͵ dE ʹ ͬ. ͑7͒ 2 E ͱE2Ϫ⌬2͑T,␪,␸͒ץ ⌬ ␭ ͑T͒

Here ͗...͘ represents an angular average over the Fermi sur- face, and f is the Fermi function. Evidently the temperature dependence of ␭(T) depends on the topology of the gap structure. For line nodes in the energy gap the penetration depth is expected to be linear in the low temperature limit, where the temperature dependence of the energy gap can be FIG. 11. Fermi surface splitting. The lack of inversion symmetry gives rise Ͼ␰ to antisymmetric spin-orbit coupling which leads in general to a splitting of neglected. Assuming that CePt3Si is both a clean (l 0) and the spin degeneracy of the electronic states. The figure depicts schematically ␭ Ͼ␰ a local ( (0) 0) superconductor, as was discussed above, the splitting of the Fermi surfaces for the generating point group C4v the penetration depth experimental result points out to the CePt3Si. The spinor states are k-dependent in the way that the spin quanti- existence of lines of nodes in the structure of the supercon- zation axes lie perpendicular to the Fermi surface and the z axis with the spin pointing in opposite direction for the two Fermi surfaces. This feature is ducting pairing state and, hence, to unconventional supercon- essential for the possible spin configuration of the pairing states which can ductivity in CePt3Si. For this material with a tetragonal crys- be constructed from two electrons on the same Fermi surface. ϭ Ϫ tal lattice a p-wave pairing symmetry, like d(k) xˆky yˆkx proposed earlier by Frigeri et al., will not be consistent with line nodes, because for materials with strong spin-orbit cou- reflection symmetry z→Ϫz, where z is the fourfold tetrag- pling all the spin-triplet states are predicted to have point onal rotation axis. This is incorporated into band structure by nodes.30 Thus, the T-linear variation of the penetration depth a Rashba-like term suggests possibly a d-wave ͑spin-singlet͒ type of pairing ⑀ ␴ →⑀ ␴ ϩ␣͑zˆϫk͒ ␴ ͑8͒ symmetry in the energy gap. k 0 k 0 • ⌬␭ ␴ ␴ A broad shoulder in (T) just below Tc , like the one with 0 as the unit matrix and as the Pauli matrices in the observed in Fig. 10a, is usually associated in polycrystalline electron spinor space. The electron bands split into two, samples to intergrain or proximity effects. However, this ef- yielding two different Fermi surfaces with opposite spinor fect is not expected to be relevant in the present case, be- orientation as depicted in Fig. 11. Obviously there is now a cause of the high quality of the polycrystalline sample and clear restriction on the possible spin configurations for zero the very small measuring magnetic fields ͑about 5 mOe͒.27 momentum Cooper pairs. ⌬␭(T) for this sample has an inflection point around 0.52 K The mixing of the spin singlet and spin triplet channel temperature that is near to the second superconducting tran- for this electron band structure leads to an non-vanishing ͑ ͒ 12 ϭ 31–33 sition 0.55 K recently found in CePt3Si. Thus, it is tempt- spin susceptibility at T 0 in any case. Thus the absence ing to relate the shoulder to this second transition. However, of paramagnetic limiting would be naturally explained. It the inflection point—or similar feature—does not seem to be was suggested, however, by Kaur and collaborators that in present in preliminary results ͑obtained by one of the au- high magnetic fields perpendicular to the z-axis a novel su- thors͒ in a sedimented powder sample, where the intrinsic perconducting phase with a helical order parameter could be behavior is thought to be more pronounced. Thus, no conclu- realized which would leave its traces in the temperature de- sions can be drawn on the second transition from the present pendence of the upper critical field.34 measurements of the penetration depth. So far the symmetry of the order parameter remains an open question. However, theoretical considerations suggest that the lack of inversion symmetry introduces several complications in the microscopic discussion of C. Symmetry aspects superconductivity.31,32,35–39 Time reversal invariance and inversion symmetry are es- sential ingredients for superconductivity. They allow us to III. SUMMARY distinguish pairing with spin singlet and spin triplet configu- ration and even and odd parity. Due to Fermion antisymme- We summarize that non-centrosymmetric CePt3Si is a ϭ try of the Cooper pair wavefunction these are uniquely com- heavy-fermion SC with Tc 0.75 K that orders magnetically ϭ ␮ bined into even parity spin singlet and odd parity spin triplet at TN 2.2 K. Specific heat, NMR, and SR studies indicate pairing states. The formation of Cooper pairs with vanishing that superconductivity and long range magnetic order coexist total momentum relies on the availability of degenerate elec- on a microscopic scale and may be originated by two differ- tron states of opposite momentum on the Fermi surfaces. It is ent sets of electrons. The NMR relaxation rate 1/T1 shows generally believed that for spin singlet Cooper pairing time unexpected features which were neither found before in con- reversal invariance provides the necessary conditions, while ventional nor in heavy fermion SC, indicative of very un- spin triplet pairing needs additionally an inversion center.2 usual shapes of the SC order parameter. Unconventional SC The absence of inversion symmetry removes, however, the is backed also by the present penetration depth studies. In distinction of even and odd parity and leads immediately to a fact, the various theoretical scenarios developed for this mixing also of the spin channels.31 compound support these conclusions. The absence of inversion symmetry gives rise to anti- This work was supported by the Austrian FWF P16370, symmetric spin–orbit coupling. In the case of CePt3Si the P18054, and by FONACIT of Venezuela under Grant No. generating point group is C4v , which implies that there is no S1-2001000639. 756 Low Temp. Phys. 31 (8–9), August–September 2005 Bauer et al.

*E-mail: [email protected] 20 R. Feyerherm, A. Amato, C. Geibel, F. N. Gygax, P. Hellmann, R. H. Heffner, D. E. MacLaughlin, R. Mu¨ller-Reisener, G. J. Nieuwenhuys, A. ͑ ͒ 1 E. Bauer, G. Hilscher, H. Michor, Ch. Paul, E. W. Scheidt, A. Gribanov, Schenck, and F. Steglich, Phys. Rev. B 56, 699 1997 . 21 Yu. Seropegin, H. Noel, M. Sigrist, and P. Rogl, Phys. Rev. Lett. 92, R. Caspary, P. Hellmann, M. Keller, G. Sparn, C. Wassilew, R. Ko¨hler, C. 027003 ͑2004͒. Geibel, C. Schank, F. Steglich, and N. E. Phillips, Phys. Rev. Lett. 71, 2 P. W. Anderson, Phys. Rev. B 30, 4000 ͑1984͒. 2146 ͑1993͒ 3 M. A. Continentino, S. N. de Medeiros, M. T. D. Orlando, M. B. Fontes, 22 M. Yogi, Y. Kitaoka, S. Hashimoto, T. Yasuda, R. Settai, T. D. Matsuda, Y. and E. M. Baggio-Saitovitch, Phys. Rev. B 64, 012404 ͑2001͒. Haga, Y. Onuki, P. Rogl, and E. Bauer, Phys. Rev. Lett. 93, 027003 4 N. Metoki, K. Kaneko, T. D. Matsuda, A. Galatanu, T. Takeuchi, S. Hash- ͑2004͒. imoto, T. Ueda, R. Settai, Y. Onuki, and N. Bernhoeft, J. Phys.: Condens. 23 L. C. Hebel and C. P. Slichter, Phys. Rev. 107, 901 ͑1957͒. ͑ ͒ Matter 16, L207 2004 . 24 H. Tou, Y. Kitaoka, K. Asayama, C. Geibel, C. Schank, and F. Steglich, 5 H. U. Desgranges and K. D. Schotte, Phys. Lett. A 91, 240 ͑1982͒. ͑ ͒ 6 J. Phys. Soc. Jpn. 64, 725 2003 . M. Besnus, M. J. Braghta, A. Hamdaoui, and A. Meyer, J. Magn. Magn. 25 R. Balian and N. R. Werthamer, Phys. Rev. 131, 1553 ͑1963͒. Mater. 104–107,1385͑1992͒. 26 I. Bonalde, W. Bra¨mer-Escamilla, and E. Bauer, Phys. Rev. Lett. 94, 7 K. Kadowaki and S. B. Woods, Solid State Commun. 58,507͑1986͒. 207002 ͑2005͒. 8 N. Tsujii, K. Yoshimura, and K. Kosuge, J. Phys.: Condens. Matter 15, 27 R. B. Goldfarb, M. Lelental, and C. A. Thompson, in Magnetic Suscepti- 1993 ͑2003͒. ͑ ͒ 9 E. Bauer, R. Lackner, G. Hilscher, H. Michor, M. Sieberer, A. Eichler, A. bility of Superconductors and Other Spin Systems, R. A. Hein ed. , Ple- ͑ ͒ Gribanov, Y. Seropegin, and P. Rogl, J. Phys.: Condens. Matter 17, 1877 num Press, New York 1991 ,p.49. 28 ͑2005͒. T. Xiang, C. Panagopoulos, and J. R. Copper, Int. J. Mod. Phys. B 12, 10 D. T. Adroja, J. G. Park, K. A. McEwen, N. Takeda, M. Ishikawa, and 1007 ͑1998͒; C. Panagopoulos, Phys. Rev. Lett. 79, 2320 ͑1997͒. J.-Y. So, Phys. Rev. B 68, 094425 ͑2003͒. 29 F. Manzano, A. Carrington, N. E. Hussey, S. Lee, A. Yamamoto, and S. 11 A. P. Makenzie and Y. Maeno, Rev. Mod. Phys. 75, 657 ͑2003͒. Tajima, Phys. Rev. Lett. 88, 047002 ͑2002͒. 12 E. W. Scheidt, F. Mayr, G. Eickerling, P. Rogl, and E. Bauer, J. Phys.: 30 M. Sigrist and K. Ueda, Rev. Mod. Phys. 63, 239 ͑1991͒. Condens. Matter 17, L121 ͑2005͒. 31 L. P. Gor’kov and E. I. Rashba, Phys. Rev. Lett. 87, 037004 ͑2001͒. 13 See, for example, M. Tinkham, Introduction to Superconductivity, 32 P. A. Frigeri, D. F. Agterberg, and M. Sigrist, New J. Phys. 6,931͑2004͒. McGraw-Hill, New York ͑1975͒. 33 L. N. Bulaevski, A. A. Guseinov, and A. I. Rusinov, Sov. Phys. JETP 44, 14 ͑ ͒ U. Rauchschwalbe, Physica B 147,1 1987 . 1243 ͑1976͒. 15 R. Movshovich, M. Jaime, J. D. Thompson, C. Petrovic, Z. Fisk, P. G. 34 R. P. Kaur, D. F. Agterberg, and M. Sigrist, cond-mat/0408149. ͑ ͒ Pagliuso, and J. L. Sarrao, Phys. Rev. Lett. 86, 5152 2001 . 35 K. V. Samokhin, E. S. Zijlstra, and S. K. Bose, Phys. Rev. B 69, 094514 16 A. Amato, E. Bauer, and C. Baines, Phys. Rev. B 71, 092501 ͑2005͒. ͑2004͒. 17 E. Bauer, G. Hilscher, H. Michor, M. Sieberer, E. W. Scheidt, A. Gribanov, 36 P. A. Frigeri, D. F. Agterberg, A. Koga, and M. Sigrist, Phys. Rev. Lett. 92, Yu. Seropegin, P. Rogl, A. Amato, W. Y. Song, J.-G. Park, D. T. Adroja, 097001 ͑2004͒. M. Nicklas, G. Sparn, M. Yogi, and Y. Kitaoka, Physica B 359–361, 360 37 ͑ ͒ ͑2005͒. V. P. Mineev, Int. J. Mod. Phys. B 18, 2963 2004 . 38 ͑ ͒ 18 C. Petrovic, R. Movshovich, M. Jaime, P. G. Pagliuso, M. F. Hundley, J. I. A. Sergienko and S. Curnoe, Phys. Rev. B 70, 144522 2004 . 39 L. Sarrao, Z. Fisk, and J. D. Thompson, Europhys. Lett. 53, 354 ͑2001͒. S. S. Saxena and P. Monthoux, Nature ͑London͒ 427, 799 ͑2004͒. 19 G. M. Luke, A. Keren, K. Kojima, L. P. Le, B. J. Sternlieb, W. D. Wu, Y. J. Uemura, Y. Onuki, and T. Komatsubara, Phys. Rev. Lett. 73, 1853 This article was published in English in the original Russian journal. Repro- ͑1994͒. duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 31, NUMBERS 8–9 AUGUST–SEPTEMBER 2005

Antiferromagnetism and superconductivity of the two-dimensional extended t-J model C. T. Shih,* J. J. Wu, and Y. C. Chen

Department of Physics, Tunghai University, Taichung, Taiwan C. Y. Mou and C. P. Chou

Department of Physics, National Tsing Hua University, Hsinchu, Taiwan R. Eder

Institut fu¨r Festko¨rperphysik, Forschungszentrum Karlsruhe, Germany T. K. Lee

Institute of Physics, Academia Sinica, Nankang, Taiwan ͑Submitted January 13, 2005͒ Fiz. Nizk. Temp. 31, 995–1001 ͑August–September 2005͒ The mechanism of high-temperature superconductivity ͑HTS͒ and the correlation between the antiferromagnetic long-range order ͑AFLRO͒ and superconductivity ͑SC͒ are the central issues of the study of HTS theory. SC and AFLRO of the hole-doped two-dimensional extended t-J model are studied by the variational Monte Carlo method. The results show that SC is greatly enhanced by the long-range hopping terms tЈ and tЉ for the optimal and overdoped cases. The phase of coexisting SC and AFM in the t-J model disappears when tЈ and tЉ are included. It is concluded that the extended t-J model provides a more accurate description for HTS than the traditional t-J model does. The momentum distribution function n(k) and the shape of Fermi surface play critical roles for establishing the phase diagram of HTS materials. © 2005 American Institute of Physics. ͓DOI: 10.1063/1.2008136͔

I. INTRODUCTION d-wave resonating valence bond ͑RVB͒ state with AFLRO is a good trial wave function ͑TWF͒ and that SC is absent due The two-dimensional ͑2D͒ t-J model was proposed to to the constraint of no double occupancy. Upon doping, the 1,2 provide the mechanism of superconductivity ͑SC͒ right af- carriers become mobile and SC sets in while AFLRO is ter the discovery of high-temperature superconductivity quickly suppressed. However, AFLRO will survive until the ͑HTS͒. This idea quickly gained momentum when varia- ␦ Ͼ hole density h 10%, which is much larger than the critical tional calculations showed that the doping dependence of density observed by experiments. SC and AFLRO coexist in 3,4 pairing correlation and the phase diagram of the antifer- the very underdoped regime.5,19–22 ranagnetic long-range order ͑AFLRO͒ and SC seem to agree 5 The discrepancies imply that the t-J model may be in- with experimental results fairly well. However, the calcula- sufficient to describe the physics of HTS. On the other hand, tion beyond variational method showed that SC of the pure there are several experimental and theoretical studies sug- 2D t-J model was not large enough to explain such high gesting the presence of the next- and third-nearest-neighbor transition temperature of the cuprates.6 Up to now, this issue hopping terms tЈ and tЉ in cuprates. For example, the topol- has still not been settled.7–9 ogy of the large Fermi surface ͑FS͒ and the single-hole dis- Interplay between the d-wave SC and AFLRO is another persion studied by angle-resolved photoemission spectros- one of the critical issues in the physics of HTS.10,11 Early copy ͑ARPES͒, and the asymmetry of phase diagrams of the experimental results showed the existence of AFLRO at tem- ´ electron- and hole-doped cuprates can be understood by in- perature lower than the Neel temperature TN in the insulating 23 perovskite parent compounds of the cuprates. When charge troducing these terms. carriers are doped, AFLRO is destroyed quickly and then SC It is suggested that the longer range hopping terms may appears. In most thermodynamic measurements for hole play important roles on the mechanism of HTS. Results of 24,25 26 doped cuprates, AFLRO does not coexist with SC12 and dis- band-structure calculations and experimental analysis appears completely around doping density ␦ ϳ5%. How- show that Tc is enhanced by the next-nearest neighbor hop- h Ј ever, recent experiments such as muon spin rotation and elas- ping t /t, and the highest Tc,max for different monolayer hole tic neutron scattering show that the ͑SDW͒ doped cuprates strongly correlates with tЈ/t. However, this 27,28 may compete, or coexist with SC.13–18 These results suggest contradicts with previous results of exact calculations that AFLRO may coexist with SC, but the possibility of in- that for the hole doped systems, introducing tЈ into the t-J homogeneous phases is not completely ruled out. model will suppress pairing. For the theoretical part of this issue, analytical and nu- We will discuss the model and the trial wave function in merical studies of the t-J model show that at half-filling, the Sec. 2, and the variational Monte Carlo ͑VMC͒ method re-

1063-777X/2005/31(8–9)/6/$26.00757 © 2005 American Institute of Physics 758 Low Temp. Phys. 31 (8–9), August–September 2005 Shih et al. sults for SC, AFLRO, and the shape of the Fermi surface in where ␮ is the chemical potential determining the number of Ј Љ Sec. 3. At last we will make a summary in Sec. 4. electrons, and tv and tv are variational parameters corre- Ј sponding to the next and third nearest neighbor hoppings. tv Љ Ј Љ and tv are not necessarily equal to the bare values t and t because the constraint strongly renormalizes the hopping am- II. THE MODEL AND THE WAVE FUNCTIONS plitude. Note that the summation in Eq. ͑2͒ is taken over the ͑ ͒ The Hamiltonian of the extended t-J model is sublattice Brillouin zone SBZ . The operator Pd enforces the constraint of no doubly occupied sites for cases with ϭ ϩ ϭϪ ͑ † ϩ ͒ finite doping. H Ht HJ ͚ tij ˜ci,␴˜c j,␴ H.c. ij For the half-filled case, ␮ϭtЈϭtЉϭ0 and the optimal variational energy of this trial wave function ͑TWF͒ obtained 1 ϩ ͩ Ϫ ͪ ͑ ͒ by tuning ⌬ and m in the VMC simulation is Ϫ0.332 J per J ͚ Si•Sj nin j , 1 s ͗i, j͘ 4 bond, which is within 1% of the best estimate of the ground- ϭ 30 where tij t, tЈ, tЉ, and 0 for sites i and j that are nearest, state energy of the Heisenberg model. For the case of pure next nearest, and third nearest neighbors and other sites, re- AFLRO without ⌬, the energy per bond is about 3 to 4% spectively. ͗i, j͘ in HJ means the spin-spin interaction occurs higher. ϭ Ϫ only for nearest neighbors. ˜ci,␴ (1 ni,Ϫ␴)ci,␴ satisfies the Upon doping, there are two methods of modifying the no-double-occupancy constraint. At half-filling, the system is TWF: one is to use a nonzero ␮ to control the filling of the 29 reduced to the Heisenberg Hamiltonian HJ . As carriers are SDW bands, the other is to create charge excitations from 31 doped into the parent compound, Ht is included in the the half-filled ground states. For the former method, the ⌬ Ј Љ ␮ Hamiltonian. TWF is optimized by tuning , ms , tv , tv , and . Note that ϭ To solve the ground-state wave function of this Hamil- for larger doping densities, AFLRO disappears (ms 0) and tonian, three mean-field order parameters are introduced:21,29 the wave function reduces to the standard d-RVB wave func- ϭ z ϭϪ z the staggered magnetization ms ͗S ͘ ͗S ͘, where the tion. For the latter method, the wave function is the ‘‘small A B ͉⌿ lattice is divided into A and B sublattices, the uniform bond Fermi pocket’’ state p͘: ␹ϭ ͚ † order parameters ͗ ␴ci␴c j␴͘, and the d-wave RVB Ns/2 (d-RVB͒ order parameter ⌬ϭ͗c ↓c ↑Ϫc ↑c ↓͘,ifi and j are j i j i ͉⌿ ͘ϭP ͩ ͚ ͑A a† a† ϩB b† b† ͒ͪ ͉0͘. p d k k↑ Ϫk↓ k k↑ Ϫk↓ Ϫ⌬ ෈  n.n. sites in the x direction, and for the y direction. The k SBZ,k Qp Lee-Shih wave function, which is the mean-field ground- ͑3͒ state wave function, is N The k-points in Q p are the momenta of the electron singlet s/2 ͑ ↑ Ϫ ↓ ͉⌿ ͘ϭP ͩ ͚ ͑A a† a† ϩB b† b† ͒ͪ ͉0͘, pairs with momenta and spin (k , k ) removed from the LS d k k↑ Ϫk↓ k k↑ Ϫk↓ k෈SBZ half-filled FS. Thus the number of holes is twice of the num- ͑2͒ ␮ Ј Љ ber of k-points in Q p , , tv and tv are identical to zero in Eq. ͑2͒ because the size and shape of FS are determined by where Ns is the total number of sites and the choice of Q . Note that no matter what k’s are chosen in ͑1͒ ͑2͒ ϩ p A ϭ͑E ϩ␰˜͒/⌬ , B ϭϪ͑E Ϫ␰ ͒/⌬ k k k k k k k k Q p , the total momentum of the wave function is zero. The with k’s can be viewed as ‘‘hidden quantum numbers’’ of the Ϫ ϩ wave function. E͑1͒ϭ͑␰ 2ϩ⌬2͒1/2, E͑2͒ϭ͑␰ 2ϩ⌬2͒1/2. k k k k k k In general, for the ground state the set Q p should be Here ⌬ ϭ 3J⌬(cos k Ϫcos k ). The energy dispersions for determined variationally. As we expected, it agrees well with k 4 x y 31 the two SDW bands are the rigid band picture for very underdoped systems. For example, there is only one point in the two-hole system. ␰ϮϭϮ͓͑␧ ϩ␮͒2ϩ͑ ͒2͔1/2Ϫ␮Ј k k Jms The variational energies for several choices of k in a ϫ with 12 12 lattice are shown in Fig. 1. It can be seen that for both the (tЈ,tЉ)ϭ(Ϫ0.1,0.05)t ͑full circles͒ and (tЈ,tЉ) 3 ϭ(Ϫ0.3,0.2)t ͑open circles͒ cases, the k with lowest energy ␧kϭϪ2ͩ t␦ϩ J␹ͪ ͑cos k ϩcos k ͒. 8 x y is ͑␲/2,␲/2͒. The k’s with the second-lowest energy are ͑2␲/ 3,␲/3͒ and ͑␲/2,␲/3͒ for (tЈ,tЉ)ϭ(Ϫ0.1,0.05)t and (tЈ,tЉ) a ϭ␣ c ϩ␴␤ c k␴ k k␴ k kϩQ␴ ϭ(Ϫ0.3,0.2)t, respectively. ϭϪ␴␤ ϩ␣ ϭ ␲ ␲ and bk␴ kck␴ kckϩQ␴ , where Q ( , ), According to the rigid-band assumption, we expect that the best choice of Q for the 4-hole system is ͕(␲/2,␲/2), 1 p 2 ϩ Ϫ␲ ␲ ͖ ␣ ϭ ͕1Ϫ͓͑␧ ϩ␮Ј͒/͑␰ ϩ␮Ј͔͖͒, ( /2, /2) . And the Q p’s for the 6-hole system with k 2 k k ͑ Ј Љ͒ϭ͑Ϫ ͒ ͑ Ј Љ͒ϭ͑Ϫ ͒ 1 t ,t 0.3,0.2 t and t ,t 0.1,0.05 t ␤2ϭ ͕ ϩ͓͑␧ ϩ␮Ј͒ ͑␰ϩϩ␮Ј͔͖͒ k 1 k / k , 2 are are the operators of the lower and upper SDW bands, respec- tively, ͕͑␲/2,␲/2͒,͑Ϫ␲/2,␲/2͒,͑␲/2,␲/3͖͒, ␮Јϭ␮ϩ Ј ϩ Љ͑ ϩ ͒ 4tv cos kx cos ky 2tv cos 2kx cos 2ky , and Low Temp. Phys. 31 (8–9), August–September 2005 Shih et al. 759

ϭ ␲ ␲ Ϫ␲ ␲ ␲ ␲ ␲ ␲ Q p ͕͑ /2, /2͒,͑ /2, /2͒,͑ /2, /3͒,͑ /3, /2͖͒. The variational energies of these two wave functions, long-range pair-pair correlation, and staggered magetization are almost identical ͑within error bars͒. Since kϭ(Ϯ␲/2, Ϯ␲/3) and (Ϯ␲/3,Ϯ␲/2) are all degenerate for the two- hole system, the wave function could also be degenerate for Ϯ␲ Ϯ␲ those Q p’s with k-points ( /2, /2) and any two of k ϭ(Ϯ␲/2,Ϯ␲/3) and (Ϯ␲/3,Ϯ␲/2) for the 8-hole system. This conjecture has been verified numerically. Thus the best TWF should be a linear combination of all these wave func- tions. For simplicity, we choose only one of the Q p in the ϫ J tϭ tЈ tЉ following calculation. The properties of SC and AFLRO are FIG. 1. Energies for two holes in a 12 12 lattice for / 0.3, ( , ) 32 ϭ(Ϫ0.1,0.05)t ͑filled circles͒ and (tЈ,tЉ)ϭ(Ϫ0.3,0.2)t ͑unfilled circles͒. k not affected by this simplification. is the ‘‘hidden quantum number’’ corresponding to the momentum of the pair removed from the half-filled Fermi surface. Note that the total momenta of all the wave functions are zero. III. RESULTS AND DISCUSSION The staggered magnetization ͕͑␲/2,␲/2͒,͑Ϫ␲/2,␲/2͒,͑2␲/3,␲/3͖͒, 1 " ͗M͘ϭ ͳ ͚ eiQ RjSz ʹ , ͑4͒ respectively. R Ns j j Figure 2 shows the choices of Q p’s for several doping densities ͑0–10 holes͒ for the (tЈ,tЉ)ϭ(Ϫ0.3,0.2)t case. The the momentum distribution function validity of the rigid-band picture has been checked by com- 1 ͑ Ϫ ͒ ͑ ͒ϭ ik• Ri Rj † ͑ ͒ paring several Q p’s for the same number of holes for these n k ͚ e ͗ci␴c j␴͘, 5 N ␴ very underdoped cases. s ij Another issue is that the choice of Q p may change the and the d-wave pair-pair correlation total symmetry of the wave function. For example, Fig. 2e 1 † shows that Q p for 8 holes is P ͑R͒ϭ ͳ ͚ ⌬ ⌬ ϩ ʹ , ͑6͒ d R Ri R Ns i i ͕͑␲/2,␲/2͒,͑Ϫ␲/2,␲/2͒,͑␲/2,3͒,͑Ϫ␲/2,Ϫ␲/3͖͒. where We can also choose ⌬ ϭc ͑c ϩ ˆ↓ϩc Ϫ ˆ↓Ϫc ϩ ˆ↓Ϫc Ϫ ˆ↓͒ Ri Ri↑ Ri x Ri x Ri y Ri y are measured for J/tϭ0.3 and ͑a͒ tЈϭtЉϭ0, ͑b͒ (tЈ,tЉ) ϭ(Ϫ0.3,0.2)t and ͑c͒ (tЈ,tЉ)ϭ(Ϫ0.1,0.05)t cases for the ϫ ave 12 12 lattice with periodic boundary condition. Pd is the ͉ ͉Ͼ averaged value of the long-range part ( R 2) of Pd(R). The optimal wave function for different densities are deter- mined by minimizing the variational energies among ͉⌿ ⌬ ͕ ͖ ͘ ͉⌿ ⌬ Ј Љ ␮ ͘ p(ms , , Q p ) and LS(ms , ,tv ,tv , ) . We will dis- cuss the results for these three cases in this section.

A. tЈÄtЉÄ0 It can be seen in Fig. 3a that in the underdoped region for the J/tϭ0.3, tЈϭtЉϭ0 case, AFLRO coexists with SC ␦ ϳ ␦ for density smaller than c 10%. The c is smaller than the weak-coupling mean-field result ϳ15%,21 but is still larger than the phase boundary of AFLRO determined by experi- ␦ Ͻ ͉⌿ ments ( c 5%). The energies of LS͘ are lower than ͉⌿ those of P͘ for all doping densities in this case. This result is also consistent with the results reported by Himeda and Ogata.22 Comparison of the VMC result with that of the weak-coupling one seems to indicate that the rigorous no- double-occupancy constraint suppresses the AFLRO faster than the constraint-relaxed mean-field approximation. ave Pd shows a dome-like shape which agrees well with the experiments except in the slightly doped AFLRO region. It is well known that the variational method usually overes- ͑ ͒ timates the order parameters. Our previous studies using cal- FIG. 2. Choices of Q p unfilled circles for several doping densities for ave tЈ/tϭϪ0.3 and tЉ/tϭ0.2 in k space. The filled circles are the occupied culations beyond VMC show that Pd will be suppressed k-points: 0 ͑a͒,2͑b͒,4͑c͒,6͑d͒,8͑e͒,10͑f͒ holes. greatly when the wave function is projected to the true 760 Low Temp. Phys. 31 (8–9), August–September 2005 Shih et al.

Ͻ␦ Ͻ For a little larger doping density 0.06 h 0.15, it can ave be seen that Pd starts to grow but is still smaller than that in Јϭ Љϭ ave Ј Љ the t t 0 case. The suppression of Pd by t and t in the underdoped regime is consistent with the results27,28 ob- tained the density matrix renormalization group ͑DMRG͒ ␦ ave method. Interestingly, for even larger h , Pd grows greatly ␦ ϳ and reaches the maximum at h 30%, and the SC region ␦ ϳ ave Ј extends to h 0.4. The maximal Pd is larger than the t ϭtЉϭ0 value at the same density by almost one order of magnitude, and about 2.5 times larger than the maximum of the optimal value of the tЈϭtЉϭ0 case. The enhancement of ave Pd may come from the deformation of the Fermi surface. The electron occupation at the k-points near ͑␲,0͒ is in- creased by a negative tЈ. The results from exact diagonaliza- tion and slave-boson mean-field theory also show similar behavior.34 The great enhancement of pairing due to tЈ may provide ␦ a possible mechanism for HTS. But the doping density max ave ϳ with maximal Pd is too large ( 30%) in comparison with experiments ͑15%͒. This discrepancy may disappear for the real ground state of the extended t-J model. From our expe- rience, if we do the calculation beyond VMC, the amplitude ave ␦ of Pd will be suppressed and max will move to a smaller value.6 If this trend is true for the t-J type models, we expect ␦ that max may move toward the more physical value. This conjecture will be investigated in the future.

͑ ͒ ave ͑ ͒ ϭ Јϭ Љ FIG. 3. ͗M͘ filled circles and Pd unfilled circles for J/t 0.3: t t ϭ0 ͑a͒,(tЈ,tЉ)ϭ(Ϫ0.3,0.2)t ͑b͒, and (tЈ,tЉ)ϭ(Ϫ0.1,0.05)t ͑c͒ for a hole- C. tЈÕtÄÀ0.1 and tЉÕtÄ0.05 doped 12ϫ12 lattice. The vertical dashed lines show the critical doping Ј ϭϪ Љ ͉⌿ ͘ ͉⌿ ͘ ͉⌿ ͘ For the lanthanum materials with t /t 0.1 and t /t densities of level crossing. P ( P ), LS ( LS ), and RVB ( LS with ϭ m ϭ0) represent the best TWFs of each region. 0.05, the behaviors are more complex. It can be seen from s ␦ Ͻ ͉⌿ Fig. 3c that for hole density h 4%, P͘ optimizes the variational energy, and the phase in this region is ARFLO but Ͻ␦ Ͻ ͉⌿ ground state. Note that the two-hole binding energy becomes not SC. For 4% h 10%, LS͘ is the best TWF with 6 ⌬ positive ͑no binding͒ in the thermodynamic limit. nonzero ms and . AFLRO and SC coexist in the ground state of this density interval. At even larger dopings, ms in ͉⌿ ͘ B. tЈÕtÄÀ0.3 and tЉÕtÄ0.2 LS vanishes and the phase becomes pure SC. The maxi- ␦ ϳ ave mum of the SC dome is at h 20%, and the maximal Pd Now we examine the phase diagram for J/tϭ0.3, tЈ/t is about 1.5 times larger than the tЈϭtЉϭ0 value. ϭϪ0.3, and tЉ/tϭ0.2, parameters for YBCO and BSCO Since the phase transition comes from the level crossing compounds. The results are shown in Fig. 3b. It was found of the two classes of states ͉⌿ ͘ and ͉⌿ ͘, it is a first-order ␦ ϳ ␦ Ͻ ͉⌿ P LS that level crossing occurs at c 0.06. For h 0.06, P͘ is phase transition. It is quite natural to have inhomogeneity in the ground-state wave function and ͗M͘ is a little larger than the system near the critical point.38 It may also lead to other Јϭ Љϭ ave 39 in the t t 0 case, while Pd is suppressed by one order more novel inhomogeneous states such as a stripe phase. of magnitude. Thus there is AFLRO but no SC in this re- Another interesting result of our study is that the non- ␦ ϭ gime. For h larger than 0.06, the RVB state (ms 0in coexistence of SC and AFLRO is much more robust for sys- ͉⌿ ͘ ave ͗ ͘ LS ) optimizes the energy. Pd increases, and M drops tems with larger values of tЈ/t and tЉ/t, such as YBCO and sharply to zero. Unlike the tЈϭtЉϭ0 case, there is no region BSCO.25 For LSCO, where tЈ/t and tЉ/t are smaller, the ͉⌿ optimized by LS͘ with nonzero ms . In conclusion, there is tendency toward coexistence is larger and the possibility of no coexistence of AFLRO and SC for the (tЈ,tЉ) an inhomogeneous phase will become much more likely. ϭ(Ϫ0.3,0.2)t case. To show that ͉⌿ ͘ and ͉⌿ ͘ belong to two different LS P D. Shape of the Fermi surface types of wave function, we calculate their overlap. ⌿ ͉⌿ ͉⌿ ͉͉⌿ ͉ ͑ ͒ 33 ͗ LS P͘/( LS P ) is only 0.0113 4 . The near- Figure 4 shows the FS of both under- and overdoped orthogonality of the two wave functions implies that the systems with the parameter sets we discussed above. For the ␦ ϭ ground-state wave functions switch at the critical density. underdoped systems ( h 6/144), there is a large FS for the ␦ ϭ ϭ ͑ ͒ The result that the critical h for negative tЈ/t is smaller tЈ tЉ 0 case Fig. 4a and a clear ‘‘Fermi pocket’’ for the than that of the tЈϭ0 case is consistent with the results tЈ/tϭϪ0.3, tЉ/tϭ0.2 case ͑Fig. 4b͒, whose ground-state 35,36 ͉⌿ ϭ evaluated by exact diagonalization, and the suppression wave function is P͘. The shape of the FS for tЈ/t of coexistence of AFLRO and SC is consistent with the Ϫ0.1 and tЉ/tϭ0.05 ͑Fig. 4c͒ is placed between the previous 37 ͉⌿ slave-boson mean-field theory. two cases. The ground-state wave function is LS͘ but the Low Temp. Phys. 31 (8–9), August–September 2005 Shih et al. 761

ave Ј ЉϭϪ Ј ϫ ͑ FIG. 5. Maximal Pd for different t with t t /2 for 8 8 unfilled circles͒ and 12ϫ12 ͑filled circles͒ lattices.

and kϭ(0,Ϯ␲). Hence the FS becomes disjoint pieces. Al- though at ϪtЈ/tϭ0.4 the FS is still connected, this tendency is already observed. The density of states starts to decrease, and this is probably the reason for the suppression of pairing beyond ϪtЈ/tу0.4.

IV. SUMMARY ͉⌿ In summary, a new wave function P͘ is proposed for FIG. 4. Contour maps of n(k): tЈϭtЉϭ0 ͑a͒, ͑d͒;(tЈ,tЉ)ϭ(Ϫ0.3,0.2)t the extended t-J model for very low hole densities. The size ͑ ͒ ͑ ͒ ϭ Ϫ ͑ ͒ ͑ ͒ ͑ ͒ ͉⌿ b , e ;(tЈ,tЉ) ( 0.1,0.05)t c , f . The hole densities are 6/144 for a , and shape of the FS and of P͘ are determined by the ͑b͒, ͑c͒, and 44/144 for ͑d͒, ͑e͒, ͑f͒, respectively. ෈ choice of pairs with the momenta k Q p removed from the half-filled system. The chosen k’s are around the ͑␲/2,␲/2͒ region in k space for hole-doped materials. The behavior of pocket-like feature is still obvious. Lack of a large FS is one ͉⌿ ͉⌿ P͘ is very different from that of LS͘ which optimizes of the possible reasons for the suppression of Pave by tЈ in ͉⌿ d the energy for the t-J model. In contrast to P͘, the FS for the underdoped region. the states of ͉⌿ ͘ is controlled by the chemical potential ␮ ␦ ϭ LS For the overdoped systems ( h 44/144), the ground- and the effective long-range hopping terms tЈ and tЉ . ͉⌿ ͘ v v state wave functions for all the three cases are LS . They There are three remarkable effects of tЈ and tЉ for the all have large FSs but with different shapes determined extended t-J model. First, the critical density where AFLRO Ј Љ mainly by the parameters tv and tv . It is clear that the dis- vanishes is moved to more physical values. Second, the tortion of the FS makes n(kϭ(␲,0)) for the tЈ/tϭϪ0.3 and phase of coexisting AFLRO and SC is suppressed. If tЈ and tЉ/tϭ0.2 ͑Fig. 4e͒ case much larger than the other two. The tЉ are large enough ͑corresponding to the YBCO or BSCO shapes of the FS for tЈϭtЉϭ0 ͑Fig. 4d͒ and tЈ/tϭϪ0.1 and ͒ ave materials , the coexisting phase will disappear. Third, Pd is tЉ/tϭ0.05 ͑Fig. 4f͒ are similar and the occupations near enhanced for the optimal and overdoped region, and sup- ͑␲,0͒ are both small. For the d-wave SC, the electron pairs pressed for the underdoped region. This resolves the conflict ͑␲ ͒ ave with momenta near ,0 contribute to SC most. Thus Pd between the DMRG and band structure calculation results. for tЈ/tϭϪ0.3 and tЉ/tϭ0.2 case is much larger than the ave The enhancement of Pd can be explained by the electron other two. occupation near ͑␲,0͒ and FS. These results offer a possible ave Our results show that Pd is closely correlated with mechanism for HTS. n(k) and thus with the shape of the FS. Figure 5 plots the The work is supported by the National Science Council ave maximal possible value of Pd for all doping densities as a in Taiwan with Grant Nos. NSC-93-2112-M-029-001-, 93- Ј ave Ј function of t . The maximal Pd is proportional to t in the 2112-M-007-009-, 93-2112-M-001-018-, and 93-2112- range 0уtЈуϪ(0.3– 0.4). Beyond these values pairing is M029-009. Part of the calculations are performed in the IBM no longer enhanced. Coincidentally, these values are about P690 and PC clusters in the National Center for High- the same value of tЈ/t for mercury cuprates, as estimated by Performance Computing in Taiwan, and the PC clusters of 25 Pavarini et al., but much larger than those reported in Ref. the Department of Physics and Department of Computer Sci- 24. Among all the cuprate series, mercury cuprate has held ence and Engineering of Tunghai University, Taiwan. We are the record of having highest Tc for almost a decade. grateful for their help. ave Ϫ Ју The decrease of Pd for t 0.4 in the overdoped re- ␦ϭ gime such as 0.31 is also likely a consequence of the *E-mail: [email protected] change of the FS. n(kϭ(␲,0)) is almost saturated at ϪtЈ ϭ0.4 and remains unchanged for larger ϪtЈ. It is not diffi- Ϫ cult to recognize that as tЈ becomes much larger than t, 1 P. W. Anderson, Science 235,1196͑1987͒. electrons will occupy separate regions around kϭ(Ϯ␲,0) 2 F. C. Zhang and T. M. Rice, Phys. Rev. B 37, 3759 ͑1988͒. 762 Low Temp. Phys. 31 (8–9), August–September 2005 Shih et al.

3 C. Gros, Phys. Rev. B 38,1196͑1988͒. 21 M. Inaba, H. Matsukawa, M. Saitoh, and H. Fukuyama, Physica C 257, 4 F. C. Zhang, C. Gros, T. M. Rice, and H. Shiba, Supercond. Sci. Technol. 299 ͑1996͒. 1,36͑1988͒. 22 A. Himeda and M. Ogata, Phys. Rev. B 60, R9935 ͑1999͒. 5 T. K. Lee and S. P. Feng, Phys. Rev. B 38, 11809 ͑1988͒. 23 A. Damascelli, Z. Hussain, and Z.-X. Shen, Rev. Mod. Phys. 75, 473 6 C. T. Shih, Y. C. Chen, H. Q. Lin, and T. K. Lee, Phys. Rev. Lett. 81, 1294 ͑2003͒. ͑1998͒. 24 R. Raimondi, J. H. Jefferson, and L. F. Feiner, Phys. Rev. B 53,8774 7 S. Sorella, G. B. Martins, F. Becca, C. Gazza, L. Capriotti, A. Parola, and ͑1996͒; L. F. Feiner, J. H. Jefferson, and R. Raimondi, Phys. Rev. Lett. 76, ͑ ͒ E. Dagotto, Phys. Rev. Lett. 88, 117002 2002 . 4939 ͑1996͒. 8 T. K. Lee, C. T. Shih, Y. C. Chen, and H. Q. Lin, Phys. Rev. Lett. 89, 25 E. Pavarini, I. Dasgupta, T. Saha-Dasgupta, O. Jepsen, and O. K. Ander- 279702 ͑2002͒. ͑ ͒ 9 sen, Phys. Rev. Lett. 87, 047003 2001 . S. Sorella, A. Parola, F. Becca, L. Capriotti, C. Gazza, E. Dagotto, and G. 26 K. Tanaka, T. Yoshida, A. Fujimori, D. H. Lu, Z.-X. Shen, X.-J. Zhou, H. Martins, Phys. Rev. Lett. 89, 279703 ͑2002͒. Eisaki, Z. Hussain, S. Uchida, Y. Aiura, K. Ono, T. Sugaya, T. Mizuno, 10 P. W. Anderson, The Theory of Superconductivity in the High-T Cuprates, c and I. Terasaki, Phys. Rev. B 70, 092503 ͑2004͒. Princeton University Press, Princeton, NJ ͑1997͒. 27 S. R. White and D. J. Scalapino, Phys. Rev. B 60, R753 ͑1999͒. 11 S. C. Zhang, Science 275, 1089 ͑1997͒. 28 G. B. Martins, J. C. Xavier, L. Arrachea, and E. Dagotto, Phys. Rev. B 64, 12 C. M. A. Kastner, R. J. Birgeneau, G. Shirane, and Y. Endoh, Rev. Mod. 180513 ͑2001͒. Phys. 70,897͑1998͒. 29 T. K. Lee and C. T. Shih, Phys. Rev. B 55, 5983 ͑1997͒. 13 B. Lake, H. M. Rønnow, N. B. Christensen, G. Aeppli, K. Lefmann, D. F. 30 ͑ ͒ McMorrow, P. Vorderwisch, P. Smeibidl, N. Mangkorntong, T. Sasagawa, S. Liang, B. Doucot, and P. W. Anderson, Phys. Rev. Lett. 61,365 1988 . 31 ͑ ͒ M. Nohara, H. Takagi, and T. E. Mason, Nature ͑London͒ 415, 299 T. K. Lee, C. M. Ho, and N. Nagaosa, Phys. Rev. Lett. 90, 067001 2003 . 32 ͑2002͒. Note that n(k) will depend on the choice of these Q p because there will be 14 R. I. Miller, R. F. Kiefl, J. H. Brewer, J. E. Sonier, J. Chakhalian, S. dips for the unoccupied k’s. Dunsiger, G. D. Morris, A. N. Price, D. A. Bonn, W. H. Hardy, and R. 33 C. T. Shih, Y. C. Chen, C. P. Chou, and T. K. Lee, Phys. Rev. B 70, Liang, Phys. Rev. Lett. 88, 137002 ͑2002͒. 220502͑R͒͑2004͒. 15 J. E. Sonier, K. F. Poon, G. M. Luke, P. Kyriakou, R. I. Miller, R. Liang, 34 C. T. Shih, T. K. Lee, R. Eder, C. Y. Mou, and Y. C. Chen, Phys. Rev. Lett. C. R. Wiebe, P. Fournier, and R. L. Greene, Phys. Rev. Lett. 91, 147002 92, 227002 ͑2004͒. ͑2003͒. 35 T. Tohyama and S. Maekawa, Phys. Rev. B 49, 3596 ͑1994͒. 16 Y. Sidis, C. Ulrich, P. Bourges, C. Bernhard, C. Niedermayer, L. P. Reg- 36 R. J. Gooding, K. J. E. Vos, and P. W. Leung, Phys. Rev. B 49,4119 nault, N. H. Andersen, and B. Keimer, Phys. Rev. Lett. 86, 4100 ͑2001͒. ͑1994͒. 17 J. A. Hodges, Y. Sidis, P. Bourges, I. Mirebeau, M. Hennion, and X. 37 H. Yamase and H. Kohno, Phys. Rev. B 69, 104526 ͑2004͒. Chaud, Phys. Rev. B 66, 020501͑R͒͑2002͒. 38 J. Burgy, M. Mayr, V. Martin-Mayor, A. Moreo, and E. Dagotto, Phys. 18 H. A. Mook, P. Dai, S. M. Hayden, A. Hiess, J. W. Lynn, S.-H Lee, and F. Rev. Lett. 87, 277202 ͑2001͒. ͑ ͒ Dogˇan, Phys. Rev. B 66, 144513 2002 . 39 A. Himeda, T. Kato, and M. Ogata, Phys. Rev. Lett. 88, 117001 ͑2002͒. 19 G. J. Chen, R. Joynt, F. C. Zhang, and C. Gros, Phys. Rev. B 42, R2662 ͑1990͒. This article was published in English in the original Russian journal. Repro- 20 T. Giamarchi and C. Lhuillier, Phys. Rev. B 43, 12943 ͑1991͒. duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 31, NUMBERS 8–9 AUGUST–SEPTEMBER 2005

Shift of the basal planes as the order parameter of transitions between the antiferromagnetic phases of solid oxygen E. V. Gomonay*

N. N. Bogolyubov Institute of Theoretical Physics, National Academy of Sciences of Ukraine, ul. Metrologicheskaya 14-b, Kiev 03143, Ukraine; National Technical University of Ukraine ‘‘Kiev Polytechnical Institute,’’ pr. Pobedy 37, Kiev 03506, Ukraine V. M. Loktev†

N. N. Bogolyubov Institute of Theoretical Physics, National Academy of Sciences of Ukraine, ul. Metrologicheskaya 14-b, Kiev 03143, Ukraine ͑Submitted January 31, 2005͒ Fiz. Nizk. Temp. 31, 1002–1019 ͑August–September 2005͒ A phenomenological model in the spirit of Landau theory is constructed and used to analyze the phase stability conditions and the conditions for phase transitions between the various magnetocrystalline structures of solid oxygen over a wide range of pressure, temperature, and external magnetic field. © 2005 American Institute of Physics. ͓DOI: 10.1063/1.2008137͔

I. INTRODUCTION properties—the heat capacity,5 thermal expansion,6 and ther- 7,8 mal conductivity —of solid O2 has not only confirmed the The discovery of antiferromagnetism as a physical phe- results of the pioneering magnetic measurements of Borovik- nomenon is inextricably connected to the Kharkov period of Romanov et al.9,10 but also yielded new results, in particular, activity of L. V. Shubnikov and L. D. Landau. There can be on the order of the ␣-␤ transition. It is now accepted ͑though no doubt that the experimental discovery of a new magnetic not conclusively proven͒ that it is a first-order, close to phase transformation between identical crystal structures not second-order, transition, and it occurs not only between dif- manifesting a macroscopic magnetic moment M or, in mod- ferent crystal structures but also between different magnetic ern terminology, a paramagnet-antiferromagnet transition, orderings. While the lowest-temperature (Tр23 K) phase at belongs to Trapeznikov and Shubnikov, who observed an Pϭ0, the monoclinic ␣ phase, is characterized by a collinear anomaly of the heat capacity of anhydrous ferrous chloride two-sublattice magnetic structure with easy-axis anisotropy, and linked it to the magnetic properties of this substance.1 the rhombohedral ␤ phase is an example of a three-sublattice Landau, apparently learning of this research from the authors 120° magnetic structure with easy-plane anisotropy and prior to publication of the results, laid out the basic concepts correlation-type ͑short-range͒ ordering. This triangular struc- of the antiferromagnetic state. Based on these concepts, he ture in solid O was predicted by one of the authors11 and is developed a phenomenological theory of such a magnetic 2 known as the Loktev structure. state of the crystal, in which one ferromagnetic sublattice is A theory of the magnetocrystalline phase transformation exactly compensated by another, so that, on the whole, spon- ␣ taneous magnetization of macroscopic regions ͑ferromag- between the spin-collinear phase and the spin-noncollinear ␤ netism͒ does not arise after the transition.2 It would later be phase was proposed in Ref. 12, based on taking into ac- ͑ ͒ found that the class of antiferromagnets, in which, by defi- count the dependence of the exchange Heisenberg pair in- nition, the total magnetization Mϭ0 although the mean teraction on the intermolecular distance in the framework of  the linear theory of elasticity. That theory showed that the spins ͗Sn͘ 0 at the sites n are finite and ordered, includes an enormous number of both metallic and insulating systems, magnetoelastic transition arising in such a model can provide with an extremely great diversity of magnetic structures ͑see, a correct description of both the crystal and magnetic struc- e.g., Ref. 3͒. tures of the ␣ and ␤ phases on the qualitative ͑and even Among the antiferromagnetic insulators a special place semiquantitative͒ level. Nevertheless, a treatment of the ␤ is held by solid oxygen. First, it is one of a very few mo- phase as an antiferromagnet with long-range dipolar mag- ͑ ͒ lecular magnets if not the only one , since the O2 molecule netic order does not completely fit in with the modern view has spin Sϭ1 in its ground state. Second, also uniquely, the of the magnetism of easy-plane structures in crystals of re- antiferromagnetism of this cryocrystal is realized in a uni- duced dimensionality. form ͑homomolecular͒ medium. Third and last, the equilib- The point is that, to a significant degree, ␤ oxygen is rium phases ͑␣, ␤, and ␥͒ of this crystal in the absence of among those degenerate two-dimensional magnetic systems external pressure P display different magnetic properties, that cannot exhibit an isotropic magnetoelastic gap, and so and the magnetic ͑exchange͒ component makes a real con- long-range spin ordering in the usual sense cannot arise in it tribution to the intermolecular interaction, significantly stabi- ͑see, e.g., Ref. 13͒. However, it has been shown14–16 that in lizing the crystal structure of the ␣ and ␤ phases of solid view of the above-listed special properties of the Loktev oxygen.4 structure, which can exist as right or left polarized, the sign It should be noted that research on the thermal of the polarization ͑and not the Ne´el order͒ can be preserved

1063-777X/2005/31(8–9)/14/$26.00763 © 2005 American Institute of Physics 764 Low Temp. Phys. 31 (8–9), August–September 2005 E. V. Gomonay and V. M. Loktev in it over macroscopic distances. This introduces a physical II. CHOICE OF THE ORDER PARAMETER characteristic that can serve as a measure of the ordering in the absence of long-range vector order. This ordering is Solid oxygen, as we have said, is a unique physical sys- called ‘‘correlation ordering,’’ since it is of the nature of tem in which the character of the magnetic ordering has a highly developed short-range correlations corresponding substantial influence on the crystal structure; the relative dis- completely to the Loktev structure. These correlations admit placements of the atoms at phase transitions induced by ex- description in the framework of the Landau phenomenologi- ternal pressure are so large that the lattice strain tensor can- cal approach. not be considered small, and the standard harmonic By the mid-1980s to the early 1990s it seemed that the approximation is insufficient to describe it. In addition, from ͑ ͒ basic experimental and theoretical research on the magnetic a symmetry standpoint not all the Fedorov space groups of ͑ and structural features of solid oxygen had been done, and the observed phases are in a subgroup relationship for ex- ␤ ␦ the most important physical parameters characterizing the ample, the and phases, the space groups of which are 5 23 ͒ crystal and magnetic ͑sub͒lattices had been determined. But D3d and D2h , respectively . All of these features make it a ͑ the development of high-pressure technique substantially ex- nontrivial problem to choose the order parameter or param- ͒ panded the realm of standard problems in the physics of eters and the form of the corresponding thermodynamic po- tential adequate to describe the whole sequence of transitions cryocrystals, ushering in the study of new crystal structures between the ␤, ␣, and ␦ phases of solid oxygen ͑we begin arising under external mechanical pressure in substances that with the lattice of ␤-O as the most symmetric of these had seemingly been studied already. Again, by virtue of its 2 phases͒. molecular structure and unusually strong magnetic interac- tions, solid oxygen has turned out to be a unique system A. Order parameter of the structural phase transition among cryocrystals in that diverse phases, both magnetic and All three phases considered below, the ␣ and ␤ phases nonmagnetic, arise at high pressure, which differ in their and the ␦ phase that arises under pressure, have a layered crystal structure and electronic properties, as well. structure in which the oxygen molecules can all be consid- A comprehensive and exhaustive review of the experi- ered to be oriented parallel to each other and perpendicular to mental data concerning the thermodynamic properties of the the close-packed atomic planes.19 The molecules solidify in solid phases of O2 and the existing theoretical models was ͑ 17 such a position exclusively on account of quantum valence recently given by Freiman and Jodl. Without going into the and exchange͒ interactions, since the other substantial details presented in that review, let us note that the theoreti- interaction—the quadrupole-quadrupole—leads to a noncol- cal studies of this topic have predominantly taken a micro- linear mutual orientation of the molecules.4 scopic or semi-microscopic approach based on a concrete If we ignore the distortions of vectors lying in the basal form for the intermolecular interaction potentials. Ap- plane and the relatively small changes of the interplane dis- proaches of this type, which to a certain extent includes the tances, then it seems obvious that each of the phases in the 12,14–16 studies mentioned above, have succeeded, for ex- sequence of phase transitions ␤→␣→␦ can be obtained and ample, in elucidating the leading role of magnetic interac- in reality arises through a simple shift of the hexagonal tions in the ␣-␤ transition and, through analysis of the lattice close-packed planes relative to each other ͑see Fig. 1͒, from dynamics, have led to the prediction of a jumplike change of an initial right-angle stacking ͑i.e., stacking such that in the 18 the lattice parameters at the ␣-␤ and ␣-␦ transitions. A direction perpendicular to the planes the atoms are located good complement to these microscopic models, which per- one above the other͒. mit rather exact calculation of the parameters of the crystal in The initial ‘‘unshifted’’ ordering of the hexagonal planes the framework of some restrictions on the range of the inter- can be regarded as some virtual ͑not observed in reality͒ actions, the forms of the potentials, etc., is the phenomeno- phase ͑the paraphase, in the terminology of Ref. 20͒, the logical ͑in essence, thermodynamic͒ theory, which gives a 1 symmetry group D6h of which includes the symmetry groups quantity picture of the phase transitions and the features that of the ␤, ␣, and ␦ phases as subgroups. appear on the macroscopic observables at those transitions. Choosing the reference unit cell to be one constructed on Below we attempt to treat the sequence of phase transforma- the basis vectors of the lattice of the paraphase, one can tions in solid oxygen specifically from the standpoint of the describe all the experimentally observed changes of the crys- general phenomenological theory in the spirit of the Landau tallographic structure with the use of a uniform finite-strain theory of phase transitions, which, as far as we know, has not tensor containing five independent components ͑since the been applied to solid O2 before. The approach is based on twofold symmetry axis in the basal plane is preserved͒.As general symmetry concepts without recourse to a concrete the finite distortions one can choose, e.g., the relative change form of the intermolecular interaction, in conformity with of the lengths of the basis vectors, the distortion of the basal Occam’s razor—simplicity without trivialization. A feature plane determined by the hexagonality parameter (&hab of the proposed approach is that the thermodynamic potential Ϫa)/a, and the relative displacement of the close-packed is of a form for which one can take into account the finite planes ␰ϭc cos ␤/a, where a,b,c, ␤ specify in the general ͑and non-small͒ lattice distortions arising in solid oxygen case a monoclinic lattice with mutually perpendicular vectors under pressure, which are not describable in the framework a, b in the basal plane. Only the last of these listed param- of the standard theory of elasticity. We shall also consider the eters ͑independently of its magnitude͒ can be used to de- question of the influence of external magnetic field on the scribe the sequence of phase transitions between the ␤, ␣, structure of the phases and the macroscopic properties of and ␦ phases using general symmetry arguments in accor- their lattices in different ranges of pressure and temperature. dance with the canons of Landau theory. The quantity ␰, Low Temp. Phys. 31 (8–9), August–September 2005 E. V. Gomonay and V. M. Loktev 765

␺ ϭ TABLE I. Transformation of the functions nm , n,m 0,1, under the op- 1 erations of the generating elements of the group D6h .

