v ‘— _—‘l ‘7 ! HEEWR. LIERAfir Michigan State University This is to certify that the dissertation entitled Interaction-induced Properties and Perturbation Theory presented by Jesfis Juanés i Timoneda has been accepted towards fulfillment of the requirements for AD— degree in Min!— .0 Major professor Dam_the_22nd_n£_September 1986. ucn' -_ Ana.— ..- . - w 0-12771 MSU LlBRARlES -_—. RETURNING MATERIALS: Place in book drop to remove this checkout from your record. FINES will be charged if book is returned after the date stamped below. INTERACTION-INDUCED PROPERTIES AND PERTURBATION THEORY By Jesus Juanés i Timoneda A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1986 ABSTRACT INTERACTION-INDUCED PROPERTIES AND PERTURBATION THEORY By I O 0 Jesus Juanos 1 Timoneda The perturbation expansions obtained from Lowdin’s projection-operator formalism are derived in a new way, using Kato’s formulation of perturbation theory. This formulation does not involve the symbolic use of the inverse of singular operators. Kato’s approach provides a convenient algebraic alternative to diagrammatic techniques for obtaining eigenvalues and eigenvectors. Different normalization criteria imposable on the wave function are easily visualized in terms of the operator that yields the perturbed state vector when it acts upon the unperturbed wave function. We use a label-free exchange perturbation method to calculate the dipole moment of interacting Re and H atoms as function of internuclear separation. In the label-free formalism, the unperturbed Hamiltonian and perturbation terms are constructed so that each is invariant with respect to exchange of electrons between the interacting atoms; then a Rayleigh- Schrodinger perturbation expansion with a fully antisymmetrized set of zeroth—order wave functions yields the interaction energy and collision- induced properties. Good agreement with ab initio results for the He...H dipole is obtained when a long-range dispersion contribution is added to the first-order overlap and exchange contributions. Jesus Juanos i Timoneda We calculate the dipole moment for the quartet state of H3 in its linear and right triangular configurations and we find that the first- order overlap-exchange and the long-range contributions are about the same order of magnitude at intermediate range. We prove that a system of n atoms in any spatial configuration has zero dipole moment when the electrons are described by s-type functions (Slater or Gaussian) and the dipole moment is calculated as the expectation value of the corresponding operator with an unperturbed, but fully antisymmetrized wave function. An analytic expression for the damped pair dipole moment is calculated in terms of Clebsch-Gordan coefficients and reduced matrix elements. We show explicitly with the first nonlocal hyperpolarizability how susceptibility densities may be cast in a form without secular divergences. The theory of generalized functions is applied to calculate the limits in the expressions for susceptibility densities and tensors. ACKNOWLEDGMENTS This dissertation is an account of the research perfOrmed under Professor Katharine L.C. Hunt’s direction. I would like to express my appreciation and gratefulness for her guidance and advice. Also, I acknowledge the financial support of the Donors of the Petroleum Research Fund, Standard Oil ($0810) and "la Caixa de Barcelona". The computations involved in this research have been carried out at a time of expansion of the computational facilites in the Chemistry Department. My thanks to those who helped me in making efficient use of them. TABLE OF CONTENTS LIST OF TABLES LIST OF FIGURES CHAPTER 1. CHAPTER 2. CHAPTER 3. CHAPTER 4. CHAPTER 5. INTRODUCTION DERIVATION OF PERTURBATION EXPANSIONS 2.1. Introduction 2.2. On Lowdin’s studies in perturbation theory 2.3. Theoretical Analysis 2.4. Discussion 2.5. Relationships among perturbation methods INTERACTION-INDUCED PAIR DIPOLE MOMENT 3.1. Introduction 3.2. Exchange perturbation theory in label-free form 3.3. Comparison with other exchange perturbation theories 3.3.1. Polarization and exchange contributions to interaction energies and pair properties 3.3.2. Density-matrix perturbation theory 3.4. Computational methods 3.5. Results and discussion INTERACTION-INDUCED TRIPLET DIPOLE MOMENT 4.1. Introduction 4.2. Dipole moment calculations 4.3. Results and discussions OVERLAP DAMPED PAIR DIPOLE MOMENT 5.1. Introduction 5.2. Generalized functions and charge-susceptibility iii densities 5.3. Nonlocal polarizability and hyperpolarizability calculation 5.4. Damped dispersion dipole moment calculation APPENDIX A APPENDIX B APPENDIX C APPENDIX D APPENDIX E APPENDIX F REFERENCES iv LIST OF TABLES TABLE CHAPTER 1 3 Orbital exponents and contraction coefficients of the Gaussian basis (103 6p) on each atom. FIGURE 1 CHAPTER 3 LIST OF FIGURES Collision-induced dipole of He...H as a function of the internuclear separation R in the range from 4.0 to 11.0 a.u. Curve A shows the ab initio results obtained by Meyer, and Curve B the ab initio results of Ulrich gt a1. Curves C and D have been obtained by adding the long-range dispersion dipole to the exchange-overlap dipole computed form the energy derivative (Eq. 13). Curve C shows the dipole in the Gaussian basis, and D the dipole in the Slater basis. Curves E and F have been obtained similarly, but the exchange-overlap dipole has been computed as an expectation value from Eq. 12; Curve E shows results from the Gaussian basis, F from the Slater basis. Curve G is a plot on the leading term of the dispersion dipole, D7R-7. The ordinate is scaled logarithmically and the dipole p(R) is given in a.u. Collision-induced dipole of He...H as a function of internuclear separation R in the range from 0 to 4.0 a.u. Curve A shows the ab initio results of Ulrich gt a1. Curve B and C show the lowest-order vi exchange-overlap dipole, obtained from Eq. 13, B in Slater basis and C in Gaussion basis. Curve D shows the exchange—overlap dipole computed as an expectation value with the zeroth—order wavefunction; results from the Gaussian and Slater bases superimpose on this scale. The dipole p(R) is given in a.u. Ratio of the correction term Apz to the exchange- overlap dipole for He...H, plotted as a function of the internuclear separation R (in a.u.). The exchange-overlap dipole has been computed as the derivative of the first-order interaction energy with respect to an applied field. Curve A shows the results from the Slater-basis calculation, and Curve B the results from the Gaussian basis. Spatial configuration of three identical nuclei (3, b, c). e=a, "/2, "/3 for the linear, right-triangle and isosceles—triangle configurations. H dispersion dipole moment. X component. Right- 3 triangle configuration. H3 dispersion dipole moment. Z component. Right~ vii triangle configuration. H3 overlap-exchange dipole moment. X component. Right-triangle configuration. H3 overlap-exchange dipole moment. 2 component. Right-triangle configuration. H3 overlap-exchange dipole moment. Linear configuration. H3 dispersion dipole moment. Linear configuration. \llll CHAPTER 1. INTRODUCTION. A major part of this work is devoted to the analysis of interactions between atoms or molecules at separations such that the overlap is not negligible. From this analysis we derive methods to calculate interaction-induced properties. These properties allow us to predict the behavior of matter interacting with radiation; alternatively the calculated properties may be compared with those computed from experimental data. We have also worked on several theoretical aspects of the methods used in our study. Perturbation theory is the framework within which most of our work has been performed. The next chapter is devoted to a theoretical analysis of perturbation theory. 0n the basis of an entirely algebraic formalism, we present formulae to calculate corrections to the perturbed wave function and eigenvalue to any order. Several normalization criteria are associated with the way in which different projection operators are manipulated. We use a reduced resolvent for which no problem about singularities arises. This powerful technique allows us to establish consistency with other formalisms and developments in perturbation theory. We conclude the second chapter by presenting a unification of several different formalisms, with common principles. In chapter 3 we use a special form of perturbation theory, yi§., exchange perturbation theory in label-free form, to calculate interaction- induced dipole moments. In particular, the collision-induced dipole moment of a diaton is computed. We compare our method with other perturbation theories and we analyze an approximation employing additive long- and short-range contributions to the pair dipole moment. Since the f0 Hellmann—Feynman theorem is not satisfied at the level of approximation used in our calculations, the dipole moment is calculated both as the expectation value of the corresponding operator and as the field derivative of the energy. The results that each strategy yields and comparison with accurate ab initio calculations are used to assess the reliability of different methods in computing pair dipoles. Our computations have been performed with Gaussisan~ and Slater-type orbitals. The work with Slater-type orbitals required us to augment the contents of the standard tables in order to calculate one of the integrals over these functions. The same perturbation method as in Chapter 3 is used in Chapter 4. We derive an expression for the overlap-exchange contribution to the triplet dipole moment. The significance of the results of this chapter for studies of many-body effects is twofold: We prove that some methods used to calculate interaction-induced properties are not reliable. Second, we show that the results for the three-body long-range and exchange-overlap contributions to the pair dipole of quartet H3 are comparable in order of magnitude at intermediate range. This result is known to hold also for the energy. The exchange—overlap dipole (upon which Chapters 3 and 4 focus) vasishes when the separation becomes very large. The dispersion dipole moment as calculated in the region where overlap and exchange effects are negligible diverges when the separation approaches zero. In chapter 5 we calculate the overlap damped dispersion dipole moment, which remains finite. Also, we present a general treatment of susceptibility densities and tensors of potential use in the study of nonlinear phenomena. We treat the problem of secular divergencies, and we give an analytic expression for the first nonlocal hyperpolarizability density, needed to compute several damped properties, including the dispersion dipole. CHAPTER 2 2.1. - INTRODUCTION Perturbation techniques are widely used ([1,2] and refs. therein) in the study of intermolecular interactions and their effects on super- molecular system properties. The wave operator that generates the perturbed state vector from the unperturbed wave function is important within this context, and in a general treatment of self-consistent field theory [3]. Lowdin has developed a projection operator formalism [3,5] that gives the wave operator in terms of a reduced resolvent operator T. Within Lowdin’s formalism, proof of the existence of T hinges on proofs of operator invertibility. Wilson and Sharma [6,7] have analysed the invertibility requirements for one form of the T operator. By relating T to an ”outer projection" of the Hamiltonian, Lowdin has recently shown that the operator T remains regular as needed for the construction of perturbed eigenfunctions [4]; he has also applied the partitioning method to derive rational perturbation approximations [4]. An alternative derivation of Lowdin’s perturbation expansions is presented in this chapter, based on Kato’s theory. This approach does not involve manipulation of inverse operators, although it does require proper manipulation of resolvents and consideration of their domains. In section 2.2 we review the aspects of Lawdin’s work necessary for comparisons [3,5], with brief reference to other work in this field [8]. In section 2.3 the calculation of the perturbed eigenvectors and eigenfunctions is treated from the viewpoint of complex variable and Hilbert space theories [9-13]. Expansions are developed in section 2.3 for perturbed Hamiltonians with arbitrary dependence on a perturbation parameter x. The Hamiltonians considered by Lowdin (in Refs. 4, p. 79 and 5) can be understood as one- parameter dependent expansions in which only the first-order correction term is present. The results and conclusions from section 2.3 are discussed in section 2.4, while section 2.5 contains a digression on the interrelation among several perturbative treatments used in quantum mechanics. 2.2. ~ ON LOWDIN’S STUDIES IN PERTURBATION THEORY The Hamiltonian H of the system under consideration is expressed in terms of the Hamiltonian H0 for the unperturbed system and the correction (1) Hm term H as H = H0 + . The wave operator U relates the perturbed eigenfunction w to the unperturbed eigenfunction $0: w = uwo. (1) When the operator t is constructed from the reduced resolvent T and the perturbation term H(l) as t = H(1) + H(1)TH(1) = H(1)U , (2) where U = 1 + mm (3) the eigenvalue x associated with dimay be written as x = x0 + 0 , (4) where the brackets with subscript zero denote an expectation value in the state vb. It is useful to express T in terms of the projection operator P that projects out the eigenfunction $0 of Ho from an arbitrary function. P projects onto a one-dimensional subspace of 356, the separable Hilbert space taken as the set of functions that are square-integrable and complex-valued on the configuration space of the Hamiltonian operator H. The projector onto the orthogonal complement with respect to the range of P is (l-P). The reduced resolvent T in Eq. 3 is the operator T(€) = [c — (1-r)u]’1 (l-P) (5) evaluated at €=x (in general e is a complex variable). T(€) may be expressed equivarently in terms of the resolvent R i (6—H)-1 of the outer projection H = (l-P)H(l-P) of the Hamiltonian H: T = (1—p) (e—fi>’1 (1—P) . (6) In general we may assume 3(x—H)—1 only when x¢o(H), where 0(H) is the discrete or point spectrum of H, even though x may be a point of constancy of E(x) or a point of continuity of E(x), where E(x) stands for a resolution of the identity [14]. In deriving conventional perturbation theory formulae, we may take T(€) with 6 in the spectrum of H, provided that € is not in the spectrum of H; in particular, we may set €=x, because x0 ensures that T(€) remains regular at x [4]. Series for t and U [5] may be derived by using the unperturbed reduced resolvent To(€), defined as T0(€) = [€-(l-P)Ho]-1 (l-P), with appropriate choice of € to obtain Brillouin-Wigner, Rayleigh-Schrodinger or "intermediary" perturbation expansions [4]. With the reduced resolvent S = [xo-(l-P)H0]-1(l-P), if || < l, where v’ = n<1> - = H‘l) — (7) o O the following expansion is obtained for t: t 2 H(l) + H(1)SH(1)+H(1)SV’SH(1) + H(1)SV’SV’SH(1) + . . . . (8) Hence V’=H(l)-o - 0 - 0 + 1)>OSH(1)>0 + OSH(1)>0 + . . . , (9) (l) + (H S )SH(1)] + [H(1)S(H(l) - O)S(H(l) - O)SH(1) — - OSSH(1)] + . - - 3 t1 + t2 + t3 + . - - (10) u = 1 + [sn(1)] + [s(n(1) - o)sa(1)] + + [s(H(1) - 0)S(H(l) - o)SH(1) (11) _ (DU) 2(1) ...= (H SH >08 H ] + — 1 + 111 + 112 + 03 + where the ith term in square brackets has been identified with ti in expression (10) or with Ui in (11). The expansions for t and U given in (10) and (11) let us calculate corrections to $0 and x0 consistently with the choice of H. A related perturbation formalism has been developed by Speisman [8], using the operator T =—élEEl . (12) o >~o We should emphasize, though, the essentially symbolic character of this expression in his work. This character becomes evident when, after taking x0 as an isolated eigenvalue of Ho, Speisman in fact equates (12) to the reduced resolvent, i.§., I’(x - x0)-1 dE(x), where the prime in the integration means that it is performed in the whole range of x except within a properly defined neighbourhood of x0. 2.3. - THEORETICAL ANALYSIS This section begins with a brief review of the aspects of Kate’s perturbation theory that are central in this analysis. Rigorous treatments of the perturbation methods, regular or asymptotic expansions and convergence properties may be found in Kato’s original papers [9-12,15]. The derivation of the perturbation series is carried out and the connection with Lowdin’s work [4,5] is established. Following Kato, let us consider HK a self-adjoint or hypermaximal, but not necessarily bounded operator such that W Hx = I xdEK(x) (13) Q where the system of projections EK (x) is the resolution of the identity corresponding to HK and the sub-index x stands for the dependence on a parameter xefli. For £€A(HK), the resolvent RK(R) is defined by Rx(“) = (“K‘“)—1 = I ‘0 ) (An explicit derivation is given in Appendix A.) In the more general case 9) ax = no + xn(1) + x2u(~' + - - - , (24) we obtain card (I) card (J) k (v ) k k (v ) k A(") = - z (—1)p z (s 1H 1 s 2 - - - 5 PH P s p+1)i‘j , i=1 j=1 (25) where card(I) = number of elements in P v I = {(v1,-",vp)353vj=n A VJZI: JZISJSP} and card(J) = number of elements in 2+1 J = {(k1,...,kp+1)3§3kj=p A kJ20,Vj:l$j$p+l} It has been proven [10] that there exists a unitary operator UK such that a) U0 = l, b) UK E0 is regular wherever EK is regular, and c) E = U E0 U—l . K K K (n) Let us now consider explicitly the terms A , for n = l and 2 12 (1) _ _ (1) _ (1) A - EOH S SH E0 A(2) = E0H(1)SH(1)S + SH(1)EOH(1)S + SH<1)SH(1)E0 - _ , (1) (1) 2_ (1) 2 (1) __ 2 (1) (1) _ (2) _ ECHO EOH S EOH S H E0 5 H EOH E0 EOH S _ (2) SH E0 . (25) If we define (1) - _ (1) al ; SH E0 19> _ sows (2)- (1) (1) _ 2(1) (1) _ (2) a1 ; SH SH E0 S H EOH E0 SH Eo aéz) 5 E0H(1)SH(1)S - E0H(1)E0H(1)S2 - E0H(2)S (27) then from (26) and (27) (1) _ (1) (1) A — a1 + 82 A(2) 2 aiz) + aéz) + SH(1)EOH(1)S - E0H(1)SZH(1)E0 . (28) Substitution of (28) into (19) yields Eszo+ x[a§11 aé1)] + x2[a§2)+ aé2)] +...