PLACE IN RETURN BOX to move thIo checkout from your record. TO AVOID FINES return on or balm duo due. DATE DUE DATE DUE DATE DUE W .1 l 'n:: 7 ’. fj- » 2 6 7 .. m msozaoz I MSU Is An Affirmdivo Action/Equal Opponunity IMMIOI'I VIBRATIONAL DYNAMICS OF DISORDERED DIPOLAR AND IONIC SOLIDS AND LAYERED SILICATES BY Joel Moritz Gales A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requireuents for the degree of DOCTOR OF PHILOSOPHY Departnent of Physics and Astronouy 1989 ABSTRACT VIBRATIONAL DYNANICS OF DISORDERED DIPOLAR AND IONIC SOLIDS AND LAYERED SILICATES By Joel Moritz Gales In this thesis we study the vibrational dynamics of two classes of materials. In the first part we look at disordered solids in which long-range Coulomb forces are present. We first present the basic features of these forces in crystalline materials, in particular, the LO-TO splitting, and the techniques used with special emphasis on the Ewald method. The computational technique, the Equation-of-Hotion (EON) method is described which we apply to a study of depleted simple cubic arrays of point dipoles. We show that a trucated approach is numerically equivalent to the exact Ewald method but computationally simpler and faster. The main results of this section are given in a study of the LO-TO splitting in (III-V) IV mixed crystals semiconductors. Using both Raman and infrared responses, we show that this splitting vanishes when the III-V ionic material no longer percolates. We demonstrate this for both a one- and two-mode behavior material and compare our results with recent experimental data on these systels. In the second part, we present a theoretical survey of the lattice dynamics of layered silicates or clay. We present the zone center frequencies, the phonon dispersion curves and the elastic behavior. We derive expression for the low-frequency zone center modes that should be useful in the extraction of force constant values from experimental data. We also study the torsional mode dynamics of ternary intercalated alloys using the EON method and develop a simple theory for the frequency as a function of varying concentration that incorporates the effects of layer rigidity. DEDICATION To my parents, who listened to it all. 11 ACKNOWLEDGMENTS I would like to thank my advisors, Prof. H.F. Thorpe and Prof. S.D. Hahanti for their guidance and. support and for our mutual perseverance through sometimes difficult periods. I am grateful to Prof. S.A. Solin, Prof. H. Repko, and Prof. H. Abolins for their service on my guidance coumittee. For his careful presentation, both in person and in print, of the somewhat subtle features of the Coulomb interaction, I would like to thank Dr. 3.". deLeeuw and for her assistance in tasks both large and small throughout my time here, I thank Hs. Janet King. Finally, for their friendship, humor, and refuge, I wish to thank Roy Day, Rich Kennan, and Dr. G. Pollack. iii TABLE OF CONTENTS LIST OF FIGURESOOOOOOOOOOOOOOOO ...... OOOOOOOOOOOOOOOOOIOV1 LIST OF TABLESOIOOOOIOOODOOOIOIDIIOIOO‘OIOOOICOOOOIIIIIOix MI. I. IntrOductionOOOCOOOOOOOOOOOOIOOOOOO‘OO0.0.0.00002 II. Equation of Motion for Dipolar Systems.........6 III. Dynamics of Periodic Dipolar Arrays A. Ewald Method for Periodic Lattices...... ....... 9 E. Zone Center Modes - Depolarizing Factors......l7 IV. Dynamics of Disordered Dipolar Systems A. IntrOductionllIOOOOIOOOOIOOOOI0.00.00.00.0000039 B. Equation of Motion / Linear Response Theory...43 C. Implementation fo the Coulomb Interaction.....52 DO Effects Of DisorderOOOOO...0.00.00.00.0000000062 V. Coulomb Effects in (III-V)bm IvgxiMixed Crystals A. IntrOductionOOOOOOOOOO0.0.00.0...0.0.0.000000073 8. Implementation of Coulomb Interaction for Ionic syste.80000.0.0.0000...0.0.0.0600000000077 C. Equation of Motion / Simulation Techniques....80 D. Construction of Mixed Crystal - Percolation ' Of DinerBOOOOO00.00....000......0.0.00.000000086 E. One-Mode Behavior (GaAs) Ge..................92 F. Two-Mode Behavior (InSb) Ge.. ...... ..........98 VI.‘ Summary and Conclusions......................107 iv APPENDICES A. Fourier Sums for Three-Dimensional Periodic Functional.OOOOOOOOOOOOOOOO00.0.0000000000000109 B. Relation between the L0 and TO Responses for Mixed Crystals...........................112 £11111 I. Introduction.......................... ...... .116 II. Crystal Structure............................120 III. Phenomenological Force Constant Models.......126 IV. Lattice Dynamics of the Periodic Model System A. Zone Center Modes............................l32 B. Phonon Dispersion Curves.....................l4l C. Elastic Behavior.............................lSl V. Torsional Mode Dynamics in Ternary Intercalant Alloys Introduction.................................160 Description of Model.........................162 Torsional Mode Frequency / Layer Rigidity....165 Simulation Results...........................l76 cocoa» REFERENCES.............................................l77 2(a) 2(b) 2(c) 5(a) 5(b) 5(c) 10 11 LIST OF FIGURES 2111.1. Model System for Dynamical Calculations ........ 25 of Ionic Systems Energy Dispersion for SC Dipole Lattice........30 Energy Dispersion for 300 Dipole Lattice.......31 Energy Dispersion for FCC Dipole Lattice.......32 Brillouin Zones for Cubic Lattices..... ........ 33 Dispersion for Simple Cubic (100) using........34 Truncation and Ewald Methods Density of States for SC Dipole Lattice........36 Density of States for BCC Dipole Lattice.... ..36 Density of States for FCC Dipole Lattice.......37 '003 for Disordered Simple Cubic (p=0.9)........59 Im 61(3) for Disordered Simple Cubic (p=0.9)...59 Im II Im Im :.(E) for Disordered Simple Cubic (p=0.9)...60 41(E) for Disordered Simple Cubic (p=0.9)...60 ct(E,E) for Disordered Simple Cubic.........61 (p=0.9) f=<2n/L) k for Disordered Simple Cubic.........61 (p=0.9) I=<2n/2L) i vi 12(a) 12(b) 12(c) 13 14 15 16 17 18 19 20 21 22 23 24(a) 24(b) 25 26(a) 26(b) DOS for Disordered SC Lattice............. DOS for Disordered BCC Lattice............ DOS for Disordered FCC Lattice............ DOS for Disordered Cubic Lattices (p=0.05) and Random Dipolar Gas DOS for Dipolar Gas with Lorentzian Tails. from Dipole Pairs DOS for Quenched Dipolar Liquid (Pick)... Conventional unit cell of zincblende..... structure Percolation Statistics for Dimers on Diamond. (III-V) IV "Mixed Crystal" on Hexagonal. Lattice Raman Response for (GaAs) Ge............ Peak Positions of Raman Response for.... (GaAs) Ge Infrared Response for (GaAs) Ge......... Raman Response for (InSb) Ge............ Raman Measurement for (GaSb) Ge (Expt.) Peak Positions of Raman Response........ (GaAs) Ge (Expt.) Peak Positions of Raman Response........ (InSb) Ge (Theory) Infrared Response for (InSb) Ge......... Longitudinal Response EOM vs KR (x=0.2) Longitudinal Response EOM vs KK (x=0.8) vii ..63 ..67 ..67 .102 .103 .104 .105 .114 .114 3(a) 3(b) 6(a) 6(b) 8(a) 8(b) 10(a) 10(b) 11(a) 11(b) 12 13 2.112111 Structure of Model Layered Silicate........ Low Frequency Zone Center Modes............ w(f) for I (0,0,k)....................... “(3) for E (0’k,0)oeoooeoooaaoomoooaoaaao Structure for y-axis Dispersion Model...... Density of States of Model Layered Silicate. Sound Velocity as calculated from Model.... Sound Velocity for Muscovite as measured... by Brillioun scattering w(i) for f = (0,k,0) for 31205 Layer........ Raman Measurements of Torsional Mode....... Torsional Mode Frequency vs x (Expt.)..... Two-Dimensional Kagomé-Intercalated Lattice Infinitely Rigid Clay Layer................. Floppy Clay Layer.......................... Torsional Mode Frequency vs x (Theory).... (Sub-linear Behavior) Torsional Mode Frequency vs x (Super-linear Behavior) (Theory).... Raman Response for Torsional Mode.......... calculated with EOM Density of States for Kagomé-IC calculated. via Direct Diagonalization and EOM viii ..121 ..133 ..144 O .145 ..147 ..150 ..154 ..154 ..159 ..161 O .161 ..164 ..169 ..169 ..171 ..172 ..174 ..175 LIST OF TABLES Table 1 Structural parameters of the model clay.......123 Table 2 Phenomological Force Constant Values..........131 ix WWQW mmmm mem In their paper "Coulomb Effects in Disordered Solids", Thorpe and de Leeuw (Th 86) developed a general technique for computing the vibrational response functions of disordered solids (such as glasses and mixed-crystal alloys) in the presence of long-range Coulomb forces. They derived moment sum rules for the transverse and longitudinal dielectric responses, a generalized LST (Ly 41) relation for the static and high-frequency dielectric constant, and most importantly, a general relation between the transverse and longitudinal responses allowing one to be calculated once the other is known. They illustrated these results by computing the response functions for a model AX: glass roughly corresponding to vitreous silica. These simulations showed, unambiguously, the existence of longitudinal optic - transverse optic (LO-TO) splitting in disordered materials for which there had been experimental evidence for nearly a decade (Ga 76, Ga 83). It had been hoped initially that these dielectric responses could be calculated theoretically, albeit approximately, as opposed to being merely computed. However, the long-range nature of the interactions coupled with the disordered nature of the materials makes this a difficult problem and we have been unable to make much progress in this regard. We must content ourselves, therefore, with results essentially empirical and descriptive in nature. We'll begin with a look at a prototypical Coulombic system which displays the characteristic longitudinal optic/transverse optic splitting of ionic systems, and provides a good introduction to the techniques used in studying these materials. We study crystalline cubic lattices of Einstein oscillators coupled via long-range dipole interactions. For these purely "optic mode” systems, we introduce the Ewald method, which transforms the slowly convergent dipolar interaction term (V(r) + l/r') into an equivalent, rapidly convergent, expression. In addition, it demonstrates explicitly the characteristic splitting near the zone center (i=0). We'll next show how the macroscopic shape of the sample can be incorporated through the depolarizing factors and reveal the connection between the longitudinal and transverse modes at small wavevector and the response of the system, of a given shape, at the zone center. Dispersion curves, w(E) vs. E, will be presented for the three cubic lattices along with the vibrational density of states. With this foundation established we can consider the effects of disorder. In this part, numerical simulations play a major role. We use the equation-of-motion (EOM) method for virtually all computations on the disordered materials. This technique computes the response of a system to various external probes and will be used primarily to calculate the density of states (DOS) and dielectric (infrared) response (Th 86). We will see how the Coulomb forces can be implemented once again using the Ewald method and introduce a simpler, more physically intuitive approach that gives essentially the same results as the Ewald method but is faster and easier to implement (Ga 89a). We use this technique to study the effects of disorder on the DOS of point dipole systems, by considering depleted cubic lattices. This section culminates with a look at a more realistic ionic system, that of a model (III-V) IV mixed crystal semiconductor. For x=0, we have a crystalline. partially ionic, zincblende structure with well-defined zone center splitting of the optic modes, for x=l, a purely covalent material. We will compute the dielectric (IR) and Raman response for two systems, one exhibiting one-mode behavior based on (GaAs)‘__xGe2x and the other with two-mode corresponding roughly to (GaSb)t«Ge and show that the 2x’ LO-TO splitting vanishes not at x=1 when the effective charge disappears but rather when the (III-V) material no longer percolates. Thus we see that a local, nearest-neighbor property can have dramatic consequences for an effect usually associated with, and indeed dependent on, the presence of long-range forces. mwummmm Consider first an isolated pair of particles with equal and opposite charges, q. They are coupled together by a Hooke spring with force constant k. We neglect explicitly the Coulomb interaction between the two charges, however the "short-ranged" part of this force can be considered to be included in the harmonic potential. Fixing the center of mass in space to prevent translational motion, the Lagrangian of the system is given by: ) , (1) where 3+ and 3_ are the displacements from equilibrium of the two charges. From the center of mass condition, m+ap+ m_3;=0, and the expression for the dipole moment, a = q (3 - 3_), we can re-write eqn. (1) in terms of the f more convenient dynamical variable 3: _ L m 42 _ ‘1 k +2 ‘- - 2 Ir)“. 2 [flux ‘2’ where m is the reduced mass given by, m+m;/(m++ m_). We now add the dipolar interaction between this dipole, the 1th, and all the others in the system, which we consider to be identical. This is given by (Ja 74): + + + + P.‘ u. 3 (pa 3..) (us 3..) vaa(1)= E [ " ’ ‘ "’ ’ "] (3) 8 - 5 . . R.. Jflt. t.) where Ru is the displacement vector between the 1th and jth dipole. Adding this term to the Lagrangian in eqn. (2) and using the Euler-Lagrange equation (Ma 70): d 0 L 0 L __ _. --—=0, (4) dt 0 z 0 3 we get the equation of motion: .. qz z! 3 (:13 3..) N” 9 z 9 J J ‘J U u... p.._Z _. =0, (5) t. o I. I 9 8 .. R.. R.. 33“. 1.3 I.) where ab: = k/m. Because of the linear nature of this equation, we have harmonic solutions of the form: fist) = Bieiwt. Substituting this into eqn. (5) we get: (wf- mi) 3 = Q,3 where D, is a matrix containing the .’ structural information and p = (p‘x,p‘y,uu, ,HN*,HNV,HN8). Since Q,is independent of mo, so are its eigenvalues and thus changing w: merely shifts the origin of the spectrum when expressed as a function of 03. Furthermore, by introducing the plasma frequency, defined by: N z N q, a d s z _ 4 n i = 4n9 L_ ‘L 4mg we- 02,-, 0 2[m++m_] 8 mp (6) i. i where Np(Nd) is the number of particles (dipoles) in the volume 0 and p is the number density of dipoles, we can write eqn. (5) as: z 9 9 fi.+w’fi.+ " ——’—- ’ " " =0 (7) t. O t. 4719 3 R5 13“. ij ij . The term in brackets in eqns. (5) and (7) is (minus) the electric field at the 1th dipole due to the Jth dipole located a distance Ru from 1 along the unit vector rn= RH/R . We can therefore write eqn (7) as: t: L: t1 2 9 2 9 up 4' (o - u. 0 Hi E(R.) (7a) 1. I. 4np where 3(3) is the total electric field at the 1th dipole due to all the other dipoles. mmuwmm memmm Periodic arrays of point dipoles are an important special case of the harmonically coupled systems we have discussed above. Rather than dealing with an interaction matrix D of dimension 3Nd, we need only consider a matrix of dimension 3s, where s is the number of dipoles per unit cell, that is, the number of elements in the smallest periodically repeated unit making up the system. In this section we will consider only Bravias lattices, where s=1, and will develop the Ewald method appropiate to this case. We return to eqn. (3), assuming there is a dipole at every lattice point defined by the set of direct lattice vectors, R. The interaction energy of a given point dipole, Do, with this periodic lattice of point dipoles is given by: F___o-R 3d) 3(30- fluid!)- i!) “(140) :32 - ] (a) where the sum is over all direct lattice vectors and R a IRI. Because of the periodicity we consider solutions iE'R of the form: 3(R)=3°e , therefore: 10 (9) As it stands, the sum converges too slowly to be of practical use. The Ewald technique re-expresses the single sum as two rapidly-convergent sums, one in real-space, the other in reciprocal space. We start the derivation with the integral representations: 1m 1 2 ix: “-83% .3 = — I dt t e (10a) R f; o 1 4 an axe -I2t. -3 = -—— J'dt t e (10b) R an”? 0 which follow from the substitution u = Rt‘flz. Substituting these relations into eqn. (9), we get the two terms: *3 as «It a.» 11 2“: * R m = ——‘—’- 2 e‘k' In: t‘” e “ (11a) a Fee 0 I {II-R 3(“0' it) I a —2 e s R5 Rec 4 9 3 m 2 = - — 2 ed" (30- i!)2 In r’” e"" ‘ (11b) {5 iso ° 0 4 . 9 . _ a = — (30° (’82 2 e‘(“?) if Int t’” e “ (11c) {E Karo ° ?-o We split the integrations into two parts, from O to n2 and n2 to at where n is a freely adjustable parameter. We will deal with the four resulting terms one at a time beginning with 1;” where: s .. _2 o 2 e‘kij‘dt ti/se mi (12) o and 12 sz 9 R 1m e 1;” 5 -——° 2 ea. In t": e" ‘ (13) ' a 3:0 n2 . With the substitution u E R t‘AL the integral in (13) can be written as: Q 2 z -u2 -s am.)2 1’1? —; Idu u e = R nR e + -§-erfc(nR) (14) R 0‘ where we have integrated by parts using a = u and 2 dt = u e“u du. Thus we have: 2:42 eff-i! z 1‘” -..- —-2 2 mt e"”" + E erfc(nR) (15) ° 4.? n' 2 3:10 Because both terms in the brackets go exponentially to zero as R-90, this term is rapidly convergent. We now consider (1) I. and begin by interchanging the order of the sum and integral. This gives: a 3 n 2p ,9. _ 2 1:95 o I“: ti/zz .ikRent n o l3 s 2 2 n 2 n 2p ,9. . _ 2 2p = o I“: ti/z eik R e mi __ o Idt tax: (16) f n o f n o where we have added the 380 term to the sum, and compensated with the second term. The sum in the first term is 9 9 unchanged when k -9 k + 3 where C is a reciprocal lattice vector and can therefore be expressed as a three dimensional Fourier sum in reciprocal space. This gives (see Appendix A): 8/2 9 gait-R e-Izt ___ :7 [g] ge'm’z'zl“ (17) where the 3's are the reciprocal lattice vectors and u is the volume per dipole in real space and thus equal to the inverse of the number density. From eqn. (16) then: 2 s n 2 Zap _ 9_ 2 4p 1;» z o o I“; e Ik GI /at _ o T). (18) o 34; . As we shall see shortly, the first term cancels with an identical term from T;” which we now evaluate. 14 2 n 2 T“) = :1 dt t3/2 ”9‘ _ V )2 et(:+?)