7 . ~.,—~. -- “WM-9:1- . ,..|'. I,"~.- . J , u'n ' ' I"'n1 u. _: .1..‘."‘ ' " - (".II‘I2I"',"‘.~..'5.I‘-'3fi- "352:" "MI 2““. 1" .\'-I1-' ,11'3 ' 11.11.1116“ '1’?» ' 'II 11"1'1‘1." -‘ II N .4. : 1.1w~.--.wmw ,- :" I' Riga. ‘ "II! ‘J L}! El. 2-":‘1 I ' I ‘ ‘~‘ .‘ ‘ u I 111.” "11?". p' .1..#z1:1‘t:11u'\h' $11.17;; 44"”: ‘5 11 11.1.3)“. \ . .' 1' 1".""I‘“.I' ..-"',,-.-:- 11w- .,- ..-' .- , .- - . I : "I :11"... 'I ' 'x "J." .\‘1 i 'l? > " I11'II"‘-‘:1'II:.I111111III'IA1-"11'5'111‘Ig11'II ‘ i" “1'3"“? 1‘11" 5.”. ilfif'”. .I' "'.A ' ,DI ‘71:}; 1,(.( .I. '. W C". I :.I.I_ ’ : 2' I~m . 1.1.«r-yafiwymnwc- g. g1! I11-||"1(.$p‘h&" 2‘31“ “‘11. :13. 'fii-Hfi‘jith 1 1“! fr “1.1, l.: . m-Z-R‘g VI ”.l‘l 1" 1") "V I... ' u.‘:.. 1'11? 111-] II" ‘1 "1r". 1 ‘1 11 1" 1 " 11111 . ’13 1'35. . \IW I "( I..I.)-‘II- xi. "'2“. . 8'" 1‘1 I 11 I"'*IIII..I";<““‘""II’ ' 1| -,QM‘1I¢$W11%I'H ., 1" ‘1 11111- 1'11 f. 11131.- W. 1 ' 1 II..- .. -. '.’ ‘ I~ _' II':|.|' I 1‘ $0 II . 1‘ -.-. , 1‘ '51? I11 i'r“ 111:. £40. .1.- 1.1-. .11 1 LI“? ' 1...” 'Il1l1l‘\ll1lfil;' . l;‘\'r {S'W 1111.314 .1 '1“i".u"5"1:11'"§l ”I11“ 1 h": 1 'I I (‘N'IRHSP {115.1.” ”1:. 1:13;” I333} .111“ :1 ' "1 "1" "’ 1-1-3.- ’-‘ I13. .&:1.Ii?fis 11. I ‘, 9-"-‘I"'."I".1‘ , ,. .‘11'I‘1I..1.{‘V4‘I~1I‘ U111: 1'11 (I 9%?" 11:“1‘1510.’ 1.\11‘( '1‘??? J'Izr"§.fi“ '1'. ‘. .‘\‘1‘E.I~'.I “Q. 1 ' .111'25': :JJ‘IJR‘I‘ ”‘ '1‘ h”! \".1 1m ”1““... “3:7?! ”1.. N\'.\’(‘ v: '1'. 1% "1‘”,',-"‘1: 1 1111"”‘1l1 I51" I'II“I . {13133110601‘ . ."_‘I. H "ILITI .‘;"- ”1.0111313; "-13 .' .h "'.' 1_ 1.51.0.1 1f 1' I . It. .- 1. 1-! .- X1‘L"I\"2‘11 ".1" ' "1:111 "_..V1I , ' III",.*..I’- "' 1 1 1.).-- . ““1 11112-11)" ' J1“. 1 I1 .‘s .0 1I | 1.,W .gw‘ U ’ng .4 A -. H“ 1“; ”7‘ I I” .,'.‘.:"‘_ . 11“" $11.45.". ':21'.1.‘:'I'E."‘1‘I}:;IIH" ' ' ‘II'h {- if“: , :1‘.‘.‘ :1‘E‘1g: ‘JKI;I.I‘,%‘%‘§L "' 'g’flr‘fib’ , ." I " "1 11 1‘11 "' "'-'-.IgI1'5}1'-111‘\$‘,1"'1‘HI mh‘fl " 1‘15“."7-"1‘ “1" "fl." . » I ‘..I|'III' H. h -_I. 1‘ y - .II ’ '1‘ . p I I . . : Ii..l'r"'I"I')>1 >1” .1”- l' '1.'1I " 11'!“ III"I‘1V‘.‘SII""II' ':;’:"P’€ '1‘“ ‘tfh‘fi'l ', [:1 1.1.7:" gzi'11‘v ”it" I} ,1 I"? ‘ £331.11 7-1-1 3.'I‘.'.3.?1\'I":.. "'7!- '1.I'I":I.-'."711' 1.. ‘1'”"6-‘1-I'31?I'I'-w'.*."- ~" I ‘g,':...-‘.':~;.:.".. .4- ..II.' .1101. 1 .111“ .1‘1','-I..I..:.Ij, -11 1 I \'l‘1".\‘ 1.11. ~1.’ ‘. , -'-‘ ., .- .I'«-. .- "11'“. «V. '4 - . - I - .. 2 1-11-1111: ”11-- - -. '-‘.':~.." i w .- i' 1111' ' ".‘v- ""11' " “ “'1 "'-1""II""' "‘ 7913116111'11.'3"!£"-.'-1v~-'II‘I "'3'" II 11 ’.' "'I‘IIII 1 I: '1‘ 1111‘" "1 ‘-".1'.".I-"11"' ‘1‘" "I "111\"""I"""II""-'\' .8I ., '1 I 1| I 1 _. .D.‘ .1 1 .' .- ._I. 'l' I- .1 .l) .' -. n I"" 1.1.. I 1. 1' H I. ‘11l'llll .1. 1.11 1‘.. '11I ' ." 'J‘Mflo ‘4‘31‘1'Q' ,1?" ‘ch 4,11! ,Iuwm1I1N-nvnwm... I1QU‘ mHuMHw? "'I' I "'II \" II "'I 11' h ‘ ' ‘ I' ‘1 ‘ I II IA .1 {I I ' 0 ‘ I I .ll' 1.-l1h- In" - 1. 1' . - ‘9 . ”2;. 1-.1' .131 -., I‘M. .1 . Iv ..\~ I " HI. . '\1 '. " 111" MIN! 1 IIII'T "-" -' I'I ‘ .I II" II ','1' 'I.'I" II.‘\'1I"I~h'I ": I“. I" 1 I. ‘9‘1'-;-“‘ LYN-1...“; I I" IIII‘". 1. 'I. )1?‘XI1". “'-' '.;2. - ' ' " '1 1- :I-I. - -‘ . I '11.: .1 1 . I ' 1'.’ "It I II I III1'1'1I1-"IIII “1'. . II1‘112xf111‘II"II" III” III III III III THESlg‘ W3 ‘1293101667 I LIBRARY l Michigan State University \_-1_, “:1 f This is to certify that the thesis entitled PROBLEMS IN ORDERED AND DISORDERED ISING SYSTEMS presented by JOHN MARSHALL THOMSEN has been accepted towards fulfillment of the requirements for Ph.D. degree in PhYSiCS WW/es Major pro ssor M. F. Thorpe Date I? ”up ({1ng 0-7639 MSU is an Affirmative Action/Equal Opportunity Institution MSU RETURNING MATERIALS: Place in book drop to LJBRAfiJES remove this checkout from .—3—_ your record. FINES will be charged if book is returned after the date stamped below. _ ._‘__7._____ r_ -, PROBLEMS IN ORDERED AND DISORDERED ISING SYSTEMS By John Marshall Thomsen A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Physics and Astronomy 1984 oi 2/9/37" hp} ABSTRACT PROBLEMS IN ORDERED AND DISORDERED ISING SYSTEMS By John Marshall Thomsen A wide range of Ising systems are studied. In one dimension, attention is focussed on calculating the transverse susceptibility. By combining transfer matrix techniques with linear response theory, it is shown that Xi can be calculated for pure and dilute spin S Ising chains. The low temperature behavior in the presence of a parallel field is discussed. The temperature dependent probability distribution of internal fields, P(h), is defined. and it iS shown to yield several exact thermodynamic results. This function is studied on a variety of Ising ferromagnets, with exact results available in one and two dimensions as well as on the Bethe lattice. Three and four dimensional lattices are studied with Monte Carlo simulations. It is shown that P(h) is very sensitive to short range order. This order is in turn sensitive more to the lattice dimension and less to the coordination number. P(h) is also studied for the Sherrington-Kirkpatrick infinite range spin glass model. An analytic result is available in the paramagnetic phase while Monte Carlo simulations are employed in the spin glass phase. The de—. velopment of the dip at h = O is investigated in some detail. Finally, some speculations are made on the possibility of extending the P(h) formalism to non-Ising systems. ACKNOWLEDGMENTS It is a pleasure to acknowledge valuable discussions with Drs. T. A. Kaplan, N. Rivier, D. H. Lee, S. de Leeuw, J. Bonner, and J. S. Thomsen. I am in particular indebted to Drs. J. Parkinson and S. D. Mahanti each for a series of discussions which included important insights and probing questions. E. Garboczi, H. He, D. Heys, Dr. D. Sahu, and P. Murray have all been good sources of advice during the process of converting general ideas into concrete results. I I would also like to thank S. Rosenbrook, J. King, and especially Dr. J. Kovacs, whose efforts allowed me to concentrate more of my time on research than might otherwise have been possible. I am indebted to Drs. D. Stump, J. Bass, S. D. Mahanti, and M. F. Thorpe who, as members of my guidance committee, must tolerate my writing for another 150 pages. As my research adviser, Dr. Thorpe has suggested to me more interesting problems than I could possibly have had time to solve. His insight helped guide me through this wide assortment of problems to produce some coherent results. I should also point out that while the preprint in Appendix A was a collaborative effort among myself and Drs. Thorpe, Sherrington, and Choy, most of the writing was done by Dr. Thorpe. I have in addition benefitted from some correspondence with Drs. Choy and Sherrington. ii Finally, I should like to thank Amelia Hefferlin whose en- couragement and assistance have helped bring this project to com- pletion. No less important have been her occasional reminders that Physics is not all there is to life (which is easy for an Ecologist to say). Most of this work was performed under the auspices of an Exxon Fellowship. iii TABLE OF CONTENTS List of Tables ...... . ..... . ...... . ............... ........ ...... .-V List of Figures ................................................ ~Vi 1. Introduction ................................................. l 2. Transverse Susceptibility in One Dimensional Chains..........6 3. P(h): The Local Field Probability Distribution Function ..... 26 4. P(h) for Bethe Lattices ...... . .............................. 33 5. P(h) for Two Dimensional Ferromagnets ....................... 41 6. P(h) for Ferromagnetic Systems of Higher Dimension .......... SO 7. The Sherrington-Kirkpatrick Spin Glass Model ..... ...........58 8. P(h) for the Sherrington—Kirkpatrick Model..................68 9. Future Work ................ .............. .......... . ........ 82 Appendix A. Local Magnetic Field Distributions I. Two-Dimensional Ising Models.................85 Appendix B. Some 2D Correlations .............................. 124 Appendix C. Details of Monte Carlo Simulations................128 List of References ....... ...... ..... . ......... . ............... 135 LIST OF TABLES Table 1. <0001>T as obtained from Monte Carlo simulations c and compared to high temperature series results .......... ...53 Table 2. Nearest neighbor locations for three dimensional lattices..134 Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure LIST OF FIGURES 1. The inverse of the transverse susceptibility per spin for a dilute (c=0.25) Ising chain in the pre— sence of three different parallel fields ............... ....25 2. Section of a Bethe lattice for z=3 (upper) and =4 (lower) ................................................ 34 3. P(h) for Bethe lattices of coordination 3, 4, 6, and w ..... 40 4. P(h) for the ferromagnetic square net ...................... 43 5. P(h) for the ferromagnetic honeycomb lattice ............... 45 6. P(h) for the ferromagnetic triangular lattice .............. 46 7. Generating the 3-12 lattice from the honeycomb lattice.....47 8. Showing P(h) at TC for the honeycomb (upper) and 3—12 (lower) lattices. ................ .... ................. 49 9. Extrapolation of <0001> to the infinite lattice limit for a simple cubic at TC................. ....... . .......... 52 10. P(h) for four different three dimensional ferromagnets....55 11. P(h) for simple cubics of dimensions one through four and for the mean field limit ......................... 56 12. The phase diagram of the Sherrington-Kirkpatrick model (reproduced from Sherrington and Kirkpatrick 1975) ....... 61 13. P(h) from the analytic result for the Sherrington- Kirkpatrick model in the paramagnetic phase (JO < J) ...... 72 vi Figure Figure Figure Figure 14. 15. 16. 17. A comparison between the Monte Carlo simulation of the Sherrington-Kirkpatrick model and the analytic result: T = ZTSG.. ......................... .. ........... ..75 A comparison between the Monte Carlo simulation of the Sherrington-Kirkpatrick model and the analytic result: T = T ...; .............................. . ....... .76 86 Results of the Monte Carlo simulations of the Sherrington-Kirkpatrick model: P(h) at four different temperatures. The T=O data has been symmetrized .......... 77 P(O) as a function of temperature in the spin glass phase of the Sherrington-Kirkpatrick model. Open circles indicate simulation results. The dashed line ‘W represents the linear hypothesis .......... . ............... /9 vii 1. Introduction A fundamental and much studied problem in condensed matter physics is the explanation on a microscopic level of the magnetic properties of various materials. As might be expected, there is frequently a competition between building a model Hamiltonian that describes a system with accuracy, and building a model Hamiltonian which is tractable. Fortunately, there exist a variety of systems which are described reasonably well by a simple spin Hamiltonian. For a system of well localized effective moments, one general form a spin-spin interaction may take is H12 = film-'3’2 (1) where 3+is a 3X3 interaction matrix and S1 and S2 are three compo- nent operators describing spins located at sites 1 and 2. The matrix elements of UIare related to the overlap of the electronic wave functions associated with the two spins. (For a justification of this Hamiltonian, see White 1983.) In certain cases, the Jzz term may dominate the matrix, reducing the interaction to the Ising (1925) form I z z 12 - -J8182 . (2) Systems well described by this form over at least a restricted H temperature range include CoC12° 2NC5H5, a one dimensional ferro- magnet (Takeda et. al. 1971), COCs3Br5, a two dimensional antifer- romagnet (Mess et. al. 1967), and DyPOA, a three dimensional antiferromagnet (Wright et. al. 1971). The advantage of working with Hamiltonians built with Ising interactions in the absence of a field is that every piece of the Hamiltonian commutes with every other piece. In the presence of external fields, the most general Ising Hamiltonian considered in this work includes one non-commuting piece: H = —%.Z.Jij3:8: - u hZZS: - uthisi (3) ll . 1"] l l where the Si are taken to be quantum spin operators. The matrix of Jij's is assumed to be symmetric with vanishing diagonal elements. The fact that the Si terms do not commute with the rest of the Hamiltonian will in general cause problems; however hX will be taken either to be zero, in which case the problems are eliminated, or infinitessimal, in which case one can work around the commutation problems. The Ising Hamiltonian has been studied in a wide variety of forms, and some of the more relevant studies are reviewed here. When first introduced, the partition function of a pure spin % chain with uniform nearest neighbor interactions was calculated (Ising 1925). Analytic results for the specific heat and parallel sus— ceptibility are now available for spins up to 3/2 (Suzuki et. al. 1967). The site diluted version has been studied by, among others, Matsubara et. al. (1973) for S=%, and Matsubara and Yoshimura (1973) for 8:1 and S=3/2. In both cases, the parallel susceptibility and the specific heat were calculated. IH The zero field transverse susceptibility of a pure spin 2 infinite chain was calculated by Fisher (1963), while Crest and Rajagopal (1974) repeated the calculation in a finite parallel field. Chapter 2 focusses on extending the transverse susceptibility calculations to include higher spins and site dilution, both with and without external parallel fields. Unusual effects arising from dilution and from the application of the parallel field will be discussed. Traditional approaches to studying Ising models focus on de— termining the free energy of the system and obtaining other thermo— dynamic quantities by differentiation. The subsequent chapters of this work deal with an alternative approach involving the distri— bution of local fields. This function, P(h), has been much studied in relation to spin glasses at zero temperature (see for instance Palmer and Pond 1979 or Kirkpatrick and Sherrington 1978). However, it will be shown in Chapter 3 that at all temperatures, the thermo- dynamics of an Ising system may be reformulated exactly in terms of the temperature dependent P(h). From P(h), one may obtain not only the magnetization and energy (and hence the free ener8Y). but also the inelastic neutron scattering cross section. Furthermore, since P(h) describes what is happening locally, it can give some intuitive feel for what is happening to the system. It was for these reasons that a study of P(h) on various Ising systems was undertaken. While the original motivation to study P(h) lay in spin glass research, it is useful to first study this function on well under- stood uniform Ising systems in order to obtain a basic understanding of how it behaves. In Chapter 4, P(h) is calculated for Bethe lattices with nearest neighbor ferromagnetic interactions. The Bethe lattice allows one to make a fairly smooth transition from a network of small coordination to the mean field limit, and hence may provide some insight into the effect of coordination number (or perhaps dimensionality) on this function. The Bethe lattice provides well behaved analytic results, but it is not a true lattice and the large fraction of spins residing on its surface makes it unphysical. For this reason, it is useful to look at the two dimensional Ising model. This model takes on a special significance in statistical mechanics as it is one of the few exactly solvable models exhibiting a phase transition at a finite temperature. The partition function of a square net with uniform nearest neighbor interactions was first derived by Onsager (1944). He extended the transfer matrix technique introduced by Kramers and Wannier (1941). Since then, a variety of solutions have been developed. The Pfaffian approach taken by Montroll et. al. (1963) seems to lend itself most readily to calculating cor- relation functions, although it may not be the quickest way to determine the free energy. This approach has since been generalized to a wide variety of two dimensional lattices (see, e.g., Hurst 1963 and Stephenson 1964). These results are drawn on in Chapter 5 to perform exact calculations of P(h) for several two dimensional Ising systems. The general features of P(h) on these two dimensional lattices are relatively insensitive to coordination number. However, there is a strong dependence of P(h) on lattice dimension. This depen— dence is explored in Chapter 6 using Monte Carlo simulations. Results in three and four dimensions are compared to the mean field limit. The models chosen, nearest neighbor ferromagnetic inter- actions on diamond, simple cubic, body centered cubic, face centered cubic, and hypercubic lattices, have all been studied in detail with high temperature series. Since these series provide suffi- ciently accurate estimates of TC for the purposes of this study (see, e.g., Domb 1974), it is possible to concentrate on P(h) with- out getting bogged down in details such as locating Tc' In Chapter 7, attention is turned to a spin glass model, the Sherrington-Kirkpatrick (infinite range) Ising model (Sherrington and Kirkpatrick 1975). This model was first introduced as an approximation to a spin glass system dominated by the long range RKKY interaction. It was hoped that the infinite range interaction would lead to an exact solution, although the solution in the spin glass phase has turned out to be more complex than originally hoped and is still under scrutiny (Parisi 1983). Nevertheless, this is one of the more tractable spin glass Hamiltonians and one can in fact obtain an analytic result for P(h) in its paramagnetic phase. There are some distinctive features associated with the small h behavior of P(h) and these will be elaborated on in Chapters 7 and 8. Finally, in Chapter 9 further studies of Ising systems are suggested as is the possibility of extending the P(h) formalism to non-Ising systems. 