Inequalities for Isothermal Linear Response

8
Aust. J. Phys., 1994, 47, 129-35 Inequalities for Isothermal Linear Response A. M. Stewart Department of Applied Mathematics, Research School of Physical Sciences and Engineering, Australian National University, Canberra, A.C.T. 0200, Australia. Abstract The inequality that Brown, Hornreich and Shtrikman (1968) reported for the magneto-electric susceptibility is shown to be valid quite generally for all isothermal linear response coefficients. Certain difficulties in the derivation are removed. 1. Introduction In 1968 Brown, Hornreich and Shtrikman reported an inequality that stated that the square of the magneto-electric susceptibility of any material was less than or equal to the product of its paraelectric and paramagnetic susceptibilities. Their derivation contained two shortcomings: first, they did not obtain explicit expressions for these susceptibilities, and second, their derivation did not appear to be applicable to materials that contain degenerate energy levels because of divergent denominators in their expression for the free energy. This rather large class of materials includes all paramagnetic systems containing unpolarised local moments. In this paper it is shown how these shortcomings may be circumvented and that inequalities (15) and (16) may be obtained that are valid for all suitably defined isothermal linear response coefficients, not only magneto-electric ones. 2. Free Energy Consider a system in thermal equilibrium at a constant temperature T to which a perturbation V is applied of the form: v = - LOsFs, s (1) where the F s are small steady applied classical multipolar electric, magnetic or exchange fields which couple to quantum mechanical operators Os of the system. As one particular example V might be given by -p.E - m.B - 2J.LB S.B ex , where p,m {=J.LB(L+2S)} and S are the operators for the vector electric and magnetic dipole moments and the spin of the system. Here E and Bare the applied electric and magnetic fields and B ex is a fictitious exchange field 0004-9506/94/020129$05.00 Aust. J. Phys., 1994, 47, 129-35 Inequalities for Isothermal Linear Response A. M. Stewart Department of Applied Mathematics, Research School of Physical Sciences and Engineering, Australian National University, Canberra, A.C.T. 0200, Australia. Abstract The inequality that Brown, Hornreich and Shtrikman (1968) reported for the magneto-electric susceptibility is shown to be valid quite generally for all isothermal linear response coefficients. Certain difficulties in the derivation are removed. 1. Introduction In 1968 Brown, Hornreich and Shtrikman reported an inequality that stated that the square of the magneto-electric susceptibility of any material was less than or equal to the product of its paraelectric and paramagnetic susceptibilities. Their derivation contained two shortcomings: first, they did not obtain explicit expressions for these susceptibilities, and second, their derivation did not appear to be applicable to materials that contain degenerate energy levels because of divergent denominators in their expression for the free energy. This rather large class of materials includes all paramagnetic systems containing unpolarised local moments. In this paper it is shown how these shortcomings may be circumvented and that inequalities (15) and (16) may be obtained that are valid for all suitably defined isothermal linear response coefficients, not only magneto-electric ones. 2. Free Energy Consider a system in thermal equilibrium at a constant temperature T to which a perturbation V is applied of the form: V= - LO.F., (1) where the F. are small steady applied classical multipolar electric, magnetic or exchange fields which couple to quantum mechanical operators O. of the system. As one particular example V might be given by -p.E - m.B - 2/1B S.Bex , where p, m {=/1B(L+2S)} and S are the operators for the vector electric and magnetic dipole moments and the spin of the system. Here E and Bare the applied electric and magnetic fields and Bex is a fictitious exchange field 0004-9506/94/020129$05.00

Transcript of Inequalities for Isothermal Linear Response

Aust. J. Phys., 1994, 47, 129-35

Inequalities for Isothermal Linear Response

A. M. Stewart

Department of Applied Mathematics,Research School of Physical Sciences and Engineering,Australian National University, Canberra, A.C.T. 0200, Australia.