1 tions of the group D6h are presented in Table I. The indicated order parameter is uniform in the sense that the relative dis- placement of any two adjacent planes is identical in both magnitude and direction.3͒

B. Description of the magnetic structure In accordance with the experimental data currently avail- able, which are presented rather completely in Freiman and Jodl’s review,17 the spin moments localized on the oxygen molecules form a noncollinear triangular structure in the basal plane, with short-range order in the ␤ phase ͑Fig. 2a͒ and an antiferromagnetic collinear structure in the ␣ and ␦ phases ͑Fig. 2b͒. The mutual orientation of the moments in adjacent close-packed layers is not uniquely determined and FIG. 1. Crystal structure of the antiferromagnetic phases of solid oxygen: can be parallel or antiparallel ͑see, e.g., Ref. 21͒. A symme- ͑ ͒ ␤ ͑ ͒ ␣ ͑ ͒ ␦ hypothetical nonmagnetic paraphase a,c , -O2 b,d , -O2 e , and -O2 ͑ ͒ ͑ ͒ try analysis of the two-dimensional magnetic structures com- f . The arrows indicate the direction of the shift the vector u . The hori- ͑ ͒ zontal lines on diagrams ͑c͒–͑f͒ correspond to the close-packed basal planes. patible with hexagonal ordering of the atoms molecules within the plane was carried out in the papers by Vitebskii et al.,14–16,22 and it was shown in a paper by Gaididei and one of the authors12 that the wave vectors specifying the which need not be small, is considered below as an order magnetic superlattices of the ␣ and ␤ phases correspond to a parameter for these transitions. minimum of the quasi-classical magnetic energy of the crys- For a rigorous symmetry analysis instead of the param- tal when the only exchange interactions taken into account eter ␰ we consider the relative displacement vector u charac- are those between nearest neighbors in the basal plane. We terizing the mutual position of two adjacent planes in the note that those results remain valid for the ␦ phase in the different phases ͑upon a displacement of the order of ͉u͉ϰ␰ two-dimensional approximation. along one of the twofold axes͒. Since the states in which two However, for studying the mutual influence of the mag- parallel atomic planes are displaced by a vector multiple of netic and structural phase transitions arising in solid oxygen the lattice period are physically equivalent and in essence under pressure it is important to take into account not only identical, the vector u is defined up to a translation vector of the intralayer but also the interlayer exchange interaction, 1͒ the lattice in the basal plane. This imposes certain symme- which is more sensitive to the relative shift of adjacent try restrictions on the dependence of the internal energy on atomic planes. Accordingly, we shall classify the magnetic u. structures of the ␤, ␣, and ␦ phases in irreducible represen- For example, let us consider the interaction potential tations of the symmetry group of the paraphase, D1 , i.e., a (12) 6h V (u) of two adjacent close-packed planes, which, by vir- substantially three-dimensional rather than a two- tue of the periodicity properties mentioned above, can be dimensional lattice, and we write the wave vectors of the written in the form of a series magnetic superstructures in terms of the vectors b1 ϭ ␲ ␲ ) ϭ ␲ ) ␲ (2 /a,2 / a,0), b2 (0,4 / a,0), b3(0,0,2 /c)of ͑12͒͑ ͒ϭ ͑12͒ ͑ ͒ ͑ ͒ V u ͚ Vnm exp ignmu 1 n,m ϭ ϩ in the reciprocal lattice vectors gnm nb1 mb2 , parallel to ␺ ϭ the basal plane. The functions nm exp(ignm•u) form a rep- resentation of the symmetry point group of the ‘‘unshifted’’ paraphase, and their linear combinations can be chosen as the order parameter of the phase transitions described.2͒ For the sequence of transitions between the ␤, ␣, and ␦ phases con- sidered in this paper it is sufficient to choose as the order parameters of the structural transition the functions FIG. 2. Orientation of the magnetic moments of solid oxygen in the basal ␺ ␺ ␺ ␤ ͑ ͒ ␣ ␦ ͑ ͒ 01 , 10 , ¯11 , the transformations of which under the opera- plane of the phase a and of the and phases b . 766 Low Temp. Phys. 31 (8–9), August–September 2005 E. V. Gomonay and V. M. Loktev

The channel of a possible transition contains only one of the ͑ (␣) ͒ three arms i.e., only one of the vectors l1,2,3 is nonzero , and in this case the normalization conditions reduce to the rela- (␣) 2ϭ 2 ͑ ␦ ͒ tion (lk ) M 0 analogously for the phase . In practice (␣) (␣) (␣) the different antiferromagnetic vectors (l1 ,l2 ,orl3 ) cor- respond to different domains of the same magnetically or- dered phase. As to a description in terms of magnetic sublat- tices, according to the experimental data in the ␣ phase the ͑ number of such sublattices is two the star k12), while in the ␦ ͑ phase it is four the star k14). The star pairs (k13 ,k15) and (k12 ,k14) correspond to the same antiferromagnetic ordering of magnetic moments in the basal plane of the crystal but differ in respect of the mutual orientation of the spins in adjacent planes: parallel for the FIG. 3. Orientation of the direct and reciprocal lattices in the basal plane. stars k12 and k13 , and antiparallel for k15 and k14 .Aswe have said above, the majority of experiments done to date do not permit an unambiguous determination of the type of star ␤ the reciprocal lattice of the hexagonal lattice with vectors (k13 or k15) that characterizes the magnetic structure in the ϭ ϭ Ϫ ) ϭ ͑ ͒ a1 (a,0,0), a2 ( a/2, a/2,0), a3 (0,0,c) see Fig. 3 . phase. Contributing to this is the absence of long-range order The experimentally observed phases are compatible with in the basal planes themselves, although the short-range or- the following representations of the magnetization vector der corresponds unambiguously to the Loktev structure ͑see ͒ ␣ M(rj) at the point rj : Ref. 17 . As to the collinear magnetic structure of the and 1. In the ␤ phase ͑Fig. 2a͒ ␦ phases, the spin configurations we have chosen above 21 ͑␤͒ ͑␤͒ ͑␤͒ agree not only with the direct magnetic measurements but M ͑r ͒ϭRe͓͑l ϩil ͒exp͑iq␤ r ͔͒, j 1 2 • j also permit explanation of the observed19 jump of the order (␤) (␤) ␣ ␦ ͑ ͒ where the orthogonal vectors l1 ,l2 satisfy the normaliza- parameter at the - transition see below . Nevertheless, it tion conditions should be kept in mind that a two-sublattice magnetic struc- ␤ ␤ ture is also compatible with the lattice of the ␦ phase. ͑l͑ ͒͒2ϩ͑l͑ ͒͒2ϭM 2, 1 2 0 Upon application of an external magnetic field a macro- where M 0 is the saturation magnetization per unit volume. scopic moment arises in the crystal. In the exchange approxi- These vectors can be regarded as a two-component4͒ order mation the magnetization vector m transforms according to parameter, and the transition wave vector q␤ coincides with the unit irreducible representation, corresponding to the vec- the representation of one of the single-ray stars ͑in Kovalev’s tor kϭ0. In the description in terms of a multi-sublattice ͒23 1 ϭ ϩ ϭ notation of the group D6h : k13 (b1 b2)/3 or k15 (b1 model the vector m is equal to the sum of the magnetic ϩ ϩ (␤) (␤) b2)/3 b3/2. The antiferromagnetic vectors l1 ,l2 can moments of all the sublattices. also be regarded as the nonzero linear combinations of sub- lattice magnetic moments that determine the transition to the magnetically ordered phase, the number of sublattices being III. THERMODYNAMIC POTENTIAL equal to three or six, depending on the star (k13 and k15 , respectively͒, of the transition. On the phenomenological level the sequence of magne- 2. In the ␣ and ␤ phases ͑Fig. 2b͒ tostructural transitions between low-temperature phases of solid oxygen can be described on the basis of the Landau 3 theory of phase transitions, i.e., by analysis of the corre- ͑␣͒͑ ͒ϭ ͑␣͒ ͑ ͑␣͒ ͒ M rj ͚ lk exp iqk •rj ⌽͑ kϭ1 sponding thermodynamic potential the Gibbs potential in the presence of external fields and the Helmholtz free energy and in the absence of field͒. 3 In view of the the strong coupling between the magnetic ͑␦͒͑ ͒ϭ ͑␦͒ ͑ ͑␦͒ ͒ ͑ ͒ and structural subsystems of solid oxygen, as internal param- M rj ͚ lk exp iqk •rj . 2 kϭ1 eters describing the thermodynamic state of the crystal one (␣) ϭ can choose the magnetic moment density M(k)(r)(k The antiferromagnetic vectors lk , k 1,2,3 form the three- ϭ␤,␣,␦), the structure functions ␺ ͑which depend on the component order parameter, corresponding to the star k12 nm ϭ shift vector of the atomic planes͒, the specific volume v of b1/2 with the three arms the crystal, and the components of the strain tensor u .As Ϫ jk ͑␣͒ b1 ͑␣͒ b2 ͑␣͒ b2 b1 ͑ q ϭ , q ϭϪ , q ϭ , the variables responsible for the phase transitions primary 1 2 2 2 3 2 ͒ (k) order parameter , besides the magnetic vectors lj ( j ϭ and each of the antiferromagnetic vectors corresponds to one 1,2,3), we choose the real part of the functions ␦ ␺ ␺ ␺ ⌬ ( ) ϭ 01 , 10 , ¯11 and the relative volume change v/v, and the of the arms. Analogously, the vectors lk , k 1,2,3 corre- ϭ ϩ remaining combinations will be assumed to be secondary spond to the three-arm star k14 (b1 b3)/2, with the arms order parameters.5͒ The main parameters responsible for the ͑␦͒ ͑␣͒ b3 phase transitions and the accompanying parameters are listed q ϭq ϩ . j j 2 in Table II. Low Temp. Phys. 31 (8–9), August–September 2005 E. V. Gomonay and V. M. Loktev 767

TABLE II. Primary and secondary order parameters implementing irreduc- K2) has a power-law dependence on temperature, while the D1 ible representations of the group 6h of the paraphase. other coefficients are assumed constant. For solid oxygen one can choose the main parameter ‘‘controlling’’ the phase transformations to be the specific volume v, since both the value of the exchange magnetic interactions and the interac- tion between adjacent atomic planes depend substantially on the intermolecular distances and, as a consequence, on v. That is the reason why we have assumed that the coefficient K2 depends on the external parameters indirectly, through the relative volume change ⌬v/v. From a thermodynamic The thermodynamic potential represented as a function point of view, however, the volume itself is an internal pa- of the internal and external parameters can be written in the rameter of the crystal, and for establishing the relation with form of a sum of terms ⌽ and ⌽ which depend only on temperature and pressure the function f has been introduced mag str ͑ ͒ the magnetic and only on the structural parameters, respec- in 4 . In spite of the fact that with variation of the external tively, and also a term ⌽ that takes the interaction between parameters the specific volume of solid oxygen varies very int ⌬ them into account: strongly and the quantity v/v certainly cannot be consid- ered a small parameter, this function can be approximated to ⌽ϭ⌽ ϩ⌽ ϩ⌽ ͑ ͒ str mag int . 3 good accuracy by the expression In principle, the thermodynamic potential ⌽ contains yet an- ⌬v 1 ⌬v 2 1 ⌬v T other component associated with the orientation of the oxy- ͩ ͪ ϭ ͩ ͪ Ϫ ͩ ͪ ͵ ␤ ͑ Ј͒ Ј ͑ ͒ f ,T ␹ ␹ P T dT , 5 gen molecules. However, we have not written it down be- v 2 T v T v 0 cause it does not directly influence the sequence of ͑ magnetostructural phase transitions under consideration, al- where the ‘‘bare’’ not including the contributions due to the ͒ ␤ though by changing from phase to phase ͑and also with phase transitions coefficient of thermal expansion P(T)at ␹ changes in temperature and pressure͒ it can contribute to the constant pressure and the isothermal compliance T gener- ally speaking change on going from one temperature and observed values of the coefficients of volume expansion, etc. 7͒ ͑see below͒. pressure range to another, and the volume change is reck- To take into account the features of the phase transitions oned from some reference state, e.g., from the state with T ϭ ϭ ͑ ͒ observed in solid oxygen, we write the structure ͑lattice͒ po- 0, P 0. The presence of an integral in the last term in 5 ⌽ is due to the fact that, according to the experimental data, the tential str in the form coefficient of thermal expansion depends substantially on ⌽ ϭϪK ͑v͓͒cos 2␲␰ ϩcos 2␲␰ ϩcos 2␲͑␰ Ϫ␰ ͔͒ ␤ str 2 1 2 1 2 temperature. In the limit of a weak dependence P(T) Ϸ ͑ ͒ 1 const, expression 5 agrees with the analogous contribu- ϩ K ͓cos 4␲␰ ϩcos 4␲␰ ϩcos 4␲͑␰ Ϫ␰ ͔͒ tion to the free energy in the Landau model.30 In an analo- 4 4 1 2 1 2 gous way the coefficient K2 can be assumed to be a linear ⌬v function of volume, with a phenomenological coefficient ϩ f ͑v,T͒ϩP . ͑4͒ v ⌬v ͑ ͒ϵ Ϫ␭ ␲␰ ϵ K2 v K0 v . In this expression the quantities 2 1,2 b1,2•u are propor- v tional to the projections of the shift vector u onto the vectors of the hexagonal lattice, f (v,T) is some function of the spe- In expression ͑4͒ for the structural contribution to the cific volume v and temperature T which models the depen- free energy of solid oxygen only two of the five independent dence of the internal energy of the crystal on the interatomic parameters governing the finite strain of the crystal ͑see distances, and P is the external hydrostatic pressure. above͒ are taken into account explicitly: the relative dis- Expression ͑4͒ is in essence an analog of the Landau- placement vector of adjacent planes, since it determines the theory expansion of the thermodynamic potential in powers order parameter of the phase transitions ␤-␣-␦, and the rela- of the order parameter, which is invariant with respect to the tive volume change, since this parameter is isomorphic to the symmetry operations of the most symmetric of the phases. external parameters—the temperature and pressure. However, unlike the orthodox theory, here the expansion is When only the exchange interactions are taken into ac- in periodic functions of the shift vector u ͓actually in Fourier count, the magnetic energy can be written in the form of a harmonics; cf. formula ͑1͔͒, where it is assumed that the sum coefficients K ,K ,... of the expansion fall off in magnitude 2 4 2 3 with increasing number of the harmonic, so that, as usual, it ⌽ ϭϪ ͑ ͒ ͑ ͑␤͒͒2Ϫ ͑ ͒ ͑ ͑␣͒͒2 mag J k13 ͚ lj J k12 ͚ lj is sufficient to keep only the first few terms of the series. In jϭ1 jϭ1 the limit of small strains, these coefficients can be expressed 3 in terms of the shear modulus6͒ c and the anharmonicity 44 Ϫ ͑ ͒ ͑ ͑␦͒͒2Ϫ " ͑ ͒ J k14 ͚ lj m H, 6 constants. jϭ1 Formula ͑4͒ differs from the standard Landau approach Ͼ ϭ in another way as well. Usually for description of a phase where J(kj) 0, j 12,13,14 are the Fourier components of transition due to a change of temperature, it is sufficient to the exchange integral with the corresponding wave-vector assume that one of the expansion coefficients ͑for example, representatives of the stars ͑see Sec. 2.2͒. The last term in ͑6͒ 768 Low Temp. Phys. 31 (8–9), August–September 2005 E. V. Gomonay and V. M. Loktev arises on account of the the interaction of the macroscopic the magnetic superstructures, since only those of them cor- magnetic moment m with the external magnetic field H. responding to the observed magnetic structures have been ͑ ͒ The phenomenological quantities J(kj) include a contri- taken into account in expression 4 . bution from the exchange excitation both between nearest Analysis of the expressions appearing in Eq. ͑3͒ shows and, generally speaking, between more remote neighbors in that the most symmetric state, which in the given case cor- the basal plane, and also the interlayer exchange. However, responds to a hypothetical nonmagnetic paraphase, corre- taking into account the experimentally established17 fact that spond to zero values of the structural and magnetic param- the exchange interaction between nearest neighbors is domi- eters: ϭ nant, we can assume approximately that J(k12) J(k13) ϭ ϭ J(k14) J0 . ␰ ϭ␰ ϭ ͑␤͒ϭ ͑␣͒ ϭ ͑ ͒ 1 2 0, l1,2 l1,2,3 0. 9 The interaction between the magnetic and structural sub- systems, as we have said, is of a magnetoelastic nature and, This unobserved phase will not be considered further. in accordance with the symmetry of the paraphase, is mod- ␤ ␣ ␦ eled by the potential: The less-symmetric , , and phases, which neverthe- less preserve the symmetry with respect to twofold rotation 2 3 ⌬v ⌬v around one of the directions a Ϫa , a ϩ2a ,or2a ϩa in ⌽intϭ⌳͑␤͒ ͑ ͑␤͒͒2ϩ⌳͑␣͒ ͑ ͑␣͒͒2 1 2 1 2 1 2 ͚ lj ͚ lj v jϭ1 v jϭ1 the basal plane, correspond to three equivalent sets of param- eters with 3 ⌬v ϩ⌳͑␦͒ ͑ ͑␦͒͒2Ϫ⌳͑␣͓͒͑ ͑␣͒͒2 ␲␰ ͚ lj ʈ l1 cos 2 1 ϭ ␰ ϭϪ␰ ␰ ϭ ␰ ␰ ϭ␰ ͑ ͒ v j 1 1) 1 2 ,2)1 2 2 ,3)21 2 . 10 ϩ͑ ͑␣͒͒2 ␲␰ ϩ͑ ͑␣͒͒2 ␲͑␰ Ϫ␰ ͔͒ l2 cos 2 2 l3 cos 2 1 2 In the ␤ phase all of the states indicated in ͑10͒ are physi- Ϫ⌳͑␦͓͒͑ ͑␦͒͒2 ␲␰ ϩ͑ ͑␦͒͒2 ␲␰ ʈ l1 cos 2 1 l2 cos 2 2 cally indistinguishable, and the multivaluedness arises be- cause of additional symmetry of the paraphase—the close- ϩ͑ ͑␦͒͒2 ␲͑␰ Ϫ␰ ͔͒Ϫ⌳͑␣͕͒͑ Ϫ ͒͑ ͑␣͒͒2 l3 cos 2 1 2 Ќ uxx uyy l1 packed state of two adjacent hexagonal planes can be ϩ ͓͑ ͑␣͒͒2Ϫ͑ ͑␣͒͒2͔͖Ϫ⌳͑␦͕͒͑ Ϫ ͒͑ ͑␣͒͒2 obtained with the aid of a shift along any of the directions 2uxy l2 l3 ʈ uxx uyy l1 indicated above.8͒ ϩ ͓͑ ͑␣͒͒2Ϫ͑ ͑␣͒͒2͔͖ ͑ ͒ ␣ ␦ ͑ ͒ 2uxy l2 l3 . 7 In the and phases the different states 10 correspond to different structural domains of the same phase, differing The coefficients ⌳(k) (kϭ␤,␣,␦) are responsible for the by the direction of the twofold symmetry axis in the basal volume effect arising at the transition between different mag- (␣) plane. For definiteness we shall consider the first of the pos- netic phases, while the terms with the coefficients ⌳ʈ and (␦) sibilities indicated in Eq. ͑10͒, corresponding to the direction ⌳ʈ are responsible for the magnetoelastic contribution, a Ϫa ͑the other two can be analyzed in the same way͒. which describes the change of the interlayer exchange inter- 1 2 For further analysis we write conditions ͑8͒ for a mini- action upon a mutual displacement of the basal planes. The ͑ ͒ ␰ ⌳(␣) ⌳(␦) mum of the potential 3 with respect to the variables last two terms, with Ќ and Ќ , describe the result of ϵ␰ ϭ␰ ⌬ ͑ 2 1/2 and v/v in explicit form the analogous condi- Ref. 12 for the dependence of the intraplane exchange inter- (k) tions for the magnetic vectors lj are trivial and are not action on the corresponding intraplane shift of the rhombic ͒ ͑ ͒ Ϫ shown : strain see Ref. 12 , uxx uyy , which lifts the degeneracy of the intermolecular distances in the basal plane. Unlike the ⌽ץ -exchange constants J(k ) introduced above, all of the coef j ϭ ␲ ␲␰͕ ͑ ͒͑ ϩ ␲␰͒ ␰ 4 sin 2 K2 v 1 2 cos 2ץ -ficients of the magnetoelastic interaction that appear in ex pression ͑7͒ are governed predominantly by the long-range Ϫ ␲␰͑ ϩ ␲␰͒ structural order. For our model it is important that these co- K4 cos 2 1 2 cos 4 efficients are substantially different for the collinear ͑␣ and ␦ ͑␣͒ ͑␣͒ 2 ͑␦͒ ͑␦͒ 2 ϩ2͓⌳ʈ ͑l ͒ ϩ⌳ʈ ͑l ͒ ͔cos 2␲␰͖ϭ0; phases͒ and triangular ͑␤ phase͒ magnetic structures. 1 1

⌽ 1 ⌬v 1 Tץ ϭ Ϫ ͵ ␤ ͑ Ј͒ Јϩ ϩ␭ ͑ ␲␰ IV. EQUILIBRIUM PHASES AND TRANSITIONS BETWEEN P T dT P v 2 cos 2 v/v͒ ␹ v ␹ 0⌬͑ץ THEM T T 2 For determination of the thermodynamic equilibrium ϩ ␲␰͒ϩ⌳͑␤͒ ͑ ͑␤͒͒2ϩ⌳͑␣͒͑ ͑␣͒͒2 cos 4 ͚ lj I1 phases and the conditions of their stability we use the stan- jϭ1 dard condition of a minimum of the free energy potential ͑3͒, ϩ⌳͑␦͒͑ ͑␦͒͒2ϭ ͑ ͒ namely: I1 0, 11 2⌽ץ ⌽ץ ϭ0, Ͱ Ͱ Ͼ0 ͑8͒ where we have taken into account that in the ␣ and ␦ phases xץ xץ xץ p p q the magnetic state is determined solely by one antiferromag- ␰ ⌬ (k) ϭ␤ ␣ (␣) (␦) with respect to the parameters 1,2 , v/v, lj , k , ; j netic vector, l1 and l1 , respectively. ϭ ͓ ͑ ͔͒ 1,2,3 all denoted x p in Eq. 8 . Here it is unnecessary to The conditions of stability against small perturbations of minimize the free energy with respect to the wave vectors of the parameters are obtained from ͑8͒; they have the form Low Temp. Phys. 31 (8–9), August–September 2005 E. V. Gomonay and V. M. Loktev 769

2⌽ץ ϭ ␲2͕ ͑ ͒͑ ␲␰ϩ ␲␰͒ ␰2 16 K2 v cos 2 2 cos 4ץ

Ϫ ͑ ␲␰ϩ ␲␰͒ϩ ͓⌳͑␣͒͑ ͑␣͒͒2 K4 cos 4 2 cos 8 2 ʈ l1 ϩ⌳͑␦͒͑ ͑␦͒͒2͔ ␲␰͖Ͼ ʈ l1 cos 4 0; 2⌽ץ 2⌽ץ Ͼ ϭ16␲2␹2 ␭2 sin2 2␲␰͑1 v/v͒ T v⌬͑ץ␰ץ ␰2ץ ϩ ␲␰ 2 ␹ Ͼ ͑ ͒ 2 cos 2 ͒ , T 0. 12

A. ␤ phase It is seen from Eqs. ͑11͒ that the possible equilibrium values of the order parameter ␰ depend on the type of mag- netic ordering realized in the crystal. In particular, if it is the triangular Loktev magnetic structure, described by the vec- (␤) (␤) 2ϩ (␤) 2ϭ 2 (␣)ϭ (␦)ϭ tors l1,2 , with (l1 ) (l2 ) M 0, and l1 l1 0, then ͑ ͒ ␰ϭ the first of Eqs. 11 has the solution 1/3, corresponding FIG. 4. Pressure dependence of the relative change of the specific volume at to the ␤ phase, in which the adjacent planes are shifted rela- room temperature. The points are experimental,17 the solid line is a linear Ϫ ␹ ϭ ϫ Ϫ11 3 tive to each other by (a1 a2)/3. The stability of this phase approximation with the coefficient T 0.27 10 cm /dyn. with respect to the long-wave shear vibrations and hydro- static pressure is determined by the relation

T periment also allows one to determine the dependence K Ϫ␭ ͫ ͵ ␤ ͑TЈ͒dTЈϪ␹ PͬϽ0; K Ͼ0; ␹ Ͼ0, eff v P T 4 T ␤ (T). For example, on the basis of an analysis of the data 0 P ͑13͒ of Ref. 6, at atmospheric pressure Ϫ4 Ͻ where for convenience we have introduced the effective con- 1.6•10 1/K, T 11 K, stant ␤ ͑ ͒ϭ Ϫ5 Ϫ Ϫ4 ͑ ͒ P T ͭ 6.66•10 T 5.72•10 1/K, 11 16 3 KϽTϽ43.5 K. K ϭK ϪK Ϫ ␹ ␭2ϩ␹ ␭ ⌳͑␤͒M 2. ͑14͒ eff 0 4 2 T v T v 0 ␣ ␣ ␤ ␭ Ͼ B. phase and the - transition Since, as will be seen from the following, v 0, the stability conditions ͑13͒ are strengthened with increasing temperature If the magnetic structure of the crystal is two-sublattice, ͉ (␣)͉ϭ ͑ ͒ and weakened with increasing pressure; taken all together, with l1 M 0 , then Eqs. 11 have a solution with an order this is in complete agreement with the observed existence parameter that, depending on the external parameters, lies in region of the ␤ phase ͑see Ref. 17͒. the range 1/3Ͻ␰Ͻ1/2, corresponding to the ␣ phase. The The appearance ͑in comparison with the paraphase͒ of a value of ␰ is determined by the following equation, which is nonzero order parameter leads to a lowering of the symme- written for convenience in terms of the variable ␩ϵ1 try: the rotations about the sixfold axis vanish, while the ϩ2 cos 2␲␰, which from a macroscopic point of view can be rotations about the threefold axis, which are weighted by the considered to be the order parameter of the phase transitions ϩ ␤ operation of translation by the lattice vector a1 a2 in the examined below, since its value in the phase is equal to basal plane, remain. Of the order parameters corresponding zero: to the transition, only those appear which transform accord- ͑␣͒ ͑␤͒ ͑␣͒ 2 ing to the irreducible representation Ag1 , i.e., hydrostatic ͭ K ϩ͓⌳ʈ ϩ␹ ␭ ͑⌳ Ϫ⌳ ͔͒M eff T v 0 ͑isomorphic͒ compression-extension in the basal plane ͑the u ϩu parameter xx yy) and the change of the interplane dis- T 3 tances ͑determined by u ). Ϫ␭ ͩ ͵ ␤ ͑TЈ͒dTЈϪ␹ Pͪ ␩ϩ K ␩2 zz ͮ v P T 4 The important characteristic of the ␤ phase is the tem- 0 2 perature and pressure dependence of the relative volume 1 2 3 ͑␣͒ 2 change of the crystal: Ϫ ͑K Ϫ␹ ␭ ͒␩ ϭ⌳ʈ M . ͑17͒ 2 4 T v 0 ⌬v E ͑ ͒ͯ ϭ ͵ ␤ ͑ Ј͒ ЈϪ␹ Ϫ␹ ⌳͑␤͒ 2 The structural stability of the ␣ phase with respect to varia- T,P P T dT TP T M 0 v ␤ 0 tion of ␰ is then determined by the condition

3 T ϩ ␹ ␭ . ͑15͒ Ϫ␭ ͵ ␤ ͑ Ј͒ ЈϪ␹ Ͻ Ϫ ⌳͑␣͒ 2 T v K ͩ T dT Pͪ K 2 ʈ M eff v P T 4 0 2 0

The linear dependence of the specific volume on P is Ϫ␹ ␭ ͑⌳͑␣͒Ϫ⌳͑␤͒͒M 2, ͑18͒ observed experimentally17 over a wide range of pressures, as T v 0 can be seen, for example, in Fig. 4. A comparison with ex- and the energetic favorability, by the inequality 770 Low Temp. Phys. 31 (8–9), August–September 2005 E. V. Gomonay and V. M. Loktev

TABLE III. Values of the main phenomenological coefficients. that this magnetostructural transition is close to second-order is apparently due to the fact that the ‘‘heavy’’ ͑inertial͒ subsystem—the molecular lattice—can undergo an ␣-␤ transformation continuously. However, it does not occur be- fore the initiation of a jumplike rearrangement of the mag- netic ordering of the ‘‘easy’’ spin ͑and hence, electronic͒ sub- system of the oxygen. We note that the scenario of a smooth ␣-␤ transformation was set forth in Ref. 17. Here it should be emphasized that the isolation of the role of magnetic interactions in the ␣-␤ transition is of a conventional character. Actually there is a simultaneous change of the crystalline ͑a shift and straining of the basal planes͒ and magnetic ͑transition from the noncollinear to the collinear structure͒ subsystems caused by the energetic fa- vorability of the new state. There is no doubt that an impor- tant role in this transformation is played by the specific vol- ume change, which in the ␣ phase is described by the expression

⌬v ⌬v ͑ ͒ͯ ϭ ͑ ͒ͯ Ϫ␹ ͑⌳͑␣͒Ϫ⌳͑␤͒͒ 2 T,P T,P T M 0 v ␣ v ␤ 1 Ϫ ␹ ␭ ␩2, ͑20͒ 2 T v

⌬ ͑ ͒ 1 1 1 where ( v/v)␤ is given by formula 15 . It is seen that at the ⌽ Ϫ⌽ ϭϪ K ␩2Ϫ K ␩3ϩ ͑K Ϫ␹ ␭2͒␩4 transition to the ␣ phase the specific volume of the crystal ␣ ␤ 2 eff 2 4 8 4 T v should decrease in a jump ͑both terms on the right-hand side 1 1 ⌬v 2 are negative͒. This decrease is due to both the difference of ϩ⌳͑␣͒ 2ͩ ␩Ϫ ␩2 ͪ ϩ ͩ ͪ ʈ M ͫ ␣ ␤ 0 2 2␹ v the magnetoelastic volumes effects in the and phases T ␤ ͓the second term in ͑20͔͒ and to the incipient shift of the ⌬v 2 basal planes ͑the last term͒. Ϫͩ ͪ ͬϽ0, ͑19͒ ͑ ͒ v It is seen from an analysis of expression 7 and Table II ␣ that the appearance of ␩0 due to the change of the mag- ͑ (␣) which can be inferred from a comparison of the thermody- netic structure the appearance of l1 0) also leads to ͑ namic potentials of the two phases at fixed values of the straining of the basal plane described by the quantity uxx Ϫ external parameters ͑cf. Ref. 12, where similar relations were uyy), but a treatment of the corresponding effects is be- written for the case Tϭ0, Pϭ0). The sign of the quantities yond the scope of this paper. on the right-hand side of inequality ͑18͒ are in essence de- Furthermore, the specific volume v actually causes the ͑ termined by the interplay of several constants and can be effective exchange interaction constants Jeff i.e., the coeffi- ͑ 2 positive or negative. According to our estimates see Table cient of M 0, which in the given model determines the stabil- III͒, the last term on the right-hand side is negligible, and the ity of the magnetic subsystem and the value of the magnetic first two terms are positive. Thus inequality ͑18͒, which de- susceptibility͒ to depend on the external parameters. We note termines the stability region of the ␣ phase, is not incompat- that in the model adopted ͓see formulas ͑4͒ and ͑5͔͒ the ible with ͑and is even weaker than͒ the existence condition influence of temperature and pressure are equivalent. How- of the ␤ phase ͓Eq. ͑13͔͒. In other words, the two phases can ever, for definiteness, in considering the ␣-␤ transition we coexist in a certain interval of temperature and pressure. shall treat temperature as the control parameter, assuming the What, then, leads to the observed ␣-␤ transition in the pressure to be fixed and equal to atmospheric. The features framework of the phenomenological description? of the ␣-␤ transition arising at other values of the external As we see from Eq. ͑17͒, the ‘‘driving force’’ of this pressure are discussed below in Sec. 4.3. transition, analogous to an ‘‘external field,’’ is the possibility The linear character of the dependence Jeff(v) is con- of lowering the free energy by a change of the magnetic firmed, for example, by the the experiments of Refs. 6 and (␣) structure, specifically: owing to the term ⌳ʈ due to the 24, the results of which were used to construct the plots of ␩ϭ ͑ 2 collinear magnetic ordering, the solution with 0, corre- the inverse magnetic susceptibility actually JeffM0)asa sponding to the ␤ phase, vanishes on the right-hand side of function of the specific volume change of the crystal in the ␣ this equation. Since, when the collinear magnetic order is and ␤ phases ͑at atmospheric pressure͒ in Fig. 5. In accor- ‘‘switched on,’’ the ‘‘driving force’’ immediately takes on dance with expressions ͑6͒ and ͑7͒ and with the assumptions some finite value, the ␣-␤ transition must inevitably occur as made above as to the short-range character of the intraplane a first-order transition ͑although close to second-order in the exchange interaction, we have for the effective exchange Landau classification͒. From a physical standpoint the fact constant Low Temp. Phys. 31 (8–9), August–September 2005 E. V. Gomonay and V. M. Loktev 771

FIG. 5. Dependence of the inverse magnetic susceptibility on the relative volume change, constructed according to the data of Refs. 6 and 24. The solid lines are the theoretical curves calculated according to formula ͑21͒ 2ϭ ϫ 8 3 ⌳(␤) 2ϭ with the parameters J0M 0 2.54 10 erg/cm , M 0 0.82 FIG. 6. Temperature dependence of the order parameter ␩ϭ1ϩ2cos 2␲␰, ϫ 9 3 ⌳(␣) 2ϭ ϫ 9 3 10 erg/cm , M 0 8.2 10 erg/cm . calculated according to the values of the lattice constants from Ref. 6 ͑points͒. The arrow indicates the temperature of the ␣-␤ transition. The solid line is the result of a calculation according to formulas ͑17͒ and ͑16͒.The ϩ⌳͑␣͒ ␲␰Ϫ⌳͑␣͒⌬ ␣Ϫ inset shows the same dependence, which shows the range of variation of ␩ J0 ʈ cos 4 v/v, O2 , ␣ J ϭͭ ͑21͒ in the phase. eff Ϫ⌳͑␤͒⌬ ␤Ϫ J0 v/v, O2 . Approximating the experimental data according to ͑21͒ al- favor the onset of a two-sublattice ͑type ␣͒ and not three- lows one to estimate the values of the constants appearing in sublattice ͑type ␤͒ magnetic structure. Because of the pres- it ͑see Table III͒. The corresponding value of the ‘‘bare’’ ence of this mutual influence of the magnetic and lattice J M 2Ӎ ϫ 8 3 exchange interaction 0 0 3 10 erg/cm , which, with subsystems, the phase transition is, as we have said, decid- allowance for the value of the saturation magnetization M 0 ϭ ͑ ͒ edly first-order: the relative displacement of the close-packed 133 G calculated from the data of Ref. 6 , gives an esti- planes varies in a jump from ␰ϭ1/3 ͑in the ␤ phase͒ to ␰ mate for the spin-flip field of the order of 230 T, in complete у␰ ͑in the ␣ phase͒. The point at which this phase transi- agreement with the value 227 T given in the review.17 cr ͑ ͒ tion occurs at fixed external parameters is determined from From relations 21 and Fig. 5 it is clearly seen that a the condition of equality of the thermodynamic potentials change of the magnetic structure can occur only with a si- ͓see Eq. ͑19͔͒. multaneous change of the crystal structure. For example, it is The temperature dependence of the order parameter ␩ obvious that at certain values of the external parameters the introduced in Eq. ͑17͒ and calculated according to that for- spin coefficient that corresponds to the larger ͑in absolute ͒ mula is shown in Fig. 6 by the solid curve. The values of the value effective exchange interaction constant is realized, phenomenological constants ͑see Table III͒ were determined ͒ ͑ 2Ͼ ץ 2⌽ץ since here / M 0 0. Hence on the basis of formula 21 mainly from the dependence of the lattice parameters on we obtain the condition of stability of the collinear magnetic pressure ͑from the data of Ref. 19͒, and the coefficient of ␣ structure corresponding to the phase: thermal expansion was calculated from the empirical relation ⌬v ͑16͒. The points show the same dependence constructed ac- ͑␣͒ ͑␤͒ ͑␣͒ ͒ ͑⌳ Ϫ⌳ ͒ Ϫ⌳ʈ cos 4␲␰Ͼ0, ͑22͒ 6 9 v cording to the values of the lattice constants. It is seen that the calculated dependence correctly reflects the character of where the relative change of volume is driven by expression the decline of the order parameter with temperature. The ͑ ͒ ͑ ͒ 20 . Inequality 22 holds in the case when the order param- transition to the ␣ phase occurs at the value ␰ϭ0.367, which, ␰ eter exceeds a certain critical value it must be acknowledged, is in fair agreement ͑considering 3 ͓⌳͑␣͒Ϫ⌳͑␤͔͒ ⌬v the substantial error that unavoidably arises in the calculation ␰у␰ Ϸ ϩ ͑ ͒ ͒ ␰ ϭ cr ͑␣͒ . 23 of the order parameter with the value cr 0.376 calculated 8 4␲⌳ʈ v according to formula ͑23͒. It should be noted, however, that Thus the transition to the ␣ phase is possible only when the range of variation of the order parameter in the ␣ phase is conditions ͑18͒ and ͑22͒ are satisfied simultaneously. The substantially less than its absolute value, as is seen in the first condition means that the lattice is ‘‘soft’’ enough to shift inset of Fig. 6. the required amount under the influence of the comparatively The temperature dependence of the relative change of small magnetic forces, and the second means that during the molar volume calculated according to formulas ͑15͒, such a rearrangement of the lattice the exchange interactions ͑20͒, and ͑16͒, is shown in Fig. 7. The points are experimen- 772 Low Temp. Phys. 31 (8–9), August–September 2005 E. V. Gomonay and V. M. Loktev

cific volume ͓the last term in ͑19͔͒, the value of which at the transition point (Tϭ23.5 K,Pϭ0) equals 8ϫ107 erg/cm3, and a negative contribution due to the appearance of the order parameter ͑all the other terms͒, the value of which is 5ϫ108 erg/cm3. Since the estimation of all the phenomeno- logical parameters is rather crude ͑not better than 10%͒, the heat of transition thus obtained corresponds to an upper bound on the experimentally observed values. Another argument in favor of a first-order transition is the coexistence region of the ␣ and ␤ phases predicted by the model, and the consequent possibility of hysteresis effects. The inset in Fig. 7 shows the change of the molar volume on cooling in the vicinity of the ␣-␤ transition point, plotted according to the data of Ref. 25. The solid curve corresponds ␤ to the theoretical approximation with a coefficient P(T) ϭ3.0ϫ10Ϫ3 1/K and, apparently, it is precisely in the coex- istence region ͑where the theory of phase transitions is tech- nically correct͒ that the ␣ phase is described by the same phenomenological parameters as the ␤ phase. On further de- crease of temperature the value of the coefficient of thermal ␤ expansion P(T) changes substantially; this cannot be inter- preted in the given model.

C. ␦ phase and phase transitions under pressure Equation ͑11͒ has yet another ‘‘crystalline’’ solution, which formally does not depend on the type of magnetic ordering. If ␰ϭ1/2 ͑or ␩ϭϪ1) then the crystal structure FIG. 7. Temperature dependence of the relative change of the specific vol- corresponds to the ͑orthorhombic͒ ␦ phase, in which the mol- ͑ ͒ ume at atmospheric pressure . The points are experimental, the solid curve ecules in every alternate basal plane lie directly above one shows the approximation according to formulas ͑15͒, ͑20͒, and ͑16͒.The dashed curve is the extrapolation of the dependence for the ␤ phase into the another ͑see Fig. 1f͒. Also entering the stability condition of lower-temperature region. The inset shows the change of the molar volume the ␦ phase are the constants determining the magnetoelastic on cooling in the vicinity of the ␣-␤ transition point ͑according to the data interactions ͓cf. Eq. ͑18͔͒: of Ref. 25͒. The solid curve is an approximation with a single fitting param- eter, ␤ (T)ϭ3.0ϫ10Ϫ3 1/K. T P Ϫ␭ ͵ ␤ ͑ Ј͒ ЈϪ␹ Ͼ Ϫ ⌳͑␦͒ 2 K ͩ T dT Pͪ K 2 ʈ M eff v P T 4 0 0 6 tal data. The dashed line shows the extrapolation of the Ϫ␹ ␭ ͑⌳͑␦͒Ϫ⌳͑␤͒͒M 2. ͑24͒ curve for the ␤ phase into the low-temperature region ac- T v 0 cording to formulas ͑15͒ and ͑16͒; this corresponds to the Analysis of Eqs. ͑17͒, ͑11͒ and condition ͑24͒ shows that thermal expansion of the lattice if the rearrangement of the there are two ways of getting to the ␦ phase in solid oxygen: lattice at the phase transition is not taken into account. directly from the ␤ phase ͑through rearrangement of the As is seen from a comparison with experiment, the pro- magnetic structure͒, and from the ␣ phase ͑through a gradual posed model satisfactorily describes the main features of the shift of the close-packed planes͒. The parameter that deter- observed phase transformation—the jump of the specific vol- mines which of these paths is realized for given values of the ume and the nonlinear dependence on temperature in the pressure and temperature is again the specific volume of the transition region. crystal, the relative change of which ͑relative to the state We note another important circumstance. A symmetry with Tϭ0, Pϭ0) is equal to ␣ ␤ analysis based on the model developed indicates that the - ⌬v ⌬v transition should be first-order, since a fundamental rear- ͑T,P͒ͯ ϭ ͑T,P͒ͯ Ϫ͑⌳͑␦͒Ϫ␹ ⌳͑␤͒͒M 2 v v T 0 rangement of the magnetic structure occurs ͑a transition from ␦ ␤ ͒ three-sublattice to two-sublattice; see above . Nevertheless, 1 from an experimental point of view the ␣-␤ transition can be Ϫ ␹ ␭ . ͑25͒ 2 T v first- or second-order,17 since the heat of transition is small ͑not more than 100 J/mole, which when expressed per unit Indeed, as we have said, the ␤ phase loses stability be- volume comes out to 5ϫ107 erg/cm3). However, the evi- cause of a change of the magnetic structure. In turn, the dence pointing to a first-order transition comes not only from collinear magnetic ordering becomes energetically favorable qualitative but also from quantitative arguments. For ex- to the Loktev structure only when condition ͑23͒ is satisfied. ample, it follows from expression ͑19͒ that the difference of In other words, the value of the basal plane shift that arises the thermodynamic potentials of the ␣ and ␤ phases is deter- simultaneously with the collinear magnetic structure must mined by the difference of two contributions which are close again exceed a certain critical value that depends on the in value: a positive contribution due to the jump of the spe- value of the specific volume. At low pressure and low tem- Low Temp. Phys. 31 (8–9), August–September 2005 E. V. Gomonay and V. M. Loktev 773

FIG. 8. Pressure dependence of the order parameter ␩ϭ1ϩ2 cos 2␲␰ at FIG. 9. Pressure dependence of the relative change of the specific volume at Tϭ19 K ͑points͒, calculated using the values of the lattice constants from Tϭ19 K, calculated according to the values of the lattice constants from Ref. 19. The solid curve is the theoretical approximation calculated accord- Ref. 19. The solid curve is the theoretical approximation calculated accord- ing to formula ͑17͒. The dashed line shows the region of the transition ing to formula ͑20͒. The dashed line shows the region of the transition between the ␣ and ␦ phases. between the ␣ and ␦ phases.