2. Transverse Susceptibility in One Dimensional Chains The one dimensional Ising model is attractive in that many exact results for the system may be obtained through straightfor- ward calculations. While this model does not exhibit any phase transitions for T>O, there are systems, such as CoC12°2NC5H5, which behave at least approximately like one dimensional Ising chains (Takeda et. al. 1971). In this chapter, a generalized method will be developed for calculating the transverse susceptibility of Ising chains by using modified transfer matrices. First, some background on the transfer matrix technique, de- veloped by Kramers and Wannier (1941), must be introduced. One starts with a Hamiltonian of the form N N _ _ A z z _ 1 z z z HN ' Jiil Sisi+1 2““ 1:1(Si+si+1) ’ (1) describing a one dimensional spin S Ising chain in the presence of an external parallel field, H = h2 2. Defining S§+1 = S? en- sures periodic boundary conditions. The partition function QN is given by QN = Triexp(-BHN>I (2) N z z z N z z _ - i - éz g2...§z exp{B[JiEISiS1+1 + zuh i§l(Si+Si+l)]} . (3) 1 2 N I - The notation g2 implies a sum over all (28+1) possible values of i S: and B is as usual 1/kBT. The (28+1)X(28+1) transfer matrix, T, is defined by its entries _ z z A z z z T 2 z — exp{B[JSiSi+l + guh (Si+S.+l)]} . (4) l SiSi+1 In terms of this matrix, QN = g ...z T T ...T T . (5) Z I S; S; 1 2 2 3 In going from equation (4) to (5), explicit use was made of the fact that all the operators in the Hamiltonian commute with each other. The operation in equation (5) is now recognized as matrix multiplication, so that N QN = :2 (T )st2 <6) 1 1 1 = TrlTNl <7) ZS+l V = 2 Ah , (8) 3:1 3 . where Aj is the jth eigenvalue of T. In the limit of infinite N, the maximum eigenvalue, AM, dominates: N QN = (AM) (II-m) . <9) Thermodynamic quantities may be obtained by the usual differentiation of RnQN. In the absence of an external field, for spin %, eXP(BJ/4) €XP(-BJ/4) T = (10) exp(-BJ/4) exp(BJ/4) with A1 ~2 cosh(BJ/4) (11) A2 2 sinh(BJ/4) so that for large N one recovers the result (Kramers and Wannier 1941), QnQN = N£n[2 cosh(BJ/4)] . (12) For spin 1, exp(BJ) 1 exp(-8J) T = 1 1 1 (13) exp(-BJ) 1 exp(BJ) giving A1 = cosh BJ + % + %q A2 = cosh BJ + % - %q (14) A3 = 2 sinh BJ where q = [(2 cosh BJ - 1)2 + 8]% . (15) For large N, anN = N£n[cosh BJ + % + %q] . (16) With this background, it is now possible to proceed with the calculation of the transverse susceptibility of an infinite, pure Ising chain. J. Phys. C: Solid State Phys. 16(1983) L237- 24-0. Printed in Great Britain LETTER TO THE EDITOR Transverse susceptibility in spin S Ising chainsi‘ M F Thorpe and M Thomsen: Department of Physics and Astronomy. Michigan State Universuty. East Lansing. MI 18824. USA Received 31 December 1982 Abstract. It is shown that the transverse susceptibility of a one-dimensmnal spin S Ising system can be calculated by combining linear response theory With the :ranster matrix technique. Closed form expressions for infinite chains are obtained for 5 = 5 and 5 = 1. and numencal results are displayed for S =- l to S = i. The perpendicular susceptibility of a spin é [sing chain has been calculated by several methods leg. Fisher 1963. Katsura 1962. and GreSI and Rajagopal 1974). However. it would appear that these methods have not yet been generalised to higher spin. The free energy and parallel susceptibility have been found for higher spins using transfer matrix techniques (Suzuki et al 1967). In this paper the spin dependence of the perpendicular susceptibility will be studied. We use standard linear response theory (see for example Grest and Rajagopal 1974) and apply it with transfer matrices to obtain exact closed form expressions for the zero field perpendicular susceptibility for spin 3: and spin 1 as well as numerical results for spin 3, 2. and 3. Our Hamiltonian has the general form H=HO+HI (1) and we wish to find the expectation value of an operator .-'1 to first order in H I. where HI... H I and A do not necessanly commute. It can be shown that If Trfexp(-I13HI))A] == 0 I2) then I” "l’ I -TrI I epr-(B - r)H..]HI~ expl - tH..; dr.~11 ' -O (1'1) = T d f 3 . Trlexm -I3Ho) l x to first order in H 1. In the Ising model for a closed cyclic chain we have .V Ho: -1 25555-. . (4) _V H1 = -_uh 215: I5) -‘ Work supported in part by the National SCIence Foundation. : Exxon Fellow. © 1983 The InSIitute of Physics L237 C8—C2 10 L238 Letter to the Editor where the S .- are quantum spins of magnitude 5 and site N + 1 is equivalent to site 1. From equation (3) we find .V r B ./ (Sf) = TILL exp[-(fi- rJHnluh 2:1 5'} expl—rH0) drSfJ/x/ Tr[exp(-/3H.,)]. (6) The tracing operation ensures that only the i = 1' term in the sum contributes. F urther- more, we n0te that all but two terms of Ho (which are -.IS;'_1.S‘,z -JS,’S,‘- 1) commute with 5}“ so that equation (6) can be rewritten as :x_. VV‘ 55-15311 3v _ , , TSf—sz-i Tgf- 57" T52-l5,:-2 . . . ngsg/‘Tr[exp(-l3Ho)] (7) l where 25,: stands for the sum over the 23 i- 1 possible values of S 5. We have defined the standard transfer matrix (Suzuki er al 1967) Tsfsf’i = exp(flJSES.-‘-l) (8). and have defined a new transfer matrix 13 T3: 5: I = (uh 2L exp[.l(fi- r)(5f-15f + Sfo-1)] X Sf exp[Jr(S,-‘-IS;' + Sfo-1)] drSf. (9) Interpreting the sums in equation (7) as a trace and cyclically permuting, we obtain (5;) = Tr{T”'2T‘]/Tr[T”]. (10) We now define the maximum eigenvalue of T and the corresponding eigenvector to be Am and lm) respectively, so that in the limit of N-> so, <.5f)=(mlT‘lm)/A§n. 111) Thus the zero field perpendicular susceptibility is u(m| T’lm) z. = .V‘—-———,—. . (12) h A?“ The (25 *- 1)2 elements of T‘ may be determined analytically from equation (9) by examining the 25 ~.- 1)3 states involved. However. since T§;_15;-, is a t'uncrion only of 55f-1 4-3,?.. 1 I, it can be shown that there are just 25 + 1 independent elements of T"r and hence only (25 + 1)2 states need be examined. A short computer algorithm was devel- oped to evaluate equation (9) analytically for larger spins. For 5 = l, we find it p ,8] sh T 7"“? ‘2‘ (1., x = .3 h _/:h_ 75inh/371 11 Letter to the Editor L239 while when S = 1 h sinh 2p] 2/1 sinh x3] ., J J ‘fih 2h ' 2 ' T‘ ___ smh l3] 2 1312 h Slnh l3! ( 14) I J 7 ' - 1 2&1 ..h 51:11 6] h Sinlli .31 Combining well known transfer matrix results with (12) and (13). F isher’s ( 1963) result for S = éis recovered: LL: 1' 3fl+____tanh(fiJ/4)‘l ‘ N112 i[,fisech 4 . j J (13) while for S = 1 we obtain it- _ (cosh l3] - l 4- iq) sinh 213] - 85inh (3] + ZBJq .V‘u‘: !q(cosh p1 + l + sq )2 “6) where q = [(2 cosh 51 - 1); 4- 3]“. (17) With Tand T" both known analytically. equation ( 11) may be evaluated numerically for larger spins. The results for S = i through 5 = 3 are displayed in figure 1. By evaluating equation (6) considering the two low lying states, all spins pointing up and all spins pointing up with the exception that Sf =5 - 1, it can be shown that? 1X a (T = 0) .v.u= =1 (18) independent of S. as is shown in figure 1. 0.7 0.6 Figure l. The perpendicular susceptibility (IL, .VuJ) pictted against reduced temperature (kgT/J) forS = i, 1. i, Zandi. 1‘ This result can be seen most easily by using ordinary second-order perturbation theory at T = i) to give _\E = 1V(uhl/2)2[(VIZS)Z/ZIS] = fix; hl. 12 L240 Letter to the Editor 0.1.0 1 L 1 l l l O 0.1 ).2 13.3 DJ. 3.3 1.6 3.7 Figure 2.. The perpendicular susceptibility (Jr -. .Vir') plotted against the reduced tempera- ture (kgT/[lSlS - l) I). This 15 the same as figure 1 except for the rescaling or the temoer~ attire axes. N0te that the vertical scale does nor begin at zero. Values or 5 are shown for each curve. The high-temperature limit ,‘(_ 5(5 1- l) _, = ___—.. (19 W 13 .3 ) suggests rescaling the temperature. The plot in figure 2 of x_ against kBTx’SlS ~ 1) indicates that in addition the location of the peak scales roughly with 5(5 -- 1) for S 2 1. The perpendicular susceptibility for an infinite spin é chain in the presence of a non-zero parallel field has already been calculated exactly by Grest and Rajagopal (1974). By including parallel field terms in the exponents in equations (8) and (9 ). similar results may be obtained numerically for larger spins. Finally. we note that some modi- fications after equation (7) will yield the perpendicular susceptibility for finite open Ising chains and hence for dilute Ising chains (Thomsen and Thorpe 1983). We are indebted to Dr S D Mahanti for several useful discussions. References Fisher M E 1963!. Math. Phys. 4124 Grest G S and Raiagopal 1974 J. Math. Phys. 15 589 Katsura S 1962 Phys. Rev. 127 1508 Suzuki M. Tsuliyama B and Katsura S 1967 J. Math. Phys. 8 124 Thomsen M and Thorpe M F 1983 to be published 13 J. Phys. C; Solid State Phys., 16 (1983) 4191-4198. Printed in Great Britain Dilute Ising chains with spin Si‘ M Thomsen: and M F Thorpe Department of Physics and Astronomy. Michigan State University. East Lansing. MI 18824, USA Received 17 February 1983 Abstract. Randomly diluted one-dimensional Ising chains of spin 5 are studied by breaking the systems down into non-interacting finite segments. Exact closed-form expressions are obtained for the specific heat and perpendicularsusceptibility of 5 = Echains and numerically exact results are given for spins up to it. 1. Introduction The problem of a dilute one-dimensional spin system with nearest-neighbour interac- tions is tractable becauseiit can be reduced to a collection of non-interacting finite segments. Matsubara and Katsura (1973) and Thorpe and Miyazima (1981) have exploited this fact to calculate thermodynamic properties for dilute XY quantum chains. Matsubara er al (1973) have investigated the specific heat and parallel susceptibility of dilute spin-i Ising chains whilst Matsubara and Yoshimura (1973) have investigated the same quantities for S > i. In this paper we use this technique to calculate the transverse susceptibility for spin-S dilute chains for the firsr time. This is an extension of previous work on pure spin-S chains by Thorpe and Thomsen (1983). The main problem to overcome is that the transverse field term in the Hamiltonian does not commute with the exchange part of the Hamiltonian. This can be handled as shown previously (Thorpe and Thomsen 1983) by introducing a new transfer matrix T". In addition we have to introduce another transfer matrix T“ in the dilute case to handle the situation when the field is on a spin at the end of a segment. Details of this calculation and results for 5 = i, 3 and it for concentrations c— = 0. i. %. fland 1 are given in § 3. In § 2 we calculate the specific heat using the transfer matrix method. This serves to introduce much of the notation although results have previously been obtained for the specific heat and parallel susceptibility for S > i by Matsubara and Yoshimura (1973). In neither of these calculations does one have to worry about non-commuting variables. For 5 = i it is well known that the specific heat per bond is independent of c. This is because each bond acts independently. We show that the specific heat per bond is weakly dependent on concentration for 5 2 1. * Work supported in part by the National Science Foundation. 2: Exxon Fellow. © 1983 The Institute of Physics 4191 14 4192 M Thomsen and M F Thorpe 2. Specific heat The transfer matrix technique was introduced by Kramers and Wannier (1941) to solve infinite and finite closed chains in the Ising model. One starts with an Ising Hamiltonian with a field in the z direction: =-leSi‘55-1-,uh 215.; (1) where if S..1 a S the chain is closed. The partition function is Q: = 1315le = Tr (D; exp{fl[‘,5izsiz*l " iuhle *' 55-1)”) (2 If one defines a (25 + 1) X (25 + 1) transfer matrix Tby 73:5!“ = eXPifilJSi’Si’q '1' iii/“Si: ‘1’ 55.1)” 13) where the 5,? refer to eigenvalues. nor operators. then ('2) becomes Q: = 2 E . . . Z (SilT’5§f>- . . cars.» .4) 5g 5g 5%, = Tr[T']. (5‘) Taking it,- to be the jth eigenvalue of T. 1.5.1 Q: = 2. A;- <6) ,. To modify this for an open chain. we note that we need only set the bond energy of the 5fo pair to zero. In analogy to the derivation of (5) we find for the open chain Q? = Tr[T" l7"] (7) with T§fsf = T55;(J = 1)) '3 CXPHBUJHSE "’ $13)]. {8‘ The partition function ('7) is generally evaluated by rorating T and T' into the represen- tation which diagonalises T. In this section we wish to find the zero field specific heat of a finite open segment. NOting that in zero field, T533 =1 for all 55 and S 1’ , and defining M as the matrix that diagonalises T. it is easy to show that ZSrI 234-1 3 o... ' ‘-li ‘ Qr - S A: '2 -Will- ‘9) 1']. i [‘1 For 5 = i, this yields Q? = 2(2 cosh ilil)" ‘. ('10) while for S = 1 we find 3-'" .- it‘-3 .- Q9=ZC _A._lA)'*7'f:-T:(A )' (11) A. A. A. 15 Dilute Ising chains with spin 5 4193 where = = coshfiJ i-l:[(cosh13’J—l)2 + 2]“? (12) Using 1 _ a 2 a ) hie-”(r H In Q. (13) it follows that in the S = 1 case Cr/ke = (r - 1) (Bl/4)z sechzffll/4) (14) (see also Stanley (1971)), while application of (13) to (11) yields an expression too complex to be of much use analytically. For general spin. the matrix calculations involved in (9) are performed by computer at three points close to each other so that the differentiation in (13) can be done numerically. 4 1 41 4,4 714 //’°/°/// ° °//// ° W —-I M W r=2 r=l r=3 r=4 Figure l. A dilute spin-S chain with segments of length r = 1. Z, 3 and 4. In a dilute chain of length N (N -> an) and concentration c, the number of seg- ments‘containing exactly r consecutive spins (as shown in figure 1) is P, = N(1 - C)ZC’. (15) Since C=2ec do we find that the specific heat per bond in the spin-l case is C/Nkac‘ = (191/4) seen: (131/4). (17) which agrees with Matsubara er al (1973). We note that the specific heat per bond is independent of the concentration c because the spin-i Ising chain can be mapped onto an equivalent system of independent bonds. This is done by defining a new variable r,- = 5,5..1. Because of the absence of loops the t.- can take on values 1'.- = t i and so in the absence of a field h, the Hamiltonian ( 1) splits up into non-interacting Hamiltonians each associated with a bond with a spin at each end. For 5 > i, the specific heat per bond is notindependent of c, as can be seen in figure 2. However, the specific heat per band depends only weakly on the concentration and this dependence does not become greater as 5 increases. A useful check is provided by noting that (”gnaw-5(a) <18) 6 where 5( 00) is the entrOpy of the system at infinite temperature and 5(6) is the entropy 16 4194 M Thomsen and M F Thorpe 10+ 10 . -. r, S a .55." ‘K S a 1 i If: a"; “1% 05 ) 05 3,),- “K B n I ..“ o 10 20 o 10 20 so i T 7-01 - 1 2a - - 1. 15> 5:: . 1_g- 5:: -. z : u. 2 ~15 :5: 1:15. 10» If! l as 1‘ ix \ ‘ A A —__.Jl 1.0 7.0 310 o 1.0 7.0 so kJ/JS’ Figure 2. The specific heat per bond (C/Nc‘k.) as a function of reduced temperature (kg 1713’) for various values of spin 5 as indicated and concentration c - 0 ( ----- ), l (- - - - -),§(-—--),i(-—--), 1( ). ForS - lthespecificheatperbondisindependent ofcasshown. Thee=011mitcorrespondstotwospinsjoinedbyonebondandotherwise isolated. at a small positive temperature 6. Since the total number of states is (2.9 + 1) m“*‘" = (23 + 1)”‘ S( on) = Nkac ln(ZS + 1). (19) On the other hand, the degeneracy of the ground state (in zero field) is 2”“(25 +1)”1 where N1 is the number of isolated spins and N2 is the number of segments of length greater than one. This gives 3(6) = Nk3(1 - c)c2 1n 2 + Nk3(1 - c)=c 1n(2.s + 1) _ (20) and hence [:ng = Nkaczflz - c) ln(2.S' + l) - (1 - c) In 2]. (21) The low-temperature behaviour of the specific heat is dominated by the lowest excited states. For S = 1, this corresponds to flipping all the spins simultaneously in any part of that segment, provided that part includes an end spin, as shown in figure 3. For S 2 1, the low-lying excitations come only from reducing by one unit the 2 component of an end spin which is also shown in figure 31. It can thus be shown that at low 1‘ For spin 1 , additional degeneracy arises for the lowest excited states in segments of length 2. However, it is still true that all the lowest-lying excitations are governed by end segment spins for S - 1. 17 Dilute [sing chains with spin S 4195 temperatures. C/ngcz == ([315): e“"’5 S = .1 (22a) C/ngc2 z [3 - lc](1 - c)(,BJ.S)2 e‘w S‘= 1 (22b) C/ngcz 2' 2(1 - c)(l3JS)2 e‘ws S 2 '2‘ 22C) llloolllllllloo ll°°\llllllll°o Figure .3. The lowesr excitations for 5 = 5 and S 2 1 dilute chains as described in the text. When c-> 1, the right-hand sides of equations (22b) and 225) vanish since end-of- segment spin excitations are no longer possible. which can be seen in figure 2 where the c = 1 specific heats go to zero much more rapidly as T-> 0 for S 2 1. This is because there is an extra factor 2 in the exponent of the next terms in (22b) and (2c) reflecting the fact that every spin has two neighbours rather than one. We conclude this section by noting that when one evaluates the high-temperature expansion of Q = Trfe ”3”] in powers of )6. it is found that C/Nk3c3 2 [15(5 + 1))3112 (23) which is independent of c for all S. 3. Perpendicular susceptibility Calculation of the perpendicular susceptibility is complicated by the fact that parts of the Hamiltonian do not commute with each Other. In the absence of any parallel field r-ol r H=-J,§.