VARiATIONAL APPROACH TO AN APPROXIMATE HAMILTONIAN FOR STRONGLY CORRELATED NARROW-BAND ELECTRONS - RELATION TO BAND CALCULATIONS AND THEORY OF EXCHANGE IN INSULATORS Thesss for the Degree of Ph. D. MITCHTGAN STATE UNIVERSITY MILTON PENHA SILVA 1975 A. L LII-‘33»! 13 V Michigan 513:?" 3 ( University r“ ww‘vV This is to certify that the thesis entitled VARIATIONAL APPROACH TO AN APPROXIMATE HAMILTONIAN FOR STRONGLY CORRELATED NARROW-BAND ELECTRONS-RELATION TO BAND CALCULATIONS AND THEORY OF EXCHANGE IN INSULATORS presented by NILTON PENHA SILVA has been accepted towards fulfillment of the requirements for PH.D. Physics degree in \I ..~ 5" ,IIMAAOgA WWW: Major professor II Ill Date / /75 0-7639 ) 'fl‘n—‘fl. # , ‘ \ r ? BIREM '1’: " IIOAG &"88II$' HOOK BINDERY INC. BRARY Bmomc I ‘ ABSTRACT VARIATIONAL APPROACH TO AN APPROXIMATE HAMILTONIAN FOR STRONGLY CORRELATED NARROW-BAND ELECTRONS - RELATION TO BAND CALCULATIONS AND THEORY OF EXCHANGE IN INSULATORS BY Nilton Penha Silva This paper contains a critique of one-electron band theory for narrow-band magnetic semiconductors, with the conclusion that the approaches in the literature are unsatisfactory from a conceptual point of view, the difficulty being connected with the question of what periodic potential, if any, is to be considered. We then present a new approach which is variational and consists of taking a trial Hamiltonian H, much simpler than the exact Hamiltonian, nevertheless containing two-electron terms as well as one-electron terms. Unlike one—electron theory, physical predictions are to be made only when the one-electron or band terms in H are combined with important two-electron terms. Unlike previous theories (due to Hubbard, Anderson) which follow the Nilton Penha Silva latter procedure recognizing the essential importance of certain two—particle terms, we choose the one-electron terms (the band-structure) variationally and simultaneously with a similar choice of the two-electron terms. This novel feature of our approach overcomes the difficulties in the previous theories, as shown by our initial investi- gations made within the narrow half-filled single-band problem. Two definitions of our trial Hamiltonian H are considered; in one, H is of the form of the Hubbard Hamiltonian, and in other, potential exchange terms are added. VARIATIONAL APPROACH TO AN APPROXIMATE HAMILTONIAN FOR STRONGLY CORRELATED NARROW-BAND ELECTRONS - RELATION TO BAND CALCULATIONS AND THEORY OF EXCHANGE IN INSULATORS BY Nilton Penha Silva A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Physics 1975 DEDICATION In memory of my father. ii ACKNOWLEDGMENTS I am deeply grateful to Professor Thomas A. Kaplan for having me as his student and giving me the opportunity of collaborating with him in physics research. His guidance has been superb. Truly, without his enthusiasm, time, effort, encouragement and friendship, the accomplishment of this thesis would not have been possible. I am very thankful to Professor Truman O. Woodruff for accepting me as a student at this Department of Physics, for being my academic advisor and for continuous encouragement. I thank very much The Rockefeller Foundation for pro- viding me with a scholarship through all this doctoral program. I am thankful to Professor Julius Kovacs, Professor Wayne Repko and Professor Gerald Pollack for their encourage- ment and continuous interest in my progress. I am indebted to Professor Robert D. Spence and Professor S. D. Mahanti for valuable comments on this work. Last but not least, I thank my wife for her invaluable assistance, encouragement, patience and support. iii LIST OF LIST OF Chapter I. II. III. IV. V. VI. TABLE OF CONTENTS Page TABLES Vi FIGURES Vii INTRODUCTION 1 HUBBARD'S DERIVATION OF THE HUBBARD HAMILTONIAN; ITS FAILURE 9 ANDERSON'S LIGAND FIELD THEORY; ITS UNACCEPTABILITY 16 VARIATIONAL APPROACH 26 EXPLORATION OF THE VARIATIONAL APPROACH WITHIN THE SINGLE-BAND PROBLEM 31 A. Zero-Bandwidth Hamiltonian; All Temperatures 33 B. Narrow Half-Filled Band: The "Best" Hubbard Hamiltonian 42 B.1. High-Temperature Limit 44 B.2. 'Low-Temperature Limit 48 3.3. Two Sites; All Temperatures 59 C. Narrow Half-Filled Band: The "Best" Modified Hubbard Hamiltonian (With Potential Exchange) 64 C.l. Two Sites; Low-Temperature Limit 64 C.2. Effective One-Electron Potential 75 SUMMARY AND DISCUSSION 80 iv APPENDICES A1. A2. A3. A4. A5. A6. A7. A8. REFERENCES Two-Site Single-Band Models Natural Conditions for a Unique Choice of Wannier Functions Solutions of Equations (60) and (61) Proof of Equality in Equation (92) Solution of Equation (101) Proof that fieff in (114) Corresponds to a Heisenberg Hamiltonian Calculation of Equation (121) Calculation of Terms in Equation (127) Page 84 92 94 99 101 105 107 110 112 Table 1. Table 2. Table 3. LIST OF TABLES Page Listed are the eigenstates of the exact 89 Hamiltonian H, (138), and E - uN for half-filled band condition; a is a function of temperature and other parameters in H (See Appendix A1.). Here, _ 2 *2 ((V V 'I’ 16t12) ] I 2 12) i 12) - E-)/2t A = (1 + c2)”2 12' 1 I — _ _ .— n = 3h + v + 2v12 - x12 , |a1> = (c 11 + + + + + + °2+°1+II°> ' Ia2> ‘ (°1+C1+ + °2+°2+)I°> ' E1 = (1/2) [(V+v C = (V12 t = h Mud ~ X12 ' + + 1+°2+ 12 12 + ~ + .—= Hx . Here E 91 2 2 S 2) i ((U - U12) + 16b12) ] , 2 _~l- ~l = v E )/2b12 and A (1 + C The singlet eigenstates of H (1/2)[(U + U 5—: ~ -1’ c = 2 C (012 ) . Eigenstates of H1 and H1 - uNl . The last 98 column is for u = b11 + (half-filled band). «HG vi Figure 1. Figure 2. LIST OF FIGURES Page The values of b U and b for the "best" 63 12' ll half-filled Hubbard Hamiltonian are shown as functions of kT, in units of V. Also is shown the ratio blZ/U' The parameters in the exact Hamiltonian were, in units of V: v = 0.2, h = ' = 12 0.05, V 0.03 (so 12 12 0.08), h11 = X12 = 0. t12 The eigenvalues of i) H-uN, ii) HH - uN 74 and iii) HHX - uN in the zero and small overlap limits for the half-filled-band grand canonical ensemble. In case i) it is I assumed that t12 < 0, v12 < 0 and |t12| Ivizl and kT << ||t12| - Ivizll . The levels in ii) and iii) were drawn roughly as in the best HH and HHX . vii I . INTRODUCTION This paper contains a critique of one—electron band theory for narrow-band magnetic semiconductors, e.g. NiO, EuS, commonly spoken of as Mott insulators. The 3' conclusion of this discussion is that the approachesl-6 in KMnF the literature are unsatisfactory from a conceptual point of View. The difficulty we are discussing is not connected with the solution of a given one-electron periodic potential ("self-consistent" or not), but it is rather concerned with the question of what periodic poten- tial, if any, should be considered. We go on to present an approach which is shown to have promise of being satis- factory (it overcomes all the objections raised in our critique). We neglect phonon effects, focussing on the Schr6dinger equation which describes the motion of the electrons in the Coulomb field of nuclei fixed in space. Nevertheless, this equation is not that of a single electron moving in a fixed potential, because of the Coulomb interaction between electrons — an electron "sees" the fixed nuclei and all the other electrons at their instantaneous positions: the Hamiltonian assumed is P 2 I 2 +V(ri)] +§§ S- (1) N p. H=Zlm 1 where V(r) is the potential due the nuclei. Hence the con- cept of a one-electron theory - a model where the electrons interact with some effective fixed potential Veff’ but not with each other - by no means obviously flows from the fundamental physics defined by (1). Nonetheless, this con- cept forms the basis of much of solid-state, molecular and atomic physics. A large portion of solid state work takes the nuclear positions to form the crystal structure. The resulting translational symmetry simplifies appreciably the one-electron problem (given the periodic potential), but does not bring the original problem (1) perceptibly closer to explicit solution. Historically, some theoretical foundation was given to the one-electron theory in the variationally best one-electron approximation, the Hartree-Fock Approximation (HFA)7’8’9. However it was recognized early10 that the common restricted version of this (plane wave solutions) was poor for the low—temperature properties of the electron gas, and presumably therefore for broad-band metals. The HFA is, again for non-zero temperature, also extremely poor and misleading for models of very narrow partially-filled-band 11-15 such as the Mott insulators in which we are insulators interested, even in its general (unrestricted) form. The use of one-electron potentials other than the HF EXJtential for the calculation of energy bands has been Wiriespread. In the remaining paragraphs of this introductory secztion we will discuss methods which were not specifically designed for narrow-band situations and which either have been applied to narrow-band materials in which we are interested or appear to have possibilities in this direction. In sections II and III, methods that were addressed explicitly to narrow band electrons will be considered in some detail. Finally in section IV we discuss our proposal as a contribu- tion to the search for a satisfactory definition of one-electron states in Mott insulators; and in section V our approach is investigated for the half-filled-single-band problem. Probably the most widely used potential for band calcu- 1/3 exchange"1 or its well-known gener- alization known as the "Xa" method16. Slater's exchange was lations is Slater's "p originally developed1 as a simplifying approximation to the Hartree-Fock method. Namely the exchange terms were first averaged and the average exchange was further simplified by replacing the electron density p in the free-electron expres- sion for the average exchange by p(r), the electron density in the crystal. This leads to an exchange term -Aop(r)1/3 where A0 is a specific positive numerical constant. The Xa method consists of replacing AG by a positive parameter A which is unspecified theoretically. The one-electron Hamiltonian G in this approach is the sum of three terms: the kinetic energy, the electrostatic potential energy due to an electron density p(r) plus the nuclei, and -Ap(r)1/3. The final one-electron picture consists of one-electron states w. 1.81 which satisfy Gwi=eiwi and the "self-consistency requirement" p(r) = LIWiIIIIZr the sum taken over occupied i states. Later it was argued that Slater's approximation1 to the HFA would actually be better than the latter in certain respects, and in particular for energy band calculationsl7. Clearly the derivation of Slater's approxi- mation as a method of improving on the HFA must be viewed as an arbitrary procedure, and the Xa method has to be seen as phenomenology. Furthermore, the consequent uncertainty in the meaning of these procedures does not decrease as one proceeds from nearly-free-electron situations to cases of narrow bands18 (because of the dependence on the free-electron case of the original argumentl). In this connection we point out that 1/3 schemes are unsatisfactory on intuitive grounds these p in connection with an N-electron atomic (or