This article was downloaded by: [University of Cambridge] On: 06 March 2012, At: 02:05 Publisher: Taylor & Francis Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Contemporary Physics Publication details, including instructions for authors and subscription information: http://www.tandfonline.com/loi/tcph20

Varieties of magnetic order in solids C. M. Hurd a

a

Solid State Chemistry, National Research Council of Canada, Ottawa, K1A OR9, Canada

Available online: 13 Sep 2006

To cite this article: C. M. Hurd (1982): Varieties of magnetic order in solids, Contemporary Physics, 23:5, 469-493 To link to this article: http://dx.doi.org/10.1080/00107518208237096

PLEASE SCROLL DOWN FOR ARTICLE Full terms and conditions of use: http://www.tandfonline.com/page/terms-and-conditions This article may be used for research, teaching, and private study purposes. Any substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any form to anyone is expressly forbidden. The publisher does not give any warranty express or implied or make any representation that the contents will be complete or accurate or up to date. The accuracy of any instructions, formulae, and drug doses should be independently verified with primary sources. The publisher shall not be liable for any loss, actions, claims, proceedings, demand, or costs or damages whatsoever or howsoever caused arising directly or indirectly in connection with or arising out of the use of this material.

COF’TEMP. PHYS.,

1982, voi.. 23, NO. 5, 469493

Varieties of Magnetic Order in Solids

Downloaded by [University of Cambridge] at 02:05 06 March 2012

C. M.HURD Solid State Chemistry, National Research Council of Canada, Ottawa. K I A OR9. Canada ABSTRACT.Mictomagnetism, metarnagnetism, asperomagnetism, sperimagnetism, speromagnetisrn, spin glass or cluster glass-this string of jargon is part of a score or more ofnew terms recently added to the lexicon ofmapetism. Few ofthem appear in standard textbooks. This article gives an elementary and descriptive review of the new kinds ofmagnetism and their relationships.We have tried to cover all the new terminology. Starting from the familiar types ofmagneticorder in solids, we review the consequences of amorphousnessand disorder for atomic moments, exchange interactions,and single-ion anisotropies.In all, fourteen different types of magnetism are considered.

1. Introduction Magnetism in solids used to be a tidy subject. Although the microscopic origins of some of the different types of magnetism were disputed, their classification seemed straightforward. Five basic types of magnetic behaviour were distinguished and associated with diamagnetism, paramagnetism, ferromagnetism, antiferromagnetism and ferrimagnetism.It had been established that, apart from closed shell diamagnetism and the diamagnetism and paramagnetism of conduction electrons, the behaviour comes from permanent, microscopic magnetic moments possessed by some or all ofthe ions in the solid; the difference between the behaviours lies in the internal arrangement of these moments. Magnetism in solids was an orderly subject that considered orderly systemswith identical magnetic ions distributed throughout a regular crystallinelattice on equivalent atomic sites. The subject was upset about a decade ago by a burst oftheoretical and experimental activity involving two related types of systems: amorphous solids, in which no two atomic sites are equivalent, and disordered solids, in which different atoms occupy irregularly the sites of a regular crystal lattice. New types of magnetic order were recognized that appear only where there is no long-range order, while others appear only in a regular crystal lattice. The subject has since expanded from the original five types of magnetic behaviour to comprise nearly three times that number, and the terminology has multiplied proportionately. It has been jokingly suggested (Nature 1973) that magnetism needs a taxonomist to cope with names like mictomagnetism, metamagnetism, asperomagnetism, speromagnetism, sperimagnetism,spin glass, cluster glass and a dozen others in current use. Clearly, the field presents a bewildering picture to an outsider. My aim is to give an elementary and descriptive review of the origms and connections of the various types of magnetic order in solids. I start from the five basic types (Section 3) and consider their relationship to nine other classes of behaviour (fig. 1).(I shall concentrate on the ground states of the systems and ignore the various types of magnetic excitations that are possible.) In each case I will give a short description of the microscopic origins of the order, together with some examples of its 0010 7514/82,2305 0469 5005.00 I