-§ e-m t. 5 fi 0 0 { i=0 s n 9 9 2 ’0 1 -lk’-Cl2/at n2 _ 2 1 -|E-C|'/at 21.10 ; Iodt t— e (19) where we have applied eqn. (17) to the real space sum. Thus: - 2 Ico+ T“) = 42 o e (k/an) + 9 9 2 4—32 It‘é‘k‘a’“ e_|f_a|2/‘n2 o |g_a|2 -54" n g (20) :1 lot» m duo where we have separated the irregular 3-0 term from the sum. As 3 goes to zero, we see that the value of this term depends on how this limit is taken, as it is a function of the angle between 30 and E. For the longitudinal mode 15 (30 | E), this term gives 4nu2/o, while for the transverse mode (30 1.3)! it gives zero. Thus we have explicitly demonstrated the zone center LO-TO splitting. The ambiguity in the limit is a consequence of the long-range nature of these forces which requires that the boundary shape be specified to completely define the problem. This is not surprising when we recall that problems in electrostatics are essentially boundary-value problems. We shall see shortly how the sample shape can be incorporated in the Ewald sum through the use of convergence factors. (2) We return now to the evaluation of T5 . From eqn. (11b): (2) 4 9 2 tk’R O s/2 -mzt. T5 = - —-- 2040' R) e Idt t e . (21) f n a’o ”2 The integral can be evaluated once again through the substitution u R t}/2 and integrating by parts using 2 s I u' and dt 8 u efi‘du. This gives finally: 2 9 2 1..., 1...: _2 2 a” { [52 _ 3W. ‘5’] D B I 8 *fi'i—o R R 16 3 x [72R e-mm + £3 erfc(nR)] } - (i5 - fi)’ (22) Equations (20) and (22) taken together give the Ewald expression for the dipolar interaction term. It is valid for non-zero E but as stated earlier, remains undefined for E = 0. The adjustable parameter n controls the convergence of the real- and reciprocal-space terms. For large n, the convergence of the first term is increased while that of the latter is decreased. For small n, the reverse is true. Its numerical value is chosen to optimize the total convergence. For the calculations on the cubic arrays, which we present in the next section, we use 7; I 2.0 a“, where a is the linear dimension of the conventional unit cell. Before presenting results for these three lattices, the simple (SC), body-centered (BCC), and face-centered cubic (FCC), we first discuss the procedure for handling the zone center (E = 0) modes. l7 al.22ns.antst.hsdea.:.anelatizins.tastsrs. We begin the analysis of the zone center modes with the modified interaction: 2 9 s M 3m - R) .9 O 0 V (p 'E=O) = lim -—- — x dd 0’ s90 :2 [R' R. ] a: a2 3: exp -s —;’+ -%’+ —; (23) a b c where the second term is an ellipsoidal convergence factor that imposes a "shape" onto the sum through a directionally dependent, exponentially decaying, weighting factor. The rigorous justification for the mathematical form of this factor is given in de Leeuw et a1. (dL 80) where it is shown that the infinite sum in eqn. (23) is equivalent to a finite sum over the supercells with a surface defined by s[(x/a)z + (y/b)z + (2/c)2] = 1. We obtain the desired expression for infinite sample size in the limit of the scale factor s, going to zero. Proceeding as before, we use the integral representations for inverse distances, eqns. (10), and split the integrals into two parts. We need only concentrate on the evaluation of the integral with limits from 0 to n2 as the other term simply lead to eqn. (22) in the limit I 9 0, once the limit s 9 0 is taken. Evaluation of T;” for this 18 case, leads to an expression similar to eqn. (16) but with the sum over R replaced by: _ s a _ s a _ s a gexp[ (t + is)Rx]exp[ (t 4- Ss)Ry]exp[ (t + EHRZ] (24) This sum can be re-expressed in terms of a three-dimensional Fourier sum. From Appendix A, part II, we get for the a = 0 term: 3/2 T! (25) (t + £2)" (t + fie)" (t + gn‘” We note that for finite C, in the limit s 9 0, we recover eqn. (17) with the C = 0 term excluded and with E set to zero. Equation (25) gives: tn _ Ia (8s0) - 1 2n ' n ta/s 11' 7E0 I dt 2 4/2 s 1/2 a 1/2 (26) s90 o (t + :s) (t + 52) (t + 3') Note that if the limit is taken, this leads to a logrithmic divergence. This cancels with an identical divergence of opposite sign from the term T;”(C=O) which is given by: 19 (u _ _ 1's (3-0) - 2 fl a/s 4n t lim —— dt x .90 0° (t+:‘)s/s(t+%.)u2(t*%.)s/2 2 9 . 2 -( /4(t+s) (yo Vz) e =0 (27) The second term of the integrand gives: #2 u' #2 _—_xs 4» __L_s + ___2‘ (28) (t+;¢) (ti-~52) (ti-Es) Adding eqns. (26) and (27) gives finally: 3 9 9 4n 2 vuwozk-OJ-O) = —; find “a (29) all (a = x.y.2) with 2 n 1/2 abc t Bx B 11' . 2 I“; 2 ex: 2 V2 2 v2 s90 o (a tts) (b t+s) (c t+s) 20 a) abc 4“ 3 I 2 9/2 2 1/2 2 us ( 3 O ) o (a +u) (b +u) (c +u) where we have used the substitution t a s/u. Similar expressions exist for the other depolarizing factors, By and Bz. The expression in eqn. (29) is the corresponding term, for the zone center modes (E=0), of the first term in eqn. (20) for finite E. For the case a=béc (spherical geometry), it is easily shown that Bx = By 8 B2 = $9 whereas in slab geometry with the slab face perpendicular to the 2-axis (a,b 9 m), we have Bx = By = 0 and B2 = 1. Thus by comparing eqn. (29) with the first term in eqn. (20) in the limit E 9 0, we see that the longitudinal mode corresponds to a uniformly polarized array with 30 perpendicular to the slab face and the transverse mode, to an array polarized parallel to the slab. Therefore, when E = 0 and hence the difference between longitudinal and transverse is undefined, we can then refer to the direction defined by the normal to the slab face in order to re-establish this distinction. For the cubic lattices, in the case of spherical geometry, one can show that the interaction energy, eqn. (23), vanishes for the case of E = 0. To do so, we first note that the lattice vectors in these structures are given by: 21 8.1 R = a (l,m,n) l,m,n are integers with no restrictions (31) ESE. R = ; (l,m,n) l,m,n all even OR all odd (32) ESL R = % (l,m,n) l+m+n even a all even OR 1 even, 2 odd (33) where a is the conventional unit cell length. We consider the ordered triple (l,m,n), all non-negative, and its various permutations and sign changes, which we designate by <1mn>. The set of lattice points so defined forms a spherical shell a distance R = (12 + m:I +n2)1/2 Ro from the origin, where R0 is a for the SC and a/2 for the BCC and FCC. (We neglect the case where l=m=n=0.) If we represent 30 by (x,y,2), then the interaction energy of the central dipole with this shell is given by: V = 2 xz+ yzi- 22 _ 3(xl‘t’Lm‘I-zn)a shell. (lam) R9 R5 22 22 22 = -l; 2;» (xz+73+sz) - 3‘” 1 +y ' :2 n ) R <1 (R/Ro) 22 .. ——6——2' 2 (xy 1m 4- xz 1n 4- ya mn) (34) (R/R ) (l n) o 0 Now: (12/30)” = (1’+-’+n’) .. 3 (1' or e’ or n') (35) (1 n) (l n) (I > where the last expression follows from the cubic symmetry. Thus we see that the first sum in curly brackets in eqn. (34) vanishes. In addition, a typical term in the second sum can be written as: x7 2 1m =2 [1m + l(-m)] = 0 (36) < n) (Inn) Thus the second sum is also zero. Since the entire lattice can be summed up in this manner, we have demonstrated that for a spherical cubic symmetric lattice, the dipolar interaction energy is zero. Because this corresponds to the case where Bx = B = B = - in eqn. (29), we have shown y 2 a’ that: 9 4n 2 V(k=0) = 53% + Z[f=o] sphere 3:0 2 u + Z[E=o] - in. = 0 (37s) n in “I. 23 where the middle two terms signify the reciprocal and real space (shape-independent) sums in eqns. (20) S (22) evaluated at i=0. Thus from the previous discussion on the connection between the longitudinal and transverse modes and the slab geometry, the total interaction energy (a a?) for these mode (with 30 along the z-axis) is given by: V‘L’dsm = 5’3- p2 + E [#0] (37b) 0 0 slab 3:10 2 2 u u 9 4 o s __ _ 1 =3 _2 + Z [3:0 ] " 5- —fl 7) -' [‘7' (1 .) 871/3] 0 ho and V"’(f-0) = 0 + 2 [2:0 ] (37c) slob 8:10 ' 2 2 u u + Z[f=o] - 3—3-7? = [4n(0-§)--4n/3]—3- 1'! 3:0 LMMM;WW Up to this point, we have taken a purely microscopic approach and our results have been fairly formal. We now show how they can be interpreted physically in the context 24 of simple macroscopic electrostatics. In what follows, we consider the lattice to be composed of a spherical ”near” region where the dipoles are treated discretely imbedded in a ”far” region where the material is treated as a homogeneous polarizable medium with an ellipsoidal outer surface. We show this schematically in Figure 1. If the medium is now uniformly polarized, a surface polarisation charge forms, both at the outer ellipsoidal one and the fictitious, spherical, inner one. The areal charge density is given by 3-3, where P is the polarization, in this case equal to polo, and Q is the unit normal to the surface. For spheres, this charge density is given by [Pl cos 6, where 6 is measured from the direction defined by the polarization vector, while for slabs, it is IPI and 0, for P perpendicular (longitudinal) and parallel (transverse) to the slab face, respectively. There is no volume polarization charge as 9-3 = 0 for constant P. The electric field due to polarization charge at the center of the inner sphere is given by: 9, . 9 it?) = ‘3" ’ “1 (r? - 2') (30) I? - f‘l’ surface where the vector surface element is given by, dz - n da. The integration can be easily carried out for the special cases of the sphere and slab, giving: E” = + (Mt/3) P 25 Figure 1. Model System for Dynamical Calculations of Ionic Systems 26 where the minus (plus) sign corresponds to the field arising from the polarization charge on the outer (inner) surface, and gang-L) = -4n 3 / Isiah", = 0 where ‘J.’ and <|> refer to the polarization perpendicular or parallel to the slab face. Since the interaction energy between a point dipole and an electric field is given by, -po°E, we see that eqn. (29) can be physically interpreted as the energy due to the surface polarization charge. The net electric field from the ”far" region, which is the sum of the fields due to polarisation charge on the (real) outer surface and the (fictitious) inner surface is given by: (netJar ) sphere 0 (39.) (net..fo.r) “aw = ~(8n/3) 3 (39b) (netJor) ““4. = (411/3) 3 . (39c) There is also, in general, an electric field (and hence interaction energy) due to the discrete charges within the sphere (the so-called ”near” field), however, in the analysis above we have shown that this vanishes for the case of cubic symmetry. Therefore the net interaction energy of the central dipole with all the other dipoles of the system in this macroscopic approach is given by: 27 vdd - “0 total - - “o . outer + inner) a r 0 (spherical surface) (40a) 2 A 8n/3 -3 (PL slab face / longitudinal) (40b) 2 “o L -4n/3 ;—- (P I slab face / transverse) (40c) We have thus recovered the previous results, eqn. (37), without recourse to the Ewald method, using simple macroscopic electrostatics. As we shall see, this approach has a greater validity than demonstrated here, and can be used to calculate the two most important quantities, the density of states (DOS) and the dielectric (IR) response, for disordered systems. This results in a significant savings in computational time when compared with the Ewald method. In fact, for the study of the mixed crystal semiconductors, this allowed us to use the on-site computer (IBM 3090VF) rather than remote supercomputing facilities. medtmmmhtmunm Assuming a harmonic solution, uo(t) = poeudht, eqn. (7a) gives: 28 w + J— 3030) = 0 (41) [w(E)' - ail; 412p . 0 For general E, -E(3°) is obtained from eqns. (20) and (22), the Ewald result, with the substitutions: i:}9 Do, (50- f)'—. (60- Inf. and (Do- §)'-+ (Do- an. Thus eqn. (41) leads to a 3 x 3 secular determinant (one equation for each of the three components: (#0)”, (p0)y, (Mo’sL’ from which the eigenvalues (0(3)2 can be determined numerically. Results for the three cubic lattices, described by eqns. (31)-(33), are presented in Figure 2(a-c). They give the dispersion curves along various paths in the first Brillouin zone, connecting the symmetry points of each lattice as labeled on the lower axis. We further identify the symmetry lines along the coordinate axes, (100), the face diagonal, (110) and the body diagonal, (111). The locations of these special points and lines within the Brillouin zone are shown in Figure 3(a-c) which have been reproduced from Callaway (Ca 76). The LO-TO splitting is clearly apparent at the zone center (P) for all three lattices. From eqns. (39), the net electric field is given #3 “my by, -8n/3 —0 [471/3 T] for the longitudinal (transverse) branch at the zone center. Thus: (a: - (0:) .. £00: (1.2.) 2 2 _ _ 1 (wk - w5)- . w (42b) Note that this holds for all three lattices. To get an idea of the utility of the Ewald method for finite I, we plot in Figure 4 the longitudinal and transverse branches along the (100) direction for the simple cubic lattice using the Ewald method (solid line) and the truncated dipolar interaction, eqn. (7), for various cut-off radii, in units of the lattice spacing. We see that the latter all start at zero at the zone center, for both the longitudinal and transverse branches, and then approach the Ewald values at varying rates. For the smallest E, even truncating at ten lattice spacings, corresponding to over 4000 dipoles, is inadequate. However, as we approach the zone boundary, the values for small cut-off (Rmneu'= 2, for example), are reasonably good. Thus we see that the Ewald method is most important at small, but non-zero(I), wavevector, For modes at the zone center or near the zone boundary, the truncated interaction is potentially still useful. We will see this empirical rule further validated when we consider the problem of disorder. For the disordered dipolar systems, we are most interested in the density of states (DOS). Thus to provide a basis for comparison, we calculate the DOS for the three periodic cubic lattices. We do so by randomly choosing a 30 1-0 E : : E I 3 0.5 r CL I.“ - I.“ 0 _. _0_5 s a s s r X M r R M X R (100) (110) (111) Figure 2(a). Energy Dispersion for SC Dipole Lattice 31 1.0 0.5 r E/Ep -0.5 (100) Figure 2(b). (110) (111) Energy Dispersion for BCC Dipole Lattice 32 1.0 : f 2 : g ; I E/Ep -0.5 i. E 5 i E E E i (100) (110) (111) Figure 2(c). Energy Dispersion for FCC Dipole Lattice 33 (a) SC (b) BCC (c) FCC Figure 3. Brillouin Zones for Cubic Lattices 3). 1.0 0.5 E/Ep -\ -0.5 Figure 4. (100) Dispersion for Simple Cubic (100) using Truncation and Ewald Methods 35 -9 9 9 9 point in reciprocal space, k = x‘b‘ + xzb2 xsba, where x‘,xz,x3 are random numbers between zero and one, and 3;,gz,g. are the reciprocal lattice vectors, and then calculating the (three) corresponding eigenvalues, ”(3)2. Note that each point in this region can be mapped to a unique and equivalent point in the first Brillioun zone. The values are then distributed into one of 150 bins from -0.5 0): to 1.0 «3:. For the results shown, in the solid line, in Figure 5(a-c), one million points were sampled for each lattice. Although all three show a ”transverse" peak on the left, there are quite large difference between the simple cubic and the body-centered and face-centered lattices. First of all, whereas the last two have band edges at - 3 and E, corresponding to the modes at the .zone center, the simple cubic band is slightly broader, the edges due to the longitudinal and transverse modes at the point (X) in the Brillioun zone along the (100) direction. Furthermore, both the BCC and FCC lattices have a dip at approximately 0.12 and a broad ”longitudinal” peak to the right. And other than the former having a slightly sharper transverse peak and less broad longitudinal peak, the two are quite similar. In contrast, the simple cubic lattice has a rather flat band to the right of the transverse peak other than a smaller peak at zero which is due to the modes at point (B) at the zone boundary along the (111) direction. 36 6 l T V r y Ewald —-- NM 4 — _ A |.I.l V Q a. ll.) 2 - _ .J x ’ .\ o 1 l 1 J \ 1 -0.50 0 0.50 1.00 E/E D Figure 5(a). Density of States for SC Dipole Lattice 1.00 E/ep Figure 5(b). Density of States for BCC Dipole Lattice 37 Ep 0(5) 1.00 Figure 5(c). Density of States for FCC Dipole Lattice 38 (The even smaller peak just to the right we feel is noise as we can find no stationary points in the dispersion curves corresponding to this frequency.) To determine to what extent these structures are related to the coordination (number of nearest neighbors) of the lattices, we have calculated the DOS, having truncated the dipolar interaction at the nearest neighbor. These are plotted in dashed lines on the same set of figures. For the SC, the result is quite flat as in the case for the full interaction. It is symmetric, with a small peak at zero (as before) but with no transverse peak. At first glance, the result for the BCC is quite different from its full interaction DOS, with an extremely steep maximum at the origin. However, we can consider the hollow to the right of this peak to be the origin of the dip seen previously and the smaller side peaks to be related to the (narrow) transverse and (broad) longitudinal peaks. For the FCC, the nearest-neighbor results are surprisingly similar to the full interaction results even to the extent of the band edges being in almost exactly the same places. Two peaks are seperated by a hollow of nearly identical depth and position as before, although the right-hand one is narrower for the truncated interaction. Thus these NI results are suggestive of those for the full interaction, 2 point we will return to when we discuss the model of Pick (Pi 77). 