ionic) calcula- tion. For large distance r from the nucleus, p(r) approaches zero exponentially; therefore the total potential in the X0 scheme does not approach the physically expected potential due to the (N-l)-e1ectron ion (e.g. if the N—electron atom is neutral, the Xa scheme gives a total potential energy that is exponentially small at large r rather than the expected -e2/r). We note that disagreement with the same physical expectation was the basis of Slater's criticism of the excited or unoccupied states in HF theory (see footnote 15), and that he claimed17 a major improve- 1/3 ment for his p method in this connection. Clearly, a failure to properly predict the large-r behavior in atoms will introduce serious uncertainties into the prediction of band-structure effects in narrow bands. A more fundamental approach based on the real physical problem of interest defined by (1) was taken by Kohn and coworkers2 who developed a new variational principle and, making certain approximations appropriate to either slowly-varying p(r) or to p(r) with a large constant part, derived a one-electron potential similar to Slater's but with A = g-Ao. However, the authors2 noted that their theory could not make justifiable predictions "at the surface of atoms"; hence one would not expect band-structure effects in narrow bands to be discussed satisfactorily within this approach in accordance with the discussion of the previous paragraph. Another approach to one-electron states in crystals which is not basically of a phenomenological nature is the field-theoretical approach described by Hedin and Lundqvist3. Here one-particle states (wave-functions fs(x) and corresponding energies 88; x stands for both space and spin coordinates) are precisely defined in principle in terms of the exact many-electron energy eigenstates, and it is shown that the one-electron Green's function (and therefore many physically observable quantities) can be determined in principle from these states. Unfortunately. this definition does not appear to be appropriate in its present form to the narrow-band regime, at least for obtaining a complete orthonormal set of one-electron func- tions, since these single-particle wave functions fs(x) 3,19 are not linearly independent for interacting electrons; for our case of interest, where the interactions are very strong, criteria for the choice of an appropriate subset of the fs(x) (which further have to be normalized) do not existzo. The final band-theoretical method that was not origin- ally presented in the context of narrow-band situations is the Wigner-Seitz potentia14. This turns out to be related to certain results of our variational calculation, and will be discussed in section V. Hubbard's derivation of the well-known Hubbard Hamiltonian HH, which is presumed to be appropriate for narrow-band systems is discussed in section II. It is construed here as involving a one-electron band calcula- tion in the sense that it involves a calculation of the one-electron operator appearing in H Anderson's "ligand H' field theory", crucial to his theory of exchange in insula- tors, is a theory of one-electron states for the "non-magnetic lattice" which is supposed to have lattice translational symmetry and consequently these states are Bloch functions. We therefore consider this as a one-electron band theory. Both Hubbard's and Anderson's band theories are made in the Hartree-Fock approximation, Hubbard's in the restricted (non-magnetic), Anderson's in an unrestricted (magnetic) HF theory. 11-15 of the HFA as In view of the known failures discussed above, we certainly must answer the obvious question, why even consider these HF band calculations of Hubbard and Anderson? The answer is that the failures result from actually applying the HFA in the way it was derived, as a one-electron approximation, to deduce physical quantities (specific heat, susceptibility, etc.); whereas neither Hubbard nor Anderson use their HF band-calculation in this way. Rather, they use their one-electron operator in combination with important two-particle terms (intra-atomic repulsion in Hubbard's case, the same plus interatomic exchange in Anderson's case) which, for narrow bands, actually dominate the band-theoretic term. Hence our results of sections II and III, namely that these HF approaches are unsatisfactory within the context of the use made of them by their authors, do not follow obviously (as far as we can see) from previously known failures of the HF approximation. It is convenient to discuss here the interesting work of MattheissZI. At first sight he might appear to be following the philosophy of Slater's group since he21 1/3 used Slater's p exchange (gnéspin-polarized) to calcu- late the band structure of KNiF3. However, he didn't follow that philosophy22 according to which one should compare the predictions of their one-electron theory with experiment and, if this results in failure, more refined one-electron approximations should be attempted. Instead Mattheiss correctly recognized the essential importance of following Anderson and Hubbard in introducing the large (intra-atomic) interaction terms along with his one-electron term and considering the latter as a small perturbation, when dealing with these very narrow band situation323. l/3 However, Mattheiss' use of Slater's p approximation for his one-electron terms is still arbitrary and unjustified, and part of the motivation for our approach (section IV) is to reduce, by theoretical means, the arbitrariness of this potential. The gist of our variational approach presented in section IV is as follows. With Anderson and Hubbard, we recognize that in our narrow-band situations certain two-particle terms (including intra-atomic Coulomb inter- actions) are much larger than the interatomic single-particle terms essential to band structure, and cannot be approxi- mated even crudely by effective single-particle terms for most of the common physical observables. Hence we do not consider applying a single-particle approximation to com- pare with experiment. Rather, physical predictions are to be made from a simplified "effective" Hamiltonian H which includes certain types of two-electron terms. The one-electron terms in H are called the band structure for the interacting system, and are chosen variationally simul- taneously with a similar choice for the two-electron terms. It is in this simultaneous variational treatment of the one- and two-electron terms that we differ from Hubbard and Anderson, and it is this novel feature of our theory that overcomes the difficulties in their theories pointed out below, as shown in section V. Aspects of the present work have been reported earlier24. II. HUBBARD'S DERIVATION OF THE HUBBARD HAMILTONIAN; ITS FAILURE. The Hubbard Hamiltonian was derived5 in connection with the problem of narrow-band electrons. Its properties have been studied by a large number of authorszs. It represents the simplest model of a crystal which contains the tendencies toward itinerant behavior and toward localization of the electrons. This Hamiltonian can be written as o + U E Nit NiI (2) + H = Z X b.. c. c H ij 0 13 10 i . . + . where {1} refers to Sites; cio and ci0 respectively creates and destroys an electron with spin 0 in the Wannier func- . O O + O O 1 . O = O 0 tion wl(£) at s te 1, N10 c10 c10 the corresponding number operator; bij is a hopping or transfer integral; and finally U is the intra-site Coulomb interaction (U>0). The one-electron term causes the tendency toward itinerancy, while the two-electron (interaction) term causes localization. One can study the properties of (2) in two different ways. In the first, one tries to calculate bij and U, according to the actual derivation of the Hamiltonian. The ”band theory" part of such a derivation is the calcula- tion of bij (the one-electron part of Hubbard Hamiltonian). 9 10 In the second, one regards bij and U as parameters and proceeds with the study of the properties of (2). It is in this latter, phenomenological, sense that most researchers have used the Hubbard Hamiltonian. In this sense (2) is not restricted to narrow-band electrons; it can be an approximation for any bandwidth, i.e., any bij/U’ including bands that are wide compared to U. Our interest here is in the first view, namely we are concerned with the principles of the calculation of bij and U. Let us outline briefly Hubbard's derivation of his Hamiltonian presented in reference 5. For clarity we will discuss this in the context of a single-s-band model. The Hamiltonian H in this model i526 + 1 + H = X X h.. c. c. + — X E v.. c. c. ,c ,c (3) ij 0 13 10 30 2 ijkl oo' ijki 10 30 lo ko where hij is the matrix element of h - kinetic energy plus the Coulomb interaction of the electron with the ions - between wi(£) and wj(£); and 2 _ 3 3 * * Vijk£ — j d r1 d r2 wi(£1)wj(£2) r12 wk(£1)w2(£2) . (4) The Wannier functions wi(£) and the Bloch functions are related by b. . ”1(5) = N Z (exp-1§°§i)w&(g) (5) 11 where N8 is the number of sites, Bi is the position of site i, and X goes over a Brillouin zone. Then the Hamiltonian (3) k can also be written as l + + H = X h a a — Z v a a .a ,a (6) £.§° £0 2 §1§2E3£4 E10 R20 E40 E30 where + _ -g . _ + a£0 - Ns E (exp+1k Bi) ci0 (7) is the Bloch function creation operator, and h = N'1 X h.. exp-ik-R . (8) k s .. 13 — —ij ' _ -2 Vk k k k ‘ NS 2 Vijke eXp i(kl R +R2 Rj —3 Rkfi 'IR) —l-—2—-3-—4 and = R. - . —ij -1 —3 If one makes the standard non-magnetic HFA to H, one gets for the HF eigenvalues sip HF _ SE — hE + XE , (10) with Ar :12. Vk' ”"3: .1: 12' 1:""12' r 1: Is." ' ”1’ 12 where vk is the average occupation number for state k. Now, with (10), eq. (6) can be written a (12) _ HF_ + 1 H — 2 (RE A 2 ango + — o'ak o'ak 0 ° a o k _4 _3 X v a k k k k 31 _2 ) R —1—2—3—4 As Hubbard5 noted, the presence of A in (12) avoids counting k the interactions of the electrons of the band twice, once explicitly in the Hamiltonian and also implicitly through EHF k C At this point it is convenient to write (12) in terms of Wannier-function creation and destruction operators, obtaining eq. (6) of reference 5: H = R bij Cio Cjo + % R vijkll. Cio Cjo' Clo' cko - Z (Zvijkg ' Vijzk) ij Cio Cko (13) where bij = Ngl E 8;? exp ikfgij (14) and v. = N“1 2 v exp ik-R. (15) 39. s .15 _l_(_ —-—32. (so vjj = 1/2 for the half-filled band). Having in mind that one is dealing with narrow-band electrons, we can imagine a picture where the Wannier functions are very much like the atomic s-functions and these form an atomic shell with a radius small compared 13 to the interatomic separation. Then Hubbard argues that the integral U : viiii should be much greater than Vijki' if i,j,k and 2 are not all equal. In other words, the intrasite interaction U is much greater than all the inter- site interactions. Neglecting then all intersite v's in the last two sums, the Hamiltonian (13) becomes the Hubbard Hamiltonian (2) plus constant terms, which are dropped. The neglect of the difference between the Coulomb inter- actions and their HFA (the second and third sum in (13), respectively) is, for macroscopic systems, better than neg- lecting the Coulomb