‘lstss-.-,Lm‘ZStaHo—H, 124) ta :-1 for an open segment of length r. Using linear response theory. Grest and Rajagopal ( 1974) calculated the perpen- dicular susceptibility for a closed spin-Ii chain. With transfer matrices, this was gener- alised to higher spins (Thorpe and Thomsen 1983). It was shown that the contribution of each spin to the magnetisation is r I -5H 1‘s X 7.51-351-) T span. 7.55.1554 . - - Tsist/Tl’ie ‘ ”l 14-5) 18 4196 M Thomsen and M F Thorpe where T53),l is as defined in (3) and 3 first. = .qu 2f [0 expmr - r)(57-15,~‘ + s;s;..>is; x expws;., s; + s; 51.1)] dTSf. (26) UsingSf 1(5 * + S,‘ ), the matrix T’ can be evaluated analytically for any 5. In par- ticular, when S= i ((W/J) sinh llil With" “3“”! (WI/051““ “31 . We now open the chain by setting 1 = O for the SSS? term. If S,- is nor at the end of a segment, then (25) becomes (SX)=SE§§.. 2 2 ”@7333 T593... 3-1 Sin (27) X 7.57.2.3“ Tx5f_l3f.l Tng 55.2 . . . T3145; T'gsg/TI'R -dHo] (28) where T335? is defined in equation (8). In the case of an end spin, (51) = (55) = 35.; 2:. - - - $2 Tstngsgsg- - - 755-33-. Trams/[THC "W1 .29) 1 2 ’1 with the new transfer matrix T" defined as p 11's.“.1 = 52 we I exp[(fl - r)Js*-ls:]5: exp[fiS‘_( 5:] dist. (30) 0 We n0te that the subscript Sf appears for formal reasons only—7‘55” 5g is. independent of Si. For 5 = l, we obtain = (OW/J) Sinh W (2147!) sinh is!) (ZW/Dsinhffl] (zmx/Dsinhffi] ' Using the above equations. we find the zero-field perpendicular susceptibility of a finite open Ising chain to be (31) x”: ;[(2'1'1'[ richer-1' 1T1) +2TrT"’T’]/TIT' ‘7‘]. (32) This expression is valid for r 2 2. For r = 1, one uses the free-spin Curie susceptibility. With all four matrices in equation (32) known analytically, it can be shown that for S = 1 X1! = [(r - 2)/2](/3/4) Sit—ch2 $131 + [(r + Zl/Z/l tanh H31- (33) Performing a sum similar to equation (16). but singling out the special case of r = 1, we find that for a dilute spin-l Ising chain Jib/Neill = (1 - 021137 + 1181 c2 sech‘llil + 1(4c - 3c?) tanhltil. (34) Setting c = 1 reproduces Fisher’s (1963) result for a spin-l pure Ising chain JxJ/Nis2 = 1,3! sech2 llil + l tanh 1131. (35) 19 Dilute [sing chains with spin S 4197 As c approaches 0 only the first term in equation (34) survives so that JxJNuzc e W as expected for No free spins. All four matrices in equation (32) are known analytically for larger 5. but the final evaluation of that expression must in general be performed numerically. The results are displayed in figure 4 where we have plotted the inverse susceptibility per spin as a ”Xi/Hip“.| ‘2 C.) 2 ,) 1 £0 .10 i 2.0 1 k3 I'USJ Figure 4. The inverse perpendicular susceptibility per spin [ILA/ctr)“ as a funcn’on of reduced temperature {kg 1715"] for various values of spin 5 as indicated and concentranon C’0( """" Li‘l---'°).i('~'-).i(---).ll l.Thec=0limitcorrespondsroa single isolated spin. function of temperature for four different spins: S = is, 1. 5%, 3. We note that 1:1 approaches zero linearly for c < l, reflecting the effect of .Vcil - c)z isolated spins [X1 = iNuzd 1 - c)25(S «'- 1)/k3 T]. For c = 1, we recover the result for the pure chain that x; = lN/f/J. At high temperatures we obtain the Curie susceptibility of NC isolated spins [1 g = lNuch(S 4- l)/k3 T] so that [J x J’Nc‘tfl‘l becomes independent of c. 4. Conclusion The perpendicular susceptibility per spin and the specific heat per bond do nor change dramatically upon dilution over much of the temperature range studied. The leading 20 4198 M Thomsen and M F Thorpe term in the high-temperature expansion of each of these quantities is independent of concentration. For the special case of spin l, the specific heat per bond is independent of concentration for all temperatures. The most n0ticeable effect of diluting an Ising chain is in the way these thermodynamic quantities behave in the zero-temperature limit. The specific heat per bond differs in the nature of its exponential approach to zero upon dilution (for S 2 1). Most striking is the divergence of X- at T = 0 when the chain is diluted, owing to the finite probability of finding isolated free spins. We note that the results for both the specific heat and the perpendicular susceptibility depend only on Ill and so are identical for an antiferromagnetic Ising chain. This is most easily seen by performing a rotation of 180° about the x axis. One can expand on this study by turning on a weak interaction between segments, particularly those separated by just one non-magnetic site. However. the independent-segment assumption and hence equation ( 16) are not valid in this model and the problem becomes significantly more complex. Acknowledgment The authors wish to thank S D Mahanti and T A Kaplan for several valuable discussions. References Fisher M E 1963 J. Math. Phys. 4 124 Grest G S and Rajagopal A K 1974 J. Math. Phys. 15 589 Kramers-H A and Wannier G H 1941 Phys. Rev. 60 252 Matsubara F and Katsura S 1973 Prog. Theor. Phys. 49 367 Matsubara F and Yoshimura K 1973 Prog. Theor. Phys. 50 1824 Matsubara F. Yoshimura K and Katsura S 1973 Can. J. Phys. 51 1053 Stanley H E 1971 lnrroducrion to Phase Transitions and Critical Phenomena (Oxford: Oxford University Press) Thorpe M F and Miyazima S 1981 Phys. Rev. B 24 6686 Thorpe M F and Thomsen M 1983 J. Phys. C. Sou'd State Phys. 16 2237 21 Much of the preceding analysis can be carried out in the presence of an external parallel field, hz. For instance, using equation (4) the standard transfer matrix for spin S is determined in a parallel field. From this one can obtain the partition function, although for S>%, the transfer matrix is not easily diagonalized analytically. Similarly, one can write down the elements of the TX transfer matrices, introduced in the X1 calculation, in the presence of a parallel field: r3 X X 3-. .- _ X “X T32 Sz - h g2 JO expliB r)T1] Si exp(rfl) 51 dr (17) i—l 1+1 1 TX, - n‘ z (a x [(B-t)F 1 3X ex (If ) sX dT (18) 2 S2 C Sz 0 e p 2 r .p 2 r r—1 1 r where Z Z Z Z Z Z Z F1 — JSi(Si_l + Si+l) + h [(Si-l + Si+l)/2 + Si] (19) Z Z Z Z Z Z r2 = JSr_1Sr + h [(sr_1 + 51)/2 + sr] . (20) For instance, for S=%, exp(BhZ/2)sinh[%g(hz+d)] sinh(BhZ/2) J+hz h2 .l; TX = (21) h sinh(BhZ/2) exp(-th/2)sinh[8(J-hz)/2] hz J-hz and 2exp(3hz/2)sinh[8(%J+%hz)] ZSinhLB(%J+%hz)1 J+2hz J+2hz l__Tx'= (22) hz 23inh[8(%J+%hZ)] 2exp(-th/2)sinh[8(%J-%hz)] J+2hz J-th 22 To calculate the transverse susceptibility for a pure chain, one needs to calculate the maximum eigenvalue, Am, and corresponding . x . . eigenvector lm> of T . Then by equation (12) of the preceding letter, "Transverse susceptibility in spin S Ising chains," one can obtain Xl exp(th/2)sinh[B(J+hz)/2] + exp(-th/2)sinh[B(J-hz)/2l = Z Z Nui 2(J+h ) 2(J-h ) + eprBJ/Ajsinh(BhZ/2) eprBhZ/2)sinh[B(J+hz)/2] 2q1 J+hz _ expg-BhZ/2)sinh[B(J-hz)/21 + 2expg-BJ/2) ] J-hz hz J X [exp(BJ/4)cosn(BhZ/2) + qll-Z (23) where q1 = [exp(BJ/2)cosh2(BhZ/2) - 23inh(BJ/2)] % . (24) This result was obtained earlier by Crest and Rajagopal (1974). The modified transfer matrix, TX, can be obtained straight- forwardly for higher spins in a parallel field, but analytic cal- culation of Xl is no longer tractable. Nevertheless, numerical results may be obtained easily. Inspection of equation (23) reveals that for hz=iJ, there is the possibility of a divergence due to vanishing denominators. In fact, an antiferromagnetic chain in this field will have a divergent zero temperature transverse susceptibility, as noted by Crest and Rajagopal (1974). This can be seen for general spin by considering the Hamiltonian associated with a single bond: 23 22 Z Z 2 H12 = Josls2 - h (31+32)/2 (25) with JO 2 —J > 0 . (26) Taking hz = 2JOS, Z Z Z H12 - (Jos2 - J05)s1 - Joss2 . (27) H12 can be minimized with respect to S? by setting S§=S, yielding a ground state energy E12 = -JOS“ , (28) independent of 8:. As long as one spin in every bond is parallel to the field, the system is in a ground state. This state leaves a large number of spins free to point in any direction, which can be seen by noting that E12 is independent of S These free spins are then able to z 2. line up in the direction of an infinitessimal transverse field, hence creating the divergence. In particular, from equation (23), one can show that at H=iJ and as T + O, xi/Nui = s(1+/§)/(10+6/§) . (29) The dilute chain transverse susceptibility in the presence of a parallel field may also be obtained using the formalism developed in the preceding paper. It should be noted that now the contribu— tion of each spin in an open segment is a function of its location in the segment. This is a direct consequence of the fact that T BHd T, are not simultaneously diagonaliéable in the presence of a field. Thus for dilute chains, analytic work stops at the writing 24 ? x X, down of the four matrices T, T , T , T , and the matrix multiplica- tion and diagonalization is performed numerically. It was seen earlier that the isolated spins contributed a divergence at T=O to Xl' Application of a finite parallel field destroys this divergence. However, for the same reasons outlined for the pure chains, an antiferromagnetic chain will have another T=O divergence in Xl when hz=32JS. A third divergence occurs at hz=.JS due to the spins at the end of open segments. Figure 1 il- lustrates one of these divergences, at hz=JS, with a plot of xll versus temperature. Since the divergence is proportional to 3, x11 approaches zero linearly, as shown. One limitation of this formalism that has been developed for the calculation of XI is that it is based on linear response theory and hence is valid only for infinitessimal hx. However, in exchange for this restriction, one has gained the ability to calculate Xl for a variety of models that are not tractable by other methods. 3.5 2.5 iJIXL-l 2 Ncu2 0.5 Figure l. 25 DIOT .’ .4 d ."’ '.". a, -.' 0’ 0.. 0’ " a’ 0' 'I '0‘ d I./ x” O" " a’.’ ..o" J “-—'-"'—. .0 """ 'l"- oooo '- -4 ,1"- “‘----------"".." ‘ ‘—‘H/Ulfl15 _/’ ----. r1/ 1,: 3:0 .4 - --- HI (.1 [20.6 I I I I I 0 0.5 l 1.5 2 2.5 kBT/IJISZ The inverse of the transverse susceptibility per spin of a dilute (c=O.25) Ising chain in the presence of three different parallel fields. 3. P(h): The Local Field Probability Distribution Function The local field probability distribution function, P(h), has been studied at zero temperature for a variety of spin glass models (see, e.g., Palmer and Pond 1979 or Walker and Walstedt 1980). One of the distinctive features of P(h) these studies bring out is a dip at h=O. The nature of P(h), measuring the distribution of effective fields acting on a single spin, suggests that it be relegated to mean field theory or to low temperature studies. It is probably for this reason that the temperature dependence of the distribution has not been studied much before this. It turns out, however, that for Ising systems one may obtain exact thermo- dynamic relations using P(h). This chapter is devoted to a formal discussion of the properties of P(h). While Appendix A contains details for a very broad class of Ising systems, for simplicity a more restrictive Hamiltonian is used here: H = —% Z J..o.o. (1) i,j 13 l J where the 01's (=il) are spin % Ising variables. The matrix de- fined by the Jij's is symmetric with vanishing diagonal elements but is otherwise arbitrary. One begins by defining a local field operator, hi: h- = Z J..o, 1 j 13 J . (2) The missing factor of % in equation (2) ensures that hioi contains 26 27 all the Oi terms in the original Hamiltonian. The local field probability distribution function at site i is defined by Pi(h) = <5(h-hi)> . (3) The notation <...> refers to a thermal average. The site averaged distribution for a system of N spins is P(h) = tPi(h) . (4) l L Alr—l This function is of course normalized: f P(h) dh i 1 . (5) Associated with P(h) are the following three exact results: M = N f tanh(Bh) P(h) dh ' (6) E = ?§ I h tanh(Bh) P(h) db (7) S(k,w) = %[P(w/2) + P(-w/2)]/[1 + exp(-Bw)] . (8) These last three equations have simple physical interpretations. The magnetization of an isolated spin in a field h is tanh(Bh). Equation (6) performs a weighted average with respect to h. Since tanh(Bh) is odd, M couples to the antisymmetric part of P(h): Pa(h). The second equation has a similar interpretation with an additional factor of % present to correct for double counting of bonds. E couples to Ps(h), the symmetric part of P(h). Finally, a beam of neutrons polarized in a direction perpendicular to the z axis interacts with the system by flipping individual spins at an energy cost proportional to the local field. Since these neutrons cannot distinguish between +2 and -z, they couple to Ps(h). The statement these equations make is significant. Without any explicit knowledge of the Jij's, one.can obtain the thermodynamics 28 of a system directly from P(h). Furthermore, one can measure Ps(h) directly, and that is sufficient to determine E. Detailed proofs of equations (6) - (8) appear in Appendix A. However, to give the reader a flavor for the proofs, the energy equation, (7), is derived here. One begins with the energy of an isolated spin in a field h: €(h) = —h tanh(Bh) . (9) In its more primitive form, this is €(h) = TrOIeXp(-BHO(h)) H0(h)I/TrOIeXP(-BHO(h))I (10) where H0(h) = -hoO and TrOmeans trace over 00 = :1. Denoting the energy of all the bonds terminating at site i by 2Ei (E1 is then the energy associated with site i), it needs to be shown that f €(h) P(h) db = 21?.1 . (11) The following definitions need to be introduced: Trl = trace over Oi = t1 Tr' = trace over all spins except Oi H. = -h.O. 1 1 1 H' = H — Hi’ which is independent of 01 Z = Tr[exp(-BH)] Then, I P(h) €(h) dh I h Tr[exp(_BH)5(h-hi)]Tr0[exp(-BHO(h)) H0(h)] (12) d Z Tr0[exp(-BHO(h))I 29 exp(—BH) Tr0[exp(-BHO(hi)) HO(hi)] 1 (13) = -Tr Z Troiexp(-BHO(hi>)1 =-l-Tr' exp(-BH') Tri[exp(-BHi)] Tr0[exP(-BHO(hi))HO(hi)] (14) z Tr0[exp(-BHO(hi))I In the last step, it is crucial that [H',Hi] = 0 so that the expo— nential can be split up. Since Tri[exp<-eHi>1 = Trotexp<-BHO>1 , (15) one can cancel a term from the denominator and the numerator. Now 00 can be relabelled Oi and the traces can be regrouped, giving I P(h) €(h) dh %-Tr[exp(-BH) Hi] (16) 2Ei , (17) which was to be shown. P(h) is fundamentally a lggal quantity coupling to local pro- ~perties such as Bi and Hi. The specific heat and other quantities which involve long range correlations require differentiation of P(h). The free energy, meanwhile, is obtained by integration of E in equation (7). Thus P(h) contains not only all the information found in the free energy, but it also contains additional information such as S(k,m). This formalism does not of course provide short cuts to solving for the thermodynamics of a system. The price one pays for the additional information in P(h) is that it in general is harder to calculate than the free energy. The remainder of this chapter will be devoted to methods for calculating P(h). 30 This part of the discussion will be limited to ferromagnetic systems with nearest neighbor interactions (although some of the techniques will be generalizable). Setting the interaction parameter J = l, the Hamiltonian becomes H = - X 0.0. (18) 1 3 where X indicates a sum over nearest neighbor pairs and the spins are assumed to lie on a lattice. One of the advantages of this simplification is that P1(h) = P(h). The first approach to calculating the distribution is a formal one starting from the definition, Pi(h) = <5(h—hi)> . (19) An integral representation of the 6 function is used, and a little algebra yields the form 2 Pi(h) = E- wS 6(h—s) . (20) s z z is the lattice coordination and the sum is over the allowed values of the local field, -z, -z+2, ..., 2—2, 2. Details of this approach are in Appendix A. It is shown there that the weights ws involve numerical factors, which depend only on coordination number (not on the lattice itself), and all possible correlations among the nearest neighbors of site 1. Hence the real work involved with this calculation becomes the determination of those correlation functions. This method will be demonstrated in Chapter 5. For a system of coordination of two or three the correlations on the nearest neighbor shell can be eliminated in favor of the 31 nearest neighbor correlation, and the magnetization, . If, for example, z=3, then the possible field values are :3 and :1 and so four equations are needed to determine the four weights, ws. These equations are obtained by using the normalization requirement, (5), the energy and magnetization relations, (7) and (6), and the fact that = 3. One then obtains 0 (W3 + w-3) + (w1 + w_1) = 1 (21) (w3 + w_3)3tanh(3B) + (w1 + w_1)tanh(B) = 3 (22) (w3 - w_3)tanh(38) + (w1 - w_1)tanh(8) = (23) 3(w3 — w_3) + 1(w1 - w_l) = 3 . (24) The first two equations yield Ps(h) and the last two, Pa(h). These equations can be solved explicitly to yield 1 3(tanh(3B) - (0001)) 3(tanh(3B) - 1) w = -- i (25) :1 2 3tanh(3B) - tanh(B) tanh(38) - 3tanh(B) 1 3<0001> - tanh(B) (l-3tanh(B)) w = -' i" (26) :3 2 _3tanh(38) - tanh(B) tanh(3B) - 3tanh(B) ’ showing explicitly in this case that the w's depend only on (0001) and . No reference is made here to the particular network other than it has coordination three and is compatible with a Hamiltonian of the form given in (18). Finally, some systems lend themselves to calculation of the moments by standard thermodynamic methods. If one knows all the moments up to , then the 2+1 equations, = Z w s (27) 32 can be inverted to find the ws directly. These methods may be generalized to sgmg other systems, such as antiferromagnets, but their generalization to disordered systems is nontrivial. Different and specialized techniques must be de- veloped for calculation of P(h) on these systems and so discussion of such techniques will be postponed until a specific disordered system is introduced (Chapter 7). 