+x2[SH(1)E0H(1)S- 13 9 +...+ {5(x“)Tr-—O} (29) where the braces in (29) include all those terms whose order in x is equal or greater than 2 and whose trace vanishes, the first of which is (1) (1) _ (1) 2 (1) SH EOH S EOH S H E0. Let us denote by EK the operator obtained by removing from E all K those terms whose trace vanishes. We may take EX 2 exl + e"2 (30) where _ (1) 2 (2) exl - E0 + x81 + x 81 + (1) 2 2) (3” ,( exZ ‘32 * “ d2 Rearrangement of the terms in the series (19) is justified by the fact that the expansions (19) and (29) are absolutely convergent when Ix] is less than a bound that must be determined in each particular case [12]. Also ex2w0 = 0 and Exwo = ex (32) Using for the unitary operator the same notation for removal of terms with vanishing trace, we write . _ - - -1 _ - - -1 Ex - UxEon - (UKE0)(E0UK ) . (33) The following relationship is trivially fulfilled when 14 co co AiltZa. andB-I1+Zb.: . 1 . 1 1:1 1:1 Q as A - B 1 A + B + Z Z aib. -l (34) 1:1 3:1 J provided that A and B are absolutely convergent. Since 6 1 + KU(1) + x2fi(2) + . . (35) E B E + E B '1 + [ z 2 x(1+J)U(1)EOU(J) ] — E0 (36) i=1 3:1 as may easily be deduced by using (34). Comparing (36) and (30), we identify e = U E = E + xa(1) t x a + . . . (37) x 1 1 x1 0 0 The expansion for UK may be written from (37) and (27) as “ 1 - xSH(1) + x2[SH(1)SH(1) — 52H(1)EOH(1) — sn‘2)] + . C II 1 + [‘xS + x2(SH(1)S - SZH(1)E°) +...]H(1)+ + [x2(—S)+...]H(2)+ . . = 1 + f1(x)H(1) + {2(x)n(2) + . . . = 1 + E: f (x)H(n) , (38) n=l n where fn(x) stands for an expansion in powers of x, the first term being of (1), H(2) order n. It contains the operators H , According to (32), 15 (39) If the traceless terms in EK had not been omitted, a different U K would have been obtained, and it would yield (see (39)) a wave function that would satisfy the normalization criterion (O lwx) = 1, rather than <¢b|$x) = K 1 . When only one perturbative term is present in the Hamiltonian, U; becomes E : 1 + f1(x)H(1) (40) K and E = E E w = 6 w = w + f (x)H(1)¢ (41) x x 0 0 x 0 0 1 0 where f1(x) = "KS + x2(SH(1)S — $2H(1)Eo) + . (42) Let T a f1(x=1) = —S+SH(1)S-SZH(1)E0+... = s’+sn(1)s-szu(1)so+ ... (43) where S’ 5 -S. Now w = (1 + TH(1))EO = (1 + w(1) + w(2) + ... ) E0 (44) where w(1) = S’H(1)E 0 16 .I l f) w(2) : su‘l’sn‘1)20 ~ s“H(1)E0HmE0 . (45) Formulae (45) are exactly the same as those given by Lowdin, as may be seen in a straightforward way, once (37) has been slightly transformed by using 2 (1) (1) u 2 (1) (1) - 2 (1) (1) (46) s H EOH E0 - s H EOH EOEO — s H 0E0 ’ (1)0 for SZH(1)E0H(1) in Ex. Let us now compare our results with the expression for the energy and by substituting 32H given in ref. [5]. If x = 1, >’ ll x0 + <¢0|H(1)W|¢0> = x0 + <¢0|H(1)(1 + TH(1))I¢(0)> (1) (1) (1) x0 + <¢0|n |¢0> + <¢0|H TH |¢0> x0 + [<¢0|H(1)|¢0>] + [<¢0|n(1)s’n(1)|wo>] + [<¢0|n(1)sn(1)sn(1)|¢o> — <¢b|H(1)82H(1)E H(1)|¢0>] + . 0 x0 + (t1)0 + (t2)o + (t3)0 + . . . (47) where 0 has been identified with the ith term in square brackets. According to (24) and (38), we may write in general t = E an(n)[1 + in fm(x) H(m)] (48) n=l m=l and if, in particular, we take only n = l and x = 1 l7 (1) (1)[ (1) (l) t = H(1)[l + fl(x=1)H 1 = H 1 + TH 1 : H u (49) The last expression above is identical to the expression given by Lowdin for the reaction operator t. Kato [12] found the expansion of wx in a slightly different way. He considered the adiabatic transformation Wx that allows us to obtain on from ¢ 0 9 w = W W0 (50) K K where W is taken as U -E and x x 0 dWK dEK = - W dx dx x (51) The formulae to be substituted in an expansion of WK at x = 0 are found by calculating the qth derivative of WK with respect to x at x = 0 and by comparing with (19), a q W = Z -}— [ £1— 1:0q'Kq : E0 + KA(1)EO + x q=0 q. dxq x x + K2{ E0 + % (4(1))230} + . . . (52) where [‘dE" ] “(1) dx x=0 18 d2 E <2) (53) , have been used. The perturbed wave function is obtained from (50). - _ (1) 2 (1) (1) _ , (1) 2 (1) 4x - 40 KSH 4b + x {SH SH *0 \H >05 R 4b 1 1 2 2 - 5 IISH( )woll 4:0 -— SH( >40} + . . . (54) Derivations of the expressions equivalent to A(n) for the energy are analogous (9:. [9,12]). 2.4. - DISCUSSION Several authors ([16] and references therein) have formulated diagrammatic representations of the Rayleigh-Schrodinger perturbation theory. The use of the rules from such techniques is a convenient way to write to any order the correction to both the energy and the wave function, without requiring an explicit calculation when the expressions are needed. A(“) and its Similarly we point out that the sum rules in expressions for equivalent correction terms to x0 are not particularly complicated. Once these expressions have been written, it is a straightforward matter to (n) obtain correction terms directly either by introducing A in the expansion of WK to obtain the wave function or by calculating several traces to obtain the eigenvalue. The method has the advantage that we may keep consistency between the order of the correction in our expansions for xx and w“ and the order in x kept in the expansion for Hx, since we have formulae that are general enough for both A(n) and its equivalent form for xx; i.g., the formulae are not restricted to Hamiltonians like HK = H0 + xH(1), but Hx may be taken to the desired order in x. Another significant aspect within this context is that w; is obtained from $0 as the result of an adiabatic transformation, WK or Ux'EO' The operator that carries out this transformation acting on an arbitrary function yields an eigenfunction of H“. The unitary operator, UK or Ux, represents how the function describing the state of the system evolves as the effects of the perturbation change it. The relationship with the variational principle has been discussed by Kato in his original papers. The degenerate case is also treated in the 19 20 references given to Kato’s work. Following Kate’s original work, the usual series in perturbation theory have been deduced, with the same normalization criterion as in Lowdin’s work [3-5]. The series have been written in terms of a reduced resolvent, for which no problem about singularities arises. This work establishes the consistency between results obtained from Ldein’s method and expressions developed strictly within the framework of complex—variable perturbation methods applicable for Hilbert spaces. A different method from that used by Kato [12] has been given to derive an expansion for the wave operator. It allows us to visualize different ways to obtain the perturbed wave function and its different normalization conditions. 2.5 - RELATIONSHIPS AMONG PERTURBATION METHODS. This section should be regarded as a complement to the main contents of this chapter. We include it here because we have seldom encountered references to Kato’s work in the literature on perturbation theory; yet Kato’s rigorous treatment of perturbation theory provides a reliable and powerful algebraic language to deal with perturbative problems. The diversity of perturbative treatments in textbooks and research papers and their apparent independence and specialized character make it difficult to develop a unified conception of several areas of nuclear, atomic and molecular quantum mechanics because of an apparent lack of common principles. We therefore present in this section the underlying fundamentals, which interrelate different mathematical formalisms through a common background. It is customary to regard interactions as corrections to an unperturbed Hamiltonian, H0. The total Hamiltonian HK is decomposed into Ho and the perturbation xH(1) associated with the interaction. The limit x = 0 correspondents to the unperturbed system whereas x = 1 corresponds to the real system. In the study of atoms, for instance, H0 may include the H(1) would kinetic energy, nuclear attraction and central potential terms; include the Coulomb repulsion minus the central potential terms, and perhaps magnetic interactions or interactions with an external field. The whole Hamiltonian is thus split into two terms formulated according to the independent-particle model. Kato’s formalism, as formulated in the references given in this chapter, yields expansions commonly called Ragleigh-Schrbdinger expansions for the perturbed eigenvalues and eigenfunctions. Besides mathematical 21 IN) (‘3 rigour, Kato’s treatment has the strength of dealing with projection operators, EK or E0, which are uniquely determined, whereas eigenvectors lack this property. Some of the results of this chapter have been obtained because of this feature, g;g;, different normalization criteria imposable on the wave function are easily visualized in terms of the operator that yields the perturbed state vector when it acts upon the unperturbed wave function. The exact energy, xx, does not appear explicity in the expansions. This fact renders the expansions for the exact eigenvalues and eigenfunctions very useful for calculations. Size consistency is another interesting feature of the expansion for the exact energy given by Eq. (47). Brueckner [17] considered the Rayleigh- Schrodinger expansion and showed that the terms having a non-linear (aphysical) dependence on the number of particles of the system cancel with each other. The linked—cluster diagram theorem was later proven by time- dependent [18] and time-independent [19] methods. This theorem states that the aphysical terms cancel through all orders of diagrammatic perturbation theory, only linked diagrams appear in the series for xx. Although it is not conventional in the theory of operators on the Hilbert space, we could consider the inverse (xx - Ho)- instead of the resolvents of either HO or H“, as taken previously. We would thus generate expansions analogous to those given in previous sections of this chapter, but they would have xx instead of xb in the resolvents. The perturbation expansions with explicit dependence on the perturbed eigenvalue xx are known as the Brillouin-Wigner expansions. The analogue to Eq. (47) so constructed is the Brillouin-Wigner expansion for the energy, and that of Eq. (49) is designed the transition matrix or T-matrix [20-22] in 23 scattering theory, whereas its expanded form is known as the Born (or Newmann) series [20~22]. "Reaction matrix" or "K—matrix" replaces the term "transition matrix" in studies on nuclear matter [23]. Eq. (47) with x0 in the resolvents replaced by xx may provide a justification for the terminology "effective interaction", as t is sometimes called. The sum of the zeroth- and first-order correction energies is the expectation value of HK in the state represented by the model function ¢b in both the Brillouin-Wigner and the Rayleigh"Schrodinger perturbation expansions (see Eq. (47)). When H0 is the central-field Hamiltonian and Hm includes at least the noncentral electrostatic interaction, the sum x0 + («no I 11(1) I¢O> is called the Hartree—Fock energy because this is the quantity minimized in a Hartree—Fock procedure [24-27]. The remaining part of the energy is the correlation energy. Methods such as CI, MCSCF, electron—pair theories,..., [28] are employed to obtain quantitative information on the terms in Eq. (47) beyond the first two. The terms in Eq. (47) of second order and beyond represent true many-body effects, for the Hartree-Fock approximation only takes into account the effects on each electron of the average field of the remaining electrons and nuclei. Thus, the electrons move independently of each other and the instantaneous motion of the electrons or the correlation between them in their mutual Coulomb field is not taken into account as such [29]. Size consistency is a desirable property of any method used in computations in chemistry because it is the differences between two quantities what are often most interesting. In contrast to the Rayleigh- Schrodinger perturbation expansion however, the Brillouin-Wigner expansion for the energy is not size consistent. Nevertheless, we know [9] that the 24 eigenvalue xx of the perturbed Hamiltonian is regular in x and power—series expandable with a non-vanishing convergence radius. The eigenvalue is an analytic function of x and xx * x0 as x + 0. Therefore, it is legitimate to think of an expansion about xx = x0 of the resolvents in the Brillouin- Wigner series. Such an expansion allows us to recast the Brillouin-Wigner series in the Rayleigh-Schrodinger form. Furthermore, Brandow proved [31,32] that this procedure allows for cancellation of all the unlinked terms in the Brillouin-Wigner series, as should be expected because the Rayleigh-Schrodinger series is size consistent. Feshbach’s operator [34,35] in nuclear physics is an effective operator that yields the exact energy when operating on a model function. Lowdin’s [3,5] treatment of the partitioning technique may be considered a development of Feshbach’s previous work. Finally, the resolvent operator (14) may be written as _ —-l_ . -1_ + R.(“) - (HK-R) - — (c - Hx+1n) = — c (a) (55) with t, n e]R,and £5£+in. The zeroth-order Green’s function operator or propagator G; (C) is obtained from 6+ (t) by replacing Hx by H0. Now (a-H +1.) = (E‘Hx+in) + “(1) (56) 0 from which it is a simple matter to obtain the Dyson equation 6+(z) = 63(c) + GS<£>H(1)G+(:) (57) closely related to the wave-operator relationship (40). The limit of £+ xx and n + 0 yields a distribution which establishes the connection between the Green's function-opeator notation of the resolvent and the singular (or Sochozki’s) generalized functions [36] of use in scattering theory. CHAPTER 3. 3.1. INTRODUCTION Interactions between colliding molecules in gases or liquids cause shifts in the charge distributions of the collision partners. These shifts result in differences between the net dipoles of colliding pairs (or clusters) and the vector sums of the dipoles of the molecular constituents, if unperturbed. Collision-induced changes in dipole moments are manifested in the dielectric and spectroscopic properties of bulk samples. For example, infrared and far infrared absorption processes that are single- molecule forbidden may be observed in compressed gases and liquids as a consequence of transient, collision-induced dipoles. Such pressure-induced far IR absorption has been observed experimentally in inert-gas mixtures [37-39], H2 [38], N2 [38,40-42], 02 [38,40,41], CH4 [43], and SF6 [44] and forbidden near IR spectra have been studied for the diatomics, triatomics such as CS2 [45], and other polyatomics. Useful information on dynamics in dense gases and liquids can be obtained by analyzing the lineshapes for single-molecule forbidden spectra (and for collision-induced contributions to allowed spectra), if the collision-induced dipoles are known as functions of intermolecular separation and relative orientation. In analyses to date, collision—induced dipoles for small molecule pairs have often been approximated as sums of classical multipolar contributions, with short-range anisotropic overlap corrections represented by parametrized exponential functions. At this stage, direct calculations of overlap effects on pair dipoles are needed in spectroscopic applications. It is also of interest to compare the calculated dipoles of van der Waals complexes in their equilibrium configurations with the dipoles determined experimentally by 25 26 molecular beam electric resonance studies of the Stark effect on rotational transition frequencies characteristic of the complex [46,47]. Calculations or measurements of collision—induced dipoles provide information on molecular interactions complementary to that obtained from potential energy surfaces, and may indicate the relative importance of classical electrostatic interactions, charge transfer, and short-range overlap and exchange effects [48]. For molecules interacting at long range, only classical-multipole polarization (cf. [48]) and dispersion effects [49] contribute to the collision—induced change in dipole moment. The net dipole for well separated molecules can therefore be determined if values of the single- molecule multipole moments, polarizabilities, and nonlinear response tensors are known [50—55]. For molecules interacting at short range, definitive results can only be obtained by ab initio calculation [56-62]. Calculations including correlation