interactions only, as we shall see. However we shall now show that this approach has an unaccep- table flaw. The explicit expression for Hubbard's transfer integral is = h.. + Z (2 .) v bij 13 k2 Vika ‘ Vikzj kg (16’ obtained from Fourier transforming (10). Consider first the very simple case of two sites and two electrons. In the usual situation where the bonding state («(w1(r) + w2(r))) lies lowest, and the wi are real, the expression (16) gives _ 3 b ‘ R + 2 v1221 12 12 v1121 + l ‘ 5 (l7) V1212 ' All the quantities on the right side of (17) but the last one v (inter-site Coulomb interaction) behave27 like 1212 14 exp -aR12 for large intersite separation R12 (a is a positive constant). But the quantity v1212 goes like l/R12 for large intersite separation. Hence the low-lying singlet-triplet splitting 4bi2/U , calculated from the Hubbard Hamiltonian, behaves as l/Ri2 for large R12 , in drastic disagreement with the exponentially small behaviour of the singlet-triplet splitting obtained directly from the exact Hamiltonian (3). Explicit calculation of these splittings can be found in Appendix Al. For an abritrary number of sites NS , (16) can be written bij ’ hij + E vikjk ' Vijij vji + 2££ Vika sz Ezgjivikkj sz ' (16‘a) For macroscopic systems NS+ m ; although X vikjk would k diverge (vikjk = R; as Rk+ 00) , these large-Rk terms are cancelled or screened by the interaction in hij of an electron with the nuclei, rendering the expression (16-a) convergent. From (lG-a) the leading behavior for large R.. = Igisgjl is seen to be bi' . At zero T, —1 a : ’Vijijvji at least for a variety of examples (e.g. a linear chain), vji approaches zero as a power of Rij and oscillates sinu- soidally with a wavelength 3 Fermi wavelength. So Hubbard's transfer integral oscillates and vanishes as a power of the intersite separation. This is again in serious disagreement with the expectation that it should be exponentially small: 15 6,28,29 perturbation theory leads to a Heisenberg Hamiltonian = - Z. J.. s.-s. (18) H . Heis ij 1] -1 —j describing the low—lying states of H, where J.. = 2 1111 Vijij) (19) and (Li) t.. = h 1] ij + v11j1 + 1 Virjz ' (20) The sum in (20) excludes the case £=i and j. Since 29 it can be shown in the narrow-band case that wi(£) is 30 31 exponentially small for I; - Ril+ w , (20) gives ti' 3 and therefore Jij being exponentially small for large RijBZ. An even worse result would have been obtained if Hubbard had merely neglected the interatomic part of the Coulomb interaction (rather than the difference between it and its HF approximation). Then the transfer integral bi' 3 would have been simply hij , which diverges for an infinite crystal, fixed i and j , in accordance with the above discussion. III. ANDERSON'S LIGAND FIELD THEORY; ITS UNACCEPTABILITY In this section we examine the concept of the ligand field problem as discussed by Anderson and crucial to his theory of exchange in insulatorsG. This problem was defined6 as that of determining "the exact one-electron states" for the crystal as a whole, excluding the effects of exchange interactions between the magnetic ions. These states are Bloch functions |k> and their energies 6k ; by their definition6 they are to be "non-magnetic", i.;. each wave function is to be a product of a spatial and a spin function, with the spatial function and energy being inde- pendent of spin. From these states one is to construct "the exact Wannier functions" as in Eq. (5). Finally, one is to carry out perturbation theory for the full crystal many-electron problem considering interionic overlap to be small. This procedure leads, according to Anderson, to the low-lying energies of the system being predominantly the eigenvalues of the Heisenberg Hamiltonian (18) where the dominant contribution to the exchange "integral" Jij is _ _ 2 J.. — 2|bij| /U + v. (21) 1) ljji 16 17 Here bij for i # j is the transfer integral or hopping inte- gral as calculated from the ligand field Hamiltonian HSC: bij a (wilnscle) , (22) U is the change in energy needed, to zero order in overlap, to transfer an electron from wi to wj . The spins of wi and wj must be specified to define (22); Anderson did not specify these - we will discuss the various possibilities below. (We have simplified the notation to be appropriate for a single magnetic band, which is sufficient for our present purposes.) The first and second terms of (21) were 6 called "kinetic exchange" and "potential exchange", respectively. The states |k>, 8k were defined6 (in the "most formally exact" method) as eigenfunctions and energies of the Hartree-Fock operator H in which the spins of all HF the magnetic ions are assumed parallel, i.e. in this most exact method HSC = HHF and HHFIk) = EEIR) . (23) Anderson motivated this HF definition by saying that certain desirable properties are expected to follow (for example the non-magnetic electrons in the system, e.g. associated with F- ions, can then be treated as core electrons whose wavefunctions will not change very much with magnetic excitations; again, the perturbation theory is rapidly convergent when use is made of "the exact l8 localized functions"). However these desirable properties were not ghgwné to occur. In fact, it has never been clear to us why a HF definition is at all appropriate: Since the wi , which define the perturbation expansion, are to lead to excited-state properties (the excitations within the Heisenberg model, with energies per particle ranging over .), and since a complete set of one-electron 3 states is necessary for perturbation theory (not just an interval 3 J1 occupied states) the zero-T HF theory would not seem to have variational significance33. The non-zero-T HFA (even in the limit T+ 0), while variational, gives physically absurd predictions for very narrow band systemslz'14’34. Neverthe- less those absurdities were apparently not so obvious as to prevent many workers from being misled. 29,35 Recently, it was shown that something is definitely wrong with Anderson's definition. We present here 29 the simpler of these arguments. Consider a two-site model having two electrons and two spatial orbitals W1 and w2, 27 like H in Slater's model where w1 and w2 are real ortho- 2 normal linear combinations of the ls-states centered at the two sites. The zero-T limit of the HF operator with occupied states wl...w£ is defined in general by its matrix elements in an orthonormal basis set ¢1,¢2,... by 2' A I¢1IHHF‘¢1"°¢1)I¢j’ (¢i|hl¢j) +k£l (¢iwklvl¢jwk) (24) where (¢wl3l¢'w') s (¢wlv|¢'w') - (¢wlvlw'¢') <25) l9 and (¢w|vl¢'w') Id3rld3r2 ¢*(rl)w*(rz)v(.r.12)¢'(51w(£2) (26) The case of Anderson's definition (22) gives, from (24) o _ b12(+,+) : (inIHHF(wl+,w2+)Iwzo) (27) h o = I = 12 (28) h12 + 2 v1121 o + We used the symmetry and reality of W1 and w2 for the hydrogen molecule which implies v also 1121 = V2212‘ h - * 3 12 — f wl(£) h w2(£) d r , (29) where _ 2 _ 2 _ _ 2 _ _ 2 h - p /2m e /I£ fill e /l£ R and v(r12)—e /r12 . 2" The first difficulty presented by the result (28) is its strong dependence on the spin of the states between which the matrix element is taken: h12 and v are of the 1121 same order (first) in the overlap between the atomic functions. Furthermore, the exact exchange integral J12 as shown in Appendix A1, is approximately J (30) —. _ 2 .— 12 ' v1221 ZtlZ/(Vllll V1212) 20 for small overlap, where t = h (31) 12 12 + v1121 ' Expression (30) is like (21), but with b12 replaced by (31); thus it is different from either of the values given by Anderson's scheme36. Clearly then his theory is not defined precisely enough for his purposes, and neither of the choices (28) is correct within Anderson's picture, the error made (which is first order in the overlap) being of the same order as terms retained. Although unsatisfactory, we note that Anderson's result is far superior to Hubbard's. Fuchikami35 came to a similar conclusion about the unsatisfactory nature of Anderson's definition37. Consider now the HF operator obtained by occupying antiparallel spin states. The corresponding transfer integral is 0' ._ b12II'I) ‘ (wlOIHHF(w1+’W2+)Iw20) (32) independent of the spin 0 . Hence the hopping integral calculated using the HF operator with occupied Wannier 21 functions having antiparallel spins leads to correct exchange splitting via (21). Fuchikami3S apparently noticed the analogous point for her similar, but slightly more complicated two-site model. However, contrary to Fuchikami35, one cannot con- clude from these special models that the HF operator for the antiferromagnetic state is correct in general (one can conclude only that using parallel spins is not always correct). In fact, in certain cases (e.g. MnO) it is impossible to have all the nearest-neighbor pairs be antiparallel. Thus it would be impossible in those cases to satisfy Anderson's requirement6 that the exact one-electron states be solutions of a Schr6dinger equation for the crystal as a whole simultaneously with this antiparallelism of all near-neighbor pairs. Furthermore, in any such HF scheme, the states do not satisfy, even approximately, Anderson's requirement of spin-independence. We can see this directly in terms of our simple two-site model. Consider the parallel-spin case. The HF equations are (wv|HHF(wl,wz)|wu) = av av“ (33) so if the eigenfunctions are written w = Z A? w. , (34) 22 the coefficients Aio are to satisfy Z'(w10 IH HHF(W1'w2 )Iw. JO,) Ajo . = e A. . (35) jo It is easily verified from (24) that the matrix elements 7 ° = I (wioIHHF(wi+’w2+)ijo ) are zero 1f 0 o and that h o = + | 11 + V1212 ' v1221 ' HF W1+'w2+) w1= (36) R11 + v1111 + v1212 ' 0 “ I' io|H thus with (28) we have all the matrix elements of HHFIW1+’w2+) in this model. The eigenstates of HHF(w1+,w2+) (1n our limited model space) are _ -% ik W — 2 (W1 + e w2) 0 k0 I k = OI”T I (37) 0' (as easily seen from the symmetry), with energies, from (28) and (31): ik 11 + v1212 ' v1221 + e h12 ° = I E:kcj = ik (38) R11 + v1111 + v1212 + 9 (R12 + 2 V1121) ° = h Finally we see that (37), (38) provide a solution to the 23 HF equations (33) with occupied states W = W , W = W (39) because from (24) and (37) (¢i|HHF(wl+’w2+)|¢j) = (¢iIHHF(w0+'Wn+)I¢j) ' (4O) (HHF(wl...w2) is invariant under a unitary transformation of the occupied orbitals, as is well-known). Thus ik + 2 e v1121 - (41) €k+ ' €k+ = V1111 + v1221 Since (41) involves the intra-atomic Coulomb integral Vllll' the difference between up and down energies is very large, even for small overlap. Clearly the intra-atomic nature of this term implies that it will occur independently of the relative orientation of the occupied spins. Pic- torially, the viiii occurs in ck+ because wi+ is unoccupied; as is well-known (ref. 8, secs. 5-2, 6-2) the HF equation for an unoccupied state corresponds intuitively to the Schrodinger equation for an ddddd electron. One might note that the wave-functions do not show this magnetic property in this model: the up and down wave-functions at site 1 are wlo = w1(£)aO (the same spatial function for both spins). However, it is clear from the above 24 discussion that this non-magnetic property of wave-functions results only from the restricted nature