Abstract

The inequality that Brown, Hornreich and Shtrikman (1968) reported for the magneto-electricsusceptibility is shown to be valid quite generally for all isothermal linear response coefficients.Certain difficulties in the derivation are removed.

1. Introduction

In 1968 Brown, Hornreich and Shtrikman reported an inequality that statedthat the square of the magneto-electric susceptibility of any material was lessthan or equal to the product of its paraelectric and paramagnetic susceptibilities.Their derivation contained two shortcomings: first, they did not obtain explicitexpressions for these susceptibilities, and second, their derivation did not appearto be applicable to materials that contain degenerate energy levels because ofdivergent denominators in their expression for the free energy. This rather largeclass of materials includes all paramagnetic systems containing unpolarised localmoments. In this paper it is shown how these shortcomings may be circumventedand that inequalities (15) and (16) may be obtained that are valid for all suitablydefined isothermal linear response coefficients, not only magneto-electric ones.

2. Free Energy

Consider a system in thermal equilibrium at a constant temperature T towhich a perturbation V is applied of the form:

v = - LOsFs,s

(1)

where the F s are small steady applied classical multipolar electric, magnetic orexchange fields which couple to quantum mechanical operators Os of the system.As one particular example V might be given by

-p.E - m.B - 2J.LB S.Be x ,

where p,m {=J.LB(L+2S)} and S are the operators for the vector electric andmagnetic dipole moments and the spin of the system. Here E and Barethe applied electric and magnetic fields and B ex is a fictitious exchange field

0004-9506/94/020129$05.00

Aust. J. Phys., 1994, 47, 129-35

Inequalities for Isothermal Linear Response

A. M. Stewart

Department of Applied Mathematics, Research School of Physical Sciences and Engineering, Australian National University, Canberra, A.C.T. 0200, Australia.

Abstract

The inequality that Brown, Hornreich and Shtrikman (1968) reported for the magneto-electric susceptibility is shown to be valid quite generally for all isothermal linear response coefficients. Certain difficulties in the derivation are removed.

1. Introduction

In 1968 Brown, Hornreich and Shtrikman reported an inequality that stated that the square of the magneto-electric susceptibility of any material was less than or equal to the product of its paraelectric and paramagnetic susceptibilities. Their derivation contained two shortcomings: first, they did not obtain explicit expressions for these susceptibilities, and second, their derivation did not appear to be applicable to materials that contain degenerate energy levels because of divergent denominators in their expression for the free energy. This rather large class of materials includes all paramagnetic systems containing unpolarised local moments. In this paper it is shown how these shortcomings may be circumvented and that inequalities (15) and (16) may be obtained that are valid for all suitably defined isothermal linear response coefficients, not only magneto-electric ones.

2. Free Energy

Consider a system in thermal equilibrium at a constant temperature T to which a perturbation V is applied of the form:

V= - LO.F., (1) •

where the F. are small steady applied classical multipolar electric, magnetic or exchange fields which couple to quantum mechanical operators O. of the system. As one particular example V might be given by

-p.E - m.B - 2/1B S.Bex ,

where p, m {=/1B(L+2S)} and S are the operators for the vector electric and magnetic dipole moments and the spin of the system. Here E and Bare the applied electric and magnetic fields and Bex is a fictitious exchange field

0004-9506/94/020129$05.00

130 A. M. Stewart

often useful in the treatment of magnetic systems. The F s are the individualcomponents of these fields.

As a result of the perturbation, the eigenvalues of the Hamiltonian H thatdescribes the system are changed from E; to E/ where, to second order inperturbation theory,

I

E/ = Ei + (ilVli) - L IUlVli)12

,j E j - E i

(2)

and the prime on the sum over j means that the j = i term is to be omitted.It is assumed for the time being that the quantum states Ii) of the unperturbedsystem are non-degenerate.