effects have been performed for the pair dipoles of He...H [57,59,60], He...Ar, He...H2, and H2...H2 [57], while calculations restricted to self-consistent field level are available for the dipoles of the inert~gas heterodiatoms Ne...Ar, Ne...Kr, and Ar...Kr [56,58], and for the Ne...HF dipole [48]. Quite large basis sets are usually needed in pair property calculations [61], with the consequence that the computational requirements are substantial. This prompts interest in approximations that yield good results in the region where overlap is small, but nonnegligible. Since numerical precision is most difficult to attain in ab initio calculations on molecules at intermediate and long range, approximations applicable near the van der Waals minimum may be used to join the known long-range forms of the interaction—induced dipole to ab initio results at 27 short range. Additionally, quantum mechanical approximations for collision~ induced properties provide information on the effects of long-range electrostatic interactions, overlap, exchange, hyperpolarization, and dispersion, which may prove useful in selecting basis sets for subsequent ab initio work. To evaluate proposed approximations for pair properties, it is necessary to compare the results with the accurate ab initio results available for properties of small molecular pairs. Comparisons of ab initio and approximate collision-induced polarizabilities of H...H in the triplet state [62] and He...He [61] have been used to test electrostatic overlap models [63], exchange perturbation methods [62,64—66], polarizability density models [67~69], and exchange antisymmetrization approximations [70,71]. In this chapter, we report a calculation of the collision-induced dipole moment of He...H as a test of label—free exchange perturbation theory [72—75] at lowest order. The label—free exchange perturbation method constitutes a direct Rayleigh-Schrddinger perturbation theory with fully antisymmetrized zeroth-order wavefunctions. By construction in terms of projection operators, the unperturbed Hamiltonian and the perturbation term are individually invariant with respect to exchange of electrons between the interacting molecules. The label-free exchange perturbation formalism is reviewed briefly in section 3.2, for application in Computing collision- induced dipoles. In section 3.3 the label-free theory is related to other exchange perturbation approximations, and the collision-induced dipole is shown to separate into polarization and exchange contributions. It is also shown that results for the pair dipole at zeroth-order in the label-free 28 theory are identical to the results of an exchange-antisymmetrization approximation developed by Lacey and Byers Brown [70]. We have calculated the He...H dipole moment in two ways: first, by evaluating the expectation value of the dipole operator with the zeroth- order pair wavefunction; and second, by computing the energy for He...H in the presence of a uniform applied electric field and then differentiating with respect to the field to obtain the dipole. Also, we have carried out the calculations at two levels of approximation for the single-atom wavefunctions. In the first, ls Slater orbitals are used on each center; and in the second, an extended Gaussian basis is used at each center. Evaluation of the dipole expectation value in the Slater basis has been reported previously by Buckingham [49] and by Mahanty and Majumdar [76]; we obtain identical results in this case and new results from the other three calculations. The methods of calculation and the selection of basis sets are described in section 3.4. Results are presented in section 3.5. We find that closest agreement with accurate ab initio results [57] is obtained from the Gaussian-basis calculations of the dipole as an energy derivative. Errors in the overlap dipole are typically 20-30% at this level of approximation. Significantly, the errors in this approximation appear to be smaller than the discrepancies between the two reported ab initio calculations of the He...H dipole [57,60]. 3.2 EXCHANGE PERTURBATION THEORY IN LABEL-FREE FORM For molecules interacting at short range, both polarization and exchange effects contribute significantly to the interaction energy and to interaction-induced changes in electric properties. The label—free exchange perturbation theory developed by Jansen [72-75] treats these effects within a direct Rayleigh-Schrodinger formalism. Projection operators are used to define an unperturbed Hamiltonian and a perturbation term that are separately invariant with respect to electron permutation. Also, within this formalism the expectation value of any dynamical variable for a cluster of interacting molecules can be separated into additive contributions from each of the molecules in the cluster. The Hamiltonian H for interacting molecules A and B with a total of N electrons is 2 . 7 Zb 1 V . ( -z-—§-—Z-——)+2—~- (1) 2 a raj b '"bj k (9) c = 2 ; <$(°)|$(1)> (10) 2a i=1 in ' 5(0) is an antisymmetrized product of unperturbed orbitals centered on A and B, and the tilde superscript indicates that the function has not been 32 ~(1) normalized. “in is the antisymmetrized product with the unperturbed orbital i replaced by the first order correction to orbital i in an applied field Fu in the u direction. The vector égl) has components "(1) "(1) ~(1) (pix ' ’iy ’ and ’12 ‘ To lowest order in the interaction, the total energy E of the A—B pair is - - <0) <0) -1 ~(0) P ~(0) E-<@0|H|¢o>—EA +13 +c < [xvi/tile >+ B 1 i=1 —1 ~(0) ,. N ~(1) —2 ~(0> “ ~(0) - +E-{2c1 < IH(£=0)IJEIQ>‘j >-<,;2c1 < IH(£=0)|§ >-g } (11) where §= awhile, jLI350)»;1 (12) 1: Thus, to lowest order, the energy is the sum of three terms: first, the (0) (0) A I EB the absence of the applied electric field; second, a field-independent term total energy E for molecules A and B at infinite separation and in equal to the A-B interaction energy to lowest order if exact wavefunctions wk and “h are used to construct 0b [77,78]; and third, a term linear in the applied field. From the energy expression, the dipole moment may be obtained by differentiating with respect to the applied field: 2% B aF - 19 «p Inn; > - -[<fl |H|~p > + ‘ a§ o 0 i=0 ’ aE 0 N {=0 33 + .811 gal: (QOIaEI@O> + <¢0|H| 8E )]E = 0 A N Q - 2c‘1<¢(°)|n(520)| z (1)) + c_2c 1 ~ 1—l §i 1 <5(°)|§(§=0)|$(°)>. (13) 2 In Eq. (13) g is the dipole moment calculated directly as the expectation value of the dipole operator with the zeroth~order, zero-field wavefunction, as in Eq. 12. The remaining terms in Eq. (13) represent the non-Hellmann- Feynman contribution, which would vanish if Q0 were exact. In the label-free perturbation formalism, the operators R0 and 7 are not Hermitian individually, although the sum of the operators is Hermitian. As a result, the pair energy at first order in the interaction formally includes the correction term <¢0|H0|01>, which depends on the first-order change Q1 in the pair wavefunction due to the perturbation V. Since this term is expected to be substantially smaller than the standard first-order energy shift (Q0|V|@b> included in Eq. 11, the correction <¢b|H0|@1> is usually grouped with the leading second—order terms [72,73]. To second order, the interaction energy may be computed from Q1 in a sum—over-states form [72], with contributions from the continuum generally nonnegligible [79]. Alternatively Q1 may be approximated by finding the stationary vector of the functional (cf. [80]) J[q,] = + + (\IIlV-EIIKIJ0> (14) subject to = 0. If SJ vanishes with an arbitrary variation 6% of Q away from the stationary vector 31, then the expansion coefficients ak for $1 satisfy 34 l <¢k|vnp0> + 2 31:30 kaa[<4’k'"o'%> + <¢Q|H0|¢k>1 8k: Ei— E ('5) 0 k in the basis @k of orthogonalized, antisymmetrized product states for A and B at infinite separation. In the approximation that the sum on the right- hand side of Eq. 16 is negligible, $1 = @1. An equivalent approximation has been made by Jansen in a direct expansion of 01 in the basis ¢k [72]. If $0 and @1 have been determined, then the sum of the unperturbed energy E0, the full first-order energy E1 (including the correction term <¢0|H0|¢1>), and an approximate second~order energy E for the A-B pair in 2 an external field can be obtained from E0 4 E1 + E2 : + . (16) The interaction~induced dipole may then be computed from the derivative of Eq. 16 with respect to E, as in Eq. 13. In this work, the calculation of exchange-overlap contributions to the He...H pair dipole is based on Eqs. 11 and 12, which represent the lowest- order effects. The exchange~overlap terms fall off exponentially with increasing R. In contrast, the second-order dispersion contribution to the He...H dipole (present in B derived from Eq. 16) varies as R".7 at long range. This contribution originates in the correlations between the fluctuating charge moments on the H and He atoms, and it dominates the exchange-overlap contribution for sufficiently large R. Accordingly, in calculating the He...H dipole for internuclear distances R > 4.0 a.u., we have added the leading dispersion dipole to the lowest~order exchange- 35 overlap dipole from Eq. 11 or 12. Higher-order polarization and exchange effects are neglected at this level of approximation for the pair dipole; overlap damping of the dispersion contribution and the effects of He— electron correlation on the He-H correlation and exchange energies are also neglected. 3.3. COMPARISON WITH OTHER EXCHANGE PERTURBATION METHODS In this section the label-free exchange perturbation theory of interaction-induced properties is related to other exchange perturbation approaches. We show that the results from the label-free theory can be separated into polarization and exchange contributions [2,81-83]. We also show that the zeroth—order expression for the pair dipole in the label-free theory is identical to the exchange-antisymmetrization approximation used by Lacey and Byers Brown in calculations of inert—gas heterodiatom dipoles [70]. 3.3.1 Exchange perturbation theory: Polarization and exchange contributions to interaction energies and pair properties In one standard form of exchange-perturbation theory [81] the nth approximation to the wavefunction On and the nth approximation to the interaction energy En of an AeB pair are obtained iteratively from @n ¢o l R0 (En " V1) @n—l (17) and E n ll <¢0|v1|¢n_1> , (18) where ¢ is the simple product of the ground—state wavefunctions ¢ and w , 0 A V1 is the perturbation term for the assignment of the first NA electrons to molecule A and the remainder to B, and R0 is the reduced resolvent |¢k><¢kl kxo Ek ’ Eo R = Z (19) 0 ‘Phe function ¢k is the kth simple—product excited state of A and B at infinite separation and Ek is the corresponding energy. The intermediate 36 37 normalization <¢OIQ> = 1 has been imposed. Symmetry forcing is accomplished by choosing the normalized antisymmetrized product of ¢A and wB as the zeroth-order approximation @0 [2,65,66]. At first order the A-B interaction pol energy is the sum of a polarization term E1 and an exchange term EeXCh: l <¢0|v1|r¢0> — <¢0|v1|¢0><¢olrl¢o> 1 + <¢0|P¢0> _ pol exch _ E1 ~ E1 + E1 ~ <¢0|V1|¢0> + <¢0|V1|¢0> + <¢0IV1|P¢0> = , . (20) 1 + <¢0|P¢0> where P is the intersystem antisymmetrizer (see Appendix B). In the label-free exchange perturbation theory, the interaction energy at first order is p E1 : «11021-21 ViAil‘I’o'i . (21) This expression for E1 may be simplified by splitting the full antisymmetrizer in @0 into terms that involve electron permutations within A, within B, and between A and B. In Appendix B, the permutation invariance of V is used to obtain E-l[<¢w|vlww\+/¢¢|§:v1\|flfli: PAB{x }>l(2°) 1 ' s A B 1 A 3’ ‘ A 3 i=2 i i 'A BJ;1 OJ 5 1"‘XN ‘ The expressions for wA and $3 have been restricted to the self~consistent field level, and the set {x1...xN} contains the orbitals occupied in the ground states of isolated A and B molecules. The operators P23 acting on {x1...xN} perform intersystem permutations only. In Eq. 22,.AA is the antisymmetrizer for electrons assigned to molecule A,.AB is the 38 antisymmetrizer for B, 01 is the parity of the ith permutation among the total of 2, and s ; l + <¢0|P¢0> . (23) Thus the interaction energy at first order is given as the sum of the polarization contribution <¢0|V1|¢0> and an exchange correction. Each of the operators/l.i in the second term of Eq. 22 projects out the single term from the ket for which the assignment of electrons to A and B is consistent with the form Vi for the perturbation. In the product ¢A¢B the first NA electrons are assigned to A and the remainder to B. Relabelling of dummy indices suffices to show that Eqs. 20 and 22 are identical. At first order, equivalent results are also obtained from variants of exchange—perturbation theory with stronger symmetry forcing [2,84-87]. The one-particle density matrix [88~90] can also be separated into two terms, one without electron exchange between A and B and a second with one or more interchange contributions. Consequently the expectation value of any single—particle operator decomposes into polarization contributions and exchange-overlap contributions [2]. McWeeny and Sutcliffe [89] and Magnasco, Musso, and McWeeny [91] have performed the separation explicitly for the one-electron part of the Hamiltonian for two interacting molecules. The analysis for the dipole moment operator is analogous. 3.3.2 Density-matrix perturbation theory Lacey and Byers Brown have evaluated the dipole moments of several diatoms by first finding the density matrix associated with the normalized, —l/2 5(0) antisymmetrized product c1 of the isolated atom wavefunctions [70]. The spinless electron density p(0)(r) in the absence of an applied field is 39 obtained from 0 -1 ~ P( )(El) = Nc1 I ds1 I dxz...dx ¢(O N )(51’52"°§N)$(0)*(§1’§2’“§N) (24) (0) where o and c1 are defined as in section 3.2, and E, s, and x denote spatial variables, spin variables, and collective space-spin variables, respectively. The exchange—overlap dipole of the atom pair AB is then r + Z r (25) B 2’ - IE p(0)(:) d3£ 't Z A B~B. A At lowest order in the label-free perturbation theory, the expectation value of the dipole is identical to the Lacey-Byers Brown result. This equivalence can be shown explicitly starting from P g=<~plzgiAinJ> (26) 121 with the pair wavefunction Q from Eq. 8. The operator Bi is the same for each of the electron assignments i and A. =1. (27) LEM” i Thus t H /\ '6' (C H '6‘ V = - I 5 9(a) d E + Z 5 + Z 5 (28) In terms of the applied field E p(1;) = 9(0)“) + g-e‘llr; + 003) (29) 40 (0) The electron density p (r) for the AB pair in the absence of an applied field (determined from Q) is identical to p(0)(r) from Eq. 24, so the field- independent dipole is equal to B from Eq. 25. It should be noted that this approach yields only the term i in Eq. 13. At this level of approximation the Hellmann-Feynman theorem is not satisfied, and the pair dipole moment obtained by differentiating the energy differs from E. The lowest—order correction to the pair electron density in an applied field satisfies N (1) - -1 ~(0) ~ 8 (r1) — NcLE1 Ids1 Idxz...de [ é (51...xN) 9i (x1...xN) * + §§1)(x ...x 5(0) (x ...xN) ] _ —2 ~(0) ~ Ngzc1 Ids1 dez...de Q (x1...§ ) Q (§1°°'§ ) . (30) N Substitution of this result for 3(1)(r) into Eqs. 28 and 29 gives an approximation for the exchange—overlap contribution to the pair polarizability. 