The experimental data19 plotted in Fig. 8 show, however, ␰ peratures the value of the critical shift cr lies in the range that in reality the pressure-induced transition from the ␣ to Ͻ␰ Ͻ ␤ 3/8 cr 1/2, and the unstable phase transforms only to the ␦ phase occurs earlier, at a pressure of Ӎ6.5– 7 GPa ͑the the ␣ phase. However, with increasing pressure an increase transition region is denoted in the figure by the vertical ͒ of the negative volume change occurs10 ͓see Eqs. ͑15͒ and dashed line͒, and the order parameter changes in a jump at ͑ ͔͒ ␰ 25 and, accordingly, cr grows to a limiting value of 1/2 that point. Such behavior of the parameter ␩ permits the after the stimulated transition from the ␤ phase to the ␦ phase assertion that, in accordance with the observations,21 the occurs under the influence of the magnetic interaction and magnetic structure of the ␦ phase, although collinear, is dif- the magnetic rearrangement. This transition, like the ␣-␤ ferent from that of the ␣ phase ͑see Sec. 2.2͒. Formally it transition, is first-order and should be accompanied by the follows from this that the corresponding magnetoelastic in- ͓ ␣ (␣) (␦) (␣) (␦) release of heat cf. the analogous expression for the phase, teraction constants (⌳ ,⌳ ,⌳ʈ ,⌳ʈ ) in the two phases Eq. ͑19͔͒: are different. The condition for loss of stability of the collin- ear two-sublattice structure containing the ␣ phase, accord- 1 5 1 2 1 ͑␦͒ 2 ⌽ Ϫ⌽ ϭϪ K ϩ K Ϫ ␹ ␭ Ϫ ⌳ʈ M ingly, has a form analogous to expression ͑22͒: ␦ ␤ 2 eff 8 4 8 T v 2 0 ⌬v 2 2 ͑␦͒ ͑␣͒ ͑␣͒ ͑␦͒ 1 ⌬v ⌬v ͑⌳ Ϫ⌳ ͒ ϩ⌳ʈ cos 4␲␰Ϫ⌳ʈ Ͼ0, ͑27͒ ϩ ͫͩ ͪ Ϫͩ ͪ ͬ. ͑26͒ v 2␹ v v T ␤ ␦ (␦) 9 3 which gives the estimate ⌳ʈ ӍϪ3ϫ10 erg/cm . Let us now consider the question of how the ␣-␦ transi- The jumplike character of the ␣-␦ transition at high pres- tion occurs. Analysis of Eq. ͑17͒ shows that the value of the sures is also discernable on the pressure dependence of the order parameter ͑i.e., the shift vector͒ depends substantially relative volume change shown in Fig. 9. The theoretical on the pressure. This fact is illustrated in Fig. 8, which shows curve ͑solid line͒ calculated according to formulas ͑25͒ and the pressure dependence of ␩ at Tϭ19 K ͑the corresponding ͑15͒ gives good agreement with the experimental data19 11͒ values of the phenomenological parameters are given in ͑points͒ up to 6 GPa, after which a significantly smoother ␹ ϭ ϫ Ϫ11 Table III; T 0.75 10 cm/dyn). It is seen that an in- dependence of ⌬v/v(P) is observed. crease of pressure causes a gradual ͑but not small͒ shift of Thus the ␣-␦ transition occurs as a first-order phase tran- the close-packed planes. When the limiting value ␰ϭ1/2 is sition. The cause of this, as for the transition between the ␣ reached, the symmetry of the crystal increases to orthorhom- and ␤ phases, is a rearrangement of the magnetic structure of bic. This also corresponds to the transition to the ␦ phase. the crystal from two collinear sublattices to four. The critical pressure above which the symmetry of the lattice no longer changes12͒ is calculated as 9 GPa. Generally V. PHASE TRANSITIONS IN THE PRESENCE OF EXTERNAL speaking, such a transition could occur as a second-order transition, since ͑under the condition that the magnetic struc- ture of the ␣ and ␦ phases is identical͒ the two phases are As we have said more than once, an important role in the found in a subgroup relationship, and the order parameter sequence of transitions between the ␣, ␤, and ␦ phases of can vary in a continuous manner. solid oxygen is played by the magnetic interaction. This 774 Low Temp. Phys. 31 (8–9), August–September 2005 E. V. Gomonay and V. M. Loktev

contribution to the change in specific volume from the ap- pearance of a mean magnetic moment. Nevertheless, even on the basis of the available data one can conclude that ͑in any case, in a defectless crystal͒ the field dependence of the spe- cific volume should be only of a quadratic character. Further- more, it can be assumed that at high enough pressures, close to the region of the transition to the ␦ phase, the ␣-␦ struc- tural transition can also be induced by a field ͑the proposed phase transition point at Pϭ5 GPa is indicated by an arrow in Fig. 10͒. The application of a magnetic field at Pϭ0 can lead to a thermomagnetic effect: specifically, a shift of the temperature of the ␤-␣ transition by an amount

͑␣͒ 2 2 ⌳ʈ M ͑1Ϫ␩͒ H ⌬ ϭ 0 ͩ ͪ ͑ ͒ T ␭ ␤ ͑ ͒␩ , 28 v P T HE which at a field Hϭ20 T should amount to around 2 K. Interestingly, according to the model, a magnetic field should shift the transition point to higher temperatures, i.e., it seem- ingly should broaden the existence region of the ␤ phase according to the predictions of Ref. 31, which were based on an analysis of the magnon spectrum. However, as is seen FIG. 10. The value of the order parameter ␩ϭ1ϩ2 cos 2␲␰ ͑a͒ and relative (␣) 2 from Eq. ͑17͒, the term of a magnetic nature (⌳ʈ M ) to- change of the specific volume ͑b͒ as functions of the crystal magnetic field 0 at Pϭ0 ͑᭹͒,2͑᭡͒, and 5 ͑*****͒ GPa. Temperature Tϭ19 K; spin-flip gether with the temperature and pressure renormalizes the field 227 T. constant Keff , which determines the stiffness of the crystal with respect to the magnetic ‘‘driving force.’’ Ultimately, the sign of ⌬T is determined by the balance of the two tenden- naturally raises the question: Can these transitions be con- cies indicated, specifically, the weakening of the exchange trolled by applying to the crystal an external magnetic field interaction responsible for the antiparallel ordering of the sufficient to alter the magnetic structure of the crystal? magnetic moments in the external field and the increase of Since all three of the phases of solid oxygen ͑␣, ␤, ␦͒ the compliance of the lattice with respect to a shift of the discussed in this paper ͑in the absence of external magnetic close-packed planes. Nevertheless, the precise behavior of field͒ are compensated antiferromagnets, the application of the lines of phase transitions and their dependence on the an external magnetic field will lead, first, to a change of the external parameters can be established only experimentally. absolute value of the antiferromagnetic vector. Taking into The possibilities of both the microscopic and the phenom- 2ϩ 2ϭ 2 ͑ enological approaches still remain limited. account the normalization conditions l m M 0 l is one of the antiferromagnetic vectors introduced above, in the ␣ It is also not ruled out that the external magnetic field ␤ ␤ or ␦ phase͒ and (l( ))2ϩ(l( ))2ϩm2ϭM 2, and also the re- can have a noticeable influence on the value of the shear 1 2 0 ͑ lation mϭH/H ͑m is the magnetization, H ϭ227 T is the modulus and, as a consequence, on the velocity of trans- E E ͒ collapse field7͒, it is easy to obtain expressions for determin- verse sound , which is determined by the constant Keff , but ing the order parameter and the phase equilibrium conditions the discussion of that question is beyond the scope of this in an external field. For this one should make the following paper. substitution in all the formulas obtained previously: 2 H 6. CONCLUSION M 2→M 2ͩ 1Ϫ ͪ . 0 0 H2 E In this paper we have analyzed on the basis of a phe- Thus the magnetic field, as it were, decreases the ‘‘driv- nomenological model the phase stability conditions and the ing force’’ ͓see expression ͑17͔͒ that leads to a shift of the conditions for phase transitions between different magneto- basal planes, but it also influences the phase stability condi- crystalline structures of solid oxygen over a wide range of tions ͓expressions ͑18͒ and ͑24͔͒ and the value of the specific pressure, temperature, and external magnetic field. volume ͓expressions ͑15͒, ͑20͒, and ͑25͔͒. In particular, as an We have shown that the sequence of observed transitions analysis of Eq. ͑17͒ shows, the application of a field in the ␣ between the ␣, ␤, and ␦ phases can be described with the use phase can lead to an additional shift of the basal planes ‘‘in of a single structural order parameter—the shift vector of the direction’’ of the ␦ phase ͑i.e., to an increase of ␰͒. The two adjacent close-packed basal planes. As the main param- corresponding curves of the order parameter and specific vol- eter that determines the dependence of the value of the shear ume as a function of external field for different values of the vector on the external thermodynamic parameters—the tem- external pressure (Tϭ19 K), calculated according to for- perature and pressure—one can use the specific volume of mula ͑17͒, are shown in Fig. 10. It should be emphasized that the crystal ͑or, from a microscopic point of view, the change the curves shown are more of a qualitative character, since of the intermolecular distances͒, the change ⌬v of which can because of the lack of data they do not take into account the be rather large (⌬vϳv). Low Temp. Phys. 31 (8–9), August–September 2005 E. V. Gomonay and V. M. Loktev 775

The character of the magnetic ordering also has a sub- approach first proposed by Landau for the description of an- stantial influence on the crystal structure. For example, it tiferromagnetically ordered systems remains applicable for follows from an analysis of the conditions for a minimum of such an unusual object from the theoretical and experimental the thermodynamic potential that collinear ordering is com- standpoints as solid oxygen, where the coupling of the mag- patible only with the crystal structure of the ␣ and ␦ phases. netic and elastic subsystems in principle cannot be neglected. Furthermore, the value of the shift vector is also influenced Therefore the Landau phenomenological theory requires a by the character of the collinear magnetic ordering ͑two- or certain modification due to the fact that the value of the four-sublattice͒. The substantial dependence of the structural relative shifts of the crystal ͑basal͒ planes is not small. In this parameter on the character and form of the magnetic struc- paper we have attempted to go beyond the standard Landau ture is what makes it so that the phase transitions between theory, which relies on the smallness of the order parameter the ␤ and ␣ phases and between the ␣ and ␦ phases can and, in particular, the linear theory of elasticity, and to de- occur only as first-order transitions, accompanied by a jum- scribe the properties of the antiferromagnetic phases of solid plike change of the value of the shift vector13͒ and also of the oxygen arising under the influence of high pressures. specific volume and other observables, e.g., the lattice pa- We are sincerely grateful to Yu. A. Freiman for acquaint- 17 rameters, the energies of the different elementary excitations, ing us with the content of his review article prior to its etc. publication; that was the stimulus for our studies. Freiman Since the ‘‘driving force’’ of the structural transforma- made a number of concrete critical comments that helped us tions considered is necessarily a change of the magnetic or- state the results presented above in a more precise way. This dering, then we assume that an external magnetic field, by work was done partially as part of a plan of scientific re- changing the mutual orientation of the spins, can ͑and search of the National Technical University of Ukraine ͑ should͒ also lead to a change of the crystal structure. In ‘‘Kiev Polytechnical Institute’’ State Registration No. ͒ ͑ ͒ particular, the proposed model predicts a change of the mag- N105U001280 . One of us E.V.G. thanks A. A. Maly- nitude of the shift vector and a displacement toward the ␦ shenko for financial and technical support during the perfor- phase on the phase diagram upon application of magnetic mance of this study. field ͑and pressure͒ to a crystal found in the ␣ phase. One * also expects a noticeable change of the specific volume in a E-mail: [email protected] †E-mail: [email protected] magnetic field, and the character of this dependence should 1͒We note that this requirement differs from the condition of translational be quadratic. It is not ruled out that the application of mag- invariance of the crystal as a whole, since we are talking about the relative netic field can lead to displacement of the temperature of the shift of only two planes, while the position of the others remains un- ␣ ␤ changed. - transition, but the lack of available experimental data 2͒The idea of choosing a periodic function of the shift vector as the order does not permit an unambiguous determination of whether parameter was put forth in Refs. 26–28. ͒ this displacement is to higher or lower temperatures. 3 Nonuniform shifts of this type usually give rise to the so-called polymor- 29 As we have demonstrated, the proposed approach allows phic structures, the best-known example of which is the case of the fcc and hcp lattices, with layer alternations of ABCABC... and ABAB... . ͑ ͒ one on a qualitative and, under certain conditions semi- 4 In the exchange approximation, with accuracy up to rotations of the mag- ͒quantitative level to describe and systematize the rather netic moments relative to the crystallographic axes. ͒ large amount of experimental material that has been accumu- 5 This is meant in the sense that the appearance of the order parameter lated to date concerning the properties of solid oxygen. How- responsible for the transition entails the appearance of parameters propor- tional to it which transform according to the same irreducible representa- ever, the domain of applicability of the model is restricted, at tion. minimum, by the assumptions that a number of parameters 6͒The shear modulus is defined by the value of the second derivative of the ⌽ remain unchanged or change little, and also by the fact that potential str with respect to the vector u, taken at the equilibrium value of the latter. the values of the phenomenological coefficients used are ͒ 7 Such a dependence of the phenomenological coefficients in the case of only estimates and not rigorous quantitative values, and, pos- solid oxygen can be due to both anharmonicity of the intermolecular in- sibly importantly, we have neglected effects due to the teraction potential and to dependence on the orientation and degree of change of the mutual orientation of the oxygen molecules. overlap of the electron shells of the oxygen molecules. The latter circum- stance can turn out to be important at high pressures. Another important circumstance that can limit the applicabil- 8͒This degeneracy is lifted if the interaction of planes farther apart than ity of the phenomenological model is its independence of the nearest neighbors is taken into account. dimensionality of the crystal, since the model is essentially 9͒The value of ␰ was calculated from the experimental data as follows: ␰ ϭ ␤ ␤ based on the self-consistent field theory. At the same time, it c•cos /a, where the quantities a,b,c, specify the monoclinic lattice of the ␣ phase. is well known that at least the magnetic characteristics of ͒ 10 We note that in solid oxygen the specific volume changes very strongly; solid O2 are close to two-dimensional, and that requires a for example, at Tϭ19 K and Pϭ0 one has vӍ20 cm3/mole, while at certain care, especially for the description of phase transi- Pϭ7 GPa, vӍ14 cm3/mole. 11͒ ⌳(␣) tions ͑see Ref. 17͒. The satisfactory agreement obtained The values of the coefficients Keff , K4 , ʈ were determined by approxi- above between the calculated and observed parameters asso- mating the experimental data19 by the theoretical dependence ͑17͒, while ␹ ␭ ciated with the volume change for practically all the low- T and v were obtained by approximating the dependence of the specific volume on the external pressure according to formulas ͑20͒ and ͑25͒. temperature phases of solid O2 is apparently due to the phe- 12͒We note that a further shift of the basal planes in the same direction would nomenological allowance for the temperature dependence of lead to a lowering of the symmetry and to a ‘‘retro’’ ͑from a symmetry the coefficients of volume thermal expansion, which most standpoint͒ transition to the ␣ phase. 13͒ likely implicitly include the contribution of correlation ef- If it were possible to ‘‘turn off’’ the magnetic interactions that lead to antiferromagnetic order in the solid oxygen crystal ͑e.g., on account of fects. spin-flip of the magnetic sublattices in an external field͒, then the phase On the whole, one can say that the phenomenological transitions mentioned could in principle also occur as second-order tran- 776 Low Temp. Phys. 31 (8–9), August–September 2005 E. V. Gomonay and V. M. Loktev

sitions, with a smooth change of the shift vector. 16 I. M. Vitebskii and N. M. Lavrinenko, Fiz. Nizk. Temp. 19,535͑1993͒ ͓Low Temp. Phys. 19,380͑1993͔͒. 17 ͑ ͒ 1 O. N. Trapeznikova and L. W. Shubnikow, Nature ͑London͒ 3384, 378 Y. A. Freiman and H. J. Jodl, Phys. Rep. 401,1 2004 . 18 ͑1934͒. R. D. Etters, K. Kobashi, and J. Belak, Phys. Rev. B 32, 4097 ͑1985͒. 19 2 L. D. Landau, Collected Works ͓in Russian͔, Vol. 1, Nauka, Moscow Y. Akahama, H. Kawamura, and O. Shimomura, Phys. Rev. B 64, 054105 ͑1969͒. ͑2001͒. 3 ͑ ͒ 20 S. V. Vonsovski, Magnetism, Vols. 1 and 2, Wiley, New York 1974 , Yu. A. Izyumov and V. N. Syromyatnikov, Phase Transitions and Symme- Nauka, Moscow ͑1971͒. try of Crystals ͓in Russian͔, Nauka, Moscow ͑1984͒. 4 B. I. Verkin and A. F. Prikhot’ko ͑eds.͒, Cryocrystals ͓in Russian͔, 21 I. N. Goncharenko and O. L. Makarova, Phys. Rev. Lett. 93, 055502 Naukova Dumka, Kiev ͑1983͒. ͑2004͒. 5 C. Fagerstroem and A. Hollis, J. Low Temp. Phys. 1,3͑1969͒. 22 I. M. Vitebskii, A. N. Knigavko, and A. A. Chabanov, Fiz. Nizk. Temp. 19, 6 I. N. Krupski, A. I. Prokhvatilov, Yu. A. Freman, and A. I. Erenburg, Fiz. 542 ͑1993͓͒Low Temp. Phys. 19, 385 ͑1993͔͒. ͑ ͓͒ ͑ ͔͒ Nizk. Temp. 5,271 1979 Sov. J. Low Temp. Phys. 5,130 1979 . 23 O. V. Kovalev, Irreducible and Induced Representations and Co- 7 B. Verkin, V. Manzhelii, and V. Grigor’ev, Handbook of Properties of Representations of Fedorov Groups ͓in Russian͔, Nauka, Moscow ͑1986͒. Condensed Phases of Hydrogen and Oxygen, Hemisphere Publishing Cor- 24 G. S. DeFotis, Phys. Rev. B 23, 4714 ͑1981͒. ͑ ͒ poration, New York 1991 . 25 A. I. Prokhvatilov, N. N. Gal’tsov, and A. V. Raenko, Fiz. Nizk. Temp., 27, 8 A. Jezowski, P. Stachowiak, V. V. Sumarokov, J. Mucha, and Yu. A. Frei- 523 ͑2001͓͒Low Temp. Phys. 27, 391 ͑2001͔͒. man, Phys. Rev. Lett. 71,97͑1993͒. 26 J. C. Tole´dano and P. Tole´dano, The Landau Theory of Phase Transitions, 9 A. S. Borovik-Romanov, Zh. E´ ksp. Teor. Fiz. 21, 1303 ͑1951͔͒. World Scientific, Singapore ͑1987͒. 10 A. S. Borovik-Romanov, M. P. Orlova, and P. G. Strelkov, DAN SSSR 99, 27 ´ 699 ͑1954͒. V. P. Dmitriev, S. B. Rochal, Y. M. Gufan, and P. Toledano, Phys. Rev. ͑ ͒ 11 V. M. Loktev, Fiz. Nizk. Temp. 5, 295 ͑1979͓͒Sov. J. Low Temp. Phys. 5, Lett. 60, 1958 1988 . 28 142 ͑1979͔͒. P. Tole´dano, G. Krexner, M. Prem, H.-P. Weber, and V. P. Dmitriev, Phys. 12 Yu. B. Gaididei and V. M. Loktev, Fiz. Nizk. Temp. 7, 1305 ͑1981͓͒Sov. Rev. B 64, 144104 ͑2001͒. 29 J. Low Temp. Phys. 7, 634 ͑1981͔͒. B. I. Nikolin, Multilayer Structures and Polytypism in Metallic Alloys ͓in 13 ͔ ͑ ͒ A. Z. Patashinski and V. L. Pokrovski, Fluctuation Theory of Phase Russian , Naukova Dumka, Kiev 1984 . Transitions, transl. of 1st Russ. ed., Pergamon Press, Oxford ͑1979͒, cited 30 L. D. Landau and E. M. Lifshitz, Theory of Elasticity, 3rd English ed., 2nd Russ. ed., Nauka, Moscow ͑1983͒. Pergamon Press, Oxford ͑1986͒, 4th Russ. ed., Nauka, Moscow ͑1987͒, 14 I. M. Vitebskii, V. L. Sobolev, A. A. Chabanov, and V. M. Loktev, Fiz. p. 28. Nizk. Temp. 18, 1044 ͑1992͓͒Low Temp. Phys. 18, 733 ͑1992͔͒. 31 A. P. J. Jansen and A. van der Avoird, J. Chem. Phys. 86, 3597 ͑1987͒. 15 I. M. Vitebskii, V. L. Sobolev, A. A. Chabanov, and V. M. Loktev, Fiz. Nizk. Temp. 19,151͑1993͓͒Low Temp. Phys. 19, 107 ͑1993͔͒. Translated by Steve Torstveit LOW TEMPERATURE PHYSICS VOLUME 31, NUMBERS 8–9 AUGUST–SEPTEMBER 2005

On the discovery of magnon sidebands in insulating antiferromagnets R. M. White*

Carnegie Mellon University, Pittsburgh, Pennsylvania 15213, USA W. M. Yen†

Department of Physics and Astronomy, University of Georgia, Athens, Georgia 30602-2451, USA ͑Submitted February 4, 2005͒ Fiz. Nizk. Temp. 31, 1020–1023 ͑August–September 2005͒ In this article, we reconstruct the course of events which led to the discovery and identification of magnon sidebands in the optical spectrum of simple antiferromagnetic insulators. © 2005 American Institute of Physics. ͓DOI: 10.1063/1.2008138͔

INTRODUCTION Schawlow had arrived earlier in the year at Stanford from Bell Labs and was assigned temporary lab space in the In a recent article in which he summarizes his remark- Hansen Microwave Lab and began establishing those pro- able career and his many seminal contributions to our under- grams which were later to be recognized with a Nobel Prize. standing of optical properties of the condensed phases,1 Don The Microwave Lab was housed in a ‘‘temporary’’ building McClure writes: dating from the Second World War and the lab was ex- ...But I was an unreconstructed academic and re- tremely crowded; the forced contact in these close quarters turned to academic life after receiving an offer from the led to considerable socialization between members of the six University of Chicago in 1962. There, among other or seven different research groups utilizing Hansen and re- things, we built a pulsed magnet capable of producing sulted in some beneficial interactions, as exemplified by our fields of 20 T and saw the Zeeman effect in the triplet discovery of magnon sidebands. states of organic molecules. Schawlow’s interests at the time centered on the optical More interesting though was the effect of these high properties of various dopants in insulating solids which fields on the absorption spectrum of MnF2 and similar could be used for solid state lasers. Yen and other members compounds whose thermodynamic of the group were given the task of understanding the ther- properties Willard Stout had been studying. We found mal dependence of the widths and positions of the sharp line that some of the spectra lines were split in the field as spectra of transition metals and rare earths in solids. These expected, but that others were not—a surprise which studies resulted in a number of publications in which the could be explained in terms of the antiferromagnetic thermal broadening and shifts of the spectra were identified structure of the crystal. While we were learning that as arising from phonon-ion interactions. Of significance to there were such things as magnons as well as this article, in Imbusch et al.,3 the isotope shifts of the Cr3ϩ in these crystals, a group of students of Art Schawlow R-lines in Al2O3 and MgO were related to changes in the at Stanford published a paper on the spin-wave side- phonon frequencies induced by the differences in the masses ͑ ͒ bands magnon side bands of MnF2 . We were in sec- of the hosts and the reflection of these changes in the phonon ond place again but our paper nicely confirmed their sideband structure. In addition, in Yen et al.,4 observation of interpretation. We never learned why they picked up phonon sideband structure was first reported for a rare earth MnF just when we did... . 3 3 3ϩ 2 based material ( H4 to P0 absorption of LaF3 :Pr ). The paper McClure alludes to concerned the first obser- White’s thesis involved extending Suhl’s theory of spin- vation of a magnon-induced sideband in a magnetically or- wave instabilities in ferrites to antiferromagnets. Although dered material.2 The identification of these sidebands by a the magnetic properties of antiferromagnets had been de- pair of young students and two post-docs illustrates how ser- scribed by Ne´el in 1932, at that time they were more of a endipity, a touch of good luck and reckless youthful enthu- curiosity than a material with technological applications as siasm often plays a role in scientific discovery. In this article, were the ferrites. This is not the case today where antiferro- we describe the events that led to these developments in the magnets play a critical role in providing the giant magnetore- hope that they will answer McClure’s curiosity on this sub- sistance ͑GMR͒ so important in magnetic recording heads. ject. White’s thesis also explored the possibility of phonon insta- bilities which developed his appreciation for analogies be- THE STANFORD HANSEN MICROWAVE LABORATORY tween magnons and . The authors became acquainted with each other at Stan- THE VARIAN LABS AND THE OASIS ford in the Fall 1962; at that time, one of us ͑RMW͒ was in midst of his PhD thesis work in Marshall Sparks’ theory The new Physics Building at Stanford, the Varian Build- group while the other ͑WMY͒ had just joined Art ing, was finished in 1963 and Schawlow’s spectroscopy Schawlow’s group in spectroscopy as a Research Associate. group moved over from Hansen. At that time, it was custom-

1063-777X/2005/31(8–9)/3/$26.00777 © 2005 American Institute of Physics 778 Low Temp. Phys. 31 (8–9), August–September 2005 R. M. White and W. M. Yen ary for a group of us Physics graduate students and post-docs 1964 was an exceptional active year for both the authors. to stop for a night cap at a local watering spot ͑‘‘The Oasis’’ White had obtained an NSF postdoctoral fellowship and or affectionately ‘‘The O’’͒ following our usual late evenings had joined C. Kittel’s group at UC Berkeley. Again guided at the lab; there normally would be a whole table reserved to by analogy with phonon sidebands he developed a theory for accommodate us and both authors were frequent attendees of the shape of the magnon sideband. This characteristic asym- these gatherings. The topic of conversation at them was wide metric line shape became critical in the subsequent identifi- ranging, even including physics problems at times! On one cation of the sideband in experimental data. of these gatherings, about the time the two publications men- Yen, on the other hand, had begun a program to investi- tioned above were submitted, a discussion on phonon- gate optical energy transfer processes in rare earth activated induced sidebands arose during which White first raised the systems and was in the process of interviewing for a perma- possibility of observing a magnon-induced sideband; our dis- nent position; he was to accept an offer from Wisconsin- course then shifted to the practicality of observing such a Madison in the Spring of 1965. We were able to revisit the feature in magnetically ordered materials and on the appro- MnF2 only occasionally during this period, but fortunately priate materials which would allow the observation of these both Darrell Sell and Rick Greene began to show an in- assisted transitions. The first requirement, of course, was that creased interest in this problem as 1964 progressed and it the material possess good optical properties and secondly eventually became the subject of their PhD theses. that its crystalline and magnetic properties be well character- ized. At that time there were no obvious ferromagnets with PHYSICS OF QUANTUM ELECTRONICS CONFERENCE AND the desired optical properties. THE DISCOVERY Eventually, White settled on the antiferromagnet, MnF , 2 These activities produced a small problem for Art which was being investigated at Bell Labs from the reso- Schawlow, since none of his grants provided coverage for nance point of view and whose magnetic properties had been these investigations. He discussed these difficulties with Yen established. In fact, Vince Jaccarino once refered to MnF2 as just prior to their departure for Puerto Rico to attend the the ‘‘fruit fly’’ of magnetics research. It is reasonably trans- Physics of Quantum Electronics Conference ͑PQEC, June parent and has a Ne´el temperature that is not too low. In ͒ 2ϩ 28–30, 1965 ; it was agreed during that discussion that the addition, the room temperature spectra of the Mn in MnF2 5 project would be transferred to Wisconsin when Yen as- had been determined by Stout. sumed his faculty position there in the Fall of 1965 but that The ‘‘luck’’ we referred to above also had to do with we would continue our investigations at Stanford until our selecting MnF2 . MnF2 has the rutile crystal structure, which return from the PQEC meeting. Schawlow also was aware does not have a center of symmetry between Mn pairs. It is that R.E. Dietz, L.F. Johnson and others at Bell Labs had these pairs that are excited in the sideband. This lack of been investigating the properties of magnetically ordered symmetry enables a relatively strong electric dipole transi- materials and suggested that Yen discuss our ideas with them tion, without which this effect might not have been observ- to resolve any conflicts if any existed. Yen’s paper on line- 6 able. This was subsequently pointed out by Tanabe et al. width studies of energy transfer mechanisms8 was scheduled White actually purchased a sample of MnF2 using his in the same session as that of Dietz et al. which was entitled advisor’s theory funding to launch this exploration. Crystals ‘‘Fluorescence from Magnetic Crystals’’ (NiF2 , CoF2 , and boules of MnF were also listed as available in a cata- 9 2 MnF2). During the question period, Bob Dietz was specifi- logue from SemiElements, a small company in Pennsylvania. cally asked by Yen whether he had observed any effects re- We purchased a whole boule from this company; though we latable to the collective magnetic excitations of the simple later obtained samples from Howard Guggenheim ͑Bell fluorides he investigated; before he could answer, Willi Lowe Labs͒ and from Bob Feigelson ͑Stanford͒; curiously, the of the Hebrew University responded for him and empatheti- samples derived from this boule proved to be the best cally stated that ‘‘magnetic interactions were notoriously samples we ever investigated. We never learned the origin of weak and there was absolutely no chance they would couple the boule as SemiElements went out of business soon after. to electronic transitions.’’ It is likely that at that point, Samples in different orientations, ␴, ␲, and ␣ were cut Schawlow wrote off our effort as a lost cause. from the boule, and initial surveys of the absorption spectra As noted earlier, however, we had strong and tantalizing were done at low temperature using a photographic B&L hints of sharp structure accompanying the lower 4T transi- 2ϩ spectrograph. Initial results were first reported at the 1963 tions of Mn in MnF2 ; before his departure for PQEC, Yen, APS Winter Meeting in Berkeley;7 in that paper, we first Sell and Greene mapped out a schedule of experimentation reported sharp structure in the vicinity of the first two excited during his absence which included He ␭-point temperature 2ϩ 4 4 states of Mn ( T1g and T2g) and very strong absorptions runs on a high resolution Jarrell–Ash scanning spectrograph 4 4 4 in the A, E region, but no clearly identifiable magnon- with emphasis on the lowest T1g state of MnF2 . Yen also assisted transitions. The paper attracted very little attention, scheduled himself for a week of vacation in St. Thomas fol- but the preliminary results were encouraging enough to war- lowing the PQEC meeting to extend the time before our rant continued interest on our part. Our activities attracted efforts at Stanford were to be shut down. As it turns out, the the attention of three members of the Schawlow group, ini- additional week was not needed; Sell’s lab book entry for the tially Warner Scott and then Darrell Sell and Rick Greene; observation of what would later be definitely identified as we began an earnest effort to obtain high resolution, low- magnon sidebands, shown in Fig. 1, was June 29, 1965, temperature spectra of MnF2 using a scanning spectrometer likely the very day of the session alluded to above! When and electronic detection in the winter of 1964. these traces were shown to Schawlow in early July, he im- Low Temp. Phys. 31 (8–9), August–September 2005 R. M. White and W. M. Yen 779

two-magnon excitation in the infrared was unknowingly be- 19 ing observed in FeF2 by I. Silvera at Berkeley and later in 20 CoF2 at Bell Labs. In fact, the Berkeley IR absorption data with its anamolous peak at 154 cmϪ1 was presented in the very same session of the 1963 APS meeting in Berkeley at which we presented our first optical spectra of MnF2 . And, as indicated by McClure’s quote at the beginning of this article, his group was also studying the optical spectra of MnF2 and had been puzzled by the observation of transitions which did not split in a magnetic field.21 Now that we under- stand this process we know that all these phenomenon are related, and their existence provides a very useful technique for studying magnon dynamics22 as well as energy transfer in solids.23

*E-mail: [email protected] †E-mail: [email protected]

1 D. S. McClure, J. Lumin. 100,47͑2002͒. 2 Ϫ1 R. L. Greene, D. D. Sell, W. M. Yen, A. L. Schawlow, and R. M. White, FIG. 1. Absorption spectrum of MnF2 in the range of 18.550 cm .The 2ϩ Phys. Rev. Lett. 15, 656 ͑1965͒. absorptions E1 and E2 are pure electronic transitions of Mn . The absorp- 3 tions ␴1, ␴2, and ␲1 are the first identified magnon-assisted transitions in a G. F. Imbusch, W. M. Yen, A. L. Schawlow, G. E. Devlin, and J. P. ͑ ͒ magnetic material. Remeika, Phys. Rev. 136,A481 1964 . 4 W. M. Yen, W. C. Scott, and A. L. Schawlow, Phys. Rev. 136, A 271 ͑1964͒. 5 J. W. Stout, J. Chem. Phys. 31, 709 ͑1959͒. mediately understood the importance of the observation, and 6 Y. Tanabe, T. Moriya, and S. Sugano, Phys. Rev. Lett. 15,1023͑1965͒. shutting down the effort was, of course, never considered 7 W. M. Yen, W. C. Scott, and R. M. White, Bull. Am. Phys. Soc. II-9,715 again. Schawlow’s enthusiasm was also responsible for ͑1963͒. 8 White returning to Stanford as an Assistant Professor in the W. M. Yen, R. L. Greene, and W. C. Scott, ‘‘Optical linewidth studies of energy transfer mechanisms between impurity ions,’’ Proceedings of the fall of 1965. Physics of Quantum Electronics, P. L. Kelley, B. Lax, and P. E. Tannen- wald ͑eds.͒, McGraw-Hill, New York ͑1966͒,p.332. 9 CONCLUSIONS R. E. Dietz, L. F. Johnson, and H. J. Guggenheim, ‘‘Fluorescence of magnetic crystals,’’ Proceedings of the Physics of Quantum Electronics,P. A complete description of the magnon sideband phe- L. Kelley, B. Lax, and P. E. Tannenwald ͑eds.͒, McGraw-Hill, New York ͑ ͒ nomenon was published in 1967.10 The fact that the peak of 1966 , p. 361. 10 D. Sell, R. Greene, and R. M. White, Phys. Rev. 158, 489 ͑1967͒. the sideband is directly related to the frequency of magnons 11 R. L. Greene, D. D. Sell, R. S. Feigelson, G. F. Imbusch, and H. J. at the Brillouin zone boundary enables one to study these Guggenheim, Phys. Rev. 171, 600 ͑1968͒. modes quite easily, which previously could only be done by 12 B. A. Wilson, W. M. Yen, J. Hegarty, and G. F. Imbusch, Phys. Rev. B 19, ͑ ͒ neutron scattering. At least half a dozen theses on various 4238 1978 . 13 J. F. Holzrichter, M. L. Report No. 1912, Stanford University, November aspects of the spectra of magnetically ordered materials from ͑1970͒. Stanford and from Wisconsin followed in the following de- 14 M. W. Passow, D. L. Huber, and W. M. Yen, Phys. Rev. Lett. 23, 477 ͑1969͒. cades; this work contributed to a comprehensive understand- 15 ing of the static and dynamic optical properties of ordered Y. H. Wong, F. L. Scarpace, C. D. Pfeifer, and W. M. Yen, Phys. Rev. B 9, 3086 ͑1974͒. magnetic systems. These studies included studies on the lu- 16 B. A. Wilson, J. Hegarty, and W. M. Yen, Phys. Rev. Lett. 41,268͑1978͒. 11,12 17 minescence properties of pure and doped MnF2 , the ob- D. J. Lockwood, R. M. White, and W. M. Yen, Phys. Rev. B 69, 174413 13 ͑2004͒. servation of induced photomagnetism, of magnon-induced 18 14 R. M. White, J. Appl. Phys. 36, 3653 ͑1965͒. broadening of optical transitions, of circular and magnetic 19 I. Silvera and M. Tinkham, Bull. Am. Phys. Soc. II-9,714͑1963͒. 15 circular dichroism and the discovery of biexciton annihila- 20 P. L. Richards, Bull. Am. Phys. Soc. II-10,33͑1965͒. tion in this material.16 And recently, we have been able to 21 P. G. Russell, D. S. McClure, and J. W. Stout, Phys. Rev. Lett. 16,176 identify sublattice splittings in MnF induced by dipolar in- ͑1966͒. 2 22 ͑ ͒ ͑ ͒ 17,18 L. D. Rotter, W. M. Dennis, and W. M. Yen, Phys. Rev. B 42,720 1990 . teractions predicted earlier by one of us RMW . 23 G. P. Morgan and W. M. Yen, ‘‘Optical energy transfer in insulators,’’ in It is interesting how often new concepts in science ap- Laser Spectroscopy of Solids,W.M.Yen͑ed.͒, Vol. 65 of Springer Topics pear almost simultaneously in different contexts. While we in Applied Physics, Springer-Verlag, Berlin ͑1989͒,p.77. were looking for what we now call an -magnon exci- This article was published in English in the original Russian journal. Repro- tation in MnF2 at Stanford, the very similar phenomenon of duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 31, NUMBERS 8–9 AUGUST–SEPTEMBER 2005

␣ Direct optical excitation of two and three magnons in -Fe2O3 Y. Tanabe

Department of Applied Physics, University of Tokyo, 7-3-1 Hongo, Tokyo 113-8656, Japan Y. Fujimaki, K. Kojima, and S. Uchida

Department of Physics, University of Tokyo, 7-3-1 Hongo, Tokyo 113-8656, Japan S. Onari and T. Matsuo

Institute of Materials Science, University of Tsukuba, Tsukuba, Ibaraki 304-8573, Japan S. Azuma and E. Hanamura*

Chitose Institute of Science and Technology and Japan Science and Technology Agency Chitose, Hokkaido 066-8655, Japan ͑Submitted December 20, 2004͒ Fiz. Nizk. Temp. 31, 1024–1031 ͑August–September 2005͒ Direct excitation of two and three magnons is observed in mid-infrared absorption and Raman ␣ scattering spectra of -Fe2O3 crystals. These polarization characteristics and the spectra themselves are shown to be understood from group-theoretical point of view. The microscopic mechanism of three-magnon excitation is proposed in addition to that of well-known two- magnon excitation process. © 2005 American Institute of Physics. ͓DOI: 10.1063/1.2008139͔

I. INTRODUCTION ϭ 24 perature TM 261 K. Below TM , the four sublattice mag- netizations are parallel to the threefold c axis of the corun- Transition-metal oxide with open d-electron shells have 6 ¯ shown many interesting magnetic and transport phenomena dum structure with the space group D3d(3 2(/m)), and this of strongly correlated d-electrons.1 Optical absorption due to phase shows pure antiferromagnetism. Between TM and the ϭ phonon-assisted two-magnon excitation was observed for Ne´el temperature TN 960 K, the four sublattice magnetiza- tions are almost in the ab plane, and the weak ferromag- transition-metal oxides of quasi-two-dimensional La2CuO4 , 2 3 4 netism is induced in the direction of the twofold a axis. Here Nd2CuO4 ,Sr2CuO2Cl2 , La2NiO4 , YBa2Cu3O6 , and 5 the space group is monoclinic C (2/m). two-leg-ladder (Ca,La)14Cu24O41 crystals. On the other 2h hand, phonon-assisted particle-hole excitation of the spinons In the present paper, we shall supplement the previous 6 report on the mid-infrared absorption due to magnons8 by was confirmed in quasi-one-dimensional Sr2CuO3 crystal. Here the oscillator strength of the optical phonon makes the supplying the observed polarization dependence and tem- optical excitation of two magnons possible through the perature dependence of the Raman spectrum. Then we can magnon-phonon interaction.7 confirm the assignment of the two-magnon excitation at the ͑ We have observed direct mid-infrared absorption by D point at the Brillouin zone edge to the effective ex- ␣ 8 change͒ dipole moment ␲S S . This will be given in Sec. 2. two- and three-magnon excitation in -Fe2O3 crystal. This i• j is due to the effective electric dipole moment induced by the The three-magnon excitation was assigned to originate from ␲DM ϫ 10 combined effect of the electronic transition of the d-electron the effective dipole moment, like in •(Si Sj). We and the exchange interaction between the neighboring Fe3ϩ will give all the terms that contribute to the three-magnon ions.9,10 The magnon energy as well as the absorption coef- excitation in Sec. 3 and discuss its microscopic mechanism ␣ in Sec. 4. Last, Sec. 5 is for conclusion and discussion. ficient of -Fe2O3 are much larger than those of MnF2 with ϭ Ne´el temperature TN 67.3 K and absorption coefficienct of a few cmϪ1 ͑Refs. 11, 12͒. ␣ II. MID-INFRARED AND RAMAN SCATTERING SPECTRUM Magnetic properties of -Fe2O3 have been well studied,13 and the origin of weak ferromagnetism was pro- The samples used were prepared from single crystals of posed from the studies of this crystal.14,15 The magnon dis- good-quality natural hematite, which were also used in Ref. persion has been obtained from inelastic neutron scattering 8. These crystals have been sliced and polished by using investigation16 and also from its theoretical analysis.17 On diamond powder and finally by aluminate powder with a ␣ ␮ the other hand, the phonon structure of -Fe2O3 is also ob- diameter 0.1 m. The samples used have thickness ranging tained by the infrared reflection spectrum,18 Raman from 50 ␮mto140␮m. scattering19 and lattice dynamics study.20 Two-magnon Ra- The measurement of the mid-infrared absorption spec- man scattering in this crystal had been a controversial point trum was performed with a Bruker IFS 113 V Fourier- but this was also clarified by the effects of pressure and transform spectrometer in the energy range 350 to isotope substitution on the Raman spectrum.21–23 Magnetic 2400 cmϪ1. We observed several prominent features on the structure of this crystal was found to change at Morin tem- high-energy side of the single phonon modes,18 as shown in

1063-777X/2005/31(8–9)/6/$26.00780 © 2005 American Institute of Physics Low Temp. Phys. 31 (8–9), August–September 2005 Tanabe et al. 781

␣ ͑ ͒ ͑ ͒ FIG. 3. Raman spectrum of -Fe2O3 at room temperature a and77K b with incident beam 633 nm from He–Ne laser. A, B and C describe Lorent- zian fittings with ͑central wave number, width at half maximum, relative intensity͒ϭ(1320 cmϪ1,60 cmϪ1,60) for A, (1460 cmϪ1,80cmϪ1,9͒ for FIG. 1. The polarization dependence of mid-infrared absorption spectrum of Ϫ Ϫ Ϫ Ϫ B, and (1540 cm 1,70cm 1,18͒ for C in ͑a͒, and (1327 cm 1,45cm 1, ␣-Fe O crystal with a crystal thickness 120 ␮m at 30 K and 300 K ͑a͒.The Ϫ Ϫ Ϫ Ϫ 2 3 110͒ for A, (1500 cm 1,90cm 1,38͒ for B, and (1580 cm 1,90cm 1, incident beam is polarized in the ab plane and in the c direction ͑b͒. 45͒ for C in ͑b͒. The thick solid line describes superposition of Lorentzian curves A, B, and C. Figs. 1 and 2. The first important characteristics is that the absorption peak just above 1500 cmϪ1 shows strong polar- ization dependence: no signal is observed for Eʈc axis while 2200 cmϪ1. From these polarization and temperature depen- the strong signal for EЌc axis. On the other hand, the signal dence, we may speculate that the signals at 1500 cmϪ1 and Ϫ1 around 2200 cm is observed under both polarizations. Sec- 2200 cmϪ1 come from the elementary excitations different ond, both of these peaks show blue shift with decreasing from others. Ϫ1 lattice temperatures, e.g., the absorption peak of 1500 cm As a matter of fact, on the basis of the assignment of Ϫ1 at room temperature shifts to 1570 cm at 30 K. Third, single phonon modes18 observed by the infrared reflection three absorption peaks are observed on the high energy side spectrum and the magnon dispersion obtained from neutron 18 Ϫ1 Ϫ1 of the single phonon modes, i.e., 1160 cm , 1050 cm , inelastic scattering, we can assign the 1160 cmϪ1 signal to a Ϫ and 920 cm 1. These energies are almost independent of the Ϫ1 ϩ ͑ Ј combination of TO4 (524 cm ) LO4 or LO2) Ϫ1 Ϫ1 lattice temperature in contrast to those at 1500-1 and (662 cm ), the 1050 cm signal to that of TO4 Ϫ1 ϩ Ј Ϫ1 Ϫ1 (524 cm ) TO2 (526 cm ), and the 920 cm signal to Ϫ1 ϩ Ј Ϫ1 that of TO3 (437 cm ) TO2 (494 cm ). Each of the transverse and longitudinal Eu modes is numbered from the Ј Ј lower to higher energy side, and LO2 and TO2 correspond to Ϫ1 Ϫ1 the A2u modes. The energies of 1500 cm and 2200 cm are nearly equal to two and three magnons, respectively, ob- tained from the neutron data.17 We are interested in the wave number region around 800 cmϪ1 to check whether a single magnon is observed in the absorption spectrum. The trans- mission spectrum is observed as shown in the inset of Fig. 2 by reducing the sample thickness from 120 ␮m to 100 and 80 ␮m. In order to confirm these assignments, Raman spectrum has been observed with Stokes shift in this energy range. This gives information to supplement the mid-infrared ab- FIG. 2. The temperature dependence of the transmission spectrum for the sorption spectrum. Figures 3a, b give the Raman spectrum polarization E in the ab plane for the same sample thickness 120 ␮masin observed by using He–Ne laser as incident beam at room Fig. 1. The inset shows the sample-thickness dependence around 800 cmϪ1 at 30 K. A single E magnon is expected at 782 cmϪ1 while combination temperature and nitrogen temperature, respectively. A dou- u Ϫ1 modes of two phonons are at 780 cmϪ1 (TO ϩLO ), 805 cmϪ1 (TO blet structure is observed with Stokes shift 1460 cm and 2 3 3 Ϫ ϩ Ϫ1 ϩ Ј 1 LO2) and 851 cm (TO3 LO1). 1540 cm at room temperature, as shown in Fig. 3a, and 782 Low Temp. Phys. 31 (8–9), August–September 2005 Tanabe et al. showing a blue shift similar to that in the mid-infrared ab- TABLE I. Representations and excitation energies of magnons at high- sorption spectrum at nitrogen temperature, as shown in Fig. symmetry points. 3b. The stronger signal at 1320 cmϪ1 is almost independent of the lattice temperature and is well assigned to overtone of ϫ Ϫ1 two LO4 phonons (2 662 cm ). These two-phonon exci- tations at 1160, 1050 and 920 cmϪ1 and at 1320 cmϪ1 obey Loudon’s selection rule of combination and overtone modes, respectively, for the infrared and Raman scattering process.25 The doublet structure around 1500 cmϪ1 is observed under the configuration b(a,c)¯b but not for b(a,a)¯b nor for b(c,c)¯b, while the signal at 1320 cmϪ1 is allowed for both configurations. This means that the elementary excitation with doublet structure around 1500 cmϪ1 has the symmetry Eg . The second task of the present paper is to understand two- and three-magnon excitations both in the mid-infrared and Raman spectrum from both group-theoretical and micro- down-sublattices. In the antisymmetric combination denoted scopic points of view. These will be discussed in the follow- by ͕͖, two contributions cancel out each other. ing sections. ͑ ͒ a Z point with aconstic Z2 and optical Z3 magnons: ͓ 2͔ϭ͓ 2͔ϭ ϩ͓͓ ͔͔ ͕ 2͖ϭ͕ 2͖ϭ͓͓ ͔͔ ͑ ͒ III. SELECTION RULES Z2 Z3 A1u Eg , Z2 Z3 A2u , 1 The mid-infrared absorption peak due to two-magnon ϫ ϭ ϩ ϩ ͑ ͒ Z2 Z3 A1g A2g ͓͓Eu͔͔; 2 excitation is observed more strongly than a single magnon ͑ ͒ ϩ ϩ absorption. This is expected simply from two reasons: ͑i͒ the b D point with acoustic D1 D3 and optical D2 D4 mag- joint density of states for a pair of excitation k and Ϫk, nons: which diverges at the high-symmetry points on the Brillouin 2ϩ 2ϭ 2ϩ 2ϭ ϩ ϩ͓͓ ϩ ͔͔ ͑ ͒ D1 D3 D2 D4 A1g Eg A1g Eg , 3 zone ͑BZ͒ edges, becomes much larger than that of a single ϫ ϭ ϫ ϭ ͓͓ ϩ ͔͔ ͑ ͒ elementary excitation near the ⌫ point; ͑ii͒ and the two- 2D1 D3 2D2 D4 2 A2g Eg , 4 magnon excitation in the antiferromagnet is accompanied by D ϫD ϩD ϫD ϭA ϩE ϩ͓͓A ϩE ͔͔, ͑5͒ rather large electronic dipole moment due to the mechanism 1 2 3 4 1u u 1u u ϫ ϩ ϫ ϭ ϩ ϩ ϩ ͑ ͒ different from a single magnon excitation, as discussed in D1 D4 D2 D3 A2u Eu ͓͓A2u Eu͔͔; 6 ͑ ͒ Ref. 9. The total symmetry of the two excitons magnons at ͑c͒ A point with acoustic A and optical A magnons: the high-symmetry points of the BZ edge may be determined 1 1 26,27 ͓ 2͔ϭ͓͓ ϩ ϩ ͔͔ϩ ϩ ͑ ͒ as given in the literature. A1 A1g A2g 2Eg A1u Eu , 7 The magnetic unit cell of ␣-Fe O consists of four sub- 2 3 2ϭ͓͓ ϩ ͔͔ ͑ ͒ A1 A2u Eu , 8 lattices both below and above TM . The effective dipole mo- ment to excite two magnons is associated with the antiferro- for each mode, and magnetic pair of the neighboring magnetic ions coupled by A ϫA ϭA ϩA ϩ2E ϩ͓͓A ϩA ϩ2E ͔͔ ͑9͒ the superexchange interaction.9 1 1 1g 2g g 1u 2u u Loudon26,27 has shown how to derive the selection rules for the simultaneons excitation of the acoustic and optical for the two-magnon absorption. The procedure is to reduce mode. ⌫ the representation at the point, whose bases are two- For D3d , the states of allowed electric dipole transition magnon states at the high-symmetry point of the BZ bound- are Eu for a- and b-polarized light and A2u for c polariza- ary, and to see whether the decomposition contains the rep- tion, and the allowed states of Raman scattering are Eg and resentation to which the dipole moment operator belongs. If A1g . We have already obtained the eigenvectors at high- it does, the excitation of two magnons at the boundary will symmetry points,10 and we also have the eigenenergies of the be allowed, intensified there by the high density of states. magnon at these points16 as shown in Table I. This conventional procedure can be applied to optical Now we can assign the two-magnon excitation from the excitations of two magnons at high-symmetry points in the experimental results of mid-infrared absorption and Raman Brillouin edge for antiferromagnets with two magnetic sub- scattering as well as from the selection rule together with ␣ lattices such as in MnF2 . However, in the case of -Fe2O3 Table I. First, we can eliminate the contribution from the Z where the magnetic unit cell contains four sublattices, we point to the mid-infrared absorption of two-magnon excita- ͑ ͒ ͑ ͒ must exclude the processes of two-magnon excitation from tion because ͓͓A2u͔͔ in Eq. 1 and ͓͓Eu͔͔ in Eq. 2 come the pair with ferromagnetic coupling. As a result, the selec- from two magnons within the up- or down-spin sublattices. tion rule is modified as shown in Ref. 10. The modified se- Furthermore, Raman scattering with Eg symmetry is absent lection rule for direct optical excitation of two magnons in from the contribution at the Z point. Second, the dipole- ␣ -Fe2O3 is summarized as follows. Symmetry put in a allowed transition with Eu symmetry is possible from the D double bracket on the right-hand side is forbidden, either and A points, as seen from Eqs. ͑5͒, ͑6͒, and ͑7͒, respectively. because the final state is antisymmetric or because it corre- However, the doublet structure of Eg Raman scattering is sponds to the creation of two magnons both within the up- or acceptable for the D point because the splitting between two Low Temp. Phys. 31 (8–9), August–September 2005 Tanabe et al. 783

ϫ acoustic magnons D1 D3 and two optical magnons D2 ing antiferromagnetic ions is accompanied by the electric ϫ ͑ ͒ D4 in Eq. 3 give the doublet structure of Eg symmetry dipole transition within the transition-metal ion or the with a splitting energy of 10 meV, in agreement with the charge-transfer transition between the empty 3d orbital of observed value. The Raman signal in Eq. ͑9͒ comes from one the magnetic ion and the 2p orbitals of the surrounding oxy- optical and one acoustic magnon, and the splitting of the two gen ion.9 We use the expression in Ref. 9 to estimate the Eg modes may be speculated to be smaller than the splitting two-magnon absorption coefficient. The charge transfer ex- 1.4 meV between acoustic and optical magnons. Although citation gives the absorption coefficient of the order of we cannot eliminate the possibility to assign the doublet Ra- 104 cmϪ1 around 2 eV. The superexchange interaction is as- man signal to the two-magnon excitation at the A point, two- sumed to be of the order of the Ne´el temperature 960 K ͑0.1 magnon excitation from the D point looks most probable. eV͒, and we use this value for the off-diagonal exchange Let us discuss the possibility of this assignment from a matrix element. Then the absorption coefficient for two- numerical standpoint. A combination energy of infrared- magnon excitation is estimated to be of the order of ϩ ϩ Ϫ1 2 Ϫ1 active (D1 D3) and (D2 D4) is 189 meV (1533 cm ), 10 cm . This value is reasonable to explain the observed 2ϩ 2 2ϩ 2 while two overtones of Raman-active D1 D3 and D2 D4 results in Fig. 1. The advantage of direct optical excitation of Ϫ1 Ϫ1 ␣ are 184 meV (1482 cm ) and 194 meV (1565 cm ), re- two magnons in -Fe2O3 over MnF2 is speculated to arise spectively. The systematic difference between the observed from the large charge-transfer matrix element between the energies and the sum of corresponding energies given in O(2p) and Fe3ϩ(3d) orbitals, and its low excitation energy Ϫ1 ⌬ ␣ Table I is about 30 cm and may be attributed to the bind- E of the order of 2 eV for -Fe2O3 . This large matrix ing energy of two magnons.28 While a combination of acous- element at the same time results in large absorption coeffi- ϫ ϩ ϫ ͑ ͒ 4 Ϫ1 tic and optical magnons D1 D4 D2 D3 in Eq. 6 may be cient of the order of 10 cm and high Ne´el temperature Ќ ʈ excited by both polarized light E c axis (Eu) and E c axis 960 K through the large superexchange integral 2J. This is (A2u), the effective dipole moment of this process is an or- in contrast to the Ne´el temperature of MnF2 , which is 67.3 der of magnitude smaller than that of Eq. ͑5͒.10 Therefore, K. the mid-infrared absorption due to two-magnon excitation The dipole moment of two-magnon excitation is de- Ќ ␲ should be observed dominantly under polarization E c axis scribed by (i, j)Si•Sj . As to the three-magnon excitation, ͑ ͒ 15 through Eu in Eq. 5 . we first remember that Moriya has derived the so-called ϫ ͑ We shall show that three-magnon excitation is accessible Dzyaloshinski-Moriya interaction d•(Si Sj) as the off- ͒ Ϫ by both incident beams with the polarization parallel and diagonal exchange interaction 2JSi•Sj between the ions i perpendicular to the c axis. The three-magnon excitation and j, modified by the ͑off-diagonal͒ spin-orbit interaction may be possible for the combination of an optical magnon ␭(S"L). The magnitude of d here may be estimated very ⌫ ␭ ⌬ ϫ Ϫ ⌬ with Eu symmetry at the point and two magnons at the roughly as ( / E) ( 2J), where E is the energy of opposite BZ edges. We may use D points as the BZ edge electronic excitation from the ground ͉0͘ to the excited state point with high symmetry. Then the three-magnon excita- ͉n͘. This suggests that there may also exist an effective dipole ␲DM ϫ tions have the following symmetry: moment such as (i, j)•(Si Sj), which may be regarded as derived from the exchange dipole moment ␲(i, j)S S E ϫ͑D2ϩD2͒ϭE ϫ͑D2ϩD2͒ϭA ϩA ϩ2E i• j u 1 3 u 2 4 1u 2u u ͑responsible for the two-magnon absorption͒ perturbed by ϩ ϩ ϩ ͑ ͒ ␲DM ͓͓A1u A2u 2Eu͔͔, 10 the spin-orbit interaction. The magnitude of may then ␭ ⌬ ϫ␲ ϫ be given roughly by ( / E) . Note here that Si Sj con- E ϫ͑D ϫD ϩD ϫD ͒ϭE ϫ͑D ϫD ϩD ϫD ͒ Ϫ Ϫ Ϫ u 1 2 3 4 u 1 4 2 4 tains, e.g., a term like (Six iSiy)S jz Siz(S jx iSjy) which ϭ ϩ ϩ gives rise to three-magnon excitation. A1g A2g 2Eg The expression for the effective dipole moment PDM of ϩ ϩ ϩ ͑ ͒ ͓͓A1g A2g 2Eg͔͔. 11 three-magnon excitation is expressed as

ϫ 2ϩ 2 ͑ ͒ ͑DM͒ ͑DM͒ As a result, three-magnon excitations Eu (D1 D3) and P DM ϭP ϩP , ͑12͒ ϫ 2ϩ 2 i*j j*t Eu (D2 D4) are excited by the incident beam both for polarization parallel (A2u) and perpendicular (Eu) to the c ͗00͉H ͉n0͗͘n0͉P ͉00͘ ͑DM͒ϭ ͩ so i*j axis. The excitation energy, which is estimated as a sum of Pi j ͚ * n E ϪE the corresponding magnon energies in Table I, extends from i0 in Ϫ Ϫ 281 meV (2266 cm 1) to 291 meV (2349 cm 1). The en- ͗00͉P ͉n0͗͘n0͉H ͉00͘ ϩ i*j so ͑ ͒ ergy difference of the observed one from these is a little Ϫ ͪ 13 Ei0 Ein larger than 30 cmϪ1 ͑the binding energy of two magnons28͒ and may be attributed to the binding energy of three mag- with the exchange dipole Pi*j defined in Ref. 9. That is, nons. ϭ ␲͑ Ј← ͒ ͑ Ј← ͒ ͑ ͒ Pi*j ͚ i;n n, jm si n n •sjm , 14 nnЈm IV. MICROSCOPIC MODEL ͑ Ј← ͒ϭ † ͗ Ј͉ ͉ ͘ ͑ ͒ The possibility of electric dipole transition due to two- si n n ͚ cinЈmЈcinm m s m . 15 Ј and three-magnon excitation has been confirmed by the mm group-theoretical consideration and the magnon dispersion When we assume that the ion i is excited by the light to the ͉ ϭ͉␬ ␯ ͉ obtained by the neutron diffraction. In the two-magnon exci- lowest orbital state n͘ T1g ͘ from its ground state 0͘ ϭ͉ tation, the superexchange interaction between the neighbor- A1g͘, we have the following matrix element: 784 Low Temp. Phys. 31 (8–9), August–September 2005 Tanabe et al.