The change of the eigenvalues following the application of the fields F s resultsin a change of the partition function Z = ~iexp(-{3E/), where (3 = 1/kT andk is Boltzmann's constant. This in turn perturbs the Helmholtz free energyF = - kT InZ, where F = E - TS, S being the entropy and E = (H) the internalenergy, given by the thermal expectation value of the Hamiltonian operator. Byexpanding in powers of the small quantity V it follows that

F - R = (V) _! ". "IVJiI2(pi

- Pj) - _1_ ((y:. - (V) )2) (3)o 0 2 L...J ~ E: _ E, 2kT ~~ 0 0 ,

J

where V j i = (jIVli), F o is the free energy of the unperturbed system,Pi = exp( -(3Ei ) / Zo, Zo = ~i exp(-{3Ei) is the partition function of the un­perturbed system, and the angle brackets with subscript 0 denote ~i Pi, theensemble average in the unperturbed state. Angle brackets without this subscriptindicate the ensemble average in the perturbed state. Lifshitz and Pitaevskii(1980) attribute the above result to R. E. Peierls. An important feature of (3)is that the change in free energy that is of second order in the perturbation isnegative-definite (~ 0) because Pi > Pj if e, > e.. The first-order term, givenin our case by - ~s F s (0 s)o, may be of either sign. An assumption implicit inthe derivation is that the magnitude of the energy perturbation per particle ismuch less than kT.

The last term in (3) may be expressed as

(V)6/2kT - L exp( -(3Ei ) (iIVli) (iIVli) /2kTZo.i

Because of the denominator, the form of the second term on the right-hand sideof (3) is not defined when the states i and j are degenerate; unpolarised magneticmaterials will possess such states. To circumvent this difficulty, the limit

(e-/3Ei - e-/3Ej)/(Ej - E i ) -+ {3e-/3Ei

must be taken, and then these degenerate terms, together with the last term of(3), result in the degenerate parts of the second term of (3) being summed overall j with E j = E i . In addition, for the non-degenerate terms, the two terms

130 A. M. Stewart

often useful in the treatment of magnetic systems. The F 8 are the individual components of these fields.

As a result of the perturbation, the eigenvalues of the Hamiltonian 'Ii that describes the system are changed from Ei to E/ where, to second order in perturbation theory,

(2)

and the prime on the sum over j means that the j = i term is to be omitted. It is assumed for the time being that the quantum states Ii) of the unperturbed system are non-degenerate.

The change of the eigenvalues following the application of the fields F s results in a change of the partition function Z = Ei exp( -(3E/) , where (3 = l/kT and k is Boltzmann's constant. This in turn perturbs the Helmholtz free energy F = -kTlnZ, where F = E-TS, S being the entropy and E = ('Ii) the internal energy, given by the thermal expectation value of the Hamiltonian operator. By expanding in powers of the small quantity V it follows that

(3)

where V ji = UlVli), Fo is the free energy of the unperturbed system, Pi = exp( -(3Ei)/ Zo, Zo = Ei exp( -(3Ei) is the partition function of the un­perturbed system, and the angle brackets with subscript 0 denote Ei Pi, the ensemble average in the unperturbed state. Angle brackets without this subscript indicate the ensemble average in the perturbed state. Lifshitz and Pitaevskii (1980) attribute the above result to R. E. Peierls. An important feature of (3) is that the change in free energy that is of second order in the perturbation is negative-definite (:=:; 0) because Pi > Pj if E j > E i • The first-order term, given in our case by - Es F s ( 0 8 )0, may be of either sign. An assumption implicit in the derivation is that the magnitude of the energy perturbation per particle is much less than kT.