3.4. COMPUTATIONAL METHODS. In this section the methods used to approximate the He...H wavefunction and to calculate the pair dipole are described. In each of the calculations we have employed a wavefunction of the form given in Eq. 8 for the He-H pair in an applied field in the z direction 3 . ~(0) _ 1 , <0) <0 —<0 w1th Q —Jr§- E opxtxfl <1) Xfle 22) xhe 23)} (32) prior to normalization. P runs over all the elements of the symmetric group 83 and oP = :1 depending on the parity of P. The functions 51 are defined by $§1> = x§1><1)w‘°)<2.3) — x§1)<2) w(°)(1.3) + xé1)(3) ¢(°)(1.2) <33.a) (1‘ 5%]i: x§°)<1) ¢[1)<2.3) — x;°)<2) w§1)(1.3) + x§°)<3) w, ’(1.2) (33 b) ~<1)_ <0) <1) _ (0) <1) <0) (1) i3 - xH (1) wz (2,3) xfl (2) $2 (1.3) + x” (3) t2 (1.2) (33.c) <0) . -__1_ <0 — (o _ — (o (o a» (1..) -- J2- { xfle 21) xHe 22) xfle (1) 4,6 32) } <34.a) (1)- _1 <1) ~(0), _ —(0) (1) w, - J2 a,“ (1.) xfle (2) 4,8 <1) xHe (2)} <34.b) <1)- _1 <0) —<1) _ —<1)' (0) $2 — fl {xfle (1)xHe (2) xfle (1)> (35) To lowest order in the atomic interaction and to first order in the applied field F , 2 H ~(0) ~(0) 3 ~(0) ~(1) (“2) r N2 (F‘z ) { + F2 2 (é I“ %| 23:1 l 0 +Fz: <<1>(.)Ip'z‘<'1'>>l()1 (36) 1:1 and 3 < He > - N2 (Fz ) {<¢(0)| “Ola (0)> + F2 2 <¢(°)|$ %|$(1)> } 1:1 3 ~(1) “He ~(0) + F, 1: < <1. I.) |<1> >1 (37) z ._ j z le 2 _ —l -l . . where N (F2) ~ c1 (1 * cmqzc1 ) and C02 is the z component of go given by Eq. 10 with N=3. If the origin of the coordinate system coincides with the H nucleus and the He nucleus is located at R on the +2 axis, “H “z 2 ‘2 ziAi (38) 1 and 11:8 = — [ (22+ 23- 2R)/\1+ (21 + 23- 2R)/\2+ (21+ 22— 2R)/\3 ]. (39) The result for (pz> is then 43 00 s HHe oo oo 2 _ oo 01 10 (“2)“1_Soo 2 [22 Zane Msnne1+F 1_Soo 2 [ RSHHe(SHHe+SHHe)+ HHe HeH oo 01 +810 )_Z;g +zHeH( HeH HeH 10 00 10 00 2 +2 zHHeSHeH zHeHe(2-SHHe ) 800 10 00 HHe 00 00 01 10 00 +ZHeHSHHe+ 00 2 [2% flSHHe](SHeH+SHeH )SHeH } (40) l- -S HHe where t. (t t t S a;: (x 5 ix £5; and z ab: (x ; izlx (Si ; t,s = 0 or 1; a,b = H or He . The field-independent term yields the approximate zero-field pair dipole (and the term linear in F2 approximates the exchange—overlap contribution to the 22 component of the pair polarizability tensor. It has been suggested [92] that calculating the energy of the He-H pair in an applied field F2 and then differentiating with respect to F2 in order to find the dipole (Eq. 13) should yield more accurate results than a direct calculation as an expectation value (from the field—independent term of Eq. 28). In general, though, no a pgigri choice between the two methods is possible, and we have performed calculations to test each of the methods. The explicit expression obtained from Eq. 11 for the He-H pair energy (to first order in the atomic interaction and first order in the applied field F2) is given in Appendix C. We have obtained wave functions for the isolated atoms in an external field at two levels of approximation. The A.0.’s x are constructed with 44 Slater bases in the first approximation and with Gaussian bases in the (0) second. In the Slater basis, X“ is the H~atom ls orbital, and the first (1) order correction to the hydrogenic wave function XH is found analytically[93]. xél) can be expressed in terms of modified 2pz and 3pz Slater orbitals: iél) =-;% ’ (1+§ )r cos a (41) In the first approximation for He [76], we use the optimal ls Slater- (0 ) (0 ) orbital for‘xHe and the unperturbed wave function ¢H He satisfies W(O) wHe "7? : expi-{(r1+r2) ][u(1)B(2)-B(1)u(2)] (42) with § 2 27/16. The first order correction to the He wave function is approximated variationally by minimizing _ 3 3 l (l) . (l) _ (l) ¢(1) l _ l _ l 27 2 (l)2 I—Jd rld r2[4 Ylee lehe + 4 Y2¢He V2¢H + (2r12 r1 r2+ 2(16) )whe l + <1) _ _,H<1) (0) using a trial wave function of the form wHe wHe . (1) _ with H — (21+22)Fz' (The subscripts l and 2 in the functional I refer to the electrons in the He atom.) The value obtained for the variational parameter is t = —0.72231, and 45 32.2: = ...;3 In the calculations at this level of approximation, the matrix elements appearing in Eqs. C.l-C.ll for the pair energy and in Eq. 40 for the pair dipole were evaluated by use of standard formulas for integrals involving Slater orbitals [94-102] (see Appendix D for a hybrid integral involving the modified Slater orbital 3pz). In the second approximation SCF calculations were performed for He and H using uncontracted bases of ten 3 Gaussian functions to determine x(o). The Gaussian orbital exponents [103] and the calculated coefficients for each of the functions are given in the Table. The calculated ground state energies are -2.8616692 a.u. for the He atom and -0.49999862 a.u. for the H atom. As outlined below, xél) was determined from a series of SCF calculations on the H atom, with a bare positive electric charge placed on the z axis at a distance ranging from 45 a.u. to 60 a.u. from the H nucleus [104]; xéi) was determined in the same way. An uncontracted Gaussian basis set of six p functions was used for each atom, together with a single contracted function x(o) formed from the ten 8 functions with contraction coefficients from the Table. The orbital exponents of the p Gaussian functions for hydrogen [105] and helium [62] are also listed in this Table. The p function coefficients obtained directly from the SCF calculation reflect linear and nonlinear polarization of the atoms by the field E and (1) field gradients E’, E"... of the point charge. To fix x , it is necessary to determine the contribution to each coefficient that is first order in the electric field Fz. This contribution was obtained by fitting each contraction coefficient ci to the expansion 46 C . = C(1)F + C€2) F3 + c (3) F F’ (45) 1 1 z 1 z i z 22' (l) The resulting Ci values are listed as the p contraction coefficients in the Table . As a check on the quality of the Gaussian bases, we have calculated the atomic polarizabilities of H and He. For the H atom the Gaussian basis yields a = 4.4989 a.u., while the exact value is a = 9/2. For He, in the Gaussian basis a = 1.3087 a.u. More accurate values for the He polarizability from near Hartree—Fock calculations differ by less than 2% from this result [61,106,107]. The integrals needed in the Gaussian basis calculations were performed using the program SOINTS, written by R. Pitzer (Ohio State University) and maintained by the Argonne National Laboratory Theoretical Chemistry Group. Function 10 ll 12 13 14 15 16 Type 47 TABLE 1 Orbital exponents and contraction coefficients of the Gaussian basis (lOs 6p) on each atom Exponent H Coefficient (Normalized) 1170.498 7.37776139.10' 173.5822 5.83581828-10‘ 38.65163 3.18288336-10" 10.60720 1.38031549-10” 3.379649 4.89937724-10‘ 1.202518 1.42487999-10’ 0.463925 3.12524622-10' 0.190537 4.13000752-10‘ 0.0812406 2.02671433-10’ 0.0285649 7.74758325-10" 3.009711 —1.5993250-10‘ 0.710128 —1.4338116-10‘ 0.227763 —8.7766919-10’ 0.0812406 -1.4154398 0.0356520 —1.3241078-10" 0.0154420 —3.0605474-1o‘ 5 4 3 2 2 l l l l 3 2 l 1 l 2 2 2 1 l l l 2 3 He Exponent Coefficient (Normalized) 3293.694 9.59977811-10 488.8941 7.61241309-10 108.772 4.11477457-10 30.1799 1.72174980-10’ 9.789653 5.70398703-10’ 3.522610 1.49210258-10” 1.35436 2.82269836-10“ 0.5561 3.59695130-10’ 0.2409 2.51503899-10’ 0.10795 5.17599605-10' 6.6 -2.4617864-10‘ 2.1957216 —1.9220719-10’ 0.5693178 —1.4916053-10' 0.4223072 -2.4362999-10" 0.2007022 -3.4883846-10' 0.0799030 —1.0301889-10’ 5 4 3 2 l 3 l 1 3.5. - RESULTS AND DISCUSSION The He...H dipole obtained by differentiating the interaction energy with respect to an applied field (Eq. 13) shows a roughly exponential dependence on internuclear separation in the range 4 a.u. S R $ 8 a.u. This range extends beyond the van der Waals minimum near 7 a.u. [60]. In the Gaussian basis, the dipole from Eq. 13 reaches a local maximum at very short range (R ~ 0.6 a.u.), and vanishes as R approaches zero, as expected. At short range, where the dipole is determined primarily by overlap and exchange effects, the polarity is He+H—, consistent with the rule that the larger atom is negative [49]. At long range, the pair dipole results entirely from dispersion effects, and its sign changes to He-H+. Asymptotically [49,53,54] p(R) = D7R—7 + DQR—9 + . . . (46) As noted in section 3.3, we have approximated the pair dipole by adding the leading dispersion term (with D7 = 120 a.u. [53,54]) to the exchange-overlap dipole calculated from Eq. 13 or from Eq. 12 (i.e., by direct calculation of ; alone). In the Gaussian-basis calculations using Eq. 13, the dipole changes sign at R ~ 8.7 a.u. Ab initio calculations of the He...H dipole including correlation effects have been reported by Bender and Davidson for the single internuclear distance R = 3.0 a.u. [59], by Ulrich, Ford and Browne for the R range from 0.5 to 20.0 a.u. [60]; and by Meyer for R between 5.0 and 11.0 a.u. [57]. At intermediate and long range, there are substantial differences between the available ab initio results. The dipole calculated by Ulrich gt gl. changes sign at R ~ 7 a.u., while the dipole obtained by Meyer does not change sign until R reaches ~ 8.5 a.u. Between 9.5 and 11.0 48 49 a.u., Meyer’s results approach the known long-range form of the dipole closely. In contrast, the dipole calculated by Ulrich gt gt. exceeds the leading dispersion dipole by a factor of ~ 6 in this R range. For He-H distances between 6.0 a.u. and 11.0 a.u., Meyer’s results are used as the basis for assessing our approximations, and the results of Ulrich gt gt. are used at shorter range. Our results for the He...H dipole are plotted for comparison with the gb initio dipole in Figs. 1 and 2. The long-range form is also shown in Fig. 1. Our approximation exhibits the same qualitative features as the gb initio dipole. Fig. 1 shows that the results obtained by differentiating the interaction energy with respect to the applied field to find the overlap dipole (Eq. 13) and then adding the asymptotic dispersion correction agree quite closely with Meyer’s results in the range from 4 a.u. to 11 a.u. In fact, the discrepancies between Meyer’s results and this approximation are smaller than the discrepancies between the two sets of gt tgitig results [57,60]. Relative to Meyer’s values, our closest approximation is typically in error by 20~30% over the range from 4 a.u. to 11 a.u. The remaining differences result from neglect of the higher-order exchange, dispersion, and orbital distortion effects, overlap damping of dispersion (gt. [108— 111]), and the effects of intra—atomic correlation in He on the He-H exchange and dispersion dipole. Some cancellation of error occurs, since the omitted effects are not all of the same sign. The error is g 5% at R 4.0 a.u., ~20% at R = 7.0 a.u. (the van der Waals minimum), and g 5% at R 11.0 a.u. It is large (roughly a factor of 2) at R ; 8.0 a.u., but near this point large relative error can result from a slight displacement of the zero of p(R) between the accurate and approximate analyses. 50 Computing the expectation value of the dipole with the zeroth-order antisymmetrized wavefunction (Eq. 12) gives a smaller exchange—overlap dipole than that found by differentiating the energy with respect to an applied field (Eq. 13). The latter approach yields better results. It is interesting that the gb initio value of the pair dipole lies between the Gaussianabasis values obtained from Eq. 12 and those obtained from Eq. 13 for exchange~over1ap effects, (with the dispersion correction added in each case), when R<9.5a.u. As R decreases from 5.0 to 1.0 a.u., the dipole moment increases rapidly. Low-order perturbation theory breaks down at short range, and the exchange—overlap contribution to the dipole is overestimated by Eq. 13. The level of agreement between the results of Eq. 13, evaluated either with the Slater basis or with the Gaussian basis, and the results of Ulrich gt gl. for R < 4 a.u. is surprisingly high, as Fig. 2 shows. This agreement must be fortuitous. The local maximum in p(R) occurs at larger R (near 1.0 a.u.) in the g9 tgtttg calculations than in the approximate work, and the value of p(R) from Eq. 13 is smaller at the maximum. In the Slater-basis calculations with Eq. 13, we did not find a local maximum in p(R) over the range of R values studied. At short range, the dipole computed from Eq. 12 is substantially smaller than the gg initio dipole. On the scale of Fig. 2, results from Eq. 12 are essentially unaffected by the choice of the Slater or Gaussian basis. The dipole obtained by differentiating the energy with respect to an applied field (Eq. 13) differs from the expectation value of the dipole moment (Eq. 12) by the correction term Apz [112] 51 at! —-| Au 2 “ <¢ I z E=0 0 322 o E=0 (39) ~ This term is nonvanishing because the Hellmann-Feynman theorem is not satisfied at lowest order in label-free exchange perturbation theory. The ratio of the correction term Apz to the total dipole calculated from Eq. 13 is plotted in Fig. 3 for both the Gaussian basis and for the Slater basis. Comparison of the dipoles calculated with Eqs. 12 and 13 provides one indicator of the uncertainty in the results. Calculations of Apz have been used previously to indicate uncertainties in the correlation contributions to molecular dipoles, in cases where the wavefunction at self-consistent field level approaches the HartreeeFock limit. In this work, Apz is definitely smaller for the Slater basis (see Fig. 3), but the results from the Gaussian basis and Eq. 13 agree more closely with Meyer’s work. At lowest order in the labe1~free exchange perturbation theory, for the He...H dipole we have found better agreement with accurate gb initio results when the dipole is determined by differentiating the pair energy with respect to an applied field than when the dipole is computed directly as an expectation value. The dipole calculated as an energy derivative is significantly larger than the dipole calculated from the expectation value (by factors as large as 4 for Gaussian—basis calculations in the range 4.0 a.u. < R < 7.0 a.u.). In this context, it is interesting to compare recent self-consistent field (SCF) results [56] for inert-gas heterodiatom dipoles with estimates based on the exchange-antisymmetrization approximation developed by Lacey and Byers Brown, since the exchange-antisymmetrization approximation is equivalent to a direct calculation of the dipole expectation value with the zeroth-order wavefunction. The SCF results 52 exceed the exchange—antisymmetrization estimates by ~ 60% for Ne...Ar for internuclear distances between 4.0 and 6.0 a.u., and by ~ 50% for Ne...Hr (with smaller differences for other pairs). Although the discrepancies are significantly smaller than for He...H, they are of the same sign. Improved agreement with SCF results might be obtained from exchange—perturbation calculations that employ energy derivatives. Results for the He...H dipole suggest that the label-free exchange perturbation method can provide a useful approximation for collision-induced properties in the region of charge overlap. Fig. 1. 53 Collision-induced dipole of He...H as a function of the internuclear separation R in the range from 4.0 to 11.0 a.u. Curve A shows the ab initio results obtained by Meyer, and Curve B the ab initio results of Ulrich gt gt. Curves C and D have been obtained by adding the long-range dispersion dipole to the exchange-overlap dipole computed form the energy derivative (Eq. 13). Curve C shows the dipole in the Gaussian basis, and D the dipole in the Slater basis. Curves E and F have been obtained similarly, but the exchange-overlap dipole has been computed as an expectation value from Eq. 12; Curve B shows results from the Gaussian basis, F from the Slater basis. Curvg G is a plot on the leading term of the dispersion dipole, D R . The ordinate is scaled logarithmically and the dipole p(R3 is given in a.u. Fig. 2. ab initio results of Ulrich gt gt. 54 3.0 1 . r 1 '1 \ ’ x. 1 B / ‘- \. \ 2.0 7 ./ \i. " - .1 '5 i C t : \ ' 9"}. 1 "3}. g A .5" \'\._ IO _. j 0" \. \.\ .. i ; \Kgbx i 95.x. \‘g--.,\., I 6 \§&::“\ 0 I 1 ‘~~t-~“¢‘:fi O l 2 3 Collision-induced dipole of He...H as a function of internuclear separation R in the range from 0 to 4.0 a.u. Curve A shows the Curve B and C show the lowest-order exchange-overlap dipole, Obtained from Eq. 13, B in Slater basis and C in Gaussian basis. Curve D shows the exchange-overlap dipole computed as an expectation value with the zeroth-order wavefunction; results from the Gaussian and Slater bases superimpose on this scale. The dipole p(R) is given in a.u. 55 A P2 4’2 0.9 0.8 r 0.7 - \ 0.5 - \ 0.4 - \ 0.1— 0 24 6 8|O|2|4|6R Fig. 3. Ratio of the correction term by to the exchange-overlap dipole for He...H, plotted as a functign of the internuclear separation R (in a.u.). The exchange-overlap dipole has been computed as the derivative of the first-order interaction energy with respect to an applied field. Curve A shows the results from the Slater-basis calculation, and Curve B the results from the Gaussian basis. CHAPTER 4. 