39 mummnnmmdmnmmmm; Alum“ The presence of periodically repeated structural units in crystalline solids allows an enormous simplification to be made in the calculation of the dynamics of harmonically coupled systems. This is possible because the eigenfrequencies (and eigenmodes) can be paramaterized by the wavevector, thus requiring only the diagonalization of a 3s x 3s matrix where s is the number of atoms in the basis. When disorder is present, be it structural as in a glass or substitutional as in an alloy, the wavevector loses precise meaning and we are faced with the diagonalization of an I x I matrix, where I is the number of atoms in the solid. In practice, one works with a number sufficiently large to characterize the disorder, usually a few hundred to a few thousand units placed within a cube. To minimize the surface effects, periodic boundary conditions are used. Our disordered system is thus a periodic one but with a very large unit cell called a supercell. One constraint imposed by this periodicity on numerical simulations is that we are .9. always working at i=0 since M1(R) =- 121(0)“k R 8 [41(0) where R is a lattice vector of the supercell. As we have seen at the zone center, it is imperative that shape effects be 40 included in the calculations which we do so as before through the use of depolarizing factors. To diagonalize directly the resulting dynamical matrices for these supercells is not a trivial task even with today's supercomputers. The Equation of Motion (EOM) method, however, provides an efficient means of calculating many quantities of interest for random systems (Th 86 and references therein). This is a variant of conventional molecular dynamics tailored for harmonic systems. The initial conditions at t=0 are determined by the particular response function it is desired to calculate, and the equations of motion are integrated forward in time with appropriate weighting factors. We discuss this method more fully in Sec. II-B. In order for a numerical algorithm to be accurate, the step size must be small enough to adequately track the highest frequency, or equivalently, shortest period oscillation of the motion. One the other hand, the simulation must be run long enough that the longest period motion is sufficiently sampled. Thus there is an advantage in minimizing the maximum frequency as this allows for a larger time step and hence a smaller total number of time steps needed for the simulation. As we have seen previously, the natural frequency of the dipoles, “5’ merely shifts the spectrum, thus the argument above suggest that we 41 set it as small as possible. Note, however, that if set w: smaller than the lower band edge (a: E (0:), then (.02 will become negative at some point. This leads to exponentially growing solutions (a —9 -1|m|, iwt —9 +|m|t) which cannot be properly simulated as it quickly leads to overflow errors in the computation. (Physically, this instability is related to the ferroelectric transition in which a soft phonon leads to a new equilibrium structure with not static dipole moment. However, whereas in our case, m5 is an adjustable parameter with little physical significance, the corresponding frequency for a real solid is related to the physical masses and force constants of the system which are, of course, given by nature and not easily tampered with.) If instead we solve a dynamical equation that is first-order in time: t o t. 4flp R 9 R 5 jfli ij ij (E03 00:, RP: 00:) with solutions, A“) = rote-mt, then this problem can be avoided. Note that this equation has the same eigenvalues DOW as before since the dynamical matrix, ii given by: 42 a s an 60-3 3 R1.) Rii Di) = n—s— " as g 1‘.) (2) ij 11 remains unchanged by this switch from a? to E. However, even for negative E, the solutions e‘Et are still oscillatory and bounded. In addition, the complex arithmetic features of FORTRAN 77 make the implementation of eqn. (1) straightforward. The only disadvantage of this transformation is that the 3's are now complex, essentially doubling the number of variables, but this is more than compensated by the decrease in the maximum frequency or rather energy (Eamg). We set Eo=0, expressing results in units of Ep 5 mi. Equation (1) is solved using a 4th order Adams-Bashforth method which is of the form: _ h - _ a ym1 - y" + 57.- (San 59fn_‘+ 37f"_2 9fn_n) + O(h ) (3) where Ens f(xn,yn) and y; = fn and h is the step size. The previous starting values are generated using the (self-starting) 4th order Runge-Rutta method (accurate to order 0(h5) which is described along with the Adams method in Koonin (Ko 86). 43 meummmmLume We first consider the density of states as it is somewhat simpler. Because eqn. (1) has the form of the time-dependent Schrfidinger equation, we find it convenient to use bra-ket notation; however this in no way implies that we are treating the dipoles quantum mechanically. Thus we re-write eqn. (1), with Eo=0, as: Q. 1 —— |p(t)> = H |u(t)> (4) n. where |p(t)> is a Bid-dimension complex (Hilbert space) vector and H, in the site basis [1,00, is given by -(Ep/4np) 0:3 In this basis, we can write the initial 1‘ state |p(0)> as: |u(0)> = 2.1,“ [1,00 (s) 1,a where the a1 a 's are in general, complex. We can introduce 9 another basis, the energy eigenbasis, IE"), which we can relate to the site basis using the completeness relation: 2 "3“"! = 1 (6) h where 1 stands for the unit matrix with the same dimension as the energy eigenbasis. Thus: 44 |1,a> e 2 IE") = Zhv‘f'“ |tn> . (7) k k Since: d 1 a? |En(t)> = u |ln(t)> = 2" Iran» (a) we have: |En(t)> = e-iEnt IE“). Thus from eqns. (6-8): -i.E t |p(t)) = 2e 0 It") “J“‘m’ (9a) n _ (1,a) -iE t -12G.1902b" e m 'En> (9b) ’ n If we now take the inner product of this state |u(t)> with its value at t-O, we get: G(t) a (10a) = II Ni',a') (1,00 -tE t e13“. '1,a2hn h“ e n (10b) 1,a n 10’“. where we have used the orthogonality condition, (1042") a 6",”. Choosing a1,“ - expuai’a) where 91’“ is a random number between 0 and Zn, we have < " ' > a .5 6 h th b k t i if a1,“ .1,a “v 1,’1 a',a‘ w ere e rac e s s gn y the average over many runs (random initial conditions). 45 Taking the Fourier transform, we get: m «(2» a J- 1 dt e"Et (G(t)) (112) a CV av '0 Q = 2“: .2“’°‘" béi’m] 7:71“: e“"'»”} (11b) 1 n a ’ -a> a) _ 1 ((24 )t _ _ _2[§;J-dte n ]-26(r sh). (11c) n '0 n which is the desired density of states, where we have used (1): 3 5 . (12) which expresses the orthonormality of the eigenvectors of H (and follows from the completeness of the site basis, |1,a>, and energy basis, IE“), vectors.) and (2), the integral representation of the Dirac 6-function. In practice, large systems are self-averaging and it is unnecessary to average over different sets of random numbers. Two technical points need to be addressed. First, it is not possible, of course, to integrate starting from -u)in a computer simulation. Note, however, that from eqn. (9a): 46 eiEnt |u(-t)> =2 ".3 <3n|p(0)> (13) therefore: ' = and O Q G(E) . .51,- [I dt eat + Idt eat ] -1» Q O = 271? [I dt 6““: + Idt em ] 0 0 Re Idt e)“ . (11.) 0 u al" Secondly, since it is impossible to integrate forever, the upper limit should be replaced by T, the total length of the simulation. Because I is finite, the delta-functions in eqn. (11c) are replaced by the broadened resolution functions: e A'(E) .. 2 Re Jdt e“:t (15) e which is peaked at E80 with ripples on the "wings". Thus the precision of our calculations are limited by the finite width of these resolution functions much as experimental 47 measurements are limited by the finite resolution of the intruments. To dampen the small oscillations on the tails, we actually calculate: r . _ 2 G(E) - %.Re J'dt e‘nt e h(t/T) (p(0)|u(t)>, (16) 0 where h is a smoothing parameter chosen, in our case, to be 3. The other quantity of interest is the dielectric response. This is the response of the system to a harmonically oscillating, peturbative, electric field. It is a function of both the frequency of oscillation (or equivalently for our first-order equation, the frequency squared (3 E2)) and the wavelength of the displacement pattern and is given by: “3.00) = so + (audio) (17) where sm’is the low-frequency electronic response of the solid. The infinity refers to fact that this is high-frequency when compared with typical phonon frequencies. We consider our dipolar system to be a pure insulator, therefore we set sm'= l, the dielectric constant of the vacuum. The problem then becomes one of determining the susceptibility, x(E,w), which relates the polarization 48 (the response) of the dipolar system at a given frequency to the external electric field (the probe) oscillating at the same frequency: pas-210.33”. (1s) (a (Note that we denote the electric field by !, reserving the letter E for the energy.) We use linear response theory following the derivation presented in Thorpe & de Leeuw (Th 66) but re-written in bra-ket notation. (See also Maradudin et a1., Chap. 6, (Ma 71).) The equation of motion for the perturbed system is given by: Et '9- i |s>e"‘ = no |s>e"‘" - Z 33 Ij.fi>e-ut (19) 1.3 O- t where H0 is the original Hamiltonian with eigenvectors, IE“), and eigenvalues, In. Multiplying on the left by (Enl e -'tEt and eliminating the common term: , we get, after some simple algebra: “..I’J” "| 2 r9 (r -n J.” h Multiplying on the left once again, this time by , and summing over the indice n, we get: 49 = = = 2 2 8,, . (21) n j,” (En-E) The polarization is given by: i 1 Pa - 5211051) = 52 i i (E -E) n DP‘ DV1 ___ 2 ‘3 . (22) (3 i J 9 fl Comparing with eqn. (18), we therefore get the desired result: x . .. (23) a” “ (sh-s) The systems we are considering are either cubic or random and therefore the susceptibility, at first glance, is isotropic and characterized by a single scalar quantity. Recall, however, that for i=0, the response is shape-dependent and that the transverse and longitudinal responses, corresponding to polarizations parallel or perpendicular to the slab face of a slab geometry, are different. (Recall that the sample shape is incorporated in 50 the potential via the depolarizing factors.) Therefore the susceptibility tensor is diagonal but with two distinct elements: 1" = xyy E x, and x” a 1L. (For a spherical sample the response at E is isotropic and z a x z a 2..) Each of these elements has the form: X8 yy xm = 52 (24.) (s -s) is.) n n z[2bd °° h” °°]+— . (24b) -3) However, because D?? is real, so are its eigenvectors, given (1.6:) by: (tn 1. Furthermore, we are interested in the imaginary (absorbtive) part of the suceptibility, thus: Im x” a a: [2 h“ '°°]I- —;— —2) (25a) 2 = 3.2 [2 h;"°°] sun-s), (25b) n 1 where we have used the formal relation: n 6(E -s) = lim (3 - r -i.r)-‘ . (26) 790 To actually calculate this quantity with the Equation of Motion method, we give each dipole an identical initial 51 displacement along the direction of the field. In bra-ket notation, we have: Ga (27) a1,” = - 6 Therefore, from eqn. (10), we have: 3 . G(t) = =2 [ 2 a1 fl b;1’n’] eflint 1 9 h 2 . =2 [Zb;1,w] e-‘tlnt (28) n 1 and O c(£)=}7aeIdt e"it G(t) 0 (ion 2 -2 [2b ' ]6(:-:) n n n 1 25(E) * 0 (aax,y) e (29) x£(E) * 0 (asz) where the last line follows from eqn. (25b) assuming the normal to the slab face is along the z-axis. To obtain 2a(E,l) for non-zero I we set .. . ,9 .. ..1th II ka e3 1.1 where k is a unit vector along k’: to obtain 9 AL ‘3‘? A aa(3,l) we set a1,“ I ka e .i where k, is a unit vector 52 perpendicular to E. In both cases, 2 = (2n/L)*(n‘,n',n.), where L is the supercell box length and the n’s are integers, and :1 is the postion vector of the ith dipole within the supercell. Thus, to summarize, we build an initial state |u(0)> (see eqn. (5)), calculate (numerically) its time development |u(t)>, and take the Fourier transform of G(t) 5 . We next consider the implementation, within the numerical simulation, of the Coulomb interaction. El.Ianlsmsnlalisn.2£.ths.anlsmh.Intszasiian. Consider a cubic supercell of length L on a side containing I identical point dipoles. We denote the lattice vectors by K I L(n‘,nz,n.) where the n's are integers. The displacement vector between two dipoles in the same supercell, say i and J, is given by i?“ 2 3.; :5. To implement the disorder, we construct perfect (cubic) lattices ofll'sites within the supercell which we then populate with point dipoles with a given probability p. This allows us to interpolate between the two extremes of periodicity (p81) and completely uncorrelated randomess (p —0 0, I —9 (n, with 1]) finite.) It should be emphasised that we are not decreasing the total number of dipole by decreasing p, but rather increasing the number of sites each dipole has available and thus the disorder. We adjust an 53 the number of sites, so that the total number of dipole in the supercell, M remains fairly constant, typically at between 160 and 200. Furthermore, note that from eqn. (1), we see the ”coupling constant”, in units of Ep(m mi), is given by Hum)”. In this way, the results for different p are directly comparable. The dipolar interaction energy of the ith dipole in a given supercell with all the other dipoles in the system is given by: )3 3. 31?: (f..+3)ltfi.-(? +3): dd(i) = ‘ ’ ‘ ” ’ u (30) 9 '9 s 9 9 o [rt j+n| Ir..+n| n ‘J where the prime over the sum over 1 indicates that the i=1 term for 330 is excluded. We can, as before, re-express the slowly convergent sum into rapidly convergent ones. Because the derivation is similar to the one for the periodic lattice, we will not present the details. The interested reader is refered to the article by de Leeuw et a1. (dL 80). We merely quote the results. -. __-_,_‘.’._j _ 313,- in.) 13,- 3“) 1 x “m ZZd—L—H Ii 1' I3 I“ n i=1 " ‘5 54 2 [7:11.ti e "7315’ 9 £2! erfc(nRij)] - 2 19-.°mn"')(i5-3)(35-i) --‘3:u' *3- R' i. ij j 15 316? 1 ii I d «uh/n)z .. + :33 Z 1‘ (35.; I) Z (333 if) e": ”u k’flo i=1 3 4n 2 + ?)— 28°! (“1’01 (31) asl where it) a fir-K, It”. 2 lit)" 5 2 (ill) (n‘,n.,n.), 951.7, and not are the depolarization factors defined previously. For our simulations we have used n I 3.0 and have truncated the real-space sum at RU < 0.75 and the reciprocal space sum at [3| < 3.5. The results of Sec. (III-C) suggest that we can, alternatively, truncate eqn. (30) at some radius, Re, about each dipole, treating the other dipoles discretely (whose net contribution will in general not vanish, unlike the case for the cubic arrays) and augment this interaction with the electric field from the ”far" dipoles. (See Figure 1.) This field, as we see from eqn. (39-III), is proportional to P, in this case, the polarization averaged over all the 55 dipoles in the supercell (= {5 {til L'). In essence, therefore, we are handling the ”far” region contribution of the Coulomb interaction using a mean-field approximation. As we shall show, this approach is quite satisfactory for calculating two of the more important response functions, the density of states and the dielectric (IR) response in the long wavelength limit (3-0) for these random systems. It should be stressed that in both approaches, lwald and truncated, the interaction of a dipole with all others in the system have been included. We test for the equivalence of the two implementations of the Coulomb forces (the exact Ewald method and the approximate truncated/auguented interaction) on a simple cubic lattice with J'8 6' = 216 sites in the supercell choosing p-0.9. We feel this provides a rather severe test of the approximation as it includes a large number of low symmetry configurations. Simulations were run, on a single configuration, for ten periods, where one period is the time needed for an oscillation of energy IP to be completed. time steps equal to one-fifteenth (:5) of a period. (cycle) were used for the DOS and the 3-0 dielectric responses. For 251(3,“ smaller steps were needed to insure numerical stability, i th and i th of a cycle for the truncated and Ewald runs respectively. for the runs using the truncated method, BC was chosen to be 0.‘ L(- 2.‘ s) and 0.5 L('3.0 a) 56 where a is the nearest-neighbor distance of the simple cubic lattice sites. These radii contain 52 and 102 dipoles, respectively. The truncated runs ran at least 50 times faster than those using the lwald method. This might be overstating the time advantage somewhat as the Ewald code was not fully optimized, although it was partially vectorized. (The truncated method was applied successfully to ionic systems (with discrete charges) by Boltjes & de Leeuw (Ga 89) although the time improvement was not nearly as dramatic. As we did not review the code in this case, we cannot explain the reasons for this. lonetheless, a quick comparison of eqn. (41) with eqn. (40) reveals that the truncated method is far simpler to implement and code.) The results are shown in Figures (6-10). There is some discrepency in the 00$ for the Re a 0.‘L run but by RC 3 0.5L the convergence to the Ewald result is nearly complete. (The overall "peakiness" is due to the finite size of the model. It is the comparisons between the three results that is of interest to us here.) For the 2.0 dielectric responses, the finite size is of less consequence and other than the case of the longitudinal response, Im 5(1) where the truncated results are slightly narrower, the agreement between the two techniques is excellent. A more serious discrepency arises for Tm st(t,l) for 3*(2n/L) h1_and is analogous to the problem with truncation 57 that we found when calculating the dispersion for the simple cubic lattice with the same method. (See Figure A) In the truncated results the energy of the mode is underestimated and the response broadened. however, as Re is increased, the results improve and for sufficiently large radius would eventually agree with the Ewald result. In Figure ll we show the truncation results for §8(2n/2L) ii! A 8 2L. There is a smaller change from the Re I 4 to the Re a 5 result, indicating that the convergence to the true (Ewald) result is farther along. As we see in Figure 6, this is also the case for the simple cuhic results. This failure of truncation suggests that correlations between dipoles separated by more that three bond lengths (the size of Re) are important for the small wavelength modes. Note that for non-zero : the net polarisation nearly vanishes and it is the ”near” region that provides most of the interaction. As : increases, the more rapid alternation in the phases of the dipoles leads to a faster cancelation of the interaction from more distance ”shells”. For the 008, the random phases of the dipoles also leads to the vanishing of 3 and in this case, the cut-off procedure is even better. In the case of the dielectric response for 2.0, this polarisation, and the macroscopic field it gives rise to, is quite important as it leads to the splitting. In addition, the success of the truncation method shows that 58 the variation of this field throughout the supercell is quite small and is accurately represented by the ”average” given by F. Encouraged by these results we now examine more systematically, the effects of disorder on the density of states. 59 2.0 r I l Ewald /‘ ——r-05L . I —-—r'-0.4L . “7 / K“1.0? I \ \ - “I 0 -1.0 -0.5 0 0.5 1.0 Figure 6. 9 l I l n Ewald 6 n. — - P . 0.5 L ‘ '— —-— P I 0.4 L E :2? .5. 3 - _ O —- _J -1.0 -0.5 O 0.5 1 0 E/Ep Figure 7. Im 11(E) for Disordered Simple Cubic (p30.9) 60 9 T l l —— Ewald __... — r - 0.5 L __—_- FIDOu‘L 6 P —J E? ‘3 2. .5. 3 _ .. o 1 ' ‘ -1.0 -0.5 0 0.5 1 0 E/Ep Figure 8. Im s.(E) for Disordered Simple Cubic (p80.9) 9 I l I 6 _ .__-— r-IOJSL ET. 2‘.’ .5. 3 .— o ‘ ' -1.0 -0.5 0 E/Ep Figure 9. Im 53(E) for Disordered Sisple Cubic (p80.9) 61 lmlflk. 5)] 0 -1.0 .’ Figure 10. Im 8l(k,E) for Disordered Simple Cubic (p80.9) f-(zn/L) i \ m...“ l '— 5p. ——-rmax-0.5 I‘- W I “~53 ,’ I .