The last term in (3) may be expressed as

(V)~/2kT - L exp(-{3Ei) (ilVli)(ilVli)/2kTZo. i

Because of the denominator, the form of the second term on the right-hand side of (3) is not defined when the states i and j are degenerate; unpolarised magnetic materials will possess such states. To circumvent this difficulty, the limit

must be taken, and then these degenerate terms, together with the last term of (3), result in the degenerate parts of the second term of (3) being summed over all j with E j = E i . In addition, for the non-degenerate terms, the two terms

Inequalities for Isothermal Linear Response 131

with Pi and Pi are equal, so (3) may be expressed in the following alternativeform, in which the terms involving degenerate and non-degenerate states havebeen separated and the divergent denominators are absent:

with

F - Fo == (V)o - !Q + (V)6/2kT , (4)

Q = ~o ~ e-{3Ei~ IVJil2 {DEj,EjkT + 2(1 - DEj,E,)/(Ej - E i )} . (5)'1, 1

The purpose of the Kronecker delta is to indicate that only terms with E j == E;are to be included in the first (Curie) term of (5) and only terms with E j i= E;in the second (the Van Vleck term). The quantity Q is real and positive-definite,because exp(-{3E i ) is greater than exp(-{3E j ) if E j > E i . It also follows thatQ ~ (V)6/ kT if the second-order free energy is to be negative-definite (see theAppendix). The two equations above may also be obtained from the first threeterms of the expansion of the expression:

z / ({f3 h/

27r) )

Zo = T \ exp - 271" 10 V(r) dr/h 0'

where V (T) is in the Heisenberg picture with imaginary time T == it and T isan operator that orders the arguments of the terms in the expansion of theexponential.

3. Susceptibilities

It is useful to consider an isothermal susceptibility R r s defined by

(Or) == (Or)O + L n., Fs .s

(6)

The explicit form of this susceptibility may be obtained by calculating (Or)by perturbation theory (Stewart 1993a). A less direct but shorter derivationmakes use of a standard result of quantum statistical mechanics (Lifshitz andPitaevskii 1980) that the change dF in free energy resulting from changes inexternal parameters and temperature is given by

dF == -8 dT + L (81-l/8Ai) dAi' (7)

where the Ai are external parameters such as the volume or the fields F s thatdetermine the energy eigenvalues of the quantum system. In the notation usedhere this gives

dF == -8dT - L (Os)dFs,s

Inequalities for Isothermal Linear Response 131

with Pi and Pj are equal, so (3) may be expressed in the following alternative form, in which the terms involving degenerate and non-degenerate states have been separated and the divergent denominators are absent:

F - Fo = (v)o - ~Q + (V)6/2kT , (4)

with

The purpose of the Kronecker delta is to indicate that only terms with E j = Ei are to be included in the first (Curie) term of (5) and only terms with E j -=f. Ei in the second (the Van Vleck term). The quantity Q is real and positive-definite, because exp( -(3Ei) is greater than exp( -(3E j ) if E j > Ei. It also follows that Q 2: (V)5/ kT if the second-order free energy is to be negative-definite (see the Appendix). The two equations above may also be obtained from the first three terms of the expansion of the expression:

Z / (r(3h/27r ) ) Zo =T\exp -27r Jo V(r)dr/h 0'

where V(r) is in the Heisenberg picture with imaginary time r = it and Tis an operator that orders the arguments of the terms in the expansion of the exponential.

3. Susceptibilities

It is useful to consider an isothermal susceptibility Rrs defined by

(6) 8

The explicit form of this susceptibility may be obtained by calculating (Or) by perturbation theory (Stewart 1993a). A less direct but shorter derivation makes use of a standard result of quantum statistical mechanics (Lifshitz and Pitaevskii 1980) that the change dF in free energy resulting from changes in external parameters and temperature is given by

(7)

where the Ai are external parameters such as the volume or the fields F 8 that determine the energy eigenvalues of the quantum system. In the notation used here this gives

132

or for the example given previously,

dF = -8 dT - p.dE - m.dB - 2J.-tB S.dBex .