4.1. INTRODUCTION Absorption in the IR region of the spectrum by non-polar fluids is a collision-induced phenomenon [115], and a number of molecular properties may be determined from the spectra [116]. Collision-induced IR spectra have been studied for atoms and non-polar molecules [37-38, 40-41, 117— 119]. Furthermore, cooperative many-particle effects in polar fluids affect the absorption line shape in this region of the spectrum. Experiments have been performed [120] in order to ascertain the importance of the collision-induced absorption relative to the absorption associated with the permanent dipole. Both the intermolecular potential and the interaction-induced dipole are needed [121] as functions of the intermolecular separation and relative orientation in the calculation of the absorption line shape. Pairwise additivity of the potential is usually assumed [122] in the simulation of collision dynamics, but higher order or nonadditive effects [81,123-124] play an important rgle in many physical situations. For example, the inclusion of three-body forces is known [125-126] to bring results of molecular dynamics simulations into closer agreement with experimental data. Molecular dynamics simulations with many-body potential functions have been suggested [127]. Some calculations [73,128] have been reported with long-range triple dipole potentials [129-130]. When the intermolecular interaction is expanded in a Taylor series, the first nonadditive part of the third-order perturbation energy of three nonoverlapping molecules is the triple-dipole potential. Nonpolar species can have a triple-dipole potential; if we consider three indentical rare 56 57 gas atoms a, b and c, an instantaneous dipole on atom a polarizes b, whose dipolar field perturbs c, and the dipole on c interacts with a. The long- range triple-dipole energy may be written as a sum-over-states expression [129] or as an integral over imaginary frequencies [130]. Bounds to the three-body long-range interaction coefficients have been obtained [131] for a number of atoms. It was established [132c] in the Drude model approximation that the triple-dipole energy term is a good approximation to the long-range nonadditive energy for the rare gas crystals, for the many-body terms that involve more than three bodies in the dipole interaction approximately cancel the remaining three-body multiple interaction terms. Nevertheless, the addition of higher order contributions than the triple-dipole term was later claimed [133] to be necessary for Xe. A combination of two-body Lennard-Jones and three-body triple-dipole and experimental exchange-overlap potentials has been used [134] in the study of atomic cluster growth. The relevance of nonadditive interactions has prompted the generalization [135] of the correction to the basis-set-superposition error for the calculation of many-body effects [136-138]. Also, nonadditive effects in the potential have been studied for hydrogen [129,139-140] and rare gases [73,135,137,141-153]. Barker gt g1. [154-158] have claimed with the support of experimental data that overlap-dependent manyebody interactions in rare gases must be irrelevant. However, the fact that the addition of long-range three-body interactions to the pair potential brings calculated properties of rare gases into agreement with experimental data does not in itself provide a high degree of physical insight [159] into the nature of the interactions. While the three-body exchange terms are considerably smaller than the two- 58 body [132d] at distances corresponding to the minimum of the van der Waals potentials, the three-body exchange contribution is significant [137,160- 161] when it is compared with the triple-dipole energy [129]. Although an overlap-dependent modification of the long-range forces [1323,162-163] has been used to explain the stability of the different rare gas solid structures, three-body exchange energy had earlier been considered crucial by Jansen gt gt. [164-171] in explaining the polymorphism of the rare gas and another isoelectronic solids. In spite of points made [172] in reply to criticisms [173], it has been suggested later [132a-b, 174] that Jansen gt _t. overestimated the nonadditive exchange contributions. The unavailability of reliable information on many body forces has prompted their accurate quantum mechanical computation [156,174], consistently to some specified order of magnitude [174]. Despite Barker’s remarks [157— 158] and the interpretive challenge posed [158] by the agreement of experimental data and results of calculations that omit three-body exchange effects, it is believed [127, 153, 174-177] that overlap~ dependent three-body forces cannot be fully disregarded and that they alone or in combination with other many-body effects may be the cause of discrepancies between experimental data and calculations of the dynamic structure factor and other properties [127,178]. Furthermore, it does not seem totally convincing to argue that overlap-dependent contributions are made irrelevant by the small statistical weight of regions where overlap is important, given that these contributions may be very large in those regions (orders of magnitude larger than in the statistically favored regions). The moment induced in a group of three atoms or molecules introduces 59 three-body effects in the ternary absorption coefficient [115,179]. This coefficient has been calculated [180] from experimental data for C02. Pair additivity in the dipole moment is usually assumed in dynamics calculations [39,181], but it has been suggested that many-body effects associated with the irreducible triplet dipole moment must be included in order for the calculated collision-induced absorption to reproduce the experimental line shapes [39,182]. Although binary collisions are the most important [183] at sufficiently low densities, the irreducible triplet dipole moment must make the leading contribution to absorption by monatomic unicomponent gases. Buckingham stated [184] that a set of three spherically symmetrical charge distributions should be polar due to induced moments, hence capable of absorbing electromagnetic radiation in the IR region of the spectrum. A model has been proposed [185] to calculate the spectrum due to ternary collisions in pure rare gases. Studies on the triplet dipole moment [186] have yielded the long range contribution as a power series of the internuclear distances. The coefficients of the leading terms have been calculated [187-188] for three hydrogen and three helium atoms, and the dispersion triplet dipole moment has been evaluated [189] for the Drude model. A variational method has been used [190] in order to explore the dipole moment of a hypothetical model neutral system with one electron in the field of three identical nuclei with a fractional charge each. The effects of applied pressure [191] and many-body effects [181,192-195] on interaction-induced depolarized light scattering [196] have also been discussed and calculations have been performed [197] to determine the triplet polarizability of helium. 60 A system of three hydrogen atoms in its quartet state is a good candidate to study three-body effects on the dipole moment with no interference from two-body contributions, and, in general to model interaction-related properties or potentials for three closed shell atoms. Nonadditivity in the potential energy has been studied [140,198-201] for spin-polarized H3, and the potential energy curves [199,201-204] and the Axilrod—Teller triple-dipole term [139,205] have been calculated at different levels of approximation. Hecht [206] predicted that gas or liquid completely spin polarized hydrogen would be endowed with superfluid behavior, and that it would recombine at densities where the three-body collisions became important unless an external magnetic field precluded such recombination. Monte Carlo calculations of bulk properties showed [207-208] that the system would exist as a gas (18,28) or liquid (3H), and liquid-to-gas and solid-to-gas phase transitions were studied [209-211] subsequently. Measurements on completely polarized atomic hydrogen at low temperatures have been reported [212-218], transport properties [219-220] have been determined, and the quantum Boltzmann equation [221] has been established for the gas in a regime where only binary collisions are important. Three-particle recombination has been considered [222] in the kinetics studies of the polarized gas, and experimental evidence for such a process has been reported [223-225]. In order to achieve Bose-Einstein condensation, magnetic confinement of the gas is under investigation [226]. Label—free exchange perturbation theory [74-75,]68,l72,227] has been applied to the study of interaction energies [73,168,171-172], interaction-induced molecular properties [76,228], and hyperfine structre 61 spectral line shifts [75]. A reformulation of the label-free exchange perturbation theory has been proposed [229] since model calculations for a non-interacting system [230] suggested that the original formulation [227] could not be satisfactory in practice. Further calculations [75,228] have shown that this method can be useful in approximating interaction-induced properties when overlap is non-negligible. We derive in Section 4.2 the expression for the dipole moment of H3 in its quartet spin state. Although the Hellmann-Feynman theorem is not satisfied at this level of approximation, for this system it can be shown that the dipole moment calculated as an energy derivative will be more accurate than the dipole calculated as an expectation value (section 4.2). It has been established [228] that the Lacey and Byers Brown method [70], which gives the lowest- order expectation value, accounts for only a small part of the exchange- overlap pair dipole moment. We conclude on the grounds of the treatment given in Section 4.2 that the Lacey-Byers Brown method provides no information at all about the dipole moment of certain triplets of identical atoms. Furthermore, the Gaussian model yields in both its original [164,166,231] and modified [232] forms an electron density which renders the exchange-antisymmetrization dipole moment [70] equal to zero for any cluster of identical atoms. The second-order dispersion triplet dipole moment has already been calculated [187-188]; the question then arises whether the total dipole can be approximated simply by adding the first-order exchange contribution. The issue of approximating interaction energies as the sum of second-order dispersion and first-order exchange contribution has been addressed by Margenau [233]. The calculations that he performed with a rather simple wave function showed that the 62 approximation was not acceptable for hydrogen, but it was almost correct for helium and it should be legitimate for heavier structures. Second- order nonadditive overlap-dependent contributions to the energy have been studied [166,170] in the Gaussian effective electron model, but it has already been established [132a,l74] that this approximation overestimates the second-order energies. The results of more accurate calculations [234] on hydrogen have shown that the first-order exchange energy between two ls atomic orbitals accounts for ~90% of the difference between the energies of the lowest singlet and triplet states. Second-order interaction energies involving exchange are considerably smaller than the first-order exchange contribution in the He...He interactions [235]. Higher—order exchange energies are expected to be a small percentage of the first-order exchange energy unless a strong bond is formed [132d]. Furthermore, it has been claimed that it is legitimate to calculate induction and first-order exchange energies independently, and this procedure has been used in the computation of interatomic interactions [236]; the results seem to support the validity of the approximation [236]. The same approximation for the second-order exchange contribution to the triplet dipole moment is adopted here. The results of our calculations discussed in Section 4.3 show that the exchange-overlap contribution to the triplet dipole moment cannot be disregarded in comparison with the long-range dispersion contribution. 4.2. DIPOLE MOMENT CALCULATIONS. In this section, the dipole moment for a system of three interacting hydrogen atoms in the lowest quartet state is obtained in the label-free exchange perturbation formalism. The Hamiltonian, H, for the system in a uniform static electric field is F {j (1) A A H = H(~=Q) + H(§) = HO+ V — ~ H(§) is associated with the interaction of the external electric field E and the hydrogen triplet. H(§=Q) is the sum of the Hamiltonian operators for the three non-interacting hydrogen atoms plus a term, 3, which contains all the electrostatic interatomic electron- nucleus attractions, and nucleus-nucleus as well as electron-electron repulsions. Each term in the Hamiltonian may be written in a form which is invariant with respect to assignment of the electrons to the interacting atoms: A p 1:]. A A A A where 0 represents one of the operators HO, V or B in (1), and " p O :2 z OpiAi - . (3) i=1 p The subscript p=a, b, c labels each hydrogen nucleus in the triplet system. The number of ways NP electrons from a total of N = §.NP may be assigned to three different centers is 63 p _ (Na+Nb+NC)! ' o N82Nb3NC. so in this case, p = 6. (4) When applied to an antisymmetrized triplet wave function 9.111 projects out the simple product term for the ith electron assignment. If the zeroth order save function, Qb’ is constructed as an antisymmetrized product of the ground-state save functions @a’ $5 and $6 for the isolated atoms: P [ 1 P . ‘I’o‘ ff: °p. Pi ‘I’a‘pb‘l’c ' 5:24“ ’ (5) 1- 1 1-1 p where f is a normalization constant and Z 0 Pi i=1 i is the intersystem antisymmetrizer. Thus, [1i is not self-adjoint and Ai‘l’o = “’1‘ The triplet wave function to first order in the applied field E and zeroth order in the interatomic interaction is given by 3 _ -1/2 (0) , -1/2 (1)_ l -3/2 (0) 0b- c1 Q + 5 [Cl 151 g i 2 gzcl é ] (5) The constants c1 and 92 are c1 = <<1>(°)| (°)> ('7) and c = 2 ‘32: <¢(°)| em> - (8) 2a 121 in ¢(0) is an antisymmetrized product of unperturbed orbitals x‘o) centered p (1) in is the antisymmetrized product with the unperturbed on p = a,b,c and é orbital i replaced by the first-order correction to orbital i in an applied field Pa in the a direction. The vector 3:1) has components 65 (l) (1) §(1) Qix , Q1 , éiz . To lowest order in the interaction, the triplet energy E is [228] = - 3};(0)=c‘1'1 <(0)> + i=1 A _ +F -'1'1<<1>{2c<¢(°)|H(F=-0)|£ g<1>>——czc12<.<°>.u(§=0)..<°>>_,} , (9) j: lJ where 6 g = <¢‘°)| z piAi]§(0)>Cll (10) i=1 With possible basis-set-dependent effects [81] confined to second and higher orders [228], the energy is given by the sum of three terms: the total energy of the three non-interacting atoms 3Eé0); the zero-field triplet interaction energy to the lowest order, and a term linear in the electric field. The dipole moment may be obtained as a field derivative of the energy (9): (C II a“ a1: A o 311%] = -81”. IHI~P0>++}F=0 = ~:0 ... (11) w = B - 2c;1 <(°)|§(F=0)| 24315 + Cl 2c2<<1>(°)m(§=-0)|q>(°)> " " i=1 Q1 For the three H—atom problem, if the term 7 were removed from the Hamiltonian H in (1) at §=0, an exact eigenfunction could then be calculated. Moreover, even if an external electric field were turned on, an exact wave function could still be obtained to any desired order in the electric field. The presence of 7 in the Hamiltonian introduces an error 8&0 in the zeroth—order wave function when 2:0. Therefore, when an electric field is applied the error has the form 8Q0(§)=60b(1+g(§)) where g(§) is some function of the electric field. The error in the dipole 66 moment calculated as A - _ _ .81! is first order in GO . Generally, the error in the field derivative of 0 the energy may be first or second order in 6&0' When 6wb(§) is of the form considered above though, (aE/aE)F=0 is second order in SQb [237]. Hence, the dipole moment as given by F5. (11) is in this case more accurate than the expectation value of the corresponding operator, as in (12). The AO’s x in the wave function 9 are constructed with Slater bases; xéo) is the H—atom ls orbital and the first—order correction to the hydrogenic wave function xél) is found analytically [93] and written in terms of modified p—Slater orbitals. We find that 9259’ according to the properties of overlap integrals with Slater orbitals on identical centers. The exchange-overlap contribution [70] to the dipole moment given by Eq. (12) may be analysed as follows. Let us consider a system of p identical atoms with a total of N electrons. Let us assign the set _ i ’ P1 - {xl, . . "XN/p} of A0 s to the first atom, the set "U N H Fla, 2‘ '3‘ z w to the second atom, and so forth. The set ,nx } is assigned to the pjib atom. The atomic P: a p {Km-m” P orbitals of the non-interacting atoms satisfy i i i 1.8 n 6 ...-S o ; , ’ : 1, so. , o 13 J J) PP J P P P ( ) i _ <(x )Pl(xj)P.> - (8 Thus, there are p2 (5 x3 )-dimensional blocks in the overlap matrix S. 67 All the diagonal blocks are (3 x 2 )-unit matrices. Consequently, we can decompose S as the sum of the unit matrix A and a matrix T which contains p (g x g )-diagona1 blocks whose elements are identically equal to zero. The inverse of the overlap matrix may be expanded in a power series (below, the implicit summation convention applies to the contracted covariant and contravariant indices, but Eq. 14c does not involve summation over i values): S-1=(A+T)-1=A‘T+T2-T3+. . . (14a) -li__i ik_ik9.. . . (S ) J— T 3+ T kT J T k? RT J , 1 t J (14b) -li_ ik_ik 9. (S ) i— 1 + T kT i T k1 l T i (14c) k=k_k k n_ 1 1- T n T an1* T menT 1 ... (14d) All the contracted indices are distinct from i (diagonal and off-diagonal matrix elements) and from 3 (off-diagonal matrix elements). If p(1.1’) = lx5(1)>(S—1)ij -(S’1)1J.(qu)ji . (16) a 1’91 Hence -e1__ 1 i k k 1 i _ _ i i _ i k 1 J ”a ’ [5 i+T RT 1 I 1T 1] ((111) i [T a” kaj T kT 1T j](qu) i .(17) Now i - i _ i p i _ i (qu) i- (x Iqulxi> - (x Iqu lxi> + (x lxi>RuP- 0 +RuPS i , (18) 68 where q: is referred to the p-nucleus and Rap is the coordinate of the p- nucleus in a given reference frame. Then Si.(q )i.= 2n 2 51.: 3 ER 1 a 1 P up iePP 1 p P up . (19) The product of a component of the dipole moment and an atomic orbital in Eq. (17) is expressed in terms of modified AO’s, and the spherical harmonics are transformed to the forms referred to canonical axes by means of the corresponding rotation matrices. The use of orthogonality relationships and specialization to s-type atomic orbitals let us write (see Appendix E) -el _ :N ”a - E ( p) Rap (20) The nuclear a—component is N - nuc N l p = Z " R -’ dv p(1,1’)‘-’ Z“ R ; (21) 0. pP upN [1,.4 1 pp up so — _ -e1 -nuc _ ”u - pa + ”a - 0 . Va (22) Therefore, the exchange-overlap contribution [70] to the dipole moment given by Eq. (12) vanishes when a cluster of hydrogen atoms is described by an approximate wave function such as @b° Furthermore, this result applies to any spatial arrangement of any number of atoms whose electrons are described by zeroth order s—orbitals, or by an effective 5 Gaussian function [164,166,231-232,242]. Thus, the dipole moment is given by -1 (0)“ 3 <1) B = -201 (23) ~ ~ ._ ~1 1-l Further elementary consideration on the properties of integrals involving Slater orbitals on identical atoms allow for additional simplification in 69 the expression of 2* namely, Eq. (23) becomes B 3‘2cil P103 ¢.