§ 3 — I )- I l 91.0 -0l.5 40 0.5 1.0 E/Ep Figure 11. Im clan!) for Disordered Simple Cubic (p80.9) 32(2n/2L) E 62 Lmummer. In Figures 12(a-c) we give the results for the density of states, calculated via the truncation method with RC = 0.5L and averaged over ten configuration, for the three cubic lattices. He chose p=0.75 (—————-), p=0.50 (——— -——), p=0.25 (—— ° ——), and for conparison, give the results for a completely random dipole gas (-— -<-) in which the equilibrium positions were chosen at random. The disordered DCC and FCC lattices have nearly identical DOS, the ”transverse” peak remaining identifiable even at p-0.5 and the ”longitudinal" shoulder having filled in the dip that existed at p=1.0. Together with the SC, all show a gradual broadening and, not unexpectedly, all approach the random result as the number of possible sites becomes larger and the discreteness of the lattice less important. Eventually, symmetric tails form, as seen in Figure 13 for the case of p-0.05 and the DOS for the three lattices have nearly converged to that for the dipole gas. (Also note the nearly factor of five increase in the bandwidth with increasing disorder.) These tails are a result of isolated, strongly-interacting dipole pairs which become more plentiful as the lattice becomes more disordered. We can explain this quantitatively by diagonalising the 6 x 6 interaction matrix of a single pair with 3 8 Rs. 63 I I T I —— p-0.75 -- - 9-0.50 —r— 9-0.25 —-— PDHOOIII A 2 r- ‘fl Ill V Cl 0. I.l.| d 1.5 Figure 12(a). DOS for Disordered SC Lattice. T I l T -— 9-0.75 — - p-0.50 —— p-0.25 —‘— PDHOOR A 2 '- Ill V O. 0. III 1 L— / P’- o A L -1.5 Figure 12(b). DOS for Disordered BCC Lattice 64 3 T I I I I --—-—— p-0.75 — — 9-0.50 —— 9-0.25 —--— random _ A 2 - |.I.l v Q a. u: 1 .— /./ -l'f 0 -1.5 Figure 12(c). DOS for Disordered FCC Lattice 65 (31) I where 0 represents a 3 x 3 matrix with all the element equal to zero and A is a 3 x 3 diagonal matrix with A“: A”: 1 and A”: -2. The eigenvalues are AL 2 -2,2 for the in-phase,out-of-phase longitudinal modes and A, a -1,1 for the doubly degenerate (3 along § or ;), transverse modes, where we have given the values in units of (lp/dnpi'). Thus the spectrum consists of a symmetric set of 6-functions and that for a given mode, there exists a one-to-one correspondence between the energy and the radius of seperation, that is: h I, E (R) = . (32) A Anna. We can therefore write a relation between the density of states from a single mode, px(l) and the density of dipoles a given distance away, p(R). He have: p(R) d3 = “Px(l) d2 (33) where p“) I lmlzp (p is the number of dipole per unit volume) and the minus sign reflects the fact that the energy 66 decreases as the radius increases. Therefore px(l) = -p(R)/(d!/dR) where from eqn. (32), dF/dR = -3x FP/(énpi‘). Thus: (kapn') ' XI px(2) = -——————- = -—§ (3‘) 3X 33 where the above result is for a single mode. The actual density of states is an average of all the modes of the pair of a given sign, thus :9 p(3) 8 RP (px(l)> I (k) / 3(1/39)’, where (A) 2 3 x 1 + a x 2 = 3. (There are 2 transverse modes, x, = 1 and 1 longitudinal mode, AL I 2 per sign, out of a total of 6.) Our final result is then: 22’ :9 p(E) = ;;§ (35) which we plot in Figure lb along with the results for the dipole gas. The agreement is quite good even for relatively small I but breaks down eventually as the probability of an isolated pair with large B, corresponsing to a low I, vansishes as larger dipolar units come to dominate the spectrum. Equation (35) is the one analytical expression we have succeeded in applying to the output of our computer calculations because it involves solving the tractable problem of a dipole pair. The three-dipole problem, and 67 0.50 I I T I I I Y i i g —— random A I — —' SC ‘19 —-— acc Q ---— FCC a. 0.25 - % Ill I : /' 0 i L L 1 1 -10.0 -5.0 0 5.0 10.0 E/Ep Figure 13. D08 for Disordered Cubic Lattices (p=0.05) and Random Dipolar Gas 0.50 I .lgr I i ‘ Pandas 93 l I Q \ a \ Ill -10 Figure 14. D08 for Dipolar Gas with Lorentzian Tails from Dipole Pairs 68 those involving larger clusters, are in principle solvable but we no longer have a simple relationship between a single radius and the energy, which was the key for the dipole pair. Although we have been considering the random dipole gas where a continuous range of radii exist, we can apply the pair analysis to the disordered discrete lattices. For example, the prominent peaks that exist on the tails for the simple cubic in Figure 13 are a result of the. longitudinal modes for dipole pairs seperated by the nearest-neighbor distance. From eqn. (32) we have: 2 Sp E (R ) = . (36) 1. an 3 (mp (Rm) However, p = N/V =p luau/V and (RM)3 = VII-u“, thus: 1 E (R )/E = -— a: 3.18 (37) I. rm 9 2119 for p-0.05. The peaks due to the transverse modes (at Flip z 1.59) are somewhat obscured by the shoulders and indeed are responsible for the deviation of the simple cubic DOS from that of the dipole gas on the left-hand side. For the DCC and FCC lattices, I!“" = sf3/2 and R 8 elf? “h respectively, where a is the linear dimension of the 69 (ICC) l . mites conventional unit cell. Also, a 2V/sa and ('00) sites = 4V/a'. Therefore the longitudinal mode for the nearest-neighbor dipole pair occurs at: 1 thump/I: = — a: 2.12 (ace) (38a) P 3np 1 EL(Rm)/Ep = up a: 2.25 (FCC) . (38b) These peaks are also seen in Figure 13 although masked by the asymmetry of the DOS for these lattices that persists to some extent even for large disorder, p=0.05. He close with a comment on the previous results of Pick and Yvinec (Pi 77) and its relation to our work. Although they considered the dynamics of point dipoles, their model started from a molecular dynamics simulation of liquid argon, interacting via a Lennard-Jones potential, which was then quenched to give the equilibrium positions of the dipoles. Because of the excluded volume effects due to the hard-core repulsion, their results, obtained using a truncated dipole interaction with no ”extra” polarisation term which as we have seen is quite adequate for the DOS, differ quite profoundly from our calculation on the dipole gas in which there is no restriction on the minimum separation of the particles. And as we have seen, the 70 presence of the Lorentzian tails, missing in the Pick results, shown in Figure 15 (reproduced from Pi 77), is a directly consequence of this. However, a liquid, to a first approximation, has a close-packed structure with twelve nearest-neigbors, as is also the case for the FCC lattice which indeed does exhibit a two-peak DOS, even when only nearest-neighbor dipole interactions are considered. The effect of disorder would be to smooth the peaks and give a result much like Pick's or our own disordered FCC lattice. (See Figure (12c) for p=0.75.) Although for identification purposes it is useful, as Pick has done, to describe the sharper low-frequency peak as belonging to the ”transverse” mode and the other to the "longitudinal" mode, the splitting is an attribute of the dielectric response 8(w) not the DOS. Note, for example, that in the case of the simple cubic, where the spliting at the zone-center is quite clear, the DOS shows no clear two-peak structure but rather an off-center "transverse" peak with a slowly decending plateau. In this section we have concentrated on the DOS as this shows the overall effect of disorder on the system. In the next section, we will consider the more selective dielectric and Raman response and the effect of disorder on the Coulomb splitting. We do so for an ionic mixed crystal system with discrete charges, studying, in particular, the effects of 7l ..ZJO 0.90 0.95 1 we no Inf-F; Figure 15. DOS for Quenched Dipolar Liquid (Pick) 72 diminishing ionicity on the LO-TO splitting as positive-negative charged pairs are replaced by neutral, covalently bonded atoms. 73 memmug_xuuaimm A... W111. Recently there has been a good deal of interest in mixed crystal semiconductors of the type (III-V)tmIva where a cation is taken from column III of the periodic table, an anion from column V, and a neutral (covalent) element from column IV. Examples of these include (GaAs)1_xGezx and (InSb)‘_xGeu. In the mixed crystals that have been more extensively studied, such as the alkali halide labflxx Cl or the III-V semiconductor Ga Pvasx, either the cation or anion of the pure crystal is substituted with another in the same column of the periodic table so that the basic character of the material reaains unchanged (see the review article by Chang Ch 71). However, in the (III-V-IV) mixed crystals, we go from a (partially) ionic (III-V) semiconductor to a purely covalent (IV) one. Accompanying this shift is a structural transition in which a zincblende structure abruptly changes to a diamond structure as the order parameter, defined as the number of cations on one sublattice minus the number on the other, goes to zero. There has been much controversy about the exact structural nature of these alloys and various models have been proposed for their formation and growth, each with a different value for xc, the critical concentration of. 74 the transition. While not addressing these concerns directly, the techniques we have developed for calculating the dynamics of disordered ionic systems is well suited to the study of an important aspect of this problem as there is experimental data by McGlinn et a1. (Hc 89) suggesting that the LO-TO splitting in optical response vansishes at this transition point. We perform our calculations on a dimer percolation model that exhibits precisely the structural transition that occurs in the actual materials, although at a critical concentration different from the experimentally determined value. Nonetheless, this model system displays many of the same optical features found experimentally giving us confidence in the applicability of our techniques. A crystal of pure GaAs or InSb has a zincblende structure, in which the lattice points form a diamond lattice, made up of two FCC sub-lattices separated by the basis vector: 3 = (a/4) x (x + ; + ;) where a is the side length of the conventional unit. (See Figure 16 reproduced from Us 70.) All the cations are on one sub-lattice, call it the A sub-lattice, and all the anions on the other (D sub-lattice). As x increased, clusters of GaAs (or InSb) form,. separate from the percolating infinite cluster, isolated by a ”buffer” of germanium. These clusters can have a different registration from the percolating cluster, that is, cations on the B sub-lattice and anions on the A. The surrounding Ge prohibits the occurence of any Ga-Ga or 75 Figure 16 Conventional unit cell of zincblende structure 76 As-As nearest neighbors which are energetically unfavorable because of the large Coulomb repulsion between identical species. At a critical value of x, the largest cluster of the III-V material no longer percolates and indeed all clusters of this type become isolated. There is no longer a prefered registration scheme, defined previously by the infinite cluster, and hence there is an equal probablity that a given cluster is either phase (A-B) or anti-phase (B-A), where the first letter gives the cation sub-lattice and the second, the anion sub-lattice. In the zone-center optic modes, for the perfect crystal, the two sub-lattices ”beat” against each other, and are in phase with the macroscopic electric field that is set up during the oscillation. With isolated clusters of mixed registration, charges of both types would be on a given sub-lattice which would destroy the coherence, leading to a degradation of the electric field and hence the splitting. Thus what might be called ”dynamical frustration” occurs as the mode pattern favors one motion while the electric field coupling forces the opposite. This non-rigorous arguement leads to the ansata that the splitting disappears not with the ionicity (xul), but rather when the (III-V) material no longer percolates, which occurs in our model, as we shall see, at xc = 0.69. Using the equation-of-motion technique to compute the optical response, both IR and Raman, we verify this for a system with one-mode behavior, roughly 77 corresponding to (GaAs) Ge, and also for a model material based on (InSb) Ge which exhibits two-mode behavior. By these terms, we mean that the net optical response of the first is given by a single (broadened) peak, whereas in the second, the InSb and Ge structures (peaks) are distinct, at each concentration. We begin with a discussion of the implementation of Coulomb forces for disordered ionic systems. anmmnmlminnunmmmnm The ideas that were developed for the dipolar systems carry over with only minor changes to ionic systems. Most importantly, the long-range Coulomb interaction can once again be separated into a near region where the discrete ions are treated exactly, and a far region that is incorporated through the average polarization of the supercell using the proper depolarization factors to introduce the macroscopic shape. Coworkers at the Univesity of Amsterdam, 3. BoltJes and s.w. de Leeuw, have computed the density of states and dielectric response (for both :30 and 3 finite), for a structurally disordered AX: ionic glass, modeled on silica, and have obtained essentially the same results as ours for dipolar systems (Ga 89a). That is, the DOS and 0(a) can be calculated quite accurately with the truncation scheme giving the same results as the Ewald 78 method for a sufficiently large cut-off radius, on the order of four bond lengths. Likewise, the truncated results for s(k,w) for the saallest possible non-zero f converge, somewhat slowly, to the proper Ewald result. (For details on implementing the Ewald method for ionic systems, see (dL 80).) We consider a cubic supercell of linear dimension L containing N ions of charge qi (i=1,2,...,N) with instantaneous positions ? For the mixed crystal we are 1. considering, qi is either +1, -1, or 0 unit charge. Electrical neutrality is always maintained by replacing a dimer of (III)o”(V)b° by one of Gez. These supercells form a simple cubic lattice described by translation vectors 3=(nx,ny,nz)L. The electrostatic interaction energy of the ith ion with the other ions in the "near” region is given by: Q - i vnmar - ‘12.: + 3 I (1) J 11 ii where 315 gives the difference in the equilibrium positions .. of the ith and jth ions and u11 = 31 - 3 the relative, J instantaneous deviation froa equilibrium. Fxpanding in a power series of the small parameter (u/R), we get: 79 2 '9 2 q u 3(u - R ) 1 11 _ 11 11 s vnear = -q1 27 R, R5 ] + 0(a) (2) J 11 11 where R11 8 |R11|. Because the equilibrium positions are chosen so that there is no net static force on any of the ions, there is no resultant linear term. The total effective force ?1 on the ith ion due to the other charges. both in the near and far regions. is equal to F1 = -61 Vtou" and given by: + + i? ___ q {:3 “11 _ ““11 in’i‘u] 1 1 “‘ 2 s s 1 “11 R11 3 4n 1 +2—2[§'-Ba]?a} (3) 4m asl where l l + Ia'az‘lj“: (4) i=1 is the dynamical polarization of the supercell. Note that the first ters in eqn. (3) is very similar to the expression for the point dipole interaction, falling off as R-.. We choose a cut-off radius of about three bond lengths: eighty-six neighboring particles are included in the ”near" region lattice sum. 80 LWflMLWW We now discuss briefly the Equation of lotion technique used to calculate the optical response of the system. In these simulations we use the usual second-order (Newtonian) dynamical equations which lead to slight modifications in the formalism. The details can be found in Thorpe & de Leeuw (Th 86). We merely quote the main results. The central entity of this technique is the Green function defined by: 1 (“um :guml G (i 1'») = (5) an ’ ’ a/s :2 2 _ a ('1'1) w Kn where t and A are the eignevectors and eigenvalues of the dynamical matrix of the disordered supercell. The m’s are the particle masses. (In the previous section on the TO! for the first-order equation, we were also dealing with Green functions. We just never called them by their proper name.) Both the density of states and the dielectric response, or more precisely, the susceptibility x, can be written in terms of this function. From Sec. II of Thorpe 6 de Leeuw, we have, - .. ESL . pm) - 3“ 1):“ m1 Im cw(1,1.m) (6) 0 81 ”(w) = DP‘ io‘qz Wu 1; co) . (7) 1a, Note that the results are now a function of frequency rather that energy (a up). Because we are interested primarily in the imaginary (absorptive) part of the susceptibility, we need only the imaginary part of the Green function for both these quantities. It is shown in Sec. VI of (Th 86) that this imaginary part can be computed numerically from: r z _ a -X(t/T) G(w) - "J;dt cos(wt) e 12; A1,“ “i,a(t) (8) 0 where u1 a(t) is the displacement along the nth axis of the 9 ith particle. Also mentioned earlier, the smoothing factor A lessens the ripples in the convolution function introduced by the finite T truncation. If we let Ai,a ani and “i,a anil'i’ where E is a unit vector in the direction of the electric field of the incident light, then we get, using slab geometry implemented via the depolarizing factors: Im (stag) - a: m] = g— G(w), E parallel to slab (9a) 2::2 ‘ Im [6"(0) - em] = 53— G(w), E perpendicular to slab. (9b) The Raman response of a material is a function of the polarizability of the electronic- degrees of freedom in 62 contrast to the ionic polarizability that gives rise to the dielectric (IR) response discussed previously. As such, a fundamental expression for it involves the complicated quantum mechanical calculation of the electronic structure of a solid. For our purposes, a phenomological approach will suffice, based on a local bond-polarizability model of Alben et a1. (Al 75). The perturbing interaction between the external radiation and the phonon is described by the polarization tensor: 9 2‘1) [’15 ° ’11 " 31] “1 ’11 (1°) i,j where r11 is the unit vector pointing from site i to site 1, 1 is the 3 x 3 unit matrix, and 3 is the displacement of i the atom on site i. Note that the symbol 0 signifies the outer product of the two unit vectors, giving, of course, a tensor. In general, the coupling coefficient A11, can depend on the bond, however for simplicity, we take A11 8 A to be the same for all bonds. Thus in the perfect tetrahedral arrangement that exists in the zincblende (and diamond) structure, the second term in brackets vanishes. That is: 2r“ =0 (11) J for the unit vectors given by: 83 3‘ = :(1/73) (+1,+1,+1) (12.) $2 = :(1/13) (+1.