A. M. Stewart

Hence (Or) = -(8F/8Fr)IT. The next step is to substitute the explicit form ofthe perturbation (1) into (5) and, noting that F r occurs twice in the second-orderpart of (5), differentiation with respect to F r of that equation leads to

where

Rrs = Xrs - (Or)O (Os)o/kT , (8)

Xrs ~o ~:>-,8Ei "2;38Ei,EJkT + 2(1 - 8Ei,E.)/(Ej - E i )}

't J

X Re{ (iIOrlj) (jlOs Ii)} . (9)

It can be seen from inspection that Xrs is real, that Xsr = Xrs, and that Xrr ~ o.The second term on the right-hand side of (8) is zero when the system is initiallyunpolarised with <Or>O or <o.>« equal to zero. Equation (9) may also bederived from the time integral of the Kubo (1957) formula

Xrs(t, t') = 2~ i ([Or(t), o,(t')])o B(t - t') ,

where the square brackets denote a commutator, the operators are in theHeisenberg picture and the function 8(x) = 1 for x > 0 and 0 otherwise.

At zero temperature the system is in its (non-degenerate) ground state 10),so the Kronecker delta gives zero and

,Xrs = 2L Re{(OIOrli) UIOsIO)}

j E j - Eo .(10)

Since the condition that the energy perturbation is much less than kT is notsatisfied at zero temperature, the calculation needs to be carried out in termsof the perturbation of the ground-state energy and wavefunction. This leads to(10) but Rrs = Xrs instead of (8). Because the ground-state is non-degenerateXrr is again positive-definite. At temperatures much larger than the separationbetween energy levels (assumed finite and N in number), both terms of (9) maybe expanded in powers of 1/ T to give

Xrs = L LRe{(iIOrlj) (jIOsli)}/NkT,j

(11)

or, by closure, assuming that the Ii) form a complete set of statesXrs = Trace{ Or Os}/NkT, or Xrs = (Or Os)o/kT, where the correlation functionis evaluated in the high-temperature limit. Hence, for example, the Curie constantof a paramagnetic ion becomes equal to the free-ion value at temperatures much

132 A. M. Stewart

or for the example given previously,

dF = -8 dT - p.dE - m.dB - 2/LB S.dBex.

Hence (Or) = -(8F/8Fr)IT. The next step is to substitute the explicit form of the perturbation (1) into (5) and, noting that Fr occurs twice in the second-order part of (5), differentiation with respect to F r of that equation leads to

(8)

where

Xrs 1

Zo L e-,6Ei L{8Ej ,EjkT + 2(1- 8Ej ,E,)/(Ej - Ei )}

• J

X Re{ (iIOrlJ) (jIOs Ii)} . (9)

It can be seen from inspection that Xrs is real, that Xsr = Xrs, and that Xrr ~ O. The second term on the right-hand side of (8) is zero when the system is initially unpolarised with <Or>O or <Os>o equal to zero. Equation (9) may also be derived from the time integral of the Kubo (1957) formula

Xrs(t, t') = 2~i ([Or(t), Os (t')])o B(t - t'),

where the square brackets denote a commutator, the operators are III

Heisenberg picture and the function B(x) = 1 for x> 0 and 0 otherwise. At zero temperature the system is in its (non-degenerate) ground state

so the Kronecker delta gives zero and

the

10),

(10)

Since the condition that the energy perturbation is much less than kT is not satisfied at zero temperature, the calculation needs to be carried out in terms of the perturbation of the ground-state energy and wavefunction. This leads to (10) but Rrs = Xrs instead of (8). Because the ground-state is non-degenerate Xrr is again positive-definite. At temperatures much larger than the separation between energy levels (assumed finite and N in number), both terms of (9) may be expanded in powers of 1/ T to give

Xrs = L LRe{(iIOrlj) (jIOsli)}/NkT, j

(11)

or, by closure, assuming that the Ii) form a complete set of states Xrs = Trace{ Or Os}/ NkT, or Xrs = (Or Os)o/kT, where the correlation function is evaluated in the high-temperature limit. Hence, for example, the Curie constant of a paramagnetic ion becomes equal to the free-ion value at temperatures much

Inequalities for Isothermal Linear Response 133

greater than the crystal field splitting, but not so large as to cause thermalpopulation of multiplet levels above the ground state.