4. P(h) for Bethe Lattices A simple system for which P(h) can be determined consists of S=% Ising spins on a Bethe lattice with constant nearest neigh- bor ferromagnetic interactions. The Bethe "lattice" is really a pseudolattice (see Figure 2). It has the unfortunate property that, even in the thermodynamic limit, a finite fraction of its sites lie on the surface. On the other hand, it has the advan- tage that its coordination number, 2, is easily varied. The z=2 case, a linear chain, will not be considered here since many of its properties are distinctly different from the z>2 Bethe lattices. Because of the large surface area, calculations must be ap- proached with caution. For instance, Eggarter (1973) pointed out that by assuming free boundaries, the Bethe lattice does not exhibit a phase transition. However, the more traditional Bethe- Peierls approach (Bethe 1935, Peierls 1936) which assumes boundary conditions determined by self-consistent fields, does find a phase transition. This latter approach will be used in this discussion. The Bethe-Peierls method, leading to the self-consistent field equation (8) below, is described in Domb (1960) and is shown to be an exact solution for the system. An effective Hamil- tonian is assumed for a cluster consisting of site 0 and its 2 nearest neighbors: Z Z H = —J 2 o o. — H o - H 2 o. . (1) z j=1 0 3 O O lj=1 J 33 34 > <2 2 é ’ Figure 2. Section of a Bethe lattice for z=3(upper) and z=4(lower). 35 H0 is an external field acting on the system and will be set to zero later in the calculation. H1 is an effective, temperature dependent field acting on the Oj's accounting for both HO and the z-l neighbors (other than 00) of Oj. It can be shown that the partition function for H is z -%,—-1—-% 222 22-. 2122 Q2 = A0 ( 1 x + Alx ) + AO (XIX + Al x ) (2) where 10 = exp(-2H0/kBT) (3) 11 = exp(—2H1/kBT) (4) x = exp(-2J/kBT) . (5) To obtain H1, one requires <0 > = <0.) 0 J ’ or equivalently, A 3 1 8 A0 575 £an = z— Efi £qu - (6) Using equation (2), some algebra yields 11/10 = [(11 + x)/(1 + 111012“1 . (7) The external field HO will now be set to zero, having served its purpose. Some rearrangement gives tanh[H1/(z-1)kBT] tanh(H1/kBT) = tanh(J/kBT) , (8) and this specifies H1 as a function of temperature. An expansion for small H1 gives a transition temperature defined by tanh(J/kBTC) = 1/(2-1) . (9) 36 Above Tc’ H = O, and the effective partition function becomes 1 Q2 = 2[cosh(J/kBT)]z , (10) corresponding to 2 independent bonds. This result is typical for spin % Ising systems with no loops, in the paramagnetic phase (see, e.g., Thorpe 1982). It should be noted that Qz determines only local quantities such as the energy associated with 00 or the local magnetization . It is 22; true that the free energy F is given by -BF = Zan. Below TC, one needs to determine H1. Introducing two new variables Y = All/(2‘1) (11) rt ll tanh(J/kBT) , (12) it is now possible to rewrite the self-consistent equation (8) as -1 l-Y 1+1z —'—:-=t <13) 1+Y 1_Yz 1 01' —1 2'2 i (l-t)Yz - 2t 2 Y + (1-c) = o . (14) i=1 By symmetry (H1 ++ -H1), one expects that if YO is a solution to this equation, Yal is also a solution. This fact can be exploited to reduce the degree of the polynomial by a factor of two. Since a fourth degree polynomial is solvable (Burington 1946), one can in principle obtain analytic expressions for H1 up to z = 10, In practice, the complexity of these solutions makes some of the 37 larger 2 results not worth pursuing analytically. For completeness sake, analytic expressions giving X1 (=exp(-2H1/kBT)) as a function of x (=exp(-2J/kBT)) are shown below for z=3 up to 2:6: 1 z=3 A1 = %[x - 2x_1 — 1 + (xFZ-x_1)(l-2x-3x2)2] (15) - .102] <16) 2: A = %[x 1 1 2:5 Define C = [1 - x + (5x2+2x+l)2]/2x Then 11 = {[c + (c2-4fi1/2}4 (17) 2:6 Define D = [x_1 + (x-2+4)%]/2 Then A1 = {[D + (DZ-4)%]/2}5 . (18) A system with infinite range interactions produces a mean field result (Stanley 1971). On the Bethe lattice this corresponds to the z +'m limit, which is readily accessible. In order to keep the energy finite, this limit must be taken holding Jz con- stant. Since Hl should sCale with Jz, l L xii xi = 1 2 , (19) NIH ,__. 1+ reducing the partition function (2) to 2 Q2 = 2cosh(HO/kBT)[2cosh(H1/kBT)] . (20) The self-consistent field equation, (8), becomes Hl/Jz = tanh(H1/kBT) . (21) Which has a nontrivial solution if and only if kBT < Jz. Hence kBTC = Jz . (22) With the fundamental equations in place, it is now possible to calculate P(h) for Bethe lattices. This is most easily done 38 by calculating the moments of the distribution. Inspection of the Hamiltonian (1) reveals that these moments may be obtained by differentiation of Q2. Taking h/J = Z 0. , ' (23) - J J (assuming zero external field), one finds a n <(h/J)n> = —' [-ZA T] Q . (24) D—4 Q) As pointed out in the previous chapter, given the moments up to <(h/J)z>, one can solve for the weights in the 2+1 6 functions associated with P(h). This process, while straightforward, yields very complex analytic expressions which by themselves do not provide much insight. For this reason, they are omitted here in favor of the numerical results below. The mean field result is easily derived. iStarting from the partition function, (20), and applying a modified form of equation (24), one obtains for H = O, 0 <(h/J)n> = (2cosh8H1)‘z [SBHI n(2coshBH1)z (25) = z(z-1)...(z-n-1)tanh“BH1 + 0(zn‘1) (26) Thus <(h/Jz)“> = tanhnBHl + 0(1/2) (27) This implies the 6 function distribution P(h) = 5(h - JztanhBHl) . (28) Using equation (21), this can be rewritten P(h) = 6(h — H1) , (29) 39 which is another way of saying that in the large 2 limit, mean field theory becomes exact. In Figure 3, P(h) is plotted against h/z for z = 3, 4, 6, and infinity. It is symmetric for T 2 Tc’ reflecting the fact that the magnetization is zero. At infinite temperature, the weights are determined by counting the number of ways to make a particular value of h. Since there are more ways to make h = O, the distri- bution is peaked there. When T = TC , the distribution is still peaked at small h, with more weight being at h = O as the mean field limit is approached. The symmetry is broken below TC when spontaneous magnetization sets in. Again the approach to mean field theory is seen as the maximum in P(h) moves from h = Jz (z = 3) to h = sz for larger 2. It is apparent that there is strong dependence of P(h) on coordination number for Bethe lattices. While the number of 5 functions is clearly determined by z, the profile of the distri- bution, particularly at Tc’ is also sensitive to 2. It is interesting to note that the nearest neighbor correlation at TC is also strongly 2 dependent: T=TC = 1/(2—1) . (30) It will be argued later that this correlation plays an important role in the profile of P(h). 4O .8 U53 ’3 6 8n H\B N. o An E\H .0 .q .m :ofiumcfivuooo mo mmoflsumfi onumm com ALVA .m ouzwflm N\: o e.gn %\E o.ouue\e o 0.0" E\H 1 «.3 A 3.3 «.9 ...,. . 3.: sum N \0 ll Aces 5. P(h) for Two Dimensional Ferromagnets There are several two dimensional Ising systems for which P(h) can be calculated exactly. These analytic results can provide useful insight into systems with phase transitions not dominated by surface effects (in contrast to the Bethe lattice). The first two dimensional Ising system to be solved was the square net with uniform nearest neighbor interactions (Onsager 1944). Taking the ferromagnetic case and setting J = 1 yields a Hamiltonian of the form H=~Z 0.0. . (1) l 3 Since the square net has coordination four, the possible values of h are :4, t2, and O. The weights of the corresponding 5 func- tions in P(h) are easily expressed in terms of correlation functions. using equations (38) and (43) and Table 1 of Appendix A. The nearest neighbors of an arbitrary site, 0, have been numbered cyclically. One finds: ”:4 = “1'6 1 I“? + 2150102> + ém‘lcb> i-%<010203> + IE<01°20304> (2) “:2 = T * 2‘01) 1 %<"1"203> ' 2‘0102’304) ° (3) W0 = g ‘ 23102> " F010? + % ' (4) Apparently. <010203> is not in the literature, nor is it easily calculable. This problem is dealt with by using the magnetization 41 42 equation ( (6) in Chapter 3) to eliminate <010203> in favor of <00), The even spin correlations can be reduced to elliptic integrals using the Pfaffian method described by Montroll, Potts and Ward (1963). While all of the two spin correlations needed appear ex- plicitly in that paper, the four spin correlation does not. Details concerning the calculation of the latter in terms of elliptic integrals are given in Appendix B. The final step, evaluation of the integrals, must be performed numerically. In Figure 4, excerpted from the preprint in Appendix A, P(h) is shown at five different temperatures. The infinite temperature distribution, as for the Bethe lattices, has roughly a Gaussian profile reflecting the fact that there are more h a 0 states than h = 14 states. However, in contrast to the Bethe lattice, a pro- nounced dip has developed at Tc' This dip will be discussed in more detail later in this chapter. Below Tc’ the asymmetry in P(h) develops very quickly, reflecting the rapid rise of the magnetization. Results for the honeycomb lattice are more easily obtained. Since the lattice sites are threefold coordinated, equations (25) and (26) of Chapter 3 apply and one needs just the magnetization, , and the nearest neighbor correlation, <0001>. The former has been calculated by Naya (1954), while the latter may be obtained by differentiation of the partition function in Green and Hurst (1964). The expression for <0001> thus obtained involves a double integral, making it somewhat awkward to evaluate numerically. 43 13/1}:=(313 0.5- c) ,_ -F/-EC= C)f3 0.5 O 1 A P(h) T/TC=I.O 0.5- V C) E3 E3 E3 El a; T/Tc= 1.2 0.5‘ 0 12 T/Tc=00 0.5-. ‘4 '2 O 2 4 h Figure 4. P(h) for the ferromagnetic square net. 44 Analytic work reduced this form to a single integral, details of which appear in Appendix B. A plot of P(h) for the honeycomb lattice in Figure 5, excerpted from Appendix A, reveals that it shares many of the features found in the square net, including having a dip at Tc' One can also investigate P(h) for a triangular net. A problem arises however since for a sixfold coordinated lattice, many more correlations are needed. While the even spin correlations can be obtained (Choy 1984), not all of the odd spin correlations can be. Hence the results in Figure 6, also excerpted from Appendix A, are incomplete. Nevertheless, one can clearly see similarities between these results and the earlier two. In particular, there is once again a pronounced dip at Tc' The existence of the h = O dip means that at TC most spins find themselves in the presence of a nonzero local field. This suggests that perhaps small domains form in the system before long range order sets in. The existence of strong short range order is consistent with the value at TC of , which for these lat- tices is 0.71:.06 (Domb 1974). One might expect that for systems which require short range order stronger than this before long range order sets in, the h = O dip should be even larger at Tc' A lattice with a larger value of <0001> at TC can be formed following a procedure described by Syozi (1972). One starts by decorating the bonds of a honeycomb lattice with two sites and then one performs a star-triangle transformation, eliminating the original honeycomb sites (see Figure 7). This process can be carried out 45 T/TC=0.0 0.5 *- (D , 'fi T/TC=0.9 g 0.5-- ? é O ..l /_ P(h) T/T - —|.0 0.5 - C 4 9 O a z 21 Z T/TC=I.2 0.5 ‘ 7 r O 2 a a 13/1}:=‘33 0.5- « O E] g g 711 '3 'l I 3 h Figure 5. P(h) for the ferromagnetic honeycomb lattice. 46 E I T/TC=0.O g 0.5 ~ g .. é / O , 0.5 ~ - 0 P(h) T/TC=I.0 0.5 - - 0.5 L T/Tc=l.2 . 0 n m a m m n m T/TC=CD 0.5 - . on... m a E a .1 ..__1 ’6 -4 -2 0 2 4 6 h Figure 6. P(h) for the ferromagnetic triangular lattice. 47 -Decoration+ Star-Triangle + Figure 7. Generating the 3-12 lattice from the honeycomb lattice. 48 mathematically, generating exact results from analagous honeycomb results. The new lattice, called the 3-12 lattice, is more open than the previous one and hence requires more short range order, <0001>T = 0.8751, before long range order sets in. Note that c for the honeycomb, <0001>T = 0.7698. A comparison of P(h) at TC c for these two lattices is made in Figure 8. As anticipated, the dip is bigger for the 3-12 lattice. ' From the two dimensional lattices, one sees that the profile of P(h) above TC reflects short range order in the system; there appears to be little effect due to coordination number. Below Tc’ the rapid onset of the magnetization dominates. Finally, it should be noted that many of these calculations can easily be modified for antiferromagnets. For instance, a square net or honeycomb structure can be broken up into two sublattices so that at T = 0 the system can order by designating one sublattice as a spin up sublattice and the other as a spin down sublattice. The spin up sublattice will have a P(h) precisely as in Figure 4 (for a square net), while the down sublattice will have the same profile but with peaks on the -h side below Tc' The resultant P(h) for the entire system is symmetric, but below Tc’ Pi(h) z P(h) . (5) The antiferromagnetic triangular net does not decompose into two sublattices and hence cannot order. Calculations by Choy (1984) indicate that P(h) for this system changes very little with tempera- ture, which is consistent with its inability to order. 49 0.3 . P(h) 0.2 ‘ '3 -l 1 3 ’1 ..4 -3 -1 H E Figure 8. Showing P(h) at T for the honeycomb(upper) and 3-12 (lower) lattices. 6. P(h) for Ferromagnetic Systems of Higher Dimension There are no exact solutions for Ising models on standard three or four dimensional lattices. However, it would still be interesting to obtain some information, if only approximate, on systems of dimension greater that two in order to gauge the effect lattice dimension has on P(h). Frank et. al. (1982) developed an approximation method which effectively allowed them to estimate the moments of P(h) for a variety of three dimensional lattices at TC. In principle, one can take their moments and recover the original distribution. In practice, the error bars associated with those moments were too large, so that negative weights appeared in the reconstructed distribution. Since it appeared that little progress would be made analytically, Monte Carlo simulations were developed. The temperature scale for these simulations was set by the values for TC based on high temperature series and reported else— where: kBTC/J = 2.7044i.OOOl (diamond), 4.5108i.0002 (simple cubic), 6.3533 1.0010 (body centered cubic), and 9.79521.0005 (face centered cubic) from Domb (1974), and kBTC/J = 6.6817i.0015 (four dimensional simple cubic) from Gaunt et. al. (1979). No attempt was made to locate TC precisely with these simulations since detailed information (such as critical exponents) was not sought. Simulations generally have the most trouble converging when T = TC. However, since it is known that P(h) is symmetric at Tc’ 50 51 only PS(h) need be extracted at the critical temperature. This function couples to the energy, or equivalently to the nearest neighbor correlation <0001>, which is less sensitive to critical fluctuations than the magnetization and Pa(h). The criterion for convergence at TC was thus taken to be that <0001> should equili- brate. This correlation can then be compared to estimates obtained from high temperature series. The numerical simulations were performed on lattices of varying sizes and with periodic boundary conditions, employing the standard Metropolis algorithm (Metropolis et. al. 1953). Technical details associated with these studies appear in Appendix C. A given system would initially be brought into equilibrium and then be allowed to remain there while data was accumulated for thermal averages. Statistical errors quoted refer to the fluctuations in the thermal average during the second half of the averaging part of the simula- tion. The simulation was performed on three different lattice sizes (four at Tc)’ with the largest system containing roughly 30,000 spins. Then the data was plotted against the inverse of the num— ber of spins in the system to allow extrapolation to infinite size. Figure 9 shows an example of an extrapolation of <0001> for a simple cubic at Tc' The size of the data points is of the same order as the statistical error. The results of these extrapolations can now be compared to high temperature series results (see Domb 1974 and Sykes 1979) to gauge the accuracy of the simulation. This comparison, reported in Table 1, shows reasonable agreement between the two results. 