-1,-1) (121) :3 = :(1/15) (-1,+1,-1) (12.) ;‘ a 2(1/13) (-1.—1,+1) (121) .where the two signs correspond to the two sites in the zincblende basis. For this same reason any diagonal element of the polarization tensor (which gives the response of the system along the same direction of the incident radiation field) also vanishes. In contrast, for the off-diagonal elements, say 01“)", we have for the contribution due to the ith site: (xy) (x) (y) (on (on + (on 011 or Zrij r“ r:L1 “i at _6“ u1 (12) j where the plus (minus) sign applies if the ith site is on the A (B) sub-lattice. Experimentally, this corresponds to the arrangement where the scattered radiation field is measured perpendicular to the incident radiation. In practical terms, for the ROM, this corresponds to 8+ 0 3+ A1,“ '6a: and ui,a ‘6asf'i’ thus even in covalent materials such as germanium with little effective charge, the Raman response is non-zero. lote that the induced motion of the atoms is along the direction of the incident beam and thus the Raman technique gives (primarily) the longitudinal response of the system. This complements the 84 IR response where the particle displacements are along the electric field direction and hence transverse to the wavevector. The EON method, however, can compute either the longitudinal or transverse response for either situation (IR or Raman) simply by setting the initial displacements either along or perpendicular to the normal of the slab face as defined by the depolarizing factors. The equations of motion are numerically integrated using the Verlet algorithm (Ve 67) that can be easily derived using Taylor expansions. We have: x(t+At) = 11(1) + 11:11)“ + 311””) (At)2 + out)’ (13.) x(t-At) = 11(1) - x'(t)At + ix"(t) (At): - out)”. (131) Adding these two equations and solving for x(t+At) gives: nun) . 211(1) - 11(1-111.) + 11"(1) (111;)2 + 0(At)‘ (11.) where x"(t) is the acceleration of the particle given by the force on the particle divided by its mass. From eqn. (14), we see that both the current and previous values are needed, and therefore the first step, x(At), must be determined before the Verlet algorithm can be started. First note that since we do not have velocity dependent 85 forces, the dynamical equations are invariant under time reversal. Secondly, we start the particles from rest. Therefore x(-At) = x(At) and from eqns. (13a-b), we see that all the odd order terms vanish. Thus: 11(111) = 11(0) + §x"(0) (At). + 0(At)‘ (15) which is of the same accuracy as the Verlet expression, eqn. (14). In contrast to our dipolar simulations which were purely long-ranged, we must include short-ranged nearest-neighbor forces in the present calculations. We model these interactions with a Born model which to second-order is described by (Th 76): - _ _ * . 1 V11 - 2(0 3) Eu“ r11] +2B(1j) (15) where 31’ 3 61 - 3’, 211 is the unit vector along the direction of the bond, and a and 3 are the bond-stretching and angle-bending force constants, respectively. The first term gives resistance only along the direction of the bond and for that reason is called the central-force term. By itself, the four-coordinated diasond lattice would be unstable. The second, isotropic, term provides the extra constraints needed for stability. (B.B.: There are problems 86 with this model that arise because of the non-rotational invariance of the second term. This is mainly a problem for acoustic modes and those optic modes involving rotations of molecular units. For our purposes here, these are not of concern where the ease of implementation is of greater importance. This subject will be dealt with in much greater detail in the second part of this thesis on the lattice dynamics of layered silicates.) qummzwum Our supercells consist of a 5x5x5 a 125 cubic array of conventional cubic unit cells each containing eight sites arranged in a diamond lattice. This gives a total of 1000 sites, 500 on the A (FCC) sub-lattice and 500 on the B sub-lattice. We start off with all the sites occupied by neutral particles which represent the germanium atoms that we replace with dipolar dimers, representing, for the purpose of this section, the GaAs and consisting of a positively and negatively charged particle separated by a distance equal to the bond length of the diamond lattice. In contrast to the case of monomer percolation, the placement procedure is a two step process. First, a site on the A sub-lattice, chosen at random, must be available, that is, occupied by a Ge atom. If this is satisfied, one of the four orientations, defined by eqn. (12), and pointing to a 87 site on the B sub-lattice, is chosen randomly. If the corresponding B-site is free, that is, occupied by a Ge, then the dimer can be ”parked" on the lattice with the positive (Ga) atom always on the A sub-lattice and the negative (As) atom on the B sub-lattice. If this B-site is filled, a new A-site is choosen and the procedure repeated. In this way, a concentration of up to 802 GaAs can be achieved. Although the mean static polarization, averaged over many samples, is zero, for any given sample there is a significant probability that such a polarization does exist. However, if we include only the dynamical polarization, as given by eqn. (4), in the force calculations (see eqn. (3)), then this spurious, finite-size effect, is of no consequence. With these rules established we can calculate the percolation statistics. We do so on diamond lattices of two different sizes, one with 63 x 8 = 1728 sites, the other with 73 x 8 = 2744, to test for convergence. Both are larger than the simulation lattices in order to obtain better statistics. We scan the ”parameter” space from x = 0.5 to 0.8, averaging over 100 configurations for each point. Once the mixed crystal is built, we enumerate the clusters of GaAs. To do this, we randomly pick a Ga from the A sub-lattice, then find its As neighbors. In turn, each of these is then checked for its Ga neighbor, making sure not to ”retreat” to the Ga already found. This 88 procedure is repeated until the only neighbors left are Ge atoms, which define the edges of the cluster. We then choose another Ga, not on any previously determined cluster, and repeat the operation until all the Ga and As have been acounted for. We next determine whether the largest cluster spans the supercell, which is the finite-size criterion for percolation. This spanning can occur along any one of the three directions and in any combination. We choose, in sequence, a plane perpendicular to a given axis and determine whether the cluster has a member within that plane. If it does not, then the cluster does not also does not span the supercell and hence, does not percolate. If a cluster does have a member in the plane and every other in the supercell (there are four such planes for each conventional unit cell), then the cluster does percolate along that direction and is given the probability of percolation of 3. We check the other two coordinate axes in the same manner. Thus the largest cluster can have a percolation probability of zero to one in increments of one-third. This procedure is performed for a given x, on 100 configurations, and the average taken to give the percolation probability for that value of x. For an infinite cluster, the (second-order) percolation transition would have a sharp step function profile. Because of the finite-size of our samples, the transition is smoothed out. 89 We determine xc, the concentration at which percolation occurs, using the midpoint of the probability of percolation which we show in Figure 17. From the tabulated data we determine the critical concentration for percolation of dimers on a diamond lattice to be xc = 0.69. The other curve, gives P(x), the probability of a Ga or As atom being on the infinite (largest) cluster and is calculated by simply dividing the size of the largest cluster by the total number of Ga and As atoms at a given x. It should be rigorously zero for x > xc, however because of finite-size effects, it tails off at the percolation point. In Figure 18 we show a two-dimensional analogue of our mixed crystal on a hexagonal net. The open symbols represent the percolating phase and the closed symbols, the isolated anti-phase clusters. The bare vertices are the neutral germanium which separate the two phases. The squares represent the cations and the hexagons, the anions. This whole unit cell is charge neutral and periodic boundary conditions hold both vertically and horizontally. (For clarity, we have not drawn the germanium bonds. It must stressed, however, that the structures we work with are fully connected.) 1.00 90 I I T I r \. \. \. \ \ 0.75 - \ 8 L 0 50 - C1 —- per: (1728) —- - lnfc (1728) ‘ —-— per: (27“) \\ 0.25 _ — — in": (2744) \\ \\ l l l l 1 I L 8.50 L 0.60 0. Figure 17 Percolation Statistics for Dimers Diamond 91 3 ”a H ”\a C :1 D: 9 II D: c 9 0 n o 1: a 1. :1 '3 9 L‘ :0 1: :1 : ‘° '3 .0 1: ’3 *9 0 c :1 fl (0 We 7 (O ‘ )3 Figure 18 (III-V) IV "Hixed Crystal" on Hexagonal Lattice 92 LWWW Since gallium, germanium, and arsenic are consecutive elements in the periodic table with masses, 69.7 amu, 72.6 amu, and 74.9 amu, respectively, for the model (GaAs) Ge mixed crystal we give the same (unit) mass to all the particles. The disorder therefore lies in the arrangement of the charges and the resulting electrostatic forces. In the units of our simulation we choose a s 1 and fl = 0.4. The unit of frequency is the maximum frequency of the uncharged diamond lattice given by (so = [g (a + 2(3) 1"2 = 2.191, where the triply degenerage eigenmode corresponds to the two-sublattices "beating" against one another along any one of the three coordinate axes. The plasma frequency of the pure GaAs crystal, which is closely related to the splitting (see eqns. (42a-b-III)), is given by (I): = (Lg- H qz. There are eight particles in a conventional unit cell of side length a =4/i3 since we choose the nearest-neighbor distance to be the unit distance. (Recall that the displacement vector between the two FCC lattices making up the diamond lattice is given by d I (a/4) x (§ + ; + £).) We take |q| = 0.3, thus we get a numerical value for w: of 0.735 = 0.153 mi. We set ‘a)= 1. The GaAs on the isolated clusters can have the same phase as the infinite cluster, that is, Ga on A and As on B, or the anti-phase, Ga on B and As on A. Since there is no 93 natural preference for one or the other, we feel the most natural arrangement is half and half, therefore we randomly reverse isolated clusters until 50% of the GaAs, on the isolated clusters, is of the anti-phase form. Only one run is neccesary for each simulation as the results vary only slightly for different configurations of the same concentration. We used a time step of one-twentieth of a period defined by T a ZR/wb and each simulation ran for sixty such periods which gives a resolution comparable to the experimental value of approximately 6 cm.‘ (full width at half-height). The run times ranged between thirty minutes (IBM 3090 CPU) for the pure (III-V) system where the Coulomb force (electric field) had to be evaluated for every particle to about three minute for the germanium where there are no long-range interactions. In Figure 19 we plot the Raman response, both transverse and longitudinal, as a function of w / as starting at the bottom with pure GaAs and proceeding, with decreasing ionicity, until we reach pure germanium at the top. Note that at all values of x (percent Ge), the response is given by a single peak. As x increases, that is, as the GaAs is replaced by Ge, the peak position shifts upwards for the transverse response (dotted line) and downward for the longitudinal (solid line) until at x=0.7, when the GaAs no longer percolates, the two essentially 94 0.8 0.9 1.0 Figure 19 Raman Response for (GaAs) Ge . 1 95 merge. This is shown more dramatically in Figure 20 where we plot the peak positions (0) as a function of x, with the solid line to guide the eye. The frequency resolution in our units is about 0.01. We can give greater support to the premise that it is mis-registration of the cation-anion sites that leads to the loss of the splitting if we now change the character of the isolated clusters by making all of them with the same phase as the percolation cluster. We see in Figure 20 that the response (+) is "repeled" compared to the previous (50-50) response. That is, the transverse frequency is somewhat lower and the longitudinal somewhat higher. Note that for a given concentration, the isolated cluster positions are identical for all three variations as they were constructed using the same random number seed. Further support can be garnered from the dielectric or infra-red response shown in Figure 21. Unlike the Raman response where the total weight remains constant for all x (due to the fact that all bond types have equal polarizability), in the IR, there is a decrease in the total response proportional to w;(x) = (1-x) (wile, a result which follows from a sum rule derived by Thorpe & de Leeuw (See eqn. (62) in Th 86.) and due to the decreasing number of charged species. In addition, there is a small transfer of response to the lower frequencies, somewhat more pronounced in the transverse response. Up to x=0.6, however, which is 96 T I l r I r I T l 1.04 - ~ 1.02 - - O 3 s " . + 1. + 1.00 - -4 1 0.98 *- 4 L l i l 1 l l l l 0 0.2 0.4 0.6 0.8 1.0 Figure 20 Peak Positions X of Raman Response for (GaAs) Ge 97 _4 0.8 Figure 21 \ 1. (1)/(1)0 Infrared Response for (GaAs) Ge 1 98 the last point for which a clear peak exists for both responses, the splitting is virtually the same as that of the Raman data. El.Ixnzflads.fisha1121.Llnfihlzfis. If we move one row down the periodic table for both the third and fifth column, we arrive at InSb. Indium has a mass of 114.82 amu and antimony, 121.75 amu. We represent these in the model with particles of equal mass given by E . The ratio of the ”bare" (uncharged) frequency of InSb, given by (ow. = ((201): + (op/31‘”, to the maximum frequency of Ge is about 185 cn"/ 302 on" = 1.63. (These and other optical data were taken from the Landolt-Bérnstein Handbook La 82). To achieve approximately this same ratio in our simulation we chose the In-Sb force constants to be 63% of the strengths quoted in the previous section for the (GaAs) Ge system. For the indium-germanium and antimony-germanium bonds we use a common convention for mixed crystal computations and set the strengths to be (0.63)"2 of the Ge-Ge bonds. There are, of course, no In-In or Sb-Sb bonds. We plot the Raman response in Figure 22 using the same convention as before, solid lines for longitudinal, dotted for transverse. Each response has (at least) two peaks, thus the term, two-mode behavior. One the right we have the 99 l l l l x=1.0 L.— x=0.9 x=0.8 x=0.7 x=0.6A 7 - x=0.54,l\ . x=0.4 m x=0.3 P N Mgo'z /' Q x=0.l 1L 11 111'; (”a”) 0.5 0.8 0.7 0.8 0.9 1.0 1.1 (1)/u)o Figure 22 Raman Response for (InSb) Ge 100 peak(s) due to the germanium which is essentially the same for both the longitudinal and transverse responses. It is a single peak from x=1.0 to about x=0.5 after which multiple structures form. These are due to the the isolate clusters of germanium of various sizes and thus different ”normal mode" frequencies. For x=0.1 it is difficult to generate the mixed crystal by replacing 90% of the germanium with (III-V) dimers, therefore we start with a pure InSb crystal and place at random, ten percent of the sites with germanium dimers. Unlike the other concentrations, we have no germanium monomers. This can be considered a slight weakness in the model but it is not serious as it occurs far from the transition point at x80.69. However, it is interesting to note that the two smaller peaks on the right for x=0.1 can be attributed to the modes of isolated dimers of germanium, with the taller one due to the two transverse modes (displacements perpendicular to the bond) and the smaller, which is roughly half the size, due to the single longitudinal mode with motion along the bond. We verify this by calculating the local density of states of a single Ge dimer in InSb using the 20H, which involves merely displacing the two Ge atoms with all others in their equilibrium positions and then following the dynamical behavior as before. These two dimer frequencies are marked along the x axis by (T) at w/wb = 0.760 and (L) at w/wb = 0.824. Clearly, the two germanium peaks for x30 are lOl converging to these values in Figure 22. The TO-LO splitting, on the other hand, is apparent on the left side of the figure. The responses broaden, decrease in height and shift to lower frequency as x increases, while simultaneously, the splitting diminishes. Raman measurements have been made on a related material, (GaSb) Ge by HcGlinn et a1. (Hc 88). We reproduce their data for the longitudinal response in Figure 23. The similarities between the measurements and the simulations are striking although our results show greater structure for the germanium which is perhaps due to our particular model. (See the last paragraph.) In the next figure, (24 a-b), we plot our peak postions (a) for both the InSb and Ge, and theirs (b) for GaSb and Ge. Qualitatively, they are quite similar, the main difference being that the actually system (GeSb) Ge undergoes the zincblende to diamond transition at around x=0.3 whereas for our percolation model, this occurs at about x=0.7. In addition, because of the ambiquity in our germanium result for x < 0.5, we do not plot these positions in (a) but do mark the germanium dimer frequencies along the y axis. In general though, we feel confident that we have captured the essentials of the process with our model and have given support to the argument that it is this loss of long-range coherence, as defined by an infinite cluster of (III-V) semiconductor of a given registration, that leads to the vanishing of the splitting. We plot our 102 x = 0.95 _/ K _. 31:0.80 A ~—:::FI73—4‘ —4'-*-“’////fi\\\_______- Ramon Inlensily I ’.‘ .0 D U) 4 M A ‘— __ _J M x : q13 A“ ,_ -1 GaSb 1 . g 100 150 200 250 300 350 Ramon Shifl (cm“) Figure 23 Raman Heasurement for (GaSb) Ge (Expt.) Raman Shift (cm'1) 1) : {1; +11}: 'GaSb" ZZOIFO ++++ + + 0 4O 60 80 (3290 Concentration (percent) A T l 1 T T T 7 T T ’ 2500L 11: ++ . 280» "C39" + a + + +11++ Jr ZGOF+ - . . 