4. Inequalities

By substituting (8), (9) and (1) into (4) and (5), it follows that the changein free energy may be expressed as

F - Fo = - L Fs(Os)o - ! L L r: r, Rrs, (12)s r s

with Rrs = -(82Fj8Fr8Fs) IT. Since the second-order free energy in (12) isrequired to be negative-definite for all values of the applied fields, it follows first,by setting all the fields except T to be zero, that Rrr 2 0 or (see the Appendix)

Xrr 2 (Or)~jkT. (13)

At zero temperature the relation is Xrr 2 o. Second, if all the fields are setto zero except for the pair F rand F s, the second-order free energy becomesthe negative of (F; Rss + F~ Rrr + 2F s Fr Rrs), which may be written in thealgebraically identical form

(JRrr r; + J;,:r Fsr+ F; (Rss - R~s/Rrr) ·

For the free energy to be negative-definite, it follows that Rss Rrr 2 R~s for allpairs of T, s. By substituting (8) into this relation we obtain

(XrrXss - X~s)(Xss - (Os)~/kT) ~ ( ,JX;;(Or)O - ~(Os)o) 2 Xss/kT. (14)

Because the right-hand side is greater than zero, and by using (13), we finallyget the inequality

2Xrr Xss 2 Xrs . (15)

By considering more than two fields to have nonzero values the theory of quadraticforms leads to relations of the type (Perlis 1952):

tc; n.; u.,Det I «; ti., ti; I~ 0 ,

s.; u; s;

where Det indicates the determinant of the matrix, or

n.; tc: Rtt + ui; «; Rst 2 0 ,

and so forth.

(16)

Inequalities for Isothermal Linear Response 133

greater than the crystal field splitting, but not so large as to cause thermal population of multiplet levels above the ground state.

4. Inequalities

By substituting (8), (9) and (1) into (4) and (5), it follows that the change in free energy may be expressed as

(12) s r s

with Rrs = -(f)2Fj8Fr8Fs) IT. Since the second-order free energy in (12) is required to be negative-definite for all values of the applied fields, it follows first, by setting all the fields except T to be zero, that Rrr 20 or (see the Appendix)

(13)

At zero temperature the relation is Xrr 2 o. Second, if all the fields are set to zero except for the pair F rand F s, the second-order free energy becomes the negative of (F; Rss + F; Rrr + 2F s F r Rrs ), which may be written in the algebraically identical form

For the free energy to be negative-definite, it follows that Rss Rrr 2 R;s for all pairs of T, s. By substituting (8) into this relation we obtain

Because the right-hand side is greater than zero, and by using (13), we finally get the inequality

2 Xrr Xss 2 Xrs . (15)

By considering more than two fields to have nonzero values the theory of quadratic forms leads to relations of the type (Perlis 1952):

Rrr Rrs Rrt

Det Rrs Rss Rst 2 0,

Rrt Rst R tt

where Det indicates the determinant of the matrix, or

and so forth.

(16)

134 A. M. Stewart

The relation (15) may also be obtained by direct algebraic manipulation of (9), but the derivation given above is shorter and more physically instructive. There are several situations as well as the magneto-electric effect (Brown et al. 1968) in which the inequalities may be useful. For example, they apply to the individual components of the paraelectric and paramagnetic susceptibility tensors of crystals. When these tensors are referred to their principal axes, the off-diagonal components are zero (Nye 1957) and so (15) and (16) are satisfied trivially, but for crystals of low symmetry whose principal axes are not known a priori, these relations provide restrictions upon the measured components. Quite generally they apply to any response to any perturbation that may be expressed in the form of (1). Another use is to demonstrate that the paramagnetic susceptibility of a complex magnetic system, when expressed in terms of its component susceptibilities and properly defined inter-lattice exchange interactions, remains positive-definite (Stewart 1993b).