(1)> (24) A “MW, “1 l i where 6’ is the operator 6 after removal of the terms which involve nucleus-nucleus repulsions. The constant c1 is given by C1 ; 6[:1“Sib-Sic—slz)c+28absacsbc ] ’ (25) where Spp, is the overlap integral between two 1s hydrogen orbitals on p,p’ = a, b, or c. The explicit formula for pa is given in Appendix F. Modified Slater-type orbitals for xél) have been taken in the directions 2 and x for a = z, x, respectively. The total dipole moment has been computed as the sum of the overlap-exhcange (Eq. (24)) and the long-range dispersion [187*188] contributions. Overlap damping, orbital distortion and higher—order effects are neglected. The calculations have been carried out for several values of H and Q (or XC for the right triangle configuration) between 3.0 a.u. and 10.0 a.u.; R, Q and KC are defined in Fig. 1 in the linear and equilateral, isosceles and right triangular configurations of the three hydrogen atoms a, b, c. This range of internuclear separations includes the van der Waals minimum of 32: H2 [243-244] and it covers the region where the pair distribution function is appreciably different from 0 and 1 [208]. All the integrals over Gaussian expansions [245] of the Slater-type orbitals which appear in the expression for the triplet dipole moment have been computed with the program ARGOS [246-248] maintained by R. Shepard (Argonne National Laboratory). - {I ‘9 4.3. RESULTS AND DISCUSSION. We have proven that the Lacey and Byers Brown [70] approximation yields zero for the triplet dipole moment when the calculation is performed for identical, spherically symmetric, neutral charge distributions described by s-type functions. This result generalizes to any number of such systems in any spatial configuration. Overlap exchange (Eq. (11)) and dispersion [118] triplet dipole moment calculations have been performed for the quartet H3 system in its right- triangle (e=w/2 rads) and linear (a=n rads) spatial configurations, with 3 a.u. $8 $10 a.u. and 3 a.u. $XC $10 a.u. (right triangle) and 3 a.u. $0 $10 a.u. (linear). The results for the triplet dipole moment in a.u. are plotted in Figs. 2-7. The X-dispersion dipole moment points from the pair to the single atom in most of the region XC > R (Fig. 2) whereas the reverse trend is observed for the Z-component in most of the region XC (B (Fig. 3). The behaviour of the X-overlap-exchange dipole moment (Fig. 4) is similar to that of the X-dispersion dipole moment (Fig. 2), but the Z-overlap- exchange dipole moment (Fig. 5) is directed form the pair to the single atom in most of the region R) XC. The overlap-exchange dipole moment (Fig. 6) for the linear configuration is directed from the single atom to the pair whilst the dispersion contribution (Fig. 7) points in the opposite direction, t.g., from the pair to the single atom. The order of magnitude of both contributions is comparable for all the configurations of Figs. 2—7. We do not present results for the isosceles-triangle configuration, but a similar trend is observed for its dispersion and overlap-exchange contributions. Therefore, the tiplet dipole moment is 70 71 substantially affected by overlap-exchange effects in addition to the dispersion contribution. An analogous result holds for the corresponding contributions to the energy [174]. The controversy over the actual rgle that the overlap-exchange contribution plays in the many-body interactions has been noted in the introduction. We have found that short-range effects in a molecular property are very important when they are compared with the long-range contribution. This theoretical result may be checked against experimental data by computing the dipole moment of a unicomponent triplet, as in Section 4.2. The collision-induced absorption spectrum may thus be calculated [182] and the importance of dispersion and overlap-exchange contributions properly analysed. Further research along this line is likely to shed some light on the relative weight of short- and long-range contributions to the many—body effects on properties of systems for which the leading terms are due to triplet and higher-order associations. The behavior of the total dipole moment (taken as the sum of the different contributions) may permit inferences [249] about the form of the triplet dipole moment time-correlation function. 72 Figure 1. Spatial configuration of three identical nuclei (a, b, c). a=w, "/2, n/3 for the linear, right-triangle and isosceles-triangle configurations. 73 ._'a X o mhwmuoocoo 345678910 R A \ A Figure 2. H3 dispersion dipole moment. X component. Right-triangle configuration. 74 ‘\ i’°"" b_ ’2' ls". .l. 1 ‘ .10'3 (N-bwmxlCDKDO 345678910 R Figure 3. H3 dispersion dipole moment. Z component. Right-triangle configuration. 75 1O \‘ -3 ‘ ,, \C‘1O \ ‘ \ -2 1 “x40 ‘\ \ . . 345678910 R _ wbwmuooco Figure 4. H3 overlap-exchange dipole moment. X component. Right-triangle configuration. 76 1O (N .2: (II 15) ‘4 (I) “0 K10“2 34 56 7 8910 R Figure 5. H3 overlap-exchange dipole moment. Z component. Right-triangle configuration. 77 1O 345678910 R (Al £> ()1 (3) ~J ‘00 “D Figure 6. H3 overlap-exchange dipole moment. Linear configuration. 10 CM .35 (11 15) ‘4 (I) (I) 78 345678910 R Figure 7. H3 dispersion dipole moment. Linear configuration. CHAPTER 5. 5.1. - INTRODUCTION. The interaction between two non-overlapping spherical neutral charge distributions is dominated by the London dispersion term [250-251]. When the charge distributions are so far away from each other that retardation effects must be taken into account, the interaction is still attractive, but its functional form in terms of the internuclear separation is slightly different. The use of the quantum theory of radiation [252] yields an attractive interaction inversely proportional to the seventh power of the separation. The retarded dispersion forces between macroscopic bodies may be calculated within the framework of classical electrodynamics [253], and the van der Waals forces between individual atoms and molecules are obtained as a special case [254-255]. Other treatments involving quatum electrodynamics have also been developed [256— 257] to account for retardation effects. The study of retarded dispersion forces covers the interaction between molecules [258], effects in dielectrics at finite temperatures [259], three-body dispersion forces [130] and the interaction between an atom and a surface [260]. The results of such theoretical studies have in common that the retarded and unretarded dispersion forces between macroscopic and microscopic bodies are given in terms of properties of the individual interacting species, 315;, in terms of susceptibilities. Therefore, dispersion forces may be viewed from a unified standpoint gig susceptibility theory [261]. The relationship among electromagnetic field fluctuations, susceptibilities, zero-point energy and long-range forces is discussed by Boyer [262]. Whether retarded (t.g., proportional to “-7) or not (i.e., proportional to 79 80 8.6), dispersion forces undergo a power-law divergence as the separation between the interacting systems goes to zero. Effects due to overlap and exchange should prevent this collapse by damping the power law which gives the dispersion energy. Second-order perturbation theory yields the dispersion contribution when the perturbation in the Hamiltonian is associated with the intersystem interaction. The matrix elements which involve the perturbation must be properly handled in order to include the overlap effects. The multipole series (which indeed is not intrinsic [88,263] to London’s theory) cannot be used to expand the perturbation, for the series expansion require that the charge distributions do not overlap. Matrix elements of operators such as er - rBI-l, where ru is the position of a particle of system a = A, B have to be evaluated. Their evaluation without use of a multipole expansion is usually carried out by separating the coordinate rA and r8 first. Several techniques (all relying on use of the convolution theorem) allow for such a separation. The Fourier transform of the potential is used by Koide [264]. His method has been developed and used to calculate damping function for Hez, Bez, HeH [265], and (32:)H2 [266], and to calculate interaction energy curves as well as dispersion damping interaction functions for He2 [267], Ar2 [268], Xe2 [269], (32:)Li2[270] and He...H2 and Ar...HCl [271]. An exchange correction term has been incorporated into the damped dispersion energy for (32:)Li2, (32f)LiNa, (32:)Li; and (22f)NaAr [272]. Other gt tgtttg computations of dispersion energies and damping functions have been performed [273] for Ar2 and L12. A universal, empirical damping function has been proposed and applied [274] to the calculation of the van der Waals potentials of (323H2, Arz, 81 (BZDNaK and Lng. Nonlocal polarizability densities [275] and Coulomb interactions in k- space have also been used to derive expressions for damped dispersion energies [276]. In other work, the Fourier transform of the transition-density matrix in the dispersion energy has been used [277] to provide a dispersion force between closed-shell atoms finite at all distances. This method has been extended [278] to represent the transition amplitudes in terms of Slater- type ortitals. Damping effects for the first-order Coulomb molecule- molecule interactions have been calculated on the basis of a two-centre or bipolar expansion fo er — rBI-1 [279], and the damped second-order Coulomb energy has been given in terms of response functions [280]. i; The full Coulomb interacting potential is also used in a reaction- field approach. A generalized form factor is related in this approach to the charge-density susceptibility with which we may calculate the second— order interaction energy. This method [281] is applied to the study of the He2 [281-282] and Ne2 [283] potentials. Damping in the dispersion energy has been treated empirically in electron-gas calculations, applied to the potential energy of H2...He, H2...Ne, H2...Ar [284] and N2...N2 [285]. Other procedures [286] are based on fitting expressions involving exponential damping to reliable results or experimental data. The dispersion contribution to the dipole moment [53-54,287-290]; has a power-law divergence (OCH-7) at zero internuclear separation. Expressions for the damped dispersion dipole moment have been given [291— 292] in terms of properties of the isolated systems. Nonlocal 82 polarizability densities [275] and the full Coulomb interaction have been used to calculate damped molecular properties such as dipoles, quadrupoles and polarizabilities [293-294] within a reaction-field theory. The damped dispersion-pair dipole is given [293] in terms of the Fourier transforms of the nonlocal frequency—dependent polarizability density of one system and the first hyperpolarizability density of the other system. We present in Section 2 of this chapter a general method to transform the nth-order charge-susceptibility density as obtained from generalized susceptibility theory [295] into a form which can easily be compared with the fully contracted nth—order nonlocal hyperpolarizability. We derive an expression for the nonlocal polarizability and hyperpolarizability densities in Section 5.3 and we give in Section 5.4 the expression for the damped dispersion pair dipole moment in terms of reduced matrix elements. 5.2.-GENERALIZED FUNCTIONS AND CHARGE-SUSCEPTIBILITY DENSITIES The expression 1 ln6(w) , (1) lim .[ = PV {*0 u_1C where PV stands for "principal value" and 8 is Dirac’s distribution, is used without proof in most work involving charge-susceptibility densities. Here we present a general method based on the theory of generalized functions [36,296] to carry out the limit involved in the charge- susceptibility density as obtained from perturbation theory. The method is equally applicable to susceptibility tensors, for the nature of the matrix elements in the numerators is irrelevant in this treatment. The method is general because it may be used with susceptibilities of any order - it is not restricted to the first order, to which Eq. (1) corresponds. This feature makes the method of potential use in the study of nonlinear phenomena, both in the nonresonant and resonant regions. The third-order nonlocal charge susceptibility density is [292] X(2)(£”a£’, E; w’, u) = [1 + P(£’,w’;£, (0)] 11!!! 75—2 £40 “40 . p0k(£”),pkn(£’).pnotg) 'Z Z{ (w +u’+ u-it in) ' (w +u- in) + k n 0k 0n 90k(5’) , Pkn (5”),pn0 (5) * («bk-out) - (no; O-in) " p (5’), (5), (5") + 0k, . Pkn POn’ . _ } . (2) (”OR'” +1£)°(ubn-uru +1: + 1n) where u and u’ are the frequencies of the field due to an external source, 83 84 ”bk is the difference of the energies (in units of 8) of the ground and excited (k) unperturbed states, p(£) is the charge-density operator and the primes in the sums stand for k # 0 and n x 0. The variables t and n are real and P(£’,w’;£,w) denotes the permutation of E, with E and of (u, - it) with (u-in) simultaneously. The matrix elements in the numerators of Eq. (2) are not relevant in the analysis that we perform. Let us consider the representative term lim 1 (*0 [(ubk’ u’ +i€)'(wbn+ w“ in)] n40 Define uOk-m’ 5 x, won+u 5 y Let l/(x.y) be a singular generalized function ( the terms distribution and generalized function are considered synonymous in this analysis). Let C3(fl?2) be the space of test functions ¢ = ¢(x,y) required to be infinitely differentiable and to vanish identically outside some finite interval; ttgt, {¢} are a class of infinitely differentiable functions with bounded support. The test functions ¢ = ¢(x,y) may also be elements of 5( [17(2), the set of infinitely differentiable functions such that k +k p a 1 2 sup llzll I—k k ¢(X.Y)| < .. . 5 5 (my) . (5) ac 1 ay 2 1:92: the set of functions which belong to C” and (together with their derivatives) vanish faster than any power when Igleta, although they do not have bounded support. Also, there must be constants K for all g and every k1, k2, p = 0, l, ..., such that k1.k2.p [a kl+k2 k llgllp- 1 k2 ¢(X.y) l 41”: .[bb (m it) (ragin) n*0 (x-a lit) (Y 02in) :11“ ...-gr“) 2 2 [¢(X,0)-¢(X,0)+¢(0,Y)-¢(0,Y) + X2+€ 2) (Y +n ) ¢(x.y) dxdy = +¢(0.0)+¢(X.Y)-¢(0.Qfldxdy - (4) The integrations in Eq. (4) are performed as follows (X-o it) (Y‘Oz i n) O 1 11m _ _ ¢(x,0) -—- dxdy = ¢(x.0) =::: -02 21 arctg (b/n) [Ea 87;:zigz) n40 -_ - £1.21. -_ . , sigma).- ozlfljfa x dx - 021" [Eb 8(y)dy PV ~a x dx]= 86 -olazuzfibfiadxdys(x,y).¢(x,y) . (5) Analogously, . (x-a i£)'(Y“o in) 11m a l 2 _ :40 [with “0’” 2 2 2 2 d“ d” ‘ n*0 (X +€ )‘(Y +n ) = - oliquadx 6(x) W [13315]?!) dy - °1°2"2EaJ1—)b¢(x’Y) 6(X.y)dxdy -(6) (x-oli€)'(Y‘ozin) (X2+£2)'(y2+n2) 1m ¢(o,0)[_a _b dxdy : = ¢(0.0) lim “-0121 arctg (a/£))'(-0221 arctg (b/n))] = {40 - n2 «0.0) . (7) (x-oli{)'(Y—ozin) 2) (114; fiafib [ ¢ +¢<0.0)—¢ - oziwjgbdy 6(y) PV -a x dx al"l-adXS(x) PVJEb y dy - 87 - zalazuzjfaijdxdys(x.y) ¢(x.y> + 2 a a x +0102» [_aijs(x,y)¢(x,y) dxdy + PVI_8JEb$L;;Xl dxdy . (9) Therefore, . 1 _ _1 _ 2 _. 1 11m (x+a i£)'(y+a in) - PV xy alazn 6(x,y) 1n[018(x)PVy + {*0 l 2 n*0 +0 5(y)Fvl ] (10) 2 x ’ where 6(x,y) is a two-dimensional Dirac distribution The representative result given by Eq. (10) is used to perform the limits in Eq. (2) to obtain x(2)(£”o£’,£;w’.m) = [1+P(£,ou,;£.u] ‘5-2 ' . . 90k(§”),pkn(§’),pn0(§) E W <~..+~'+~>-<~..+~> + Pou‘i')-"kn(5”)-Pno‘5) + ("Gk-w, ) . (”Onfi") (r’) (r), (5”) + POk ~ 'Pkn ~ FDn } (ubk‘u’)'(wbn-u-u’) ’"2{ °11°21P0k(5")'Pkn(5')'Pno(5)'““bk+“'+°'”bn+”) +012°2290k(5'"'Pkn‘L"’)"’no(5)'““bk"""“bn*“’) 88 Em} ' o +"13"23"0k(E ) Pkn (5) + “{Pm‘i >‘Fk.<£ >°Pno<£><-°u$<~ok+w +w>PV.,On..,, - ‘0216(wbn+w) PV _ ’. H. _ _’ P0k(5) Pkn(5 ) Pno(5)( °125(“0k ” ) PV 0 +m ___l__. -0228(w0n+u) PV ,) l _P0k(5 ) 'pkn(5)-p0n(§; )(‘0135(00k“m ) PV LOT-3;, - *0 6(0 ‘u-u’) PV 1 ) }] (ll) 23 On “bk-u, ' where 012 and 021 are equivalent to al and 02 in Eq. (10) and l = 1, 2, 3 h term in the sum within brackets in Eq.(2). refers to the it The procedure that we have followed here may be applied to a term with n factors in the denominator, t.g., to the (n+l)-order nonlocal susceptibility density. The analogue of Eq. (10) in such a case would involve the principal value of the inverse of the product of n variables, plus a combination of n, (n-l), (n-2),..., l-dimensional Dirac distributions each multiplied by the principal value of the inverse of the product of 0, l, 2,..., (n-l) variables (respectively) that are not contained in the corresponding Dirac distribution. 