240 r Lo - - 100 Figure 24(a) Peak Positions of Raman Response (GaAs) Ge (Fxpt.) 104 Figure 24(b) 1.1 r I 1 I I r *T I I 1 "Ge" [3 I] E] El E] E] 0.9 r ‘ L _- 7.1 0.7 - ‘ A A ”InSb" A J?- A + + A A ‘l' + + ¢ A a A L l I 1 l l l l 0.50 0.2 o 4 0.5 o a 1 o X Peak Positions of Raman Response (InSb) Ge (Theory) 105 I "2,9) I l l x=0.8 X=0£7 X=0.6 x=0.5 x=0.4 /\ x=0.3 ‘ x=0 2 A 1 1 x=0.1 / +_ + k M ' ‘ x=0.0 , l l J 1 0.5 0.6 0.7 0.8 0.9 1.0 U) (I) / 0 Figure 25 Infrared Response for (InSb) Ge 106 simulations of the infra-red response in Figure 25. The results are by now, self-explanatory. These peak postions once again are identical with those of the Raman data. 107 mmmmma. In this first part of the thesis, we have considered the implementation and effects of Coulomb forces in disordered materials. Because of the subtleties involved with this long-range interaction, we have presented, somewhat deliberately, the basic ideas and techniques. These include, of course, the LO-TO splitting and the Ewald method. An important result of this part, is the demonstration, empirically, that for the important zone-center modes, the exact Ewald method can be replaced by an approximate truncated interaction technique if care is taken to include the long-range bulk and surface contributions from the dynamic polarization field set up by the particle motion. We have shown this using a pure dipolar disordered system for the density of states and the dielectric responses. Once established, we have applied this technique to a disordered ionic system based on the class of mixed crystal semiconductors of the type (III-V) IV. We have shown, with both Raman and IR response, that the LO-TO splitting disappears at the percolation concentration for dimers on a diamond lattice for both one-mode and two-mode behavior materials. An account of this work will be submitted to Physical Review Letters. Although simulation techniques have long been applied to dynamical problems in condensed 108 matter physics, most systems under consideration have involved only short—ranged bonding forces. It is hoped that the techniques developed in this thesis will play a role in encouraging more research and ultimately, greater insight into the properties of disordered ionic materials where long-range Coulomb forces are present. 109 Appendix A mmmwmm In this appendix we give a derivation of eqn. (17-III). Consider the function: .4. _+ H?) = ge‘k (if ’) nil—i") (Al) where the sum is over the complete set of lattice vectors and f(?) is an arbitrary function. It is easy to see that F(?) is periodic, that is, F(?+R') = F(?) for any lattice vector R'. Thus F(?) can be expressed in a three-dimensional Fourier series: + 13'? 1(1) . gF-o e (12) where the sum is over all reciprocal lattice vectors G and , .9 = ~151M- e": ’ 11(3) (113) where the integral is over the Wigner-Seitz cell and u is the volume of this cell. Substituting eqn. (Al) into (A3) 110 gives: . '9 . _ 1‘3 = 13§I113 8.111%) (3 1)) Mil-f) (14) c 43-? = eat-(1L?) where we have used e since G-R is an integral multiple of Zn. From the combined effect of the sum and the integral, we have: . .‘9 , r» = ‘— 11? 9-1124?) ’ “-3) . (15) a o All. Space 9 -tr2 We now choose f(r) = e . The integral is easily evaluated in spherical coordinates giving: 9 z e-'k+al /‘t . (A6) The desired result, eqn. (17), then follows from eqns. (A1), (A2), and (A6). after taking 3 = 0 and making the change in the summation from G to —-G. If instead we choose: a a a a z a ((3) = e-(t+s/a )x ‘e-(t+s/b )y e-(t+s/c )2 (A7) then from eqn. (A5), we have for the case of 380: 111 (I) . 2 2 h = 1_ I 1111 e-ucxx) e-(t+s/a )11 x a 0 -O a) O I ., H ,. I .. H -a1 -a> where the integrals over y and z are of the exact same form 2 as the that over x. Because e Ax is an even function, we can replace the complex exponential by its real (symmetric) part, cos(Gxx). The resulting integral is evaluated in Dwight (Dw 61) as: It 2 2 [ ] ‘G / 4 (t+s/a ) (A9) t + s/aa This gives: ex: 1 1'! F» (s) = — ° ° (tn/3’)“ (tn/15‘” (tn/o”)MI z s s z z z e-Gx1/ 4(t+s/a ) e-Gy,/ 4(t+a/b ) 2-631/ 4(t+s/c ) (A10) In the limit, a -+ 0, we recover eqn. (A6) for the case 3:0, For 3 -+ 0, we get the desired result, eqn. (25). 112 Appendix B mmmmmmmmmm In this appendix we look at an interesting general result obtained by Thorpe S de Leeuw (Th 86) that relates the longitudinal response to the transverse. The derivation is worked out in Sec. IV of their paper and is a consequence of the fact that the shape-dependent effects, contained in the depolarizing factors, are separable from the other terms of the potential. We merely state the final result. st(w) s 8(1) 61(0)) Taking the imaginary part of both sides we get: - .. 2 _1_ Im [81(0),] - so Im [85(0)] Im [51(w)] a s 8 . (82) °° (Re 113(11):)” + (II 15(11):)“ Since the real part of 11(w) can be determined from the imaginary part, through a Kramers-Kronig (El 71) given by: 113 a) 1m [stuo' )] 11(11')2 , (113) 2”" Re [stun] = em + I s s o (w') + m then the longitudinal response, which can be determined separately by simulation, can also be obtained from the simulation results of the transverse response. We have verified this for the two values of x = 0.2, far from the percolation transition , and x = 0.8, when all the GaAs is in isolated clusters. We present the results in Figures. (26a-b). Except for the optic peak for the case of x = 0.2, the agreement is quite good. Barrow peaks give difficulties because the computed peak heights are sensitive to the resolution of the simulation and hence not always accurately reproduced. 114 o I I l A 0 0.2 0.4 0.6 0.8 1.0 1.2 (1)/(1)o Figure 26(a) Longitudinal Response EOH vs KR (x=0.2) 0 0.2 0.4 0.6 0.0 1.0 1.2 (1)/u)o Figure 26(b) Longitudinal Response EOH vs KR (x=0.8) 115 mwumm ll6 mehsim Clays or layered silicates are a ubiquitous component of many soils and sediment deposits throughout the world and hence have long been of interest to soil scientists and geologists. In addition, the ability of a class of these materials, the smectite clays, to absorb a wide range of intercalants, both inorganic and organic, and act as catalytic agents, has made them of great interest to chemists (Pi 83). Recently, physicists have begun to study clays as they are layered materials with highly anisotropic properties, and provide an additional system, somewhat complementary to graphite, for the study of intercalation and two-dimensional phenomena (So 88). Both materials provide hexagonal ”cages” for the intercalants; however, whereas each graphite layer is a simple honeycomb lattice with only two atoms per unit cell, a single clay layer is a complicated structure with a large unit cell, consisting of both tetrahedral and octahedral molecular units. This makes the study of the dynamics of layered silicates, in particular the intercalant dynamics, more difficult as we must contend with a much larger number of force constants and isolate the low-frequency modes of greatest interest from the much more numerous and less important ”internal” modes of the clay. There have been previous studies of the zone center (3.0) modes, relevant to infrared and Raman 117 spectra, by Ishii et al. (Is 67) and Gupta et a1. (Gu 88). In this paper, we extend these studies to finite wavevector, but more importantly, we explain the results in terms of simple dynamical models and thus gain insight into the intercalant-clay interaction. Specifically, we derive approximate semi-empirical relations for the relevant low-frequency zone center modes (IR and Raman) in terms of the intercalant force constants and show how these models can describe the essential features of the phonon spectra both within and perpendicular to the clay layer. Although this conceptual "simplification" of the clay is the primary result of our study, we hope that the more detailed results, in particular the phonon dispersion curves we provide, will both encourage and guide those making measurements on these materials. The model clay used in this study is typical of many naturally occuring minerals. After presenting the crystal structure of these layered silicates, we describe the phenomenological force constant approach used to model the atomic interactions. We base our calculations on both bond-stretching and bond-bending potentials, but in contrast to the study by Gupta et al. (Gu 88), we avoid the use of tangential two-body bond-bending potentials as these give spurious and inconsistent results for both the phonon spectra and the elastic (long-wavelength acoustic) response. Using the eigenmode information from 3=0 calculations, we' 118 formulate a simple four-unit linear model for the clay (one for the intercalant, two for the tetrahedral layer and one for the octahedral layer) and derive the frequency relations described in the previous paragraph. 'We next present phonon dispersion curves along the symmetry axis of rotation (y-axis), within the plane of the clay layer, and along the reciprocal lattice vector (z-axis) perpendicular to the xy-plane. As a summation to the section on the crystalline materials, we present the results for the elastic behavior, illustrating the highly anisotropic nature of the crystal within the plane perpendicular to the clay layer and the nearly isotropic hexagonal nature within this layer. These results compare quite well with experimental Brillouin scattering measurements by Vaughan and Guggenheim (Va 86) and a pure continuum (macroscopic) model giving us confidence in the soundness of our microscopic approach. In the last part of this section we present a simple, heuristic model for the torsional mode frequency of a ternary alloy system. Using a probabilistic approach we derive expressions for the frequency in terms of the effective force constants. We introduce the effects of layer rigidity using simple rules that account for variations in local environment. To test these ideas, we simulate the disorder using the equation-of—motion method, described previously, on a two-dimensional intercalated Ragomé lattice, to calculate the Raman response of the 119 alloy. A good agreement with the above phenomenological model is found. 120 LL-lealatmmu. In Figure 1 we show the structure of a typical layered silicate. We define the x-axis to be within the plane in the direction of the bias (slant). The s-axis is perpendicular to the plane. Each layer consists of three parts, the upper and lower of tetrahedral units and the middle of octahedral units. Such materials are designated as 2:1 phyllosilicates. (There are also layered silicates with one tetrahedral and one octahedral section, the so-called 1:1 silicates.) The tetrahedral sections consist of AO‘ tetrahedra, connected at the vertices so as to form a planar Kagomé lattice, with the apex oxygen atoms forming a coplanar hexagonal lattice above. A hydroxyl group at the center of each hexagon gives rise to a triangular lattice providing the bottom faces of AO‘(OH)z octahedra. An inverted tetrahedral layer is then attached to the top faces of these octahedrals to complete the clay layer. The intercalants are positioned between two clay layers "encaged" by the hexagons of the Kagomé lattice. These crystals are monoclinic (space group C:h(C2/m)) with the upper tetrahedral layer shifted 2a/3 relative to the lower one where a is the 0-0 distance. Although there is freedom in the stacking of the layers, for the calculations presented in this paper the layers are placed such that the Kagomé lattice of the lower tetrahedral layer lies over the 122 Kagomé lattice of the upper tetrahedral layer below. The intercalant therefore interacts with the clay layers through twelve equivalent oxygen atoms. The structural properties, including the bond lengths, lattice vectors and density are summarized in Table 1. The (t) and (0) refer to the- 0-0 bond distances in the tetrahedral and octahedral layers. The 0-0 (t) bond distance is computed from the Si-O bond distance (1.62 A) taken from Liebau (Li 85). We now discuss the chemical species ' within the tetrahedral and octahedral units. These clays are classified in part by their octahedral components. In the dioctahedral mineral group, one out of every three octahedra is vacant while in trioctahedral clays, all such sites are occupied. Typically the tetrahedra contain silicon with an occasional partial substitution of aluminium. The octahedral sites contain either magnesium (tri), aluminium (di), or iron (di), exclusively or in mixture. For a listing of structural formuli, the reader is refered to Pinnavaia (Pi 83). In naturally occuring clays, the intercalant is typically an alkaline earth or alkali ion, often hydrated. The charge of the intercalant balances an equal but opposite charge within the clay layer giving rise 123 Table 1 Structural parameters of the model clay. Bond Distance (in A) 0-0 (t) 2.64 (E a) 0-0 (0) 2/153 = 3.05 Si-O 13/8a = 1.62 Cs-O V3/2a = 3.23 3 2a x 6 273a y -’ A A c 2/3a x + 3.99s 2 s p 3.34 gm/cm 124 to a fairly strong ionic component to the binding which results in stiff, brittle clays such as muscovite (mica). In the pyrophyllite-talc group, however, there is no intercalant due to charge neutrality of the clay layer itself. Clays of this type are very soft and powdery. For the calculations in this paper we use the structural unit cell formula Cs((Siz)Os}2[Hg3(OH)z], where the curly brackets represent the tetrahdral (silicate) layer and the square brackes, the octahedral (brucite) layer. This is not a charge neutral compound and does not occur in nature as such, but by neglecting the Coulomb force and the species variation within the clay, we can essentially reduce the number of atoms needed for a lattice dynamics calculation by half. As we shall see later, for the "soft” clays such as talc and the smectites of ”intermediate” stiffness, the correction to the low frequency zone center modes, where the ”long-range” Coulomb effects are greatest, is of a few percent. However, for the ”hard” mica-like clays, with large effective charges, the effect can be ocloser to ten percent. The reason for not including them at the beginning is that it is difficult to assign accurate effective charges to the individual atoms. Furthermore, an exact lattice dynamical calculation with Coulomb interactions would require the use of the Ewald method which would greatly complicate the computation for such a large unit cell material. (Note however that the ”short-range" part of the 125 Coulomb interaction is already included in the phenomological forces.) Finally we note that a mismatch between the natural lengths of the tetrahedral and octahedral layers causes the tetrahedra to rotate a few degrees, distorting the Kagomé lattice. (The planar layer goes from a C to a C symmetry.) We neglect this av 3v distortion in our present calculations. 126 ULWMMM In these lattice dynamics calculations we have used a combination of phenomenological force constant models to represent the bond-stretching and bond-bending components of the interatomic interactions. Bond-stretching is incorporated through a central-force term of the form: C II NI” 9 |_' '34 l 11 " I — 1151 — 1111]: (1) CF _ 1 z a - E a (ri uij) + 0(u ) + e where a is the radial force constant, r1 = R1 + 111 is the instantaneous position of the ith atom, rij is the unit vector between the two equilibrium particle sites, and 311 is the relative displacement of the two atoms from their equilibrium positions. Note that only motion along the equilibrium bond contributes to ch. We model the angular bonds with a three-body bond-bending potential first suggested by Resting (Re 66). This potential is particularly well suited for describing covalent bonds with strong directional character. It is given by: 127 z [(3 ° 3 ) - Rz cos6 ] (2) where r is the force constant, R is the equilibrium bond distance, and 6 is the equilibrium bond angle. Here, the ith particle is the vertex atom, the other two being the ”leg” atoms. Expanding to second-order in the displacements we get: + 2(1..- '13.) (r. - 3.)]1» 0(u”) (3) 13 1. 1.): 1.1 In our initial calculations, following Gupta et a1., we used a two-body bond-bending (de Launey) term (dL 56) given by: 4 + s C + s u ° u - s 2 - m ) - ’23 [111-1 ( r11 “11) ] (4) where 3 is the angular force constant and 61L is the displacement perpendicular to the equilibrium bond. (The central-force and de Launey potentials generate the so-called radial and tangential forces, respectively, and 128 when combined are sometimes described as the Born model.) Although easier to implement than the Keating potential, it suffers from a number of problems. First of all, the de Launey term is not rotationally invariant and the pivoting of a rigid tetrahedron, such as occurs with the torsional mode, coats energy not only due to the distortion of the basal plane but in addition, and incorrectly, due to the fixed rotation of the internal O-Si-O bonds. To compensate for this, Gupta et a1. used an unphysical, negative value for the 0-0 de Launey tern. With these simple two-body force terms, they were able to reproduce reasonably well the values of the 3:0 phonon frequencies calculated by Ishii et al (Is 67) who used the more complicated Urey-Bradley potentials. when we used this same force model to calculate phonon dispersion curves (qln) for the full clay, it led to spurious softening in finite wavevector modes involving motion of the octahedral layer. This is due to the instability introduced through the negative force constants. The lack of rotational invariance of the de Launey potential has an additional detrimental effect on the sound velocity results which we will discuss in greater detail in the section on the elastic behavior of the clay. The Keating potential described previously, however, is rotational invariant, and suffers from none of these disadvantages. For the reasons mentioned in this section, virtually all angular forces have been modeled using this potential. 