5. Diamagnetism

The arguments that have been used so far in this discussion apply only to perturbations that are linear in the applied fields. They do not apply, for example, to diamagnetic effects, whose Hamiltonian is quadratic in field, and so (13) does not constrain the diamagnetic susceptibility to be positive. The perturbing Hamiltonian for diamagnetism, e2 A 2 /2m with B = 'V X A where A is the electromagnetic vector potential, gives rise via the first term on the right-hand side of (3) to a change in free energy that is positive-definite. The atomic diamagnetic susceptibility tensor, defined to be X~s = -(f)2 F /8Br 8Bs) IT, is obtained with A = B X r /2, appropriate for a uniform magnetic field B, to be

y2+z2

XD =_L( I-xy rs 4m -xz

-xy

z2+x2

-yz

-xz

-yz I )0' :r:-+y2

(17)

The free energy change is therefore also given by -~'Er 'Es Br Bs X~s. Consequently the total free energy change resulting from the application of a magnetic field to such a system is -~'Er 'Es Br Bs (X~s+R,.s), where the first (diamagnetic) term in this expression is associated with the first term on the right-hand side of (3) and the second (paramagnetic) term is associated with the last two terms of that equation.

The atomic diamagnetic susceptibility tensor must satisfy the inequalities (15) and (16) to make the free energy change positive-definite. We demonstrate that the elements of this tensor have the required properties. Consider, for example, Xxx Xyy-X;. From (17) this is proportional to «(z2) (r2) + (x2)(y2)_(xy)2). But (x2) (y~) 2:: (xy)2 by the Schwarz-Cauchy inequality (Hardy et al. 1952), so therefore inequality (15) is satisfied. By expanding the positive quantities (X±y)2) for the pairs of arguments (x,y), (y,z) and (z,x), and by multiplying them together, it can be shown that inequality (16) is satisfied too. However, the elements of the sum of the paramagnetic and atomic diamagnetic susceptibility tensors, which would be measured in an experiment, need not satisfy these inequalities.

Inequalities for Isothermal Linear Response 135

Acknowledgment

Dr M. P. Das and Professor D. J. Mitchell are thanked for helpful discussions.

References Brown, W. F., Hornreich, R. M., and Shtrikman, S. (1968). Phys. Rev. 168, 574. Hardy, G. H., Littlewood, J. E., and Polya, G. (1952). 'Inequalities', 2nd edition (Cambridge

University Press). Kubo, R. J. (1957). J. Phys. Soc. Japan 12, 570. Lifshitz, E. M., and Pitaevskii, L. P. (1980). 'Statistical Physics', 3rd edition, Part 1 (Pergamon:

Oxford). Nye, J. F. (1957). 'Physical Properties of Crystals' (Oxford University Press). Perlis, S. (1952). 'Theory of Matrices', p. 94 (Addison-Wesley: Massachusetts). Stewart, A. M. (1993a). Phys. Rev. B 47, 11242. Stewart, A. M. (1993b). Rare Earths '93 Processing and Utilisation Workshop. Papers, p. 129

(Australian Mineral Foundation: Adelaide).

Appendix

To demonstrate explicitly that Q ~ (V)6/ kT, it suffices to show that

Ej=Ei

~ L e-r;Ei L Iltjil2 ~ (Vo)2. Zo i j

(AI)

Consider the positive quantity:

~ Le-r;Ei[(ilVli) - (VoW· Zo i

(A2)

By multiplying out the square bracket it follows that

~ Le-r;Ei(ilVli)2 ~ (~ Le-r;Ei(iIV1i))2 = (V)5· Zo i Zo i

(A3)

Since (AI) contains the positive squares of the off-diagonal elements, as well as the diagonal elements, (AI) is true a fortiori. An identical argument is used to prove that Xrr ~ (Or)6!kT.

Manuscript received 5 November 1993, accepted 21 January 1994