89 We obtain the expression for the linear (or second order) charge susceptibility density given by Linder and Rabenold [295] when either Eq. (10) or Eq. (11) is particularized to the case n = 1. Furthermore, the limits in the nth -order charge susceptibility tensor [297] and the nth order conductivity tensor [298] are easily obtained when we take £=n in Eq. (10), t.g., when we carry out the transition to purely real frequencies by taking a common imaginary component of all frequencies [297-298] to zero. 5.3.—NONLOCAL POLARIZABILITY AND HYPEHPOLARIZABILITY DENSITY CALCULATION Response tensors may be given as sum-over-all-states [293,299-300] or as restricted sum-over-states expressions [293,300-303]. The expressions given as sums over all states have apparent divergences (secular divergences) which should be removable [300] because their origin is a number of redundant terms coming from a phase factor in the wave function. This phase factor is expanded in powers of the field. Here, we first derive and expression for the first hyperpolarizability by removing the secular divergences from a sum-over-all-states formula [300]. Secondly, we calculate the reduced hyperpolarizability and polarizability which we shall need in section 5.4 for the calculation fo the damped dispersion dipole. The derivation is restricted to systems whose ground-state dipole moment is zero. The first nonlocal hyperpolarizability BuBY(5,§’,5”;iu,-iu,0) is [304] [F _(rnofltr awn“In [Pg 9* >1 0 (u m‘+iw)(mho+iu) (r,r’, r’ ’;iu,- 11.1.0): 132 Z[l+c.c. ]{— m,n 8,3, [P u(r)]0m[PB (r’ )] manY(E”)]n0 (u w'+iu)qho + [P(r’)] [Pam )1 [P(r”)l + fl~ 0m mm I.” n0 } ’ (12) (me-iu) Un0 where Pu(§) is the u—component of the polarization and c.c. indicates that one must take the complex conjugate of the expression following it. We now split the unrestricted sum in Eq. (12) and perform the integration 90 91 over the g”-variable to obtain the reduced hyperpolarizability A ’.o _o ) 8187(5’!’ ,10, 1:.) . .[P (1')] [p] [P (r’)] A 9.- _- _ ‘2 u~ W1kfl“ n0 Bwr(r,[ ,1... 1...) — F. [l+c.c.]{m’n (”no+i“’)(“’no+i°) (13.1) + Z.[Pu(§)]om[PB(g’)lmnlurlno (13.2) m,n (“ho+lw)“ho .[P(1;’)] [P (5)] [ J + z B %‘9 E. )4" ”I “0 (13.3) m,n ”hm 1” ”ho ll] (uw+iu)im . n iw(uno+iw) ° n 1:.) (duo +_2:[PB(£’)]oo[Pu(§)]on[ur]no } (13 7) n uh0(-iu) , O 0 Where 2: stands for the sum over m and n with the restriction m,n m s 0 and n t 0. In the integration to obtain Eq. (13) we have used I dE"[PY(E”)loo = [urloo = o (14) where pr is the recomponent of the dipole moment [293], t.g., the reduced A _ hyperpolarizability BnBY(£,§’;iu,-iu) is specialized to a system without a ground-state dipole. After partial-fraction decomposition of the terms (13.4) and (13.5) above and some rearrangement in the sum [l+c.c.], 92 cancellation occurs and we obtain the reduced hyperpolarizability as given in Eq. (15) . [P (r)]0m[uY]m[PB(§’)]no (5.5:;iu.-iw>=2l1+° C 1 {5 (“Midway“) 8&81 [P n(r)10m[PB (r’)]mn [u ran (“ho+1°)”no [PB(5’)JOmIPu(;)]mn[urlno (”mo-i“) mn0 [Pu(E)]om[UY1molpfl (5’)]00 ”mo(”ho+1”) 1 £3g(£)]00[PB(E’)JQE[”Y]m0 } O (15) — Z w (u -io) m m0 m0 Integration over the 5— and E’-variables in Eq. (15) yields the (fully reduced) first hyperpolarizability in complete agreement with the expression given by Buckingham [303]. This first hyperpolarizability, BuBY(i°’-iu)’ gives the static nonlinear dipole in the presence of the electric fields Hawk-‘1‘"t and Eé.°)ei°t. The static nonlocal hyperpolarizability density obtained from Eq. (15) by taking w = 0 agrees in full with the expression given by Hunt [293] in formula (2.37). Now the first-order correction to the wave function for a system perturbed by a uniform, static electric field in the u-direction is given by 93 (fllu '0) a [P lmo (l) ’ u _ an N )=£——N’>=ZI a m ‘“mo m m m0 olm) . (15) Reshuffling of indices in Eq. (15), the use of Eq. (16) and Fourier transformation allow us to write the hyperpolarizability in k-space as M(I)IP (k)|n> 7K wno +11») A 88711 (15,15’;iu,-iu)- [1+C. c. (1) a “(4!“ IPLQS )Im)|m> fl(uh0‘iw) A 9 kBk Yem(g,g :1w,’1w) = [l+c.c] {-5 . fl (who-iu) .<0|P(E)|m> 51(who+iw)(who+iu) n,m <0|P(§)I0><0|uu|m> +iw) + 1:2 E": m ”mo(“’mo <0|p(5’)|0><0|uu|m> + _. 1}. (19) ull0(ulllo 1”) The multipole moments Q:(k) are defined [264] by the equation m _ 4n 1/2 §21+123 . o,(k)i. 53Zu‘5iii) Y:(°a'¢e) , . JR(kru) (20) a 2 1. . . th . J1 is the Q spherical Bessel function, and (ra,eu,¢a) is the position vector of the particle a. where Zn is the electric charge of particle d, Then using the Rayleigh expansion of the plane wave exp(ig.£a)in the Fourier transform of the charge density p(k), we may write .. 9. , a " pm: 2 X "(34)? W141 Y,”(e.¢)* 020:) . (21) ~ [=0 m=-k With 5 = (k, 90 ¢)' Let the function «I»: (k,u) be defined by 1 l u + u no q}; (k,u) a if 2’ |n> . (22) n 95 We rewrite Eq. (19) with the aid of Eqns. (21) and (22) and we specialize it to spherically symmetric systems. Thus, a R 1’ (ng’iiwfiu) = [l+c.c.] Z Z Z GULF) . t A k k B Y BY“ l’lizo m:_n m!:_1) B [’Y'i'ke'.¢'>*¢E*<¢1mnIn?“"i:<“"iw>>5m~.m+m’ ‘M’fl * ’ a y * m, ’ m —’ "YJ;(9,¢) YE,(9 9¢ ) < ¢1m19|01’(k )|Ql(k’ 1m)>6m”’m’+m6£,i’t1 ’ , , t m _. m’ , . ‘Y';(99¢) YIE’(9,¢ ) <¢l(ka 1U)|Uu|§1!(k ’1w)>69_’1’:16m,m”+m,] at H -” 4" m +1y3(e.¢)*Y'{' (e’.¢')<0|vu|9'i' ("”iw)><°'°3(")'°> +— :10 [(-—1> I}, +YT”(9.¢)*Y3(6’.¢’)*<0|v¢'°lf’,(k”i“’)>(ed€ 6|cr> = 8.6.6 = [(2e+l)(2f+1)]1/2 W(abcd;ef), (32) where denotes a Clebsch—Gordan coefficient and W(abcd;ef) is the Racah coefficient, closely related to the 6-j symbol. we obtain [305] the damped dispersion dipole moment for a spherically symmetric system A _' g_ I“ I“, , ” .x+x’. . , , . (”disp)z' 1‘ "3 od‘I’Od-k dk xzx’=01 J)‘(kR)J)" (k R ) . 112w) sz'nnl/z 4" (xx 00|10> - [l+c.c’.]- L=0 1/2 [. 15%— 11:.» cm mum.)- [<0l|00(k)||0><0llulld>1(k’.iu)> + <0||c)0(k')||o><<1>l(k,i..,)llulI0>]esL’x + E C(1,1’,L).A(L,n’,x’)[(2£+1)(2x’+1)]1/2 W(11’xL;nx’) 1,1’=0 .{h(X.L.1)[<¢lllQn(k)Iléh,(k’aiw)>+<¢llIQ£,(k’)II§£(k,-iu))]3%r + +—1gLLL5%- <§1(k,-im)llull§ .(k’.iw) }]» (33) (22+1) where A(a,b,c) H)” [QWEQMD 11/2 [a b c] 4w 0 o o ’ (34) the double-bar matrix elements are the reduced matrix elements generated A A by the use of the Wigner-Rckart theorem, and we have taken R = Notice that the Clebsch-Gordan coefficients involved in Eq. (33) imply the existence of the factors 6 £,£’:l and 62.x’il . Therefore, we may rewrite Eq. (33) as (u‘ )- E “L Ex 5 (kmjdk’j (k mj'd... cw, k’ i... 9., >. 1.) (35) d‘sP z 1,L=0 x=l1-L| *1 (l+x+L= 2) where G(k,k’;iu;£,x,L) in Eq. (35) is defined by comparing this equation 100 and Eq. (33). An analysis analogous to Koide’s for the damped dispersion energy [264] holds for Eq. (35) (or Eq. (33)). Consequently, we may say that (pA. ) vanishes as R goes to O and as R goes to infinity. Computations disp z of (”disp)z and the dipole moment overlap-damping functions may be performed with the analytic expression given by Eq. (33). APPENDIX A. A derivation of Eq. (21) is carried out in this Appendix. Let us rewrite Eq. (20) (n) _ _ “’1 _L (1) on. A _ ( 1) Zni ér 90(9)H . . . 30(9)d9 (9.1) R (9) - E + s + (9— )s2 + (9— >253 + (A 2) 0 - xo-l >~O x0 ’ ' ' ' 30(9) 2 s + (9—90) 52 + (9—90)zs3+ . . . (9.3) S E 80(xo) (A.4) From (A.l) and (A.2), taking into account definitions (A.3) and (9.4), A‘“) - (-1)“"1 _1- § I E0 + s (9) J “(1) '“' [E 0 + - Zni F xO-l 0 ' ° ' xo-R +so(9)] d9 . (9.5) In particular E (1) 1 (1) 1 o A = . a (9)9 a (9)49 = . [ _ + 2w1§r0 o 2111i. N)“ + 80(1)] d1: E (1) __Q_ + 50(9)]H [ *0-1 10] 102 1. -— 80(1)H(1——Qd£ + xo-D. xo- —Qd9. 2"1¢r))‘0-9‘ ___. -.__ (1) _1_ (1) + . 4%. H 30(9)d9 + Zni §; 50 (9)H 30(9)d9 = -SH(1)E - E H 0 0 S (A.6) E (2)- _:l (1) (1) _ _:l 0 - 2.1 (fir 90(9):; 90(9)!) 90(9)d9 27:91 [W1 + E <1) 0 (1) [ t— 9 + So(l) ] H + 50(1) 1H 2 + S 0(1) 1 d1 0 z _-_1§ _E_0 H (hi9 H(1)119) 2191 90—9 xo-D. xo-D. 1" d9.- E E __Q (1) (1)__Q _ li.>O-9 H so”)H 90-9 d“ E (1) 0 (1) - . S (9.)H H @1‘ 0 xo-R. xo-D. E (1) (1) 0 - . S (9.)H S (1)11 d9. - g. 0 0 )0- E E - __l__ __9_ (1) __Q_ 2111 § xo-B. H xo-D. H(l)so(9.)d1— r E0 -§%§xo—o—9.- F H(1) SH(£)H(1)S (9)d9— 103 E (1) 0 xO-R H(1)So(£)d1 - _1_ ’ 2ni 4i.so(n)" (l) (1) SO(Q)H So(l)dfl = _1_ - 2ni éi'so(9)a _ (1) 2 (1) _ 2 (1) (1) 30H 3 a 30 s H 80H 30 + (1) (1) _ (1) (1) 2 + SH SH E0 EOH EOH s + + EOH(1)SH(1)S + SH(1)E0H(1)S (A.7) where Cauchy’s integral theorem has been used in both (A.6) and (A.7). It should also be noted that 50(1) fulfills _QE _ . 9+1 The generalization to obtain (21) is now straightforward. APPENDIX B Expectation values of operators in label-free form In this Appendix, a decomposition of the antisymmetrizer for an N- electron system is used to evaluate the expectation value of a permutation- invariant operator 9 in the state 0. The function Q is assumed to be a normalized, fully antisymmetrized product of N molecular orbitals on system A A and Ni on system B, with N=NA+NB. The N-electron antisymmetrizer.fl is related to an idempotent projection operator 0 by ,4 "1%: zapszN‘efl (3.1) P N’ and GP = 11, depending upon the parity of the permutation P. The operator 0 commutes where the sum runs over the permutations in the symmetric group S with any operator 9 that is invariant with respect to each of the permutations. In particular [0.11] = 0 (3.2) where H is the full N-electron Hamiltonian for the "supermolecule" AB; and if the unperturbed Hamiltonian H0 and the perturbation V due to A-B interaction are written in label-free form (Eqs. 2 and 30f chapter 3.2), [0,190] = [ON] = 0 ‘ (8.3) The operator 0 may be decomposed into terms that perform the antisymmetrization within systems A and B and terms that perform the antisymmetrization between systems. Lagrange’s theorem applied to the symmetric group S and its subgroup SN ® SN ensures that 0 may be split in N this way [113,114]: 104 105 N !N ! R 1 A 3 AB = __, Z a P = —, O 0 (ll + Z o.P. ) (8.4) (NA+NB)‘ P P (NA+NB). A B i=1 1 1 In Eq. 8.4, Ci is the antisymmetrizing projection operator for molecule A with NA electrons, defined by analogy tattlfor the full N-electron system, and.cg_is the corresponding operator for molecule B. n.denotes the identity operator. PiAB exchanges one or more particle labels assigned to A with labels assigned to B; it does not involve permutations within the set of electrons assigned to A or to B. The number 1 of operators PiAB is n 9 = Z (NA) ("3) . (3.5) 5:1 J .5 Since the upper limit n on the number of electrons exchanged is min (NA’NB)’ N! 9 o NA'NE' l = - l. (B.6) It is convenient to define P by 9 P = 2 0.3.“ . (3.7) . 1 1 i=1 Then the expectation value of Q in the state 0 is «FIRM» = N! <6{xl...xN} |n| 00%...pr / s (3.8) where the xi’s are orbitals located on A or B and 106 N! (UD‘I"'XN} |O{x1...xN}> (MA!)2 (NB!)2 8 = N! < 5A 03 (1+3) {x1"'xN} |OAOB (1+3) alum”) > NA! NB! < 0A OB {x1...xN} '09 GB (1+3) {xlmxfl} > < ¢A¢B l (1 + P) ¢A¢B > . (8.9) In Eq. 8.9 wk is the normalized antisymmetrized product wavefunction for system A (and *3 for B). An analysis similar to that for 5 shows _ l - s <¢A¢B|n|flAflB (1+3) alum”) > . (3.10) Substituting the form 2 QiAi for 0 yields _ 1 <~9|n|w> - S [<¢A¢B|nllwAwB> + p 9 93 + (wAwBIifzniAilflAflB {:1 0ij {x1'"xn}>]’ (3.11) where/l1 is the projection operator that assigns the first NA electrons to molecule A and the remainder to B. Each of the operators Ai in the second matrix element projects out a single P‘jAB term from the ket. Eq. 22 in section 3.3.1 is obtained by identifying n with V. Appendix C Pair energy of He...H The first-order pair energy E for the He...H system _ (0) (0) _ _- E - EH + EHe + V(Fz-0) + F2 {T1 + T2 + T3} ”ze (C.l) from Eq. 11, with -1 —1 -1 {-2+2‘2 anus mh—a V(Fz:0) : + +[++ 2} ’ (C.2) T = .2. (M) 1 . <3§°’->’ T2 =5 - +<><§|xne>1-<§.><>3|xne>- - + 3&3)[ + “9'31...” ] 9 “W __2_ _ -1 1 _ —1 1 -1 1 T3 - s 2 2 + 2 ’ 9% < + Whig) 107 108 -1 1 -1 1 + 2 + 2 + + < + “uni.” + + —1 1 2 -1 1 -1 1 <3).er Ixne> - - <3e> + <§e>> - Quasar}:Ime><><§e>+m‘z=°> “me”? + >]’ (0.5) and _ < I > p2 = ffl_§fl§_ 2 - 2R (C.8) _ 3 - [d rlxq(1)g(1)xr(1) (c.9) = I d3r1d3r2x§(l)xv(2) r1; xh(1)xu(2) (C.10) _ 3 (WV _ Id rlxq(l))%(l) (0.11) 109 is used in Eqs. C.1-C.6. Also, the internuclear distance is R, r12 is the interelectronic distance, and rq is the distance from an electron to nucleus q; g is any one—particle operator and q, v, m, u = H or He. The functions (1) (0) . . 1 xh and Xq from the main text are abbrev1ated as xq and xq, respectively. Zq is the atomic number of nucleus q. APPENDIX D . . -1 H . The hybrid integral (lsflelsflelr12|gpzlsfle> needed for the calculations in Chapter 13 is not one of the 79 hybrid integrals in terms of auxiliary functions compiled in Ref. 67. If we adopt the same conventions and symbology as in Ref. 67, the calculation proceeds in the following way. . . . . = H = Define the charge d1str1butions Ra - lsflelsHe and nab - lsHe 3pz where a - He and b E H. Q is a E —type distribution according to the standard ab two-centre charge distributions given in Table III of Ref. 67. Furthermore, fl = k w(£.n)exp(-m£)exp(-Bn) (D.1) 4“ _1_______ 22 E152 ‘n 2)w(£.n) - 2’ Z on fit n J: 4Jr§a (C‘n)(1‘{n)({ “n ) (D-Z) n= 0 j: —0 w(€an) = 5%;F§3~ (c-n)(1-:n) (3.3) k = 3 2(:1:)3/2(:: pz)”2 (v.4) (§)2(z—n)n = §;§r§5 (11:)3/2(:? 2) 1/2 ; 8XP(‘P£-PTn)(€ -n)2 (1— (n) (n. 5) and M = 0 = q. The expression (0.5) expands by one the number of terms given in Table III of Ref. 67. The one-centre distribution is ”He = [ISHe], and l-exp(-Sa(z-n))[1+% ;a(€-n)l = g “+3)U“ae (v.6) Therefore, u0 = 1, 111 = 1/2 110 111 _ (5‘3) 1 3 Is 3/2 z-1/2 4 .. 1 File [ISHellsHe3pz]: (:H 8) (:HP ) pH [0 dcj;ldn[1—e 9E3 “933: ((+3 ) kle P“*"“’4rl—~ (z- “)2 (1— an): _ 1 ls 3/2 3pz —1/2 4 .. 1 _ _ - HJHH (:He) (:H ) P3 jldsj_1dn{. 1m1 lzmz 1:112:11 4w(2£+l) 2 . C (llzl;m1,m2,m1+ m2) Yfi,m + m2(6.¢) (E5) 1 where 2 stands for the sum over at most two possible values 1=|1211| for 1, gig., i + 1 and [£2 — 1|. 2 E.3. Product of a component of the dipole moment and a Slater-type orbital. Let the normalized complex Slater-type orbital on centre a be 1 (n.1.m>a= (zza)"*’[<2n)sl‘l/Zrn’le"a’ Ylm(ea’¢a) . (E6) Then 1 - fl! 1/2 n+- , ‘1/2 n -{ r _ rnl.(n.R2.m2)a — ( 3) (2(8) 2[(2n)-] r e a Ylm1(ea.¢h)Y£2méea.¢h) - _ 2 [2124-1 11/2 1=Il211I _Ei:l— C(1121;000).C(112£;m1,m2,m1+m2) . [2n+2)(§p+ll 1/2 2{a (n+l,9.,m1+m2)a , (E7) where Eqns. (E3, ES, ES) have been used and (n+1, k, m1+m2)a’ is a modified Slater-type orbital on a defined by 3 _1 .. (n+1,£,ml+m2); = (2:8)“+5[(2n+2)s] /2rne ‘a’Yl,ml+m2(ea,¢h). (as) 0.4. Transformation of spherical harmonics under rotations. Let (X’, Y’, 2’) be the canonical reference frame. 0 denotes the rotation (passive interpretation) carrying the system (X, Y, 2) into coincidence with (X’,Y’,Z’). The transformation law is 114 t o _ X YXLJ(6 ,¢ ) — E Dmu(fl).Ym(e,¢), (89) where 0;” (Q) is the rotation matrix [241] defined by the convention given by Rose [238] and Messiah [239]. E.5. u—component of the electronic contribution to the dipole moment. The use of 8.1 through 8.4 let us write Eq. (17) in chapter 4 as e1__'ik_ik9. i_ ”a“ [611+TkTi T117931] (‘11) i _ _ i i _ i k 1 j _ [ T 3* T kaJ T k? llT 3 ] (r ) i _ k k 1 =3; (_12 )-212 z (1711* P up p21 up l€Pp k i k l P . . . + z z 2 [T1.- T113".- 14".“ p=1 iEPp jGPp J J (i#J) 1 J p . . . - z z z z [411 .- T113“.- le-rk p=1 iEPP jGPp 1=|1i21| 9 (1:3) A [ 21i+1 21 + 1 (2ni+2)(2ni+l)J1/z 2(i [ ((1131319) [(Biifiiw + p . . . + Z Z Z [T1 .41ka .+T1kwlen .] Ru p=1 iePp j¢PP 3' J J P J J J. . _ . <(n 1 m )|(nillimi )> - T1 w T .) + 1 T1.] - 9. T J.] 1/2 ] . C(l 111:000)C(1Ril;m1,mi, m1+ mi) . J _ 1- _ x: Z 11“., .(n1)n} ml). ’ ' J J m- m. 1111 mJ’ m m 1 1 (E.10.l.l) (E.10.1.2) (E.10.2.l) (E.10.2.2) (E.lO.2.3) (8.10.2.4) 115 p . . . — Z 2 z [7T1.+T1ka.-T1kwknTn.] - p21 iePp j¢Pp J J J 21.+1 1/2 Z 2., z - ‘ C(11.i;000) m; mJ n=|£itll[ 2Q+1 ] 1 L(2ni+2)(2ni+1)]1/2 . C(liil;m1,mi,ml+mi) 2:1 . J J J — - It” -1 “i '1 . < (n R m ’)l(n.1 m.)’> . D ., .(Q )'0 , (fl ), (8.10.2.5) 1 1 mJ mJ mimi ' where n.5 n.+1 ; m.E m +m. , and m.’ E m +m 3 1 1 1 l 1 1 l 1 The terms in (8.10.2.1) come from the sum (8.10.1.1). The sums (8.10.2.2) and(8.10.2.3) include the terms in (8.10.1.2) with the orbitals i and j on the same nucleus. The sums (8.10.2.4) and(8.10.2.5) include the terms in (8.10.1.2) with orbitals i and j on different nuclei. 8g. (87) has been used in (8.10.2.3) and (8.10.2.5). The inverse of Eq. (89) has been used in (8.10.2.4) and (8.10.2.5). Two spherical harmonics in the canonical reference frame on i and j satisfy (Y IY . .> = 6 j . (811) “imi iJmJ ”1” Furthermore, = 0 , (812) 1 1 1 m1 1 1 1 when 1i = IlJtll. The use of 8qns. (811), (812) and some orthogonality 116 relationships in Eq. (810) and its specialization to s—orbitals (i.e., 1i = 0, Vi) yield P p . . . . 11:" = E z (-Ru ) - z Ru 2 z [T1.TJi—T1k-rk.TJi] + P p=1 P p=1 Pie? j¢P J p . . . . + Z .Z X l?.,_p-['1".'rJ.-T1 7k.TJ.] = p=1 1€Pp j¢Pp J l R J 1 N P - p Z (-Rup) , q. e. d. (813) 9:1 APPENDIX F The dipole moment for the H3 system in its lowest quartet state is given by Eq. (24) in Section 4.2. The explicit expression for the transition element in Eq. (24) is 3 (0) A, (1) - 1 2 _ 2 <¢ |v ligl Qiu >-6{(SbCS 8C 2) (S2-SZL) + —1 1 -1 1 + (Sab~SaCSbC)((xb|rc 'xfia>++) + + (Sbc—Sacsab)(+—< | ‘1; 1 >+< l '1 | 1 > — ab xcxb r12 “taxi xbxc r12 x'uaxc xaxb r12 Xucxs —2< Ir-ll 1 >—< lr-ll 1 >) + J‘c’Sa 12 xuc’ia xcxb 12 Xaxuc -+< |r_1| 1 >-< I ’ll 1 >) + JTBS: 12 Xc’ina cha r12 xcxub +s (< '1 >-2< '1 1 >+< "1 1 >+ ch xbxa'rlzb‘aaxc m'rm'xuaxc xcxlrm'xuaa. + < Ir-ll 1 >-< | ’1| 1 >+< | ’II 1 >—< | -1| 1 >) + xbxa 12 xaxhc Xaxs r12 xaxuc chs r12 xaxab xsxc r12 Xaxub +s (< | ’II 1 >—< | ”1| 1 >+< | ’1| 1 > - ac xbxb r12 xhaxb “ext r12 Xuaxb xbxa r12 xficxb - < Ir-ll 1 >+< |r‘1| 1 >—2< Ir—ll 1 >+ xaxb 12 xucxs “bxc 12 Xaxub thb 12 thhb + magma.» 