͗n0͉P ͉00͘ϭ⌸͑n0,00͒͗SϪ1M 0͉␴ ͉SM ͘ ͗0͉S ͉0͘, Im ⌸͑m0,00͒ i*j s i s • j ϩ ␨͑ ͒ ͗ ͉␶†͉ ͘ ͑16͒ ͚ i 0m Ϫ 0 j m m E j0 E jm

⌸͑n0,00͒ϭ ͚ ␲͑i;nЈ←n, jm͒ 1 1 ϫͩ S Ϫ Q S ͪ , ͑27͒ nnЈm 3 i S͑2SϪ1͒ j• i ␬4 ␯ʈ ← ͒ʈ6 ͗ T1g si͑nЈ n A1g͘ ϫ . ͑17͒ where we have made use of ͗SϪ1ʈ␴ʈS͘2S ␴† ␴ ϭ␦ ϩ͓ ͔ Ϫ ͑ Ϫ ͒ ͑ ͒ In terms of operator equivalent for off-diagonal ␣ ␤ ␣␤/6 S␣ ,S␤ /4S Q␣␤/2S 2S 1 , 28 29,30 H elements, so may be replaced by 1 1 ͗6 ͉H ͉␬4 0s␯͘ϭ␨͑ ͒͗ ͉␴†͉ Ϫ 0͘ Q␣␤ϭ ͑S␣S␤ϩS␤S␣͒Ϫ␦␣␤ S͑Sϩ1͒, ͑29͒ A1gM s so T1gM 0n SMs S 1M s 2 3 ͗A ͉␶†͉␬T ␯͘, ͑18͒ • 1g 1g and 6 ʈ ʈ␬4 ͗ A1g Vso T1g͘ where ␨͑0n͒ϭ , ͑19͒ ͗Sʈ␴†ʈSϪ ͘ ͗A ʈ␶†ʈ␬T ͘ ͒ ϭ ͑ ͒ 1 • 1g 1g ͑Qi Sj ␯ ͚ Qi;␯␤S j␤ . 30 • ␤ ␬ ʈ␶ʈ ϭ ʈ␶†ʈ␬ ϭ ) ͑ ͒ ͗ T1g A1g͘ ͗A1g T1g͘ i , 20 Here we evaluate ⌸ϭ(n0,00) by choosing ͉␬T aϩ ͘ in the ͗ Ϫ ʈ␴ʈ ͘ϭϪ͗ ʈ␴†ʈ Ϫ ͘ϭͱ͑ ϩ ͒ ͑ ͒ 1g 1 S 1 S S S 1 2S 1 /2. 21 trigonal basis32 as ͉n͘, because this is the orbital state excited ͑ ͒ The reduced matrix element of the spin-orbit interaction op- by Pi*j of Eq. 13 in the magnetically ordered phase. Ex- d ␰ erator Vso is found in Ref. 31: panding the 3 orbital in the trigonal field in terms of real , ␩, ␨ and the u,v orbitals in the O field,33 we find that ⌸ ͗6 ʈ ʈ␬4 ͘ϭ ␨ ͑ ͒ h A1g Vso T1g 6i , 22 ϭ(n0,00) is complex. As a consequence, both the vector ϫ with ␨ being the spin-orbit coupling parameter for the 3d product term (Si Sj) and the tensor term Qi•Sj and Qj•Si orbital. The operator ␶ is defined as term in Eq. ͑27͒ can contribute to the three-magnon excita- tion. The three-magnon excitation due to Q S and Q S ␶ ͉ ͘ϭ ͉␬ ␯͘ ͑ ͒ i• j j• i ␯ A1g i T1g , 23 ϫ terms as well as Si Sj term is always accompanied by the ͉␶†ϭϪ ␬ ␯͉ ͑ ͒ single magnon excitation by S or S with a common coeffi- ͗A1g ␯ i͑ T1g , 24 j i cient. We have not yet observed clear evidence of the single- ␯ϭ␣ ␤ ␥ Ϫ for , , , and magnon excitation around 800 cm 1 because the tail part of ͗n͉H ͉0͘ϭ␨͑0n͒*͗SϪ1M 0͉␴͉SM ͘ ͗␬T ␯͉␶͉A ͘, large absorption coefficient due to single phonons below so s s • 1g 1g Ϫ1 ͑25͒ 700 cm prevent clear observation of the single magnon, as shown in the inset of Fig. 2. The first term in Eq. ͑27͒ cor- where Sϭ5/2. ␲DM ϫ responds exactly to (i, j)•(Si Sj) obtained in Refs. 8 We now have and 10 simply following Moriya’s derivation of the DM in- ͗ ͉␨͑ ͒␴† ␶†͉ ͗͘ ͉ ͉ ͘ teraction. These spin operators are rewritten in terms of the ͑DM͒ 00 0n i • i n0 n0 Pi*j 00 P ϭ͚ ͫ i*j E ϪE magnon creation and annihilation operators through the Bo- n i0 in ϫ goliubov transformation. Then we realize that the Si Sj ͉ ͉ ͉␨ ␴ ␶ ͉ term as well as the Q S and Q S terms contribute to the ͗00 Pi j n0͗͘n0 ͑0n͒* i i 00͘ i• j j• i ϩ * • ͬ 8 E ϪE three-magnon excitation. Note also that the coefficients of i0 in both spin-product operators in the transition dipole moment ⌸͑n0,00͒ in Eq. ͑27͒ is naturally real. From this derivation, we can ϭ ␨͑ ͒ ͫ ͗ ͉␶† ͉ ͘␴† ␴ ͚ 0n ͚ Ϫ 0 ␣ n i␣ i␤S j␤ answer a question where the oscillator strength of three- n ␣␤ Ei0 Ein magnon excitation comes from. That is, when one of the two ⌸͑n0,00͒* 3ϩ Ϫ ͗ ͉␶† ͉ ͘␴† ␴ ͬ ͑ ͒ Fe ions participating in two-magnon excitation is per- Ϫ 0 ␣ n i␤ i␣S j␤ , 26 Ei0 Ein turbed by off-diagonal elements of the spin-orbit interaction, three-magnon mid-infrared absorption becomes possible us- because ␨(0n)*ϭ␨(0n). ing the effective exchange dipole moment ␲(i, j). The mag- We then obtain nitude of three-magnon absorption due to this process is es- Re ⌸͑n0,00͒ timated as follows. The spin-orbit coupling parameter P͑DM͒ϭ͚ͫ i␨͑0n͒ ͗0͉␶†͉n͘ Ϫ i ␭(0n)ϭ␨(0n)/2S is related to ␨, the atomic one, by ␭(0n) n Ei0 Ein ϭϪ␨/S. When we use ␨ϭ410 cmϪ1 for Fe2ϩ of magnitude ⌸͑ ͒ ␨ ⌬ ϭ Ϫ1 Re m0,00 † for and the excitation energy E 8000 cm , Ϫ͚ i␨͑0m͒ ͗0͉␶ ͉mͬ͘ 2 Ϫ2 Ϫ3 Ϫ j (␭(0n)/⌬E) is estimated to be of the order of 10 –10 . m E j0 E jm Here either one of the Fe3ϩ ions which contribute to two- Ϫ1 Im ⌸͑n0,00͒ magnon excitation may well accompany the third magnon ͑ ϫ ͒ϩ ␨͑ ͒ ͗ ͉␶†͉ ͘ • Si Sj ͚ i 0n Ϫ 0 i n 2S n Ei0 Ein excitation under the additional action of spin-orbit interac- tion. Taking into account this fact, the ratio of the absorption 1 1 coefficient I (3 magnons)/I (2 magnons) is estimated to ϫͩ S Ϫ Q S ͪ abs abs 3 j S͑2SϪ1͒ i• j be of the order of 10Ϫ2 in agreement with the observed re- Low Temp. Phys. 31 (8–9), August–September 2005 Tanabe et al. 785 sults. At the present stage, we cannot judge theoretically *E-mail: [email protected] which plays the more important role in the three-magnon excitation, Re ⌸ or Im ⌸ in Eq. ͑27͒. 1 M. Imada, A. Fujimori, and Y. Tokura, Rev. Mod. Phys. 70,1039͑1998͒. V. CONCLUSION 2 J. D. Perkin, J. M. Graybeal, M. A. Kastner, R. J. Birgeneau, J. P. Falck, and M. Greven, Phys. Rev. Lett. 71, 1621 ͑1993͒; J. D. Perkins, R. J. We conclude that two- and three-magnon excitation is Birgeneau, J. M. Graybeal, M. A. Kastner, and D. S. Kleinberg, Phys. Rev. induced around 1500 cmϪ1 and 2200 cmϪ1, respectively, by B 58, 9390 ͑1998͒. 3 J. D. Perkins, D. S. Kleinberg, M. A. Kastner, R. J. Birgeneau, Y. Endoh, the electric dipole moment in the antiferromagnetic ͑ ͒ ␣ K. Yamada, and S. Hosoya, Phys. Rev. B 52, R9863 1995 . -Fe2O3 . Two magnons at the D point have been assigned 4 M. Gru¨minger, D. van der Marel, A. Damascelli, A. Erb, T. Nunner, and to contribute to the mid-infrared absorption and Raman scat- T. Kopp, Phys. Rev. B 62, 12422 ͑2000͒. tering processes from the polarization dependence and the 5 M. Windt, M. Gru¨minger, T. Nunner, C. Knetter, K. P. Schmidt, G. S. Uhrig, T. Kopp, A. Freimuth, U. Ammerahl, B. Buchner, and A. Rev- group-theoretical consideration. colevschi, Phys. Rev. Lett. 87, 127002 ͑2001͒. In some crystals with a center of inversion such as the 6 H. Suzuura, H. Yasuhara, A. Furusaki, N. Nagaosa, and Y. Tokura, Phys. ͑ ͒ layered perovskites, e.g., La2CuO4 , the mid-infrared absorp- Rev. Lett. 76, 2579 1996 . 7 tion due to two-magnon excitation through the exchange di- J. Lorenzana and G. A. Zawatzky, Phys. Rev. B 52, 9576 ͑1995͒; Phys. Rev. Lett. 74, 1867 ͑1995͒. pole moment is not allowed, but phonon-assisted two- 8 S. Azuma, M. Sato, Y. Fujimaki, S. Uchida, Y. Tanabe, and E. Hanamura, magnon excitation was observed using the oscillator strength Phys. Rev. B 71, 014429 ͑2005͒. ␣ 9 ͑ ͒ of the phonon. However, this is not the case with -Fe2O3 , Y. Tanabe, T. Moriya, and S. Sugano, Phys. Rev. Lett. 15,1023 1965 . 3ϩ ͑ ͒ ͑ ͒ 10 Y. Tanabe and E. Hanamura, J. Phys. Soc. Jpn. 74, 670 ͑2005͒. whose unit cell contains four Fe ions with, e.g., 1 up, 2 11 ͑ ͒ ͑ ͒ 3ϩ S. J. Allen, Jr., R. Loudon, and P. L. Richards, Phys. Rev. Lett. 16, 463 down, 3 down, and 4 up Fe spins on the c axis. Al- ͑1966͒. though centers of inversion are located at the midpoints be- 12 R. Loudon, Adv. Phys. 13, 423 ͑1964͒. tween the 2nd and 3rd and between the 4th and 1st Fe3ϩ 13 T. Moriya, in Magnetism I, G. T. Rado and H. Suhl ͑eds.͒, Academic ͑ ͒ ions, two magnons may be excited from the antiferromag- Press, New York and London 1963 , chap. 3, and references cited in that 3ϩ article. netic pair of the 1st and 2nd or 3rd and 4th Fe ions, with 14 I. Dzyaloshinski, J. Phys. Chem. Solids 4, 241 ͑1958͒. no inversion symmetry around the midpoints. Note that these 15 T. Moriya, Phys. Rev. Lett. 4,228͑1960͒. are the first-neighbor pairs. Actually, we have more impor- 16 E. J. Samuelsen and G. Shirane, Phys. Status Solidi 42, 241 ͑1970͒. ͑ ͒ 17 E. J. Samuelsen, Physica ͑Amsterdam͒ 43, 353 ͑1969͒. tant 3rd- and 4th-neighbor antiferromagnetic pairs of ions 18 ͑ ͒ 16 S. Onari, T. Arai, and K. Kudo, Phys. Rev. B 16, 1717 1977 . with larger coupling if we take neighboring unit cells into 19 I. R. Beattie and T. R. Gilson, J. Chem. Soc. A 5, 980 ͑1970͒. account. No inversion symmetry is found for each of these 20 W. Kappus, Z. Physik B 21, 325 ͑1975͒. The character table of the Z point pairs, which allows them to make significant contributions to in Table 1 contains errors which should be replaced by the correct one in the direct optical excitation of two magnons in this crystal.10 Ref. 34. 21 T. R. Hart, S. B. Adams, and H. Temkin, in Proceedings of the 3rd Inter- On the contrary, we find that many layered perovskite crys- national Conference on Light Scattering in Solids, M. Balkanski, R. C. C. tals such as La2CuO4 and YBa2CuO6 have the center of Leite, and S. P. S. Porto ͑eds.͒, Wiley, New York ͑1976͒,p.259. inversion between the antiferromagnetic pair of ions, so that 22 T. P. Martin, R. Merlin, D. R. Huffman, and M. Cardona, Solid State ͑ ͒ only phonon-assisted two-magnon excitations are possible. Commun. 22, 565 1977 . 23 M. J. Massey, U. Baier, R. Merlin, and W. H. Weber, Phys. Rev. B 41, The polarization dependence and the absorption spec- 7822 ͑1990͒, and references cited in this paper. trum itself of the two-magnon excitation do not change so 24 F. J. Morin, Phys. Rev. 78, 819 ͑1950͒. much above and below the Morin temperature as shown in 25 R. Loudon, Phys. Rev. 137, A 1784 ͑1965͒. 26 R. Loudon, Adv. Phys. 17, 243 ͑1968͒. Figs. 1 and 2, where the sublattice magnetization rotates by 27 I. Inui, Y. Tanabe, and Y. Onodera, Group Theory and its Applications in ␲/2. This may be explained as follows: the magnon Physics ͑Springer Series in Solid State Sciences 78͒, Second Corrected dispersion17 and optical response of two magnons10 can be Printing, Springer-Verlag, Berlin, Heidelberg and New York ͑1996͒, chap. well described by Heisenberg model. This is consistent with 13. 28 R. J. Elliot and M. F. Thorpe, J. Phys. 2, 1630 ͑1969͒. the fact that the spin anisotropy energy and anisotropic ex- 29 K. Shinagawa and Y. Tanabe, J. Phys. Soc. Jpn. 30,1280͑1971͒. change interaction are much smaller than the isotropic ex- 30 T. Fujiwara and Y. Tanabe, J. Phys. Soc. Jpn. 32,912͑1972͒. change energy.13 As a result, optical response of two- and 31 Y. Tanabe, Suppl. Prog. Theor. Phys. 14,17͑1960͒. 32 three-magnon excitation depends only on the relative angle See the appendix of K. Eguchi, Y. Tanabe, T. Ogawa, M. Tanaka, Y. Kawabe, and E. Hanamura, J. Opt. Soc. Am. 22, ͑January 2005͒. between the neighboring spins but not on the absolute angle 33 S. Sugano, Y. Tanabe, and H. Kamimura, Multiplets of Transition-Metal of the sublattice magnetization relative to the crystal axis. Ions in Crystals, Academic, New York ͑1970͒. This can explain the other mystery why the polarization 34 E. R. Cowley, Can. J. Phys. 47, 1381 ͑1969͒. characteristic of two-magnon excitation does not change at This article was published in English in the original Russian journal. Repro- TM . duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 31, NUMBERS 8–9 AUGUST–SEPTEMBER 2005

Circular dichroism and Raman optical activity in antiferromagnetic transition-metal fluorides K. R. Hoffman

Department of Physics, Whitman College, Walla Walla, Washington 99362, USA D. J. Lockwood*

Institute for Microstructural Sciences, National Research Council, Ottawa, ON K1A 0R6, Canada W. M. Yen

Department of Physics and Astronomy, University of Georgia, Athens, Georgia 30602, USA ͑Submitted February 17, 2005͒ Fiz. Nizk. Temp. 31, 1032–1041 ͑August–September 2005͒ ͑ ͒ The Raman optical activity ROA of magnons in rutile-structure antiferromagnetic FeF2 (TN ϭ78 K) is studied as a function of temperature and applied magnetic field. For exciting light incident along the c axis, ROA is observed for magnons but not for phonons. In zero field, a small splitting (0.09 cmϪ1) of the two acoustic-magnon branches is observed for the first time by inelastic light scattering. The splitting in applied magnetic field is found to reduce with increasing temperature in accordance with theory. No ROA is detected for two-magnon excitations. In optical absorption measurements performed over thirty years ago, a very small circular dichroism ͑CD͒ was observed in the magnon sidebands of other simple rutile antiferromagnetic fluorides (MnF2 and CoF2). The origin of this CD was not understood at the time. The Raman studies of the one-magnon Raman scattering in FeF2 have demonstrated that in zero field the degeneracy of the antiferromagnetic magnon branches is lifted by a weak magnetic dipole-dipole interaction, as predicted by Pincus and Loudon and by White four decades ago. The source of the observed CD in the magnon sidebands can now be traced to this same magnetic-dipole induced splitting. © 2005 American Institute of Physics. ͓DOI: 10.1063/1.2008140͔

I. INTRODUCTION ported at that time the observation of circular dichroism ͑ ͒ ͑ ͒ CD in certain magnon sidebands of MnF2 Ref. 8 and ͑ ͒ ͑ ͒ In 1965, the discovery of the one-magnon sideband in CoF2 Ref. 10 but not in FeF2 Ref. 9 ; CD occurs in the ͑ ͒ the visible absorption spectrum of MnF2 Ref. 1 and the absence of an applied magnetic field and corresponds to zero two-magnon absorption of the far-infrared spectrum of FeF2 field MCD though the origins of the dichroism may arise ͑Ref. 2͒ spurred many investigations into the optical proper- from different sources. The authors were able to eliminate ties of antiferromagnetic insulators. The theoretical frame- experimental artifacts as the source of the observed CD, but work regarding these two processes was developed by were not able at that time to identify the physical basis for Tanabe et al.3 and Allen et al.4 In these materials, coupling the unexpected observation. Recently, we revisited this prob- between the various intrinsic excitations of the lattice, i.e., lem and showed that the source of the CD is to be found in excitons, magnons and phonons, leads to the formation of the dipolar interactions that exist between the antiferromag- cooperative transitions which can account for much of the netic sublattices, which render them inequivalent.11 spectral structure. An article discussing this work appears Raman optical activity ͑ROA͒ characterizes the differ- elsewhere in this special issue of Low Temperature Physics. ence in the Raman line intensity when excited with right and In addition to absorption and luminescence measurements, left circularly polarized light. The earliest experiments to magnon lines were also observed in Raman scattering combine Raman spectroscopy with chirality in solids were measurements.5,6 Raman scattering measurements permit the performed in the 1970s on CdS under resonance excitation physical properties of magnons to be studied without the conditions.12 Attempts were also made to observe a similar additional contributions due to electronic states associated effect in antiferromagnetic fluoride materials but the experi- with optical emission or absorption. A comprehensive review mental sensitivity was insufficient. At about the same time of Raman theory and experimental results for antiferromag- ROA was also measured in a chiral molecule.13 The chiral nets is presented in the book by Cottam and Lockwood.7 structure of some molecules makes them naturally suited to The magnetic character of the magnon transitions led to couple more strongly to either right or left circularly polar- the utilization of circularly polarized light to explore the di- ized light. The theoretical14 and experimental developments chroic properties of these materials. In the early 1970s, a in this field have been reviewed in the literature.15,16 A series series of papers reported on their magnetic circular dichroism of Raman papers in the 1980s on the magnetic properties of ͑MCD͒ in an applied magnetic field.8–10 The authors re- dilute magnetic employed the use of circular

1063-777X/2005/31(8–9)/8/$26.00786 © 2005 American Institute of Physics Low Temp. Phys. 31 (8–9), August–September 2005 Hoffman et al. 787 polarization to isolate spin flip transitions of Mn ions.17,18 For simple antiferromagnets in their ordered state, the The improvement in detector sensitivity over a period of ground state of each sublattice is non-degenerate and is iden- twenty years afforded us the opportunity in the 1990s to tical to the other except for their sense of magnetization with revisit the question of ROA in antiferromagnetic insulators. respect to the unique crystalline axis. If the state for the ‘‘up’’ We were able to observe ROA in the one-magnon line of or A sublattice is given by ͉L,M͘ then the corresponding 19–21 ͉ Ϫ FeF2 and several phonon lines in rutile-structure ‘‘down’’ or B sublattice is represented by L, M͘. For pure materials.22,23 In addition, we demonstrated for the first time electronic transitions, the appropriate dipolar transition op- that Raman scattering could be used to measure a zero field erators for circular polarized radiation transform as LϮ splitting of the magnon branches in antiferromagnetic ϩ2SϮ , where the ͑ϩ͒ and ͑Ϫ͒ subscripts refer to right and 24 FeF2 . These results were then utilized to reinterpret the CD left circular polarizations, respectively. If ions in one of the measurements completed in the early 1970s that were al- sublattices have a nonvanishing transition probability be- luded to above.11 tween two states for one of these operators, then the ions in In the following Sections, we present our current under- the opposite sublattice will have an identical element for the standing of the dichroic properties of antiferromagnetic insu- corresponding conjugated operator. In the absence of a mag- lators possessing the rutile structure. The focus of this review netic field, the sublattices are degenerate in energy and no will be on ROA measurements to explore the magnetic prop- CD should be observed. However, weak interactions between erties of these materials. We start with a summary of the zero the sublattices, such as magnetic dipole-dipole interactions, field dichroism measurements in MnF2 , CoF2 , and FeF2 be- can produce a small anisotropy between the two sublattices cause the CD studies prompted one of us ͑W.M.Y.͒ to ex- that results in CD. At low temperatures, a very small zero- plore the use of ROA in antiferromagnetic fluorides. We then field CD signal was observed in two of the three compounds present a more thorough review of the ROA measurements, investigated and it is the origin of these signals that will be techniques, and analysis in FeF2 . Here we mention first the addressed in this paper. work on phonons before focusing on the one- and two- magnon scattering results, including a revised assessment of the results from the temperature dependence and zero mag- III. EXPERIMENTAL RESULTS FOR CIRCULAR DICHROISM netic field studies of FeF2 . MEASUREMENTS A finite CD was observed at low temperatures in two of II. OPTICAL DICHROISM IN RUTILE ANTIFERROMAGNETS the three difluorides studied. Signals were observed in the 6 4 magnon sidebands of the A1g ground state to T1g and 4 4 4 2 2 Dichroism in a material is defined as the difference in ( A1g , Eg) absorptions of MnF2 and of the T1 to A, T1 2 the absorption coefficient of a medium for two related polar- ( P) of CoF2 ; no CD signals were observed in FeF2 .As izations; thus CD arises from the difference in the absorption noted earlier, the origins of CD and MCD are different; as a for right- and left-circularly polarized radiation. When these consequence, CD can be subtracted from MCD in order to differences are induced by the application of an external determine the magnetic field induced splittings. When this is magnetic field, the effect is termed MCD. In both CD and done the, the residual or pure MCD signal can be fitted ac- MCD, a unique direction is necessary in order to define the curately with a differential line shape; this indicates that in sense of circulation in the circularly polarized light; this is our cases the MCD arise simply from the lifting of the sub- provided by some unique crystalline axis or by the direction lattice degeneracy, as per our expectation. of the applied magnetic field. Materials showing CD are nor- Figure 1 shows the absorption of a magnon sideband of 2 2 mally called ‘‘optically active’’ and such crystals are rela- the ( A, T1) transition of CoF2 along with the observed CD tively rare, whereas MCD is quite common in paramagnetic and MCD in a 2 T applied magnetic field. The CD is also systems possessing a unique axis. MCD is conjugated to Far- purely differential in shape and is estimated to be an additive aday rotation by the Kramers–Kronig relations and consti- intrinsic splitting of the order of ⌬␻ϭ0.015 cmϪ1. The CD tutes the resonant rather than the dispersive phenomena in in CoF2 becomes unobservable in the vicinity of 10 K. the transformation. Figure 2 shows the absorption spectrum of the so-called ␴ ␴ 4 The compounds that we wish to discuss here are the 1 and 2 sidebands accompanying the ground state to T1g ␣ common difluorides, MnF2 , CoF2 , and FeF2 , all of which transition in MnF2 . These transitions are -polarization ac- possess the rutile crystal structure. Below the Ne´el tempera- tive. A small CD signal is observed that is clearly related to ture TN these materials are describable as simple two sublat- this transition. The zero-field splittings for the sidebands ⌬␻ ␴ ϭ Ϫ1 ⌬␻ ␴ tice antiferromagnets and in principle they are ideal systems were determined to be ( 1) 0.07 cm and ( 2) on which to study magnetically dichroic phenomena. The ϭϪ0.05 cmϪ1, respectively. ⌬␻ ␴ ground states of ions in each sublattice as well as the collec- The negative sign in ( 2) signifies that the sense of ͑ ͒ ␴ tive magnetic excitations magnons are identical to each the splitting in this sideband is out of phase with that of 1 . other in every respect except their polarization. The sole re- Such a change in the sense of splitting is also observed when ͓ ͔ maining energy degeneracy, which will respond to an applied uniaxial stress is applied to MnF2 along the 110 direction in magnetic field, is that of the two sublattices; hence, the ori- the basal plane. The behavior of the CD was studied as a gin of magnetic dichroism in these systems is directly related function of basal plane stress allowing the intrinsic strain in ⌺ϭ⌬ to the lifting of this energy equivalence both in the electronic the basal plane of ordered MnF2 to be estimated: d/d ͑exciton͒, the spin ͑magnon͒ or the combined ͑magnon side- ϭ1.2ϫ10Ϫ5. This result highlights the advantages of chiral band͒ systems. techniques for exploring interactions in antiferromagnets. 788 Low Temp. Phys. 31 (8–9), August–September 2005 Hoffman et al.

FIG. 3. The electronic levels of the magnetic ions in antiferromagnetically ordered rutile materials responsible for the observed ROA signal. The two sets of lines correspond to the two antiferromagnetic sublattices. In both cases a magnon is generated during the scattering process.

R L I␣ϪI␣ ⌬ ϭ ͑ ͒ ␣ R L , 1 I␣ϩI␣ R L where I␣ and I␣ represent the scattered light intensity with ␣ Ϫ FIG. 1. The CD spectrum of the one-magnon sideband near 22,800 cm 1 of polarization due to incident right (R) or left (L) circularly CoF2 at 4 K and the corresponding absorption spectrum of the sideband. polarized light. The normalization term permits a simple comparison of data from different spectroscopic systems. One major difficulty of ROA measurements is the sensitivity Ϫ1 FeF2 was studied in the 21500– 28500 cm spectral re- of the technique to intrinsic or extrinsic depolarization ef- gion. The observed transitions without exception exhibited fects. Care must be taken to account for any birefringent MCD signals consistent with a simple sublattice splitting and effects in the steering and focusing optics. Additionally, the with a linear dependence on magnetic field. No CD signals crystal itself can contribute to significant depolarization ef- were encountered in any of the ␣-active transitions investi- fects that appear as CID in spectra that normally would not gated. The sensitivity of our apparatus places an upper limit be expected to produce any difference. We will discuss this of 10Ϫ3 cmϪ1 on any splitting that might occur in this com- effect in more detail in the next Section of the paper. pound at 4 K. Fundamentally, the source of Raman scattering is the modulation of the induced polarizability of a medium by fundamental excitations in the medium. To couple the inci- IV. RAMAN OPTICAL ACTIVITY dent light to the medium, its susceptibility is expanded in terms of phonon or magnon operators. This results in a Raman optical activity is observed experimentally in the modulation of the polarizability at the frequency of the fun- difference in a Raman line intensity when excited by right damental excitation. In the case of magnons, the inclusion of and left circularly polarized light, respectively. The circular spin operators results in a change of the overall angular mo- intensity difference ͑CID͒ is normally used to quantify mentum component along the z axis: ⌬m ϭϮ1. Figure 3 ROA:14 j illustrates the important mechanism for magnon light scatter- ing from antiferromagnetic fluorides. The energy levels shown in the picture belong to the transition metal ions on the two sublattices. These transitions obey the same selection rules that contribute to CD or MCD signals. The circularly polarized light couples to different sublattices in the Raman process as well, thus providing a way of distinguishing be- tween the two sublattices. The figure shows energy levels for both sublattices in the medium and illustrates that left and right circular polarizations create magnons on different sub- lattices. Symmetry arguments can be applied to determine selec- tion rules for the various scattering processes and are sum- marized in the form of Raman scattering tensors. The scat- tering tensors for the tetragonal antiferromagnetic fluorides are presented in Table I. The tensors have been transformed so that they are ap- FIG. 2. The absorption and CD spectra of the so-called ␴ and ␴ sidebands 1 2 propriate for ͓R,L,z͔ incident polarizations and ͓x,y,z͔ scat- of MnF2 at 2 K; these traces connect the observed CD to the magnon sidebands. tered polarizations. The Raman tensors describe the scattered Low Temp. Phys. 31 (8–9), August–September 2005 Hoffman et al. 789

TABLE I. The Raman tensors for circularly polarized incident light that is analyzed for linearly polarized light scattered at 90°.

electric field amplitudes and thus the scattered light intensity was measured using a Stanford SR-400 dual-gated photon in a specific polarization is calculated through multiplication counter and stored on a computer. The signals counted on the of the tensor by its conjugate value. To describe the expected two gates, corresponding to the two incident polarizations, CID spectrum, we need to calculate the difference between were stored separately, permitting both IRϪIL and IRϩIL to two intensity terms. For example, the A1g phonon mode be easily calculated. Low temperatures were achieved using should not generate a CID signal for scattered light polarized a Cryosystems closed-cycle refrigerator. A permanent mag- along the x axis: net assembly located inside the refrigerator applied a con- RϪ L 2Ϫ 2 stant magnetic field parallel to the crystal c axis. The mag- Ix Ix a a ⌬ ͑A ͒ϭ ϭ ϭ0. ͑2͒ nets were drilled through so that the incident laser beam x 1g Rϩ L 2ϩ 2 Ix Ix a a could be focused in the crystal collinear with the applied field and allowed to pass on through the refrigerator. In the case of the magnon scattering we can calculate the ϩ The crystals used in this study were the rutile-structure CID for each sublattice separately. In the case of the ⌫ 3 fluorides FeF and MgF . These materials are uniaxial, with band, we obtain the following CID expression: 2 2 FeF2 undergoing an antiferromagnetic phase transition at ͑␧*␧ϩ␧*2ϩ␧*␧ϩ␧2͒Ϫ͑␧*␧Ϫ␧*2ϩ␧*␧Ϫ␧2͒ T ϭ78 K. The MgF crystal orientation was determined ⌬ ͑⌫ ͒ϭ N 2 z 3 ͑␧*␧ϩ␧*2ϩ␧*␧ϩ␧2͒ϩ͑␧*␧Ϫ␧*2ϩ␧*␧Ϫ␧2͒ from x-ray Laue diffraction measurements before the cuboid sample was cut ͑to within a 1° accuracy͒ and polished. The ␧ 2ϩ␧2 * c-axis orientation for the FeF2 sample was found using a ϭ . ͑3͒ 2͑␧*␧͒ polariscope. In addition, a polished crown glass cube was used as a reference material to check for instrumental arti- ⌫ϩ Likewise for the 4 magnon band we get a similar re- facts. sult: Rayleigh and Raman CID in MgF2 and FeF2 ␧*2ϩ␧2 ⌬ ͑⌫ ͒ϭϪ . ͑4͒ Our measurements of the magnon and phonon CID in z 4 2͑␧*␧͒ FeF2 generated several signals that were difficult to under- R The negative sign reflects the coupling of left circular stand initially. Figure 5 shows the difference spectrum I Ϫ L R L polarization to this particular sublattice. Notice that if the I as well as the separate I and I spectra from FeF2 with energies of the two magnon bands are equal and the line no applied magnetic field. What is curious about these mea- profiles equivalent, the total CID is the sum of these terms surements is that the magnon exhibits no apparent differen- and will be zero. The application of a magnetic field lifts the tial scattering but the B1g phonon mode does. When the degeneracy of the magnon branches. As a result, the CID of spectrum is enlarged to encompass all of the phonon modes the two branches will no longer completely cancel. This re- and the Rayleigh line, we observe a similar CID spectrum for sult is analogous to the differences observed between MCD the A1g and B2g modes and the Rayleigh line but the sign of and CD spectra. The important distinction is that the Raman the difference signal is opposite to that of the B1g mode. spectra are determined only by the energies of the magnon Referring back to the Raman tensors, no CID is expected bands with no interference from the excited states of the ion. from any of the phonon modes nor should Rayleigh scatter-

V. RAMAN OPTICAL ACTIVITY EXPERIMENTS The experimental arrangement, shown in Fig. 4, used the standard 90° Raman scattering geometry. The incident beam was the 488-nm line of a continuous-wave Spectra Physics 166 argon–ion laser. The focused laser power density at the sample was 150 kW/cm2, well below the damage threshold of the materials used. A Conoptics 370 electro-optic modu- lator ͑EOM͒ was used to alternate between right and left circular polarization. The scattered light was dispersed by a FIG. 4. The experimental setup for the ROA measurements. The laser light Spex 1401, 0.85-m double monochromator and detected us- was directed vertically through the sample to align the beam path with the ing a cooled RCA 31034-A02 photomultiplier. The signal entrance slit of the spectrometer. 790 Low Temp. Phys. 31 (8–9), August–September 2005 Hoffman et al.

FIG. 5. The ROA spectrum of FeF2 at 4 K. The presence of a CID spectrum in the B1g phonon mode indicates a misalignment of the incident laser and the crystalline c axis.

FIG. 6. The low-temperature ROA spectrum of FeF2 in the presence of an ing exhibit CID. MgF2 is also a rutile-structure fluoride, but applied magnetic field. The upper trace shows the conventional Raman spec- it does not exhibit any magnetic ordering. It showed a similar trum. The lower trace is the measured CID spectrum. differential scattering pattern. An isotropic glass sample was used to check for instrumental artifacts and none were found. ⌬␻͑ ͒ϭ ͓ ␮ Ϫ ␹ ͔ ͑ ͒ To explain the CID spectra of the phonons we looked T 2 g B J ʈ H 5 carefully at the effects of birefringence in the crystal on the where H is the applied magnetic field along the crystal c ␹ ␮ scattered intensity. Light propagating along the c axis will axis, ʈ is the parallel susceptibility per spin site, B is the experience an isotropic index of refraction so that no change Bohr magneton, J is the dominant nearest neighbor ex-

in the polarization occurs as it moves through the sample. A change, and g is the Lande g factor. At low temperatures, ␹ʈ slight misalignment results in distinct indices of refraction is nearly zero, so parallel and perpendicular to the plane containing the c axis. ⌬␻͑ ϭ ͒ϭ ␮ ͑ ͒ As a result, circularly polarized light becomes elliptically T 0 2g BH 6 ␹ ϭ ␮ polarized as it moves through the sample. The changes in whereas at temperatures near TN , ʈ g B /2J so: polarization affect the relative amplitude of the x and y com- ⌬␻͑TϭT ͒ϭg␮ H. ͑7͒ ponents of the incident light. The crucial point is that N B changes of Ix and Iy are different for left and right circular This theory predicts that the splitting between the branches polarization so that any scattering that depends on only one should change by a factor of two when going from low tem- of these terms will exhibit a CID spectrum. This result per- perature to the phase transition temperature. Using param- mits a very sensitive tool for aligning the samples for ROA measurements so that intrinsic birefringence effects can be eliminated.

Raman optical activity of magnons in FeF2 Figure 6 presents the conventional Raman and ROA spectrum of the one-magnon line of FeF2 in an applied mag- netic field. The modest applied field, BϽ0.6 T, is too small to observe the splitting of the magnon branches within the magnon linewidth using conventional ͑linearly polarized light͒ Raman techniques. However, the magnon line exhibits a clear CID spectrum. To determine the energy difference between the two branches we fit the data using offset Gauss- ian line shapes to describe the one magnon scattering peak. In this manner we can calculate the energy difference as a function of temperature. Figure 7 shows the temperature de- pendence of the frequency splitting between the two magnon branches in FeF2 . As the temperature increases, the energy splitting between the branches decreases, while the linewidth increases resulting in a larger uncertainty for the energy split- FIG. 7. The temperature dependence of the magnon branch splitting ob- tained from ROA measurements in FeF2 . The theoretical points are calcu- ting between the branches. ͑ ͒ ͑ 28 lated using applied fields of 0.59 filled diamonds and 0.55 T filled Theoretically, the energy splitting between the two circles͒. The error bars on the theoretical points derive largely from the branches is given by uncertainty in the applied magnetic field strength. Low Temp. Phys. 31 (8–9), August–September 2005 Hoffman et al. 791

excitation conditions.17,18 These transitions were observed in a forward scattering geometry where the scattered light was analyzed for the circular polarization as well. The intensities of the two spin flip signals are much lower than the single spin flip signals consistent with a higher-order interaction.