of the basis set used to define our model; if for example one enlarged the basis set to include 25 states wi,wé (giving 4 spatial orbitals), then the HF wave-functions wig will be of the form wig = w:o(£)ao where v = 1,2 (corresponding to ls and 23), k = O,n again, but w:0(£) will be expected to depend on 0 even for zero overlap38. Finally we should address ourselves to the apparent conflict between the conclusions just reached about the unsatisfactory nature of Anderson's theory, and statements recently made supporting it35’39'40. Of these works we are principally concerned with those which consider the crystal as a whole. In fact there is no conflict with the 35 O I , because a) her Wannier functions were work of Fuchikami determined using experimental output rather than by Anderson's HF method and b)she considered only 3d cation bands (without higher ones). Hence she did not deal with the aspect of Anderson's method that we have criticized. fkfl? avoidance of this criticism is of course not without cost: shereplaced a fundamental theoretical question, that of determining theoretically and therefore predicting the best Wannier functions,by a phenomenoloqical procedure. We also note that a glaring defect in Fuchikami's calcula- tion is the use of a value of U appreciably and apparently arbitrarily reduced (in quantitative terms) from the 25 carefully calculated value UO . (The value of the exchange integral is very sensitive to this change since the kinetic exchange and potential exchange terms very nearly cancel if the U0 value is used.) The physical effect invoked as the cause of this reduction is real 35 (it's called "electron rearrangement" by Fuchikami and was referred to as a "correlation effect" by Hubbard et al.41). Such an effect can be included in a perturba- tion theory provided higher bands (e.g. 4s cation bands) are included in the basis set - in such a case the magnetic properties of the Wannier functions would show up in Fuchikami's HF method, as discussed earlier in this section. 40 Gondaira and Tanabe seem to have followed Anderson's HF method; unfortunately they did not define their HF operator (in that the occupied spins were not defined) and therefore their consideration cannot be considered as a valid investigation of Anderson's approach. Finally the treatment of Anderson's approach by Huang Liu and Orbach39 .35 followed that of Fuchikami and so is subject to the deficiencies of that method. IV . VARI AT IONAL APP ROACH Our initial approach consists of finding the varia- tionally best Hamiltonian of the form ~ ~ ~ ~ ~+ H “ X bvi,vj Cvio Cvjo + Z Uv Nvi+ Nvi+ (42’ ~ where v is a band index, i,j label atomic sites, Cvio creates an electron in the Wannier function wvi , with ~ ~+ spin a , N . = c 5 and b . . U ar varia- V10 v10 V10 ’ v e the vi,vj tional parameters. Further, wvi are taken as orthonormal linear combinations of a basis set of one-electron func- tions (which can be chosen arbitrarily), the coefficients being also variational parameters. In other words, we take the trial Hamiltonian essentially of the form finally assumed by Mattheiss21 as a sum of a band structure Hamiltonian, namely the one-electron operator H(1) in (42), plus a Hubbard-type of interaction term (involving the Uv)° But instead of choosing the band structure Hamiltonian arbitrarily, and in a way conceptually associated with nearly-free electrons, we choose it variationally, treating it on the same footing as the U-terms. To accomplish this we use the so-called variational principle of statistical mechanicsg, namely, that the trial 26 27 free energy at temperature T and chemical potential u satis- fies the inequality “1 > FIH] (43) tr 5(H-H) - B F[H] in tr exp-8(H - uN) with ‘ exp-8(H - UN)/tr exp-8(H - UN) (44) '0 II for all Hermitean H. Here H is the Hamiltonian taken as exact and which is being approximated by H. By this principle, the best estimate of the free energy is the minimum of F[H], and this is the criterion for determining the best parameters in H. Clearly the resulting one-electron Hamiltonian ~(1), namely the one-electron part of the "best" (42) is by definition variational, it is a property of the crystal as a whole and is non-magnetic. Further we will see that the "best" (42) gives the correct behavior for single narrow- band models provided the exact kinetic exchange, _ 2 . . K° ' : "2tij/(V- " ) I (1%)) (45) 13 1111 Vijij dominates the potential exchange Vijji' Therefore for such ~(1) models the (one-electron) eigenstates of the best H could provide a definition of the ligand field states which represents a major improvement over the HF definition6'35. However the limitation to cases with dominant Kij is unsatis- factory on general grounds (in fact there are cases where 28 39). Further for more than this is expected to be violated one narrow band (e.g. a "degenerate" band such as a 3d-band in NiO), the trial Hamiltonian (42) will be unsatisfactory on the additional ground that Hund's rule, essential to magnetism in insulators, would be violated43. These difficulties should be overcome by adding potential exchange type terms to (42); also generalizing the one-electron terms (with no additional serious increase in difficulty) we obtain as our trial Hamiltonian ~+ ~ ~ ~ N H = Z bvi,uj Cvio cujo + z Uv Nvi+ vi+ _ _ U . . c , _ . . where U . . are additional variational parameters which VlIUJ physically appear as effective potential exchange matrix elements. It should be realized that the interaction terms in (46) are of a much simpler form than the (exact) Coulomb interactions. To find the best H we must solve the stationarity equations (3/3l)F[H] = 0 which may be written42 ‘0? 3 tr 3 (H - H) = o (47) >’ with fixed temperature T and chemical potential u ; here 1 stands for any one of the variational parameters. The derivative of the density matrix 5 , (44), with respect 29 to A is of the fOrm 3d _ 1 _ ” __ -B(H - uN) 3A ’ z (1 0 tr) 3A 3 where 2 = tr e-B(H - UN) (49) and ~ -B(H + 51 §§-- uN) ~ 3 -B(H - uN)_ 1. 1 3A -B(H - uN) (50) 57 e - 1m. Ei'le -e ], 6A+O . . 44 Wlth the use of the expan81on A+B _ A 1 1 x1 e — e [1+]o dx1 B(xl) + [o dxl [o dx2 B(xl) B(x2)+...] (51) (A and B are operators, and B(x) E e-XABe+XA) , the expression (48) becomes 8 3 '02 = -s (1 - 5 tr) 0 I; dx QA(x) (52) >2 where QA(x) is defined as 91‘“) E exB(H - uN) %%_e-XB(H - nu) . (53) Then it follows that the stationarity equation (47) becomes ~ 3H = A I: dx - <§—> o . (54) 30 We have used the following notation for the thermal average of an operator A: S tr 02 w (55) The fact that H and %% do not commute with each other in general, makes (54) much more complicated to calculate than if they did. We have so far investigated this approach only for the single-band problem, taken for simplicity of discussion to be an s-band. We further have limited ourselves to the narrow-band or small-overlap region. Our investigations are described in subsequent sections. It is worthwhile mentioning that recently it has been shown42 that a large class of approximations in statistical thermodynamics that are based on (43) yield expressions for the macroscopic quantities of the system that are con- sistent from both the statistical mechanical and thermo- dynamical points of view contrary to previous claims. A question in this regard arises because the variationally best parameters that appear in H are generally T and u dependent, unlike an exact microsc0pic Hamiltonian. That result was important in the considerations of the present paper.(1t answers questions like, should the approximate entropy be calculated as -%T F[H(b)] or -ktr B(b) ln B(b) , guaranteeing that they are the same. By H(b) and 5(b) we mean the best H and the respective density matrix according to (43)). V. EXPLORATION OF THE VARIATIONAL APPROACH WITHIN THE SINGLE-BAND PROBLEM The exact Hamiltonian for this problem is given by (3); and according to our approach the trial Hamiltonian should be: ~ ~+ ~ ~ ~ H = 2 b.. c. c. + U 2 u. N. i,jo 13 10 30 i 1+ 1+ ij 00' 13 Although the Bloch functions wk(£) are physically invariant for a single band (aside from a phase factor exp i Yk) , the Wannier functions wi(£) which are to be constructed as g g exp i(—_¢§i+yk) wk(£) (56-a) wi(;_)=NS are not invariant because of Yk . Then, in principle one should ask for the best wi in the sense of (43), leading to an enormous number of variational parameters. However, it seems natural to demand the Wannier functions to have, in common with the atomic ls functions ai , the properties 31 32 of reality and invariance under inversion through 3i ; and also to go continuously to ai as the interatomic separation goes to infinity. Then one proves (Appendix A2) that for the single band case Yk is constant, leaving no physical arbitrariness in the choice of the Wannier functions, the variational parameters being only bij' U and Uij . (We can then drop the tildes on the right side of (56). Naturally the best parameters will be found to be functions of the various integrals in the exact Hamiltonian and of B and u . In particular, the leading terms of the expression for bii and U are expected to be zero order in the overlap; similarly bij and Uij (for i # j) are expected to be mainly of the first and second order in the overlap, respectively. One can see that this is very plausible: Suppose the exact Hamiltonian contains only the terms X h N + V 2 N N + X. h E c+ c - (1/2) Z'v i 11 1 i 1+ 1+ ij 1) o 1o 30 ij 1331 + + _ . . 9 ga?iocio'cjo'cjo , where V _ Viiii . Then the m1n1mum of the trial free energy is attained for bii = hii’ U = V, hij = bij and Uij = Vijji, in which case min F[H] = F[H] . The quantities hii and V are clearly zero order in the overlap, and hij'vijji for 1 # j are first and second order in the overlap45. So that in the case of the full Hamiltonian (3), the parameters bij and Uij (i i j) should be zero in the limit where the overlap goes to zero. In accordance with this expectation, 33 we will assume in the remainder of this paper that bij and Uij+ 0 as the nearest overlap A+ O ; more precisely, we shall assume that bij = 0(1) and Uij 0(A2) . The latter assumption will be shown to be a consistent one in each of the cases studied. A. Zero-Bandwidth Hamiltonian; All Temperatures To get some insight into more complicated cases we studied first the simple case in which we keep only terms 13 in the exact Hamiltonian H to zeroeth order in overlap (3); in other words, we consider the atomic limit or zero-bandwidth exact Hamiltonian: _ l Ho — E [buni + v NHN1+ + ngi vijij NiNj] (57) where N1 = Ni+ + Ni+ , h11 = hii , and V E v. In iiii accordance with the discussion of the previous paragraph, the trial Hamiltonian is taken of the form H = E [biiNi + UNi+N 0 Z Hi ' (58) 1 1+] This is also the atomic limit of the Hubbard Hamiltonian. The consistency of the assumption bij = Uij = O , i i j , will be seen in secs. B and C below. 34 Because (58) is function of number operators only, 8H 572 commutes with HO and (54) becomes: 3H0 ~ . 