129 These include the O-Si-O and the Si-O-Si bonds within the silicate layer, the O-Hg-O bonds within the brucite (octahedral) layer, the 0-0-0 bonds in the Kagomé basal plane and the O-Cs-O intercalant-silicate layer bond. The one exception is the O-H angular bond responsible for the librational motion of the hydroxyl unit for which we use the de Launey potential. The values for the silicon-oxygen and magnesium-oxygen force constants are taken from Ishii et :1. One shortcoming of our model is the absence of dihedral force terms that are of importance in the accurate calculations of isolated tetrahedron vibrational frequencies and thus of some relavance here. However, we are more interested in the intercalant-oxygen interaction and the (analytic) results we give are relatively insensitive to the precise values of the tetrahedral and octahedral bonds. It is only necessary that these be significantly stronger than the intercalant bonds. This requirement is well satisfied. (The Si-O bond-stretching force constant is nearly two orders of magnitude larger.) The values for the Cs-O and O-Cs-O intercalant bonds and the 0-0-0 basal plane bond are chosen to give ”reasonable" results for the low-frequency modes but we emphasize that it is not the values of the force constants in themselves that are important but rather the relationship between them and the frequencies. The hydroxyl group bonding parameters were determined by fitting the 130 typical librational and vibrational experimental values of 340 cm-‘.and 3675 cm-‘ respectively (Is 67). The force constants used in the calculations are given in Table 2. Once the geometry of the crystal has been described and the values of the force constants given, it is neccesary to construct and then diagonalize the dynamical matrix, which relates the force on each atol to the displacements of the other particles. Both these steps are carried out on the computer, the latter using a standard IHSL routine, the former using a program written by one of us that can be used on any structure whose dynamics are described by 2- and 3-body potentials. The output of this program, in the form of w.(3) (eigenvalues) and 3.(i) (eigenvectors), gives the phonon dispersion curves, elastic properties, and density of states of the model clay. In the following section, we describe the low frequency zone center phonons and analyze then with a simple model. 131 Table 2 Phenomenological Force Constant Values Bond Strength (in 10‘ dyne/cn) Si-O 35.08 Hg-O 3.20 Cs-O 0.40 H-O 75.50 O-Si-O 3.89 Si-O-Si 1.50 O-Hg-O 0.40 O-Cs-O 0.60 0-0-0 0.02 H-0 0.60 (2-body) 132 mmwummmm menfln In general the eigenmodes (particle displacements) of our full clay are quite complicated, given the large number (22) of atoss involved. However for the low frequency aodes, our lattice dynamics calculations show that to a first approximation, we can group the atoms into four structural units, indeed, the same four units described in the crystal structure section. These are the intercalant, the two silicate layers, and the middle brucite (octahedral) layer. Coupled together with effective force constants, this simplified model can be used to derive analytic expressions for the lowest infrared and Raman mode frequencies. The (approximate) displacement patterns are shown in Figure 2(a-d). The masses of the intercalant, silicate layer, and brucite layer are m (133 amu for Cs), I, (136 amu), and m. (107 amu), respectively. The effective force constants are designated by k and k I 3, the first giving the coupling between the intercalant and the silicate layer and determined directly from the structural geometry and the actual intercalant force constants, “Esq: and 7M“), while the second gives the coupling between the silicate and brucite layers and is determined empirically (although it is related to au94Y) Because the effective 133 Figure 2 Low Frequency Zone Center Hodes / \ u / \ e e e o my) /\ s *— m\ /\ / ‘ I \ / u kI \ / I /\ /\ /\ /\ l [I \v/ \ . / \v/ \ (xy) (2) . / \ “g k. / \ a /e\/e\ ‘ ‘— /e\/e\ + (xy) (3) x \ / ux kx \ / (fly) 0 I (s) O I I / \ k! / \ (a) ir (planar) (b) ir (z axis) / \ us \\\f///\\\f/// .———. ks \ / /\ /\ / V k. / \ u s /e\/e\ ' kx \ : / kI / \ (c) Raman (planar) (d) Torsional (Ra) (S) (B) (S) (I) 134 couplings are different for the planar and z-axis motion, we further distinguish these by the superscripts. We now show how the effective force constants are related to the bond strengths (Table 2), considering first the central-force term. Let 3 8 'J.‘ - 35 be the relative A displacement of the two particles defining the bond and r the unit vector along the bond direction. The force on the 1th particle is given by: F = -a (; ' 3) E. In the approximate modes we consider, the net restoring force is along the displacement vector, therefore we need only consider the force component along this vector. Because in all cases the displacement is along one of the coordinate axes, we have: Ft = -ar:u, where i = (x.y.z). Therefore the contribution per bond to the effective central-force coupling is given by kcIF = a rt. For the Resting potential, for the case where 3‘ = 3. = 3, the force on the vertex atom is given by: FK= -y I; - 3] E, where E a ;s+ ;.. For the angular bonds we consider, (r‘)x = -(r.)x, (rA)y = -(r.)y, and (rs): = (r.)'. Therefore, (Fx)x = (FR)y = 0 and (FR): = -47 r: u, hence, h“ = 47 r: 6“. where 61: is the Kronecker delta. The intercalant is bonded to a given silicate layer through six central-force and three angular bonds. The brucite layer, although distributed over seven atoms is also considered as a single unit and also coupled to the silicate layer, through the six Hg-O (apex) bonds. In both cases the six unit vectors defining the bonds are 135 given by: 3‘ = [:E, o, E] (5.) 32 = [:E, :12, Thus we get for the total effective force constants: II“ 1 (5... km" = 2 a (5.) I “-0 k‘“ - 2 a 4 (6b) 1 " as-o ro-cs-o (xy)_ km - 2 (am-8 (6°) (2) k. = 2 a._. . (6d) We see that the coupling in the x- and y-directions are equal and therefore the corresponding modes should be nearly degenerate. In addition, we find that the z-axis coupling is further stiffened by the action of the intercalant angular bond and hence this mode should be greater than the planar ones. From our force constant values we get: (z) ¢ xy) k1 - 0.8 x 10‘ dyne/cm and k = 3.2 x 10‘ dyne/cm. we now consider these modes in greater detail. Figure 2a shows schematically the planar mode in which the intercalant beats transversely against the clay layer. If we consider the intercalant to have a positive charge and 136 the clay layer to have a compensating negative charge, then we see that this motion sets up an oscillating dipole moment within the xy-plane and therefore couples to electromagnetic radiation polarized with wavevector along the s-sxis. Figure 2b shows the corresponding beating mode with the dipole noment along the z-axis. For both of these cases, the equations of motion for the intercalant and the brucite layer give: < xy.2) z 'xwxur - 2kI (u!+us) (78) _ (xy.z) _ m.wbu. - 2k. (u. us) (7b) where “1’ “3’ u. > 0 and «3 = w; 5 m. We define: us E n3 “1 and u. a n. “1' With these substitutions, dividing eqn. (7a) by eqn. (7b) gives: ( xy.m) 7, [11].; ("I-"'5’ m m m (ny) r k! (1 + n8) Because there is no center-of-mass motion for this a = 0 mode, we have: 2 m n + n n = m (9) Solving for n. and substituting into eqn. (8) gives the following quadratic equation for n.: 137 26-3-1 n: + {[263] - 1]- (ti-:1 + 11“} ., . {(2.33 - 1} = o (10) c xy,z) / k! ( xy,z) where K E (k m ). He deternine h. from the planar mode (Figure 2c) in which the silicate layers beat, in oppostion, against the stationary intercalant and brucite layer. Because no dipole moment is generated this mode is not IR active, but because the polarizability of the electrons within the silicate-brucite-silicate layer are different at the 90' and 270' phases of the motion, this mode is Raman active. From the equation of motion of the silicate layer for this mode we get: 3 = m who I . (11) The (average) frequency for this mode is 88.0 cm", therefore we get k:‘y’ = 5.40 x 10‘ dyne/cm, implying that ch. 8 2.70 x 10‘ dyne/cm. This value is reasonable given that dug-o =- 3.2 x 10‘ dyne/cm. lots that this is also the (2) value of k- Returning to eqn. (10), we have: KW = 6.75 and Kw = 1.69, which gives (from the quadratic formula using the negative branch): 77:”) = 0.33 and n;”'= 0.25. From eqn. (7a) we have: 138 1/2 2k! (DIR = ;— (1+Y)8) (12) I thus (0:1)" = 52.1 cm_1 and not; = 101.0 cm”. These should be compared with the actual values from the full lattice dynamics calculation: 0);?" = 51.2 cm.1 and (2) I. = 99.6 cm-1. ‘The relative errors are 1.76% and 1.41%, respectively. Considering the simplicity of the model this agreement is very good. More importantly, eqns. (6) and (10)-(12) can be used to extract the values for the phenomenological intercalant force constants, 0110-0 and ro-rc-o (and as-s)’ from experimental measurements of the low frequency zone center modes. If we simplify the model further by considering the silicate-brucite-silicate layer to love as a single unit, we have: (xy,z> z mxcoxuI - 2kI (1+7?) uI (13) where n = 'r/(H-'1) = 1.35 (M = 512 amu). This gives 0);” s 52.5 cm“ and not; = 105.0 cm“. The first agrees quite well with the previous value while the second is quite a bit different. We infer from this that modes in which the particle motion is perpendicular to the plane probe to a greater degree, the internal structure of the clay layer, which is certainly reasonable. 139 We now present a derivation of the frequency of the torsional mode. This node is unique in that its frequency can be calculated exactly due to the simplicity of the eigenmode, shown in Figure 2d. We see that only the oxygen atoms in the silicate basal plane love, and only within the plane. Concentrating on the coupling of the oxygen atoms to the (fixed) intercalant, we find there is no contribution from the intercalant-basal plane angular force and a contribution to the total effective force constant from the intercalant-oxygen stretching bonds of §~a . For the Cm-O planar 0-0-0 angular bonds of the Kagoné lattice we have: :.=[:-.::.01 was») [we] - ransom“ as] 35 [[w] - ra-sonu- es) where u is the magnitude of the displacement of a given oxygen atom. To find the contribution from these bonds we first calculate the potential energy of this distortion from - 1 3 eqn. (3). We have. vo-o—o - 2 [92' ] u . We have two of these bonds per basal plane oxygen, thus the effective 140 torsional mode force constant is given by: 4 k - 5- ace-o + 18 y -o (15) and 2 (TOP) cor" - k / “Oxy . (16) Note that from eqn. (15), we see that the basal plane angular force is much more effective than the stretching force in stiffening this mode. In addition, if “c.43 has been previously determined, then eqns. (15)-(16) will give the value of ro-o-o‘ Finally we address the problem of the the Coulomb forces. The "long-range" portion of this interaction gives rise to an additional force on the particles due to the macroscopic electric field generated by the particle displacements. We can estimate this effect quantitatively by noting that the shift in frequency (squared) is of the order of the plasma frequency (squared), wi, given by (4n/0) Z (qt/mi), where 0 is the unit cell volume (equal to 2.54 x 10.22 cm“) and the sum is over the atom in the cell. For our purposes, we consider the clay to consist of two dynamical units, the intercalant with a mass of 133 amu (Cs)and the remainder (silicate-brucite-silicate layers) with mass 379 amu. Assuming a relatively small charge transfer‘ between the two units of 0.4 e (e = 4.3 x 10'“’ statcoulombs), we get Aw z:w: / 2w equal to approximately 141 1.5 cm-‘ for a) = 75.0 cm”, a typical frequency of interest. For muscovite, where the effective charge is as large as 1.0 e, we get a shift of about 9.5 on" for the same base frequency. For such materials it is probably advisable to include the Coulomb interactions. Note however that not all modes will be affected. For example, the Raman active frequencies do not create a dipole moment and thus do not produce an electric field to couple to the charges. The results of this section could still be applicable to these materials if it were possible to subtract the contribution of the long-range forces so that we could work with only the ”bare” frequencies. This is relatively simple for cubic (or isotropic) structures (once the effective charges are known) where, for the zone center modes, we have shown, a): = w: - 3 w: and a): I to: + go): where T and I. refer to the transverse and longitudinal polarizations and ab is the frequency without the charges. For our monoclinic structures, however, we are unfamiliar with the corresponding expressions. We stress though, that overall, many of the results of this paper are qualitative and quite robust, with relevance to materials of both small and large ionicity. Lzhamnimnimmms. We now discuss the finite wavevector response of the layered silicate, shown in Figure 3(a,b). Although our clay 142 has over sixty branches, we will consider only a few, concentrating entirely in the region below 250 cm”. For a solid such as this, consisting of many weakly bound structural units, most of the modes will be relatively dispersionless and more characteristic of the seperate units themselves. Our interest is with modes that involve inter-unit motion, in particular, the motion of the intercalant against the clay layer. Through the use of simple models, including the ones introduced in the previous section, we can explain many of the prominent features that occur in the phonon dispersion curves. We first consider the phonons with propagation vector perpendicular to the clay layer, along the z-axis. One striking feature, that is characteristic of phonons in layered materials, with wavevectors perpendicular to the planes (Be 88), is the large number of virtually dispersionless modes. But in addition, there are two groups of curves, one in two sections, the other in four, that are characteristic of the zone-folded acoustic mode of a linear chain. Specifically, these features are well described by the two-unit (intercalant - clay) and four-unit (intercalant - silicate - brucite - silicate) chain models used previously. These two models relate to the transverse and longitudinal modes of the full clay respectively. The dispersion curves from these models, using the same effective force constants determined previously from the 143 zone-center analysis, are plotted on Figure Be, as dashed lines for the first model and as dash-dotted lines for the second. We see that the acoustic modes are very well reproduced, the lower optic modes moderately reproduced, and the higher optic modes only qualitatively reproduced. However, in general, the approximation of grouping the individual atoms into larger dynamical units remains useful even when extended to finite wavevector. We next consider phonons with wavevector within the clay layer along the two-fold rotation axis. The dispersion curves as shown in Figure 3b demonstrate a quite different behavior from those along the z-axis, as one might expect for an anisotropic material. In the previous case, the phonons probe the interaction between the layers, ignoring for the most part, the structure within a given layer. In this case, however, the phonons are greatly influenced by the interaction of the individual intercalants with the hexagonal "cage” of the clay, and therefore the detailed structure of the intercalant - silicate basal plane boundary must be adequately modeled. Indeed, to first approximation, we can concentrate exclusively on the dynamics of a single interface. The atoms of the clay layer are strongly connected to each other as well as being coupled indirectly through their weaker bonds to the intercalant which are placed in the ”interstitial” regions. The model structure having these features is shown in Figure 4. The values of (1) (cm“) 250 200 150 100 144 Figure 3(a) .10 0.15 0.20 qa/ 1t w(k) for k = (0,0,k) 0.25 (0 (cm") 145 250 200 150 - ’7/ qa/ K Figure 3(b) w(k) for k = (O,k,0) 146 the stretching and angular clay-intercalant bonds are chosen to give the infrared zone-center frequencies calculated using the two-unit model (eqn. 13). The ”atoms” within the clay layer are coupled via central-force and de Launey terms '(tbe latter is neccesary for transverse rigidity) with strengths equal to the 81-0 bonds. Note that this model allows for both transverse and longitudinal modes. The resultant dispersion curves are overlayed on Figure 3b with the transverse modes displayed in dashed lines and the longitudinal in the dash-dotted pattern. The characteristic features of the acoustic and optic branches of this model are clearly present in the full clay. The quantitative agreement with both the transverse and longitudinal optic modes and the longitudinal acoustic mode is good. The characteristic optic and acoustic branches can be explained as a hybridization between the acoustic mode of a monoatomic linear chain made up of the larger ”clay” atoms and the dispersionless optic mode characteristic of isolated intercalants vibrating between them. The interaction between these two modes that gives rise to the aforementioned hybridization leads to the greatly flattened acoustic mode and the strongly dispersive, sigmoidal shaped, optic mode. We can conclude that one of the characteristic structural features of the clay, namely, the loose ”traping” of the intercalant in the hexagonal holes of the basal plane has a definite effect on the underlying structure of the 147 \ / lntercalants Figure 4 Structure for y-axis Dispersion Hodel 148 phonon dispersion curves for in-plane propagation vectors. The one shortcoming of this model is its failure to properly describe the transverse acoustic modes of the full clay. Obviously, it is neccesary to treat the transverse rigidity of the clay layer with more subtlety than this simple ”ball and spring” model allows. We consider these dynamics, in particular the softening of the transverse acoustic mode with particle displacement perpendicular to the layer (TAL), in greater detail in the second part of the next section. To give a general overview of the lattice dynamics we plot the density of states for all modes less than 1000 cm'4 in Figure 5. This was calculated by the usual technique of calculating the frequencies for a given k chosen at random in the Brillioun zone and creating a histogram. We took a sample of 5000 points in reciprocal space which gives approximately 100,000 total frequencies. We neglect in this graph the single peak at 3676 cm-‘ due to the O-H stretching mode of the hydroxyl. It shows a number of sharp peaks as a consequence of the relatively dispersionless intra-molecular modes. Just below 1000 cm-‘:Ls the peak due to