VI. MAGNETIC DIPOLE-DIPOLE INTERACTIONS In 1963 Loudon and Pincus25 examined the effect of the classical dipole-dipole interaction between magnetic mo- ments on the spin wave spectrum of a simple uniaxial anti- ferromagnet. In the absence of an applied magnetic field, as discussed above, the spin wave branches are degenerate in these systems. The magnon dispersion relation is given by ␻ϵ␥͓ 2 ϩ ϩ 2 2 2͔1/2 ͑ ͒ HA 2HAHE 2HEb k 8 ␥ϭ ␮ where g B is the gyromagnetic ratio, HA is the uniaxial single-ion internal anisotropy field, and HE is the effective ϭ Ϫ1/2 FIG. 8. The zero-field ROA spectrum of the one magnon line in FeF2 .The exchange field; k is the wave vector and b az , with a solid line is a fit modeling the small CID apparent in this spectrum. the nearest-neighbor distance and z the number of nearest neighbors. On inclusion of the dipolar fields, Eq. ͑8͒ splits into two modes with their dispersions given by: eters obtained from the experiment or the literature the the- ␻ ϭ␥͓ 2 ϩ ϩ 2 2 2͔1/2 ͑ ͒ oretical curve shown in Fig. 7 was obtained. The theoretical 1 HA 2HEHA 2HEb k , 9 curve accurately describes the temperature variation of the ␻ ϭ␥͓ 2 ϩ ͑ ϩ 2 2͒͑ ϩ ␲ 2 ␪ ͔͒1/2 energy splitting between the magnon bands, but overall there 2 HA 2 HA HEb k HE 4 M s sin k ͑ ͒ is a general offset between the measured data and the theo- 10 ␪ retical curve. The primary uncertainty for this measurement where M s is the sublattice magnetization and k is the angle arises from determining the actual value of the magnetic field between the easy axis (z) of magnetization and the direction inside the refrigerator. The magnetic field of 0.59 T for the of propagation k of the spin wave. Only the frequency of the permanent magnet gap was measured at room temperature. second mode is affected by the dipolar interaction. For spin The magnet will be at a much lower temperature during the waves with kϭ0, Eqs. ͑9͒ and ͑10͒ reduce to Raman measurement, but the affects of the lower tempera- ␻ ϭ␥͓H ͑H ϩ2H ͔͒1/2 ͑11a͒ ture on the gap field strength is not known. 1 A A E ␻ ϭ␥ ϩ ϩ ␲ 2 ␪ ͒ 1/2 ͑ ͒ These measurements highlight the sensitivity of this 2 ͓HA͑HA 2HE 8 M s sin 0 ͔ . 11b technique in measuring magnon branch splittings. The line- In the case of longitudinal spin waves ␪ ϭ0, and there is widths of the two peaks at 62 K were greater than 10 cmϪ1, 0 no dipolar effect. For the transverse spin waves, however, but the separation between the peaks was only 0.4 cmϪ1. ␪ ϭ␲/2, and the dipolar effects are at a maximum. For a simple antiferromagnet, we would not expect to 0 Subsequently, Harris26 has shown that Eq. ͑10͒ is an ap- observe any CID spectrum in zero applied field ROA mea- proximate result and has derived the full expression for the surements. However, careful measurements revealed a small spin wave dispersion in the presence of a dipolar field. How- but repeatable CID spectrum in FeF . Figure 8 shows the 2 ever, the difference between Eq. ͑10͒ and the complete result CID spectrum and the separate IR and IL Raman signals of only becomes important when 4␲M is comparable to H , FeF . Care was taken to eliminate all of the depolarization s E 2 which is not the case here. effects by using the Rayleigh scattering spectrum to ensure The Loudon and Pincus calculation is based on a mo- alignment of the crystal c axis with the incident laser light. lecular field approach that neglects the affect of the Lorentz The solid-line fit determined the zero field splitting between field of one sublattice upon the other sublattice. When this the two magnon branches to be ⌬␻ϭ(0.09Ϯ0.02) cmϪ1.To effect is included, White27 showed that the resonant frequen- our knowledge this result is the first measurement of a zero cies for kϭ0 are given by: field splitting of the magnon branches in antiferromagnets using Raman scattering. Again the high sensitivity and reso- 8 ␻͑kϭ0͒ ϭϮ␥ͫH ͩ H Ϫ ␲M ͪ ϩ2H ͑H lution of this technique is highlighted in obtaining this result. 1,2 A A 3 s A E Finally, ROA measurements of the two-magnon scatter- 1/2 ing peak in FeF showed a null result. The lack of any CID 2 ϩ4␲M N ͒ͬ , ͑12a͒ signal confirms that the dominant mechanism for two- s x magnon scattering involves exchange coupled magnons on 8 opposite sublattices. Other possible sources of two-magnon ␻͑ ϭ ͒ ϭϮ␥ͫ ͩ Ϫ ␲ ͪ k 0 3,4 HA HA M s scattering are higher order processes that would be expected 3 to be much weaker than the one-magnon scattering. Multiple 1/2 ϩ ͑ ϩ ␲ ͒ͬ ͑ ͒ spin-flip transitions have been observed in dilute magnetic 2HA HE 4 M sNy 12b semiconductors in the paramagnetic phase under resonance 792 Low Temp. Phys. 31 (8–9), August–September 2005 Hoffman et al. and for k0by TABLE II. Magnetic parameters at low temperature for various transition- metal fluorides possessing the rutile structure. The calculated dipole-dipole induced splittings, ⌬␻ ϭ and ⌬␻ ϭ␲ , are also given. ␻͑ ͒ ϭϮ␥ͫ 2 ϩ ϩ 2 ͑ Ϫ␥ 2͒ k 0 k /2 k 1,2 HA 2HAHE HE 1 k

8 1/2 ϩ ␲M ͓H ϩH ͑1Ϫ␥ ͔͒ͬ , ͑13a͒ 3 s A E k

␻͑ ͒ ϭϮ␥ͫ 2 ϩ ϩ 2 ͑ Ϫ␥ 2͒ k 3,4 HA 2HAHE HE 1 k

1 1/2 ϩ8␲M ͩ Ϫsin2 ␪ ͪ ͓H ϩH ͑1Ϫ␥ ͔͒ͬ s 3 k A E k ͑13b͒ 27 ␥ where N is the demagnetization dyadic, and k is defined as (1/z)͚␦ exp(ik"␦), where the ␦’s are the vectors to the z nearest neighbors. Notice that for the uniform spin precession (kϭ0) in a ϭ ϭ ϭ sphere (Nx Ny Nz 1/3), the Lorentz field just cancels the surface demagnetizing field. Therefore, there is no shift in ample, the sideband involves a combination of an exciton on the resonant frequency as compared to non-dipolar cases. one sublattice with essentially zero dispersion25 with a mag- The symmetric occurrence of N and N in Eq. ͑12͒ is due to Ϫ x y non on the other with nearly 50 cm 1 dispersion; k conser- the fact that the normal modes labeled 1, 2 and 3, 4 are vation requires the creation of exciton-magnon pairs with characterized by net transverse magnetizations that are lin- equal and opposite wave vectors from throughout the Bril- early polarized in the x and y directions, respectively. Notice louin zone. Although the splitting is mostly uniform across also that although the 1, 2 spin wave modes are independent the Brillouin zone, the CD is, in effect, sampling the deriva- of ␪ they are shifted by the Lorentz field. tive of the joint density of states. It follows that an accurate Since the samples used in the optical studies had rectan- calculation of the CD requires the use of Eq. ͑13͒, but the gular cross sections in the basal plane, there will be dipolar theory needs to be extended to include specific symmetry splittings of the magnons. Thus, from Eq. ͑12͒, points in the Brillouin zone and an appropriate summation ␲ Ϫ over the entire zone. 4 M sHA͑Nx Ny͒ ⌬␻ ϭ ϭ␻ Ϫ␻ ϭ . ͑14͒ Conceivably, another possible cause of CD in these com- k 0 1 3 ͱ ϩ 2 2HAHE HA pounds could be a slight tilting of the spins away from their The demagnetizing factors are given by integrals over ordered positions along the c axis. However, at low tempera- the surface.27 In order to obtain estimates of these splittings tures and in zero applied magnetic field, all experimental we approximate Nx and Ny by 1/2 and 1/4, respectively, results to date point to collinear spin alignment in all of these based on the sample dimensions. The parameter values used compounds.29 in the calculations and the resulting splittings are listed for MnF , CoF , and FeF in Table II. 2 2 2 Raman circular intensity difference In the Raman equivalent of the CD measurements dis- VII. COMPARISON OF THEORY WITH EXPERIMENT cussed above, a ⌬␻ of (0.09Ϯ0.02) cmϪ1 was observed in Optical circular dichroism FeF2 at low temperatures using a 90° scattering geometry. From Eq. ͑11͒, the theoretical splitting for magnons propa- The calculated splittings given in Table II are quite small gating at approximately 45° to the c (z) axis is ⌬␻ compared with the magnon energies in these compounds. ϭ0.15 cmϪ1, which, considering the expected large uncer- Given the nature of the natural linewidths, splittings of this tainty in M , is in good agreement with the experimental magnitude are difficult to measure and they have been re- s results. Including the Lorenz field correction, Eq. ͑14͒ gives solved only recently using Raman spectroscopy.11,15 ⌬␻ϭ0.065 cmϪ1 ͑see Table II͒, which is in even closer The calculated values for ⌬␻ in MnF and CoF ͑see 2 2 agreement with experiment. The zero-field splitting of the Table II͒ are of similar magnitude to those found experimen- magnon branches observed in FeF is thus attributed to the tally for the magnon sidebands in the visible region: 2 Ϫ Ϫ magnetic dipole-dipole interactions that can be important for 0.02 cm 1 for CoF and 0.07 and 0.05 cm 1 for MnF . 2 2 long wavelength magnons.27 Thus the origin of the observed splittings reported in the CD Another way to quantitatively describe the interaction is measurements is well encompassed by the effects of the to define it in terms of an effective anisotropy field, HЈ, that magnetic dipole-dipole interaction in these simple rutile an- is treated in the same manner as an external magnetic field. tiferromagnets. The size of the splittings detected by sponta- In this approach the energy of the magnons shows up in neous magnetic CD cannot be compared directly with those much the same form as presented in Eq. ͑6͒ except that the calculated in Table II, as the processes involved in the for- splitting is now due to HЈ. mation of magnon sidebands are not as simple and direct as ⌬␻ ϭ ͒ϭ ␥ ϭ Ϫ1 ͑ ͒ those encountered in Raman scattering. For MnF2 , for ex- ͑T 0 2 HЈ 0.09 cm . 15 Low Temp. Phys. 31 (8–9), August–September 2005 Hoffman et al. 793

Ϫ Using the parameter value ␥ϭ1.05 cm 1/T ͑Table II͒ we *E-mail: [email protected] calculate the anisotropy field to be: HЈϭ(0.04Ϯ0.01) T. By comparing Figs. 8 and 6 we see that the CID spectra 1 R. L. Greene, D. D. Sell, W. M. Yen, A. L. Schawlow, and R. M. White, ϭ Ј ϭ Phys. Rev. Lett. 15, 656 ͑1965͒. have opposite signs for the case of H H and H Happlied 2 J. W. Halley and I. Silvera, Phys. Rev. Lett. 15,654͑1965͒. ϩHЈ, respectively. Since the energy splitting of the two 3 Y. Tanabe, T. Moriya, and S. Sugano, Phys. Rev. Lett. 15,1023͑1965͒. magnetic sublattices switches sign with the external field, the 4 S. J. Allen, R. Loudon, and P. L. Richards, Phys. Rev. Lett. 16,643 direction of the applied field must be opposite to the direc- ͑1966͒. 5 P. A. Fleury, S. P. S. Porto, L. E. Cheeseman, and H. J. Guggenheim, Phys. tion of the effective anisotropy field. Going back to the tem- Rev. Lett. 17,84͑1966͒. perature dependence of the magnon branch splitting we can 6 P. A. Fleury and R. Loudon, Phys. Rev. 166,514͑1968͒. now correct for the presence of the effective anisotropy field 7 M. G. Cottam and D. J. Lockwood, Light Scattering in Magnetic Solids, ͑ ͒ in the material. Equation ͑5͒ becomes Wiley, New York 1986 . 8 F. L Scarpace, M. Y. Chen, and W. M. Yen, J. Appl. Phys. 42,1655͑1971͒. ⌬␻ ͒ϭ ␮ Ϫ ␹ Ϫ ͒ 9 M. Y. Chen, F. L. Scarpace, M. W. Passow, and W. M. Yen, Phys. Rev. B ͑T 2͓g B J ʈ͔͑Happlied HЈ 4,1332͑1971͒. ϭ ␮ Ϫ ␹ ͑ ͒ 10 Y. H. Wong, C. D. Pfeifer, and W. M. Yen, in Magnetism and Magnetic 2͓g B J ʈ͔͑0.55 T͒. 16 Materials (1972), C. D. Graham, Jr. and J. J. Rhyne ͑eds.͒, AIP Confer- The second theoretical curve shown in Fig. 7 uses this ence Proc. No. 10, AIP, New York ͑1973͒, p. 1960. 11 new value for the magnetic field and gives better agreement D. J. Lockwood, R. M White, and W. M. Yen, Phys. Rev. B 69, 174413 ͑2004͒. with the measured data. 12 W. M. Yen and Y. Oka, Bussei Kenkyu 13, 507 ͑1972͒. 13 L. D. Barron, M. P. Bogaard, and A. D. Buckingham, J. Am. Chem. Soc. 95,603͑1973͒. 14 L. D. Barron and A. D. Buckingham, Mol. Phys. 20, 1111 ͑1971͒. 8. CONCLUSIONS 15 L. D. Barron, Molecular Light Scattering and Optical Activity, Cambridge University Press, Cambridge ͑1982͒. The magnon sideband CD observed in rutile-structure 16 L. D. Barron and L. Hecht, in Circular Dichroism: Principles and Appli- antiferromagnets can be quite readily explained as being the cations, J. Nakanishi, N. Berova, and R. W. Woody ͑eds.͒, VCH Publish- result of magnetic dipole-dipole interactions breaking the de- ers, New York ͑1994͒. 17 Petrou, D. L. Peterson, S. Venugopalan, R. R. Galazka, A. K. Ramdas, and generacy of the two spin wave branches. These optical mea- S. Rodrı´guez, Phys. Rev. B 27, 3471 ͑1983͒. surements performed three decades ago are seen now to rep- 18 D. L. Peterson, D. U. Barthololmew, A. K. Ramdas, and S. Rodrı´guez, resent the first observation of such splittings, as first Phys. Rev. B 31, 7932 ͑1985͒. 19 K. R. Hoffman, W. Jia, and W. M. Yen, Opt. Lett. 15, 332 ͑1990͒. predicted by Loudon and Pincus in 1963. It would be infor- 20 2 ␪ K. R.;Hoffman, W. Jia, and W. M. Yen, Proceedings of the 12th Interna- mative to perform further experiments to confirm the sin tional Conference on Raman Spectroscopy, J. R. Durig and J. F. Sullivan ␪ dependence of the splitting by varying the angle 0 from 0 to ͑eds.͒, Wiley, New York ͑1990͒, p. 864. ␲/2, i.e., the incident light direction is varied from along to 21 K. R. Hoffman, W. M. Yen, and D. J. Lockwood, Proceedings of the 14th perpendicular to the easy axis, and also the sample shape International Conference on Raman Spectroscopy, N. T. Yu and X. Y. Li ͑eds.͒, Wiley, Chichester ͑1994͒, p. 1078. effect on the splitting. 22 K. R. Hoffman, W. M. Yen, and D. J. Lockwood, Proceedings of the 13th The ability to measure an effective anisotropy field in International Conference on Raman Spectroscopy, W. Kiefer et al. ͑eds.͒, FeF highlights the sensitivity of ROA measurements for de- Wiley, Chichester ͑1992͒,p.766. 2 23 termining small splittings between the magnon branches in K. R. Hoffman, W. M. Yen, D. J. Lockwood, and P. E. Sulewski, Phys. Rev. B 49, 182 ͑1994͒. antiferromagnets. The measurement of such a splitting due to 24 D. J. Lockwood, K. R. Hoffman, and W. M. Yen, J. Lumin. 100,147 ͑ ͒ magnetic dipole-dipole interactions in FeF2 has permitted a 2002 . final resolution of questions concerning the origins of CD in 25 R. Loudon and P. Pincus, Phys. Rev. 132,673͑1963͒. 26 A. B. Harris, Phys. Rev. 143,353͑1966͒. optical measurements completed over thirty years ago. The 27 M. G. Cottam and D. J. Lockwood, Light Scattering in Magnetic Solids, high sensitivity of ROA also permits the study of antiferro- Wiley, New York, ͑1986͒, Chapter 5. magnetic branch splittings for low applied magnetic fields. 28 D. D. Sell, R. L. Greene, and R. M. White, Phys. Rev. 158,149͑1967͒. 29 R. A. Cowley and W. J. L. Buyers, Rev. Mod. Phys. 44,406͑1972͒. The precise experimental work of Professor F. L. Scar- 30 ͑ ͒ J. P. Kotthaus and V. Jaccarino, Phys. Rev. Lett. 28, 1649 ͑1972͒. pace and Dr. Y. H. Russ Wong from thirty years ago is 31 T. Dumelow, R. E. Camley, K. Abraha, and D. R. Tilley, Phys. Rev. B 58, noted with appreciation. We thank Professors R. M. White 897 ͑1998͒. and M. G. Cottam for useful discussions and the National 32 P. L. Richards, J. Appl. Phys. 35, 850 ͑1964͒. Science Foundation for continued support spanning these This article was published in English in the original Russian journal. Repro- three decades. duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 31, NUMBERS 8–9 AUGUST–SEPTEMBER 2005

Magnetoelasticity and domain structure in antiferromagnetic crystals of the iron-group dihalides V. M. Kalita, A. F. Lozenko,* S. M. Ryabchenko, and P. A. Trotsenko

Institute of Physics of the National Academy of Sciences of Ukraine, pr. Nauki 46, Kiev 03028, Ukraine ͑Submitted December 31, 2004; revised May 4, 2005͒ Fiz. Nizk. Temp. 31, 1042–1058 ͑August–September 2005͒ We study the formation of antiferromagnetic magnetoelastic domains in easy-plane antiferromagnets of the iron-group dihalides, in which the infinite degeneracy of the spatial orientation of the antiferromagnetic vector in the basal plane is lifted on account of the spontaneous magnetostriction. In these crystals the domains differ from each other not only by the direction of the antiferromagnetic vector L in them but also by the related directions of the principal axes of the spontaneous magneostrictive strain. The system of antiferromagnetic domains turns out to be identical to the system of elastic domains. It is shown that the processes of magnetization and the induced striction in a magnetic field in the multidomain antiferromagnetic state are interconnected. Data on the field dependence of the induced magnetostriction in the multidomain state of the easy-plane antiferromagnets CoCl2 and NiCl2 are presented and analyzed. It is shown that although the magnetostriction in the cycles of imposition and removal of the magnetic field includes both reversible and irreversible contributions, the reversible being the main one. It is shown that the magnetoelastic-striction domains are responsible for reversibility ͑equilibrium͒ of the multidomain state. The field dependence of the magnetostriction of the uniform and multidomain states and the behavior of the magnetization of the crystals are described. For description of the rearrangement of the multidomain state, the approximation of a continuous distribution of domains with respect to the orientations of their L vector in the easy plane of the crystal in the absence of external magnetic field. It is shown that the matching of the elastic fields of the system of domains and the elastic fields of defects can bring about the formation of a reversible multidomain state, i.e., it can make such a state energetically favorable. The results of the analysis are in satisfactory agreement with the experimental data. © 2005 American Institute of Physics. ͓DOI: 10.1063/1.2008141͔

I. INTRODUCTION properties both in the elastic subsystem of the crystals—the values of the elastic constants in the direction perpendicular Many of the fundamental aspects of the magnetoelastic- ity resulting from the lowering of the symmetry upon the to the layer are an order of magnitude smaller than those magnetic ordering of high-symmetry magnets with multiple within the layer—and in the magnetic subsystem—the ex- spatial degeneracy of the directions of the order parameter change interactions of the magnetic ions in the layer are have not been adequately studied. First and foremost, this much larger than those between ions of adjacent layers. The applies to the formation of magnetic domains in them. This inter-sandwich stacking of different crystals of this group can paper is devoted to an examination of these questions for the be somewhat different. For example, CoCl2 and NiCl2 have 5 ͑ 1 example of layered easy-plane antiferromagnetic crystals of spatial symmetry D3d isomorphic to CdCl2), while CoBr2 3 ͑ 2 iron-group dihalides. has space group D3d isomorphic to CdI2). The majority of the crystals discussed3–5 are easy-plane ͑ ͒ ͑ ͒ II. BASIC PROPERTIES OF LAYERED ANTIFERROMAGNETIC EP two-sublattice antiferromagnets AFMs with hard axis 2ϩ IRON-GROUP DIHALIDES C3 . The Me ions lying in one plane belong to one sublat- tice, and the ions of adjacent planes belong to the other. A. Properties of the crystals Ќ ϭ Ϫ Owing to the EP anisotropy L C3 , L s1 s2 is the antifer- Easy-plane antiferromagnetic crystals of iron-group di- romagnetic vector, and s1 and s2 are the sublattice magneti- ͑ halides MeX2 Me is a metal ion, Co or Ni, and X is a zation vectors. The anisotropy due to the departure of the halogen ion, Cl or Br͒ have a layered crystal structure in vectors s1 and s2 from the EP is much larger than the anisot- which each layer of Me2ϩ metal ions are surrounded by two ropy within the EP, and for a number of these crystals it is layers of halogen ions XϪ, coupled with the Me2ϩ ions by even large compared to the exchange. As to the anisotropy ionic-covalent bonds. The set of such sandwiches, coupled within the EP, according to the neutron scattering datea3 the together by van der Waals bonds, for a crystal. A threefold easiest directions for L in the EP correspond to the three ͑ ͒ symmetry axis C3 passes perpendicular to the layers, and in equivalent C2 axes of one set sixfold spatial degeneracy , the basal plane, which coincides with the plane of the layers, while the three C2 axes of the other set are harder, i.e., there there are two sets of twofold axes C2 , three equivalent axes is some anisotropy. In NiCl2 this anisotropy, although ex- in each. The layeredness leads to anisotropy of the physical tremely weak, is nevertheless manifested in the angular de-

1063-777X/2005/31(8–9)/13/$26.00794 © 2005 American Institute of Physics Low Temp. Phys. 31 (8–9), August–September 2005 Kalita et al. 795 pendence of the low-frequency branch of the antiferromag- entations of L. A spontaneous anisotropy of the magneto- ͑ ͒ 4 netic resonance AFMR , in CoCl2 no such manifestation striction arises in the crystal, with the principal axes of the could be discerned. The assertion that the in-plane anisotropy strain tensor lying in the EP in the directions perpendicular is extremely weak can also be made from the observations of and parallel to L. The in-plane magnetoelastic anisotropy the magnetization, which is actually independent of the mag- due to the magnetostriction lifts the spatial degeneracy for netic field direction in the basal plane.5 Thus the vector L the directions of L, leading to the formation of a gap for fast ϭ ϩ can take practically any direction in the easy plane by over- oscillations of L and M in the plane (M s1 s2). Impor- coming this small anisotropy. tantly, this ‘‘anisotropy’’ for the crystals studied turns out to One should also take into consideration the possible be much greater than the initial in-plane magnetic crystallo- modification of the in-plane magnetic anisotropy by the graphic anisotropy discussed above. fields of defects of the corresponding structure of the AFMs Comparison of the field of this anisotropy ͑magnetoelas- studied. Point defects of a different type ͑vacancies or impu- tic field͒ with the exchange field shows that in order of mag- rity atoms͒ can enhance or weaken the 6-fold anisotropy nitude the magnetoelastic energy amounts to some percent of mentioned and can even cause a change of sign. Disloca- the exchange energy. It is expected that when the values of tions, including screw, and disclinations will also alter the the elastic constants of the crystals are taken into account, anisotropy mentioned in the region around them. Thus de- the values of the anisotropic spontaneous magnetostriction in fects can suppress the weak in-plane anisotropy if its local them will be anomalously large, as was observed in Ref. 11: modulation caused by them exceeds its small initial value. it reaches values of 10Ϫ4 –10Ϫ3. A magnetostriction of simi- The parameter ␪ in the Curie-Weiss law for the tempera- lar magnitude observed in rare-earth magnets is sometimes ture dependence of the paramagnetic susceptibility and also called ‘‘giant.’’ 12 ͑ the temperature of AFM ordering the Ne´el temperature TN) Observation of the LF AFMR in CoCl2 has revealed in these crystals is mainly determined by the intralayer fer- practically no manifestations of angular dependence ͑depen- romagnetic exchange, while the interlayer AFM exchange is dence on the direction of H in the EP͒ of the frequency in considerably weaker.5 either the single-or the multidomain state. The magnetic-field dependence of the LF AFMR can be described on the as- B. Spontaneous lowering of symmetry. AFMR. Domains sumption that the value of the MS gap is independent of field at the transition to the single-domain state. Consequently, in It follows from the neutron diffraction observations that the multidomain state, domains with different directions of the AFM ordering in CoCl2 , CoBr2 , and NiCl2 in the pres- the MS coexist without creating mutual elastic stresses that ence of orientational degeneracy of the directions of the vec- influence the value of the gap of the LF AFMR. tor L in the plane is accompanied by the separation of the The multidomain character is manifested in the magne- 3 crystal into antiferromagnetic domains with different direc- tization of the structures.13 At high magnetic fields in the tions of L in the EP. Within each domain L is uniform. single-domain state the samples behave as Ne´el two- According to the data of Ref. 3, the values of the angles sublattice AFMs with linear dependence of the magnetization between directions of L are predominantly equal to 60°, re- on field. In the multidomain state one observes a nonlinear flecting, it would seem, the orientation of L within each do- trend of these dependences with a noticeable decrease of the main with its easiest axis. However, even the neutron diffrac- magnetic susceptibility for H→0.13 These features of the tion data do not imply that the distribution functions of the magnetization had not been explained before. directions of L of the different domains are completely con- It was shown in Refs. 14–19 that the AFM domains in centrated only near these directions, and the residual stresses the crystals studied have a magnetoelastic nature. In those are not represented. In a magnetic field H lying in the EP a papers, and also in Refs. 20 and 21, it was established that it ͑ ͒ gradual transition to a uniform single-domain state with is the elastic aspect of the multidomain state that is respon- Ќ L H occurs. It is reached in a field H ff , much less than the sible for its essential reversibility. Below we give a brief sublattice collapse fields. When the field is removed the mul- review of the papers mentioned and carry out additional tidomain state is is restored, though not completely. The vol- proofs of the magnetoelastic nature of the AFM domains in ume of the domains with L normal to the imposed field is layered iron-group dihalides. greater than the volumes of domains with other orientations of L. C. Causes of the formation of domains in layered AFMs AFMR studies of these crystals4,6–8 have shown that the low-frequency ͑LF͒ branch of the AFMR in them has a finite The formation and stability of AFM domains have been gap, the value of which is much larger than would be ex- discussed in many papers.22–25 The multidomain character of pected from the small in-plane anisotropy. Besides the data4 the AFM is not due, as in the case of ferromagnets, to mag- ͑ ͒26,27 for NiCl2 , where the angular dependence of the frequency of netostatic fields of a dipolar nature The cost in ex- the LF AFMR was found to contain a weak contribution due change energy in the domain walls should make this state to the angle between H in the EP and the C2 axes, no mani- energetically unfavorable. Therefore the AFM domains are festations of this anisotropy in the corresponding angular de- most often metastable, thermodynamically nonequilibrium, pendences of the frequency of the LF AFMR has been dis- and are not restored after the crystal has been brought to the cerned in other crystals of this group. Such behavior of the single-domain state by an external magnetic field. LF AFMR is typical9,10 for the spontaneous lowering of the The AFM domains arising at orientation phase transi- symmetry on account of the magnetoelastic ͑ME͒ interac- tions in uniaxial AFMs was studied in Refs. 28 and 29. In a tion, in an AFM with infinite spatial degeneracy of the ori- magnetic field the transition from the state with easy-axis 796 Low Temp. Phys. 31 (8–9), August–September 2005 Kalita et al.

AFM orientation of the sublattice spins to the state with and strains, then the energy cost of the structural defect can spins tilted toward the field and oriented almost perpendicu- be reduced by having the signs of their elastic fields match lar to the easy axis, both these states are observed simulta- those of the stresses caused by the junction of the domains. neously, i.e., an intermediate AFM state is realized.29 This In that case the total energy of the crystal can even decrease multidomain state formed in the magnetization of an AFM is provided the cost of the exchange energy in the domain walls described with allowance for the magnetostatic energy. is less than the benefit in the elastic energy. Various mechanisms have been proposed22 to explain the The possibility of stabilization of the multidomain mag- formation of an equilibrium multidomain state in an AFM, netoelastic state owing to the benefit in the surface energy of but apparently none of them applies to all AFMs. The most a sample without any changes of its outward form was con- universal, at first glance, is the entropy mechanism,30 accord- sidered in Refs. 20 and 21. This mechanism may be appli- ing to which the energy cost due to the exchange interaction cable to AFM samples of very small size. Therefore, we shall in the wall is compensated by a lowering of the free energy not analyze it in detail here. T⌬S owing to the increase in entropy in the multidomain In the analysis of the induced MS of the crystal as a state (T is the temperature and ⌬S is the entropy increase͒. whole, it is necessary to distinguish two regions: the low This contribution vanishes at T→0, although in a number of field region, where the observable, aggregate MS is mainly cases it does permit an explanation of the formation of an due to rearrangement of the multidomain state, and the high equilibrium multidomain state at high temperatures, close to field region, where in the single-domain state the MS that the Ne´el point.22 appears is due solely to the rotation of the sublattice magne- In many AFMs the equilibrium ͑or almost equilibrium͒ tizations toward the magnetic field and their ultimate col- ϭ multidomain AFM state exists in the temperature interval lapse at a field H ff 2HE . For some of the crystals consid- → ered, the values of 2HE are small and accessible in from TN to T 0. It is most often linked with the influence of defects. Domains of this kind have been observed by dif- experiments. ferent methods, e.g., they have been visualized optically.31,32 AFM domains are formed upon twinning of the crystal, III. FIELD AND TEMPERATURE DEPENDENCE OF THE when the domain structure is a combination of antiferromag- INDUCED MAGNETOSTRICTION AND THE MAGNETIZATION OF THE EASY-PLANE AFMs netic and structural domains. The formation of a multido- main state is possible when the order of succession of the A. Dependence of the induced MS in crossed magnetic sublattices is disrupted, e.g., in the presence of edge disloca- fields tions, when the defect is a half plane of atoms with magnetic The magnetostriction of CoCl2 , CoBr2 , and NiCl2 has moments belonging to one of the magnetic sublattices. In been studied by means of a capacitive dilatometer.36 The Refs. 33 and 34 it was shown that a screw dislocation in an crystals were grown by the slow cooling of the melt in a AFM leads to the formation of a spin disclination, which, sealed quartz ampoule. Prior to sealing, the ampoule contain- when the anisotropy is taken into account, leads to the for- ing the ‘‘chemically pure’’ ͑grade ChDA͒ powder was mation of AFM domains. Defects can also stabilize the mul- pumped down at a temperature close to the point for tidomain state of a kinetic nature that arises in the course of dessication. The crystals are easily delaminated along the the AFM ordering, when domains form because there are plane perpendicular to the C axis; the directions of the C 35 3 2 many centers for the formation of the AFM state at TN . axes in the basal plane of the crystals was not monitored. The simplest case of the influence of defects on the for- Samples in the form of 5ϫ5ϫ1 mm rectangular slabs were mation of a multidomain state involves the so-called ‘‘met- cut from flat wafers obtained by cleaving. The variation of allurgical defects,’’ which, distorting the lattice, locally alter the length of one of the sides of the slab in magnetic field the direction of the anisotropy fields and the local orienta- was measured. tions of the vector L. We shall not consider that case but To obtain data on the reversible and irreversible ͑after a restrict the investigation to multidomain states formed as a cycle of imposition and removal of the magnetic field͒ parts result of the ME interaction of the magnetic subsystem with of the induced MS, we used a system of crossed supercon- the elastic fields of defects in rather perfect crystals, which ducting magnets. The sample was mounted so that the fields has been fully studied. of both coils lay in plane of the sample, one in the direction It is not obvious why it is optimal for a system of AFM along the side of the slab measured by the dilatometer, the domains to form in which each of the domains suffers spon- other perpendicular to it. The maximum magnetic fields taneous ms that is anisotropic in the plane, with different equaled 65 and 13 kOe. This set of solenoids made it pos- directions of the principal axes of the strictive strains. If the sible to investigate the MS while rearranging the domain directions of the striction are different in different directions, structure and also to determine the part that lies ‘‘hidden’’ in then stresses should arise on the walls of adjacent domains to the spontaneous MS of the sample before the magnetic field counteract the differently directed spontaneous MS in them. is applied. The sample temperature was set at values ranging And although the directions of the principal axes of the strain from 4.2 to 70 K. tensors can be matched for two adjacent domains, but it can- Figure 1 shows the curves of the relative elongation ␧ ϭ⌬ ͑ not be done where more than two domains come together at (H) l /l for CoCl2 and NiCl2 the curve for CoBr2 is a point. Such a point is surrounded by elastic stresses that analogous to that for CoCl2) in an imposition-removal cycle add to the energy cost to the crystal. However, if the place of the crossed magnetic fields at a temperature of 4.2 K. Here where three or more domains come together coincides with a l is the length of the sample and ⌬l is its change in the structural defect, which is also a source of elastic stresses field. Curves 1 and 2 in Fig. 1 correspond, respectively, to Low Temp. Phys. 31 (8–9), August–September 2005 Kalita et al. 797

spontaneously strained, as is reflected in the observed gap in the spectrum of the LF branch. This means that the domains retain the spontaneous MS without creating substantial stresses on each other, although the spontaneous MS of the whole crystal in the multidomain state is close to zero. It is also seen in Fig. 1 that the signs of the contribution from the induced MS in the single-domain state are opposite to the signs of the initial spontaneous MS for both crystals and both orientations of the magnetic field. Therefore, the

␧ʈ(H) curve has a maximum and the ␧Ќ(H) curve a mini- mum near the field at which the single-domain state is estab- ␧ ␧ lished in the sample. In NiCl2 the ʈ(H) and Ќ(H) curves are practically antisymmetric at the transition from the mul- tidomain to the single-domain state. In CoCl2 , as in CoBr2 , this antisymmetry is noticeably distorted. For CoCl2 and CoBr2 the field of the maximum Hmax is constant larger than ␧ Ͼ͉␧ ͉ the field of the minimum Hmin , and ʈ(Hmax) Ќ(Hmin) . These results were explained in Ref. 14 as follows. The strains of different domains add together, forming the strain of the crystal in the direction of measurement. Since for the spontaneous MS the strains of the crystal along L and per- pendicular to it must be equal in magnitude and opposite in sign, the total spontaneous MS anisotropic in the basal plane for the crystal in the multidomain state averages to zero for HϷ0 if it is assumed that the domains are uniformly distrib- uted over equivalent directions of L. After the induced tran- sition of the crystal to the single-domain state for LЌH, the dependence of the induced MS on the field is determined solely by the canting of the sublattice magnetizations toward H. For the data shown in Fig. 1, this stage begins at H у ϭ ͑ Ͼ 10 kOe and ends at H H ff the data for H H ff are not ͒ ͑ ͒ ͑ ͒ shown in Fig. 1 . The differences between the curves during FIG. 1. Relative elongation of CoCl2 a and NiCl2 b during the imposi- tion and removal of a magnetic field lying in the easy-plane perpendicular to imposition and removal of the field in Fig. 1, corresponding the direction of the measured side of the crystal ͑curves 1 and 2͒ and the to the remanent spontaneous MS of the crystal as a whole, subsequent imposition and removal of a field along the measured direction reflect the partial irreversibility of the rearrangement of the ͑curves 3 and 4͒. domain structure in this cycle.

the measurements of ␧Ќ(H) during the imposition and re- moval of the field H perpendicular to the measured side of B. Field dependence of the MS at low fields ␧ the crystal, and curves 3 and 4 to the measurements of ʈ(H) Figure 2 shows the dependence of the MS of the multi- H during the subsequent imposition and removal of a field domain state of CoCl2 and NiCl2 on the square of the mag- along that direction. It is seen that the striction is anisotropic netic field strength during the removal of the field at low with respect to the field direction and that the curves of the fields. For HϽ2.5 kOe the curves can be described by the MS thus obtained have the form of hysteresis loops with a expression remanent MS, the sign of which is determined by the direc- ␧ ϭ␧ ϩ␣ 2 ͑ ͒ tion of the removed field. ʈ,Ќ rʈ,Ќ ʈ,ЌHʈ,Ќ, 1 The curves shown in Fig. 1 correspond to the variation ␧ of the MS during the rearrangement of the multidomain state where rʈ,Ќ is the remanent striction in the plane for mea- and a transition to a uniform state under the influence of surements along ͑ʈ͒ and transverse to ͑Ќ͒ the magnetic field, ͉␧ Х͉␧ ͉ ␧ ХϪ␧ ␣ ХϪ␣ magnetic field. In the AFM state an anisotropic spontaneous rʈ rЌ , but rʈ rЌ ; ʈ Ќ are empirical param- MS is realized for these samples. However, in the multido- eters. The ␧(H) curve for the first imposition of the field is ͑ ͒ ␧ ϭ main state the differently directed spontaneous strains of the also described by Eq. 1 , but with r 0. The curves drawn individual domains average out almost to zero over the crys- according to Eq. ͑1͒ are shown by the solid lines in Fig. 2. tal as a whole. At high fields the spontaneous MS of the The quadratic character of the field dependence of the MS crystal is restored in a transition to the single-domain state. shows that the rearrangement of the multidomain state, i.e., After the field is removed, the multidomain state is largely the increase of the volume of domains with the favorable restored, and the MS decreases to a small remanent striction. orientations of the sublattice spins ͑relative to the direction This remanent MS is also absent prior to the first imposition of H͒ and the decrease of the volume of domains with an of the field. In the multidomain state, as we have said, the unfavorable orientation occurs mainly through the motion of AFMR data11 indicate that the domains individually are domain walls. 798 Low Temp. Phys. 31 (8–9), August–September 2005 Kalita et al.

͑ ͒ FIG. 3. Field dependence of the magnetostriction of CoCl ͑a͒ and NiCl ͑b͒ FIG. 2. Dependence of the induced magnetostriction of CoCl2 a and NiCl2 2 2 ͑b͒ versus the square of the magnetic field strength during removal of the crystals; the measurements were made during removal of a field lying along field. the direction of measurement of their relative elongation.

ϭ129 kOe,38 and it was not reached in our experiments; we C. Behavior of the MS in the uniform state therefore used the published value for H ff in determining the The rearrangement of the multidomain state is com- value of the parameter ␰ in Eq. ͑2͒. pleted in a magnetic field higher than the characteristic value of the field for bringing the sample to the single-domain state ͑in our case Ϸ10 kOe), by a transition to a single-domain C. Temperature dependence of the MS state with LЌH. Figure 3 shows the relative elongation of Figure 4 shows the field dependence of the relative strain the crystals as a function of the magnetic field strength for a of CoCl2 in the direction along the magnetic field, which was field Hр65 kOe parallel to the side being measured. The oriented in the EP, for temperatures from 4.2 to 24 K. The MS of the single-domain state can be described by the influence of temperature is manifested in the value of the MS 17,37 expression strain: it decreases with increasing T. The field of the tran- sition to the single-domain state ͑the region of the maximum ␧ ϭ␧͑S͒ Ϫ␰ 2 ͑ ͒ on the curves͒ also decreases with increasing T, as does the ʈ͑H͒ ͑1 ͑H/H ff͒ ͒, 2 field H ff . where ␧(S) is the striction that would be realized in the crys- tal at H→0 if the sample were in a uniform state, and ␰ is an empirical parameter. The value of ␧(S) should be regarded as spontaneous anisotropic MS, which can be obtained from the assumption continuation of the ␧ʈ(H) curve into the region → ␰Ͻ ␰Ͼ H 0. For NiCl2 and CoCl2 the cases 1 and 1 are realized, respectively. As was shown in Refs. 14, 17, and 37, this is due to the different relative contribution to the MS from the inter- and intra-sublattice magnetoelastic interac- tions. In CoCl2 an orientational phase transition is reached in ϭ the given field interval, at H H ff , when the spins of both sublattices become parallel along H. After that, the field de- pendence of the MS changes in connection with the parapro- cess. According to Ref. 1 and our data, H ϭ32 kOe in ff FIG. 4. Field dependence of the magnetostriction of CoCl2 for the first CoCl2 . In NiCl2 , according to the AFMR data, H ff imposition of the magnetic field at different temperatures. Low Temp. Phys. 31 (8–9), August–September 2005 Kalita et al. 799

␧ FIG. 5. Magnetostriction ˜ ʈ of the uniform state of CoCl2 , normalized to its spontaneous value, versus the square of the magnetic field in the plane, normalized to the collapse field, for different temperatures.

FIG. 6. Magnetostriction of CoCl2 versus the square of the magnetic field during its removal at different temperatures in the region of the multidomain state.

Figure 5 shows the MS in CoCl2 , normalized to the value of the spontaneous anisotropic MS, i.e., ˜␧(T,H)ʈ (S) ϭ␧(T,H)ʈ /␧ (T), for different temperatures as a function of the square of the magnetic field ͑for field directed parallel E. Field dependence of the magnetization in the to the side being measured͒, normalized to the square of the multidomain state ˜ 2 collapse field at the corresponding temperature: Hʈ The magnetization was measured using an LDJ-9500 vi- ϭ 2 2 Hʈ /H ff(T). It is seen that, when normalized in this way, brating . NiCl2 samples with dimensions of 5 the curves for different temperatures in the single-domain ϫ4ϫ0.2 at the were studied. The applied field did not ex- state all coincide. This means that in the uniform region the ceed 12 kOe, which was sufficient for the transformation of ͑ ͒ ␰ MS obeys relation 2 with a parameter this is independent NiCl from a multidomain to a single-domain state.17 The ␧ 2 of T. Then the variation of ʈ(H) with temperature should samples used in the magnetostriction and magnetization ␧(S) be due to the influence of temperature on (T) and H ff(T) measurements were grown at different times. Because NiCl2 only. This is easy to understand, since the induced MS of the crystals are hygroscopic, the samples were stored in a dessi- single-domain state is due solely to the canting of the sub- cated medium. In view of the possible differences in the lattice spins. growth and storage conditions, it cannot be said for certain In the region where the rearrangement of the multido- that the samples were identical from the standpoint of the main state occurs, the dependence of MS on magnetic field possible number of defects. Measurements made in the has a different character. Figure 6 shows the dependence of cycles of field imposition and removal have shown the ab- the induced MS of CoCl2 on the square of the magnetic field sence of remanent magnetization. Consequently, our AFM ␧ ϭ strength during the first imposition of the field ( r 0). They crystals, like those in Ref. 39, did not have ferromagnetic ͑ ͒ ␣ obey expression 1 but with a coefficient that depends on inclusions. T. Here the trend of the MS can be represented in the form Figure 7 shows the magnetization curve of a NiCl2 Ќ single crystal during the imposition of a field H C3 .Itis 2 seen that the magnetization depends nonlinearly on H. The Hʈ Ќ ␧ ϭ␧͑S͒͑ ͒ , ͑ ͒ dashed line in Fig. 7 shows the straight line expected for the ʈ,Ќ T 2 , 3 Hd dependence of the magnetization on H in a single-domain EP AFM with no in-plane anisotropy.40 The experimental depen- dence m(H), on the contrary, has a characteristic ‘‘sag’’ be- where Hd is a parameter having dimensions of magnetic field low this straight line in the region where the rearrangement strength. For agreement with experiment the value of Hd of the multidomain structure occurs. Initially the difference should be chosen the same for all temperatures and equal to between the calculated and experimental m(H) curves grows 5.7Ϯ0.3 kOe. If an entropy mechanism of the multidomain and the curves diverge from each other, but then ͑for H state or a mechanism involving the influence of the surface Ͼ5 kOe) they draw closer together and practically merge at elastic energy is realized in the crystal, then the parameter HϷ10 kOe. The m(H) curve of the multidomain state for → Hd would be temperature-dependent. H 0 is well approximated by the expression 800 Low Temp. Phys. 31 (8–9), August–September 2005 Kalita et al.

ϭ account that for H He the state of the crystal is uniform ͑ ϭ almost uniform for our calculation using He 10 kOe), ϭ ϭ ⌬ ϭ ϭ Ϫ ϭ ϭ e3(H He) 0, and e3 e3(H 0) e3(H He) e3(H ϭ ϭ ͑ ͒ ⌬ Ͼ 0) e3 . According to Eq. 5 , A 0, and the formation of domain walls is unfavorable for the exchange interactions, Ͼ and therefore e3 0. It follows from the arguments presented ⌬ Ͻ ͉⌬ ͉Ͼ ⌬ ϭ ϭ above that e4 0, and e4 e3 . Since e4 e4(H 0) Ϫ ϭ Ͻ e4(H3), we obtain e4(H 0) e4(H3). Such a feature of the energy source e4 of the multidomain character is a nec- essary condition for the formation of an equilibrium magne- toelastic multidomain state. The energy benefit of the multidomain state at Hϭ0 can be determined from Eq. ͑5͒. Its value is equal to the square of the figure enclosed between the straight line of uniform mag-

FIG. 7. Magnetization of NiCl2 as a function of the magnetic field lying in netization and the m(H) curve in Fig. 7. According to our the EP. The dots are experimental ͑result of the first imposition of the field͒, calculations, the ratio of this energy benefit to the exchange the dashed line is the expected dependence for a single-domain crystal with energy of the inter-sublattice interactions in NiCl equals reversible domain structure. 2 5.7ϫ10Ϫ4. Thus one can conclude that in the presence of defects causing local elastic stresses, the multidomain magnetoelas- H2 ϭ␹ ͩ ϩ ͪ ͑ ͒ tic state is energetically favorable and, consequently, should m dH 1 2 , 4 Hm be reversible, i.e., in a state of thermodynamic equilibrium. That state is realized in the given AFM. where ␹ is the magnetic susceptibility of the multidomain d Before turning to an analysis of the multidomain state, state for H→0, and H is an empirical parameter. A fitting to m let us analyze the behavior of the MS of a uniform state. experiment gives ␹ ϭ0.44 e.m.u./(g kOe), H ϭ4.3 d • m Moreover, at small values of H the contribution from the Ϯ0.6 kOe. ff canting of the sublattice spins in the rearrangement of the By integrating m(H) over H, one can determine the multidomain state cannot be neglected. work done by the magnetic field in magnetizing the crystal. The work done by the field in taking a unit volume of an 18 AFM from the multidomain to the uniform state will be the IV. PHENOMENOLOGICAL DESCRIPTION OF THE MS OF difference between the work of magnetization of a uniform THE UNIFORM STATE and a multidomain segment. For the numerical integration we use the experimental dependence shown in Fig. 7. We In the phenomenological description of the induced MS compute of the uniform state we shall assume that its value depends only on the orientation of the sublattice spins.19,37 We shall 1 He take into account only the magnetoelastic interactions which ⌬Aϭ ͫ ␹ H2Ϫ ͵ m͑H͒dHͬ, ͑5͒ e e 2␳ 0 are anisotropic in the easy plane. When the symmetry of these crystals is taken into account, the sum of the magneto- where ␳ is the density of the crystal, ␹ is the magnetic e elastic and elastic contributions to the energy can be written susceptibility of the uniform state, which, line the magneti- in the form zation m(H), is normalized per unit mass of the sample, and H is the field at which the uniform state is attained ͑we used e ϭ ␥ ͑ ͒͑ Ϫ ͒͑ Ϫ ͒ ϭ E ͚ ␣␤ T n␣xn␤x n␣yn␤y Uxx Uyy the value He 10 kOe). ␣у␤ The energy E per unit volume of a multidomain antifer- romagnet in a magnetic field can be written as a sum E ϩ ␭ ͑ ͒͑ ϩ ͒ ϩ ␦ ͑ ͒ ϭ ϩ ϩ ϩ ͚ ␣␤ T n␣xn␤y n␤xn␣y Uxy ͚ ␣␤ T e1 e2 e3 e4 , where e1 is the AFM exchange energy, ␣у␤ ␣у␤ e2 is the Zeeman contribution, e3 is the domain-wall energy, 1 and e is the energy of matching of the local elastic fields of 2 4 ϫ͑n␣ n␤ ϩn␣ n␤ ͒͑U ϩU ͒ϩ C ͑U defects of the crystal and the elastic strains due to spontane- x x y y xx yy 2 11 xx ous MS domains. The local matching of the elastic strains of ϩU2 ͒ϩC U U ϩ͑C ϪC ͒U2 , ͑6͒ the spontaneous MS of the domains with defects promote the yy 12 xx yy 11 12 xy formation of multiple domains, and therefore e4 can be where n␣x ,n␣y and n␤x ,n␤y are the direction cosines of the called the energy ‘‘source’’ of the multidomain character. An sublattice magnetization vectors s␣ and s␤ , the x and y axes equilibrium multidomain state corresponds to a minimum of lie in the EP, ␥, ␭, and ␦ are temperature-dependence param- ␣ ␤ the energy E. In that case the energy increments e3 and e4 eters of the magnetoelastic interactions, the indices , ϭ upon the formation of the multidomain state must have op- 1,2 enumerate the sublattices, and the Uij are components posite signs. of the strain tensor. The terms containing the constants C11 Let us consider the difference of the energies E of a and C12 describe the elastic contribution to the energy of the crystal in the states prior to application of the magnetic field crystal. In Eq. ͑6͒ the terms with ␣ϭ␤ and ␣␤ pertain to ⌬ ϭ⌬ ϩ⌬ ϩ⌬ ϩ⌬ and in the field He , i.e., E e1 e2 e3 e4 . the intra- and inter-sublattice magnetoelastic interactions, re- ⌬ ϭϪ ϭ ϩ⌬ Hence A ͓e3(H 0) e4͔, where we have taken into spectively. Low Temp. Phys. 31 (8–9), August–September 2005 Kalita et al. 801

In the uniform state and in a magnetic field oriented in the EP, the spins of both sublattices lie in that plane and are tilted in the same way with respect to H. Then the direction ͑ ͒ cosines in 6 are proportional to H/H ff . The dependence of the strain on H is easily found by minimizing ͑6͒ with re- spect to Uij . The expressions obtained for Uyy , when H lies along y or x, have the form14,17,37 2␦ ␥ ͑ ʈ ͒ϭϪ 11 ϯ 12 Uyy H y,x ϩ Ϫ C11 C12 C11 C12 ␦ 2␥ H2 ϩͩ 12 Ϯ 11 ͪͩ Ϫ ͪ ͑ ͒ ϩ Ϫ 1 2 2 , 7 C11 C12 C11 C12 H ff where the upper sign corresponds to Hʈy, when the field is parallel to the strain being measured, Uyy , and the lower FIG. 8. Scheme of the orientations of s1 and s2 and of L and M of a domain sign to Hʈx, when the field is perpendicular to that strain. relative to the magnetic field H. In the MS measurements the length of the side of the crystal at Hϭ0 is taken as the reference point, and the iso- tropic contributions to the strain in ͑7͒ are not reflected in the V. DESCRIPTION OF THE REARRANGEMENT OF THE measurements. Therefore, expressions for the MS of the rela- MULTIDOMAIN MAGNETOELASTIC STATE OF AN AFM tive elongation of the sides of the crystal in the easy plane of an EP AFM in the directions along and transverse to the A. System of magnetoelastic domains in an AFM magnetic field take the form It follows from the director presented that for a qualita- 2 tive discussion of the multidomain nature of these crystals Hʈ Ќ ͑S͒ , ␧ ʈ ϭϮ␧ ͩ Ϫ ͑ ϩ␩ʈ ͒ ͪ ͑ ͒ the distribution over orientations of the L vectors of the d ,Ќ 1 2 1 ,Ќ 2 . 8 H ff AFM domains in the EP can be treated as continuous. Gen- erally speaking, because of the influence of the small in- Here the MS of the uniform state at Hϭ0 should satisfy the ␧ ϭ ϭ␧(S) ␧ ϭ ϭϪ␧(S) plane anisotropy it should be anisotropic, but taking that an- relations dʈ(Hʈ 0) and dЌ(HЌ 0) , where (S) isotropy into account, in view of its smallness and the ␧ ϭ(2␥ Ϫ␥ )/(C ϪC ). The values of ␩ʈ Ќ in ͑8͒ are 11 12 11 12 , presence of three equivalent ‘‘easiest’’ axes in the EP, it does expressed in terms of the parameters of the magnetoelastic not affect the magnetic-field dependence of the MS in the interactions: multidomain state.17 We shall therefore disregard the in- ␩ ϭ␥ ␥ Ϫ␥ ͒ϩ␦ Ϫ ͒ ␥ Ϫ␥ ͒ ʈ 12 /͑2 11 12 12͑C11 C12 /͑2 11 12 plane anisotropy to avoid complicating the calculations. We shall also adopt an approximation of the probability density ϫ͑ ϩ ͒ C11 C12 , distribution of the domains over orientations of L by assum- ing that prior to the first application of the field the domains ␩Ќϭ␥ /͑2␥ Ϫ␥ ͒Ϫ␦ ͑C ϪC ͒/͑2␥ Ϫ␥ ͒ 12 11 12 12 11 12 11 12 ͑the L vectors in them͒ are distributed equiprobably over all ϫ ϩ ͒ ͑C11 C12 . directions in the EP. The orientation of the L vectors of the domains in the EP ͑ ͒ The dependence 8 is in good agreement with the empirical will be specified by an angle ␸ between the magnetic mo- ͑ ͒ relation 2 . ment vector M of the domain and the vector H ͑see Fig. 8͒, ␦  ␥  For 12 0 and 12 0 the derivative the angle ␸ being defined in the EP. For H→0 the angle ␸ -should be taken to be the angle between the direction corre ͒ 2 ͑ץ ␧ ͑ 2͒ץ2͒Ϫ ͑ץ ␧ ͑ 2͒ץ dʈ H / Hʈ dЌ H / HЌ , sponding to limH→0M and the direction of H. The density of ␩ ␩ 16,19 ␸ since ʈ Ќ , as it characteristic for CoCl2 . In that the domain distribution p( ) we define as the ratio of the ␰ϭ ϩ␩ Ͼ ␸ compound 2(1 ʈ,Ќ) 1, and the intra-sublattice mag- volume of domains with orientation to the volume of the ␲ ͐␲/2 ␸ ␸ϭ netoelastic interactions are dominant. In NiCl2 the slopes of crystal. It is normalized, i.e., 1/ Ϫ␲/2p( )d 1. ␧ 2 ␧ 2 the dʈ(H ) and dЌ(H ) are, to within the error limits, As we have said, the MS of a uniform easy-plane AFM equal in magnitude and opposite in sign. Here ␰ϭ2(1 is anisotropy in the easy plane. It is equal to ␧(S) in the ϩ␩ Ͻ Ϫ␧(S) ʈ,Ќ) 1, which corresponds to dominance of the inter- direction perpendicular to L and to in the direction sublattice magnetoelastic interactions which are isotropic in along L. Using p(␸), we write an expression for the MS of the EP.17 This difference of the magnetoelastic interactions in the crystal as a whole in the multidomain state: these crystals is apparently due to the fact that in NiCl2 the 1 ␲/2 orbital moment of the ions is almost completely frozen, and ␧ϭ ͵ ͓␧͑S͒͑ ͒ 2 ␸ϩ␧͑S͒͑ ͒ 2 ␸͔ ͑␸͒ ␸ ␲ ʈ H cos Ќ H sin p d . the single-ion anisotropy is much less than the intra- Ϫ␲/2 41 ͑20͒ sublattice exchange. In CoCl2 , on the contrary, total freez- ing of the orbital moment of the ions does not occur, and For a domain with ␸0 the magnetic field moment M is ͉ ͉ϭ␹ ␸ therefore the single-ion anisotropy in it is comparable to the not parallel to H and is equal to M eH cos . Its projec- 41–44 ϭ␹ 2 ␸ intra-sublattice exchange, and that leads to dominance tion on H is equal to M H eH cos . Therefore the mag- of the intra-sublattice magnetoelastic interactions. netization of the crystal as a whole will be 802 Low Temp. Phys. 31 (8–9), August–September 2005 Kalita et al.