52 .404 .34-l -32‘ 0.0 .2 I .0 .8 1.0 Figure 9. Extrapolation of <0 0 > to the infinite lattice limit for a simple cubic at Tc' 53 Table l. T as obtained from Monte Carlo simulations and c compared to high temperature series results. Lattice High T Series 1 Simulation Diamond .437 t .003 .44 i .01 SC .3307 t .0001 .34 t .01 BCC .2732 i .0002 .28 t .01 FCC .2474 t .0001 .26 t .01 Hypercubic .188 t .003 .20 i .01 54 Based on this, one would expect the data for P(h) to be accurate to within a few (3 or 4) percent, which is sufficient to determine the general nature of P(h). Figure 10 displays P(h) for four different three dimensional lattices. All of these lattices show an extremely flat profile near Tc' Whether there is a slight dip or bump at h = 0 cannot be conclusively determined from this data, but it is clear that these profiles differ from those of the two dimensional lattices. It is also evident that in three dimensions as in two, coordination number plays very little role in the profile. To see the dimensional dependence explicitly, it is construc- tive to look at the simple cubics in dimensions one through four (i.e., the linear chain, the square net, the simple cubic, and the hypercubic). These lattices all have coordination number equal to twice their dimension. The infinite dimension limit yields the mean field result discussed in Chapter 4, and will be included with this group too. These lattices have critical temperatures scaling roughly with coordination 2, providing a consistent way to normalize all the temperatures at which P(h) was measured (either analytically or by simulation). More exactly, one finds O S TC/z S l (1) where the lower limit corresponds to the linear chain and the up— per limit to mean field theory. The graphs in Figure 11 show P(h) for each of the five systems measured at each of the five (renormalized) transition temperatures. The profiles for any group of lattices above their transition 0.5 1 0.0 J 0.5 1 0.0 — 0.5 1 0.0 4 0.5 ‘ 0.0 n 0.5 y 0.0 J 0.5 1 0.01 0.5 1 0.0 - 0.5 1 0.0« 0.5 1 0.0« 0.5 4 0.0 -1 Figure 10. P(h) for four different three dimensional ferromagnets. 55 DIAMOND T/‘l’c = 0.0 I * r T/Tc=0.9 a T/Tc = 1.0 a—U—J—Z T/Tc=1.2 (.1 a 3 9L2: 171'c = on -4-2024 0.5 '0.0d 0.5 ‘ 0.0 0.5 0.0 0.5 0.0 0.5 0.0 0.5 ‘ 0.0 0.5 ‘ 0.0 0.5 0.0 0.5 0.0 0.5 1 0.0 4 J l T/Tc = .. 42-8404 312 h 56 .uHEHH tame“ came mnu 00m tam 430w :mzoucu mco mcowmcmsflc OJ '1 N\__ —_ o —1— '- _—_ o ——9- __1_ fi— 0- _J_ __L ., — l _ fl _ l T111 4 .11--.._I__ _ l _ _ fillrli I _ m _ 1. 11.11 _‘-._-..-4_.--_-,-_1 _ _. _ L . . 1.34011 111-31-. 4.1.111 _ . _ 1 _ L p _ L 11L 1!! x. p - L III. P L _ _ I. . 1.30.. u is ovum." as...» u is 22.1 22°» 3. 22... 1 Each 1 S. o a 3? u w: IO (p IO {p 10 Fr 1O f no mo moansu oaaefim now Acvm .HH oszwflm E am an 3 ”E. 57 temperatures are very similar provided they are measured at the same renormalized temperature. The most dramatic example of this is seen by looking at the one, two, and three dimensional data at a temperature corresponding to TC in three dimensions. All three profiles are extremely flat. Below Tc’ the symmetry is broken and the profiles lose much of their similarity. In the case of the four dimensional hyper- cubic, the approach to mean field theory can be seen by noting the development of a peak away from h = Jz. Clearly the coordination of a lattice affects P(h) in that it determines the number of 0 functions. It additionally helps set the temperature scale. Beyond that, the profile of P(h) is sensi— tive more to the dimension than to the coordination. At Tc’ this dimension dependence comes chiefly through <0001>T , reflecting c the amount of local order required before long range order can be established. Thus in four dimensions, P(h) at Tc is more nearly Gaussian than in one or two dimensions because less short range order is required in four dimensions for long range order to set in. 7. The Sherrington-Kirkpatrick Spin Glass Model Having looked at P(h) for a wide variety of ordered systems, it is now useful to turn to a disordered system, in particular, to a spin glass model. The model used is based on the Edwards and Anderson (1975) concept of a spin glass arising from some form of random interactions among spins. In the spin glass phase, the spins appear to be randomly oriented, but they are in fact frozen in. If one defines a quantity q by q = lim , (l) t+m then it becomes nonzero in the spin glass phase. Edwards and Ander- son argued that at least within mean field theory, this model is capable of reproducing some of the characteristics of experimentally observed spin glasses such as a cusp in the susceptibility which is rounded by the application of a field. Whether or not the Edwards and Anderson approach to spin glasses is an accurate description of experimentally observed spin glasses is not the focus of this investigation. Rather, as before, known results of a model will be exploited to calculate P(h). This form of P(h) will look substan- tially different from those found in the ferromagnets. Having ob- tained P(h), one can then perform a neutron scattering experiment on a disordered Ising system (preferably a spin glass) to link up with experiment. The spin glass model to be investigated is that introduced by 58 59 Sherrington and Kirkpatrick (1975). They assumed a Hamiltonian of the form H = -% Z Jijoio. (2) i,j 3 with the matrix of Jij's symmetric and having vanishing diagonal elements. The Jij's are otherwise independently distributed random variables with mean Jo and variance J2. Since this system has infi— nite range interactions, one might hope some sort of mean field theory would work. The infinite range interaction also means some variables must be rescaled with the number of spins, N: 32 JZN (3) O JON . (4) 3 The energy scale is then set by 3 or 30. When this model was introduced, Sherrington and Kirkpatrick proposed a solution based on the replica method. Starting from the identity in z = %ig(Zn—l)/n (5) where Z = TrIeXP(-BH)] (6) the calculation involves looking at Zn for integral n. Some algebra reduces this to n replicas of the original system with an effective interaction between a spin in one system and the same spin in a rep- lica.. The new system, exactly solvable, depends on the integer n. One then analytically continues Zn to n=O, recovering the free energy of the original system. While there appear to be no problems with this approach in the 6O paramagnetic phase, it yields incorrect results in the spin glass phase (e.g., negative entropy). The analytic continuation of Zn with respect to n and replica symmetry breaking are thought to be at the heart of the problem. The phase diagram obtained by the rep- lica method is shown in Figure 12. Due to the problems with the calculation outside the paramagnetic phase, the spin glass-ferromag- netic phase boundary may not be correct. However, since most subse- quent studies have been limited to JO=O, this appears to be the most accurate phase diagram available. Although Parisi (1980a) has devised a method for obtaining exact convergent series for various thermodynamic quantities in the spin glass phase, the less complete solution due to Thouless, Anderson and Palmer (1977), hereafter referred to as TAP, is more relevant to this study of P(h). Looking at the SK model for 30:0, TAP compared it to a Bethe lattice with the coordination number, 2, tending to infinity. The interactions on the Bethe lattice were taken to be nearest neighbor only and with a random distribution identical to that for the standard SK model. In the paramagnetic phase, TAP showed that the difference between the free energy per spin associated with such a Bethe lattice and the free energy per spin of the SK model with z spins, vanishes in the infinite 2 limit. They were also able to show this is true just below the transition temperature and for temperatures near zero. The TAP equations reproduce the replica result in the paramagnetic phase, E/N = —%BJZ . (7) 1-4 1.2 0.3 kBT/J 0.6 Figure 12. 61 PARA FERRO ‘ SPIN CLASS I l I I I l I 0.2 0.4 0.6 0.8 1 1.2 1.4 30/3 The phase diagram of the Sherrington-Kirkpatrick model (reproduced from Sherrington and Kirkpatrick 1975). 62 Unfortunately, their equations do not lead to exact, closed form solutions below TSG' They do however yield some information about the zero temperature form of P(h) for small h. First, a digression is required to a discussion which appeared in a paper by Palmer and Pond (1979). They assumed that P(h,T=O) a ht ' (8) and tried to determine t. Starting with the system in a local mini- mum, the N sites are relabelled in order of increasing [hi . The spins at sites 1 through n are flipped, where 1< i 2n(n/N) (17) t+1 - 2 ° 64 Since £n(n/N)l) . (20) is in general not equivalent to the definition used in equations (3) and (4) of Chapter 3, P(h) = Z<0(h—hi)> l N . 1 However, at T=0 for a system that orders, that is, when Pi(h) is of the form Pi(h) = 5(h-h2) . (21) these two definitions yield essentially the same result. PTAP(h) has the advantage that at low temperatures, it appears 65 to be less temperature sensitive than P(h) so that TAP can neglect temperature dependence in their calculation. However, PTAP(h) is a mean field quantity and will not necessarily yield exact results. Returning to their calculation, they first define mi = . (22) Using the fact that the Jij's scale with J = 3//; and are all hence small, and using standard Bethe lattice techniques, they arrive at 2 2 -1 ZJ..m. - m.BZJ.. (l-m.) = k T tanh m. . (23) J 1j j 1 j 13 j B 1 This equation is simplified by approximating 2 2 ~ .2. 2 E Jij (l-mj) - (1-mj) E Jij J J 2 <1-q) 32 (24) where q is the Edwards-Anderson order parameter. It will turn out to be consistent to take the low temperature behavior of q to have the form q = 1 - a(kBT/3)2 . (25) To further simplify notation, define h: s = z Jijmj . (26) J Then equation (23) can be rewritten * -1 hi = akBTmi + T tanh mi . (27) While an exact expression for q is '72 1 2 q = mi = N?- [fpi(h) tanh Bh dh] 9 it is plausible within the mean field limit to write q = f m2 PTAP dh* , (28) 66 where m(h) is obtained by inverting (27). No attempt is made either here or in TAP to rigorously justify (28). Combining equations (28), (25), and (19) one arrives at H 1 2 h* dh* = I [l-m ] -§' aE—'dm . (29) O H * Equation (27) will provide h (m) and its derivative so that 2 2 k T k T 1 0[—§—} = L%%J ‘. [l-mz] [0m + tanh-1m ] [ a + ] dm . (30) IO 1-m2 Integration and rearrangement yields 22n2+1 + 2n2 3 a = <0.0. ><0. 0. > 1 1+2 1 1+1 1+1 1+2 (1) tanhBJi+1,1+2 tanhBJi’1+1 (2) This is valid when site 1+1 is the only site crossed by the path from i to 1+2 (see Chapter 4 or Thorpe 1982). In this model, the temperature scale is set by kBT ~ 3 = J/N (3) so that BJiJ. ~ 1M? (4) and one can approximate tanhBJij 3 BJij . (5) Starting from the integral representation of the 0 function, Pi(h) = Ida exp(-ih0) < exp(ihi9) > (6) or Pi(6) = , (7) 68 69 and using h1 = E Jijoj . (8) (where the sum is over the z neighbors of site i), it follows that Pi(8) = = = <§ cos(Jije)[l + intan(Jij5)]> = <§ cos(Jij6)[1 + inJij6]> . (9) P(B) is obtained by averaging over all configurations of {Jij}' That average will be denoted by an overbar. Thus P(B) = 51(8) j 13 J 1J ----- z . cos(Jij8) [ 1 + 12(Jij6<0j>) + 12.212211 J..J. 62<0.0 > 2! 1j 1k 3 k 13 z(z—l)(z—2) J..J J. 03 + 3! 13 ik 1l .4 z(z:l)(z-2)(z-3) 4 + 1 4! JijJikJilJime <0j0k010m> + ...] + 0(l/z) . (10) The order l/z corrections, arising from averaging over all the cos(Jij9) terms independently of the Jij<...Oj...> terms, are neglected. In the paramagnetic phase, the thermal average of an odd number of spins is zero. Furthermore, using equations (2) and (5), 70 ~ 2 <0j0k> - B JijJik (11) Thus, in the large 2 limit, 2 ___ 2 4 ___ 4 ----- z z 2 2 2 z 4 2 4 P(B) — cos(Jij8) [1 - 578 [Jij 6 + 278 [Jij] 6 - ...] . (12) This last step has used fi——— 2 2 2 [J2 ] JijJik 3 ij (13) which holds since Jij and Jik are independent random variables. The infinite sum in (12) is the Taylor expansion of cosine, so that ----- z 2 P(G) - cos(Jij6) cos(BzJij9) . (14) Using notation and scaling introduced earlier, J2. = J2 + J2 ij 0 = 32/2 + 33/22 . (15) For large 2, '72 2 2 ij - 3 (16) Furthermore, cos(J 6) z = (1 - ABZI— + )2 ij 2 ii 2 2 6 3 2 z [1 - 22 + 0(1/z )] = exp(-0232/2) (17) and hence P(e) = exP(-6232/2) cos(8326) . (18) 71 Taking the inverse fourier transform, one obtains 1 exp(—0232/2 - 02/232) cosh(Bh) . (19) m3 P(h) = Equation (19) is valid throughout the entire paramagnetic phase and interestingly contains no explicit Jo dependence. Figure 13 shows a plot of P(h) at three temperatures. It was assumed that 30 < 3 so that (19) would be valid at k T = 3, B which is then the spin glass temperature. The infinite temperature limit is, as expected, a Gaussian: P(h,T=w) = exp(—h2/232) . (20) V203 When the temperature is decreased, the Gaussian broadens and flat— tens, a phenomenom made clearer by looking at the small h expansion, 2 2 2 h 2 2 P(h) = exp(-B J /2) [1 --———(1 - B 3 ) F211.) 232 AA 2 2 4 4 + 32(1/8 - B 3 /4 + B 3 /24) + ...] . (21) The fact that the h2 term vanishes precisely at TSG indicates how extremely flat this function becomes for small h. This calculation of P(h) cannot be repeated in an ordered phase for several reasons, among which are <0.) 1 0 J <0.0 > 1 <0.0.><0.0 > . j k j 1 1 k Without being able to express higher order correlation functions in terms of the nearest neighbor correlation functions, there appears to be little hope for finding an analytic expression for P(h) in 72 0.5 — kB'r/CI = 1 0 4 ---- kBT/3'= 2 0.3« P (h) 0.2 « 0.1 1 0 h/J Figure 13. P(h) from the analytic result for the Sherrin ton- Kirkpatrick model in the paramagnetic phase ( o < 3). 73 the spin glass phase of the model. For this reason, a Monte Carlo simulation was employed to study an ordered phase, specifically, the spin glass regime. The probability distribution for the Jij's was taken to be of the form p(Jij) = a0(Jij - J) + (l - a)0(Jij + J) , (22) where a = %(1 + Jo/J) . (23) This form ensures J.. = J 1j o and ?.=.I2 13 Since it was explicitly seen in the paramagnetic phase that once Jo and 32 have been fixed, the details of the Jij distribution are irrelevant, it is clear that in this phase one is free to choose the particular form of p(Jij) for convenience. The same statement cannot be proved in the spin glass phase but is believed to be true there owing to the infinite range of the interaction. The advantage of the distribution defined in (22) is that now Jij can be represented as a single binary digit. This fact was exploited to increase the effective memory available for the computer simulation and to increase the execution speed of the simulation. The details of this technique are provided in Appendix C. Unless otherwise noted, the data displayed in this chapter results from an average over four different 1020 spin systems. 74 In Figure 14 the results of a simulation for Jo = O at T = 2T are compared to the analytic calculation, and a similar comparison at T = TSG is made in Figure 15. While there is some noise in these simulations, it is clear that they are in essential agreement with the analytic result. 1 In Figure 16, P(h) as obtained from the simulation is plotted at four temperatures. In keeping with the trend established above TSG’ a dip at h = O has now developed below TSG' The data at T = O was accumulated by taking forty different {Jij} configurations and searching for one state in each that was stable against single spin flips. Kirkpatrick and Sherrington (1978) and Palmer and Pond (1979) have already studied the zero temperature problem in detail, so no attempt was made to be more careful in searching for local minima. Those authors, in fact, report lower values for P(h=O,T=O), and it is believed that for an infinite system, P(0,0) = 0. However, even using states stable against only single spin flips, the small h slope obtained from this simu- lation, 0.311/32, is in excellent agreement with the TAP (1977) prediction of O.3O7/J2 discussed in the previous chapter. Hence the slope appears to be less sensitive than the intercept to the particular local minimum one is in. As the temperature is reduced from TSG/Z to 0, most of the change in P(h) occurs for small h. The large h regions are relatively stable. This suggests that the formation of the h = O dip is an important characteristic of the spin glass phase. To study P(O) in more detail, first note that above TSG’ 75 0.4 —— SIMULATION ---- EXACT 0.3 -1 P(h) 0.2 -1 \ Figure 14. A comparison between the Monte Carlo simulation of the Sherrington-Kirkpatrick model and the analytic result: T = ZTSG' .4 76 0.4 — SIMULATION ‘ ---- EXACT 0.31 P(h) 0.2 - 0.1 — 0 . -6 6 Figure 15. A comparison between the Monte Carlo simulation of the Sherrington—Kirkpatrick model and the analytic result: T = TSG' 77 0.4 - — kBT/J = 2.0 ""’" kBT/J =1.0 0.3 _‘ """ k317i: 0.5 P (h) " , 0.2 “ 1' . " J 0.1 - C) I 1’50 1 I I ‘1 -6 -4 -2 0 2 4 Figure 16. Results of the Monte Carlo simulations of the Sherrington-Kirkpatrick model: P(h) at four dif- ferent temperatures. The T=O data has been symmetrized. 