1982 Taylor & Francis Ltd

Downloaded by [University of Cambridge] at 02:05 06 March 2012

470

C. M. Hurd

Vcrrieties of mtrgnetic order in solids

47 1

Downloaded by [University of Cambridge] at 02:05 06 March 2012

occurrence.I have tried to cover all the new terminology. It is convenient to begin with a brief review of the two essential ingredients for magnetic order-magnetic ions and a coupling between them-before turning to the main description in Sections 3 and 4. 2. Requirements for magnetic order The interpretation of magnetic effects in solids has developed from two concepts. The first is that a discrete magnetic moment can be associated with ions in a solid. (Induced moments are produced by an externally applied magnetic field; spontaneous ones are present even in its absence.) The second concept is that these microscopic moments interact mutually not only through the ordinary dipole-dipole forces analogous to that felt when two bar magnets are pushed together-for these forces are far too weak to be important-but through quantum mechanical forces. These socalled exchange forces depend on the separation of the magnetic ions as well as their geometrical arrangement, leading to the variety of magnetic order in solids. 2.1. Magnetic ions

The electron is the carrier of magnetism. As well as charge, it has an intrinsic angular momentum (‘spin’) that leads to an intrinsic magnetic moment (the Bohr J T-l). The origin of a permanent magnetic moment on magneton, pB =9.27 x an isolated atom was dealt with in an earlier Conrernportrry Physics article (Allen 1976). Briefly, an atom has a net moment when an inner d- or f-electron shell is incompletely filled so that the spin and orbital momenta of the electrons in the shell do not cancel exactly. (The electrons in such shells are frequently called ‘magnetic electrons’.) The Periodic Table shows five groups in which this occurs: the iron group (incomplete 3d shell), the palladium group (4d shell), the lanthanide group (4f shell), the platinum group (5d shell) and the actinide group (5f shell). The idea that a moment can persist when the atom becomes an ion in a solid, even a metal, is basic to the interpretation of magnetic order in solids. There is a strong interaction between electrons in the outer or valence shells ofneighbouring ions, particularly in a metal, leading to an energy band of delocalized states, but the d- or f-shells of ions in such circumstances are af€ected to different extents. The f-shell, which is the more localized and tightly bound in the ion, is less affected by neighbouring ions and it generally retains its atomic characteristics. The origin of its moment can thus be envisaged as in the isolated atom, with an integral number of Bohr magnetons per ion expected.This picture is particularly appropriate to a magnetic ion in an insulator, and much is known about its behaviour (Bates and Wood 1975),but in metals the delocalized electrons complicate the situation. Not only are they an extra source of discrete and itinerant magnetic moments-leading to Pauli paramagnetism-but their interaction with the magnetic ions undermines the atomic concept of a local moment of fixed magnitude. The interaction between an isolated magnetic ion and the delocalized electrons in a metal is described by the s-d interaction models of various types. Where local d-levels and s-conduction states have overlapping energy, it is convenient to talk in terms ofs-d mixing rather like the hybridization that occurs in pure transition metals, but where the mixing is weak it may be treated as a simple s-d exchange. The mixing approach leads to a virtual bound state model, where the 3d moment is modified by the so-called s-d inreruction. In this picture, which was previously described in Contemponrry Physics by Bell and Caplin (1975),an itinerant (s) electron becomes a temporary resident in a (d) atomic-like state about the ion before tunnelling back into the delocalized states.