1 ␲/2 ϭ ͵ ␹ 2 ␸ ͑␸͒ ␸ ͑ ͒ m eH cos p d . 21 ␲ Ϫ␲/2

For NiCl2 the field H ff is much greater than the width of the region of fields in which the rearrangement of the multi- (S) domain state occurs. Therefore we can assume ␧ʈ (H) Ϸ ␧(S) Ϸ const, ʈ (H) const. The MS in NiCl2 is almost com- (S) pletely antisymmetric ͑see Fig. 1͒. Therefore ␧ʈ (H)Ϸ (S) (S) Ϫ␧Ќ (H)ϭ␧ . Then the relation between the mean mag- netization and the MS in the multidomain state will have the form18

1 ␧ʈ͑H͒ m͑H͒ϭ ␹ Hͩ 1ϩ ͪ . ͑22͒ 2 e ␧͑S͒ If the experimental values18 of the relative elongation of the crystal along H in the region of rearrangement of the multi-

domain state, ␧ʈ(H), are substituted into Eq. ͑22͒, we obtain a curve for m(H) that is close to the experimental curve for the magnetization of the crystal, presented in Fig. 5. One can also carry out the inverse procedure, substituting m(H) into Eq. ͑22͒. Then one can obtain from the magnetization data the expected behavior of the relative elongation of the crys- tal. The curve thus obtained for the striction also turns out18 to agree well with the experimental dependence. Thus the magnetic-field dependences of the magnetiza- tion and the MS in the multidomain AFM state are interre- ␧ ␧ lated and reflect the process of the rearrangement of that state FIG. 9. Field dependence of the reversible rev and irreversible unr compo- ͑ ͒ ͑ ͒ under the influence of the field. This is a consequence of the nents of the magnetostriction of CoCl2 a and NiCl2 b . magnetoelastic nature of the AFM domains in these crystals. At low magnetic fields the domain distribution that gives satisfaction agreement with the experimental data on the pertains to the irreversible component of the rearrangement field dependences of the MS and magnetization has the form of the multidomain state, while the second is due to the re- 1 ␧ H2 1 versible component. The MS of the crystal as a whole can be ͑␸͒ϭ ͫ ϩ ͩ r ϩ ͪͩ 2 ␸Ϫ ͪͬ ͑ ͒ p ␲ 1 4 ␧͑S͒ 2 cos . 23 written as the sum Hd 2 ␧͑ ͒ϭ␧ ͑ ͒ϩ␧ ͑ ͒ ͑ ͒ During the removal of the magnetic field or its repeated im- H rev H unr H , 25 position in the same direction the dependence of the magne- ␧ where rev(H) is the reversible contribution to the resultant tization at low fields will have the form ␧ striction, and unr(H) is the irreversible component; both ␧ 2 1 r H components depend on the magnetic field and the stage of its mϭ ␹ Hͩ 1ϩ ϩ ͪ . ͑24͒ 2 e ␧͑S͒ H2 imposition-removal cycle, although, starting at a certain d value of the applied field, ␧ (H) should apparently cease to ␧ → unr If the remanent striction r 0, then the magnetic suscepti- depend on it. We assume that in the removal of the field the ␹ bility d at the beginning of the rearrangement of the domain irreversible part of the MS in the whole field interval in structure, for H starting from zero, should be smaller by half which the AFM state exists, remains constant and equal to than the susceptibility of the uniform state, ␹ ϭ␹ /2. The ␧ ϭ␧ ϭ d e the remanent MS: unr(H) r const. experimental value of ␹ is close to ␹ /2, although it is ␧ ␧ d e The rev(H) and unr(H) curves obtained for CoCl2 and slightly larger, reflecting the existing irreversible contribu- NiCl2 on the basis of these ideas from the experimental data tion to the structure of the multidomain state. Furthermore, ͑Fig. 1͒ are shown in Fig. 9. The reversible component does ϭ one would expect that Hd Hm . The difference of the values not have hysteresis in the cycles of imposition and removal of the parameters H and H is 15%. As we have said, the ␧ d m of the crossed magnetic fields. For CoCl2 the rev(H) curve crystals used in our investigation of the MS and magnetiza- becomes more non-antisymmetric in comparison with the tion were grown at different times, and their defect state, curves obtained directly from experiment, whereas for NiCl2 which determines the values of H and H , could have been ␧ d m the rev(H) curve remains antisymmetric. The irreversible different. ␧ component unr(H) is hysteretic for both crystals. The closed ␧ ͑ ͒ part of the loop for unr(H) in Fig. 9 in the first quadrant was constructed on the basis of the properties of antisymme- B. Reversible and irreversible components of the try of the MS at low fields, since the data for that part is rearrangement of the multidomain state missing in Fig. 1. The distribution of domains over orientations of their L We approach the magnetization in an analogous way, ͑ ͒ ␧ vectors 23 contains two terms. The first is due to r and writing it as the sum Low Temp. Phys. 31 (8–9), August–September 2005 Kalita et al. 803

͒ϭ ͒ϩ ͒ ͑ ͒ m͑H mrev͑H munr͑H , 26 where mrev(H) and munr(H) are the reversible and irrevers- ible contributions to the resonant magnetization, respectively. We define the irreversible fraction of the MS as the ratio ␦MSϭ␧ ␧(S) ␦(S)ϭ unr r / . For our NiCl2 samples unr 0.16. We shall assume that an equal fraction of the sample volume consists of domains that contribute to the irreversible MS. This frac- tion of the domains during the removal or the repeated im- position of the field in the same direction will be magnetized ␹ ͑ Ќ with the susceptibility e of the uniform state with L H). Comparing the experimental value of the magnetic suscepti- ␹ → bility d of the multidomain state for H 0 and its expected ␹ FIG. 10. Field dependence of the model magnetostriction of the uniform value e/2, we find the fraction of the irreversible contribu- ␧ ʈ Ќ and multidomain ␧ʈ Ќ states of CoCl . The light solid curves show the tion to the magnetization, ␦m ϭ2␹ /␹ Ϫ1. The value ob- d , , 2 unr d e dependence of ␧ ʈ Ќ , the heavy solid curves ␧ʈ Ќ , and the points are ex- ␦m Ϸ ␦MS d , , tained for NiCl2 , unr 0.14 turned out to be close to unr . perimental. Let us determine the field dependence of the reversible contribution to the magnetization of NiCl2 . The relative vol- ␧ ϭ Ϫ␦ where ume of domains that contribution to it is rev 1 unr .A plot of m (H) for NiCl is given in Fig. 7 (m was cal- ϭ &͒ rev 2 rev H0 Hd /͑2 and culated per unit mass of the sample͒. The magnetic suscep- → ␹ ␲/2 H2 tibility of this part of the domains for H 0is e/2. The ͑ ͒ϭ ͵ ͩ 2 ␸ͪ ␸ I H/H0 exp 2 cos d . value of Hm determined from mrev(H) better satisfies the Ϫ␲/2 2H0 condition H ϭH . d m We see that expression ͑28͒ leads to a ␦-function-like The increase of the volume of domains with the favor- concentration of the distribution function around ␸ϭ0 as the able orientation in magnetic field and the decrease of the field increases for HϾH . volume of the other domains should be regarded first as the d result of the motion of the domain wall in the field. The motion occurs on account of the stresses acting on the walls C. Rearrangement of the reversible component of the on account of the magnetic field.19 In equilibrium the pres- multidomain state sure exerted on the walls by the field will be compensated by Expression ͑28͒ for the density of the distribution func- ‘‘reaction forces,’’ which ensure equilibrium of the multido- tion enables one to obtain the field dependence of the revers- main structure at H0. If a defect lies in the path of the ible part of the relative elongation of the crystal during the wall, then the wall will bend, and surface forces arise which imposition of a field over the whose interval of fields of will prevent further displacement of the wall. The parameters transition of the multidomain state to the uniform state. The of those, essentially quasi-elastic forces are determined by field dependence of the reversible MS of the multidomain the characteristics of the elastic interaction of the wall and state in the approximation of an ideal multidomain state ␧ ϭ defects. For each value of the field there exists a certain ( unr 0) will be described by the expression balance of the volumes of domains with different directions ␲ 2 2 ␧ʈ͑H͒ 1 /2 H cos ␸ of L. If defects are responsible for the multidomain charac- ϭ ͵ Ϫ ͑ ϩ␩ ͒ ͫͩ ʈ ͪ ␧͑S͒ ͑ ͒ 1 2 1 2 ter, then equilibrium in a field corresponds to the appearance I H/H0 Ϫ␲/2 H ff of certain elastic stresses in the crystal due to the mismatch H2 cos2 ␸ of the elastic fields of the defects and the elastic fields of MS ϫ 2 ␸Ϫͩ Ϫ ͑ ϩ␩ ͒ ͪ 2 ␸ͬ cos 1 2 1 Ќ 2 sin strains in the domains. H ff The irreversible part of the rearrangement of the do- H2 mains is most likely due to the part of the domains where the ϫexpͩ cos2 ␸ͪ d␸. ͑29͒ H2 orientation L is not maintained by counterbalancing quasi- 0 elastic forces. To distinguish the reversible and irreversible An analogous expression can be written down for the components of the rearrangement of the multidomain state, MS perpendicular to the applied field. Ͻ ͑ ͒ we write the distribution function of the domains in the form For H/H0 1 Eq. 29 gives a quadratic field depen- of a sum, dence of the MS, similar to ͑3͒. ͑␸͒ϭ͑ Ϫ␦͒ ͑␸͒ϩ␦ ͑␸͒ ͑ ͒ Figure 10 shows a comparison of the experimental p 1 prev punr , 27 ␧ ␧ ʈ rev(H) and Ќ rev(H) curves for CoCl2 with the curves where ␦ is the fraction of the volume occupied by domains ␧ʈ(H) calculated according to formula ͑29͒ and ␧Ќ(H) cal- ␧(S) whose magnetic moment M is oriented along the magnetic culated by an analogous formula. The parameters , H ff , field and do not cause ‘‘reaction forces.’’ and ␩ʈ were taken from the MS data for the uniform state, An expression for the distribution function of the do- Hd from the data for the reversible component of the MS, mains over the angle ␸ was obtained in Ref. 19: and the parameter ␩Ќ was chosen so as to satisfy the ob- served degree of non-antisymmetry of the field dependences 1 H2 ͑␸͒ϭ ͩ 2 ␸ͪ ͑ ͒ of the longitudinal and transverse MS in CoCl2 . The thin prev ͑ ͒ exp 2 cos , 28 I H/H0 2H0 curves in Fig. 10 show the field dependence of the MS for 804 Low Temp. Phys. 31 (8–9), August–September 2005 Kalita et al.

FIG. 11. Model curves of the magnetization, normalized to its limiting value, and magnetic susceptibility, normalized to its value in the uniform

state, as functions of the field in the multidomain state. FIG. 12. Dependence of the magnetic susceptibility of NiCl2 on field during its removal in the region in other words rearrangement of the multidomain structure. the uniform state, calculated for the same parameters accord- ing to Eq. ͑8͒. The calculated curve for the longitudinal MS is in good agreement with the experimental data over the points where more than two domains come together, elastic entire range of fields causing rearrangement of the multido- stresses will arise. However, if those places coincide with the main state. There is some disagreement with experiment for locations of defects of the crystal structure that create a stress the calculated curve of the transverse MS, but it does not of the opposite sign around them, then the additional elastic exceed the error of measurement. energy due to these stresses will be lessened, and the system Let us use Eq. ͑28͒ to calculate the dependence of the of AFM domains can turn out to be energetically favorable. magnetization of the crystal as a whole in a magnetic field Let us demonstrate such a situation for the hypothetical ex- lying in the EP. If the irreversible component of the rear- ample of a distribution of magnetoelastic domains with sev- rangement of the multidomain state is neglected (␦ϭ0), the eral equivalent ‘‘easiest’’ directions of the L vectors in the expression for the magnetization takes the form EP. Such a structure was considered in Ref. 15 for the case m͑H͒ 1 H ␲/2 H2 ϭ ͵ 2 ␸ ͩ 2 ␸ͪ ␸ of an equilibrium distribution of domains with L vectors ͑ ͒ cos exp 2 cos d , M s I H/H0 H ff Ϫ␲/2 2H0 lying at an angle of 120° (60°) and is illustrated in Fig. 13. ͑30͒ The easy plane of the crystal is represented by a set of hexa- gons. Their boundaries are shown by solid lines. The spon- where M is the magnetization of the uniform state of the s taneous MS of the hexagons leads, for example, as in the crystal for HϭH . ff figure, to elongation of the hexagons along their diagonal In Fig. 11 the dashed curve shows m(H)/M calculated s that is parallel to L. An exception is the region where three according to Eq. ͑30͒ for the same parameters of the distri- domains come together, where the uncompensated ͑for the bution function as were used in the MS calculation. The solid example shown, tensile͒ mechanical stresses cannot be elimi- straight line shows m(H)/M for a uniform state with LЌH. s nated. These regions are formed where domains belonging to The calculated curve for the multidomain state has a charac- different ‘‘triads’’ come together ͑one such triad is shaded in teristic sag below the dependence for the uniform state and Fig. 13͒. As a result of the strain the hexagons become biax- agrees with the experimental data17 on m(H) for NiCl ͑see 2 ial, with C symmetry of the local extension at those sites. Fig. 7͒. 3 Differentiation of Eq. ͑30͒ with respect to H gives the ␹ calculated magnetic susceptibility calc(H). The dotted curve ␹ ␹ Ͻ in Fig. 11 shows a plot of calc(H)/ e . For H Hd it is 2 ϳ proportional to H , it has a maximum for H Hd , and with further increase in field it tends toward the value for the ␹ ␹ uniform state e . The maximum on (H) is due to the rapid ␸ Ϸ change of p( ) near H Hd . Figure 12 shows the experi- mentally obtained curve of the magnetic susceptibility for 45 NiCl2 .

D. Magnetoelastic mechanisms for the influence of defects on the formation of AFM domains In the treatment above we assumed an equiprobable dis- tribution function for directions of the L vector of the AFM domains at Hϭ0. To present a picture of the mutual arrange- FIG. 13. Hypothetical structure of magnetoelastic domains in an easy plane ment of domains, domain walls, and defects for such a dis- AFM. The two-headed arrows show the directions of the sublattice magne- tizations in the domains. The solid lines show the boundaries of the domains tribution is rather difficult. To fill a plane with domains hav- prior to the spontaneous striction, and the dotted lines after. The shaded triad ing boundaries with matched MS strains is impossible. At of domains is an element forming the domain walls of the crystal transition. Low Temp. Phys. 31 (8–9), August–September 2005 Kalita et al. 805 eץ In the ideal crystal such domain structure in the form of ϭ ͑ ͒ϭ ͑ ͒ϩ ͑ Ϫ ͒ Ϫ x0i x0 j ץ ͭ a set of deformed hexagons leaves the domains partially E ͚ e rij ͚ e r0ij ͚ ij ij ij ͑xi x j͒ stressed. The presence of the triangles of extension at the eץ points where three domains come together means that the Ϫ ͑ Ϫ ͒ ␧͑S͒ ͮ Ϫ y 0i y 0 j ץ strain does not conserve the volume of the domains. How- ͑y i y j͒ ever, the estimates made in Ref. 15 show that the cost in 2eץ 1 magnetoelastic energy for these nonuniform strains of the ϩ ͭ ͑ Ϫ ͒2 Ϫ ͒2 x0i x0 j ͑ץ ͚ hexagons is only a few percent of the magnetoelastic energy 2 ij xi x j of the uniform state. Therefore, the structure illustrated in 2eץ Fig. 13 can be regarded as a rather good approximation of Ϫ ͑ Ϫ ͒͑ Ϫ ͒ Ϫ ͒ x0i x0 j y 0i y 0 j ͑ץ͒ Ϫ ͑ץ 2 the ME domains with the minimal elastic stresses and with xi x j y i y j 2eץ .the minimal cost in elastic energy ϩ ͑ Ϫ ͒2ͮ ͑␧͑S͒͒2 ͑ ͒ Ϫ 2 y 0i y 0 j . 31 ץ If tensile point defects are placed at junctions of three ͑y i y j͒ domains, then the matching of the strains generated by the ͑ ͒ defects and the ME strains of the domains will lead to partial The first sum in 31 is equal to the interaction energy E0 or complete compensation of the stresses created by these between defects prior to the spontaneous anisotropic MS. In sources. The volume of the region of the crystal deformed by the approximation that the anisotropic strain is independent ͑ ͒ the defect and the energy cost associated with the defect of the local positions of the defects, the second sum in 31 decrease. Point defects that appear during growth increase should be equal to zero. The expression multiplying the ␧(S) the elastic energy of a crystal. Superposition of the location square of in the third sum will be denoted by k. The interaction energy between defects after the anisotropic MS of the defects with the junctions of three domains decreases takes the form this contribution to the energy, i.e., it makes the multidomain ϭ ϩ ␧͑S͒͒2 ͑ ͒ state preferable. Analogous elastic matching can be achieved E E0 k͑ . 32 if the junction of three domains is threaded by disclinations 46 If we consider growth defects, which can be assumed perpendicular to the EP. We emphasize that in both the case almost equilibrium, then the energy of interaction between of point defects and the case of disclinations the stress that them will satisfy a minimum principle. In that case kϾ0, and these defects exert on the domain walls is independent of the uniform spontaneous anisotropic MS, which is advanta- temperature, which account for the observed temperature- geous for a defectless crystal, increases the interaction en- importance of the parameter Hd in CoCl2 . ergy of the defects. That is, the dependence of the interaction The geometry structure of the domains illustrated in Fig. energy of the defects on the magnetostrictive displacements 13 presupposes the presence of defects isotropically extend- crates a quasi-elastic restoring force. ing the crystal in the EP. In that case the defect does not Since defects cannot move through the crystal, it follows specify the directions for the L vectors in the domains but from ͑32͒ that they will prevent the formation of the single- preserves the orientational degeneracy in the EP. If the spon- domain state of an AFM with anisotropic MS. The crystal taneous MS of the domains is opposite in sign, then a region will break up into regions with uniform, but differently di- of compression will form at the junctions of three domains. rected, MS in them. Thus the two mechanisms for the ME For stabilizing such a structure one needs defects of another interaction of defects and domains promote the equilibrium type, e.g., vacancies. of the multidomain AFM state. The ‘‘triad’’ model has a different distribution function of the domains over orientations of L than that considered in VI. CONCLUSION Sec. 5.1. Nevertheless, all of the qualitative features of the Analysis of the data on the induced MS and magnetiza- behavior of the induced MS and magnetization of the crystal tion of multidomain AFM state in layered easy-plane AFMs in the region of rearrangement of the multidomain state will of the ion dihalide group shows that the system of AFM be present in this model also, although the details will be domains in them is due to the ME interactions of spontane- somewhat different. ously strained domains and the elastic fields of defects. In the The mechanism of surrounding defects by triads of do- multidomain state the spontaneous MS of the crystal as a mains does not take into account the actual interaction of the whole is close to zero in the absence of magnetic field, while defects, which can promote equilibrium of the multidomain the domains retain their spontaneous MS. The spontaneous state. lowering of the symmetry upon the establishment of AFM 19 Let us consider the interaction of the defects in the order in the presence of many-fold degeneracy for spatial simplest case of a planar isotropic medium, when the energy orientation of the antiferromagnetic vector comes about in of interaction between the neighboring defects ͑numbered i the given case through the formation of a nonuniform mul- and j) depend only on the distance between them, e tidomain state. This multidomain state is reversible in the ϭ e(rij). Suppose that during the equilibrium of the uniform imposition and removal of magnetic field, i.e., it is a state of AFM order a uniform spontaneous anisotropic MS occurs thermodynamic equilibrium. which is is positive and equal to ␧(S) along the x axis and It is known that the spontaneous lowering of the symme- negative, and with the same magnitude, along the y axis, try in an AFM occurs because the uniform anisotropic MS Ϫ␧(S) ͑both lie in the EP͒. The total energy of interaction lowers the magnetic energy of the crystal linearly as a func- between defects in this case will be tion of the magnitude of the strictive strains. In that case the 806 Low Temp. Phys. 31 (8–9), August–September 2005 Kalita et al. elastic energy of the crystal increases quadratically in those 15 V. M. Kalita and A. F. Lozenko, Fiz. Nizk. Temp. 27,489͑2001͓͒Low strains. As a result, the total energy of the crystal reaches a Temp. Phys. 27, 358 ͑2001͔͒. 16 minimum at some equilibrium spontaneous MS. In real crys- V. M. Kalita and A. F. Lozenko, Fiz. Nizk. Temp. 27,872͑2001͓͒Low Temp. Phys. 27, 645 ͑2001͔͒. tals the elastic energy is supplemented by the interaction be- 17 V. M. Kalita, A. F. Lozenko, and P. A. Trotsenko, Fiz. Nizk. Temp. 28,378 tween the elastic fields of defects separated by distances ͑2002͓͒Low Temp. Phys. 28,263͑2002͔͒. much greater than the lattice parameter. Therefore the cost in 18 V. M. Kalita, A. F. Lozenko, S. M. Ryabchenko, P. A. Trotsenko, and T. M. elastic energy upon spontaneous MS may be lessened if the Yatkevich, Fiz. Tverd. Tela ͑St. Petersburg͒ 46, 317 ͑2004͓͒Phys. Solid State 46,326͑2004͔͒. MS strains are not uniform but matched with the elastic 19 fields of the defects. This may turn out to be sufficient to V. M. Kalita, A. F. Lozenko, S. M. Ryabchenko, and P. A. Trotsenko, Zh. E´ ksp. Teor. Fiz. 126, 1209 ͑2004͓͒JETP 99, 1054 ͑2004͔͒. compensate the cost in exchange energy in the domain walls. 20 E. V. Gomonaj and V. M. Loktev, Fiz. Nizk. Temp. 25, 699 ͑1999͓͒Low In other words, the formation of a multidomain magnetoelas- Temp. Phys. 25, 529 ͑1999͔͒. tic state in a crystal containing defects is a consequence of 21 H. Gomonay and V. Loktev, J. Phys. C 14, 3959 ͑2002͒. the presence of long-range elastic forces due to the defects. 22 M. M. Farztdinov, Physics of Magnetic Domain in Antiferromagnets and ͓ ͔ ͑ ͒ This leads to a nonuniform elastically strained state of the Ferrites in Russian , Nauka, Moscow 1981 . 23 A. Hubert, Theorie der Domanenwande in Geordneten Medien, Springer- crystal, which entails inhomogeneity of the magnetic state. It Verlag, Berlin ͑1974͒, Mir, Moscow ͑1977͒. is that nonuniform state that gives a minimum of the total 24 V. G. Bar’yakhtar, A. E. Borovik, and V. A. Popov, Zh. E´ ksp. Teor. Fiz. energy of the crystal upon spontaneous MS. We note that this 62, 2233 ͑1972͓͒Sov. Phys. JETP 35, 1844 ͑1972͔͒. should be the case not only in easy-plane AFMs but also in 25 A. N. Bogdanov and I. E. Dragunov, Fiz. Nizk. Temp. 24, 1136 ͑1998͒ ͓Low Temp. Phys. 24,852͑1998͔͒. ferromagnets with spontaneous magnetostriction. In the lat- 26 ͑ ͒ S. V. Vonsovski, Magnetism, Vols. 1 and 2, Wiley, New York 1974 , ter, however, the formation of the method state is stimulated Nauka, Moscow ͑1971͒. 27 primarily by the magnetostatic forces, while the contribution V. G. Bar’yakhtar, A. N. Bogdanov, and D. A. Yablonski, Usp. Fiz. Nauk due to defects and magnetoelasticity turns out to be second- 156,47͑1988͒. ary or, at least, masked. 28 L. Neel, Ann. Phys. 3, 137 ͑1948͒. 29 V. G. Bar’yakhtar, A. E. Borovik, and V. A. Popov, Zh. E´ ksp. Teor. Fiz. 9, ͑ ͔͒ * 634 1969 . E-mail: [email protected] 30 Y. Y. Li, Phys. Rev. 101,1450͑1956͒. 31 V. V. Eremenko, N. F. Kharchenko, and L. I. Beliy, J. Appl. Phys. 50, 7751 ͑1979͒. 32 1 J. W. Leech and A. J. Manuel, Proc. Phys. Soc. London, Sect. B 69, 210 N. F. Kharchenko and A. V. Bely, Izv. AN SSSR, ser. fiz. 446, 1451 ͑1956͒. ͑1980͒. 2 33 ͑ ͒ H. Yoshizawa, K. Ubukoshi, and K. Hirakawa, J. Phys. Soc. Jpn. 48,42 I. E. Dzyaloshinski, JETP Lett. 25, 414 1977 . ͑1980͒. 34 A. S. Kovalev and A. M. Kosevich, Fiz. Nizk. Temp. 3,259͑1977͓͒Sov. 3 M. K. Wilkinson, J. W. Cable, E. O. Wollan, and W. C. Koehler, Phys. J. Low Temp. Phys. 3, 125 ͑1977͔͒. Rev. 113,497͑1959͒. 35 I. M. Lifshitz, Zh. E´ ksp. Teor. Fiz. 42, 1354 ͑1962͓͒Sov. Phys. JETP 15, 4 M. O. Kostryukova and L. M. Kashirskaya, JETP Lett. 9, 238 ͑1969͒. 939 ͑1962͔͒. 5 A. S. Borovik-Romanov, ‘‘Antiferromagnetism,’’ in Progress in Science 36 Z. A. Kaze , M. V. Levanidov, and V. I. Sokolov, Prib. Tekh. E´ ksp., No. 1, ͓ ͔ ͑ ͒ in Russian , Izd-vo AN SSSR, Moscow 1962 . 196 ͑1982͒. 6 ´ A. F. Lozenko and S. M. Ryabchenko, Zh. Eksp. Teor. Fiz. 65,1085 37 V. M. Kalita, A. F. Lozenko, and S. M. Ryabchenko, Fiz. Nizk. Temp. 26, ͑ ͓͒ ͑ ͔͒ 1973 Sov. Phys. JETP 38, 538 1974 . 671 ͑2000͓͒Low Temp. Phys. 26, 489 ͑2000͔͒. 7 34 ͑ ͒ K. Katsumata and K. Yamasaka, J. Phys. Soc. Jpn. , 346 1973 . 38 J. Gunzbourg, S. Papassimacopoulos, A. Mieden-Gros, and A. Allain, J. 8 J. Magarino, J. Tuchendler, A. R. Fert, and J. Gelard, Solid State Com- Phys. C 32, 125 ͑1971͒. mun. 23,175͑1977͒. 39 D. Billerey, C. Terrier, A. J. Pointon, and J. P. Redoules, J. Magn. Magn. 9 A. S. Borovik-Romanov and E. G. Rudashevski , Zh. E´ ksp. Teor. Fiz. 47, Mater. 21, 187 ͑1980͒. 2095 ͑1964͓͒Sov. Phys. JETP 20, 1407 ͑1965͔͒. 40 ͑ ´ ͒ ͑ ͒ 10 E. A. Turov and V. G. Shavrov, Fiz. Tverd. Tela ͑Leningrad͒ 7, 217 ͑1965͒ L. Neel’ L. Neel , Izvestiya AN SSSR Ser. Fiz. 21, 890 1957 . 41 ͑ ͒ ͓Sov. Phys. Solid State 7, 166 ͑1965͔͒. M. E. Lines, Phys. Rev. 131, 546 1963 . 42 ͑ ͓͒ 11 A. F. Lozenko, P. E. Parkhomchuk, S. M. Ryabchenko, and P. A. Trot- V. M. Kalita and A. F. Lozenko, Fiz. Nizk. Temp. 28,91 2002 Low senko, Fiz. Nizk. Temp. 14, 941 ͑1988͓͒Sov. J. Low Temp. Phys. 14, 517 Temp. Phys. 28,66͑2002͔͒. 43 ͑1988͔͒. W. J. De Haas, B. H. Schultz, and J. Koolhaas, Physica 7,57͑1940͒. 12 A. K. Zvezdin, V. M. Matveev, A. A. Mukhin, and A. I. Popov, Rare-Earth 44 C. Starr, F. Bitter, and A. R. Kaufmann, Phys. Rev. 58,977͑1940͒. Ions in Magnetically Ordered Crystsals ͓in Russian͔, Nauka, Moscow 45 V. M. Kalita, A. F. Lozenko, P. A. Trotsenko, and T. M. Yatkevich, Fiz. ͑1985͒. Nizk. Temp. 30,38͑2004͓͒Low Temp. Phys. 30,27͑2004͔͒. 13 D. Billerey, C. Terrier, A. J. Pointon, and J. P. Redoules, J. Magn. Magn. 46 A. M. Kosevich, Theory of Crystal Lattice, Wiley, New York ͑1999͒, Mater. 21,187͑1980͒. Vishcha Shkola, Kharkov ͑1987͒. 14 V. M. Kalita, A. F. Lozenko, S. M. Ryabchenko, and P. A. Trotsenko, Ukr. Fiz. Zh. ͑Russ. Ed.͒ 43, 1469 ͑1998͒. Translated by Steve Torstveit LOW TEMPERATURE PHYSICS VOLUME 31, NUMBERS 8–9 AUGUST–SEPTEMBER 2005

Investigation of the anomalies of the magnetoelectric and magnetoelastic properties of single crystals of the ferroborate GdFe3„BO3…4 at phase transitions A. M. Kadomtseva, Yu. F. Popov,* S. S. Krotov, G. P. Vorob’ev, and E. A. Popova

M. V. Lomonosov Moscow State University, Vorob’evy gory, Moscow 119992, Russia A. K. Zvezdin

Institute of General Physics, Russian Academy of Sciences, Moscow 119991, Russia L. N. Bezmaternykh

L. V. Kirensky Institute of General Physics, Siberian Branch of the Russian Academy of Sciences, Krasnoyarsk 660036, Russia ͑Submitted January 21, 2005͒ Fiz. Nizk. Temp. 31, 1059–1067 ͑August–September 2005͒ The transformation of various properties of gadolinium ferroborate single crystals at phase transitions, both spontaneous and induced by magnetic fields up to 200 kOe, is investigated theoretically and experimentally. Particular attention is paid to elucidating the role of magnetoelectric interactions and the change in them at spin-reorientation transitions accompanied by a change of magnetic symmetry. With that goal the magnetoelastic and magnetoelectric properties of the system are investigated over a wide range of temperatures for two orientations of the magnetic field, Hʈc and HЌc, and a fundamental difference of the character of the field dependences of the magnetostriction and electric polarization is found. In the framework of a symmetry approach a description of the magnetic structures and their transformations in the system GdFe3(BO3)4 is proposed, and an interpretation of the experimentally observed properties is given. © 2005 American Institute of Physics. ͓DOI: 10.1063/1.2008142͔

INTRODUCTION In Ref. 7 the magnetic behavior of gadolinium ferrobo- Rare-earth ferroborates RFe (BO ) , isostructural to the rate was investigated in the framework of three-sublattice 3 3 4 ͑ ͒ natural mineral huntite, have a rhombohedral structure and ordering of the triangular type of the iron subsystem. The 3ϩ Ͻ belong to the hexagonal space group R32 (D7). spins of the Fe ions in the temperature interval 10 K T 3 Ͻ These compounds have attracted particular interest in re- 38 K were found to lie in the ‘‘easy plane’’ at an angle of cent years in connection with the prospects for their practical 120° to one another. Below 10 K, according to Ref. 7, under use, particularly in laser technique1,2 and for optical second- the influence of an interaction with the gadolinium sub- 3ϩ harmonic generation,3 and also because of the discovery of system, the Fe spins are reoriented from the easy plane in unusual magnetic properties and diverse phase transitions in the direction toward the c axis, forming a cone with axis them.4–6 According to measurements of the temperature de- along c. Since nothing was said about any change in the unit pendence of the magnetic susceptibility7 and heat capacity,8 call at the magnetic phase transition, and the system re- three phase transitions are observed in gadolinium ferrobo- mained antiferromagnetic over the whole temperature inter- ϭ val, the opening angle of the cone ͑the tilt angle of the Fe3ϩ rate: a structural phase transition at TC 156 K, a magnetic 3ϩ ϭ spins toward the c axis for compensation of the magnetiza- ordering of the Fe ions at TN 38 K, and a spin- reorientation phase transition at T ϭ10 K.7 The existing tion of the gadolinium subsystem͒ should ‘‘follow’’ the in- R ϩ data in the literature nevertheless do not permit one to reach crease of the magnetic moment of the Ge3 ions with de- definite conclusions as to the nature of the phase transitions creasing temperature. For an orientation of the external observed in these compounds. magnetic field Hʈc at TϽ10 K a field-induced reorientation ϩ In regard to the magnetic structure of gadolinium fer- of the Fe3 spins from the c axis to the easy plane was roborate the data in the literature are contradictory. In Ref. 9 observed.5,7 it was stated on the basis of a study of antiferromagnetic We carried out additional experimental and theoretical ͑AFM͒ resonance that the magnetic phase transition of the investigations of the transformation of various properties of 3ϩ Fe ion subsystem at TN corresponds to two-sublattice spin gadolinium ferroborate single crystals at the spontaneous and ordering of the easy-plane type, and, as the temperature is field-induced phase transitions at high magnetic fields up to lowered below TR , the interaction of the iron and gado- 200 kOe. Particular attention was paid to investigation of the linium subsystems brings about a reorientation of the Fe3ϩ role of magnetoelectric interactions and their changes at the spins from easy-plane to easy-axis ͑the c axis͒. The magnetic spin-reorientation transitions. An interpretation of the antifer- unit cell of the system is doubled along the c axis in com- romagnetic properties of the system was given in the frame- parison with the crystalline unit cell in hexagonal coordi- work of a thermodynamic approach. Studies of the magneto- nates. elastic and magnetoelectric properties of the system for two

1063-777X/2005/31(8–9)/7/$26.00807 © 2005 American Institute of Physics 808 Low Temp. Phys. 31 (8–9), August–September 2005 Kadomtseva et al.

FIG. 1. Isotherms of the dependence of the longitudinal magnetostriction of ʈ the GdFe3(BO3)4 single crystal for H c. FIG. 3. H-T phase diagram obtained for GdFe3(BO3)4 from measurements of the longitudinal magnetostriction and electric polarization for Hʈc.

orientations of the magnetic field: Hʈc and HЌc ͑for which at present there is absolutely no information in the literature͒. Our studies of the magnetoelectric properties of We expected that gadolinium ferroborate, which has a non- GdFe3(BO3)4 single crystals have shown that in the course centrosymmetric space group, would exhibit substantial of the spin-reorientation transition at TϽ10 K for a field manifestations of the magnetoelectric interactions and their orientation Hʈc there are also jumps of the electric polariza- change at the spin-reorientation transitions, which are ac- tion ͑Fig. 2͒, evidently due to the change of the magnetoelec- companied by a change of magnetic symmetry. tric interactions as a result of the change of magnetic sym- metry. The spin-reorientation transition induced by a ʈ EXPERIMENTAL RESULTS magnetic field H c is accompanied by hysteresis with re- spect to the the field, which indicates that the transition is a We studied the electric polarization Pi and magnetostric- first-order phase transition. Figure 3 shows the H-T phase ␭ tion i as functions of magnetic field up to 200 kOe in the diagrams obtained from measurements of the magnetoelec- temperature interval 4.2–50 K for Hʈc and HЌc by the tech- tric and magnetoelastic properties for Hʈc, which demon- nique described in Ref. 10, and also the temperature depen- strated good agreement with the values of the threshold dence of the thermal expansion of GdFe3(BO3)4 . fields. The character of the field dependence of the magne- Figure 1 shows isotherms of the dependence of the lon- tostriction and electric polarization differs strongly for Hʈc gitudinal magnetostriction of the gadolinium ferroborate and HЌc. Figure 4 shows the dependence of the longitudinal single crystal at Hʈc. It is seen that in the low-temperature magnetostriction for Hʈa. It is seen that at Tϳ5 K the mag- region TϽ10 K at a certain critical value of magnetic field netostriction initially depends weakly on field, and then, at crit HϭH (Hʈc) there are jumps in the magnetostriction, Hcritϳ37 K, it increases in a jump to 1ϫ10Ϫ5. With increas- where, according to Refs. 7 and 9, a magnetic-field-induced ing temperature the threshold field decreases monotonically. spin reorientation from the c axis toward the easy plane was Starting at temperatures Tϳ10 K and on up to 38 K the Ͻ Ͻ ϭ observed. At high temperatures 10 T TN 38 K the mag- magnetostriction increases in a sharp jump even in weak netostriction does not exhibit anomalies and depends qua- magnetic fields ϳ2 kOe and then behaves monotonically. crit dratically on the field. The threshold fields H at which the The magnetic-field dependence of the electric polarization is magnetostriction jumps appear decrease with increasing tem- of an analogous character ͑Fig. 5͒, and the threshold fields perature and agree with the fields at which the magnetization agree with each other, as is seen on the H-T phase diagrams jumps were observed at the field-induced reorientation of the ͑Fig. 6͒ constructed from the data of the ␭(H) and P(H) 7 spins from the c axis to the easy plane. measurements. Thus for Hʈc and HЌc a strict correlation of

FIG. 2. Isotherms of the longitudinal electric polarization in a gadolinium FIG. 4. Isotherms of the longitudinal magnetostriction for a GdFe3(BO3)4 ferroborate single crystal for Hʈc. single crystal at Hʈa. Low Temp. Phys. 31 (8–9), August–September 2005 Kadomtseva et al. 809

FIG. 5. Curves of the dependence of the electric polarization on magnetic ʈ FIG. 7. Dependence of the thermal expansion for a GdFe (BO ) single field H a for a GdFe3(BO3)4 single crystal at various temperatures. 3 3 4 crystal along the c axis in the region of the structural transition TC ϭ156 K. the magnetoelectric and magnetoelastic properties is estab- lished. It should be noted that in measurements of the tem- perature dependence of the thermal expansion we observed a netic structures that, depending on the concrete form of the sharp ␭-shaped anomaly at a temperature of 156 K ͑Fig. 7͒. exchange interactions ͑and also the features of the aniso- Recent measurements11 have shown that at Tϭ156 K there tropic interactions͒ can appear in the ferroborate is also a jumplike change of the dielectric constant, which GdFe3(BO3)4 below the magnetic ordering temperature. We apparently indicates the onset of spontaneous electric order- begin with a discussion of the features of the magnetic phase ͑ ing at that temperature. transition at the temperature TN at which the moments of the 3ϩ ͒ Fe ions order . We recall that the unit cell of RFe3(BO3)4 3ϩ ͑ THEORETICAL PART in the rhombohedral scheme contains one R ion position 1a) and three Fe3ϩ ions ͑position 3d). If the space diagonal In view of what we have said above, it is best to take a of the elementary rhombohedron is assumed to be directed symmetry approach to the theoretical description of the prop- vertically, then the crystal structure of the RFe3(BO3)4 sys- erties of the GdFe3(BO3)4 system. On the one hand, it al- tem will be a set of rhombohedrally shifted identical hori- lows one to explain the behavior of that system in a unified zontal layers ͑oblique with respect to the vertical but parallel way and to identify the active and passive degrees of free- to each other͒ alternating along the vertical. The distance dom responsible for the change of its magnetic and crystal- between two adjacent equivalent layers is one-third of the line symmetry, the corresponding order parameters, and the vertical period c of the structure, which is the height of the interactions of the latter both among themselves ͑leading to unit cell of RFe3(BO3)4 in the hexagonal crystallographic phenomena such as weak ferromagnetism, intrinsic magne- coordinates ͑Fig. 8͒. But then the hexagonal unit cell of the toelectric effect, etc.͒ and with external fields. At the same ͑ RFe3(BO3)4 system which is doubly body-centered and time, it is exceptionally efficient for predicting new effects tripled in volume as compared to the rhombohedral cell͒ con- ͑ according to the ‘‘possible-impossible’’ principle from con- tains three formula units (Zϭ3), or, in other words, three siderations of whether the space-time symmetry in the be- translationally equivalent R3ϩ ions ͑position 3a) and, ac- ͒ ϩ havior of a particular system is broken or preserved . cordingly, nine Fe3 ions ͑position 9d). Since the rhombo- The group-theoretical approach used here ͑in the spirit of ͑ ͒ hedral truly primitive crystal cell of GdFe3(BO3)4 contains the Landau theory of phase transitions͒ permits one to de- scribe in an exhaustive way all the possible types of mag-

FIG. 6. H-T phase diagram obtained from measurements of the longitudinal FIG. 8. Fragment of the Bravais lattice of the system RFe3(BO3)4 with the magnetostriction and electric polarization for Hʈa, for the single crystal rhombohedral ͑primitive͒ and hexagonal ͑unit͒ cells shown explicitly. The ͑ ͒ 3ϩ GdFe3(BO3)4 . lattice sites ellipses correspond to the positions of the R ions. 810 Low Temp. Phys. 31 (8–9), August–September 2005 Kadomtseva et al. only one R3ϩ ion, which will be magnetic for RϭGd, Nd, TABLE I. Basis functions of the irreducible representations ͑structural, magnetic, and exchange͒ of the group R32. and our system is antiferromagnetic at all temperatures be- low TN , we shall assume that the elementary translations of m m m the magnetic Bravais lattice a1 a2 a3 of the GdFe3(BO3)4 system will be related to the elementary translations ͑ ͒ a1 ,a2 ,a3 of its crystal lattice Fig. 8 by the relations mϭ ϩ mϭ ϩ mϭ ϩ ͑ ͒ a1 a2 a3 , a2 a3 a1 , a3 a1 a2 . 1 It follows from relations ͑1͒ that the magnetic primitive cell of GdFe3(BO3)4 will be a rhombohedron doubled in volume in comparison with the high-temperature primitive crystal cell. In fact, the volume Vm of the magnetic cell is found mϭ m ͓ mϫ m͔ϭ ϩ ͓ ϩ from the relation V a1 • a2 a3 (a2 a3)• (a3 a1) ϫ ϩ ϭ c c (a1 a2)͔ 2V , where V is the volume of the primitive crystal cell. In principle, since in our case the appearance of antiferromagnetic order in the system from the standpoint of magnetic symmetry corresponds to a change of the Bravais lattice on account of the appearance of anti-translations in it ͑doubling of a certain number of primitive translations of the ͒ ͑ ͒ Note: The primed representations are odd with respect to translation along initial crystal lattice , besides the case 1 considered above, the c axis. in terms of exchange symmetry there are two more possibili- ties, fundamentally different from the first: 1͒ when only one of the primitive translations is doubled, while the other two ͒ remain unchanged; 2͒ when some pair of the three transla- sentations . Under the condition that in hexagonal tions doubles, while the third translation remains unchanged. coordinates the vector of the AFM structure appearing at the ´ ϭ͕ ͖ One is readily convinced that the two cases mentioned would Neel point is equal to k 0,0,1/2 , we obtain the following lead, as a result of a phase transition, to the ‘‘loss’’ of a relations: ϭ) Ϫ ͒Ϫ) Ϫ ͒ threefold symmetry axis and, accordingly, to a change of the B1 ͑S1 S2 ͑S4 S5 , symmetry class of the system from hexagonal to monoclinic. ϭ ϩ Ϫ Ϫ Ϫ ϩ For the existing ideas about the hierarchy of exchange inter- B2 S1 S2 2S3 S4 S5 2S6 , actions ͑intralayer and interlayer͒, in our system the realiza- B˜ ϭ)͑S ϪS ͒ϩ)͑S ϪS ͒, tion of these possibilities is improbable. For that reason we 1 1 2 3 5 shall restrict the discussion to case ͑1͒. If one is working with ˜ ϭ ϩ Ϫ ϩ ϩ Ϫ B2 S1 S2 2S3 S4 S5 2S6 , the hexagonal unit cell ͑which is more natural from an ex- ϭ ϩ ϩ Ϫ Ϫ Ϫ perimental standpoint͒ for this system, the change of mag- L S1 S2 S3 S4 S5 S6 . netic symmetry ͑1͒ leads to doubling of the height of the ϭ ϩ ϩ ϩ ϩ Ϫ ͑ ͒ F S1 S2 S3 S4 S5 S6 . 2 GdFe3(BO3)4 unit cell. Since no concrete change of the crystal symmetry ͑micro- and, accordingly, macroscopic͒ of Since we are interested not only in the magnetic but also ϭ the system at TC 156 K has been established the magnetoelastic and magnetoelectric properties of the 1͒ experimentally ͑while such a change could in principle oc- GdFe3(BO3)4 system, in the corresponding column of the cur with a doubling of the unit cell, as, e.g., in the case of table of irreducible representations of the group R32 ͑Table ͒ BiFeO3), we assume that the doubling of the unit cell of the I we also give the basis functions of the ‘‘nonmagnetic’’ ͑ GdFe3(BO3)4 system occurs at the Ne´el temperature TN . representations corresponding to structural macro-variables As we have said, at high temperatures the rare-earth fer- of the system—to the components of the strain tensor and roborates belong to the rhombohedral space group R32 electric polarization vector͒. 7 In accordance with formulas ͑2͒, the pair of functions (D3). Taking the change of the unit cell at TN into account, for describing the magnetic properties of our system ͑con- ͕B1 ,B2͖, corresponding in the exchange approximation to a struction of the corresponding irreducible representations of two-dimensional representation ͑odd with respect to a shift the group R32) we consider six Fe3ϩ magnetic moments by a period along the c axis͒, will itself describe an Ϫ characterized by spins S1 S6 . The first three of them— exchange-noncollinear AFM structure of the triangular type, 3ϩ S1 ,S2 ,S3—belong to one horizontal layer, and the other with the closest three moments, those belonging to the Fe three—S4 ,S5 ,S6—lie in the adjacent layer parallel to the ions found at the center of the sides of the elementary tri- first at a distance of c/3 from it. As is done in Landau theory, angle of Gd ions of one layer, directed at an angle of 120° to in describing the magnetic phase transition it is necessary to one another, and the moments of the next-nearest layer anti- Ϫ go over from the individual spins S1 S6 to their symme- parallel to the moments of the first layer. Another pair of ͑ ˜ ˜ trized combinations basis functions of the corresponding ir- functions, ͕B1 ,B2͖, corresponding in the exchange approxi- reducible representations of the group R32), each of which mation to yet another two-dimensional representation ͑even describes one of the possible types of magnetic ordering— with respect to vertical translation͒, will describe another either collinear ͑in the case of a one-dimensional representa- exchange-noncollinear AFM structure, also of the triangular tion͒ or noncollinear ͑for two- and three-dimensional repre- type, that is possible from the standpoint of symmetry. Being Low Temp. Phys. 31 (8–9), August–September 2005 Kadomtseva et al. 811

͑ ͒ ordered as in pair of functions ͕B1 ,B2͖) antiferromagneti- modes of second order, which describe the simplest charac- cally within each layer, the moments of adjacent layers will teristic interactions of the different magnetic modes of our be coupled ferromagnetically, and the resulting magnetic system, structure will be antiferromagnetic, with coincident magnetic ⌽ ϭ␣ ͓L ͑B ϩB ͒ϩL ͑B ϪB ͔͒ and crystal cells. If one goes by the magnetizations of the rel.int. 1 X 1Y 2X Y 1X 2Y ͑ ϩ␣ ϩ ͒ ͑ ͒ layers, the function L describes two-sublattice and formally 2LZ͑B1X B2Y . 5 speaking, six-sublattice͒ collinear AFM ordering in which all Furthermore, we write the main intrinsic ͑in the absence the moments of one layer are coupled ferromagnetically and of magnetic field͒ magnetoelectric contributions of both ex- those of the next layer are also coupled ferromagnetically change and relativistic nature, among themselves, but are opposite to the moments of the initial layer, etc. ͑this corresponds to the case k ⌽ ϭ␤ ͕ ͑ " ͒Ϫ ͑ " ͖͒ϩ␤ ͕ ͑ 2Ϫ 2 ͒ ME 0 PX L B2 PY L B1 1 PX Lx LY ϭ͕0,0,1/2͖). The function F is the magnetization vector of Ϫ ͖ϩ␤ ͑ Ϫ ͒ the magnetic unit cell. It follows from what we have said that 2PYLXLY 2LZ PXLY PYLX the dominant role of the antiferromagnetic interlayer ex- ϩ␤ P L ͑B ϪB ͒. ͑6͒ change ͑brought about by the interaction of nearest spins 3 Z Z 1Y 2X lying in neighboring layers͒ leads to the appearance, as the We also take into account the simplest magnetoelastic con- temperature is lowered, first of 1D magnetic ordering along tributions: the corresponding helical chains, parallel to the c axis, of ⌬⌽ ϭ␥1u L2 ϩ␥2u ͑L2 ϩL2 ͒ϩ␥1͑u ϩu ͒L2 iron ions surrounded by distorted oxygen octahedra.6 De- Mel 1 ZZ Z 1 ZZ X Y 2 XX YY Z pending on the character of the intralayer exchange of near- ϩ␥2͑ ϩ ͒͑ 2 ϩ 2 ͒ϩ␥ ͓͑ Ϫ ͒ 2 uXX uYY LX LY 3 uXX uYY est spins of the Fe3ϩ ions ͑and, hence, the exchange between ϫ͑ 2 Ϫ 2 ͔͒ϩ ϩ␥ ͓ ͑ 2 Ϫ 2 ͒ individual ordered chains͒—ferromagnetic or LX LY 4uXYLXLY] 4 uYZ LX LY antiferromagnetic—our system on the whole in the exchange ϩ2u L L ͔. ͑7͒ approximation will be either a collinear antiferromagnet XZ X Y ͑which, in the main approximation, exhibits traits of a two- The establishment of collinear AFM order in the system ͒ sublattice antiferromagnet or an exchange-noncollinear an- at the Ne´el temperature TN means, in the ideology of the tiferromagnet of the triangular type. There are no other pos- Landau theory, that the following inequalities hold: sibilities for the aforementioned change of magnetic ⌳͑2͒͑ ͒ϭ ⌳͑2͒͑ ͒Ͼ ⌳͑2͒͑ ͒Ͼ symmetry ͑1͒. 1 TN 0, 2 TN 0, 3 TN 0, Following Ref. 9, in our discussion below we shall give ͑ ͒ ⌳ 4 ͑T ͒Ͼ0. ͑8͒ preference mainly to the case of collinear ordering. Then the 4 N magnetic ordering that appears at the temperature TN will be Assuming that in the temperature region 10 KϽT characterized by a vector L. Nevertheless, for the sake of Ͻ38 K the system is an easy-plane antiferromagnet, we must ͑ ␩(2) Ͼ completeness since more-detailed studies of the system are assume that the inequality 1 (T) 0 holds in that tempera- still warranted͒ we shall write the thermodynamic potential ture interval. Then the point of the spin-reorientation transi- ͑ ␩(2) of the system the subsystems of ordering moments of the tion TR will correspond to the condition 1 (TR) 3ϩ ͒ ϩ␩(4) ϭ 12 Fe ions to second order in the exchange approximation 1 (TR) 0, and the system goes from easy-plane to Ͻ ␩(2) with allowance for the characteristic energy of all possible easy-axis at temperatures T TR , for which 1 (T) magnetic modes of our system: ϩ␩(4) Ͻ 1 (T) 0. 1 1 1 1 It follows from the results of Ref. 8 that as the tempera- ⌽͑2͒ϭ ⌳͑2͒L2ϩ ⌳͑2͒͑B2ϩB2͒ϩ ⌳͑2͒F2ϩ ⌳͑2͒ ture is lowered in the region Tϳ20 K the subsystem of mo- ex 2 1 2 2 1 2 2 3 2 4 ments of the rare-earth ions Gd will begin to play a role in ϫ͑B˜ 2ϩB˜ 2͒. ͑3͒ the magnetism, the interaction of the subsystem of Fe spins 1 2 with it having a substantial influence on the spin- We obtain the relativistic ͑magnetically anisotropic͒ contri- reorientation transition in our system, although no actual bution taking into account not only the independent magnetic magnetic ordering occurs in the rare-earth subsystem itself. modes, after explicitly separating out the terms correspond- Let us give some special attention to this fact. We start for ing to the case kϭ͕0,0,1/2͖, the temperature region TϾ10 K, i.e., where the system is 1 1 1 still an easy-plane antiferromagnet. Since the magnetic unit ⌽͑2͒ ϭ ␩͑2͒ 2 ϩ ␩͑4͒ 4 ϩ ␩ ͑ Ϫ ͒2 rel.ind 1 LZ 1 LZ 2 B1Y B2X cell of the GdFe3(BO3)4 system is doubled at the phase tran- 2 4 2 sition at the Ne´el point ͑and then contains two Gd ions͒, for 1 1 describing the interaction of the disordered subsystem of mo- ϩ ␩ ͑B ϩB ͒2ϩ ␩ ͑B ϩB ͒2 2 3 1X 2Y 2 4 1Y 2X ments of the Gd ions with the ordered system of moments of the Fe ions, we introduce an antiferromagnetic vector L˜ for ϩ͑ Ϫ ͒2 ͑ ͒ B1X B2Y , 4 the gadolinium subsystem also. Taking into account the sym- ͓in Eq. ͑4͒ we have explicitly taken into account the aniso- metry of the sites of the two gadolinium ions in the magnetic tropic contributions of both second and fourth order only for unit cell of our system, we easily see that the vectors L˜ and the mode L of interest to us, which, as we know,12 will be L correspond to equivalent representations of the exchange responsible for possible spin reorientation͔ but also the symmetry group of our system, and so the energy of the f -d mixed contributions ͑‘‘mixing’’ the different magnetic exchange interaction, Eϭ␬L˜ "L will be an invariant of the 812 Low Temp. Phys. 31 (8–9), August–September 2005 Kadomtseva et al.