78 P(h=O,T) = 1 exp(+82J2/2) (24) /203 so that at TSG = J/kB, P(h=O T ) = -l——-e-% (25) 30 55:1 It is also easily shown that _ _1 BPgh-OzT) = l e 2 (26) 18C V_fi Since it is believed that the specific heat and energy are continuous at TSG for this model, these values hold for just under TSG also. If one adds to this the assumption that P(0,0) = O , (27) there are three constriants on P(h=O,T) for O 5 T g T G' One solution, S but certainly not the only solution, is to try the linear form k _1 -l——-e 2 B T . (28) P(h=O,T T (29) SG Pi(h) 1 P(h) T < TSG . (30) The detailed nature of this broken symmetry is not as of yet clear, although one can say that it is substantially more complicated than the two sublattice state of the antiferromagnet. Following the temperature development of P(h) has allowed the tracking of its development from a Gaussian function to a function with a hole in the center. It has been seen that the dip develops precisely at TSG and the system very quickly establishes the large h states. Below about TSG/2, most of the action in P(h) occurs in the small h region. There has been some recent work (Mézard et. al. 1984) sugges- ting that the relevant order parameter in this system is not q (the Edwards—Anderson order parameter), nor is it q(x), a generalized probability function introduced by Parisi (1980a). Rather it has been suggested that the system is best described by a probability distribution governing the distribution q(x). That is, by performing a configurational average to calculate q(x), one is throwing away valuable information in the spin glass state. One might be con— cerned then that the Monte Carlo simulations performed configurational averages and that what should have been measured was the probability distribution of the distribution, P(h). However, it has been noted 81 earlier that one cannot calculate q directly from P(h); one needs to know Pi(h) at all sites. Thus q is not expected to couple strongly to P(h). It is not unreasonable that even though q and q(x) may not be well defined thermodynamic variables, P(h) may nevertheless be well defined for this model. 9. Future Work While the nature of disorder in Ising systems has been examined here, the major development has been that of a new tool, P(h), to study Ising systems at all temperatures. This function leads to exact thermodynamic results as well as having a measurable symmetric component. An understanding of this function for a wide variety of Ising ferromagnets is in hand. However, only one disordered system, the Sherrington—Kirkpatrick spin glass model, has been studied. While some useful conclusions were drawn, it was not possible to determine which general characteristics of P(h) were common to Ising spin glasses in general and which were peculiar to infinite range spin glasses. Therefore, it would be useful to look at an Ising spin glass model with short range interactions. It is reasonable to expect that a system that orders will have a dip at h = 0, but beyond that it is difficult to generalize on the basis of the Sherrington-Kirkpatrick model. One can extend the definition of P(h) for Ising systems to a P(H) for classical KY and Heisenberg systems. While the relation— ship between this function and the inelastic neutron scattering law is not clear, inspection of the proof in Chapter 2 reveals that a similar relation between the energy and P(h) should hold. It was crucial in that proof that all parts of the Hamiltonian commute with each other, and this is the case for these classical systems. 82 83 Extension to these systems would allow for the simulation of a Heisenberg spin glass. One possibility would be the EuxSr1_XS system. Eu2+ ions have spin 7/2 and lie on an FCC lattice. They have ferromagnetic nearest neighbor interactions and antiferromag- netic next nearest neighbor interactions. EuS acts as a ferromagnet, but dilution with the nonmagnetic Sr2+ ions enhances the competition between the nearest and next nearest neighbor interactions, creating a spin glass at appropriate concentration and temperature. This system has been well studied both experimentally and by simulation (see, e.g., Krey 1981). To the extent that spin 7/2 can be approxi- mated as classical, EuxSrl_XS then provides a prime candidate for studying the temperature dependence of P(h) in a well characterized system. A connection between S(k,w) and P(h) could be extremely useful in comparing simulation results to experiment. Another system worth investigating consists of classical XY spins on a triangular lattice with nearest neighbor antiferromag- netic interations. A study by Lee et. al. (1984) indicates a very rich phase diagram in the H-T plane. With this single system it would be possible to sample several different phases. It may also be possible to extend the definition of P(h) to the spherical model (Berlin and Kac 1952). This would provide access to the well studied spherical infinite range spin glass model (Kosterlitz et. al. 1976). A more difficult extension of this formalism is to quantum Heisenberg or XY models. Problems arise early on since the local field operator, K. = 22.1.15. (1) 1 . 1] j J is not diagonalizable. Hence, P10?) = <60? - Ki» - (2) is not well defined. Preliminary efforts to relate the Ising P(h) to P(IHI) or to P(hz) for quantum systems have not yielded a use- ful formalism. It could be that there exists a relation between the energy and some generalization of the Ising P(h) that is more complicated than a simple integral. Any such hypothesized relation— ship can be checked against the known results for the infinite quan- tum KY chain (see, e.g., Lieb et. al. 1961) or for finite length quantum Heisenbero systems (Bonner and Fisher 1964). An extension of the P(h) formalism to quantum systems, while difficult, has the potential for providing new insight into these models. Appendix A. Local Magnetic Field Distributions I. Two-Dimensional Ising Models Local Magnetic Field Distributions I. Two-Dimensional Ising Models M. Thomsen+ Department of Physics and Astronomy Michigan State University East Lansing, MI 48824 U.S.A. M. F. Thorpe++ Cavendish Laboratory, Madingley Road Cambridge CBB OHE, U.K. T. C. Choy and D. Sherrington Physics Department, Imperial College London SW7 ZBZ, U.K. Copyright 1984 The American Physical Society +Exxon Fellow ++Permanent address: Department of Physics and Astronomy, Michigan State University, East Lansing, MI 48824, U.S.A. 85 86 Abstract We show that the Statistical mechanics of Ising models :an be conveniently reformulated in terms of the local magnetic field probability distribution function P(h). It is shown that for arbitrary exchange interactions J44, ~J which may or may not be random, boch thermodynamic quantities like magneti- zation, specific heat etc.; and the neutron scattering law S(k,0) can be obtained from P(h). Indeed S(k,m) provides a direct measurement of the symmetric part of P(h) which also determines the energy, specific heat etc. while the magnetization can be obtained from the antisvmmetric part of P(h). As an example, specific results for P(h) are presented for the honeycomb, square and triangular lattices with constant nearest neighbor interactions. All three lattices exhibit a pronounced dip in the center of P(h) at the transition temperature. 87 I. INTRODUCTION At various stages in the development of the theory of magnetism, the local field probability distribution function P(h) has been used}.5 Despite the fact that P(h) has a simple physical interpretation, it has 323 been widely adopted. This is probably because its usage has been associated with mean field type theories and as such it is regarded as an approximation. In this paper we show that an exact, useful and complete description of the thermodynamics of a rather general class of Ising models can be obtained in terms of P(h)- These models have localized spins Si associated with sites R1. The spins can have different magnitudes and the sites R1 do not have to form a crystalline lattice. The Hamiltonian involves only the S2 components and.contains arbitrary exchange interactions Ji and external 1 1 fields Hi' The range of the Jij is also quite general. All these results are scattered around in the literature of the early 1960'31-4 and are A summarized by Southern.3 We collect them together in this paper. We show that a measurement of the inelastic neutron scattering law S(k,m) at a particular temperature provides a direct measurement of the symmetric part of P(h) at that temperature. The total magnetization of the system can be obtained directly from the antisymmetric part of P(h), although the converse is not true. Finally the total energy can be obtained from the symmetric part of P(h), if all the Hi = O. Other thermodynamic quantities of interest, like the free energy, specific heat etc. can be ‘H 88 obtained from the energy. The above connections are very powerful because they require a knowledge of only P(h) and the temperature and no other infor- mation about the system. In the case where the Hi # 0, the energy, and heats the other thermal thermodynamic quantities, can still be found from P(h) in simple cases. Such an example would be a uniform field where all the Hi a H, and a knowledge of P(h), the temperature and H would lead to the energy. The layout of this paper is as follows. In the next section we develop the general formalism for S s 3. This is for pedagogical reasons and the rather obvious generalizations to arbitrary spin are given in the Appendix. In Sec. III we illustrate these results for various spin 3 nearest neighbor Ising models; the one-dimensional chain and the honeycomb, square and tri- angular net in two dimensions. These results are exact and could have been obtained at any time in the past 20 years. Surprisingly they do not appear to be in the literature. The results in two dimensions are rather interesting as P(h) changes from a roughly Gaussian shape at high temperatures to develop a pronounced dip in the center at the transition temperature Tc. It is-unclear how general this phenomenon is. We define a quantity that measures the fluctuations in the local field h and show, not surprisingly, that it has a maximum at Tc. 89 The attraction of P(h) is that it provides an alternative description of a thermodynamic system with perhaps a more intuitive interpretation than the free energy. It contains more information than just the free energy and, for example,the neutron scattering law S(k,m) can also be derived from it. We note that S(k,m) cannot be derived from the free energy. Our original motivation for this work was that computer experiments on spin-glass models at zero temperature often focus on P(h). Indeed it is a "natural" quantity to compute. Such experiments have universally demon- scrated a zero-field minimum in spin-glass models with competing interactions in their ground or lowblying metastable scate5.: A similar minimum has . . . . . - .- 7 3 . oeen found in Simulations of amorphous antiferromagnets. ’ Because or the interescing new results found for simple Ising models, we have deferred a discussion of more complex spin systems with competing interactions until a subsequent paper. II. GENERAL FORMALISM We define a spin 8 Ising system with the Hamiltonian H. 0. (l) H. U H U P: LJ HIvI H H where the Ji" Hi are arbitrary and the factor 3 is to prevent double counting. This Hamiltonian describes a completely general spin 8 Ising system with 0i - : 1, an arbitrary range of exchange parameters Jij and an external field Hi that can vary from site to site. We define the Jii * O. The sites can be inequivalent and we make no assumptions about the existence of a crystalline lattice or translational invariance. Although we develop these results for spin 8, all the results of this section easily generalize to arbitrary spin . ..r" 90 as shown in the Appendix. There are a number of ways of decomposing the Hamiltonian (l). The most obvious is where L4. 1 This is not useful in discussing local fields whereit is necessary to decompose (l) as H--hioi+H' (4) where The first term in Eq. (4) contains all terms involving 3i and other terms are lumped together in 3'. Notice that hi as defined in (3) and hi as defined . l . 1n (5) differ in the factor-E in front of the exchange term.. This means that some care must be taken in calculating the energy of the system later in this section. We consider the thermal average ( 00, > where O is any operator not involving site i; '1 91 < 0": > - < Tjr [0°11 exp (311131)] / TirEexpmhicifl > - < 0 tanh Bhi > . (6) 7 Putting 0 equal to the unit operator we find that 9 < 31 > - < tanh Bhi > a < tanh 3(é Jij Jj + Hi) > . (7) J We note in passing that replacing the thermal average of the tanh in (7) by ’3 the tanh or the thermal average leads to mean field theory. Putting 0 equal to hi in (6) leads to Jijgi‘j 11 La. ('1. 1 a < hi tanh 8h, > (8) It is not possible to obtain the energy from (8) alone for H1 # 0 because of the missing factor-% in front of the exchange as required for the decomposition in Eqs. (2) and (3). However the energy can be obtained as (9) l , - 2 < (hi + H1) tanh 3hi > . x The magnetization M and the energy E are given by I". 92 K n Hd\4 A Q P. V (10) We introduce the probability distribution Pi(h) for the local magnetic field at site i by Pi(h) - < 0(h - hi) > (ll) where both the "internal" and external fields are included in hi via Eq. (5). It is of course possible to define a similar quantity Pi'(h) with only the internal field contribution counted; ' I — I — P1 (h) <50. §Jijoj)> Pi(h 111) . (12) For the rest of this paper we shall use (11). From the definition of P,(h), we see that jPi(h)dh- 1 , (13) <'TJ J+H)n>=‘r'nP(h)dh (1') ‘1; ij 1 1 ,1“ 1 1 “ and,from (7) and (9), < oi > a I tanh 3h Pi(h) dh (13) J (16) A. h tanh 3h Pi-I’h) dh - 1% Hi < I; > m I I NIH ’4- 93 The distribution function P(h) for the whole system with N sites is obtained from (11) via P(h) - §- 2 Pi(h) . (17) 1 It can be seen from (15) that a knowledge of P(h) and the temperature is sufficient to determine the magnetization ; u . N ; tanh 3h P(h) 1h . (18) It is not necessary to know anything else about the syscem; in particular it is not necessary to know the Jij or the Hi. It is this sense in which P(h) provides a description of the system in the same way as the free energy does. Because tanh 8h is an odd function of 8h, only the antisymmetric part Pa(h) of P(h) contributes to M: l . Pam) =- -2- [P(h) - 14-10] . (19) (20) u a N j tanh 3h Pa(h)dh In general the energy cannot be obtained directly from P(h). However ii, all the Hi are known to be zero, then from (10) and (16) N I E a --5 J h tanh 3h P(h) dh (21) and E can be calculated directly from P(h) without any knowledge of the Jij' 94 Because (h tanh 8h) is an even function of h, only the symmetric part Ps(h) of P(h) contributes to E ; 930:) %IP(h)+P(-h)1 . (22) le I I h tanh 3h 9,01) dh . (23) J In this case the free energy F is given by L) If the Hi are sufficiently simple, the free energy can still be found. For example if the magnetic field is constant at every site Hi - H,the energy given in (9) becomes \1 1’ E - -13 J (h + H) tanh 3h P(h) dh (25) and the free energy F(H,T) is given by {24) with the integration done at constant H. It can be seen that under appropriate conditions the local magnetic field probability distribution P(h) leads to the magnetization M and the energy B. All other thermodynamic quantities like the free energy, specific heat etc. have to be found by integrating or differentiating M or E. This is because M and E are determined from the expectation values of local operators whereas other thermodynamic functions are not. Thus M and E play a special role when the statistical mechanics is done via P(h). This point appears not to have been appreciated by Klein and Brout4 who attempted to write the free energy as an integral over P(h) [see their equation (2.5)]. ‘u The function P(h) contains more than just thermodynamic information; it . 10 is directly related to the inelastic neutron scattering cross section, +09 .. 1 I" _' q .. .5 -> * . Sac...» a ,—‘- - e ‘M at; exp[i 14- (R. - R.)]< J.‘ LYCc) > (26) in j ij i 3 i J -Q where x is any direction perpendicular to the z axis. In the previous part . z of this section we suppressed the z superscript on the 3i operators. This cross section is k independent because Ising systems have no dynamics, so T - . . . .1 that transrorming to raiSing and lowering operators , < o. 370;) + 37.7fm > . <27) 1. l 1. cl. (‘1 1 a "' .L ( , ) a" .0 The time dependence of these operators is given by 01 (t) - exp(th) oi exp(-th) - oi exp(2ithioi) (28) where the h, are given by (S). Combining (27) and {28) we find that f“ s(i€,w) - '21? e‘i‘“ dt 2 < m(Zithioi) > 1 +09 3 ‘L' g efil‘”t dt 7 < cos 2th. + i 0, sin 2th. > . 2H J i i i 1 Using equation (6), this can be rewritten as 96 M 3(E,w) =«J= { e‘l‘”c dt 3 < cos 2th. + i tanh 3h. sin 2th.> 2w } 1 i 1 i +ao a 3%.j [ e-1wt dt E(cos 2th + i tanh 8h sin 2th ) Pi(h)dh J 1 . é} [P(w/Z) + P(-«n/2)] / [1 + exp(-3w)] . (29) which shows that the neutron scattering law provides a direct measurement of the symmetric part P‘Ch) of P(h); 3 3(E,d) . u PSCm/2) / £1 + sxp(-dw}3 (30) The thermal factor [1 + exp(--13m)].l ensures detailed balance. From the properties of P(h), it is easy to show that 1 r w EE' 1 S(k,m) [1 + exp(-Bm)] dw . 1 (31) and if all the Hi = O r '% } S(k,») [l - exp(-Bw)] ndw 8 -E - (32) Both (31) and (32) hold for arbitrary Jij' The expression (31) could be useful in normalizing neutron data taken at different temperatures, if this data were to be used to find E. 97 III. TWO-DIMENSIONAL ISING MODELS In this section we present results for P(h) for some pure two dimensional Ising systems.12 This is straightforward but does serve to illustrate some of the points made in the previous section and provides some interesting results. It is convenient to redefine the Hamiltonian (1) with only a nearest neighbor Jij which is set equal to l ; where the sum goes over nearest neighbor pairs only and the factor % is to prevent double counting. In what follows we use a slightly modified form of 1 the notation of Choy and Sherrington.L3 As all sites are equivalent we can ignore the distinction between Pi(h) and P(h). Therefore P(h) I < 6(h - G.) > (34) l J I «\«1 H La. where the sum over j goes over the z nearest neighbors of atom i. This can be conveniently rewritten as m p z P(h) 8 fL' ! exp(-ihd) dB < exp(i 9 S J.) > ” } j;l -Q 2 3.1; . exp(-ih5) d9 < H (l + i 0, tan a) > cos2 9 (35) -@ where we have used the an integral representation of the Dirac delta function with the usual convergence factors implied at :w. ‘1'}. 98 Multiplying out the brackets in the product, this becomes a ct 6(h - s) (36) where the r sum is in steps of l and the 3 sum is in steps of l. The final result has the form 2 P(h)- E w~ 5(h-s) (37) 33-2 ° with 2 "s“; 3 3:5 C (33) 2 r-O r z and the ars are given by the generating function a 2% J exp(-ih6) d6 cosz-r a sinr a 1r 1 3 z 1 =-- ' a o’h - s . 39 2 g rs \ ) ( ) 2 s--z The factor l/2z is included so that the a:s are integers. They are easily obtained for a given 2 by doing the simple integrals involved in (39) and are . . ' . - . l3 -. . given in Table l. The Cr are correlation functions defined by q - 3 < 3. J. . . . o > (40) c » ..‘ . 1. 2. (1]...L) J r 7‘1 where the among the neighbors 0 r4 0 ‘99 round brackets denote that only the distinct products of r operators 2 nearest neighbors are to be taken. If we label the z nearest of an atom cyclically from 1 to 2, then for the linear chain 1 r4 ' E < a; > a 2 < J, > (41) a Z < 3 3. > = < I, 3, > . For the honeycomb lattice 12 = 3) (J a : < 7 > a 3 < :1 > (i) > (32) ' ;. <<:J 3 > a 3 <31 :2 > (13) a E < 31 3 3‘ > a < 3 37 :3 > . a 4 < 51 > (1) c a R" >al ’ 2 F r. < I“ IJ 4 < 3* 32 > + < 31 73 > (13) c a 7 < 3 I 3, > a a < 1 3q 3 > 3 (13k) l J a l - 3 C; a f < 3L Jj 3‘ it > a < Jl 32 J3 Ia > J . kljki) and for the triangular net (2 a 6) e -Z (oi>-6