Downloaded by [University of Cambridge] at 02:05 06 March 2012

412

C. M. Hurd

During its stay, it experiences the intra-atomic exchange force that couples its spin to that of other resident electrons. Whence the ion’s moment. But it is important to recognize that an ‘isolated’magnetic ion in a metal is a system comprising the ion (‘delectrons’) plus the surrounding itinerant electrons (‘s-electrons’)involved in the s-d interaction. Furthermore, the s-electron population around the ion is spin polarized, opposite to the ion’s net moment, by the effects discussed below. This ‘cloud’ of antiferromagneticspin polarization resonating about the magnetic ion is called Kondo binding or Nagaoka’s compensation after two of its investigators. Below a characteristic temperature ( TK,the Kondo temperature; Bell and Caplin 1975)’ the antiparallel spin cloud neutralizes the ion’s moment and reduces its observable magnitude to zero in the theoretical limit. The Kondo binding is destroyed by increasing temperature and by overlap of clouds from neighbouring ions. The Kondo regime therefore relates ideally to an isolated magnetic ion at absolute zero. This ideal can be approached approximately in a very dilute magnetic alloy at very low temperatures. The opposite extreme occurs when the concentration of magnetic ions is so great that the unfilled shells of neighbouring ions interact sufficiently to form a narrow energy band. The electrons responsible for the ion’s magnetism are then also itinerant to an appreciable extent and the atomic concept of a permanent, localized moment is lost. This occurs in some 3d transition metals and their alloys. The combination ofionic moments and itinerancy ofthe magneticelectrons is called itinerant electron magnetism, and to picture its origins we must consider the constraints acting on itinerant electrons in a narrow band. Firstly, there are exchange effects due to the Pauli principle: electrons of parallel spin stay out of each other’s way to a greater extent than those of antiparallel spin. A pair of electrons of like spin localized on an ion are lower in energy than a pair with opposite spin by an amount called the intra-atomic exchange energy (V). (This is about 1.5eV per spin for 3d electrons.) Consequently, there is a statistical correlation for electrons of like spin, with each surrounded by a void due to the local depletion of parallel-spin electrons. This is called an exchange or Fermi hole. An electron therefore sees a more attractive potential in the presence of parallel spin electrons because the repulsive Coulomb effects are reduced by the exchange hole. (Hence the origin of Hund’s rule for atoms: the ground state of an incomplete shell in a free atom is that of maximum spin.) Secondly, there are dynamical correlation effects: Coulomb repulsion tends to keep electrons ofwhatever spin as far apart as possible, so there is an additional void surrounding any electron-the so-called correlation hole. The cost in energy of the exchange and correlation effects can be reduced only by an increase in an electron’s kinetic energy. Electrons can keep out of each other’s way only by increasing their confinement in space, but the greater an electron’s localization, the higher its kinetic energy. Itinerant electron magnetism is the result of the competition between exchange and kinetic energies. A magnetic ion in this picture is a consequence of the exchange hole. Although the magnetic electrons have some itinerancy, a localized exclusion effect operates for a given spin in the small region ofan ion’s incomplete shell. We can picture a constant interchange ofelectrons by quantum tunnelling between energy states of the shell and the more delocalized ones ofthe band. The narrower the energy bandwidth, the more localized the electrons in it. When the bandwidth is comparable with U,some electrons stay long enough on an ion to interact and to align their spins so that the ion appears to have a local moment. Because of the local exclusion effect, the delocalized electrons are drawn from a pool having a preferred spin orientation, so the ion’s

Downloaded by [University of Cambridge] at 02:05 06 March 2012

Vtrrieties of mtrgnetic order in solids

473

moment tends to be self perpetuating. On the time scale of a magnetic experiment, it appears to be permanent and localized on the ion. Since d electrons have a relatively large exchange interaction compared to their kinetic energy, metals and alloys of the iron, palladium and platinum groups show particularly varied types of magnetic ordering, especially ferromagnetism, antiferromagnetism, and exchange-enhanced paramagnetism, which derive from a reduction in exchange energy at the expense of increased kinetic energy for the magnetic electrons. The Stoner band model of itinerant electron magnetism was one of the earliest formulations of this competition. In this model, the energy states of the itinerant magnetic electrons are split into up- and down-spin bands, separated by the exchange interaction. The Stoner model assumes that the splitting is proportional to the spontaneous magnetisation via the Stoner parameter 1.The criterion for ferromagnetic splitting, IN(&),> 1, expresses the fact that the gain in exchange energy due to splitting exceeds the cost in increased kinetic energy for electrons that have to be promoted to higher energy levels (fig. 2). (N(&),is the density of states at the Fermi energy and is inversely proportional to the change in kinetic energy.)

t

E

Fig. 2. Showingthe density ofelectron statesin energy N(E)for itinerant spin-upand spin-down electrons in an itinerant ferromagnet (Section 3.3). The ferromagnet is known as ‘strong’ (S) or ’weak‘ (W), depending on the position of the Fermi level.