3H0 <31 (HO-Ho)> — <3A > = 0 (59) In particular for A = bll , making the restriction that bii is independent of i,as is hii by the lattice transla- tional symmetry, (v-U) é () + (h ) Z () 11'b11 J 1 ' _ . + 5 £2 vj£j£( ) — 0 , (50) similarly for A = U (V'U) g ( ' ) + (“11'b11) g ( ' ) 1 ' _ = + E E2 Vj£j£( % the ground state of this V l-' is a critical one as noted by Hamiltonian (containing and v interactions only) 12 consists of alternating empty and double occupied sites. Thus for n > % the variational approximation Héb) is extremely poor since the approximate ground state con- tains singly-occupied sites (U = V>0) ; however, as we have said, the variational approximation is quite good for rather large n(2%) . B. Narrow Half-Filled Band: The "Best" Hubbard Hamiltonian In the narrow band region the overlap between the atomic functions is very small and all the terms in H, (3), which depend on the overlap are small compared to the others. For convenience we separate the exact Hamiltonian H in two parts: H = H + H' (77) o where H0 is the zero overlap part given by (57)and H' is I defined by (2 means a sum with all indices different) 43 H' = Z X [h.. + v....(N.- + N.-) + X v. . N 1c? c. 0 ij 1] 1131 10 30 k#i,j 1k3k k 10 30 ' + + + f g Ej Vijjilcio Cjo + °j6 Cio)cio Cjo ’ Nio Njo] + i X Z. v .. (cT- c - + c+- c -)c+ c (78) 2 o ijk i13k 10 k0 k0 jo io jo 1 1 + + + 5 go. Ejkl Vijkl Cio cjo' Cfio' Cko ' We begin our exploration of the narrow-band case with a trial Hamiltonian which is of the form of the Hubbard Hamiltonian (2). In other words we take the Hamiltonian (56) w1th Uij = 0 , and call 1t HH ~ I + HH “ Z (bllNi + U Ni+Ni+) + Z Z. bij Cio cjo (79) 1 o 13 which for convenience is written as H=HO+T (80) where H0 is (58) and T is ~ I + TEX): b..c. c. . (81) 0 ij 1] 1o 30 As we argued before, the best bij(i#j) are expected to be proportional to the overlap in the narrow band limit, so that when H'+ 0, T+ 0 too. We can then say that T is "small" compared to H0 and expansion techniques can be used to solve the stationarity equations (54). 44 We propose to solve these stationarity equations for A=b11, U and bij(i#j) in two limiting cases: 1) high temperatures, in the sense that kT>>Ibij| for all i and j; 2) low temperatures, in the sense that kT<< nearest neighbor lb. I . In both cases |b..|<< U . 1 13 j B.l High—Temperature Limit It is convenient to define Ho 5 Ho - uN (82) ~ 88 -8(8 +T) and R E e 0 e o -B(HH-UN) -880 ~ such that e = e R . Consequently i = 2 <fi> ~ 1 ~ ~ 0 = i p R (86) 0 o where ~ ~8fio Zo = tr e (87) ~ -BHO ~ 00 = e /Z0 (88) and <...> 5 tr 5 (...) . (89) 45 With all these definitions the stationarity equations are written as ~ 1 ~ ~ o f0 dx0 (90) ~ ~ ., afiH . - O = o , for k ¢ n . (96) ko no 0 With (94), QA(X) defined in (53) becomes ' x + QA(X)=AA(X)+B g X bkz Io d3 [9k15(’5) 9k25(x) cko C20 AA(X) k2 -A() -()+ 1 (97) A x gkzo S cko C20 where x88 ~ -x88 AA(x) s e ° gE-e ° . (98) In accordance with previous discussion we assume the best parameters bij(i # j) to be proportional to the overlap between atomic functions i and j; we then calculate (90) neglecting terms which are higher than first order in the overlap. In so doing we easily find that for A = b11 and A = U the stationarity equations (90) become identical to (60) and (61) respectively; we already have the solutions for those equations: 47 U = v (99) b = h + E v . . . (100) 1 . ll 1 3#1 1313 This last one holds only in the half-filled band regime. For A = bij(i # j) eq. (90) becomes 1 + r [0 dx g O + ijo 1o 30 1 1 x + + + Bag. £2 bkz Io d“ {Io dS< [9k26'('5) 9k25"x’ cko'clo'gij3(x)ciocjo - g -(x) c+ c g - (s) c+ c ](H - H )> - (101) ijo io jo kzo' ko' 10' H o l + + ~ -Io dso + + + ~ — + (qua-(X) cko' Czo' cio cjo>o ‘3 " “3%) ' 0 - we then proceed as in Appendix A5 to solve (101) for bi'; 3 we find b = [1 - :iiii f(BV)l-l t (102) ij V ij where tij is given in (20), and f(BV) is -8V 1 - (1 + BV)e f(BV) = _ _ ; (103) 1 + 8v e 88v _e 8v 48 f(8V)+ 1(0) as 8V+ 00(0). We see that Eq. (102) is con- sistent with our assumption bij = 0(A). Then, particularly if kT<>Ibij| for any pair 1 and j means that the electron can go from site to any other singly-occupied site with equal probability; then the same discussion of section V.A on this matter applies here. The value of bll is also understandable: as can be shown, ” (b) it again makes both Hamiltonians H and HH have the same number of electrons to first order in bij (the best HH) (in the half-filled band case). B.2 Low-Temperature Limit In order to calculate the quantities appearing in the stationarity equations (54), we must first discuss the eigenstates of H When the number of electrons N is H’ equal to the number of sites NS , the ground states of HQ with U>0 have one electron on each site, and no doubly-occupied or empty sites, the degeneracy being 2N. Adding small bij terms to it will remove the degeneracy; 6,29,48 (for second order degenerate perturbation theory the energY) shows the resulting energy levels are the same as those of a Heisenberg Hamiltonian (except for a constant) with the exchange integral aij = -2b§j/U . 49 One can divide the complete set of eigenfunctions of H0 in two sub-sets: G - the ground states; and E — the excited states. Consider the following: o sbll :1 v m (I z fiH|fi> = Enln> (105) where Ino> is in G , and En and In> are the eigenvalues and eigenfunctions of H , resulting from the removal of H the degeneracy of the ground state of HO . These contain the low-lying eigenstates. If one defines idempotent operators P and Q such that P? s G (106) and Q? s E for all wavefunctions W so P + Q = 1, then one can write the eigenfunctions In> as 29 |fi> = E [1 - Q :_1___ Q $1 P B > , (107) n ~ 0 HH-En where Cn is the normalization constant, which can be con- sidered real. One then can make the following expansion29 ” ~ ~ k Q:—1:0 = ozl—zo Z (-)k[Q:—l:o (T - AEn)Q] . (108) HH-En Ho-eo k=0 HO-eo 50 ~ where AEn = in - so . By keeping only the first term (k=0) in the series (108), one obtains the approximate eigenfunctions of H in formal second order degenerate H I perturbation theory:29 In>= c (1-0.,1~ n H -E 00 0%] Plfio> . (109) In the class of the operators Q , one can distinguish (V) ’ with v = 1, 2, 3, ... which project th the operators Q functions onto the v excited states of the unperturbed Hamiltonian HO , the energy of these states being Nsb11 + vU . Since our perturbation term T , (81) , involves a hopping cf cjO from site j to another 1 , it is clear 10 that the operator Q in (109) is in fact a Q(1) operator. For clarity, it is convenient to rewrite (109) with this new information: |fi> = E [1 - 0(1) 1 0‘1) &] Plfio> . (110) n ~ ~ -e Ho 0 The orthonormalization condition = 6 gives nm b2 ~ ~ ~ ~ ~ _ ' 00 00- c2 = [1 + 1— 1 1 = [1 + Z —£1.A13] 1 (111) n U2 0 o i' 2 with 2 b. O . . — X —61 An 5 ij nm \ 51 where 13 _ ~ + + ~ An _ go' . (113) * Here we have considered b.. = b.. = b.. . 1] 31 13 2 b. O O 0 ~ ' It is clear from (112) that In0> and X _%1 A33 13 are, respectively, eigenfunctions and eigenvalues of the ~ operator Heff ~ _ 1 ~ (1)~ Heff — U PTQ TP (114) which, in Appendix A6 is shown to be identical to the Heisenberg Hamiltonian, with Jij = ~2bij/U , except for constant terms. It follows from (105) and (110) that 2 ~ b.. .. E = - 2' —$1 A13 + N b U n s n .. (115) 1] ll ' These are then equal to Heisenberg Hamiltonian eigenvalues, except for a constant. (By constant here we mean inde- pendent of Ifio>) . To simplify the calculation we are about to make, let us consider only nearest neighbor hopping bij=b12 . Then (115) becomes ~ biz En = “‘6‘ An + Ns b11 ' (116) where = ij An 2 Z An , (117) 52 denoting sum over nearest-neighbor pairs. For an arbitrary number of sites one does not know how to calculate the eigenvalues and eigenfunctions of the Heisenberg Hamiltonian, which means, in our language here, that we don't know how to evaluate An and (30> for an arbitrary NS . Nevertheless as we will show, our calcu- lation of the best parameters can be accomplished without knowing explicitly (50> and An . Since we are interested here in temperatures which are low compared to U and lblzl , we perform the traces in the stationarity equations (54), only over the low-lying eigenstates of H We then get H O biz ~ 2 B_U— (An+A£) ~ BHH ~ b12 ~ ~ e { [T (An " A2,) - <2|H|£>l n,£ 1 ~ ~ + Io dx} = 0 . (118) Because of the restriction on the traces, bllN appears only as a constant term (=b11Ns): thus bll cannot be determined variationally within the present approximation; consequently we drop it out of the calculation. (But see section B.3). The last term in (118), containing the integral is found to be 1 Bi (A —A ) ~ ~ ~ _ U n m _ ~ 3H ~ ~ ~ Io dx — z e 1 m 2 3A E12 (A _A , U n m ~ b2 ~ BHH ~ ~ ~ + kT X (E — 12 A ) (“'51- |¢><¢|H|n> , (119) 53 where |fi> and |$> are eigenstates of H which originate H from G and E , respectively. Then we proceed and calcu- late the several matrix elements in (118) and (119). We started with the matrix element of the exact Hamiltonian H between the eigenfunctions of the trial Hamiltonian HH. In accordance with our previous discussion we assume that b12 is 0(A), and since we are doing this perturbation theory which is second order in T , we neglect terms in this matrix element which are of higher order than A2 . Dropping the P operators which appear in (110), because P|no> = Ifio> , and defining, for nearest neighbor i and j only 12 ‘ Vijij , t z t.. and X (120) V 12 ‘ 13 12 ’ Vijji we have, using (110)49 , ~ ~ ~ ~ ~ ... 2 ~ (1 ~ ~ = Cncm [ - fi' + + .17 <5 (i- 0(1):: 0‘“ 1).: >1 (121) U 0 O _ 1 ' - 1 ‘ 6nm [2 £2 szkz 2 x12 + Y An] ' (122’ where, for convenience we defined 2 Y z biz (V-v ) _ 2b12‘12 + 1.x ‘ 52‘ 12 U 2 12 ° (123) Details of the above calculation are found in Appendix A7. Before proceeding with the calculations it is worth rewriting (118), taking in account (119) and (122): 54 8 Big (A +A ) ~ 2 E U n 9. ~ 3 H ~ b12 n,£ e [<"lfix— In> ('0‘ + Y) (An'Az) + ~ hi2 -1 ~ afiH ~ + kT g (Ed) " T An) <$IHIn>] = 0 . (124) Notice that the constant terms of (122) (specially the long range Coulomb interaction terms Z vkkkfi) have been ' 2 b cancelled out. Since now we know that —l3;+ Y) is second U 3H order in A , we just have to evaluate <fiI-a-A—Hlfi> to the lowest order in A). Straightforwardly we find ~ 2 a b . a - _ 1 - ~ (1) (1) ~ ~ _ 12 — 2 Z — ——2 An (125) U 1 U and 35 2b H ~ 2 ~ + (1) ~ ~ 12 = - - Z X = - —A ~(126) abl2 U o o 10 30 0 U n The second set of terms inside the square bracket in (124) has an overall factor kT assumed to be higher order than A . Then we only have to calculate these terms up to order of A2 . With (110) we can rewrite it as 2 b b12 ¢ U '1 I?“ I" ~I l A ) <5 ¢><¢ H fi> n a (127) -1 3H “2 ~ .11. " (1) A ) Cn <$|n(1-%0(1"i)|fio>. 