    l (i) C =- y > a 1 1 2 (f‘ < 3i Jj 6 < 01 02 > + 6 < 31 J3 > + 3 < 31 04 > J) . c a .; < giJij > a 6 < 315233 > + 12 < 013234 > + 2 < 31 03 (13k) C = 7 < o. o. o, o, > = 6 < o o o o > + 6 < o J o J- A (ijkl) i 3 a Z l 2 3 4 l 2 3 3 + 3 < Jl 02 3 Us ‘2 a E < o. o. o, J, o > a 6 < 3 o, J J; 3- > 3 (ijkim) i J a L m l - 3 4 3 c6 = ) =<3132 33 '34 35 36> (ijklmn) o- > 3 (43) 5 (44) 101 z . - . . . From the ars and the cr’ the amplitudes or the various fields ws occurring in (37) can be calculated using (38). Many of these correlation functions Cr can be obtained from the literature for two-dimensional lattices. No exact calculations of the Cr exist in higher dimensions. We include the linear ' chain only because it is very simple. In what follows h 3 SJ, where J has previously been set equal to one. 14 For the linear chain , s, 8 tanh” K so that from (38) we have 1 2 w = - l - tan . K o 2< h ) (4+6) W 'W -$ o l (47) m = < 3 > o 102 where 00 is the spin at the site of interest. This is achieved by rewriting (6) as 1.5-1.8. < 0 00 > 8 A < 0(0l + s + 03) > + B < O o J 3 > (48) 2 l 2 3 where [tanh 3K + tanh K] J-‘It-J (49) - . ,, _ ..,1 B 8-: Ltann sh - 3 tanh 34 Putting 0 = 1 yields < 31 J, 33 > = m(l - 3A) / 3 (50) and putting 0 . 01 yields < 31 32 > = (a - A) / (2A + 3) . (51) Combining all these results we find that19 h . l_ 3(tanh 3K - s) * 3m(tanh 3K - 1) i1 2 3 tanh 3K - tanh K ‘ tanh 3K - 3 tanh K (52) h 3.1 35 — tanh K - m(l - 3 tanh K) =3 2 3 tanh 3K - tanh K ‘ tanh 3K - 3 tanh K 12,13 The reduced energy 5 can be written as an elliptic integral and the 90 reduced magnetization m is a known function or K. These results are illustrated in Fig. l. 103 For any lattice at high temperatures P(h) obeys a binomial distribu- tion. The field is h if %(z + h) nearest neighbor spins are up and %(z - h) are down. This occurs with probability 2C1“Z + h) / 2Z so that at infinite temperature P(h) . E J? zcl 5(h - s) . 2 ‘§(z + s) s = -z As 2 gets large this approaches closer to a Gaussian distribution. When the temperature is lowered P(h) at first flattens and then develops a pronounced dip just above TC for the honeycomb lattice as shown in Fig. 1. At TC, the values of VS can be obtained by putting 5c - 4vr79, mc a O and exp(2Kc) - 2 + /3 in (52). The distribution function rapidly develops an 3 with asymmetry below Tc as the magnetization increases from zero as (Tc - T) B - 1/8. At zero temperature there is a single peak at h a 3 as we would expect when the alignment is complete. Note that P(h) contains the critical exponents a (through a) and 3 (through m) in Eqs. (52). All the critical exponents of interest can be obtained from these two.21 As the coordination increases, so does the number of correlation functions that must be known. For the sguare net, the three spin correlation function is still proportional to the reduced magnetization via a relationship l5-18 . . . However.it lS necessary to compute two other indepen- similar to (50). dent even spin correlation functions as they cannot all be reduced to just 5 via relations like (48) and (49). These involve evaluating elliptic integrals. For the triangular net the situation is much more complex but a similar reduction to elliptic integrals for even spin correlations can be made starting - - . 22 . . trom the Pfaffian form given by Stephenson . The details are complicated and ‘33 2 are given elsewhere‘ . While the magnetization is given by Potts , the remain- 104 ing odd spin correlations can be found using the methods of Barry, Minera and Tanakazs. However, we found their results to contain errors and have not pursued this further. Consequently, Fig. 3 for the triangular net is not complete for T < Tc' The results for the square net and the triangular net are shown in Figs. 2 and 3. They are very similar to those for the honeycomb lattice given in Fig. 1, showing a rather universal behavior for P(h) for all two- dimensional lattices. Results have been published previously13 by two of us for the antiferroe magnetic triangular net which show that there is very little change in P(h) between T 8 n and T 8 O apart from some flattening. This is related to the fact that this model does not have a phase transition and has a finite entropy at zero temperature. The antiferromagnetic honeycomb and square lattices map on to their ferromagnetic counterparts as they are both bichromatic. Using the definition (ll) for Pi(h)’ these functions will be identical above Tc for both sublattices and equal to the ferromagnetic counterparts. However below TC, the asymmetry will develop on opposite sides of h 8 O for the two sublattices and so P(h) for the whole syscem is juSt the symmetric part PSCh) of P(h) defined in (22) for the ferromagnetic counterpart. This leads to zero net magnetization via (18) as expected. In order to get the staggered magnetization it would be necessary to know Pi(h) at the up and down sites separately and not just the average. The one-dimensional Ising model can also be regarded as having a dip in P(h) at Tc if TC is identified with zero temperature. At zero temperature wO = 0 and w2 = w_2 8 8 from (46). If the distribution P(h) is to be characterized by a single parameter, the most useful is V8[h2-(h)2]/z (s3) 105 where the bar denotes an average over P(h). From the definition of the ct, this can be rewritten as V8l+(2c-c2/z <54) 2 l ) and is shown in Fig. 4 for ferromagnetic interactions in the honeycomb, square net and triangular net. At high temperatures v - l for all lattices as the correlation functions c1, c2 become zero. Not surprisingly this quantity, which measures the local fluctuations, has a maximum at T5 and then decreases rapidly to zero as the temperature goes to zero. The increase in v as the temperature is lowered from infinity towards Tc is due to the flattening and then the dip in P(h). The rapid drop in vbelowTé is due to the asymmetry caused by c1 and hence the magnetization. For antiferromagnetic interactions, the quantity v keeps going up as the temperature is lowered. For two sublattice antiferromagnets, like the honey- comb and square net, the local magnetic field distribution function is given by the symmetric part (see Eq. 22) of the distribution for the corresponding ferromagnet, (i.e. all exchange interactions changing sign). Thus v increases monotonically from 1 at high temperatures to z at zero temperature. For the triangular net antiferromagnet v increases from 1 at high temperatures to 1.097809 at zero temperature.13 Finally in this section. we consider the neutron scattering law (26) and (30). The scattering only takes place at discrete energies. In Table 2 we have included the thermal factor in (30) to give S(k,w) for the square net. At very low temperatures, only the "spin wave peak" is seen at w 8 8.10 This corresponds to flipping a spin where all its nearest neighbors are parallel. As the temperature is raised, other peaks have nonzero weights. '1!‘ 106 Even at Tc’ we note that 63% of the weight is still in the spin wave peak. However there is some weight at w 8 4, which corresponds to a spin flip where 3 nearest neighbors are up and l is down. The peak at w 8 0 will combine with the elastic peak and the peaks at negative frequencies are related to those at positive frequencies by detailed balance factors. At infinite temperature the thermal factor is 8 for all frequencies and the neutron scattering law becomes symmetric. There is a useful sum rule for spin 8 (additional to (31) and (32)) which shows that the total integrated intensity is independent of temperature, as is clear from Table 2; J 302,11) dw =- 1 . ' (55) This is easily proved from the definition (26). IV. CONCLUSIONS We have shown that the statiscical mechanics of Ising models can be described through the local magnetic field probability distribution function P(h). This function determines both the neutron scattering law S(k,n) and the thermodynamic quantities of interest for a large class of Ising systems. We have calculated P(h) for the ferromagnetic honeycomb, square net and tri— angular net and have shown that in all cases a pronounced dip develops at TC. In a subsequent paper, we plan to extend these results to random spin systems and spin-glass models and also to look at systems described by classical rather than Ising spins. 107 ACKNOWLEDGMENTS We should like to thank J. H. Barry, R. Haydock, T. A. Kaplan and J. Stephenson for useful discussions and the N.S.F. and S.E.R.C. for financial support. One of us (M.F.T.) would like to thank the Cavendish Laboratory for its hospitality. 108 APPENDIX The results of Sec. II can be generalized to case of arbitrary spins Si whose magnitude can vary from site to site. We will adopt a numbering scheme for equations such that equations in the text and the Appendix corre- spond. The Hamiltonian is aa-éf J..S?S?-)H.S‘T‘ . (Al) - .. ij i j 1 i i ij 1 Note that this J;. differs from the jij in the main text by a factor 4 in the limit when the spin becomes 3. This is because we found it more convenient to use Pauli operators in the main text. In a similar way there is a factor 2 different in the definition of Hi' Rather than redefine these quantities we prefer to start from (Al). The reader will easily discover places where factors of 2 appear between equations in the main text and the corresponding equations in the Appendix. In order to proceed it is necessary to insist that the Jii 8 O. The most obvious decomposition of (Al) is h S? ~ (A2) Pl‘l H. P where J 57‘ + 3. (A3) (_J.‘ ‘v'l but the most useful for our purposes is H = -hi 3",: + H' (A4) where J.. S§+H. . (A5) P LJ. p4 "l L: H 109 . . . . . z . Note that it is important that H be linear in all the 31 so that terms like 2 2 z z z 2 . . . . . i) S , (SiSj) etc. are not allowed in the damiltonian if the J present formalism is to be used. 2 2 (Si) . (s z . . . . The thermal average < O 51.- > where O is any operator not involv1ng Slte i is given by z — z 2 g 3 z < o si > =- < Tjr Lo 31 exp(8hiSi)] / i: [exp(ph5i)] > 8 < O 3- (,Sh.)> , (A6) a. 1 1 - . 26 . - . where the modified Brillouin :unction is defined by 3()=(S+l)rh’f3+1=)7 3 n’/"‘ S x 2 tot L, 2 x j 2 cot (x _, , and hence B ( l h /2 \ 3 x) 2 tan x . Putting 0 equal to the unit operator in (A6) we find that Z a . B sh.)> ; A7 < s,( 1 ( ) 1 putting 0 equal to h; leads to =- i 1 J ij i 1 i a ' B c > , 1. l The magnetization M and the energy E are given by (10) as before. It is convenient to define a P(h) for the whole system as - fl P(h) aL 33 (3111 91m) 1 r . 3 (3h) . (All) 1 1 1 1 and, further, an average Brillouin function 3(Bh) by 4 l Mam-E 233 (an) . i i This takes account of a varying spin magnitude. It is only necessary to know the number of sites with each spin magnitude in order to know B(Bh). From (7) and (9) we have 7- -9, < 51 > } 3 (Sb) Pi(h) dh (A15) 1 S and F: a _ i f, H’ 3 - 1'. z “‘1 2 } h psiflsh) Pi(h) dh 2 Hi < si > . (A16) The magnetization M is given by M = N f 3 (8h) P(h) dh (Al8) 111 and,as before,it is not necessary to know the Ji or the Hi in order to find 1 M from P(h). Again as before, the energy cannot be obtained directly from P(h). How- ever if all the H are known to be zero, then from (10) and (A16) 1 E 8 -‘% t h 3 (8h) P(h) dh , (A21) the free energy F is given by 1“ SF . - : m(lsi + 1) - ;' :de , (A14) l J O and equation (ll) generalizes in the obvious way with tanh 3h - 3(8h). The inelastic neutron scattering cross section is given by 2 +m 502,11) . 7 [ -co where x is any direction perpendicular to 2. Using arguments similar to those in Sec. II, we find that * l ' -iwt ~ + - - + S(k,w) - _— I dt » < s. s (c) + s. s. (c) > Zn i i i i i =- 2 j [ 9,.(11) + Pi(-'u) J 33 (8m) / [ 1 - exp(-8w) ] 1 " 1 = 2N E P(w) + P(-w) ] 3(Bw) / [ 1 - exp(-.6w) 1 (A29) 112 or S(k,m) 8 4NPs(w) 3(Bw) / E l - exp(-3w) ] These lead to the sum rules (A30) _l_ v . l - exp(-8m) . 2N )1 S(k,.u) 2 30311)) d 1 (A31) and if all the Hi 8 O 1 — _ . -§ S(k,n) L 1 - exp(-dw) j a dw . 5 (A32) These sum rules are very general and can be proved without introducing P(h). The second one is particularly easy to establish as the magnitude of the spin does not enter explicitly. Using the identity‘ < A(O) B(t) > 8 < B(t - 18) A(0) > for any operators A,B; we see that 3(k,o) : 1 - exp(-ém) I 113 Only the diagonal terms contribute for Ising models so that + i ,+ - - + E < 1 >1 . his1 J E 5i , hiSi ] > =- 4 ,‘7 < 11.5.2 > i i i \J O 8. 10. 16. 17. 114 REFERENCES W. Marshall, Phys. Rev. 118, 1520 (1960). H- B. Callen, Phys. Lett. 4, 161 (1963). B- Mfihlschlegel and H. Zittartz, Zeit. fflr Physik 122, 553 (1963). M. W. Klein and R. Brout, Phys. Rev. 122, 2412 (1963). B. W. Southern, J. Phys. C‘g, 4011 (1976). D. Sherrington, A.I.P. Conf. Proc. 22, 224 (1975) reporting work by S. Kirkpatrick, L. R. Walker and R. H. Walstedt, Phys. Rev. Lett. 28, 314 (1977); Phys. Rev. 3 22, 3816 (1980); K. Binder, Z. ffir Physik 3 26, 339 (1977); R. . Palmer and C. .. Pond, J. Phys. P‘g, 1451 (1979): P. T. Bantilan and R. G. Palmer, J. Phys. P 11, 261 (1981). S. N. Khanna and D. Sherrington, Solid State Comm. 36, 653 (1980). J. R. McLenaghan and D. Sherrington, to be published. T. C. Choy, D. Sherrington, M. Thomsen and M. F. Thorpe, to be published. W. Marshall.and S. W. Lovesey, "Theory of Thermal Neutron Scattering" (Oxford University Press, London, 1971). See for example, L. Schiff in "Quantum Mechanics" (McGraw-Hill, NY, 1968). L. Onsager, Phys. Rev. 62, 117 (1944). T. C. Choy and D. Sherrington, J. Phys. A 16, L265 (1983» J. Phys.111§, 3691 (1983)- See for example, M. F. Thorpe in "Excitations in Disordered Systems," N.A.T.0. Advanced Study Institute Series 878, edited by M. F. Thorpe (Plenum, NY, 1982), p. 91. These kinds of identities were first considered by M. E. Fisher, J. Math. Phys. 4, 124 (1963). M. F. Thorpe and A. R. McGurn, Phys. Rev. B fig, 2142 (1979). P. H. Stillinger, Phys. Rev. 131, 2027 (1963). "t 18. 19. 20. 26. 27. 115 J. Stephenson, J. Math. Phys. 1, 1123 (1966). These results can also be derived by using the four equations: [ P(h) dh . 1 , f h tanh 8h P(h) dh . 3s , f h P(h) dh - 3m and f tanh 8h P(h) dh - m . See for example, L. Syozi in "Phase Transitions and Critical Phenomena" edited by C. Domb and M. S. Green (Academic Press, NY, 1972) Vol. 1, p° 269. See for example, 3. Stanley in "Introduction to Phase Transitions and Critical Phenomena” (Clarendon Press, Oxford 1971). J. Stephenson, J. Math. Phys. 2, 1009 (1964). T. C. Choy (to be published). R. B. Potts, Phys. Rev. §§_392 (1952) . These involve using the star-triangle transformation on the correlation functions, see J. H. Barry, C. H. Mfihera and T. Tanaka, Physica 1111’ 367 (1982). See for example, C. Kittel ”Introduction to Solid State Physics" (J. Wiley, NY, 1971), p. 506. This modified Brillouin function BS(x) differs from that usually used, which is s i2§_:_£l 3§_:;11. _ .1. .JL Bs(x) ZS coth [ 28 x ] ZS coth ( ZS ) , by including a prefactor of S, and a factor S in both arguments so that 85(x) 8 S-Bs(Sx). This saves having to write a lot of factors Si in subsequent formulae. D. N. Zubarev, Soviet Phys. Usp. 3, 320 (1960).. 116 Table Captions Table l The coefficients a:s defined in Eq. (39) are given for z 8 2, 3, 4, and 6. Table 2 o o n o a n ’+ a The inelastic neutron scattering intenSity per site S(k,m)/N at various special temperatures for the square net. The frequency is in units of J and the numbers in the table give the weights 2Wg/[l + exp(-3m)] of the five 3 delta functions. 117 _ _2 _ 02 .t m2 2 m2 _2 22 o s 2 m T c2 .7 - _ _ . o m s dl m MI 1 m T. — o N n d: c #1 ma .1 c N _ — -2n121 212w! - _ _2 _ _2 _2 m _s _ m _ _ _ _ c..- e o2 m. N- a SN 3 .VI 2 N c e .V on _ o d -s - :72..- .2... o u u .2 .2 _ a 1.7..- ,- . n n M233. c 118 d\T T 8 0 T 8 Tc T 8 m 8 1 0.6324 0.0625 4 0 0.2254 0.25 0 0 0.0849 0.375 -4 0 0.0387 0.25 -8. 