group R32. The mean energy of the f -d exchange, averaged when the vectors L of all the domains align counter to the ˜ ϭ␬ ˜ ⌬␭ over the vector L, i.e., ͗E͘ ͗L͘•L, because of the disor- field, the magnetostriction exhibits jumplike behavior a ϳ␥ 2 ʈ dered nature of the gadolinium subsystem, does not itself 3L . For the field direction H c the magnetostriction be- contribute to the thermodynamic potential of the system. haves monotonically, in accordance with expression ͑7͒. Nevertheless, the antiferromagnetic order induced by the f -d Analysis of the magnetoelectric contributions ͑6͒ permits exchange in the gadolinium subsystem can give an effective a detailed interpretation of the behavior of the planar com- contribution to the thermodynamic potential on account of ponent of the electric polarization vector ͑Fig. 5͒.Onthe the corresponding correlations of the longitudinal ͑lying in whole, the theory ͓expression ͑6͒ and Table I͔ permits the the easy plane͒ component of the vector L˜ in the second assertion that a quadratic magnetoelectric effect should be order of perturbation theory: observed in our system. It is noteworthy here that the discus- sion of the magnetoelectric effect in the framework of the 1 ⌬⌽ ϭ␬2 ˜ 2 ͑ 2 ϩ 2 ͒ ͑ ͒ symmetry space group R32 with the invariant contributions cor ͗L ͘ LX LY , 9 ␧ ͑6͒ taken into account does not permit a satisfactory descrip- where ␧ is the characteristic mean energy of spin waves of tion of the behavior of the vertical component of the electric the gadolinium subsystem. It is seen that the correlation con- polarization vector, PC . The similarity of its behavior with ␭ tribution to the thermodynamic potential of the system will that of the magnetostriction C argues, in particular, that for be positive independently of the sign of ␬. The physical the magnetically ordered state the vector PC should be an ⌬⌽ ͑ ͒ meaning of the energy cor is that the correlations tend to invariant i.e., our system belongs to a polar class .Inthe make it unfavorable for the vector L to lie in the horizontal absence of unambiguous analysis of the change of the crystal plane. As the temperature is lowered and the antiferromag- symmetry of the system at the point of the structural phase ϭ netic vector of the gadolinium sublattice L˜ induced by the transition TC 156 K, this remains an open question. ͑ ͒ f -d exchange increases, an anisotropic contribution of oppo- It follows from expression 6 that the symmetry of the ϳ␤ site sign will appear in the thermodynamic potential of the system admits yet another magnetoelectric contribution 2 system and compete with the initial ͑easy-plane͒ one. Then at ͑of a relativistic nature͒, which is responsible for the spin- reorientation transition in an electric field. Indeed, for the the corresponding temperature TR the resultant anisotropy constant will change sign, and an ‘‘easy plane to easy axis’’ state in which the antiferromagnetic vector lies in the basal ͑ Ͼ spin-reorientation transition will occur in the system ͑it is plane for the region T TR) the transverse component of the known5 that this transition is not observed in the case of the electric field applied in the basal plane induces the appear- nonmagnetic ions RϭY, La͒. ance of a component of the vector L directed along the c axis Ͻ and, accordingly, in the temperature region T TR this con- tribution brings about a reorientation of the antiferromag- DISCUSSION OF THE RESULTS netic vector from the easy axis to the easy plane under the Our proposed expression ͑7͒ for the magnetoelastic en- influence of a transverse electric field. ergy of the system allows an easy explanation of the experi- Finally, the features of the symmetry of our system are mentally observed behavior ͑Figs. 1 and 4͒ of the magneto- such that, in principle, under certain conditions ͑in particular, striction for the magnetic field directions Hʈc and HЌc, in the presence of antiferromagnetic intralayer exchange͒ an both for the temperature region TϽ10 K and for 10 KϽT antiferromagnetic phase transition could occur at the Ne´el Ͻ38 K. Indeed, at a temperature Tр10 K, when the vector point with the formation of a fundamentally noncollinear L is directed along the c axis and the field Hʈc, at a certain structure of the triangular type, described by a two- ͑ value of the field Hcr a spin-reorientation transition occurs, dimensional representation by the basis vectors B1 and B2 and the vector L lies in the plane without changing its length. mentioned previously͒. In this case it would not require very The lower the temperature, the larger the value of the thresh- much effort to describe the spin-reorientation transition and old field. Starting from formula ͑7͒, one can conclude that at so forth in the framework of the thermodynamic approach. ␭ ϳ ␥2 that point the magnetostriction has a jump, and C ( 1 However, the realization of an easy-plane triangular AFM Ϫ␥1 2 Ͻ Ͻ 1)L . structure in the temperature interval 10 K T 38 K would At the same field orientation Hʈc but for TϾ10 K the lead, for both the cases of weak and strong easy-plane vector L lies in the easy plane, and the magnetostriction anisotropy,13 to the existence of yet another spin- behaves in a monotonic manner ͑quadratically in the field; reorientation phase transition in the high-temperature region Fig. 1͒. For the field direction HЌc in the temperature region at moderately high fields HЌc ͑completely achievable in our TϽ10 K the magnetic field H acting on the gadolinium sub- experiments͒. Since no such transformation was observed in system disrupts the mechanism of the spin reorientation that principle in our studies, we shall assume, as in Ref. 8, that is observed at temperature TR and tips the vector L over into the features of our system are such that collinear AFM or- the basal plane, but perpendicular to the field ͑the in-plane dering is realized at the Ne´el point. We complete the discus- anisotropy is extremely small͒, so that in this case the mag- sion of the properties of the thermodynamic potential of the ␭ ⌬␭ ϳ ␥2Ϫ␥1 2 ͑ ͒ netostriction C will exhibit a jump C ( 2 2)L . system 3 by calling attention to the presence of specific At temperatures TϾ10 K the vector L lies in the plane, cross contributions in it that are capable of ‘‘admixing’’ and in the general case there will be six types of domains in exchange-noncollinear components to the fundamental col- the system ͑in accordance with the three equivalent easy di- linear AFM mode both above and below the spin reorienta- rections in the plane͒, so that the mean magnetostriction over tion point. One can see that each of these noncollinear com- the sample will be equal to zero. At low magnetic fields, ponents corresponds to a different chirality ͑right or left Low Temp. Phys. 31 (8–9), August–September 2005 Kadomtseva et al. 813

‘‘handedness’’͒.14 This, in particular, attests that the magne- 3 A. M. Kalashnikova, V. V. Pavlov, R. V. Pisarev, L. M. Bezmaternykh, tooptical properties of our system will be different on differ- M. Bayer, and Th. Rasing, JETP Lett. 80, 293 ͑2004͒. 4 ent sides of the spin reorientation point. A. G. Gavriliuk, S. A. Kharlamova, I. S. Lyubutin, I. A. Troyan, S. G. Ovchinnikov, A. M. Potseluko, M. I. Eremets, and R. Boehler, JETP Lett. On the whole, the above-mentioned features of the be- 80,426͑2004͒. 5 havior of the system GdFe3(BO3)4 permit the assumption J. A. Campa, C. Cascales, E. Gutierrez-Puebla, M. A. Monge, I. Rasines, that it is a multiferroic compound. and C. Ruiz-Valero, Chem. Mater. 9, 237 ͑1997͒. 6 In closing we thank M. N. Popova and A. N. Valil’ev for Y. Hinatsu, Y. Doi, K. Ito, M. Wakeshima, and A. Alemi, J. Solid State Chem. 172, 438 ͑2003͒. steady interest in our research, and K. I. Kamilov and A. V 7 A. D. Balaev, L. N. Bezmaternykh, I. A. Gudim, V. L. Temerov, S. G. Kuvardin for making the thermal expansion measurements. Ovchinnikov, and S. A. Kharlamova, J. Magn. Magn. Mater. 258–259, This study was supported by the Russian Foundation for 532 ͑2003͒. 8 Basic Research ͑#04-02-116592͒. R. Z. Levitin, E. A. Popova, R. M. Chtsherbov, A. N. Vasiliev, M. A. Popova, E. P. Chukalina, S. A. Klimin, P. H. M. van Loosdrecht, D. Fausti, and L. N. Bezmaternykh, JETP Lett. 79, 423 ͑2004͒. * 9 E-mail: [email protected] A. I. Pankrats, G. A. Petrakovski, L. N. Bezmaternykh, and O. A. 1͒After this article went to press, the preprint of an article entitled ‘‘Evidence Bayukov, Zh. E´ ksp. Teor. Fiz., 126, 887 ͑2004͓͒JETP 99,766͑2004͔͒. 10 for structure differentiation in the iron-helicoidal-chain in GdFe3(BO3)4 ,’’ Yu. F. Popov, A. M. Kadomtseva, D. V. Belov, G. P. Vorob’ev, and A. K. by S. A. Klimin, D. Fausti, A. Meetsma, L. N. Besmaternykh, P. H. M. van Zvezdin, JETP Lett. 69, 330 ͑1999͒. Loosdrecht, and T. T. M. Palstra ͑to be published in Acta Cryst. B͒, came 11 A. N. Vasil’ev, private communication. out, which contained a detailed description of the change of the crystal 12 K. P. Belov, A. K. Zvezdin, A. M. Kadomtseva, and R. Z. Levitin, Orien- ͓ ͔ system at the point Tc ; however, taking that change into account would tational Transitions in Rare-Earth Magnets in Russian , Nauka, Moscow not alter the basic conclusions found in our paper. ͑1979͒. 13 M. F. Collins and O. A. Petrenko, Can. J. Phys. 75, 605 ͑1997͒. 14 E. A. Turov, A. V. Kolchanov, V. V. Men’shenin, I. F. Mirsaev, and V. V. Nikolaev, Symmetry and the Physical Properties of Antiferromagnets ͓in 1 D. A. Keszler, Curr. Opin. Solid State Mater. Sci. 1, 204 ͑1996͒. Russian͔, Fizmatlit, Moscow ͑2001͒. 2 M. H. Bartl, K. Gatterer, E. Cavalli, A. Speghini, and M. Bettinelli, Spec- trochim. Acta, Part A 57, 1981 ͑2001͒. Translated by Steve Torstveit LOW TEMPERATURE PHYSICS VOLUME 31, NUMBERS 8–9 AUGUST–SEPTEMBER 2005

Incommensurate magnetism in the coupled spin tetrahedra system Cu2Te2O5Cl2 O. Zaharko,* H. M. Ronnow, A. Daoud-Aladine, S. Streule, F. Juranyi, and J. Mesot

Laboratory for Neutron Scattering, ETHZ & PSI, CH-5232 Villigen PSI, Switzerland H. Berger

Institute de Physique de la Matie`re Complexe, EPFL, CH-1015 Lausanne, Switzerland P. J. Brown

Institute Laue-Langevin, 156X, F-38042 Grenoble Cedex, France ͑Submitted January 25, 2005͒ Fiz. Nizk. Temp. 31, 1068–1072 ͑August–September 2005͒ Neutron scattering studies on powder and single crystals have provided new evidence for unconventional magnetism in Cu2Te2O5Cl2 . The compound is built from tetrahedral clusters of ϭ 2ϩ S 1/2 Cu spins located on a tetragonal lattice. Magnetic ordering, emerging at TN ϭ18.2 K, leads to a very complex multi-domain, most likely degenerate, ground state, which is characterized by an incommensurate ͑ICM͒ wave vector kϳ͓0.15,0.42,1/2͔. The Cu2ϩ ␮ 2ϩ ions carry a magnetic moment of 0.67(1) B /Cu at 1.5 K and form a four-helix spin arrangement with two canted pairs within the tetrahedra. A domain redistribution is observed when Ϸ ʈ a magnetic field is applied in the tetragonal plane (Hc 0.5 T), but not for H c up to 4 T. The excitation spectrum is characterized by two well-defined modes, one completely dispersionless at 6 meV, the other strongly dispersing to a gap of 2 meV. The reason for such complex ground state and spin excitations may be of the Cu2ϩ spins within the tetrahedra, intra- and inter-tetrahedral couplings having similar strengths and strong Dzyaloshinski–Moriya anisotropy. Candidates for the dominant intra- and inter-tetrahedral interactions are proposed. © 2005 American Institute of Physics. ͓DOI: 10.1063/1.2008143͔

INTRODUCTION magnetic susceptibility and specific heat measurements5 in- ͑ ͒ dicated an onset of antiferromagnetic AF order at TN Reduced dimensionality, geometrical frustration and low ϭ18.2 K, thus confirming the importance of inter-tetrahedral spin values lead to quantum fluctuations often resulting in couplings. Recent neutron diffraction studies8 revealed the interesting new ground states and spin dynamics.1 The most details of the magnetic order: it is incommensurate and very famous examples are based on triangular units ͑triangular complex. and kagome´ lattices2͒ in two dimensions ͑2D͒ and tetrahedral In spite of considerable progress in experimental studies, clusters ͑FCC and pyrochlore lattices3͒ in 3D. The nature of the relevant intra- and inter-tetrahedral magnetic interactions the ground state in such systems is a subject of current strong in Cu2Te2O5Cl2 remain a puzzle due to the rather complex interest, especially for the extreme quantum mechanical case 3D exchange topology ͑Fig. 1͒. The intra-tetrahedral spin Sϭ1/2. interactions are mediated via the superexchange paths ϭ The copper tellurate Cu2Te2O5X2 (X Cl, Br, space Cu–O–Cu and can be described by J1 and J2 exchange con- group P-4; Ref. 4͒ belong to a new family of such com- stants. It was suggested9 according to the Goodenough rules, pounds. Their structure can be viewed as a stacking of layers that the J2 interaction should be weakly antiferromagnetic if 2ϩ of Cu4O8Cl4 clusters along c. The four Cu ions within not ferromagnetic, while J1 is antiferromagnetic and rather such a cluster, Cu1 (x,y,z), Cu2 (1Ϫx,1Ϫy,z), Cu3 (y,1 weak.9 From the band structure calculations10 it is expected Ϫ Ϫ Ϫ Ϫ x, z), and Cu4 (1 y,x, z), form an irregular tetrahe- that within the ab layers the Ja and Jd inter-tetrahedral cou- dron with two longer ͑Cu1–Cu2, Cu3–Cu4͒ and four shorter plings are substantial ͑see Fig. 1͒ and mediated by the halo- edges. The tetrahedra have a 2D square arrangement within gen p orbitals. Between the adjacent layers the tetrahedra the ab layers. interact through the super-superexchange paths The Cu2Te2O5Cl2 system attracted much attention as the Cu–O...O–Cu and the corresponding inter-tetrahedral cou- first magnetic susceptibility data fitted well a model of iso- plings are as well important. lated tetrahedra.4 The observed broad maximum between The nature of the magnetic ordering transition is also 20–30 K and a rapid drop at lower temperatures indicated a still unclear. A strong spin-lattice coupling near TN has been strength of the intra-tetrahedral coupling Jϳ38.5 K. Raman suggested from magnetic susceptibility and thermal conduc- spectroscopy, however, indicated a substantial inter- tivity studies.11 However, the infrared spectroscopy study,12 tetrahedral coupling along the c axis5,6 and the data analysis as well as high-resolution neutron diffraction study8 gave no have been performed7 in terms of a dimerized model with the evidence of any lattice distortion. two pairs of Cu2ϩ spins: Cu1–Cu2 and Cu3–Cu4. Further We present neutron diffraction and inelastic scattering

1063-777X/2005/31(8–9)/5/$26.00814 © 2005 American Institute of Physics Low Temp. Phys. 31 (8–9), August–September 2005 Zaharko et al. 815

were performed on a larger 1 cm3 crystal on TriCS (␭ ϭ1.18 and 2.4 Å͒ for three field directions Hʈa, b, and c. The inelastic neutron scattering experiment was carried out on 15 g powder on neutron time-of-flight spectrometer FOCUS (␭ϭ2.8 and 4 Å͒ at SINQ and on a large 1.5 cm3 crystal on the IN8 thermal neutron three-axis spectrometer at ILL. IN8 was configured with doubly focusing monochro- mator and analyzer witha5cmgraphite filter in k f and fixed final neutron energy of 14.7 meV.

RESULTS Diffraction ϭ ͑ ͒ Below TN 18 K tiny magnetic peaks appeared Fig. 2 ϭ FIG. 1. The xy projection of the Cu2Te2O5Cl2 crystal structure (z in DMC neutron diffraction patterns of Cu2Te2O5Cl2 , as al- Ϫ 10 1/2—1/2) with the J1 , J2 and possible intertetrahedral exchange paths. ready reported in Ref. 8. Moreover, the onset of magnetic The double-line segment is normal to the incommensurate component of the order at T can be followed from the structural peaks in the Ј 8 N wave vector k . HRPT data. The lattice constant a(b) significantly decreases with temperature above TN and changes very little below TN ͑Fig. 2, inset͒. This implies that the spin-lattice coupling is results, which supply new information on the magnetic substantial, but no changes of the crystal structure could be ground state, spin-orbit coupling and spin dynamics. We determined from the neutron patterns. hope this will provide a better starting point for theoretical To determine correctly the magnetic ground state it is 13–17 modeling. very important to elucidate the magnetic symmetry. The ͓001͔ projection of neutron diffraction pattern of a typical Cu Te O Cl single crystal is presented in Fig. 3. Up to 16 EXPERIMENTAL 2 2 5 2 magnetic satellites of the ͑0,0,0͒ reflection have been ob- High-purity powder and single crystals of Cu2Te2O5Cl2 served. The reflection set denoted by black circles arises were prepared by the halogen transport technique, us- from four arms of the star of the incommensurate ͑ICM͒ ϳ ing TeCl4 and Cl2 as transport agents. Neutron powder dif- wave vector k ͓0.150,0.422,1/2͔. The reflections denoted fraction ͑NPD͒ patterns were collected in the temperature by dotted circles correspond to the star of another wave vec- range 1.5–30 K on the DMC instrument, with a neutron tor kЈϳ͓Ϫ0.150,0.422,1/2͔. The k and kЈ vectors are not wavelength of ␭ϭ4.2 Å and on the high-resolution HRPT related by the symmetry elements of the group P-4, and instrument (␭ϭ1.889 Å) at SINQ, Villigen, Switzerland. could belong to growth crystallographic twins. Since for sev- The neutron single crystal diffraction ͑NSCD͒ experiments eral crystals studied the kЈ reflections are absent, we con- on two crystals of dimensions 2.5ϫ3ϫ5 mm and 2ϫ3.5 clude that the k and kЈ sets are independent. Interestingly, ϫ6 mm were carried out using the diffractometers TriCS at the intensity ratio between the two first magnetic reflections SINQ (␭ϭ1.18 Å) and D15 (␭ϭ1.17 Å) at the high-flux for the k and kЈ sets is different, implying that the magnetic ILL reactor, France. The NSCD with applied magnetic field structures associated with these two stars are different.

␭ϭ FIG. 2. The 18 and 4 K DMC NPD patterns of Cu2Te2O5Cl2 ( 4.2 Å), arrows point to magnetic reflections. Inset: temperature evolution of the lattice constants from HRPT NPD data (␭ϭ1.889 Å). 816 Low Temp. Phys. 31 (8–9), August–September 2005 Zaharko et al.

͓ ͔ FIG. 3. The 001 projection of the reciprocal space of a Cu2Te2O5Cl2 single crystal. The black circles correspond to the k magnetic reflections and the dotted circles—to the kЈ set; ␣ϭ0.15, ␤ϭ0.422.

We further tried to clarify if the magnetic structure is single-k or multi-k. In the case of a single-k magnetic struc- ture the k(kЈ) set contains contributions of four configura- tion domains. The configuration domains must all have the same structure but possibly different populations. Each of them could have two 180 deg domains and two chiral do- mains. In the case of multi-k structure, the four arms of the star build one magnetic structure. It is possible to distinguish the single-k or multi-k cases by studying the variation of magnetic intensities as a func- tion of an applied magnetic field H. We performed such study for fields along a, b, and c crystal directions. For H ␻ Ϫ Ϫ Ϫ ͑ ͒ c hkl FIG. 4. The TriCS scans of the ( 0.15, 0.42, 1/2) a and (0.42, along the intensities of the magnetic reflections with Ϫ0.15,1/2) ͑b͒ magnetic reflections (␭ϭ2.4 Å) at Hϭ0 T and Hϭ0.5 T lϾ0 increase and for lϽ0 they decrease, but all reflections along a. persist up to 4 T. This implies a change of the spin arrange- ment with respect to the zero-field magnetic structure, but without domain redistribution and/or meta-magnetic transi- should be considered. Since the available number of experi- ʈ ϳ tion. For H a a transition occurs at Hc 0.5 T: the intensi- mental observations does not allow to refine with confidence ties of the (Ϯ␣,Ϯ␤,Ϯl/2) reflections vanish, while the all of them, we imposed physically sound constraints: iden- ϩ magnetic reflections of the type (Ϯ␤,Ϯ␣,Ϯl/2) change tical moment value for all four independent Cu2 ions and a their intensities smoothly ͑Fig. 4͒. Flipping the field to Ϫa circular envelope of the helices. This lowered the number of results in the same behavior. Switching off the field restores independent parameters to 12. partly the vanished peaks. A similar picture is observed for Very good fits were obtained for a model ͑Fig. 5͒ in ϩ Hʈb, but now the (Ϯ␤,Ϯ␣,Ϯl/2) reflections vanish at which the four Cu2 moments in each tetrahedron form two ϳ0.5 T. Our results imply that 0.5 T applied in the tetrago- nal plane is enough to depopulate the domains with the propagation vector nearly normal (90Ϯ10 deg) to the field direction. This supports the idea that the magnetic structure is a single-k and not a multi-k structure. The model for the kЈ magnetic structure has been re- cently elaborated in Ref. 8. The only symmetry constraint is imposed by the commensurate component of the wave vec- tor. It implies that the ab layers of tetrahedra alternating along c carry oppositely oriented spins. The magnetic mo- ments of the four Cu2ϩ ions in the unit cell can be indepen- dent and a generalized helix characterizing the spin arrange- ment of each of them in the crystal is expressed as ϭ " ϩ␺ ϩ " ϩ␺ S j A cos͑k rj ͒ B sin͑k rj ͒. The spin components are modulated by the wave vector k, and rj is the radius vector to the origin of the jth unit cell. A and B are orthogonal vectors, which define the magnitude and direction of the axes of the helix, while ␺ defines its FIG. 5. The xy projection of the Cu2Te2O5Cl2 magnetic structure with the phase. For the most general case 27 independent parameters spin tetrahedra at the zϭ0 layer. Low Temp. Phys. 31 (8–9), August–September 2005 Zaharko et al. 817

Raising the temperature above TN , spectral weight shifted downwards to a broad quasi-elastic peak. For single crystal neutron spectroscopy a single crystal was aligned with ͑3,1,0͒ and ͑0,0,1͒ in the horizontal scatter- ing plane, such that ͑0.42,0.15,1/2͒ and the equivalent mag- netic Bragg peaks were accessible given the wide (ϳ5 deg) vertical resolution. Scans performed along the Q ϭ(h,h/3,3/2), ͑0.45,0.15,1͒, and ͑0,0,1͒ directions revealed two well defined excitation modes. One mode is completely dispersionless at 6.0 meV and shows no variation in intensity as a function of Q. The other mode displays strong disper- sion along both accessible directions from a maximum en- ergy close to the flat mode down to a minimum energy gap of 2.1 meV at the same positions in Q as the ICM magnetic Bragg peaks.

FIG. 6. Inelastic powder neutron scattering of Cu2Te2O5Cl2 integrated be- While more detailed modeling of the excitation spectrum Ϫ1 Ϫ1 tween 0.8 Å and 3.3 Å for four temperatures between 2 and 30 K. With is under way, several important conclusions can be read di- ϭ incident energy Ei 10.4 meV inelastic focusing was adjusted to provide optimal resolution at 5 meV of 0.45 meV ͑full-width-half-maximum͒. rectly off the figure. If the system would be a collection of very weakly coupled tetramers, one would expect a series of essentially dispersionless modes. The strong dispersion ob- canted pairs: Cu1–Cu2 and Cu3–Cu4. The two spins of the served in our measurements imply strong inter-tetrahedral pair share a common ͑A, B͒ plane, but the associated helices coupling both within the ab plane and along the c axis. have different phases ␺. The difference between the phases Secondly, a classical ordered magnetic structure with con- defines the canting angle between spins of the pair ␣. The tinuous symmetry of the order parameter should have gap- ␣ ϭ canting angle for the first pair is 12 38(6) deg and for the less spin waves emerging from the magnetic Bragg peaks. ␣ ϭ second pair 34 111(14) deg. The amplitude of the mag- The rather large energy gap must involve strong anisotropy 2ϩ ␮ netic moment carried by each Cu ion is 0.67(1) B at 1.5 terms in the Hamiltonian, whose origin remains to be deter- K. It is interesting that the vector sum of the spins of one pair mined. is the same for all tetrahedra in the crystal (m12 ϭ ␮ ϭ ␮ 1.27(6) B , m34 0.76(14) B), while the local magnetic SUMMARY moment of the tetrahedra is not zero and changes from one The presented results of neutron diffraction and inelastic unit cell to another. For an isotropic exchange the spin state neutron scattering evidence new details of the magnetic or- of the tetrahedra would be zero, so our model might suggest dering in the Cu Te O Cl compound. The idea of a single- that the two intra-tetrahedral couplings J and J are differ- 2 2 5 2 1 2 k magnetic structure is strongly supported by the observed ent and that the J interaction is stronger. 2 magnetic domains redistribution in an applied magnetic field. Such a particular magnetic ground state might be a con- The presence of two different k and kЈ structures suggests sequence of one dominant interaction or result from contri- that a number of ground states with equal or close energies butions of several inter-tetrahedral couplings of similar might exist. The discovered relation between the incommen- strengths. We tried to attribute the observed reciprocal wave surate component of the wave vector and the inter-tetrahedral vector kЈϭ(Ϫ0.15,0.42,1/2) to some specific exchange coupling J invites a theoretical revision of the Cu Te O X inter-tetrahedral path in the structure and found a simple cor- a 2 2 5 2 system. The peculiar features of the spin excitation spectrum relation not to kЈ, but to the vector (Ϫ0.15,Ϫ0.58,0). As deserve further study. this vector is related by a lattice translation to the vector The work was performed at SINQ, Paul Scherrer Insti- (Ϫ0.15,0.42,0), the two vectors mean a different choice of tute, Villigen, Switzerland, at ILL reactor, Grenoble, France. unit cell of the same magnetic structure. The ICM compo- We thank Prof. F. Mila, Dr. M. Prester, and Prof. A. Furrer nent is orthogonal to a specific set of planes containing the ϩ for fruitful discussions and Swiss NCCR research pool Cu2 ions. One of these planes, presented by a double-line MANEP of the Swiss NSF for financial support. segment in Fig. 1, passes through the Cu2–Cu4 ions of the adjacent tetrahedra. This corresponds to the J coupling, a * which according to Ref. 10 is substantial and is mediated by E-mail: [email protected] the halogen orbitals. Based on these considerations we sug- gest that Ja is the dominant inter-tetrahedral coupling. 1 Quantum Magnetism, Vol. 645 of Lecture Notes in Physics, Springer- Verlag, Berlin, Heidelberg ͑2004͒. INELASTIC SCATTERING 2 F. Mila, Eur. J. Phys. 21, 499 ͑2000͒. 3 A. P. Ramirez, ‘‘Geometrical frustration,’’ in Handbook of Magnetic Ma- Several inelastic neutron scattering studies have been terials, Vol. 13, Elsevier Science ͑2001͒. performed to investigate the excitation spectrum of the 4 M. Johnsson, K. W. Tornroos, F. Mila, and P. Millet, Chem. Mater. 12, Cu Te O Cl system. Spectra from powder revealed in the 2853 ͑2000͒. 2 2 5 2 5 ordered phase below T a spherically averaged density of P. Lemmens, K.-Y. Choi, E. E. Kaul, C. Geibel, K. Becker, W. Brenig, R. N Valenti, C. Gross, M. Johnsson, P. Millet, and F. Mila, Phys. Rev. Lett. 87, states extending to a maximum at 6 meV, above which no 227201 ͑2001͒. significant scattering was detected up to 15 meV ͑Fig. 6͒. 6 C. Gross, P. Lemmens, M. Vojta, R. Valenti, K.-Y. Choi, H. Kageyama, Z. 818 Low Temp. Phys. 31 (8–9), August–September 2005 Zaharko et al.

Hiroi, N. M. Mushnikov, T. Goto, M. Johnsson, and P. Millet, Phys. Rev. 12 A. Perucchi, L. Deiorgi, H. Berger, and P. Millet, Eur. Phys. J. B 38,65 B 67, 174405 ͑2003͒. ͑2004͒. 7 J. Jensen, P. Lemmens, and C. Gross, Europhys. Lett. 64, 689 ͑2003͒. 13 W. Brenig and K. W. Becker, Phys. Rev. B 64, 214413 ͑2001͒. 14 8 O. Zaharko, A. Daoud-Aladine, S. Streule, J. Mesot, P.-J. Brown, and W. Brenig, Phys. Rev. B 67, 64402 ͑2003͒. 15 ͑ ͒ H. Berger, Phys. Rev. Lett. 93, 217206 ͑2004͒. K. Totsuka and H.-J. Mikeska, Phys. Rev. B 66, 54435 2002 . 16 9 M. H. Whangbo, H. J. Koo, and D. Dai, J. Solid State Chem. 176, 417 V. N. Kotov, M. E. Zhitomirsky, M. Elhajal, and F. Mila, Phys. Rev. B 70, 214401 ͑2004͒. ͑2003͒. 17 V. N. Kotov, M. E. Zhitomirsky, M. Elhajal, and F. Mila, J. Phys.: Con- 10 R. Valenti, T. Saha-Dasgupta, C. Gros, and H. Rosner, Phys. Rev. B 67, dens. Matter 16, S905 ͑2004͒. 245110 ͑2003͒. 11 M. Prester, A. Smontara, I. Zˇ ivkovic´, A. Bilusˇic´, D. Drobac, H. Berger, This article was published in English in the original Russian journal. Repro- and F. Bussy, Phys. Rev. B 69, 180401͑R͒͑2004͒. duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 31, NUMBERS 8–9 AUGUST–SEPTEMBER 2005

Antiferromagnet-ferromagnet phase transition in lightly doped manganites I. O. Troyanchuk* and V. A. Khomchenko

Institute of Solid State and Physics, NAS, 17 P. Brovka St., Minsk 220072, Belarus V. V. Eremenko and V. A. Sirenko†

B. Verkin Institute for Low Temperature Physics and Engineering of the National Academy of Sciences of Ukraine, 47 Lenin Ave., Kharkov 61103, Ukraine H. Szymczak

Institute of Physics, PAS, 32/46 Lotnikow Str., Warsaw 02-668, Poland ͑Submitted February 18, 2005͒ Fiz. Nizk. Temp. 31, 1073–1080 ͑August–September 2005͒

Magnetic and structural phase diagrams of the La0.88MnOx ,La1ϪxSrx(Mn1Ϫx/2Nbx/2)O3 , Nd1ϪxCaxMnO3 , and Bi1ϪxCaxMnO3 series, constructed on the basis of x-ray, neutron powder diffraction, Young’s modulus, magnetization and resistivity measurements, are presented. It is shown that the main factor controlling the antiferromagnet-ferromagnet phase transition in the manganites is a type of an orbital state. The results are discussed in the framework of structurally driven magnetic phase separation model. © 2005 American Institute of Physics. ͓DOI: 10.1063/1.2008144͔

I. INTRODUCTION ferromagnet phase transitions can occur going through a mixed state of phases with different orbital dynamics. Mixed-valence manganites with a perovskite structure The recent magnetic phase diagrams of the are the model objects for the physics of strongly correlated La1ϪxSrxMnO3 and La1ϪxCaxMnO3 systems were con- electronic systems. The interest in the study of manganites is structed assuming a homogeneous canted magnetic state in a due to a variety of phase states and transitions and intrinsic low doping range.5,6 On the other hand, there are numerous correlation of the crystal structure, magnetic, and transport experimental data which indicate the existence of phase properties. The nature of the interplay between the crystal separation in manganites. The results of nuclear magnetic structure, magnetic, and transport properties of manganites is resonance,7,11 neutron diffraction,11,12 muon spin still a matter of discussion in spite of numerous investiga- relaxation,13 x-ray absorption,14 scanning tunneling 15 16 tions. Several models have been proposed to explain a mag- spectroscopy, and electron microscopy experiments give netic state evolution under hole doping as well as a metal- evidence of magnetic and structural inhomogeneities, but the transition at the Curie point. In the double- driving force of magnetic phase separation in manganites is exchange model of Zener, simultaneous ferromagnetic and still not fully clear. In order to contribute to the solution of metallic transitions have been qualitatively explained by the this problem we have investigated the features of the fact that electrons tend to move between Mn3ϩ and Mn4ϩ antiferromagnet-ferromagnet phase transition in low-doped ions having the same spin orientation, therefore electron de- La0.88MnOx ,La1ϪxSrx(Mn1Ϫx/2Nbx/2)O3 ,Nd1ϪxCaxMnO3 , localization favors the ferromagnetic order.1 More recently and Bi1ϪxCaxMnO3 manganites. Millis et al. pointed out that double exchange alone cannot account for many of the experimental results.2 They showed II. RESULTS AND DISCUSSION that a Jahn-Teller-type electron-phonon coupling should play an important role in explanation of the colossal magnetore- A. La0.88MnOx system sistance effect. Another mechanism of antiferromagnet- Tentative magnetic phase diagram of the La0.88MnOx ferromagnet phase transitions in manganites was proposed ϫ(2.82рxр2.96) manganites is shown in Fig. 1. The most by Nagaev.3 He assumed that the intermediate phase can be strongly reduced sample La0.88MnO2.82 is an antiferromagnet described as a inhomogeneous magnetic state driven by an with a Ne´el temperature of 140 K. Its properties are found to electronic phase segregation. In this scenario the ferromag- be similar to the properties of stoichiometric LaMnO3 . Both netic regions contain an excess of holes and are metallic. compounds have very close unit cell parameters, the same Goodenough et al. argued that the magnetic properties of magnetization value, and close temperatures of both mag- 4 ϳ ϳ manganites were determined by the type of orbital state. netic (TN 140 K) and orbital orderings (TOO 750 K). According to the rules for 180° superexchange, if the elec- The existence of orbital ordering in the A-type antiferromag- tronic configuration correlates with vibrational modes, netic structure of La0.88MnO2.82 is corroborated by neutron 3ϩ 2Ϫ 3ϩ 17 Mn uO uMn interactions are antiferromagnetic in diffraction measurements. With increasing oxygen content case of the static Jahn-Teller effect and ferromagnetic when up to the xϭ2.85 sample, the magnetic and orbital ordering the Jahn-Teller effect is dynamic. Thus, antiferromagnet- temperatures lower while the magnetization increases

1063-777X/2005/31(8–9)/6/$26.00819 © 2005 American Institute of Physics 820 Low Temp. Phys. 31 (8–9), August–September 2005 Troyanchuk et al.

approximately 650 K. The DTA measurements revealed the release of in the range 650–730 K. Neutron dif- fraction data indicate the coexistence of orbitally ordered OI and orbitally disordered O phases at Tϭ700 K. Another thermal anomaly connected with the transition to the mono- clinic phase is observed in the temperature range 915рT р960 K. With the increase of the oxygen content to x ϭ2.83, the temperatures of both orbital order-disorder and orthorhombic-monoclinic phase transitions decrease signifi- cantly. The range of coexistence of OI and O phases becomes broader, while the width of the anomaly associated with the temperature-induced orthorhombic-monoclinic transition re- mains practically constant. Starting from the xϭ2.84 sample, the differential thermal analysis does not show any signifi- cant heat effect, which could be interpreted as a transition to a pure orbitally disordered state; however, the anomaly re- lated to the transition from an orthorhombic to a monoclinic phase remains well pronounced. Neutron diffraction data р р FIG. 1. Magnetic phase diagram of the La0.88MnOx (2.82 x 2.96) sys- coupled with Young’s modulus measurements indicate the tem. Orth—orthorhombic crystal structure, M—monoclinic crystal struc- existence of predominantly static Jahn-Teller distortions at ture; PI, AFI, FI, and FM—paramagnetic insulating, antiferromagnetic insu- lating, ferromagnetic insulating, and ferromagnetic metallic states, room temperature and two-phase character of the crystal respectively. Areas 1 and 2 correspond to the concentration regions where an structure above Tϳ470 K. Inhomogeneous structural states antiferromagnetic or ferromagnetic phase predominates, respectively. are observed up to 650 K. Above this temperature the mono- clinic phase is stabilized. A further increase of the oxygen concentration leads to the broadening and gradual disappear- slightly. Results of the neutron diffraction measurements car- ance of the anomaly which relates to the transition to the ried out for the xϭ2.84 sample confirm the appearance of a orbitally ordered state. The neutron diffraction study per- ferromagnetic component. A further increase of the oxygen formed for the La0.88MnO2.87 compound indicates that the concentration leads to a significant enhancement of the fer- value of the MnO6 octahedron distortion increases with de- romagnetic contribution. The transition temperature to the creasing temperature to 200 K. However, even in the case of ϭ paramagnetic state begins to increase and the transition be- T 200 K, where the worst agreement factors for one-phase comes broader. Neutron diffraction data obtained for the x structural model have been observed, the introduction of the ϭ2.87 sample indicate that ferromagnetic coupling becomes second orthorhombic phase was unsuccessful. Apparently, predominant. No long-range antiferromagnetic order has even at 200 K, the orbitally ordered clusters are still too I been observed for this compound. At the same time, the re- small and separate to distinguish the O phase in the diffrac- fined magnetic moment is lower than that expected for the tion experiment. The temperature of orthorhombic- full spin arrangement. Besides, the relatively large magnetic monoclinic phase transition gradually decreases as the oxy- anisotropy at low temperature assumes the presence of an gen content increases and starting from the xϭ2.91 sample, anisotropic magnetic coupling which differs from the isotro- the monoclinic phase is stabilized ͑Fig. 2͒. It is necessary to pic ferromagnetic one. This can be attributed to existence of mention that the x-ray and neutron diffraction experiments either short-range antiferromagnetic clusters or a spin-glass can reveal a two-phase structural state rather in the case of phase. No pronounced thermomagnetic irreversibility indi- macroscopic structural phase separation. In the cases of local cating the anisotropic magnetic interactions is observed start- structural inhomogeneities or nanometer scale structural ing from the xϭ2.92 sample. The values of magnetization clusters, these experiments give only an average picture of a estimated for the monoclinic compounds are close to those structural state.18 Thus, the correlation between the orbital expected for full spin alignment. The ground state of all the state and magnetic properties of the La0.88MnOx manganites orthorhombic compounds 2.82рxр2.90 is insulating. It is prominent. The static Jahn-Teller distortions are respon- should be noted that the appearance of metallic conductivity sible for the A-type antiferromagnetic structure, while dy- does not coincide with the transition to monoclinic phase. namic orbital correlations lead to ferromagnetism. Simultaneous first-order magnetic transition and metal- It is worth noting that there are two alternative models of у insulator transition at TC are observed for x 2.92 com- orbital state corresponding to ferromagnetic ordering in man- pounds. ganites: 3D dynamic d3z2Ϫr2 orbital correlations and stag- A strong correlation between the magnetic and structural gered ordering of d3z2Ϫr2 and dx2Ϫy2 orbitals. Neutron dif- р р properties of La0.88MnOx (2.82 x 2.96) manganites is ob- fraction studies have shown that LaMnO3 undergoes a served. The hypothetical structural phase diagram of structural transition from OI-orthorhombic to р р ϭ 19 La0.88MnOx (2.82 x 2.96) constructed using x-ray, neu- O-orthorhombic phase at TJT 750 K. The MnO6 octahe- tron diffraction, Young’s modulus, resistivity, and DTA data dron in the O-orthorhombic phase becomes nearly regular, 19 is shown in Fig. 2. For La0.88MnO2.82 , the sharp anomalies i.e., the orbital ordering disappears. However, x-ray absorp- of the Young’s modulus and resistivity are associated with tion near the edge structure and the extended x-ray absorp- the removal of cooperative orbital ordering; it is observed at tion fine structure at the MnK-edge measurements have re- Low Temp. Phys. 31 (8–9), August–September 2005 Troyanchuk et al. 821

͑ FIG. 3. Magnetic phase diagram for La1ϪxSrx(Mn1Ϫx/2Nbx/2)O3 series Ais antiferromagnet, F is ferromagnet, P is paramagnet, SG is ; OI and O are orbitally ordered and orbitally disordered phases, respectively͒.

romagnetic order. Below we show that this assumption is р FIG. 2. Crystal structure phase diagram of the La0.88MnOx (2.82 x correct. I р2.96) system. O , O, and M are orbitally ordered orthorhombic, orbitally The second possibility lies in description of the orbital disordered orthorhombic, and monoclinic phases, respectively. Areas 1 and 2 correspond to the concentration regions where the static Jahn-Teller dis- state as a hybridization of the d3z2Ϫr2 and dx2Ϫy2 orbitals as 2 2 2 2 tortions or dynamic orbital correlations predominate, respectively. cos(␪/2)͉3z Ϫr ͘Ϯsin(␪/2)͉x Ϫy ͘. Such an orbital order- ing was recently proposed experimentally and theoretically in the ferromagnetic insulating phase of La0.88Sr0.12MnO3 23–25 and Pr0.75Ca0.25MnO3 . The difficulties in determination vealed that the MnO6 octahedrons in LaMnO3 remain of priority of the present models are conditioned by the fact Ͼ 20 3ϩ tetragonally distorted at T TJT . The empty Mn elec- that staggered ordering of d3z2Ϫr2 and dx2Ϫy2 orbitals can tronic d states were shown to be unaltered through the Jahn- exhibit itself in experiments in the same way as 3D dynamic Teller transition. The lowest energy for the e electron cor- g d 2Ϫ 2 orbital correlations. responds to the three possible distortions giving to three 3z r degenerate vibronic states, d 2Ϫ 2, d 2Ϫ 2, and d 2Ϫ 2, being x r y r z r B. La Sr Mn Nb O system the electronic orbitals of the vibronic state. The thermally 1Àx x„ 1ÀxÕ2 xÕ2… 3 excited electron jumps between these states above TJT and is A hypothetical magnetic phase diagram of localized in an ordered state below TJT . The orbital ordering La1ϪxSrx(Mn1Ϫx/2Nbx/2)O3 is shown in Fig. 3. The parent proposed for LaMnO3 arises then from the ordering of the LaMnO3 compound shows the spontaneous magnetization ͑ ␮ local Jahn–Teller distortions. The high temperature O- value at 5 K corresponding to magnetic moment of 0.07 B orthorhombic͒ phase can be described as a dynamical locally per Mn3ϩ ion. The Ne´el point, where spontaneous magneti- distorted phase with the strong antiferrodistortive first- zation develops, is 143 K. According to Ref. 26 the sponta- neighbor coupling.20 neous magnetization has a relativistic nature. Substitution of A similar situation seems to be observed for Mn with Nb leads to an enhancement of the spontaneous Mn4ϩ-doped manganites. The atomic pair-density function magnetization whereas the temperature of transition into р р of La1ϪxSrxMnO3 manganites (0 x 0.4), obtained by paramagnetic state slightly decreases. In accordance with the pulsed neutron diffraction, indicates the existence of tetrago- magnetization data the La0.8Sr0.2(Mn0.9Nb0.1)O3 and nally distorted MnO6 octahedrons even in the rhombohedral La0.7Sr0.3(Mn0.85Nb0.15)O3 samples are ferromagnets with ␮ metallic phase, when the crystallographic structure shows no the magnetic moment per chemical formula around 2.3 B 21 ␮ JT distortions. This is possible only in the case of the dy- and 2.6 B , respectively. Neutron diffraction study has re- 3ϩ namic orbital correlations described above. One can assume vealed the magnetic moment of Mn in the parent LaMnO3 4ϩ ␮ 27 that when one puts non-Jahn–Teller Mn ions in the back- antiferromagnetic compound to be close to 3.5 B , whereas 3ϩ 3ϩ 5ϩ ground of the Mn ions, the eg orbitals of all the Mn ions Nb is a diamagnetic ion, and hence the expected moment 4ϩ ␮ surrounding the localized hole (Mn ) tend to be directed should be close to 3 B per formula unit, in rather good towards it, forming an orbital .22 Due to the strong agreement with the observed value. The Nb doped sample antiferrodistortive Mn3ϩ first-neighbor coupling,20 dynamic (xϭ0.3) has a well-defined Curie point: 123 K. Both Curie correlations of the d3z2Ϫr2 orbitals should arise. point and spontaneous magnetization start gradually to de- According to the rules for 180° superexchange the dy- crease when the Nb content exceeds 15% of the total number namic orbital correlations lead to ferromagnetic interaction of sites in the manganese sublattice. The magnetic state between the Mn3ϩ ions.4 Hence, one can expect that ferro- changes cardinally as the concentration of niobium reaches magnetism in manganites can arise even in the absence of 25%. We have observed the magnetic susceptibility of the Mn4ϩ ions, if only the JT effect is dynamic. For instance, xϭ0.5 sample dramatically decreases. ZFC magnetization Mn substitution with non-Jahn-Teller diamagnetic Nb5ϩ, shows a peak at 30 K. Below this temperature FC magneti- Al3ϩ,Sc3ϩ, etc, ions should result in the appearance of fer- zation practically does not change. Taking into account the 822 Low Temp. Phys. 31 (8–9), August–September 2005 Troyanchuk et al.