55 A careful examination of (127) shows that within the second order in A , the only eigenfunctions |$> that enter in the calculation are the ones resulting from the removal of the degeneracy of the first excited state of Ho , with the same number of electrons NS , which lies higher by U . If we then express |$> in terms of |$o> , as we did in (110) for In> and Ifio> , we have I$> = 5¢ [1 - (P + 0(2)) :—l:— (P + 0(2))T]Q(l)l$o> H -e o 1 = 0¢ (1 - s i 0(1))|$ > (128) where s s 9 ~ 1~ p + Q(2) ~ 1~ 0(2) . (129) Ho-el Ho-sl The normalization constant C and energy E are ¢ found to be b2 -1 52 = [1 + —lZ-(B + B')] (130) 0 U2 ¢ c) b2 E = u + -13 (B - 23') (131) ¢ U ¢ ¢ where B¢ E 22 Z <$Olc10 cjO P 030' Cio'|$o> (132) 00' 56 B$ E 2) X <~ lo? (2) + 00' Then we proceed as in Appendix A8, and get the final form of the stationarity equations (124) b2 8 02(AH+A2)bi2 biz HQ, 9 {T An(An-A)L) (T + Y) + b b ~ 12 .12 _ _ _ —_221§1 +kTU_2_An [U (V V12) tizl} “0" an and b2 12 B U (A“+A£) b12 biz nize {:fi"An(An'Az’ (-6— + Y) (133) b ~ '1 _12 _ _ _ _ 3F[H] + kT U Ant U (V V12) t12];} ‘ ° ‘ ablz ° From these two equations it follows easily that 2 b 12 _ —E— + y — 0 b .12 _ - _ . . (the Am, defined in (117) and (113), are not needed!). Recalling the definition of Y in (123) we find from (133') that the best parameters b12 and U are: (1 + ~13) (134) 57 U = (V-v ) (1 + —l3) (135) 12 K 12 where 2 K = _ 2t12 12 V-v12 is the exact nearest neighbor kinetic exchange; Kij was defined in Eq. (45). U = w is seen to be a solution of (133'), which on physical grounds can be seen to lead to higher free energy. It can be seen from (110) that In> is a linear com- bination of singly—occupied-site functions Ifio> and functions g C10 cjolfio> which have one doubly-occupied and one empty site and all the other sites are singly- occupied; the ratio of this mixture is easily seen to be b12/U . If a similar perturbation theory is done on the exact Hamiltonian H , the same sort of mixture in the wavefunctions is found, the ratio of the mixture being equal to tlz/(V - v12) . Our results (134) and (135) for the best parameters show that the minimal principle (43) matched the wavefunctions of both the exact and trial Hamiltonians, since b12/U = t12/(V - V12) . (136) 58 Another important feature of our results (134) and (135) is that the low-lying eigenvalues of both Hamiltonians have been matched; in other words the appropriate Heisenberg exchange integrals are the same (despite the lack of Ui' in H): 210i2 2ti2 — = — = K O U (V-vlz) + v1221 12 + X12 (137) Hence we have accomplished Anderson's objective of defining a non-magnetic one-electron Hamiltonian whose transfer integral leads via the kinetic exchange to the correct low-lying exchange splittings when IK12I>>X12. However, the above approach fails to give a physically acceptable approximation to the properties of H when 0 12 > |K12| .5 This is seen as follows. We assumed, in the course of making our calculation, that U>>|b12l>>kT , X leading to (134) and (135); but when X12>IK (135) gives 12| . U<0 , which is inconsistent with our initial assumption. The only alternatives to that assumption are 00) . Solving then the above system of equations we get 2v v 12 12 U=v+—__—|t l—-—_—— K (154) V _ v _ 12 biz ' V-v t12 S V-v K12 (155) 12 12 70 _ _ 12 U12 ’ x12 V-v K12 (156) 12 V b = h + v - s v' - —£3— It | (157) 11 11 12 12 v-v12 12 to second order. These results hold for any ratio Xlz/IK12 Because of the complexity of the equations (142) - (145), we have not been able to estimate analytically the order of magnitude of the error made in determining the best parameters. Presumably it is smaller than second order in the overlap. We made a numerical calculation which agrees with the values found above. Although for each best parameter the leading order would be satisfactory, we have to know them up to second order in overlap, in order to solve for 012 (which is essentially second order in the overlap). See for instance, equation (142) or (153). We note the consistency of our original assumptions,.U, bll’ b12' U12 are respectively zeroeth, zeroeth,~first and second order. With the results (154) - (157), it can be seen that the wavefunctions of the exact and trial Hamiltonians have again been matched (compare C and C' in Tables 1 and 2, respectively, in the limit of small overlap). Also the low-lying splittings have been made identical to second order ~ in overlap; in other words, the Heisenberg exchange JHx ~ appropriate to H is equal to J HX appropriate to H H ~ _ _ 2 _ _ JHx — 2b12/U + 012 — K12 + X12 — JH (158) 71 To check it see (Al.13), (Al.11) and (154) - (157). The chemical potential necessary to fix the number of electrons in the best fiHX equal to two is, according to (Al.10), equal to hfl< . (159) In Appendix A1 we calculate the chemical potential to make the exact Hamiltonian H have two electrons; and for kT << (Itlzl - (Vizll I ItlZI > Ivizl (160) the expression (159) above is the correct one to second order. For the other situation, namely kT << lltlzl - Ivi2I| I Itlzl < (Viz) (161) the exact chemical potential is not the same as in (159), differing from that in terms which are first order in the overlap. The former case where ltlzl > [viz] is the more realistic one at least for the hydrogen molecule. Also for the conditions (160), the separation in E - uN between the ground state of the system (a singlet) and the lowest of the (NS i 1) levels has been matched perfectly (to second order): This separation GHX for ~ (b) . U _ _3_ _ 4132 h HHx is 3 2 U12 |b12| + 52 as seen from t e 72 right hand column of Table l; the corresponding separation + . . . _ v _ _3_ _ for H , under the condition (160) is 6 — 5 2 x12 |t12| 4t§2 . From (154) - (157) it can be seen that to second V-V12 ' order, 6 = 6 - (162) It is interesting to notice that the lowest of the (NS i 1) levels is made up of bonding states, for t12 < 0 , or antibonding states, for t12 > 0 ; since the sign of b12 is the same of t12 , correspondly the same happens with the Hamiltonian Hg?) . Now with (161) the degeneracy of this level is between bonding states for (NS 1 l) and antibonding for (NS 1 1), depending on the sign of viz and , - ~(b) t12 , whereas in the HHX above (there being no degeneracy between bonding and anti- the situation is the same as bonding states). This would appear to be a reason why the results are better for (160) than for (161). But in any case the leading term in the expression for the best U (namely U = V) matches the position of the (NS 1 1) levels with respect to the ground state, to zero order in overlap. Further, the position of the highest levels, namely those with NS and (Ns 1 2)partic1es (all of which are degenerate for HHX-uN) matches the average position of the respective levels in the exact model. See Figure 2 to clarify these points. Figure 2. 73 ~ The eigenvalues of i) H - uN , ii) HH - uN and iii) HHX - uN in the zero and small overlap limits for the half-filled-band grand canonical ensemble. In case i) it is assumed that t 0 , v' < 0 and 12 < 12 It and kT << ||t - Iv' 12| > lviz' 12| 12II ° The levels in ii) and iii) were drawn roughly as in the best HH and fiHX . 74 .fl'l 6"?) 2 4t 2vl2’+2Iv'2l+2X'2+ —-'-2 4M2; _2 2 V"'I2 ‘u" + —-.°= -1- )('2'+I12’l+4lv I "2" ”I2 2 I2 55", v if? 'I2+2xI2+"I2' u lbl "2'"I2°;- "2'+ I2 '2 'I2"'2XI2" l'I2""2"'I'2' -11." 2 ~13. "2"”(2' -V- =v.2+2Iv"2I '22 -U 2 —v- -:-v— v 2g--+2III2I4-2x|2v-‘_—-"'|2 203.1312 .9. .L 2 "' 2 uI2"’."’I2| I ""2""'2"’I2"“’I2l 75 C.2 Effective One-Electron Potential It is interesting to try to find an effective one-electron potential W(r) such that = b.. (163) 1 J 13 where K E p2/2m is the kinetic energy operator. The most interesting case is in connection with our best Hamiltonian ~ HHx containing the effective potential exchange. Since we have so far found this Hamiltonian only for two sites our discussion here will be only on the effective potential for this case. The diagonal element of (163) namely b11=h11 + v1212 , to the leading order in the overlap can be reproduced if one uses for W the Wigner-Seitz potential st . This is defined as the potential "seen" by an electron in a Wigner-Seitz cell. For two sites, the Wigner-Seitz cell associated with an ion is the half-space from the mid-point of the site axis, including that ion. In this case, st is the potential due to the singly positive ion in the cell that the electron finds itself, plus the potential due to the other atom. We can write then that 3 wi(£f)d r' = Z I _ ( I- 1 WS k j C811 j E B‘j gfj —I_r_ £1 1 .TE:E;T)] wk(£) . (164) 76 With (164), the diagonal element of (163) is 2 2 _ 3 _ e e — f d r w1(£)(K )wl(£) all space I£_Bll '12"le 3 2 "3(E')d3r' + f d r w1(£) f (165) cell 1 all space Igfg'l w (3') + f d3r wi(£) f d3r' 1 cell 2 all space Irfir'l which to zero-order in the overlap becomes 2 2 2 2 r w (r)w (r') + f d3r I d3r' 1 ’ 2 ‘ = ‘ LIE-51' '12"le |£-£' ' = h11 + v1212 = b11 ' (166) If the Wigner-Seitz potential is used to calculate the off-diagonal element of (163) it does not reproduce our best b12 given by (155): = K + 6 12 (167) 12 where K12 s (168) 77 and 912 : ' 2 2 = - d3r w1(£)w2(_r_) e - f d3r wl(£)w2(_r_)—e——— cell 1 _r_-§1| cell 2 Ig-gzl 2 2 W n --l d3r w1‘£""2‘£’ [_3_ _ ez d3r. 2(r) cell 1 gfgzl all space Iger'l 3 e2 2 wi‘E'MBr' - f d r w1(r)w2(r)[————-— - e ] (169) cell 2 — _ {-51 all space IE'E' which to the lowest order in overlap is 3 e2 3 e2 612 = - d r wl(£)w2(£)—————— — I d r w1(£)w2(£)——————.(170) cell 1 lgfgll cell 2 £~§2 Our best transfer integral b12’ to lowest order in the overlap, is V V I e2 I b = —:-—- t = —_—-— {K - } 12 v v12 12 v v12 12 1 Ir-Rll 2 = K12 + Y12 (171) where 2 Y12 = V—_-v {V12K12 " V} (172) 12 rfigll =_ [2f w(_1;)w(_r_)-———dr-K] WS 2 V V12 cell 1 l 2 IE—Bll 12 V I e2 3 + _ [ W (9W (E) ———-—- d r V v12 cell 2 1 2 |£7§1| e2d3r - I W1(£)w2(£) ] cell 2 lg-gzl which does not appear to be zero in general. In fact it seems to be of first order in the overlap, and therefore it is not expected to be negligible, although we have not made a numerical estimate. That an improvement over the Wigner-Seitz potential is needed can be seen intuitively when one realizes that for small overlap most of the hopping integral comes from the region where the electron position r is in between the ions. For the WS potential, the "other" electron discontinuously jumps from one ion to the other as the electron of interest (at r) crosses the midpoint (which seems unreasonable). The quantity t12 is easily seen from eq. (31) to be the matrix element between w1 and w2 of K and the potential due to the nuclei plus one electron 2 distributed equally over the two ions (in w1 and wg), a much more reasonable picture. However, the factor 79 V/(V - v12) increases the difficulty of finding explicitly the corresponding potential W (which we have not been able to do), this W appearing to be a complicated one and a function of the kinetic energy. VI. SUMMARY AND DISCUSSION A new approach to the problem of narrow-band magnetic semiconductors has been presented and investigated in the context of the half-filled single band problem. The approach consists of using the minimal principle of statistical mechanics for the free energy, to find a "best” Hamiltonian H, containing one-electron terms as well as certain important two-electron terms, but much simpler than the exact Hamiltonian. It is important to realize that if H consisted of only a one-electron term, the thermal Hartree-Fock approximation would have been obtained from the variational principle, as the best one-electron Hamiltonian, which has been proved to have its failures for the narrow-band problem. It has been recognized that one-electron approxima- tions are definitely