the Si-O bond stretching of the tetrahedron. At about 330 cm-‘ is the O-H librational mode which has a strong neutron signal due to the presence of the hydrogen. Heasurements by Reumann et a1. (Re 89b) show this structure is sensitive to the presence (or lack of) vacancies in the octahedral layer giving a two-peak structure for dioctahedral phylosilicates 149 such as montmorillonite and a single-peak for the trioctahedral type materials (this model). At present, the simple de-Launey force model used for the O-H bond is too crude to duplicate this behavior. It appears possible that the double peak is a "mixed-crystal" effect as the number of vacancies might vary from cell to cell and give rise to different local environments for the hydroxyl group and hence slightly different frequencies. Finally note that there is a strong peak at about 50 cad due to the flat longitudinal acoustic mode, essentially due to the intercalant "rattling" within the hexagonal cage. 150 1 1 1 .J_ 1 L 1 l o 200 400 500 800 1000 (0 (cm") Figure 5 Density of States of Model Layered Silicate 151 LWW Elasticity theory is concerned with the long wavelength acoustic dynamics of a solid. In this regime, A >>a where a is the typical atomic separation and A the wavelength of the phonon, and all atoms are moving essentially in phase. Therefore the relative distortion of the atoms within a given unit cell is small and as a result, we can ignore the discrete nature of the material, describing the distortion from equilibrium by the "average” or center-of-mass motion of a macroscopically small though microscopically large region. If we assume a linear relation between the distortion (strain) and the resulting restoring force (stress), the harmonic normal modes of the system are given by the solutions of the secular determinant (Lo 34): A A 2 ' cijkl. kjkk " 9" 65,1, I = 0 (17) (ijkl = mm) A where ci is the elasticity tensor, k5 is a component of jkl the unit vector along the propagation vector, p is the mass density, and v is the velocity of propagation. Because each indice can take on three different values, the "dynamical" matrix in this case is 3 x 3 with three eigenvectors (and values) corresponding to the three acoustic branches. Although the tensor is of fourth-rank and has 81 components, 152 the number of independent elements is twenty-one for the most anisotropic crystal and only thirteen for the monoclinic clay structure. Furthermore it is possible to group pairs of indices and identify these by a single number. We use the convention: xx +4 1, yy +4 2, zz +4 3, yz +4 4, zx +4 5, xy +4 6. In this form, the components of the elasticity tensor are called the elastic moduli. In principle, these moduli can be related analytically to the force constants (see Born and Huang (Bo 54)), although this procedure is difficult to carry out for all but the simplest lattices. Alternately, they can be determined experimentally through ultrasonic or Brillouin measurements (Va 86). We, however, can determine these velocities from the slopes of the acoustic modes, generated by our full lattice dynamical calculations. This data, for our clay, is presented in Figure 6a, for propogation vectors both within and perpendicular to the layer. Note the nearly isotropic behavior within the layer (x to y) and the highly anisotropic response between layers (z to -z). As a result of monoclinic (hexagonal) symmetry, one can show that the two lowest velocities along the z-axis must (nearly) correspond with those from the lowest curve along the x- and y-axes. It was the failure of our previous models with de Launey angular forces to satisfy these symmetry requirements, that alerted us to the problems with these potentials. In fact, the Keating potential was first 153 introduced precisely to correct this very problem with the de Launey (Born) dynamical models, although it was the simpler diamond structure that was initially considered (Re 66). In Figure 6b, we reproduce the acoustic velocity data from Vaughn and Guggenheim for muscovite, a common mica. (N.B.: The y-axis in Figure 6b is incorrectly labeled. The velocities are actually in km/s.) The symbols are the measurements (from Brillion scattering) and the solid curves are the calculated velocities from an elastic constant model. That is, the elastic moduli are determined from a subset of the experimental measurements and these moduli then inserted in eqn. (17) to determine the velocities (eigenvalues) for any given k. The fact that our curves, computed from the full dynamical matrix, are quite similar, gives us confidence that our sicroscopic model is capable of correctly describing the dynamics of these layered silicates. For our structure with no torsional distortion, the simpler hexagonal symmetry model (with only 5 independent constants) would have been sufficient, but in general this will not be the case. Indeed, the experimental data shows greater anisotropy within the plane for a real material where this distortion is present. We now briefly sketch how the elastic moduli (for. the hexagonal case) are determined from the propagation velocities. The value for cum: cu) (in gigapascals) is 154 7 L a s — E §. > i 3 _ ' i I . I : 1 ' 1 1 l l x y z x -2 Figure 6(a) Sound Velocity as calculated from Model .0 1 ‘ 1 D 1 1 1 C. 1 1 L I “C. 0?: o m o Mnntumu M0000 " no | =1 ; \ é ' ‘~ * I \\.. j IO// U . 2' ~ 2 {l \ 3:1] \‘1 i .. 2/ a = 7 \ " g . “% > . 3 . 3 .f E Tunsverie Modes I -0 '0 1° 00 IO ’0 3 =0 )0 0| CO ’9 0 '0 80 4O 00 ’0 00 'Ol '10 '30 10° '00 'IO Acausuc Propagauon Direcuon (Dogma from Ann Figure 6(b) Sound Velocity for Muscovite as measured by Brillioun scattering 155 where p is the density in kg/ma and iven b 2 8 Y 9" Lazy) V Lucy) is the velocity of longitudinal waves within the xy-plane in 2 meters per second. Likewise, c” - pvu” where vb“) is the velocity of longitudinal waves along the z-axis. From the transverse waves with wavevector within the plane, we 2 2 get c“ - pvflxy) and c“ Figure 9 Two-Dimensional Kagomé-Intercalated Lattice 165 1;. WMWLLMW From Figure 8a, we see that even with alloying, the torsional frequency can be defined, to a first approximation, by a single number giving the peak position. (It is also characterized by the width of the response, which we ignore, however.) From the point of view of the oxygen atoms, in the crystalline limit (x=0 or 1), the intercalants act as fixed posts between which they oscillate back and forth along the line defined by bonds. This frequency is given by: ‘0 w = V'2k / m (1) Tor oxy where k is the O-IC (stretching) force constant and moxy is the oxygen mass. Note that the intercalant masses do not appear as they are stationary, nor does the O-O coupling since all oxygen atoms move in phase so as not to distort this bond. We now randomly substitute different intercalant species. If we consider the oxygen to be independent of one another, then each would oscillate at a frequency given by eqn. (1) but with k‘ + k2 substituted for 2k, where k and I. k2 are the two intercalant-oxygen force constants, not neccesarily equal. (We have implicitly assumed that the intercalants remain essentially stationary even after the alloying. This is strictly true for the pure cases, x=0 166 and 1, however it remains a good approximation for intermediate x.) We now assume that the major effect of the mutual oxygen bonding, related ultimately to the internal silicate bonding, is simply to average the various force constants so that the effective frequency of the torsional mode in the alloy is given by: 1.1”” = V2 11(111/ m (2) Tor o oxy where k.(x) is the effective force constant at concentration x, given by the average force constant k. We can make this approximation because the 0-0 force constant is much larger that the fluctuations in the intercalant—oxygen coupling due to the alloying. In the case where the O-IC force constants retain their pure case values k and k , we have: A a k.(x) = k = (1-x) k‘ + x k. . (3) This gives rise to a nearly linear relation between a)?“ and x. We can now introduce the important effects of layer rigidity via the force constants. In Figure 10a, we show schematically, an infinitely rigid clay layer stop to large intercalants, labeled A, with a smaller intercalant B 167 between them. In this case, the equilibrium bond distance for the O-A bond remains unchanged from its value in the pure system. However, for the O-B bond, this distance is increased due to the pillaring effect of the surrounding A intercalants. The effective coupling is weaker than in the pure case and thus he should be replaced by ‘1: km where I. is a reduction factor less than or equal to one. If the clay layer were completely "floppy", as shown in Figure 10b, then it (the layer) would conform to the various heights of the different intercalants in such a way that the couplings would always be given by their pure case, the so-called ”Vegard's Law" regime. We incorporate these ideas into our model by adopting the following rules. If a 8 atom has an A atom as its nearest-neighbor on the (triangular) intercalant sub-lattice, then the O-B force constant along that bond is reduced by the factor f.- Since the intercalants are arranged randomly, the probability of a B atom having an A atom neighbor is given by (l-x). The corresponding probability of having a B atom neighbor is of course given by x. The O-A force constant, at this stage, is unaffected by its local environment. Thus we have: k.(x) = k (1-x) k‘ + x [(l-x)#. + x] k. 1‘ {(l-x) + [#.x + (1-£_)ler1 (a) 168 where r a (k./kA) < 1. We plot the corresponding frequency w (x) = 72k (x)/m in Figure 11a for r=0.6 and #?80.5 e e oxy I ( ) and r=0.8 and fin=0.25 (- - -). Both show sub-linear behavior with the second exhibiting a minimum for x < 1. (We will explain the origin of the symbols in the next section.) It is easy to see that the value of k.(x) given by eqn. (4) is always less than that given by eqn. (3). If a larger A atom is surrounded, totally or in part, by smaller B atoms, this can lead to a pillaring effect which produces an enhancement in the O-A force constant due to the pushing down of the layer on the intercalant, then we get: k.(x) = k‘ [ (l-x) {(1-x)°1 + xf‘} + k f!- x{x-1 + (l-x)fi.}] (5) A where we have assumed this effect is the same for all six oxygen-intercalant bonds about the A atom I and is proportional to the number of B nearest-neighbors on the intercalant sub-lattice. The enhancement factor 4‘ is greater than or equal to one. In order for a polynomial to show both sub- and super-linear behavior, it must have a point of inflection, that is, the second derivative must vanish at some value of x. Since eqn. (5) is only second order (quadratic) in x this is impossible and thus can give only pure super- or sub-linear behavior. In Figure llb we 169 ”Clay" 0 Intercalants Figure 10(a) Infinitely Rigid Clay Layer "Clay" lntercalants Figure 10(b) Floppy Clay Layer 170 choose the parameters r=0.6, £‘=1.6, £.=0.5, to give a super-linear variation with x. Our basis approach can continued, considering, for example, the effects of the next-nearest-neighbors, which gives a cubic equation for k.(x), with each step leading to a higher-order polynomial approximation and richer behavior for the w vs. x curve. The cost is that new parameters are introduced at each stage. Nonetheless, we can test for the validity of this parameterization procedure and the effective force constant approximation, eqn. (2), by comparing our predictions against the actual behavior of the model system, calculated using the equation-of-motion (EOH) method. Lilnlamfisulm We use a 32 x 32 intercalated Kagomé lattice with 1024 sites consisting of 256 intercalants and 768 oxygens. The intercalant were placed randomly within the hexagonal sites for x from zero to one in increments of 0.1 and the force constants determined using the local environment rules presented in the previous section. We assume that the torsional mode induced in the alloy is the same periodic torsional mode as in the pure system. This is equivalent to assuming the polarizability of the intercalant bonds is unchanged by the alloying and the same for both O-A and O-B bonds. The torsional frequency was determined using the peak height of the response, an example of which is shown in 171 I l T I I I I l I 1'0 b-O.6 r-o.so ‘ —— r-O.8 f-0.25 0.9 - o 3 \ 3 0.8 r 0.7 1 1 n 1 1 L 1 1 1 0 0 2 0 4 0.6 O 8 1.0 X Torsional Mode Frequency vs x (Theory) Figure 11(a) (Sub-linear Behavior) 172 1 . 0. w/w0 0 . o.711h111111 Figure 11(b) Torsional Hode Frequency vs x (Theory) (Super-linear Behavior) 173 Figure 12. These results are represented as the individual symbols on Figure 11(a-b) for the same choices of parameters. The fits are rather good although both the sub- and super-linear model overestimate the actual frequencies for x < 0.5. Note that the minimum for the dotted curve in Figure 11a is rather nicely reproduced. The resolution in the figures is given by $0.03 frequency units. The simple assumptions we made in deriving our probabilistic model appear to be reasonable ones although admittedly this model provides more of a convenient (but physically motivated) parameterization of the problem rather than a fundamental explanation, which could be provided, for example, by expressions for the adjustment factors I‘ and I. in terms of the local spacing between layers. It should be emphasized that these numerical calculations are exact (within the limits of the model) and make no assumptions about fixed intercalants or oxygen-oxygen bonds. To demonstrate the accuracy of this method we plot in Figure 13, the density of states for a pure, periodic Kagomé-IC lattice using both the direct diagonalization of the 8 x 8 dynamical matrix sampled throughout the Brillioun zone and the EOH using a 56 x 60 supercell. All features are quite well reproduced by the simulation. 174 P — "' X I d O > I— P h. O4 06 0.8 10 12 14 (1)/(Do Figure 12 Raman Response for Torsional Hode calculated with EOH 175 5 I I I I I I I r I I I 4 - - Diag —°-— EOM 3360 3 .- Figure 13 Density of States for Kagome-IC calculated via Direct Diagonalization and EOH 176 Him In this part of the thesis we have surveyed the lattice dynamics of a typical layered silicate, with particular emphasis on the intercalant-clay layer dynamics. With the help of simple models we have developed approximate formulas for the low frequency zone-center modes in terms of the intercalant-basal plane force constants providing a procedure for these quantities to be determined from experimental data. We have extended these calculations to finite wavevector and studied the elastic properties of these materials for the first time from a microscopic viewpoint with a particular emphasis on the dispersion of the transverse acoustic mode. In the second part we have introdued a simple model for the torsional mode frequency for ternary alloy systems that we have checked with exact simulation results. In general, much work needs to be done in the experimental realm and it is hoped that this study will encourage experimentalists, expecially those involved in neutron spectroscopy, to make more detailed measurements of this class of layered solids. BIBLIOGRAPHY Ab Be 80 Ch dL dL Dw E1 Ga Ga Ga Ga Gu Is 75 88 54 71 56 80 61 71 76 83 89a 89b 88 67 177 REFERENCES R. Alben, D. Weaire, J.E. Smith, Jr., and M.H. Brodsky, Phys. Rev. B 11, 2271 (1975) M. Bernasconi et al., Phys. Rev. B 354 12089 (1988). M. Born and X. Huang, Dynamical Theory of Crystal Lattices (Oxford University Press, London, 1954). I.F. Chang and 8.8. Mitra, Adv. Phys. 20” 359 (1971) J. de Launey, Solid State Physics, Vbl. 2, edited by F. Seitz and D. Turnbull (Academic Press, New York, 1956), pp. 219-303. S.W. de Leeuw, J.W. Perram, and E.R. Smith, Proc, R. Soc. London, Ser. A 111, 27 (1980). H.B. Dwight, Tables of Integrals and other Mathematical Data, 4th ed. (Macmillan, New York, 1961). .R.J. Elliot, J. A. Rrumhansl, and P.L. Leath, Rev. Mod. Phys. 464 465 (1971). F.L. Galeener and G. Lucovsky, Phys. Rev. Lett. 324 1477 (1976). F.L. Galeener, A.J. Leadbetter, and M.W. Stringfellow, Phys. Rev. B 214 1082 (1983). J.M. Gales, B. Boltjes, M.F. Thorpe, and S.W. de Leeuw, (submitted to Phys. Rev. B) (1989). J.M. Gales (unpublished). H.C. Gupta, S.D. Mahanti, and S.A. Solin, Phys. Chem. Minerals 16, 291 (1988). M. Ishii, T. Simanouchi, and M. Hakshira, Inorganica Chemica Acta 11;, 38 (1967). La Li Lo Ly Ja Ke Ma Mc Pi Pi So Th Th 82 85 34 41 74 66 70 71 88 89 89b 77 83 88 76 86 178 Landolt-Bérnstein, Numerical Data and Functional Relationships in Science and Technology New Series Group III Vol. 17a, (Springer-Verlag, Berlin, 1982). F. Liebau, Structural Chemistry of Silicates, (Springer-Verlag, Berlin, 1985). A.E.H. Love, A Treatise on the Mathematical Theory of Elasticity, 4th ed. (Cambridge University Press, Cambridge, England, 1934). R.M. Lyddane, R.G. Sachs, and E. Teller, Phys. Rev. 524 673 (1941). J.D. Jackson, Classical Electrodynamics, 2nd ed. (Wiley, New York, 1974), Chap. 7. P.H. Keating, Phys. Rev. 151, 637 (1966). J.B. Marion, Classical Dynamics of Particles and Systems, 2nd ed. (Academic Press, New York, 1970). A.A. Maradudin, E.W. Montroll, G.H. Weiss and J.P. Ipatova, Solid State Physics Supplement 3 (Academic, New York, 1971), Chap. 6. T.C. McGlnn, M.V. Klein, L.T. Romano and J.E. Greene, Phys. Rev. B 11, 3362 (1988). Newman et al., Phys. Rev. B 124 657 (1989). Dan Neumann (private communication). R.M. Pick and M. Yvinec, in Proceedings of the International Conference on Lattice Dynamics, Paris, 1977, edited by M. Balkanski (Flammarion, Paris, 1977). T.J. Pinnavaia, Science 2294 365 (1983). S.A. Solin and H. Zabel, Adv. Phys. 11, 87 (1988). M.F. Thorpe, in Physics of Structurally Disordered Solids, 1976, edited by S. Mitra (Plenum, Mew York, 1976). M.F. Thorpe and S.W. de Leeuw, Phys. Rev. B 11, 8490 (1986). Va Ve Wa We Yo 86 67 86 70 85 179 M.T. Vaughan and S. Guggenheim, J. Geophysical Res. 21, 465 (1986). L. Verlet, Phys. Rev. 112g 98 (1967) R. Wada, W.A. Kamitakahara, M. Suzuki, in Proceedings of Symposium K, 1986 Fall Meeting of the MRS, edited by M.S. Dresselhaus, G. Dresselhaus, S.A. Solin (Materials Reasearch Society, Pittsburgh, 1986), p. 174. Wert and Thompson, Physics of Solids, 2nd ed. (McGraw-Hill, New York, 1970). B.R. York, S.A. Solin, N. Wada, R.H. Rayathatha, I.D. Johnson, T. Pinnavaia, Solid State Commun. 11‘ 475 (1985).