In addition to the idea that a localized moment on an ion is the time-averaged result of a dynamical process, we must also note the existenceof unequal moments ofthe same species in a given solid. This originates from the asymmetric charge distribution of an unfilled d- or f-shell ion, which interacts differently with the charge on neighbouring ions (the ‘ligands’). The exchange forces felt‘ by magnetic electrons in an ion’s incomplete shell therefore depend on the separation and geometrical arrangement of the ligands, particularly the number and arrangement of nearest neighbours of the same species. This dependence of an ion’s moment of its environment is found in crystalline materials-neither Ni nor Co ions, for example, are magnetic in Au but both acquire magnetism when the number of their like nearest neighbours increases sufficiently-but it is most pronounced in amorphous or disordered solids where the atomic sites are inequivalent. There is typically a distribution in magnitude of the ionic moments in these materials that ranges from a narrow distribution for f-shell ions to a very broad one for d-shell ions. In fact, a d-shell ion can sometimes be present in both magnetic and nonmagnetic states in the same material.

Downloaded by [University of Cambridge] at 02:05 06 March 2012

474

C. M. Hurd

2.2. Coupling between moments A magnetic state arising from spontaneous moments in a solid is the net effect of competing influences. The orientation of each moment tends on the one hand to be randomized by its thermal energy while on the other it tends to be aligned by an ordering influence that depends on the magnetism involved. We can distinguish two types of magnetism: the non-cooperative type, where the individual moments behave independently and are unaware ofeach others’ existence, and a cooperative type, where the mutual interactions between the moments are intrinsically important. In a noncooperative magnetic state, the external applied field does the ordering, but in a cooperative one the ordering results from the exchange couplings between the moments. The external field is then frequentlyjust a means of making the microscopic ordering evident on a macroscopic scale. The quantum mechanical coupling between moments in cooperative magnetism is described metaphorically in several different ways depending on the context, but they all derive (see the table) from the influence ofthe Pauli principle and its manifestation as the Fermi hole described in Section 2.1. We can distinguish two classes of exchange coupling. Direct (or contact) exchange operates between moments on ions that are close enough to have significant overlap of their wave functions; it gives a strong but shortrange coupling which decreases rapidly as the ions are separated. Intlirecr exchange, o n the other hand, couples moments over relatively large distances. It acts through an intermediary which in metals can be the itinerant electrons or in insulators can be

Hierarchy of exchange coupling.

P

P The Pauli exclusion principle is the basis of all exchange forces.

E: An exchange interaction is a metaphorical description of the effects of the Pauli principle on the Coulomb repulsion between fermions. I: Indirect exchange is a coupling between quantum systems so far apart that some intermediary must be involved.

D Direct exchange is a coupling between quantum systems close enough to have overlapping wave functions.

R RKKY is an indirect exchange where itinerant electrons are the intermediaries.

S: Superexchange is an indirect exchange where the intermediary is a ligand.

DM,: The Dzyaloshinsky-Moriya coupling occurs when the spin information between the indirectly coupled systems is upset asymmetrically by spin-orbit effects. In this version, itinerant electrons are the intermediaries. DM,: As in DM, except in this version the spin-orbit coupling occurs at an intermediate ligand.

Vcirieties of magnetic order in solids

475

nonmagnetic ions in the lattice. The coupling is then known as the RKKY or superexchange effect in metals or in insulators, respectively. We have seen that the effective electrostatic interaction between two electrons depends on the relative orientations of their magnetic moments (or spins). This is conveniently expressed as a spin-dependent coupling that is assumed to be isotropic and to depend only on the distance between interacting ions. Thus if ions i and j , separated by a distance rij, have spins Si and Sj,respectively, the exchange energy HEis written:

H~= - C J(rij)S i * S j

Downloaded by [University of Cambridge] at 02:05 06 March 2012

ij

where J , is called the exchange parameter. For direct intra-atomic exchange-as between two electrons belonging to the same atom-J is positive leading to Hund’s rule. But for direct inter-atomic exchange, J can be positive or negative depending on the balance between Coulomb and kinetic energies as described in Section 2.1. In indirect exchange, J can be positive or negative, as in superexchange between magnetic ions in insulators (discussed at the end of this Section), or oscillatory as in the so-called RKKY interaction. The RKKY interaction-named after its principal investigators, Ruderman and Kittel, Kasuya and Yosida-is unique because J oscillates from positive to negative as the separation between the ions changes. It is restricted to materials containing itinerant electrons, which are the intermediaries in the coupling. A magnetic ion induces an oscillatory spin polarization in the conduction electrons in its neighbourhood. The reason why this polarization is oscillatory is that the conduction electrons try to screen out the magnetic moment on the ion by the means of their spins (just as by means of their charge they try to screen out the charge on an ion) but their wavefunctionshave a limited range of wavelengths (or wave numbers). In the simplest case ofa degenerate gas offree electrons the highest wavenumber available is 2kFwhere kF is the wave number of an electron on the Fermi surface. The process is analogous to the representation of a non-oscillatory function by means of an incomplete set of Fourier components: there is finally a residual oscillatory behaviour in the representation. The strength of this screening polarization decreases with increasing distance from the ion, but its effect has a relatively long range. This modulated spin polarization in the itinerant electrons is felt by the moments of other magnetic ions within range, leading to an oscillatory, indirect coupling. The existence of RKKY coupling means that in a disordered metallic system, where the separation between magnetic ions is random, positive or negative coupling between moments can be found. This leads to possible conflicts in the system on a microscopic scale as moments try to respond to antagonistic constraints (fig. 3). This is called frusfrution,a concept that has wider implications in other branches of science where it is sometimes called ‘structural disequilibrium’. A frustrated system is one which, not being able to achieve’astate that satisfies entirely its microscopic constraints, possesses a multiplicity of equally unsatisfied states. A frustrated system therefore has no unique microscopic arrangement for its ground state; there is an essentially infinite number of equivalent states that can be adopted. As a result, a frustrated magnetic system shows metastability, with hysteresis effects, time-dependent relaxation towards an equilibrium state, or dependence on the sample’s thermal or magnetic histories. Frustration is also possible in some ordered insulating magnetic materials. Thus MnO has an f.c.c. lattice of Mn2+ions in which nearest neighbours of a given ion are nearest neighbours

C. M. Hurd

Downloaded by [University of Cambridge] at 02:05 06 March 2012

476

Fig. 3. A triangular lattice of magnetic ions that can b e ’ u p or down-spin illustrates ‘frustration’. When the exchange parameter J between all moments is positive (a),the constraintthat all moments be parallel to their neighburs can be satisfied,so the system is nonfrustrated. But if J is always negative, or positive and negative (b), there is no arrangement that satisfies all the microscopic constraints.The system is thus hstrated.

.of one another, and since the n.n. coupling is antiferromagnetic, TNis lowered and the ordered structure complex. Equation (1) has shortcomings, particularly for the interpretation of the newer types of magnetic order, because it involves only the interatomic distances as structural information. It does not account for an exchange interaction that is anisotropic-that is, depends on the direction in the material. Such anisotropy arises bom a coupling between an ion’s moment and the lattice and can have two sources. One is the anisotropic geometrical arrangement of the ligands, leading to an anisotropic electric field at the ion (the ‘crystal field’). The other is spin-orbit coupling. Let us consider first the case when the ligand field effect is dominant. Imagine an amorphous material where the crystal field varies from point to point. Each magnetic ion will have a preferred alignment for its moment along a local ‘easy axis’ determined by the local crystal field. This is called single-site anisotropy, and to include its effect in equation (1)it is convenient to assume the case of uniaxial symmetry. The crystal field energy at the ith site is then of the form -DS;,where D is the axial crystal field strength and S, is the total spin of the ion along the local easy direction z (Kanamori 1963).This must be summed over all sites and added to equation (1) to give the energy for the case with both the local anisotropy and exchange effects:

H = -ID&); i

-zJ(rij)S;Sj ri

In an amorphous solid, D is fixed and positive but the easy axes, zi, are randomly distributed in direction. The first term is usually insignificant compared with the second for d-shell ions in an amorphous metal but the reverse usually applies for f-shell ions. Spin-orbit coupling is the other source of anisotropy. This coupling, which is a relativistic effect,acts on an electron’s intrinsic moment due to the effective magnetic

Downloaded by [University of Cambridge] at 02:05 06 March 2012

Vurieties of rntrgneric order in solitls

477

field of its own orbiting charge (Cullity 1972, Hurd 1975). The net spin and orbital motions of the unpaired electrons in a magnetic ion are therefore coupled. Generally, the orbital motion is also coupled to the crystal field via the Coulomb force acting between the charge distribution in the unfilled orbital and the electric field from neighbouring ligands. So there is effectively a coupling between an ion’s moment and the crystal field. Spin-orbit coupling is thus a component of the single-site anisotropy described above, but here we are concerned with its role leading to anisotropic exchange between pairs of magnetic ions, coupled through superexchange. Superexchange was introduced to describean interaction between moments on ions too far apart to be connected by direct exchange but coupled over a relatively long distance through an intervening, non-magnetic ligand (White and Geballe 1979). Different forms of superexchangehave been postulated for different circumstances, but we consider the prototypical case of coupling between the moments on a pair of metal cations separated by a diamagnetic anion, as in a magnetic insulator. This trio can be regarded as a trimeric molecule, where superexchange is a consequence of the hybridization of overlapping orbitals. Figure 4 is a sketch of two cases: an axiallysymmetric molecule with strong spin-orbit coupling on one of the cations, and an axially-asymmetric molecule with spin-orbit coupling on the anion. The former is illustrated by the rare earth-iron (R-Fe) interaction in a garnet. The femc ion has a half-filled 3d shell and so has a spherically symmetric charge distribution ( S state ion). The triply-ionized rare-earth ion, on the other hand, is not symmetric and has a strong spin-orbit coupling; its charge distribution is coupled to its moment. The ion’s moments are coupled via superexchange,so turning the Fe moment alters the overlap ofthe R cation in the molecule(fig.4). This changes the magnitude ofboth the Coulomb and exchange interactions between the cations, leading to a coupling which depends on the moment’s orientation. Thus J in equation (1) depends on the orbital state ofthe rate earth ion, and the exchange energy is anisotropic.

““‘+WR3+ $+ S2’

Fig. 4. Superexchange combined with spin-orbit coupling can have two consequences sketched here: anisotropicexchange, as in the ferric-rare earth interaction in garnets, and the antisymmetric exchange (or Dzyaloshinsky-Moriya coupling) of P-MnS.

Downloaded by [University of Cambridge] at 02:05 06 March 2012

478

C.M.Hurd

The second example (fig. 4)is illustrated by 8-MnS. This has a cubic zinc blende structure but the ionic arrangement in certain planes is asymmetric with respect to the line of cation centres. When there is strong spin-orbit coupling of the anion, this arrangement allows the so-called DzyaloshinskpMoriya (D-M) interaction to couple the moments on the cations (Moriya 1963). The interaction arises because the spin information carried between the cations by the delocalized electrons of superexchange-or by itinerant electrons in the RKKY version of D-M coupling (Smith 1976)- is upset by the spin-orbit coupling in the anion orbitals. The upset depends on the direction of electron transfer imagined between the cations, and its net effect is zero when averaged over all equivalent molecular configurations of ions in the crystalunless the anion breaks the inversion symmetry with respect to the mid-point of the cations’centres. Then the combination of superexchangeand spin-orbit coupling gives a D-M interaction between the cations of the form HDM=dij(Six Sj) in place of equation (1)(Keffer 1962). This mechanism is important in systems called weak ferromagnets (not to be conhsed with the ‘weakitinerant ferromagnets’of Section 3.3) where the Si and Sj form separate sublattices with antiferromagnetic alignment (Moriya 1963).The lattices are equivalent but not exactly antiparallel, leaving a net magnetization. (Examples are given in Section 4.6.)In some cases HDMis opposed by a strong tendency towards ferroor antiferromagnetic alignment from exchange of the type expressed in equation (1). The D-M energy can then be reduced by canting of the sublattices, but at a cost of increased exchange energy. The cant angles in these systems are small. But in systems where the D-M and exchange energies are not in conflict the cant angles can be up to n/2,as in P-MnS sketched in fig. 4 (Keffer 1962).