bad and that two-electron Coulomb interactions are very important for narrow-band electrons. Hubbard and Anderson, working not inconsistently with this view, presented theories in which one-electron-approximation terms (Hartree-Fock) are to be used in combination with 80 81 important two-electron terms, to make physical predictions. Although on the right track, their theories were shown to be unsatisfactory (sections II and III), essentially because they made a one-electron approximation first and then combined the resulting one-electron term with two-electron terms. Instead, we find the variationally best Hamiltonian H with both types of terms: the best one-electron term is found simultaneously with the best two-electron terms. The investigations discussed in this paper in the context of the single-band model showed that our approach overcomes the difficulties cited in the previous theories. In particular we studied two cases. In one, the trial Hamiltonian was taken to be of the form of the Hubbard Hamiltonian, and in the other, effective potential exchange terms were added to it. In the first case the low-T results matched the low-lying exchange splittings appropriate to both trial and exact Hamiltonians despite of the lack of potential exchange terms in H . Hence we accomplished Anderson's objective (not attained by him) of defining a non-magnetic one-electron Hamiltonian, whose transfer integral leads via the kinetic exchange to the exact low-lying splittings to leading order in the bandwidth. However this first approach fails to give a physically acceptable approximation to the properties of the exact Hamiltonian when the potential exchange dominates the 82 kinetic exchange. In the second case, explored so far in the simple case of two sites, the same matching of low-lying splittings is obtained, but for any_ratio of potential exchange to kinetic exchange, which certainly is an important improvement over the first trial Hamiltonian. Further, the centers of gravity of the (NS 1 1) - electron levels are matched to zero order in overlapSI. (The first trial Hamiltonian erred in zero order for those levels). As far as the near future of the exploration of this new approach is concerned we have at least three immediate problems to consider. The first is the generalization of the results of the best Hubbard Hamiltonian augmented by the potential exchange, to an arbitrary number of sites. The second is connected with the problem of superexchange, where one has closed-shell ions and non-closed—shell ions (MnO, for example). Here we hope to find that the appropriate set of Wannier functions (the best in our procedure) will achieve the rapid convergence of the perturbation theory originally desired by Anderson. The third has to do with adding higher cation bands; our approach applied to a model containing such bands should enable one to take into account the quantitatively important 35 "electron rearrangement" effect , as it modifies the one as well as the two-electron parameters. 83 Finally, we would like to point out that a rather remarkable suggestion concerning the broad-band or high-density limit results from our narrow-band con- 10, the restricted HF approxi— siderations. As is known mation for the electron gas leads to the incorrect result that the density of states is zero at the Fermi energy, and the source of this error is the long range of the Coulomb interaction via the exchange integral. Thus it would seem that the addition in H, of a short range interaction (like the Hubbard intra atomic term) to the one-electron operator would not be able to overcome this fundamental difficulty. However, this View may be quite incorrect. We essentially saw (Sec. II) that the restricted HF approximation led to a hopping integral bi' J narrow-band region, the source of the error again being that behaved very badly in the the long range of the Coulomb interaction via the exchange contribution to the HF eigenvalues. But, as we saw, the addition in H of the Hubbard intra atomic interaction was quite sufficient to overcome the diffi- culty found in the narrow band region. Thus the possibility that it might similarly overcome the diffi- culty in the region of large ratio of bandwidth to intra atomic interaction no longer appears to be an unlikely one. It is under investigation. APPENDICES APPENDIX A1 TWO-SITE SINGLE-BAND MODELS The exact single-band Hamiltonian Specialized for the two sites case is + N N ) + v N N + H = h11(N1+N2) + V(N1+N1+ 2+ 2+ 12 1 2 . _ _ + + + g [h12 + v12 (N10+N20)](cloC20 + cZoclo) + + + + + + x12 g [3(C16026 + C25C15)(Cloc20 + CZOClo) ' N1oNzo] (A1°1) where v12 5 v1212 ' : v12 _ v1121 (Al.2) x12 5 v1221 This Hamiltonian can be straightforwardly diagonalized, the eigenfunctions and the eigenvalues being shown in Table 1. There we also display the H - uN eigenvalues for the special case of = two electrons (in average) in the system. The chemical potential which fixes the number of electrons equal to two is determined by the equation: tr e-B(H-UN)(N-2) = 0 (Al.3) 84 85 or ] e - %B(V+2v -X ) B(t -v' ) -B(t -v' ) Ba 1 + e 12 12 SE? 12 12 + e 12 12 B(t +v' ) -B(t +v' ) 38a 48a [e 12 12 + e 12 12 1 e ;} - e = o (Al.4) where we have put X _ y, _ 12 u — h11 + 2 + v12 —§— + a (Al.5) and a is now to be determined. The solutions for some special interesting cases can be immediately found: 1) kT >>It12|+lvi2| a = o (Al.6) It12|>|V12| “ = “sviz 2) kT<<||t12| - Ivizll It | = (c c + + + + + + °2+°1+H°> ' Ia2> “ (C1+C1) + °2+°2+)|°> ' 89 |> H|> = EI> E - uN |0,0> = [0) 0 0 - = 2 I1 ,+> 2 (c1+-c :+)|o> h t + g V' t1 +V' _ _% c+ 11 12 1212 12 > = 4. ll ,+> = 2 l5(cl++c 2+)|O> + - ' - 1+ +> _ 2_% h11 t12 V12 C+t 12 v12 C ' ' ‘ (C1+C 2+)l0> [2-0> = A[ a > - C a > - - - - . l 1 l 2 1 2h11+X12+E v 2v12+2X12+E -2a lzfo> = A[Ia1> + C|a2 >1 2h + + 11 X12+E+ -V-2v12+2X +E -2a 12 I2 O> = 2 h( + c+ c+ 0+ > - - _ C+ + |2'++> ‘ C1+ C2+IO> _ c+ + '2'*+> ‘ C1+ C2+'°: 2h11+V12x 12 'V’V12'2C 2 ++> = 2;5 c* + I ' (C1+C 2+ C2+ C1+)|°> 3- +> = 2";5 I ' (C1+ C2+)C 1+ C2+“)> ”_t v. . I3 +> = 2_%(C+ _ + ) + + 12 12 C‘tlz'V12'3C ' 1+ C2+ C1+C2+lo> + '3 '1) = H%(C ++C2+)C1+C 2+|°> n+t +v' + + I3+ +> = ”C(c c+ ) + + 12 12 5 C112 V12 30 ' 1+ C2+ C1+C2+'°> _ + + + + |4,o> — c1+c1+c2+c2+|o> 4hll+2V+4v12 -4a -2x12 Table l 90 ~ + . Here E" = Table 2. The singlet eigenstates of HHX 2 (1/2)[(U + U + ((U — U12) 2 8 ~ C' = ( - fi'—)/2b and 3' = (l + C'2)-15 . U12 12 91 Singlet H I) = E. > Eigenfunctions HX + + + + + ~,— A [(C1+C 2+ + C2+C1+) C (C1+C1+ + C2+C 2+)]|°> 2b11 + E + + + + + + + + ~.+ A [(C1+C 2+ + C2+C1+)+C (C1+C1+ + C2+C 2+)”0> 2b11 + E -2 - (C1+C 1+ C2+ C2+"0> 2b11 + U Table 2 APPENDIX A2 NATURAL CONDITIONS FOR A UNIQUE CHOICE OF WANNIER FUNCTIONS IN THE CASE OF SINGLE BANDS. The Bloch functions, in terms of the atomic functions a(£f§n) are as follows: -;5 i£.3n TE(£) = (NSFE) g e a(£é§n) (A2.l) where 13.31“ Pk E 2 e Aim) — m A(n) a 1 51(5) a(£-§n) d3r = A(-n) . (A2.2) We have assumed a(£) =a*(£)=a(I£]) . Let S be a unitary symmetry operation of the crystal such that 551 = 3i. ; then one can show with (56-a) and (A2.l) that -'k-R.+iy -% 1— —1 k Sw (1;) = N‘ 2 e — \v (1‘) Bi 5 k SE = N 2 e —-W (r) (A2.3) s k E_— In particular let S be inversion through an origin at the lattice p01nt: in which case SEi = Bi = O and Ys‘lk = Y-k . 92 93 And further, demand that w .(r) be invariant under S (as the atomic function): This requires n (A2.4) where nk is an integer. If we demand the Wannier functions to be also real, as are the atomic functions, we must have = 2m n (A2.5) where mk is an integer. From (A2.4) and (A2.5) it follows that = R n (A2.6) with 2k being another integer. However, demanding that the Wannier functions go to the atomic functions as the overlap, (A2.2) for n # 0 , goes to zero we finally get: Yk = constant- APPENDIX A3 SOLUTIONS Cfi‘ EQUATIONS (60) and (61) Because of (63) and (64), the thermal averages in equations (60) and (61) are very simple to evaluate. For instance, 2 can be written as j 1 3+ 3+ Z tr pO NiNj+Nj+ = 2 tr pO NHNi+ + tr p0 NiNj+Nj+ — = 2g + (tr 5.N.) 2 tr 6.N. N. 1 1 . . + + 3#1 J J J = 2g + fi(N-1) 6 where g = tr 50 Ni+ Ni+ = tr 51 Ni+ Ni+ and fi = tr 5 N = tr ~ N . o 1 pl 1 These quantities g and a can be easily calculated with the help of Table 3. It then follows that “B(Zb -2u+U), a = e 11 /E and -B(b 'U) -B(2b -ZU+U) fi = 2(e 11 + e 11 )/§ 94 (A3.1) (A3.2) (A3.3) (A3.4) (A3.5) 95 where “B(b ‘U’ "B(Zb -2U+U) E = l + 2 + e 11 + e 11 . In the same way we did for (A3.l),we can show that eq. (60) and (61) can finally be written in terms of g and n as .1[2§ + fi(1-fi)] = o (U-V)(n-2)g + [hll-b + n X V1j13 1 . 1 J#1 and (U-v)(l-§) + [hll-bll + fijgl vljlj](fi—2) = o . From these two equations we find (U-V)f(fi,~) = o and [h -b + fi 2 v . .]f(fi,§) = o 1 . ll 1 j#l 1313 where f(fi,§) = fi(1-fi) + 5(35—2) ~2§2 . If f(n,g) = 0 , U and bll will be undetermined although there will be’a relation between them. Let us study (A3.11). Assume f(fi,§) is zero and solve for S in terms of 5 . The solutions are n = l+§ and fi = 2g . From (A3.4) and A3.5) n can never be equal to l + g ; (A3.6) (A3.7) (A3.8) (A3.9) (A3.10) (A3.11) 96 and also n can only be equal to 2g if n = a = 0 , which of course is of no interest here. Hence flfi,g) # 0 and (65) and (66) follow immediately. 97 Table 3. Eigenstates of fil and fil - uNl . The last mud column is for u = bll + (half-filled band). 98 (al-u~1)|> = (El-uN1)|> ; El El uNl E1 -(b11+%)N1 |o> o o |+> bll bll - u - u/z Ii) b11 b11 ’ “ ‘ ”/2 |++> 2b11 + U 2b11 - zu + U 0 Table 3 100 Straightforwardly we can show that ~ + + + [H0. ckO(S)] = U Nk0 ck0(s) - u Cko(s) (A4.8) [RO, c£0(s)] = -U NkO c£0(s) + u c£0(s) . (A4.9) Then we can rewrite (A4.5) in the following form 3 SE Ak£o(s) = BUo + + + + ijij<9ija Cio Cjo (Ni5+Nj3) cjo Cic>o + + + k;i,j vjkik (gijo(x) Cio CjO Nk Cjo Cio>o] l x + +Bg bji Io dx {yo ds [o - + + —o] ‘Il d5 < ’( ) '(S) + c+ c d> o gijo X gjio Cjo Cid i0 jo o + = 0 (A5 1) le x jo id id jo 0 ° where _ l ' d = X (hkk - bkk) Nk + 3 2 Vk£k2 NkN2 (A5.2> k k,£ We have used the fact that gjiE(X) gij5(X) = 1 by the definition (93), and also that U = V. 101 102 To evaluate the thermal averages in (A5.l) we recall the fact that 5 = H 5 , where o k k ~ -B[fi - uN ] ~ 0k = e k k /z (A5.3) ~ _ ~ - -B[Hk - uNk] (A5.4) z z z - tr e k and Hk = bkk Nk + U Nk+Nk+ ; (A5.5) and because (99), (100), and (63), for half—filled band we have fi - UN k = V/2 [-N + 2N k kf Nkil . (A5.6) k The traces in (A5.l) should be taken over the eigenstates of fi - uNk , which are shown in Table 3. But before doing k that, it is convenient to write: = Vijij[NiNj - (Ni+Nj)] +kgi jvkikiHNiijNk - 1)- 2 Nk] + 2' v [(1/2) N N - N ] . (A5.7) . . 1 2 2 kl£#1lj k k k k We have considered 2 v . . = 2 v . . . Above, 2 k#i.j klkl k¢i.j kaJ k.2¢i.j means sum with k#i,j and £¢i,j and the prime means k#£ . 103 Then eq. (A5.l) becomes 1 X f dx(hji — b.. + Z + v. . ) + o o 31 k#i,j jk1k 130 10 30 30 10 o + + ijij< gijo‘x x) Cio Cjo (N13 + NjS) Cjo Cio>ol + l _ + + + 8 gb ji Iodx'{}ijij If: ds (< gijo (s) ch ci0 ci0 cjO (N.N. — N. - N.)