state at first decreases and then around xϭ0.1 increases. We have observed two magnetic phase transitions in the range 0.06рxр0.1 as temperature decreases. The Nd1ϪxCaxMnO3 solid solutions contain two types of magnetically active sublattices: and manganese ones. At first we discuss the Nd contribution into magnetic properties. The f - f exchange interaction in rare-earth sublat- tice is as a rule rather weak in comparison with d-d interac- tion between manganese ions. One can expect that neody- mium magnetic moments should order as a result of f -d exchange interactions between neodymium and manganese sublattices. The study of magnetic properties of Nd1ϪxCaxMnO3 samples confirm this viewpoint. According to neutron diffraction data the magnetic moments of neody- mium ions start to be ordered slightly below TN . The mag- FIG. 4. Magnetic phase diagram of the Nd Ϫ Ca MnO low-doped manga- ␮ 1 x x 3 netic moment of the Nd ion is about 1.2 B at 2 K and is nites: wF is weak ferromagnet, F is ferromagnet, P is paramagnet, Teff is the directed opposite to weak ferromagnetic vector in NdMnO , effective temperature of the reorientational phase transition. 3 while in the sample xϭ0.12 the orientation of the Nd and Mn magnetic moments is the same. In the range 0.06рx р0.10 metamagnetic behavior was observed in large mag- character of M(H) dependence we have concluded that the netic fields ͑triple hysteresis loops with a negative remanent ϭ sample x 0.5 can be considered as spin glass with T f magnetization͒. ϭ30 K. We can explain the collapse of long range ferromag- According to our hypothesis, samples in the range 0.06 netic ordering by a diamagnetic dilution of the Mn sublattice. рxр0.10 consist of antiferromagnetic ͑weak ferromagnetic͒ According to resistivity versus temperature measurements and ferromagnetic phases which are exchange coupled at the La1ϪxSrx(Mn1Ϫx/2Nbx/2)O3 samples are semiconductors. Be- boundary. The neodymium sublattice in both weak ferromag- low the Curie point a large value of the magnetoresistance is netic and ferromagnetic phases orders nearby the Ne´el point observed. ͑Curie point͒. However, the orientation of neodymium mag- The results presented here deal with the facts that the netic moments in these two phases is different: f -d exchange 3ϩ 5ϩ Nb-doped La1ϪxSrx(Mn1Ϫx/2Nbx/2 )O3 samples enriched is positive for the ferromagnetic phase, whereas it is negative ϩ with Mn3 ions are ferromagnetic and show a large magne- in weak ferromagnetic phase. The ferromagnetic phase toresistance. It is worth noting that the possibility of the ex- strongly affects magnetic properties of weak ferromagnetic istence of ferromagnetic ordering in the manganites, despite phase due to exchange coupling at the boundary. This inter- ϩ the absence of Mn3 ions, reject the double exchange and action may induce a reorientational transition from antipar- the electronic phase separation concepts. The result obtained allel orientation of neodymium moments and weak ferro- indicates an important role of ferromagnetic superexchange magnetics vector to parallel one. We believe that nearby via oxygen scenario of magnetic interactions in manganites. certain temperature the ground state of the Nd3ϩ ions be- According to the superexchange mechanism the comes degenerate because opposite contributions from ex- ϩ ϩ ϩ ϩ Mn3 –O–Mn3 and Mn3 –O–Mn4 180° magnetic inter- change coupled ferromagnetic and weak ferromagnetic actions are strongly ferromagnetic for the orbitally disor- phases at Nd site become equal. According to theoretical ϩ ϩ dered state whereas the Mn4 –O–Mn4 ones are strongly consideration this state should be unstable thus leading to antiferromagnetic.4 The Curie point associated with magnetic structure transformation.29 A neutron diffraction ϩ ϩ Mn3 –O–Mn3 positive superexchange may be close to study carried out for the xϭ0.08 sample is in agreement with room temperature for manganites with perovskite structure this interpretation of the low-temperature phase transition. ϩ because our samples contain diamagnetic Nb5 ions, which On the basis of magnetization data we propose an H-T should strongly decrease the Curie point. The stoichiometric magnetic phase diagram of the Nd0.92Ca0.08MnO2.98 com- 3ϩ compound LaMn O3 also shows ferromagnetic interactions pound ͑Fig. 5͒. Depending on the prehistory in the wide ϩ between Mn3 ions when cooperative Jahn-Teller distortions range of magnetic field the phases with parallel or antiparal- have vanished at Tϭ750 K. The orbital ordering changes the lel orientation of neodymium and manganese sublattices in character of superexchange magnetic interactions, which in weak ferromagnetic phase can be realized. One can see that the orbitally ordered state become anisotropic.4,28 the value of magnetic field required for the change of relative orientation of the Nd and Mn magnetic moments in the weak ferromagnetic phase increases as temperature rises. The width of a field range in which the hysteresis is observed C. Nd1ÀxCaxMnO3 system practically does not depend on a temperature. This type of The hypothetical magnetic phase diagram of the magnetic phase diagram is in agreement with crossover of Nd1ϪxCaxMnO3 system at low Ca doping level is presented energy sub-levels of Nd ions. in Fig. 4. Neutron diffraction shows that the samples with xϽ0.08 consist mainly of antiferromagnetic phase while at D. Bi Ca MnO system xу0.08 the ferromagnetic component dominates. Under hole 1Àx x 3 doping the temperature of the transition into paramagnetic Figure 6 presents a magnetic phase diagram of the Low Temp. Phys. 31 (8–9), August–September 2005 Troyanchuk et al. 823

ferromagnetic state was observed to occur in the rare-earth manganites Sm1ϪxBaxMnO3 and Y1ϪxCaxMnO3 (xϳ0.12).32,33 It should be pointed out that at approximately this concentration of rare-earth ions, the ferromagnetic-spin glass transition takes place in Bi1ϪxCaxMnO3 . There is more than one opinion on the nature of ex- change interactions in manganites. The antiferromagnetic state certainly forms through oxygen-mediated superex- change interactions of the type Mn– O– Mn. Most research- ers believe that the ferromagnetic state in manganites is cre- ated through double exchange, i.e., via direct carrier transfer between various lattice sites. In order for such an exchange mechanism to operate, manganese ions in different valence states must be present and the electrical conductivity must be high. The presence of manganese ions of different valencies is not a sufficient condition for high electrical conductivity;

FIG. 5. The H-T magnetic phase diagram for the Nd0.92Ca0.08MnO2.98 com- indeed, the 3d orbitals of manganese and the 2p orbitals of pound. oxygen should also overlap strongly. It is believed that this parameter is controlled by the Mn– O– Mn bond angle.4,30 The larger the lanthanide ion, the larger should be the Bi1ϪxCaxMnO3 manganites. As the calcium content in the Mn– O– Mn angle, the wider the 3d band, and, accordingly, Bi1ϪxCaxMnO3 system increases, the latter passes through the higher the magnetic ordering temperature and the electri- three different magnetic states, namely, ferromagnetic (x cal conductivity. It was observed that the magnetic state of р р р 0.1), spin-glass (0.15 x 0.25), and antiferromagnetic the manganites also depends on the difference between the Ͼ (x 0.25). In the case of antiferromagnetic compositions, ionic radii of the rare-earth and the lanthanide ions. A large the magnetic-ordering and structural-transformation tem- difference between the radii lowers, as a rule, the magnetic peratures vary only weakly within the concentration interval ordering temperature as a result of competition between vari- from xϭ0.25 to 0.6. The ferromagnetic ordering in BiMnO 3 ous exchange interactions characterized by a large difference is most likely due to cooperative ordering of the d 2Ϫ 2 x y in the Mn– O– Mn angles. This is why the spin-glass state orbitals.30,31 With orbital ordering of this type, according to sets in in the Sm Ϫ Ba MnO system, wherein the average the Goodenough–Kanamori rules, ferromagnetic ordering 1 x x 3 radius of the Sm and Ba ions is far larger than that between becomes more energetically favorable than antiferromagnetic 32,33 ordering. We may recall that rare-earth manganites exhibit the Y and Ca cations in the Y1ϪxCaxMnO3 system. How- ever, in all rare earth manganites, the Mn3ϩ –O–Mn4ϩ ex- orbital ordering of the dz2 type, which stabilizes the A-type 4 change coupling in the orbitally disordered phase is appar- antiferromagnetic structure. Orbital disorder in BiMnO3 sets in, apparently, at a fairly high temperature, near 760 K. Re- ently ferromagnetic. The Mn– O– Mn angles in placement of bismuth ions by calcium results in the forma- bismuthbased manganites are fairly large, which is supported 31 tion of quadrivalent manganese ions, which should be ac- by structural studies and the quite high Curie temperature companied by destruction of orbital ordering due to the of BiMnO3 . Hence, in the case of an orbitally disordered appearance of non-Jahn-Teller Mn4ϩ ions in the lattice. phase, one can expect the ferromagnetic part of exchange However, the orbitally disordered phase in manganites interactions to be dominant, which is at odds with experi- should be ferromagnetic,4,28 whereas we observed a state of ment. Therefore, we believe that, in contrast to the rare-earth the spin-glass type. A direct transition from the antiferromag- manganites, no orbitally disordered phase forms in the р р netic to the spin-glass phase without passing through the BiMnO3 system in the concentration interval 0.1 x 0.3. The spin-glass state forms in the Bi1ϪxCaxMnO3 system most likely as a result of competition between ferromagnetic interactions in BiMnO3-type clusters and antiferromagnetic coupling in clusters in which the Mn3ϩ orbitals are frozen in random orientations. As the Ca2ϩ concentration increases, a new type of antiferromagnetic clusters, apparently due to charge ordering, appears. The existence in Bi0.75Ca0.25MnO3 of large clusters, charge-ordered in a similar way to those in Bi0.5Ca0.5MnO3 , is suggested in studies of its elastic proper- ties. Despite the presence of the spin-glass-type ground state, there is a certain fraction of states characterized by short- range order of the type of a charge-ordered phase, which is indicated by the fact that the Young modulus minima for the xϭ0.25 and 0.35 compositions are close in temperature. We FIG. 6. Magnetic phase diagram of the Bi1ϪxCaxMnO3 manganites: A is antiferromagnet, F is ferromagnet, P is paramagnet, SG is spin glass, CO is believe that the extremely high stability of the orbitally and charge-ordered state. charge-ordered states in bismuth-based manganites derives 824 Low Temp. Phys. 31 (8–9), August–September 2005 Troyanchuk et al. from the strongly anisotropic character of the Bi– O covalent J. Hejtmanek, K. Dorr, and M. Sahata, Phys. Rev. B 67, 094403 ͑2003͒. 11 bonding. P. A. Algarabel, J. M. De Teresa, J. Blasco, M. R. Ibarra, Cz. Kapusta, M. Sikora, D. Zajac, P. C. Riedi, and C. Ritter, Phys. Rev. B 67, 134402 ͑2003͒. CONCLUSIONS 12 M. Hennion, F. Moussa, G. Biotteau, J. Rodriguez-Carvajal, L. Pinsard, and A. Revcolevschi, Phys. Rev. Lett. 81, 1957 ͑1998͒. The magnetic and structural phase diagrams of 13 R. H. Heffner, L. P. Le, M. F. Hundley, J. J. Neumeier, G. M. Luke, K. La0.88MnOx ,La1ϪxSrx(Mn1Ϫx/2Nbx/2)O3 ,Nd1ϪxCaxMnO3 , Kojima, B. Nachumi, Y. J. Uemura, D. E. MacLaughlin, and S.-W. ͑ ͒ and Bi Ϫ Ca MnO manganites have been proposed. It has Cheong, Phys. Rev. Lett. 77, 1869 1996 . 1 x x 3 14 been shown that the magnetic properties of the samples un- C. H. Booth, F. Bridges, G. H. Kwei, J. M. Lawrence, A. L. Cornelius, and J. J. Neumeier, Phys. Rev. Lett. 80, 853 ͑1998͒. der study are determined by the type of their orbital state. 15 M. Fath, S. Freisem, A. A. Menovsky, Y. Tomioka, J. Aarts, and J. A. The dynamic correlations of d3z2Ϫr2 orbitals favor ferromag- Mydosh, Science 285, 1540 ͑1999͒. netic ordering in the manganites, while A-type antiferromag- 16 M. Uehara, S. Mori, C. H. Chen, and S.-W. Cheong, Nature ͑London͒ 399, ͑ ͒ netic structure is typical for the static Jahn-Teller distortions. 560 1999 . 17 B. C. Hauback, H. Fjellvag, and N. Sakai, J. Solid State Chem. 124,43 It has been argued that concentrational transition from an ͑1996͒. antiferromagnetic to a ferromagnetic state occurs via the for- 18 T. Shibata, B. Bunker, J. F. Mitchel, and P. Schiffer, Phys. Rev. Lett. 88, mation of inhomogeneous state due to structural phase sepa- 207205 ͑2002͒. 19 ration mechanism. J. Rodriguez-Carvajal, M. Hennion, F. Moussa, A. H. Moudden, L. Pin- sard, and A. Revcolevschi, Phys. Rev. B 57, R3189 ͑1998͒. This work was partly supported by Belarus and Ukrai- 20 M. C. Sanchez, G. Subias, J. Garcı´a, and J. Blasco, Phys. Rev. Lett. 90, nian Fund for Basic Research ͑Grant F04MS-004 and Grant 045503 ͑2003͒. 10.01/001 F05K–012͒ Polish-Ukrainian Academies research 21 D. Louca, T. Egami, E. L. Brosha, H. Roder, and A. R. Bishop, Phys. Rev. ͑ ͒ ͑ B 56, R8475 ͑1997͒. project No.4-2003 , and by MES of Ukraine Grant M/257- 22 ͒ T. Mizokawa, D. I. Khomskii, and G. A. Sawatzky, Phys. Rev. B 63, 2004 . 024403 ͑2001͒. 23 Y. Endoh, K. Hirota, S. Ishihara, S. Okamoto, Y. Murakami, A. Nishizawa, *E-mail: [email protected] T. Fukuda, H. Kimura, H. Nojiri, K. Kaneko, and S. Maekawa, Phys. Rev. †E-mail: [email protected] Lett. 82, 4328 ͑1999͒. 24 S. Okamoto, S. Ishihara, and S. Maekawa, Phys. Rev. B 61, 14647 ͑2000͒. 25 R. Kajimoto, H. Mochizuki, and H. Yoshizawa, Physica B 329–333,738 ͑2003͒. 1 C. Zener, Phys. Rev. 82,403͑1951͒. 26 R. L. Rasera and G. L. Catchen, Phys. Rev. B 58, 3218 ͑1998͒. 2 A. J. Millis, B. I. Shraiman, and R. Mu¨ller, Phys. Rev. Lett. 77, 175 27 E. O. Wollan and W. C. Koehler, Phys. Rev. 100, 545 ͑1955͒. ͑1996͒. 28 J. B. Goodenough, J.-S. Zhou, F. Rivadulla, and E. Winkler, J. Solid State 3 E. L. Nagaev, Phys. Rep. 346, 387 ͑2001͒. Chem. 175,116͑2003͒. 4 J. B. Goodenough, A. Wold, R. J. Arnott, and N. Menyuk, Phys. Rev. 124, 29 A. K. Zvezdin, V. M. Matveev, A. A. Mukhin, and A. I. Popov, Rare-Earth 373 ͑1961͒. Ions in Magnetically Ordered Crystsals ͓in Russian͔, Nauka, Moscow 5 G.-L. Liu, J.-S. Zhou, and J. B. Goodenough, Phys. Rev. B 64, 144414 ͑1985͒. ͑ ͒ 2001 . 30 I. O. Troyanchuk, N. V. Kasper, O. S. Mantytskaya, and S. P. Pastushonok, 6 G. Biotteau, M. Hennion, F. Moussa, J. Rodrı´guez-Carvajal, L. Pinsard, A. Zh. E´ ksp. Teor. Fiz., 105, 239 ͑1994͓͒JETP 78, 212 ͑1994͔͒. Revcolevschi, Y. M. Mukovskii, and D. Shulyatev, Phys. Rev. B 64, 31 T. Atou, H. Chiba, K. Ohoyama, Y. Yamaguchi, and Y. Syono, J. Solid ͑ ͒ 104421 2001 . State Chem. 145, 639 ͑1999͒. 7 ͑ ͒ G. Allodi, R. De Renzi, and G. Guidi, Phys. Rev. B 57, 1024 1998 . 32 I. O. Troyanchuk, D. D. Khalyavin, S. V. Truhanov, and H. Szymczak, 8 G. Papavassiliou, M. Fardis, M. Belesi, T. G. Maris, G. Kallias, M. Pissas, J. Phys.: Condens. Matter 11, 8707 ͑1999͒. D. Niarchos, C. Dimitropoulos, and J. Dolinsek, Phys. Rev. Lett. 84, 761 33 R. Mathieu, P. Nordblad, D. N. H. Nam, N. X. Phuc, and N. V. Khiem, ͑ ͒ 2000 . Phys. Rev. B 63, 174405 ͑2001͒. 9 G. Papavassiliou, M. Belesi, M. Fardis, and C. Dimitropoulos, Phys. Rev. Lett. 87, 177204 ͑2001͒. This article was published in English in the original Russian journal. Repro- 10 M. M. Savosta, V. I. Kamenev, V. A. Borodin, P. Novak, M. Marysco, duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 31, NUMBERS 8–9 AUGUST–SEPTEMBER 2005

Odd magnetic dichroism of linearly polarized light in the antiferromagnet MnF2 N. F. Kharchenko* and O. V. Miloslavskaya

B. Verkin Institute for Low Temperature Physics and Engineering, National Academy of Sciences of Ukraine, pr. Lenina 47, Kharkov 61103, Ukraine A. A. Milner

Weizmann Institute of Science, Rehovot 76100, Israel ͑Submitted July 11, 2005͒ Fiz. Nizk. Temp. 31, 1081–1088 ͑August–September 2005͒ The two-sublattice tetragonal antiferromagnetic crystal manganese fluoride is found to exhibit ʈ ʈ dichroism induced by a magnetic field H C4 Z for linearly polarized light propagating along the tetragonal axis of the crystal also. The observed effect is odd with respect to the sign of the z projection of the field and the operation of reversing the directions of the sublattice magnetic moments of the antiferromagnet. The effect can be used for optical visualization of 180- ͑ ͒ degree time-reversed antiferromagnetic domains in MnF2 .©2005 American Institute of Physics. ͓DOI: 10.1063/1.2008145͔

INTRODUCTION spectrum of an antiferromagnet was made on the MnF2 crystal.6 Manganese fluoride is a well-known model case of a Manganese fluoride is attracting attention not only in purely spin two-sublattice tetragonal antiferromagnet connection with opportunities for basic research.16,17 In re- ͑AFM͒. It has long served as a test object for checking the- cent years there has been intensive study of the features of 1–3 oretical ideas about the magnetic and elastic properties of the magnetic hysteresis properties of magnetic heterostruc- AFM crystals and their optical properties that result from the tures characterized by exchange-shifted magnetic hysteresis interaction of photons with the magnetic subsystem of the loops and containing ferromagnetic ͑FM͒ nanolayers or nan- AFM and are manifested in the transparency region4,5 and in odots deposited on a crystalline slab or single-crystal film of the absorption6–12 and scattering12–14 of light. an AFM. Research is being done on the mechanisms of ex- Since the resultant orbital moment of the five 3d elec- change interaction between neighboring layers and the trons of divalent manganese in MnF2 is equal to zero and the mechanisms giving rise to uniaxial magnetic anisotropy in ϩ spin of the Mn2 ion is maximum, all of the electric-dipole the FM layer. Manganese fluoride, in particular, is used for transitions within the 3d5 configuration are doubly the AFM base of these structures.17 The presence of 180- forbidden—by parity and by spin. The lowest-energy al- degree ͑time-reversed͒ AFM domains in the AFM layer sub- lowed electric-dipole transitions are of the type 6S(3d5) stantially alters their unidirectional magnetic properties. Ϫ6P(3d44p), and the charge-transfer transitions with the Therefore, in the creation of FM/AFM structures with speci- participation of the fluorine ions lie in the vacuum fied hysteresis properties and also for creating structures with ultraviolet.15 The zero value of the orbital moment in the provisions for switching those properties, it is very desirable ground state of the ion and the remoteness of the allowed to have the possibility of visual monitoring and purposeful electric-dipole optical transitions make for an extremely manipulation using 180-degree AFM domains. small value of the Faraday magnetooptic effect. Even in the The magnetic symmetry of AFM transition-metal fluo- visible range the contribution to the magnetic resonance tran- rides allows effects which are sensitive to opposite orienta- sition effect is predominant over the electric-dipole tions of the AFM vector. Because the anti-inversion opera- 5 contribution. tion¯ 1Ј is not among the symmetry elements of these For intra-configuration d-d transitions the spin selection crystals, the forbiddenness is lifted from unconventional rules are softened by the spin-orbit coupling. The parity for- magnetooptic effects—birefringence linear in the magnetic biddenness is lifted in the presence of local static and dy- field or the linear magnetooptic effect ͑LMOE͒, and namic breaking of inversion, in particular, for the simulta- quadratic-in-magnetic-field rotation ͑QMR͒ of the plane of neous excitation of an exciton and magnon or of an exciton polarization of light. The LMOE and QMR have been 19 and odd phonon. In the exciton-magnon excitations the for- observed in the isostructural crystals CoF2 and FeF2 . With biddenness of the electric-dipole transition mechanism is the help of those effects, the 180-degree, or collinear antifer- lifted with respect to both parity and spin. Because of these romagnetic domains have been visualized by an optical 20 mechanisms the absorption spectra of MnF2 throughout the method; previously those domains had been observed in visible region exhibit weak excitonic magnetic-dipole transi- these same crystals by a topographic method utilizing polar- tions and their stronger electric-dipole satellites: exciton- ized neutron diffraction.21 magnon and exciton-phonon.6–12 We note that the first obser- In manganese fluoride neither the LMOE nor the QMR vation of an exciton-magnon satellite in the absorption have been observed. The present study is devoted to the de-

1063-777X/2005/31(8–9)/6/$26.00825 © 2005 American Institute of Physics 826 Low Temp. Phys. 31 (8–9), August–September 2005 Kharchenko et al.

tection of the LMOE in the MnF2 crystal. In the transparency resented as a superposition of circular and linear dichroisms. region the expected effect is too small to be measured by the MLD, unlike MCD, is odd with respect to reversal of the standard method. However, it was expected that the LMOE directions of the magnetic moments of the sublattices. might manifest itself in the regions of the optical transitions, LMOE is described by an axial c-tensor qija which is in effects which are odd in the field: magnetic linear birefrin- symmetric with respect to permutation of the first two indi- gence and magnetic linear dichroism ͑MLD͒. The detection ces and has components that change sign under the time of LMOE could make it possible in principle to make optical inversion operation. The components of the tensor are iden- tically zero in all antiferromagnets whose magnetic space observations of the collinear AFM domains in MnF2 , where ␶ ¯ they have been visualized previously only by methods of group contains the antitranslation •IЈ or anti-inversion IЈ. neutron topography. A test for the effects in question could The tensor qija has the same symmetry as the tensor describ- be the presence of properties characteristic of the LMOE— ing the piezomagnetic effect. The matrices of the tensor for oddness with respect to reversal of the field direction and all magnetic crystals described by Shubnikov groups are oddness with respect to switching of the antiferromagnetic given in Ref. 18. For a crystal with symmetry 4Ј/mmmЈ the ϩ Ϫϭ ϩ ϭ states (AFM and AFM IЈ•AFM ), connected to each tensor qija has only three nonzero components, qyzx qxzy ʈ other by the time inversion operation IЈ. The effect can be and qxyz . For H C4 the LMOE is described by increments to ⌬␧ ϭ⌬␧ manifested most clearly in the longitudinal experimental ge- the components of the dielectric tensor, xy yx ʈ ʈ ͒ ϭ ometry, H k C4 (k is the wave vector , for which in the qxyzH. The field-induced dichroism for linearly polarized absence of magnetic field there is neither linear birefringence light is determined by the imaginary part of the tensor qxyz . nor linear dichroism. In this paper we report the observation of magnetic field- The inevitability that LMOE will be manifested in the induced dichroism of linearly polarized light in the region of MnF2 crystal in the longitudinal experimental geometry can three groups of absorption bands of the MnF2 crystal, de- be seen from the following considerations. The magnetic noted in the literature as the A, C, and D groups. These point group of the AFM crystal, 4Ј/mmmЈ, differs from the groups are formed by the intra-configuration electronic tran- sitions from the ground state 6A (6S ) of the Mn2ϩ ion to point group of the paramagnetic crystal, 4/mmmЈ•IЈ, by the 1g 5/2 circumstance that its tetragonal axis is an axis of symmetry excited states split by the crystalline field and the spin-orbit only in combination with the time inversion operation. Since coupling: 6 6 Ϫ4 4 ͑ the symmetry group of the magnetic field does not contain A1g( S5/2) T1g( G)—the A group in the Ϫ1 the operation 4Ј, in a magnetic field HʈC the tetragonal 18480 cm region͒ 4 6 6 Ϫ4 4 4 ͑ crystal must, according to the Curie-Neumann symmetry A1g( S5/2) A1g , Eg( G)—C group in the Ϫ1 principle, lower its symmetry to rhombic mmЈmЈ and be- 25300 cm region͒ 6 6 Ϫ4 4 ͑ come optically biaxial. The appearance of LMOE is due to A1g( S5/2) T2g( D)—the D group in the Ϫ1 ͒ the lifting of the energy quasi-degeneracy of the magnetic 28050 cm region . sublattices by the magnetic field. The unit cell of MnF2 con- tains two magnetic ions, Mn2ϩ(1) and Mn2ϩ(2). The fluo- EXPERIMENT AND DISCUSSION rine ions surrounding the manganese ions form distorted oc- The manganese fluoride single crystals from which the tahedra with symmetry mmm(D2h). The octahedra for the samples were prepared were grown at the P. L. Kapitza In- 2ϩ ͓ ͔ Mn (1) sites are slightly elongated along the 110 axis, stitute of Physics Problems. The samples were in the form of ϩ and those for the Mn2 (2) sites are elongated along ͓11¯ 0͔. parallelepipeds with a base of around 2ϫ2 mm and thick- ϩ ϩ It can be assumed that the Mn2 (1) and Mn2 (2) sites form ness 1.63 mm ͑for the spectral measurements in the A group two interpenetrating optically biaxial subsystems possessing of bands͒ and 0.19 mm ͑for measurements in the region of birefringence in the direction of the crystallographic axis C4 . the C and D groups͒. With a double monochromator a spec- Since the subsystems are rotated with respect to each other tral resolution in the MLD spectra of 2 cmϪ1 or better was ␲ around C4 by an angle of /2, their birefringent properties in obtained. The measurements were made by a modulation the direction of that axis are completely compensated. At method with the use of a Jobin Yvon piezooptic modulator. temperatures below the Ne´el temperature TN the spin mo- The modulation frequency of the polarization of the light ments of the magnetic ions lie along the tetragonal axis C4 beam was 18 kHz. There was a provision for rotating the ʈ antiparallel to each other. In a magnetic field H C4 the sub- modulator around the optical axis of the apparatus so that the systems with spins directed along the field and the sub- azimuths of the polarization axes of the light beam at the systems with spins directed counter to the field are no longer extremal phases of the modulation could be aligned parallel equivalent, and the compensation of the birefringent proper- to specified orthogonal directions in the crystal. The samples ties is broken. The dichroism and birefringence arising in were mounted in a capsule that was placed in a cold finger magnetic field should be odd with respect to reversal of the located in the evacuated cavity of a superconducting sole- magnetic field direction and at low fields should be propor- noid. The temperature of the sample was varied from 6 to tional to the field strength, and their signs should be different 100 K by means of a heater mounted on the cold finger, and for the antiferromagnetic states AFMϩ and AFMϪ of the the measurements were made with a semiconductor ther- crystal. Together with the expected MLD, in the longitudinal mometer glued to the capsule containing the sample. geometry of the experiment there is also the usual magnetic The angle between the C4 axis of the crystal and the ͑ ͒ circular dichroism MCD , which was observed in Refs. 9 direction of the light beam and also the angle between the C4 and 10. The presence of the two types of dichroism is axis and the axis of the solenoid were monitored visually by equivalent to elliptical dichroism, which can always be rep- observation of the conoscopic figures. The axis of the light Low Temp. Phys. 31 (8–9), August–September 2005 Kharchenko et al. 827

beam made an angle of less than 1° with the tetragonal axis of the sample. The cone angle of the irised light beam was ͑ ͒ not more than 1°. The optic axis of the crystal the C4 axis was deviated from the magnetic-field direction by an angle close to 2°. The plane formed by the C4 axis and the H vector was close to the ͑110͒ plane. The presence of transverse magnetic-field components Hx and Hy gave the hope that, owing to the longitudinal ϭ magnetization M z CzxyHxHy induced by them, which is sensitive to the orientation of the antiferromagnetic vector, a uniform antiferromagnetic state will form in the sample22 when it is cooled in a magnetic field Hx ,Hy ,Hz to tempera- tures below TN . The discovery of the LMOE has made it possible to monitor a measure of the uniformity of the anti- ferromagnetic sample. As subsequent experiments showed, FIG. 1. Spectra of magnetic linear dichroism of MnF2 in the region of the 6 6 Ϫ4 4 4 ͑ ͒ ϩ optical transition A1g( S5/2) Eg , A1g( G) the C group of bands ,ob- antiferromagnetic states close to the single-domain AFM ϭϩ Ϫ tained at a field strength Hz 61 kOe for different states of the sample. and AFM states actually do form in the sample. The sample Curves 1 and 2 were obtained for different parts of a sample that had not was heated to a temperature of around 100 K, i.e., above the been subjected to thermomagnetic treatment and contained collinear AFM ϭ domains. Curve 3 was obtained after a single-domain state AFMϩ had been Ne´el temperature TN 67 K, and then a magnetic field was Ϸ Ϸ ϭϩ created in the sample. applied (Hx 1.5 kOe, Hy 1.6 kOe, Hz 50 kOe or Hz ϭϪ50 kOe), and the sample was cooled in that field to a temperature of 6 K. 9 The spectral measurements of the dichroism ⌬D(␭) the dichroism differed less. In our experiments the qualita- ϭ Ϫ ϩ tive relationship between the values of the MCD in the ex- (I1 I2)/(I1 I2) were carried out at a temperature of 6 K ͑ citon and exciton-magnon bands of these groups was the here I1 and I2 are the intensities on emergence from the sample, of the light beams which on incidence had been same, although the quantitative difference for the bands in polarized along the crystallographic directions ͓110͔ and the C group was smaller. The spectra of the MLD measured ͓11¯ 0͔, respectively͒. In measurements of the circular di- in this study reflected the structure of the groups of bands, but the relationship between the contributions of the exciton chroism, I1 and I2 correspond to right and left circular po- larization of the incident beams. The spectral resolution in and exciton-magnon transitions was different. the measurements was close to 2 cmϪ1, and the error of Figure 1 shows the results of a measurement of the MLD in the C group in different states of the sample. Curves 1 and determination of the absolute values of the photon energies ϭϩ was around 5 cmϪ1. The values given here for the energies 2 were obtained in a field Hz 61 kOe with light passing of the optical transitions and the oscillator strengths are the through different parts of the sample, which had not been standard values taken from Ref. 8. subjected to thermomagnetic treatment and was separated into AFMϩ and AFMϪ domains. It is seen that in one part The values of the MCD and MLD are proportional to the ͑ ͒ splitting of the absorption bands in the magnetic field. The curve 1 the MLD is close to zero; light passes through splitting of the exciton absorption band differs little from that domains of both AFM states, and the path of the light in the AFMϪ domain is slightly longer than the path in the AFMϩ of the absorption band at the intra-ion electronic transition, ͑ ͒ ϩ which is equal to domain. In the other part curve 2 the AFM domain is predominant. Curve 3 was obtained at the same field but with ⌬ ϭ ␮ Ϫ ͒ϭ ␮ Ϫ ͒ϩ ϩ Eex 2 ␤H͑g0S0 gexSex 2 BH͓S0͑g0 gex gex͔. the sample prepared in a uniform AFM state beforehand. ϭ Ϫ Figure 2 gives the spectra of the MLD of this same Here g0 ,S0 and gex ,Sex S0 1 are the g factors and spins ␮ group, obtained in a single-domain sample in magnetic fields of the ground and excited states, and B is the Bohr magne- ton. The splitting of the exciton-magnon band, of different strength. One notices that in the absence of field the linear dichroism actually vanishes in the region of some ⌬ ϭ ␮ Ϫ ͒Ϫ ␮ EexϪm 2 BH͑g0S0 gexSex 2 Bg0H absorption bands but remains well noticeable in the region of other bands. The cause of this behavior of the linear dichro- ϭ2␮ H͑g Ϫg ͒S B 0 ex ex ism can be linked to different sensitivity of the optical tran- is much smaller, since the g factors of the ground and excited sitions to the deviation from the ideal sample orientation 2ϩ ʈ ʈ states of the Mn ion are close in value. Splitting of the C4 k H and to the presence of macroscopic strains in the exciton-magnon bands was not observed in the spectroscopic sample. The variation of the value of the dichroism at the experiments.7,8 It was determined from the value of the MCD extrema of the spectrum with increasing field is close to lin- in the band.9,10 The value of the dichroism is proportional not ear. only to the splitting but also to the intensity of the band, and Figure 3 shows the spectra of the MLD in the region of the magnetic dichroisms of the exciton and exciton-magnon all three absorption groups. Prior to the measurements a suit- bands can therefore be close. In the experiments the MCD of able thermomagnetic treatment of the samples was carried the exciton band in the C group was approximately 3 times out to prepare a uniform AFM state in them. The spectra of 10 larger than the MCD of the exciton-magnon satellites. In the MLD and the absorption spectrum of the MnF2 crystal in the A group the MCD of the exciton-magnon band was the region of the A group of bands, formed by the so-called ␴ ␴ larger than the MCD of the exciton band, while the values of 1 and 2 magnon satellites of the electronic magnetic di- 828 Low Temp. Phys. 31 (8–9), August–September 2005 Kharchenko et al.

FIG. 2. Spectra of magnetic linear dichroism of MnF2 in the region of the C group of bands, measured for different values of the magnetic field strength. The sample was found in the AFMϩ state.

6 6 Ϫ4 4 pole transitions A1g( S5/2) T1g( G) are presented in Fig. 3a. The electronic transitions ͑not seen in the ␣ spectrum͒ occur upon the absorption of photons with energies of 18418 cmϪ1 and 18435 cmϪ1 and have close values of the oscillator strength (3.4ϫ10Ϫ11 and 3.0ϫ10Ϫ11). Electrons 6 3 2 from the almost unsplit ground state A1g(t2geg) of the man- ganese ion make a transition to the orbital components B2 4 4 and B3 of the excited state T1g( G), which is split by the spin-orbit coupling and the orthorhombic crystalline field. The electronic configuration of the excited states is close to 4 1 t2geg . The arrows indicate the positions of the maxima of the bands of their electric-dipole satellites ͑18476 and 18484 cmϪ1). These absorption bands ͑the oscillator strengths are 1ϫ10Ϫ9 and 1.4ϫ10Ϫ9) are due to the simul- taneous excitation of an exciton and a magnon, which propa- gate through different sublattices in opposite directions along the crystallographic axes ͓001͔ and ͓111͔. Here the absorp- tion of light is excited primarily by excitons and magnons with the maximal momenta, which correspond to the points Z and A in the Brillouin zone. The induced linear dichroism in this group reaches its ϫ Ϫ2 ϭ largest value (2.5 10 at a field strength Hz 50 kOe) in the vicinity of the first exciton-magnon transition. The sign FIG. 3. Spectra of magnetic linear dichroism induced by a longitudinal of the field-induced dichroism changes when the direction of ʈ magnetic field H C4 :k in the MnF2 crystal in the region of the optical 6 6 Ϫ4 4 ͑ ͒͑͒ 6 6 the magnetic field is reversed while the direction of the an- transitions A1g( S5/2) T1g( G) the A group of bands a ; A1g( S5/2) Ϫ4 4 4 ͑ ͒͑ ͒ 6 6 Ϫ4 4 ͑ tiferromagnetic vector remains the same. This property is Eg , A1g( G) the C group of bands b ; A1g( S5/2) T2g( D) the D ͒͑͒ demonstrated in Fig. 3a. A sign change of the MLD also group of bands c . The absorption spectrum is shown in the lower part of each panel. Prior to the measurements a uniform AFM state was created in occurs when the antiferromagnetic state is changed from the samples. The curves with data points in the form crosses were obtained ϩ Ϫ AFM to AFM . The switching of the AFM states occurred, at Hϭ0. as expected, after the sample was heated to a temperature above the Ne´el point and then cooled in magnetic field. The prepared AFM state of the sample persisted as metastable in magnetically reversed ͑time-reversed͒ AFM state formed, the oppositely directed field up to a strength of 61 kOe ͑the and the value of the magnetic dichroism decreased. maximum field strength in the experiment͒ if the sample Figure 3b shows the MLD spectra and the absorption temperature was low enough (Tϭ6 K). This is attested by spectrum of the C group of bands, formed by the transitions ϩ 6 6 Ϫ4 4 4 the fact that the MLD spectra recorded for the AFM and A1g( S5/2) Eg , A1g( G). The electronic configuration Ϫ 3 2 AFM states at the same magnetic field are almost com- of the excited electronic states is close to t2geg . The energies ͑ ͒ ͑ ͒ pletely identical up to the sign ; see Fig. 3b. However, if the of two electronic excitonic magnetic-dipole transitions, M 1 ͑ ͒ ͑ Ϫ1 sample temperature was high enough 25–30 K , then in the and M 2 25245 and 25260 cm ), with oscillator strengths geometry we used, a partial switching of the AFM state oc- 0.5ϫ10Ϫ8 and 0.9ϫ10Ϫ8, and their satellites ͑25295 and curred as the field was increased—domains of the other, 25311 cmϪ1), with oscillator strengths 0.3ϫ10Ϫ7 and 0.9 Low Temp. Phys. 31 (8–9), August–September 2005 Kharchenko et al. 829

ϫ Ϫ7 10 , are indicated, along with the transition M 3 at 25404 cmϪ1, which apparently is also purely electronic ͑see Refs. 8 and 10͒. According to Ref. 8, the excited electronic 4 states M 1 and M 2 are orbital components of the Eg state, split by the z component of the orthorhombic crystalline 4 ͑ ͒ field, while the M 3 state comes out of Ag Ref. 10 . The field-induced dichroism is clearly seen not only in the region of strong electric-dipole exciton-magnon satellites but also in the region of weak magnetic-dipole transitions having a large g factor. We note that these transitions are much less visible in the ␣ absorption spectrum than in the dichroism spectra. The largest value of the MLD is reached at an energy of 25310 cmϪ1 in the region of the second exciton-magnon band, where it is close to 1.7ϫ10Ϫ2 in a magnetic field H ϭ50 kOe. The spectral dependence of the MLD in the region of the D group of bands is shown in Fig. 3c. This part of the spec- trum of the MnF2 crystal is formed by optical transitions 6 6 4 4 between the states A1g( S5/2) and T2g( D). The fraction of 3 2 the t2geg configuration in the configurations of the excited states of the ion is appreciable, close to 0.83. The absorption band with maximum near 28026 cmϪ1 is attributed to an exciton-magnon transition.8 The oscillator strength of this transition in the ␴ spectrum is 0.17ϫ10Ϫ7. Odd MLD was recorded in the region of this band. The figure shows the MLD spectra obtained for the AFM state of the sample in magnetic fields of opposite directions. The curves demon- strate the change of sign of the MLD for the same AFM state when the direction of the magnetic field is reversed. The maximum value of the dichroism, scaled to 50 kOe, in this band is close to 3.5ϫ10Ϫ3. We note that the amplitude of the anomalies on the curves are noticeably different. The differ- ence is most likely due not to a partial switching of the AFMϩ state and the formation of AFMϪ domains but rather to the circumstance that in the higher energy region of the spectrum, the more-stringent tolerances on the deviations of the light beam from the tetragonal axis and on the allowable macroscopic stress in the sample are exceeded. A similar but ͑ ͒ less pronounced asymmetry is also observed in the spectra of FIG. 4. Spectrum of magnetic circular dichroism MCD induced by a lon- gitudinal magnetic field in the region of the optical transitions. The curves the C group. 6 6 with filled dots are the MCD, the dotted curves are the MLD: A1g( S5/2) Ϫ4 4 ͒ ϭ ͑ ͒ 6 6 Ϫ4 4 4 The stated properties of the MLD, viz., oddness with T1g( G)(A group , H 50 kOe a ; A1g( S5/2) Eg , A1g( G)(C ͒ ϭ ͑ ͒ 6 6 Ϫ4 4 ͒ ϭ ͑ ͒ respect to a reversal of the field direction and with respect to group , H 50 kOe b ; A1g( S5/2) T2g( D)(D group , H 33 kOe c . switching of the AFM states, attest to the fact that the ob- The absorption spectrum is shown in the lower part of each panel. served dichroism is indeed a manifestation of a linear mag- netooptic effect. 18456– 18457 cmϪ1 a feature which was also observed in The values of the MLD in all groups of absorption bands Ref. 9 is seen in the MCD spectrum ͑but not in the MLD͒. is of the same order as the usual magnetic circular dichroism. In the exciton-magnon bands the MLD is noticeably Figure 4 shows the MCD spectra which we obtained in the larger than the MCD, and therefore the values of the g fac- region of the A, C, and D groups of bands. The form of the tors for the exciton-magnon satellites and for magnons, MCD spectra in the region of the A and C groups is close to which were determined on the basis of the values of the the form of the MCD spectra measured previously in Ref. 9 MCD of the corresponding absorption bands,9,10 must be cor- (A group͒ and 10 (C group͒, but the differences in the details rected with the MLD taken into account. In our opinion, the are significant. The dispersion curves of the MLD and MCD question of spontaneous magnetic circular ͑and linear͒ di- are close for the A and D groups but noticeably different for chroism, which is discussed in Refs. 9 and 11, requires ad- the C group. In the A group both the MLD and MCD spec- ditional careful experimental study. tra, in addition to the features corresponding to exciton- magnon transitions, have a feature at 18465– 18470 cmϪ1, CONCLUSION which is the position of the magnetic-dipole excitonic tran- In the antiferromagnetic manganese fluoride crystal a sition that was predicted and observed in Ref. 6. Near longitudinal magnetic linear dichroism is observed in the ex- 830 Low Temp. Phys. 31 (8–9), August–September 2005 Kharchenko et al. citon and exciton-magnon absorption bands: a magnetic field 7 V. V. Eremenko and Yu. A. Popkov, Phys. Status Solidi 12,627͑1965͒; induces a different absorption of linearly polarized light Philip G. Russell, Donald S. McClure, and J. W. Stout, Phys. Rev. Lett. propagating along the magnetic field vector parallel to the 16, 176 ͑1966͒; D. D. Sell, R. L. Greene, and Robert M. White, Phys. Rev. ͑ ͒ ͑ ͒ tetragonal axis of the crystal. It is found that the magnetic 158, 489 1967 ; D. D. Sell, J. Appl. Phys. 39, 1030 1968 ; T. Tsuboi and P. Ahmet, Phys. Rev. B 45, 468 ͑1992͒. linear dichroism is odd with respect to reversal of the mag- 8 R. S. Meltzer and L. L. Lohr, J. Chem. Phys. 49, 541 ͑1968͒;R.S. netic field direction and to switching of the antiferromagnetic Meltzer, M. Lowe, and D. S. McClure, Phys. Rev. 180,561͑1969͒. state, i.e., to a reversal of the directions of the magnetic 9 Y. H. Wong, F. L. Scarpace, C. D. Pfeifer, and W. M. Yen, Phys. Rev. B 9, moments of the sublattices of the antiferromagnetic crystal. 3086 ͑1974͒; P. W. Seltzer, W. C. Egbert, D. L. Huber, and W. M. Yen, The largest MLD, close to 0.5ϫ10Ϫ3 kOeϪ1, was observed ibid. 12,961͑1975͒. 10 in the region of the first exciton-magnon satellite in the A R. W. Schwartz, J. A. Spencer, W. C. Yeakel, P. N. Schatz, and W. G. ͑ ͒ group of bands of the absorption spectrum. The value of the Maisch, J. Chem. Phys. 60, 2598 1974 . 11 D. J. Lockwood, R. M. White, and W. M. Yen, Phys. Rev. B 69, 174413 effect is sufficient for visual observation of AFM domains in ͑2004͒. the MnF2 crystal with the use of modern techniques for op- 12 V. V. Eremenko, Introduction to the Optical Spectroscopy of Magnets ͓in tical visualization of low-contrast objects and the methods of Russian͔, Naukova Dumka, Kiev ͑1975͒. high-resolution spectroscopy. 13 P. A. Fleury, S. P. S. Porto, and R. Loudon, Phys. Rev. Lett. 18,658 It is clear that the MLD and MCD spectra complement ͑1967͒; D. J. Lockwood and M. Cottam, Phys. Rev. B 35, 1973 ͑1987͒. 14 each other and can be useful for determining the wave func- M. G. Cottam and D. J. Lockwood, Light Scattering in Magnetic Solids, ͑ ͒ tions of the excited states and for revealing subtle differences Wiley, New York, 1986 . 15 W. Parkinson and E. W. Willams, J. Chem. Phys. 18,534͑1950͒. in the behavior of excitons and magnons in magnetic field. 16 H. Mariette, F. Kany, and J. M. Hartman, Appl. Surf. Sci. 123Õ124,710 ͑1998͒; N. S. Sokolov, Y. Takeda, A. G. Banschikov, J. Harada, K. Inaba, * E-mail: [email protected] H. Ofuchi, M. Tabuchi, and N. L. Yakovlev, ibid. 162Õ163,469͑2000͒. 17 J. Nogues and I. K. Schuller, J. Magn. Magn. Mater. 192,203͑1999͒, Zhi-Pan Li, O. Petracic, J. Eisenmenger, and I. K. Schuller, ArXiv cond- ͑ ͒ 1 R. A. Erickson, Phys. Rev. 90,779͑1953͒; A. S. Borovik-Romanov and matter/0501158 2005 . 18 H. Grimmer, ‘‘Magnetic properties,’’ in Physical Properties of Crystals, V. V. Eremenko, N. F. Kharchenko, Yu. G. Litvinenko, and V. M. Nau- Vol. D of Internatonal Tables for , A. Authier ͑ed.͒, Pub- menko, Magneto-Optics and Spectroscopy of Antiferromagnets, Springer- lished for the International Union of Crystallography by Kluwer Academic Verlag, New York ͑1992͒; N. F. Kharchenko, Ferroelectrics 162,173 Publishers, Dordrecht-Boston-London ͑2003͒, p. 105. ͑1994͒. 2 Jimia Zhao, Andrea V. Bragas, David J. Lockwood, and Roberto Merlin, 19 N. F. Kharchenko, V. V. Eremenko, and L. I. Belyi, JETP Lett. 28,325 ͑ ͒ ArXiv cond-mat/0311329 2003 . ͑1979͒; N. F. Kharchenko, A. V. Bibik, and V. V. Eremenko, JETP Lett. 42, 3 ͑ ͒ Y. Shapira, S. Foner, and A. Misetich, Phys. Rev. Lett. 23,98 1969 ; 553 ͑1985͒. R. L. Melcher, Phys. Rev. 2,733͑1970͒; K. L. Dudko, V. V. Eremenko, 20 ͑ ͓͒ N. F. Kharchenko, V. V. Eremenko, and L. I. Belyi, JETP Lett. 29,392 and L. M. Semenenko, Fiz. Nizk. Temp. 1,68 1975 Sov. J. Low Temp. ͑ ͒ Phys. 1,33͑1975͔͒. 1979 ; V. V. Eremenko, N. F. Kharchenko, and L. I. Belyi, J. Appl. Phys. ͑ ͒ 4 A. S. Borovik-Romanov, N. M. Kre nes, A. A. Pankov, and M. A. Tala- 50, 7751 1979 . 21 laev, Zh. E´ ksp. Teor. Fiz. 64, 782 ͑1974͓͒sic͔. I. R. Jahn and H. Dachs, M. Schlenker and J. Baruchel, J. Appl. Phys. 49,1996͑1978͒; M. Schlen- Phys. Status Solidi B 57,681͑1973͒. ker, J. Baruchel, and B. Barbara, J. Magn. Magn. Mater. 15–18 ͑1980͒. 5 N. F. Kharchenko and V. V. Eremenko, Fiz. Tverd. Tela ͑Leningrad͒, 9, 22 A. V. Bibik, N. F. Kharchenko, and S. V. Petrov, Fiz. Nizk. Temp. 15, 1280 1655 ͑1967͓͒Sov. Phys. Solid State 9, 1302 ͑1967͔͒. ͑1989͓͒Sov. J. Low Temp. Phys. 15,707͑1989͔͒. 6 R. L. Greene, D. D. Sell, W. M. Yen, A. L. Schawlow, and R. M. White, Phys. Rev. Lett. 15,656͑1965͒. Translated by Steve Torstveit LOW TEMPERATURE PHYSICS VOLUME 31, NUMBERS 8–9 AUGUST–SEPTEMBER 2005

PERSONALIA

Viktor Grigor’evich Bar’yakhtar „on his 75th birthday… ͓DOI: 10.1063/1.2044810͔

On August 9, 2005 the prominent theoretical physicist magnetic field, a phenomenological description of relaxation Viktor Grigor’evich Bar’yakhtar, Member of the National processes in magnets, and a theory of the superconducting Academy of Sciences of Ukraine, turned 75. In the course of transition under pressures altering the topology of the Fermi over forty years he has published a paper that has been semi- surface. His scholarly activity is reflected in numerous ar- nal in various fields of theoretical physics, carried on an ticles, conference proceedings, and monographs. He has cre- enormous amount of scientific organizational activity within ated an active and successful school of theoretical physicists. the NAS Ukraine system, served on government and inter- Bar’yakhtar has been awarded the Order of Merit of national committees and scientific councils and on the edito- Ukraine in Science and several State and Academy of Sci- rial boards of journals, and taught at the Universities of ences prizes, orders, and medals. Kharkov, Donetsk, and Kiev. He is the Director of the Insti- Viktor Grigor’evich possesses great magnetism owing to tute of Magnetism of the NAS Ukraine, which was founded the wealth of his scientific ideas and his sterling personal by him. qualities. We heartily congratulate him on his birthday and Bar’yakhtar obtained ground-breaking results in the wish him many more years of good health, creative activity, theory of magnetism and phase transitions. In particular, he and charisma. constructed a theory of magnetoacoustic resonance, a theory Editorial Board of Low Temperature Physics of the intermediate state near a first-order phase transition in

1063-777X/2005/31(8–9)/1/$26.00831 © 2005 American Institute of Physics LOW TEMPERATURE PHYSICS VOLUME 31, NUMBERS 8–9 AUGUST–SEPTEMBER 2005

Radii Nikolaevich Gurzhi „on his 75th birthday… ͓DOI: 10.1063/1.2044811͔

August 11, 2005 is the 75th birthday of the prominent ticles dominate or compete with processes leading to mo- theoretical physicist and Laureate of the State Prize of mentum relaxation. The numerous results obtained on that Ukraine, Radii Nikolaevich Gurzhi. The quantum kinetic topic have been presented in monographs and text books and equation for electrons, obtained by Gurzhi when a graduate have found wide experimental confirmation. Since his arrival student, became the basis for his well-known results in metal in 1974 at the Institute for Low Temperature Physics and optics which extended the concepts of Landau’s theory of the Engineering of the National Academy of Sciences of Fermi liquid into the region of relatively high frequencies. Ukraine, his ideas have been applied to new solid-state Gurzhi’s most important scientific ideas were announced objects—nanostructures of reduced dimensionality. He has during his work at the Ukrainian Physico-Technical Institute shown that lowering the dimensionality of the space radi- ͑from 1958͒. In his first paper, published in 1963, a system of cally alters the character of electron-electron scattering in quasiparticles in a solid was treated as a Poisueille fluid conductors. whose behavior is described by the equations of hydrody- His colleagues and students who are celebrating this namics. The term ‘‘Gurzhi effect’’ has long been used in the birthday value his special combination of scholarly and hu- literature for manifestations of hydrodynamic phenomena in man qualities: his ability to frame unexpectedly simple ques- electrical conduction. Another unexpected effect predicted tions in complex problems, and his informal, intelligent, and by Gurzhi and observed experimentally is the role of anhar- human style of personal contact irrespective of station. We monicities of arbitrarily high order in the thermal conductiv- esteem and love our dear Radii Nikolaevich and wish him ity of ferromagnetic insulators. Gurzhi and his students have long years of fruitful scientific activity. studied a wide range of problems in solid-state kinetics under conditions such that normal collision processes of quasipar- Editorial Board of Low Temperature Physics

1063-777X/2005/31(8–9)/1/$26.00832 © 2005 American Institute of Physics LOW TEMPERATURE PHYSICS VOLUME 31, NUMBERS 8–9 AUGUST–SEPTEMBER 2005

ERRATA

Erratum: Evgenii Stanislavovich Borovik on the ninetieth anniversary of his birth †Low Temperature Physics 31, 183 „2005…‡ V. V. Eremenko and V. S. Borovikov Due to a translation error, on page 183, in the third paragraph of the text, the title of the Candidate’s Dissertation has been corrected. The correct title should read The Heat Conductivity of ...

1063-777X/2005/31(8–9)/1/$26.00833 © 2005 American Institute of Physics