3. Five basic types of magnetic order In this Section we review briefly the basic magnetic states familiar from any standard textbook (Cullity 1972, for example) and shown in the top row of fig. 1. 3.1. Diamagnetism We can regard diamagnetism as the result of shielding currents induced in an ion’s filled electron shells by an applied field. The currents are equivalent to an induced moment on each ion in the substance. Ideal diamagnetism is a noncooperative magnetism (Section 2 -2) characterized by a negative, temperature-independent magnetic susceptibility x (fig.5 ) . Diamagnetism is part ofall magnetic states, but it is usually neghgiblecompared with the magnetism arising fiom any spontaneous moments in the system. Examples ofdiamagnetic solids are Cu and NaCl fiom the classes ofcrystalline metals and insulators, and SiO, fiom the class of amorphous solids.

3.2.Ideal paramagnetism This is another non-cooperative magnetism but it arises from spontaneous moments which, in the ideal case, are identical and located in isotropic surroundings ( D = O in equation 2), sufficiently separated to be independent of the others’ existence. The orientation of each moment is on the one hand randomized by its thermal energy k,Twhile on the other it is aligned by an externally applied magnetic field. The magnetic order is thus the net alignment achieved by the applied field in the face of the thermal disarray existing at the ambient temperature. The measured magnetization is the time-averaged alignment of the moments, for their orientations fluctuate constantly and are in no sense fixed (as suggested by the arcs in fig. 6 et seq.). Ideal paramagnetism

Viirieties of magnetic order in solids

479

-I AMORPHOUS

CRYSTALLINE 0

0

0

0

0

0

0

0

0

0

0

0

0

0

0

0

0

0

0

0

Downloaded by [University of Cambridge] at 02:05 06 March 2012

0

O

00

oooo

o o o 0

Cu ,Na Cl

O

o o 0

0

S iO2

Fig. 5 . Diamagnetism.

0-0 CRYSTALLINE

I

AMORPHOUS

Fig. 6. Ideal paramagnetism.

is characterized by a susceptibility x which varies inversely with temperature T ( x = C/T;the Curie law, fig. 6). C is called the Curie constant. Ideal paramagnetism is the exception rather than the rule because there is normally an appreciable exchange coupling between the ionic moments in a solid. Curie's law is only a special case of the more general CurieWeiss law, x = C/(T- e), where 8 is a constant which expresses the interionic coupling. Curie-Weiss behaviour (fig. 6) thus implies a cooperative magnetism that can dominate if the thermal energy of the moments in the paramagnet is reduced sufficiently. As temperature is reduced, this domination occurs at what is, for practical purposes, a discrete temperature (T= TORD). Below TORD, a different magnetic state exists. When it is ferro- or ferrimagnetism, then ToRo is called the Curie temperature (TJ; it is called the Nee1 temperature ( TN)for antiferromagnetism;the.spin-glass temperature (7&)for the spin-glass state; and the

C. M. Hurd

480

freezing temperature ( TF)for mictomagnetism. Sketches of metallic paramagnets in crystalline and amorphous forms are given in fig. 6. In each case the group 1B metal is, of course, the nonmagnetic constituent.