0 0.0186 0.0625 Fig. Fig. H rig. Fig. 119 Eigure Captions For the honeycomb lattice, the bar graph at the left and the line graph in the upper right show the behavior of P(h) as a function of temperature. The ordinate gives the weight wh in the delta functions. The table in the lower right gives the value of wh (defined in Eq. 37) at special temperatures. Same as Fig. 1 except for the square net. The parameter x 8 l/w - l/rz. Same as Fig. 1 except for the triangular net. The results for T < Tc are not complete as we have not been able to compute the required odd spin correlation functions. The fluctuation v in the local field as a function of reduced temperature T/Tc for ferromagnetic interactions in the honeycomb, square net and triangular net. 120 a auawqm ; m _ .. m- 8.8.2. £315 0 m-.. s m m s o men-one» 38.2. o .-., . t r .00 memoirs. 38.8. o .3 81¢: 8.8.9. memoirs _ n; - . s -- o 8; up; one m m A m . .00 minute w S S w o a a o _ . is . no . a . «0 V. «N l O a m no \ - m mousse. A. W O m A .8 m const» A \. 121 a yuan—m .. s N o N- s- 88.8.3. 853.0183. o ...z s m a! o 8.9!. $828.3-.. 0 «-3 x m men-onus 33881-6 o o; . o .00 3.0.3. $82.01..-. 0 N3 8"»; ammo-0.3. mmvnmmdumkn . .i Isl: . . o 8i are c... s s . .80 o -00 2... o .00 . o ...,-o 8090.032. $005.0 0 P; 2800.an 0003.0 0 .7; 338.032.... 03.2.00 0 «-3 3.00.05 08000.0 0 0; 233.0303. 0330.0 0 «3 05000.33 0003.0 0 3.. 8090.049. «0003.0 . 03 8A or; 0.» m a»: 1 x. N. 0.0 0.0 . 0 .. N0 2 - «.0 I mwuflfl_b lull.ll i mw.flv weanlil Nun: ..... . I Our-Ill]. 1 mo _ _ _ « aupaqm Busts “812“. _m an m. _m _u . 1 m. use: crush - a L 0.0 use P mAu mAV mgu szn 0A0 nAu 123 q 323 _. uh: ca 0.. d d bwz mduaozfimh Ill ...mz wmwzor ........... > Appendix B. Some 2D Correlations In this appendix, some details are given for the calculation of correlation functions needed in Chapter 5. First the calculation of a four spin correlation on the square net will be discussed and then the second part of this appendix will discuss the one honey— comb correlation needed. As mentioned in Chapter 3, most of the square net correlations needed are found in Montroll, Potts, and Ward (1963), the exception being the four spin correlation. Rather than rederive or rewrite what has already appeared in MPW, their notation will be adopted here and then enough information will be presented so that ig_conjunction with MPW, the interested reader can calculate this particular correlation. MPW number the spins using rectangular coordinates so that in their notation, what is needed is ca 5 <012010021001> ' (1) It is useful to rewrite this as C4 = <012011011010021011011001> ' (2) In analogy to MPW equation (49), one obtains the form c = 24 P(Y‘1+Q) P(Y) (3) 4 where z = tanh(BJ) , (4) P(A) stands for the Pfaffian of matrix A, and Y and Q are matrices 124 125 which are defined below. It can easily be shown, as in MPW equation (50), that P(Y) = x4 (5) where x=z -z . (6) The matrix (Y-1+Q) can be written down almost directly. The columns and rows of the matrix are labelled in MPW's notation to aid the reader in seeing how this particular form came about: J _1_ 2 (14) Turning to the honeycomb lattice, one starts with the partition function (Green and Hurst 1964) JN°Z' (15) Z = C°[cosh3(BJ) or . Zn Z = In c + 3N2n cosh(BJ) + Zn 2' (16) where c is a constant and 2N 2n 4 J d6 I d¢ Rn{1 + 3y in Z' =-;%z O 0 8H- 2 2 - 2y (l-y )[cose + cosm + cos(6-¢)]} ; (17) y E tanh(BJ) . (18) Noting that, unlike the convention used in Chapter 5, Green and Hurst double count their bonds, 1 Bin 2' — tanh(BJ) +-§ 3(BJ) ' = tanh(BJ) +-l sech2(BJ) §&2_§_ . (19) 3 By One can show Bin 2' _ -l-J2fld8 [2"d¢ A + Bcose (20) By 81T2 O O C + Dcose + EsinB with 127 A = 12y3 (21) B = -3<4y - 8y3) (22) C = (1+3y4) - 2y2(1-y2)cos¢ (23) D = -2y2(1—y2)(1+cos¢) (24) E = -2y2(1-y2)sin¢ - . (25) Stephenson (1964) pointed out that TT _L_ exp(in8) _ n 2 _ 2 _ 2 -% . 2n [_nde c + Dcose + EsinB “ a (C D E ) ' (26) _1_ a 3 [(c2 - D2 - E2)2 - C1/(D - 1E) . (27) This relation is used to perform one of the integrals in equation (20), giving , ZTT _1_ an 2 L] de [A + BCD ] (CZ—DZ—E2)-2 By 4n 0 D2+E2 2n +A%;-[ d6 ED 2 (28) O D +E Now is expressed (through equations (28) and (19)) in terms of a pair of single integrals, a more suitable form for obtaining numerical results. Appendix C. Details of Monte Carlo Simulations The basis for the Monte Carlo simulations of the spin glass and ferromagnetic systems was the algorithm introduced by Metropolis et. al. (1953). Given a system at a temperature T, the algorithm generates an ensemble of states {Si}, which can be used to deter- mine the thermal average of properties of the system. From the ith member of an ensemble, Si’ the next member is generated by creating a new state St. In Ising systems this is typically done by flipping one spin in Si' Given Et = energy of St (1) E1 = energy of Si (2) AB = Et - Ei , (3) Si+l is defined in the following way: = st 1) AE < O + Si+1 2) AE > 0 Choose a random number p such that O < p < 1. t a) p < exp(—AE/kBT) + Si+l = S b) p > eXp(-AE/kBT) + 81+ S 1 i . The initial state, 81’ may be chosen at random or taken to be the last state generated by a simulation at a nearby temperature. In either case, one generally excludes some of the initial members: of the ensemble from the averaging process, which is equivalent to allowing the system to come into equilibrium before making 128 129 a measurement. In the simulations described below, a system change consists of flipping a single spin. A Monte Carlo step is defined by N attempted spin flips, where N is the number of spins in the system. Most of the rest of this appendix will describe the spin glass simulation, with some details of the ferromagnetic simulations appearing at the end. The Hamiltonian for the Ising Sherrington-Kirkpatrick model is H = - NIH Z J.,o.o. (4) izj lJ l J where as usual the superscript, z, is suppressed for convenience. p . th Let Oj— denote the value of OJ in the p state of the ensemble. Suppose a trial state is arrived at by flipping the ith spin so that its new value is O: = -05. Then AB = -Z J..o?(oF - oi) . l 1 J J J = 2( E J..oP) o? . lJ J l - P P ‘ Zhi Oi (5) At this point, the difficulties involved in doing a simulation of the SK model are apparent. First one has to have stored in an accessible way, the N(N-1)/2 independent values of Ji" Second one must be able to calculate hi, which has N-l terms, efficiently. Modifying an array packing scheme suggested by Jacobs and Rebbi (1981), these difficulties can be overcome to a degree. Neglecting the fact that the Jii's are zero, the information 130 describing this system is binary in nature: J.. = t1 13 09::1 1 The Jii = 0 problem is easily treated separately in the computer program. Rahter than taking an entire computer word, which is sixty bits on a Cyber 750 system, to store one value of, say, 0:, one can store sixty spin values in one word. These individual bits can be accessed and modified through such FORTRAN commands as SHIFT, OR, XOR, and AND. The last three correspond to the logical functions whose operations are defined below. AND(A,B) OR(A,B) XOR(A,B) 301 301 BAOl 000 001 001 (6) 101 111 110 The corresponding FORTRAN functions, operating on two words, perform these logical operations on a bit by bit basis. For example, AND(1,7) = 1 (7) because 7 = 1112 and 1 = 0012. Only in the least significant bit (LSB) do both words have a 1 so that 1 appears only in the LSB of the answer. In fact, a projection operator, projecting out the LSB of a word M, can be defined by P1(M) = AND(1,M) . (8) The SHIFT(M,Nb) function performs a cyclic permutation of the bits Of word M. Nb bits to the left. Hence one can define a projection operator for the Nth LSB of a word M by 131 PN(M) = AND(SHIFT(1,N-1),M) . (9) Similar use of the OR function will allow the definition of a par- ticular bit in a word. Hence each bit of the sixty bit word is both definable and accessible. The values of N spins can be stored in an array with N/60 words. This is why N was taken to be a multiple of 60 for the spin glass simulations. Similarly, one could store the N(N-1)/2 values of Jij in an array of approximately N2/12O words. However, for ease in calculation of the h? , the full matrix of Jij values was stored in an array of N2/60 words. A simple rescaling of the energy allows one to let a binary 0 represent —1 in the physical system. The calculation of the hi is achieved by noting that when Jij and 0? have the same value, 0? contributes positively to h: , while negative contributions arise when Jij and 0? have opposite values. Hence the product Jijog can be represented by the XOR function. Furthermore, the FORTRAN XOR function will perform these multipli- cations sixty at a time. What remains is to sum up over all j. Apparently there is no easy way of doing this, and so each bit of the product word is projected out separately and the results are added as integers. The memory limit of the Cyber 750, 127,000 words, is reached for an SK system of about 2400 spins. The practical computation limit, based on cpu time, is 2040 spins, while systems of 1020 spins were sufficiently large for most calculations. A typical measurement at one temperature involved averaging over four 1020 spin systems; that is, four different Jij matrices were generated. 132 Each system was brought into equilibrium by running for sixty Monte Carlo steps and then thermal averages were performed over two hundred steps. Such a run requires about 37,000 words of storage and five minutes of cpu time on the Cyber 750. Not all-of the advantages to using binary methods for simulating the infinite range spin glass model hold for simulating short range ferromagnetic systems. First, for uniform nearest neighbor systems, it is more efficient to devise a way of locating nearest neighbors than to store the entire Jij array. Secondly, the number of multi- plications involved in calculating 111 no longer scales with the system size. However, with long range order being important in ferromag- netic systems, it is useful to be able to work with larger systems, particularly near Tc' Hence the amount of central memory associated with storing the spin configurations becomes important. There is a trade-off between efficiency of use of storage space and ease with which spin values can be accessed. One compromise between these two optimizations is to utilize the sixty bit length of each word as an extra dimension. As an example, the spin con- figuration of a 3OX3OX30 simple cubic lattice was stored in an array of 3OX30 words, with each word in the array containing information about 30 spins. The chief difficulty in programming a Monte Carlo simulation of these ferromagnetic systems is finding an efficient way to locate the nearest neighbors of a spin. For a D dimensional simple cubic system, the method is intuitive. A spin located at (11.12.....1D> . 133 where iD refers to the bit number in word (i1,i2,...iD_1) , has nearest neighbors located at (ili1,iz,...,iD), (il,i211,...,iD) , ...,. A body centered cubic lattice, in three dimensions, can be represented by two interpenetrating simple cubic lattices. Locations on the BCC lattice are thus described by four integers: (L,i1,i2,i3) where L=1,2 is the sublattice number and (i1,12,i3) locates the position on that sublattice. With this decomposition it is straightforward to write out a table which describes the nearest neighbors of a given spin. Similarly, a face centered cubic structure can be decomposed into four simple cubic sublattices and‘ a diamond structure into eight. The nearest neighbor locations which are obtained in this decomposition are shown in Table 2. The periodic boundary conditions used in these simulations are imposed with the aid of modular arithmetic. Typical runs on ferromagnetic systems began with 1000 Monte Carlo steps to bring the system into equilibrium. This was followed by 1000 steps (2000 steps at TC) during which thermal averages were computed. 134 Table 2. Nearest neighbor locations for three dimensional lattices. Site (1,i,j,k) (2.i.j.k) Site (l.i.j.k) (2,i,j,k) (3,i,j,k) (4,i,j,k) Site (1:1!jak) (2,i,j,k) (3,i,j,k) (4,i,j,k) (5.1.j.k) (6,i,j,k) (7,i,j.k) (8,i,j,k) BCC Nearest Neighbors (2 .k)(211.Jk)(2111k)(2 j.k-1) (2 ,j-—1,k) (2, i- -1 ,j, k- 1) (2, i, j- *1 fl) (2,i—1,j-1,k-1) (1,1, j,k) (l, i+1, j, k) (1, i, j+1, k) (1, i, j,k+1) (1, i+1, j+1, k) (1, i+l ,j, k+1) (L i, j+1, k+1) (1, 1+1 ,j+1, k+1) FCC Nearest Neighbors (2,i,j,k) 2.i-l.j.k) (2.1.3.k-l) (2.1-1 j,k- l) (Byisjyk) (39i-19jak) (3,i’j“lvk) (3’i-11j-12k) (Avivjvk) (Arivj’l’k) (4,i,j9k-l) (4,i9j- “lik 1) ,i,j,k+1) (1,1 ,i,j,k+l) (3,1 aivj-10k) (Aai ',j,k) (l,i+1,j,k) +1 ,j, k+1) ' ' '1,j-1,k+1) +1 ,j— —1 ,k) (l,1,j,k) (1,i+1,j,k) (1,i,j+l,k) (l,i+1,j+l,k) (2,i,j,k) (2,i,j+19k) (Ztlijik-l) (2,i,j+19k-1) (4,i,j,k) (401+19j0k) (4919J9k'1) (491+19j9k‘1) (1,i,j,k) (1,i,j+l,k) (1,i,j,k+1) (1,1, j+L k+1) (2,i,j,k) (2,i- 1, j,k) (2,i,j+l,k) (2.1-1, j+1, k) (3,i,j,k) (3,i-1,j,k) (3,i,j,k+l) (3,i-1,j,k+1) DIAMOND Nearest Neighbors (S,i,j-1,k) (6,i-1,j-1,k-1) (7,i-1,j-1,k) (8,i,j—l,k-1) (5,i,j-1,k) (6,i,j-1,k) (7,i,j-1,k) (8,i,j-1,k) (S,i,j,k) (6,i,j,k-1) (7,i,j-1,k) (8,i,j-1,k-1) (S,i,j,k) (6,i-1,j,k) (7,i-1,j-1,k) (8,i,j-1,k) (1,i,j+l,k) (2,i,j+l,k) (3,i,j,k) (4,i,j,k) (l,i+1,j+1,k+1) (2,i,j+l,k) (3,i,j,k+1) (4,i+l,j,k) (1,i+1,j+1,k) (2,i,j+l,k) (3,i,j+l,k) (4,i+1,j+1,k) (1,i,j+l,k+l) (2,i,j+l,k) (3,i,j+1,k+1) (4,i,j+1,k) LIST OF REFERENCES LIST OF REFERENCES Bantilan F T and Palmer R G 1981 J. Phys. F.ll 261 Barry J H, Mfinera C H, and Tanaka T 1982 Physica 113A 367 Berlin T H and Kac M 1952 Phys. Rev. §§_821 Bethe H A 1935 Proc. R. Soc. London Ser. A 15_ 552 Binder K 1977 Z. Physik B 26 339 Bonner J and Fisher M 1964 Phys. Rev. $355 640 Burington R S 1946 Handbook 9f_Mathematical Tables and Formulas (Sandusky: Handbook Publishers, Inc.) Callen H B 1963 Phys. Lett. 4_161 Choy T C 1984 submitted to J. Math. Phys. Choy T C and Sherrington D 1983a J. Phys. A‘lg L265 Choy T C and Sherrington D 1983b J. Phys. A.16 3691 Choy T C, Sherrington D, Thomsen M, and Thorpe M F 1984 to be published Domb C 1960 Adv. Phys. 2_251 Domb C 1974 "Ising Model" in Phase Transitions and Critical Phenomena vol. 3, C Domb and M S Green eds. (New York: Academic Press) Edwards S F and Anderson P W 1975 J. Phys. F.§ 965 Eggarter T P 1974 Phys. Rev. Big 2989 Fisher M E 1963 J. Math. Phys. 4 124 Frank B, Cheung C Y, and Mouritsen O G 1982 J. Phys. C.14 1233 Gaunt D S, Sykes M F , and McKenzie S 1979 J. Phys. A l2_871 Green H S and Hurst C A 1964 Order-Disorder Phenomena (New York: John Wiley and Sons) 135 136 Grest G S and Rajagopal A K 1974 J. Math. Phys. l§_589 Hurst C A 1963 J. Chem. Phys. 38 2558 Ising E 1925 Z. Physik §1_253 Jacobs L and Rebbi C 1981 J. Comp. Phys. 4; 203 Katsura S 1962 Phys. Rev. 121 1508 Khanna S N and Sherrington D 1980 Solid State Comm. 36 653 Kirkpatrick S and Sherrington D 1978 Phys. Rev. B 11 4384 Kirkpatrick S, Walker L R, and Walstedt R E 1977 Phys. Rev. Lett. 38 514 ‘ Kittel C 1971 Introduction £2_Solid State Physics (New York: John Wiley and Sons) Klein M W and Brout R 1963 Phys. Rev. 132 2412 Kosterlitz J M, Thouless D J, and Jones R C 1976 Phys. Rev. Lett. 36 1217 Kramers H A and Wannier G H 1941 Phys. Rev. §Q_252 Krey U 1981 Z. Physik 3.3g 231 Lee D H, Joannopoulos J D, and Negele J W 1984 Phys. Rev. Lett. 52 433 Lieb E, Schultz T, and Mattis D 1961 Annals of Physics l§_466 Marshall W 1960 Phys. Rev. 118 1520 Marshall W and Lovesey S W 1971 Theory 2: Thermal Neutron Scattering (Oxford: Oxford University Press) Matsubara F and Katsura S 1973 Prog. Theor. Phys. 49 367 Matsubara F and Yoshimura K 1973 Prog. Theor. Phys. §Q_1824 Matsubara F, Yoshimura K, and Katsura S 1973 J. Can. Phys. §l_1053 McLenaghan J R and Sherrington D 1984 to be published Mess K W, Lagendijk E, Curtis D A, and Huiskamp W J 1967 Physica.§4_126 137 Metropolis N, Rosenbluth A W, Rosenbluth M N, Teller A H, and Teller E 1953 J. Chem. Phys. 21 1087 Mézard M, Parisi G, Sourlas N, Toulouse G, and Virasoro M 1984 Phys. Rev. Lett. §2_1156 Montroll E W, Potts R B, and Ward J C 1963 J. Math. Phys. 4 308 MUhlschlegel and Zittartz H 1963 Z. Physik 11; 553 Naya S 1954 Prog. Theor. Phys. _1 53 Onsager L 1944 Phys. Rev. §§_117 Palmer R G and Pond C M 1979 J. Phys. F 2 1451 Parisi G 1980a J. Phys. A 1§_L115 Parisi G 1980b J. Phys. A.13 1101 Parisi G 1980c J. Phys. A.£3 1887 Parisi G 1983 Phys. Rev. Lett. 29 1946 Peierls R E 1936 Proc. Cam. Phil. Soc. §2_477 Potts R B 1952 Phys. Rev. §§_392 Schiff L 1968 Quantum Mechanics (New York: McGraw—Hill) Sherrington D 1975 A.I.P. Conf. Proc. 22 224 Sherrington D and Kirkpatrick S 1975 Phys. Rev. Lett. 35 1792 Southern B W 1976 J. Phys. C 2_4011 Stanley H E 1971 Introduction to Phase Transitions and Critical Phenomena (Oxford: Oxford University Press) Stephenson J 1964 J. Math. Phys. §_1009 Stephenson J 1966 J. Math. Phys..Z 1123 Stillinger F H 1963 Phys Rev 131 2027 Suzuki M, Tsujiyama B, and Katsura S 1967 J. Math. Phys. §_124 Sykes M F 1979 J. Phys. A 12_879 138 Syozi I 1972 in Phase Transitions and Critical Phenomena vol. 1 C Domb and M S Green,eds. (New York: Academic Press) Takeda K, Matsukawa S, and Haseda T 1971 J. Phys. Soc. Jap. 39 1330 Thomsen M and Thorpe M F 1983 J. Phys. C.1§ 4191 Thomsen M, Thorpe M F, Choy T C, and Sherrington D 1984 to be published (in Phys. Rev. B) Thorpe M F 1982 in Excitations in_Disordered Systems NATO Advanced Study Institute Series B78, M F Thorpe,ed. (New York: Plenum) Thorpe M F and McGurn A R 1979 Phys. Rev. 3.29 2142 Thorpe M F and Miyazima S 1981 Phys. Rev. B 24 6686 Thorpe M F and Thomsen M 1983 J. Phys. C 1§_L237 Thouless D J , Anderson P W, and Palmer R G 1977 Phil. Mag. 35 593 Walker L R and Walstedt R E 1980 Phys Rev B 22 3816 White R M 1983 Quantum Theory of Magnetism (New York: Springer- Verlag) Wright J C, Moos H W. Colwell J H, Mangum B W, and Thornton D D 1971 Phys Rev B.3 843 Zubarev D N 1960 Soviet Phys. Usp. 320