> 1 j 1 j o + -o) l + + _I odSo + -o] + _[2 z .Vkiki + 1/2£ {#1 j vkzkfil If: ds(< gij5(s)c jg Ciocioc jO>o _ + + ‘(X) g --(s) c. c. c. c. > ) - S gijoX gjig S jocio 1oC jg o + + + 013 = 0 (A5.8) The next step is then to evaluate the thermal averages and then do the integrals. As an example we do the first one: 104 1 X 1 dx o o 130 10 30 30 10 o xBVN.- -xBVN.- + = Z I dx o o 10 o 30 30 o l 1 L— (2/22) I dx (e’z8V + eXBV) (1 + e<2 X’BV) 0 8V (4/22) (eléBV + EEVZl) . (A5.9) Similarly we calculate the others and finally find the result given in (102). APPENDIX A6 PROOF THAT fieff IN (114) CORRESPONDS TO A HEISENBERG HAMILTONIAN The operator fie in (114) can be written as ff ” = - _ii + + = Heff Z. X . U F[Ciocjocjo'cio'1P 13 00 2 (A6.l) — - 2' 2 Eli P[N — N N - c+ -c+-c ]P — ij 0 U i0 i0 jg iocio jo jg The P operator can be dropped if we remember that fieff is to operate on the ground level only; in other words on states containing only one electron on each site. The relationsbetween the above creation and destruction . . 29 operators W1th the sp1n operators are6’ : + — S — S +'S Ci+ci+ ‘ 1+ ’ ix 1 iy + . c. c. = S. = S. -18. 1+ 1+ 1- 1x 1y 1 §*Ni+ Ni+) ‘ Siz ' S1nce in our case Ni+ + Ni+ = 1 then _ l Ni+ — f + Siz _ l _ Ni+ ‘ i Siz 105 106 With (A6.2) and (A6.3) in (A6.1) we can write 2 b.. .. _ _ 1 _ _ _ _ Heff - i5 —U-J— [1 (15+Siz)(35+5jz) (L. s12) (15 sjz) - si+sj_ - si_sj+] = = ' Z. U [:5 _ 2§-i°—S-j] ' l] The eigenvalues En' (115), of EH calculated through second order degenerate perturbation theory, differ from the eigenvalues of the Heisenberg Hamiltonian with __ _ 2 Jij 2bij/U, by the constant I 2 Nsbll + Ej bij/ZU - APPENDIX A7 CALCULATION OF EQUATION (121) Taking the expression for the exact Hamiltonian given in (3) we have for the first term in (121) — 5 Efi Vk£k2 + (A7.l) [no >- - with only singly-occupied sites, give back another function |m0> with also only singly-occupied sites. The first term in (A7.l) is clearly I ~ = ~ , Z vkfikkanm because NkN£|n0> Ino> , the second term, k2 NIH C+ [5 koc 20 C20 Cko the curly bracket, can be written as after adding and subtracting inside c+ l ' 7 Z vk££k[§0 (m o'CkoC 20 C20 Cko' |fi0> - 6 ] . (A7.2) k2 nm The potential exchange integral Vklik is a quantity of the order of the square of the overlap between k and 2 . It's clear that overlap between non-nearest neighour sites is smaller than nearest neighbour overlap; then neglecting 107 108 terms which are higher than second order in the nearest neighbour overlap we can then write (A7.2) as X X X — — X 6 . (A7.3) 12 00. 0 kg 20 £0 kg 0 2 12 nm Be a s 2' + c c+ is ro t'o 1 t H = Pi 0(1)&P of which [E > is an ei nf nctio we eff U o ge u n then get ~ ~ _ l ' _ 1 l ] . (molHlno> ‘ 6nm ['2' 1E2 Vk2k2 5: X12 + 5' X12 An (“'4’ The second matrix element in (121) can be written as 2b [ X X h.. 2 . . ] I I l < > . I O 10 |O IO 10 O X Z v + (A7.5) + .. ii'i ' ' ' <13) 00' j j 1o 30 30 10 ~ + + ~ + . . < N . _ . ' . > . <§j> £#i,j Vlkjk g0. mol k clocjoc30 CIO'Ino ] . + + ~ . . . It 18 clear that ciocjocjo'cio'lno> 15 an e1genfunct1on of N13, Nja and Nk , with eigenvalues 0,1 and 1, respectively (k#1,j). Then it follows that A 6 . (A7.6) (1)~ ~ = H O Tlno> b12t12 n nm 109 The third term in (121), after neglecting higher order terms, becomes (1) (l)~ ~ _ 2 ~ + + ~ Tlno> _ b12V X 2 X <2 1 mOIT Q . . ' <1,j> k 00 H Q + 12 k£ k£k£ oo' o 10 30 k R 30 1o 0 _ 2 2 1 ' 12 n Cm oo' 2k,£#ij x o 10 30 30 10 o ] A 6 . (A7.7) l 0" Finally, we obtain the result shown in (122). APPENDIX A8 CALCULATION OF TERMS IN EQUATION (127) With (127) and (128) we can write a BHH 3A (1) > .(A8.1) And for A=U this becomes {-l <fi l (1) U 0 Q(l)TInO > + l <§O|TPH|fiO>]} + ~ ~ ~ ~ 1 ~ TQ Ni+Ni+lYo>[ -U (YOIH U + higher order terms. (A8.2) Evaluation of (A8.2) is similar to the ones we did before, as we see if we rewrite it in the form _1, ~ (1) (1) ~ (1) (l ) (1) ~ ~ UmaTo Nmuo “%>+UTJTQ NHN.N Ho Thy _1 ~ (1) (1) ~ ~ ~ U2 - Then it becomes 2 b t b 12 12 12 [ U + —55 (V v12)] An , (A8.4) 111 and the whole second set of terms in (124) is, for A = U: b b 12 12 For A = b (A8.1)turns out to be 2 <fi [of c I? >[ — l - <.. o 10 jo o o 0 U o o 13> (A8.6) - l <~ In o‘l)'1'~|fi >1 = -[b———12(V-v ) - t 1 A U Yo 0 U 12 12 n ' and the whole second set of terms in (64) is, for A — b , b kT 12 l—fi-(V v -U_ 12) — t12] , (A8.7) LIST OF REFERENCES 10 11 12 13 14 REFERENCES J. C. Slater, Phys. Rev. 81, 385 (1951); 83, 538 (1951). P. Hohenberg and W. Kohn, Phys. Rev. 136 B, 864 (1964); W. Kohn and L. J. Sham, Phys. Rev. 140 A, 1133 (1965). L. Hedin and S. Lundqvist, Solid State Phys. 23, (1969). F. Seitz, Modern Theory of Solids (McGraw-Hill Book Co., New York, 1940) p. 329- J. Hubbard, Proc. Roy. Soc. A 216, 238 (1963). P. W. Anderson, Solid State Phys. 14 (1963). By this term we mean the standard thermal Hartree-Fock approximation. For the special case of zero-temperature, see e.g., ref. 8; and for the general case see ref. 9. J. C. Slater, Quantum Theory of Molecules and Solids (McGraw-Hill Book Co., New York, 1963) Vol 1. For a recent review see A. Huber in Mathematical Methods in Solid State and Superfluid Theory, Scottish Universities Summer School (1967) (Ed's R. C. Clark and G. H. Derrick), Plenum Press, N.Y. (1968). J. Bardeen, Phys. Rev. 50, 1098 (1936). C. Herring, Magnetism, Vol IV (ed. by G. T. Rado and H. Suhl, Academic Press, New York and London (1966)) p. 310-311. T. A. Kaplan, Bull. Am. Phys. Soc. 13, 386 (1968). T. A. Kaplan and P. N. Argyres, Phys. Rev. B1, 2457 (1970). R. Bari and T. A. Kaplan, Phys. Rev. B6, 4623 (1972). 112 113 15 - The conclusion that the HFA is unsatisfactory for such 16 17 18 19 20 systems has been reached by Slater, Mann, Wilson, Wood, Phys. Rev. 184, 692 (1969), sec. IX, on grounds which apparently differ from those of ref. 12 but really are very similar. Namely, it was noted by Slater (reference 8 secs 5-2, 6-2) that atomic excited states which con— serve the number of electrons are not given at all correctly by the HFA. Whereas Kaplan's point (ref. 12) that the zero-point entropy is incorrect in the HFA in the atomic limit, is essentially the statement that the low-lying spin-flip excitations in a small magnetic field (which conserve electron-number) are given very badly by the HFA(even for a hydrogen lattice in the atomic limit). See J. C. Slater, T. M. Wilson, J. H. Wood, Phys. Rev. 119, 28 (1969). J. C. Slater, Phys. Rev. 165, 658 (1968). Despite wide study of this method in atoms (see, e.g. ref. 16 and 17). They are actually overcomplete as can be seen from the simple exactly soluble two-site two-electron Hubbard model. This contrasts with the broad-band or high-density situation where the interactions presumably can be treated perturbatively in a calculation of the single-particle propagator. In this case the fs(x) which grow out of the non-interacting set f:(x) will have large norms (~1) whereas the others will have small norms. 21 22 23 24 25 26 27 28 114 It is these "large" fs(x) that are apparently taken (ref. 3) to correspond to the results of a band calcu- lation. L. Matheiss, Phys. Rev. 23! 3918 (1970). See J. C. Slater et al, of ref. 15. For more recent work along these lines see D. J. Newman, J. phys. c _6_, 2203 (1973). Nilton P. Silva and T. A. Kaplan, Bull. Am. Phys. Soc. 18, 450 (1973); Nilton P. Silva and T. A. Kaplan, AIP Conf. Proc., No. 18, Magnetism and Magnetic Materials (1973), p. 656. J. Kanamori, Prog. Theo. Phys. 10, 275 (1963); M. Gutzwiller, Phys. Rev. Letters, 19, 159 (1963); Y. Nagaoka, Phys. Rev. 111, 392 (1966); D. I. Khomskiy, Phys. Metal Metallor. 12, 31 (1970); E. H. Lieb, and F. Y. Wu, Phys. Rev. Letters, 20, 1445 (1968); A. A.Ovchinnikov, Sov. Phys. JETP 10, 1160 (1970); D. Cabib, and T. A. Kaplan, Phys. Rev. B1, 2199 (1973); R. A. Bari, and T. A. Kaplan, Phys. Rev. B6, 4623 (1972); and many others; to find other refs. see H. Shiba, and P. A. Pincus, Phys. Rev. B5, 1966 (1972). This model has been used quite commonly, e.g. by Slater (ref. 27) for H and by Mattheiss, Phys. Rev. 123, 1219 2 (1961). Reference 8, Chapters 3 and 4. E. C. Kemble (1937), "The Principles of Quantum Mechanics", Sec 480, McGraw—Hill, New York. (Reprinted by Dover, New York, 1958). 29 30 31 32 33 34 115 T. A. Kaplan, unpublished lecture notes for a course in Magnetism given at Michigan State University in 1971 and 1972. The exponential behavior of wi(£) was found under the assumption of zero phases Yk of the Bloch functions (see App. A2). Also see W. Kohn, Phys. Rev. 115, 809 (1959) for a similar result in the case of one-dimensional bands of arbitrary width. Since Hubbard's explicit intent was to consider a model appropriate to transition metals, where presumably it is essential to have a broad band (45) cross the narrow band (3d), our proof that Hubbard's derivation leads to grossly incorrect results is perhaps unfair because we consider an isolated s-band (more appropriate to an insulator). However Hubbard did apply his model to insulators (Proc. Roy. Soc. A 181, 401 (1964)). In any case, it is important to realize that this type of Hartree-Fock approach fails. See discussion in Kaplan and Argyres, page 2457 of ref. 13. In particular, Kaplan (ref. 12) noted that in the thermal HFA the entropy + 0 as T+ 0 for zero bandwidth as compared to the exact result Nk 2n (28+1), where S is the ionic spin; this is essentially the reason for the fact that as T+ 0 the spin-susceptibility x+ 0 in HF whereas the exact result is x+ ot=(the Curie law). These show that the 35 36 37 38 39 40 41 42 43 44 45 116 non-zero-T HFA gives results that are vastly different from the picture in the Heisenberg model, where in the atomic limit Jij+ 0 , so that there are N non-interacting spins. See T. A. Kaplan, AIP Conf. Proc. No. 5, Magnetism and Magnetic Materials (1971) p. 1305. N. Fuchikami, J. Phys. Soc. Japan 88, 871 (1970). It is interesting to note that t12 is the average of the two biz (+,+). Fuchikami (ref. 35) gave unique values for biz (+,+); she presumably considered biz (+,+). Furthermore, this 0 dependence has been shown explicitly (T. A. Kaplan, unpublished). Nai Li Huang Liu and R. Orbach, AIP Conf. Proc. No. 10, Magnetism and Magnetic Materials (1972) p. 1238. K. I. Gondaira and Y. Tanabe, J. Phys. Soc. Japan 81, 1527 (1966). Hubbard, Rimmer and Hopgood, Proc. Phys. Soc. 88, 13 (1966). Petros N. Argyres, T. A. Kaplan, Nilton P. Silva, Phys. Rev. 88, 1716 (1974). Hund's rule was implicitly assumed in ref. 21. S. V. Tyablikov, Methods in the Quantum Theory of Magnetism, (Plenum Press, New York, 1967) p. 196. ~ -2aRij 'aIE-Bil Actually vijji ~ in a Rij e for ai(r) ~ e '“Rij v in which case Aij ~ e ; thus vijji = 0(Aij )where v is less than but arbitrarily close to 2 and it is in this sense that the term second order is used. See ref. 13. 46 47 48 49 50 51 52 117 R. S. Tu and T. A. Kaplan, Physica Status Solidi (b) 88, 659 (1974). R. A. Bari, Phys. Rev. 88, 2662 (1971). See also D. J. Klein, W. A. Seitz, Phys. Rev. 88, 2236 (1973). It can be seen that the second order corrections to the wave-functions omitted in Eq. (109) contribute to Eq. (118) only in higher order than the terms we keep. That this might occur in situations of physical interest has been noted in ref. 39. Under certain conditions (|vi2I