Downloaded by [University of Cambridge] at 02:05 06 March 2012

3.3. Ferromagnetism

Ferromagnetism is a cooperative magnetism (Section 2.2) leading to long-range colinear alignment of all the moments in the system (fig. 7). There is thus a magnetization even with no external field (the spontaneous magnetization). In ideal ferromagnetism, every ion has an identical, spontaneous moment and occupies an identical crystallographic site. The constraints acting on a moment are dominated by the inter-ion exchange coupling (IJI >> 101)and the exchange parameter is everywhere positive (J >0). To maximize its magnetostatic energy, a crystalline sample usually divides into domains which are spontaneously magnetized, nearly or completely to saturation, along a direction of easy magnetization determined by D . An external field H can change the size of these domains, enlarging those of favourable orientation between D and H at the expense of others, but it can make little difference to the intrinsic magnitude of magnetization within a domain. A ferromagnet is an example ofmagnetism where an external field isjust an agent to make evident on a macroscopic scale the ordering that exists microscopically. The field dependence of the reduced magnetization M/M,is typically (fig. 7) sharply rising at lower fields, as the domains with more favourable alignments expand at the expense of the others, and saturating when the maximum domain alignment is achieved. (This is called technical saturation.) Thereis a slight field dependence above saturation (which is called the paraprocess range in the German and Russian literature) due to the applied field's effect on the alignment within a domain. With increasing temperature in any cooperative state, a point is reached at which the moments' thermal energies are comparable to, and eventually exceed, the exchange energies. This occurs at the Curie temperature T, in a ferromagnet. The spontaneous magnetization u, decreases with temperature (fig. 7) to disappear at T,. Above T,, an

IJblDl J>O C K W A LLIN E

AMORPHOUS

Fe

Feeob.Feh

Fig. 7. Ideal ferromagnetism.

Downloaded by [University of Cambridge] at 02:05 06 March 2012

Vm-ieties of riitrgiieric order in soliils

48 1

ideal ferromagnet becomes a paramagnet and obeys a Curie-Weiss law; 1/x is linear in T. Iron, Co and Ni are examples of elemental, crystalline ferromagnets, and various compositions of Fe, Ni, Co, B (known by the trade name ‘Metglas Alloys’) are typical amorphous ferromagnets (Cahn 1980). Iron, Co and Ni, to different degrees, are also examples of itinerant electron ferromagnets(Section2.1). In the Stoner picture (fig.2), the ferromagnetismcomes from the unequal populations of up- and down-spin itinerant magnetic electrons. When the majority-spin band is completely full in this picture-as for S in fig. 2-the system is referred to as a strong itinerantferrornagnet;‘strong’meaning that hrther band splitting cannot increase the magnetism. When itinerant magnetic electrons are contained in both bands (as for W in fig. 2), the system is called a weak itinerantfenornagnet. Then the spontaneous magnetism depends crucially on the shape of the N(E)curve close to the Fermi energy.This can be sensitive to both applied field and temperature changes, leading to metamagnetism (Section 4.1) which may be described as thermal spontaneous mognetizutioii as in Y2Ni, (Gignoux et (11. 1980). Note that this ‘weak itinerant ferromagnetism’should not be confused with the ’weak ferromagnetism’of Sections2.2 and 4.6. Ferromagnetism and superconductivity should be incompatible if the scattering of electrons by the ordered magnetic ions interferes with Cooper pairing. The onset of long-range ferromagnetism can indeed destroy superconductivity, but very recent work on some exotic ternary compounds of holmium molybdenum sulphide (HOMO,$,) and particularly erbium rhodium boride (ErRh,B,) has shown that ferromagnetism and superconductivity can coexist over limited temperature ranges. The result is a strange state that is still controversial (Mook et al. 1982)but appears to be a mixture of normal but very small ferromagnetic domains and superconducting regions with sinusoidally modulated magnetic moments. It was suggested about 25 years ago that magnetic moments in a superconductor would find long-range ferromagnetism less favourable than a ferromagnetism broken into extremely small domains-this is so-called c.r.~ptc!frrromuyrietisrlr(Anderson and Suhl 1959kbut these arguments assumed a strong exchange coupling between the spins of the itinerant electrons and those of the magnetic ions. Such is not the case in the above compounds where the 4felectrons, tightly bound in the ion, have only a small interaction with the superconducting electrons. It is currently believed that electromagnetic rather than exchange interactions govern the competition between ferromagnetism and superconductivity in these compounds. 3.4. Ant@woinugnetisrn Antiferromagnetism, like ferromagnetism, is a cooperative magnetism of longrange order among identical, spontaneous moments. Ideally, the magnetic ions occupy crystallographicallyidentical sites. The exchange coupling dominates the constraints on the moments (IJI >> 1 0 1 )but the exchange parameter is negative (J