117 118 The notation =I33F1X§(1)8(1)Xfi: (1) (F2) -1 _ 3 3 —1 (qurlrlzlxnxu> _ Id rld r2x§(l)xv(1)r12xh(l)xb(2) (23) is used in Eq. (81). r12 is the interelectronic distance; g is the electron-nucleus distance; q, r, v, m, and u label the different nuclei a, 1 b, and c. xéo) and x(1) from the main text are abbreviated as xh and xh. The remaining symbols retain the same meaning as in Section 4.2. [ 1]. [ 2]. [10]. [11]. [12]. [13]. [14]. [15]. [16]. [17]. [18]. [19]. BIBLIOGRAPHY. P. Arrighini, "Intermolecular forces and their evaluation by Perturbation theory". Springer Verlag. Berlin, 1981. B. Jeziorski and W. Kolos in "Molecular Interactions". H. Ratajczak and W. J. Orville-Thomas eds., vol.3 pp.1 gt seq. John Wiley & Sons, 1982. . P. 0. stdin, J. Math. Phys. 3, 1171 (1962) . P. O. Lowdin, Int J. Quantum Chem. 31, 69 (1982). See also: P. O. Lowdin, in ”Perturbation theory and its applications in Quantum Mechanics" Ed. by C. H. Wilcox, pp. 255 et §_g. Publication n9 16 of the Mathematics Research Center. United States Army. The University of Wisconsin. J. Wiley and Sons, Inc. , New York, 1966. P. 0. Lawdin, J. Math. Phys. 6, 1341 (1965). P. 0. Lowdin, Phys. Rev. 139A, 357 (1965). P. 0. Lowdin, J. Chem. Phys. 43, $175 (1965). P. Q Lowdin, Int. J. Quantum Chem. 38, 137 (1968); 58, 685 (1971). . P. O. L5wdin, J. Math. Phys. 3, 969 (1962) . R. G. Wilson and C. S. Sharma, Int. J. Quantum Chem. 33, 927 (1981) [ 7]. . G. Speisman, Phys. Rev. 193, 1180 (1957) R. G. Wilson and C. S. Sharma, Int. J. Quantum Chem. 33, 183 (1982) . T. Kato, Progr. Theoret. Phys. 3, 514 (1949) T. Kato, Progr. Theoret. Phys. 3, 95 (1950) T. Kato, Progr. Theoret. Phys. 3, 207 (1950) T. Kato, J. Fac. Sci. Univ. Tokyo Sect.I 3, 145 (1951) F. Riesz and B. 82. Nagy, "Legons d’ Analyse Fonctionnelle”. Akadémiai Kiado. Budapest, 1968. M. H. Stone, Amer. Math. Soc. (Colloquium Publications) 13, 184 (1932) ' T. Kato, Trans. Amer. Math. Soc. 39, 195 (1951) W. R. Salzman, J. Chem. Phys. 43, 3035 (1968) K. A. Brueckner, Phys. Rev. 199, 36 (1955). J. Goldstone, Proc. Roy. Soc. (London) 5333, 267 (1957). N. H. Rugenholtz, Physica 33, 481 (1957). 119 [20]. [21]. [22]. [23]. [24]. [25]. [26]. [27]. [28]. [29]. [30]. [31]. [32]. [33]. 120 N. F. Mott and H. S. W. Massey, "The Theory of Atomic Collisions". Clarendon Press. Oxford, 1965. R. G. Newton, "Scattering Theory of Waves and Particles" McGraw Hill Book Co. New York, 1966. M. L. Goldberger and K. M. Watson, "Collision Theory". New York, 1964. K. A. Brueckner and C. A. Levinson, Phys. Rev. 33, 1344 (1955). v. Fock, z. Physik g1, 126 (1930). L. . C. Slater, Phys. Rev. 33, 210 (1930). C. C. J. Roothaan, Rev. Mod. Phys. 33, 69 (1951). G. C. Hall, Proc. Roy. Soc. (London) 3333, 541 (1951). A. Szabo and N. S. Ostlund, ”Modern Quantum Chemistry: Introduction to Advanced Electronic Structure Theory". MacMillan Publishing Co. New York, 1982. At this point we briefly analyse the nature of long-range interactions between molecules. If the instantaneous motion of the electrons is intercorrelated, we may expect that the charge distributions of two interacting atoms far away from each other will relax in such a way that the electron-correlation effects will yield minimum energy. The relaxation of the electron charge distribution is sometimes referred to as an interaction between instantaneous virtual electric moments on each interacting system. Although the time average of such moments is zero, the average of their effect on the other system is not. The effect of the virtual electric moments is to yeild a charge distribution on each centre polarized towards the other. Hence, each nucleus undergoes an attraction towards the centroid of its own electron charge distribution, and a net attractive force, 1.9., van der Waals force, exists between the nuclei [30,31]. One way of incorporating the effect of the virtual electric moments in the efpression for wx (8g. (54)) is through a multipolar expansion of H , even though such an expansion is not intrinsically associated with the theory of long-range interactions. We shall refer to this point in chapter 5. R. P. Feynman, Phys. Rev. 33, 340 (1939). R. F. W. Bader in "The Force Concept in Chemistry". Edited by B. M. Deb. Van Nostrand Reinhold Co. New York, 1981. B. H. Brandow in Proceedings of the International School of Physics "Enrico Fermi". Course 36. Edited by C. Bloch. Academic Press. London, 1966. B. H. Brandow, Rev. Mod. Phys. 33, 771 (1967). [34]. [35]. [36]. [37] [38] [39] [40] [41} [42] [43] [44] [45] [46] [47] [43] [49] [50] [51] [52] [53] [54] 121 H. Feshbach, Ann. Phys. N. Y. 3, 357 (1958). H. Feshbach, Ann. Rev. Nucl. Sci. 3, 49 (1958). V. S. Vladimirov, "Generalized Functions in Mathematical Physics". Mir Publishers. Moscow, 1979. Z.J. Kiss and R.L. Welsh, Phys. Rev. Letters 3, 166 (1959) 0.8. Bosomworth and H.P. Gush, Can. J. Phys. 33, 729,751 (1965) G. Birnbaum and R.D. Mountain, J. Chem. Phys. 31, 2347 (1984). R. Heastie and D. H. Martin, Can. J. Phys. 30, 122 (1962). ~ 8.8. Cohen and G. Birnbaum, J. Chem. Phys. 33, 2443 (1977). J.D. Poll and J.L. Hunt, Can. J. Phys. 33, 1448 (1981). G. Birnbaum, L. Frommhold, L. Nencini, and H. Sutter, Chem. Phys. Letters 133, 292 (1983). G. Birnbaum and H. Sutter, Mol. Phys. 13, 21 (1981). P.A. Madden and T.I. Cox, Mol. Phys. 53, 287 (1981) W. Klemperer, Ber. Bunsenges. Phys. Chem. 33, 128 (1974) and refs. therein; J. Mol. Struct. 33, 161 (1980). Principles, design and properties of molecular beams as well as the photochemistry of some van der Waals molecules are discussed by 0.8. Levy in "Quantum Dynamics of Molecules". Edited by R.G. Woolley. NATO-ASI Series B, Physics, Plenum Press. New York, 1980. J. M. Steed, T.A. Dixon, and W. Klemperer, J. Chem. Phys. 33, 4095 (1979). See also feature article by R.8. Miller, J. Phys. Chem. 33, 3301 (1986). P.W. Fowler and A. D. Buckingham, Molec. Phys. 33, 1349 (1983). A. D. Buckingham, Propriétés Optiques et Acoustiques des Fluides Comprimés et Actions Intermoléculaires (Centre National de la Recherche Scientifique, Paris, 1959) p. 57. L. Galatry and T. Gharbi, Chem. Phys. Letters 33, 427 (1980). K.L.C. Hunt, Chem. Phys. Letters 33, 336 (1980). M. Giraud and L. Galatry, Chem. Phys. Letters 33, 18 (1980). W. Byers Brown and D. M. Wishnant, Mol. Phys. 35, 1385 (1973). ~ 0. M. Whisnant and W. Byers Brown, Mol. Phys. 36, 1105 (1973). [55] [56] [57] [58] [59] [50] [51] [62] [63] [64] [65] [56] [67] [68] [69] [70] [71] [72] [73] [74] [75] [76] 122 D.P. Craig and T.Thirunamachandran, Chem. Phys. Letters 33, 14 (1981). G. Birnbaum, M. Krauss and L. Fromhold, J. Chem. Phys. 33, 2669 (1984). W. Meyer, in Proceedings of the NATO Advanced Research Workshop on Phenomena Induced by Intermolecular Interactions, G. Birnbaum, ed. (Plenum, New York, 1985). R.L. Matcha and R.K. Nesbet, Phys. Rev. 160, 33 (1967). C.F. Bender and 8.R. Davidson, J. Phys. Chem. 33, 2675 (1966). B.T. Ulrich, L. Ford, and J.C. Browne, J. Chem. Phys. 33, 2906 (1972) . P.D. Dacre, M01. Phys. 33, 17 (1982). R.L.C. Hunt and A.D. Buckingham, J. Chan. Phys. 33, 2832 (1980) R.L. Clarke, P.A. Madden, and A.d. Buckingham, Mol. Phys. 33, 301 (1978). P.R. Hilton and D.W. Oxtoby, J. Chem. Phys. 33, 1824 (1981). J.N. Murrell and G. Shaw, J. Chem. Phys. 33, 1768 (1967). A.T. Amos and J.I. Musher, Chem. Phys. Letters 1, 149 (1967); J.I. Musher and A.T. Amos, Phys. Rev. 133, 31 (1967). D.W. Oxtoby, J. Chem. Phys. 33, 1184 (1978). D.W. Oxtoby, J. Chem. Phys. 33, 5171 (1980). R.L.C. Hunt, J. Chem. Phys. 33, 393 (1984). A.J. Lacey and W. Byers Brown, Mol. Phys. 33, 1013 (1974). D.W. Oxtoby and W.M. Gelbert, Molec. Phys. 33, 535 (1975). L. Jansen, Phys. Rev. 162, 33 (1967); L. Jansen and 8. Lowardi, Chem. Phys. Letters, 1, 417 (1967). 8.8. Polymeropoulos, J. Brickmann, L. Jansen and R. Block, Phys. Rev. A 33, 1593 (1984). A.B. Ritchie, Phys. Rev. 131, 125 (1968); J. Chan. Phys. 49 2167 (1968). A.A. Rumyantsev, Zh. 8ksp. Teor. Fiz. 33, 926 (1973).[Sov. Phys.-J8TP 33, 459 (1974)]. J. Mahanty and C.K. Majumdar, Phys. Rev. A 33 2334 (1982). [77] [78] [79] [80] [31] [82] [83] [84] [85] [86] [87] [88] [89] [90] [911 [92] [93] [94] [95] [96] [97] 123 H. Margenau and P. Rosen, J. Chem. Phys. 31, 394 (1953) N. Rosen, Phys. Rev. 33, 255 (1931). (Corrections in Phys. Rev. 33, 2099 (1931)). J.O. Hirschfelder, Ann. Rev. Phys. Chem. 33, l (1983). S.T. Epstein, "The Variation Method in Quantum Chemistry", Academic Press, 1974. P.R. Certain and L.W. Bruch, in "MTP International Review of Science", Phys. Chem. Series 1, volume 1: Theoretical Chemistry, W. Byers Brown, ed. (Butterworths, London, 1972). U. Landman and R. Pauncz, Chem. Phys. Letters 3, 489 (1971). J. 0. Hirschfelder and P.R. Certain, Int. J. Quantum Chem. S 3, 125 (1968). H. Eisenschitz and F. London, Z. Physik 33, 491 (1930). A. van der Avoird, J. Chem. Phys. 32, 3649 (1967). J.D. Hirschfelder, Chem. Phys. Letters 1, 325 (1967). J.D. Hirschfelder, Chem. Phys. Letters 1, 363 (1967). R. McWeeny, Proc. Roy. Soc. London A 333, 242 (1959). R. McWeeny, and B.T. Sutcliffe, Proc. Roy. Soc. London A 273, 133 (1963). P.D. Dacre and R. McWeeny, Proc. Roy. Soc. London 1 313, 435 (1970). V. Magnasco, G.F. Musso, and 8” McWeeny, J. Chem. Phys. 33, 4617 (1967). R.D. Amos, Chem. Phys. Letters 33, 89 (1982). P.S. Epsetin, Phys. Rev. 33, 695 (1926). R.S. Mfilliken, C.A. Rieke, D. Orloff, and H. Orloff, J. Chem. Phys. 13, 1248 (1949); 9:. also Ref. 95. A. Lofthus, Mol. Phys. 3, 105 (1962). M. Hotani, A. Amamiya, 8. Ishiguro, and T. Himura, ”Table of Molecular Integrals”. (Maruzen, Tokyo, 1955). M.P. Barnett and C.A. Coulson, Phil. Trans. 1 333, 221 (1951). [98] [99] [100] [101] [102] [103] [104] [105] [106] [107] [108] [109] [110] [111] [112] [113] [114] [115] [116] [117] [118] 124 M.P. Barnett, in "Methods in Computational Physics" vol. 2, B. Alder, S. Fernbach, and M. Rotenberg, eds. (Academic Press, New York, 1963). C.C.J. Roothaan, J. Chem. Phys. 13, 1445 (1951). K. Rfidenberg, J. Chem. Phys. 13 1459 (1951). C.C.J. Roothaan and Rfidenberg, J. Chem. Phys 33, 765 (1954). K. Ruedenberg, C.C.J. Roothaan, and W. Jaunzemis, J. Chem. Phys. 33, 201 (1956). D.B. Knowless, J.N. Murrell, and J.P. Braga, Chem. Phys. Letters 113, 40 (1984). A.D. McLean and M. Yoshimine, J. Chem. Phys. 33, 3682 (1967). S. Huzinaga, J. Chem. Phys. 33, 1293 (1965). J.L. Dodds, R. McWeeny, and A.J. Sadlej, Mol. Phys. 33, 1779 (1977). 8.0. Roos and A.J. Sadlej, Chem. Phys. 33, 43 (1985). A. Koide, J. Phys. B 3, 3173 (1976). M. Krauss, D.B. Neumann, and W.J. Stevens, Chem. Phys. Letters, 11, 500 (1980). A. Koide, W.J. Meath, and A.R. Allnatt, Chem. Phys. 33, 105 (1981). P. Malinowski, A.C. Tanner, K.F. Lee, and B. Linder, Chem. Phys. 33, 423 (1981). G.H.F. Diercksen, B.O. Roos, and A.J. Sadlej, Chem. Phys. 33, 29 (1981). C.D.H. Chisholm, "Group Theroetical Techniques in Quantum 33emistry", Academic Press, New York (1976). J.C.C.M. van Duijneveldt-van de Rijdt and F.B. van Duijneveldt, Chem. Phys. Letters. 13, 425 (1972). J. van Kranendonk, Physics 23, 825 (1957). G. Birnbaum, Proceedings of the International School of Physics ”Enrico Fermi". Course LXXV. Edited by J. van Kranendonk. North Holland Publishing Co.; London, 1980. I. R. Dagg, L. A. A. Read and W. Smith, Can. J. Phys. 33, 1431 (1982). U. Buontempo, S. Cunsolo, P. Dore and L. Nencini, Can. J. Phys. 125 [119] For a list of refs. on experiments carried out with liquids, see [120] [121] [122] [123] [124] [125] [126] [127] [128] [129] [130] [131] [132] [133] [134] [135] [136] [137] [138] B. Catlow and J. Yarwood, J. M01. Struct. (Theochem) 133, 415 (1986). A. I. Baise, J. Chem. Soc. Faraday Trans. II 33, 1907 (1972). S. Weiss, Can. J. Phys. 33, 584 (1976). F. Sullivan, R. 0. Mountain and J. O’Connell, J. Comput. 138 (1985). Phys. 31, A. D. Buckingham and B. D. Utting, Ann. Rev. Phys. Chem. (1970). g1, 287 H. Margenau and N. R. Kestner, "Theory of Intermolecular Forces". Pergamon Press; Oxford, 1969. J. A. Barker, R. A. Fisher and R. 0. Watts, M01. Phys. 31, 657 (1971). M. Wojcik and 8. Clementi, J. Chem. Phys. 33, 5970 (1986). J. J. Salacuse, W. Schommers and P. A. 8gelstaff, Phys. Rev. A 33, 1516 (1986). C. Hoheisel, Phys. Rev. A 33, 1998 (1981). B. M. Axilrod and 8. Teller, J. Chem. Phys. 11, 299 (1943). A. D. McLachlan, Mol. Phys. 3, 423 (1963). J. M. Standard and P. R. Certain, J. Chem. Phys. 33, 3002 (1985), and refs. therein. a) K. F. Niebel and J. A. Venables; b) J. A. Barker; c) R. J. Bell and I. J. Zucker; d) J. N. Murrell, in Rare Gas Solids. Edited by M. L. Klein and J. A. Venables. (Academic Press. London, 1976). Vol. I. J. A. Barker, M. L. Klein and M. V. Bobetic, IBM J. Res. Develop. 33, 222 (1976). I. L. Garzén and 8. Blaisten-Barojas, Chem. Phys. Lett. 133, 84 (1985). B. H. Wells and S. Wilson, Chem. Phys. Lett. 131, 429 (1983). B. H. Wells and S. Wilson, M01. Phys. 33, 1295 (1983). . Wells and S. Wilson, Mol. Phys. 33, 199 (1985). . Wells and S. Wilson, Mol. Phys. 32, 21 (1986). [139] [140] [141] [142] [143] [144] [145] [146] [147] [148] [149] [150] [151] [152] [153] [154] [155] [156] [157] [158] [159] [160] [161] 126 R. J. Bell, J. Phys. B Atom. Molec. Phys. 3, L63 (1970). W. Kolos and A. Leé, Chem. Phys. Lett. 13, 167 (1972). P. Rosen, J. Chem. Phys. 31, 1007 (1953). A. Shostak, J. Chem. Phys. 33, 1808 (1955). 8. W. Schmid, J. Schwager, Y. C. Tang and R. C. Herndon, Physica 31, 1143 (1965). D. R. Williams, L. J. Schaad and J. N. Murrell, J. Chem. Phys. 32, 4916 (1967). R. F. W. Bader, 0. A. Novaro and V. Beltréh-Lépez, Chem. Phys. Lett. 3, 568 (1971). 0. A. Novaro and V. Beltrén-Lépez, J. Chem. Phys. 33, 815 (1972). L. W. Bruch and I. J. McGee, J. Chem. Phys. 33, 409 (1973). L. Gonzélez-Tovany, V. Beltrén-Lépez and 0. Novaro, J. Chem. Phys. g9, 1694 (1974). A. J. c. Varandas, M01. Phys. 1g, 817 (1983). M. Bulski and G. Chalasifiski, Theoret. Chim. Acta (Berl.) 33, 199 (1980). M. Bulski, Chem. Phys. Lett. 33, 361 (1981). J. P. Daudey, O. Novaro and M. Berrondo, Chem. Phys. Lett. 33, 26 (1979). M. Bulski and G. Chafasifiski, Chem. Phys. Lett. 33, 450 (1982). J. A. Barker and M. L. Klein, Chem. Phys. Lett. 11, 501 (1971). J. A. Barker and M. L. Klein, Phys. Rev. 3 Z, 4707 (1973). J. A. Barker, R. 0. Watts, J. K. Lee, T. P. Schafer and Y. T. Lee, J. Chem. Phys. 31, 3081 (1974). J. A. Barker, Mol. Phys. 33, 755 (1986). J. A. Barker, Phys. Rev. Lett. 33, 230 (1986). B. J. Alder, Ann. Rev. Phys. Chem. 33, 325 (1973). . McGinnies and L. Jansen, Phys. Rev. 131, 1301 (1956). . O’Shea and W. J. Meath, Mol. Phys. 33, 1431 (1974). [162] [163] [154] [165] [166] [167] [168] [169] [170] [171] [172] [173] [174] [175] [176] [177] [178] [179] [180] [181] [182] [183] 127 K. F. Niebel and J. A. Venables, Proc. 365 (1974). Roy. Soc. (London) 5 333, S. Carter and J. N. Murrell, Mol. Phys. 33, 1269 (1976). L. Jansen, Phys. Rev. 133, A1292 (1964). Rev. E. Lombardi and L. Jansen, Phys. 133, A1011 (1964). L. Jansen, Adv. Quantum Chem. 3, 119 (1965). L. Jansen and E. Lombardi, Disc. Faraday Soc. n9 49, 78 (1965). Lett. L. Jansen and E. Lombardi, Chem. Phys. 1, 417 (1967).-a- G. G. Chell, J. Phys. C (Proc. Phys. Soc.) 1, 266 (1968). E. Lombardi and L. Jansen, Phys. Rev. 133, 822 (1968). E. Lombardi, L. Jansen and R. Ritter, Phys Rev. 133, 1150 (1969). L. Jansen and E. Lombardi, Chem. Phys. Lett. 1, 33 (1967). C. E. Swenberg, Phys. Lett. 341, 163 (1967). W. J. Meath and R. A. Aziz, Mol. Phys. 33, 225 (1984). J. A. Barker and M. V. Bobetic, J. Chem. Phys. 33, 6306 (1983). R. A. Aziz in Inert Gases. Edited by M. L. Klein. Springer Series in Chemical Physics 34. (Springer Verlag. Berlin, 1984). R. A. Aziz, J. Chem. Phys. 33, 5787 (1980). P. A. Egelstaff, W. Glaser, D. Litchinsky, E. Schneider and J. B. Suck, Phys. Rev. 4 33, 1106 (1983). J. van Kranendonk, in Propriétés Optiques et Acoustiques des Fluides Comprimés et Actions Intermoleculaires. CNRS. Paris, 1959. G. Birnbaum, W. Ho and A. Rosenberg, J. Chem. Phys. (1971). §§, 1039 D. Frenkel, in Proceedings of the International School of Physics "Enrico Fermi". Course LXXV. Edited by J. van Kranendonk. North Holland Publishing Co.; London, 1980. A. Raczyfiski and G. Staszewska, Mol. Phys. 33, 919 (1986). R. L. Welsh, in MTP International Review of Science. Spectroscopy. Physical Chemistry. Series One. Vol. 3.Edited by D. A. Ramsay. Butterworths. London, 1972. [184] [185] [186] [187] [188] [189] [190] [191] [192] [193] [194] [195] [196] [197] [198] [199] [200] [201] [202] [203] [204] I 205] 128 A. D. Buckingham in Propriétés Optiques et Acoustiques des Fluides Comprimés et Actions Intermoléculaires. CNRS. Paris, 1959. S. Weiss, Chem. Phys. Lett. 13, 41 (1973). C. G. Gray and B. W. N. Lo, Chem. Phys. Lett. 33, 55 (1974). P. H. Martin, Mol. Phys. 33, 129 (1974). L. W. Bruch, C. T. Corcoran and F. Weinhold, Mol. Phys. 33, 1205 (1978). L. W. Bruch and T. Osawa, Mol. Phys. 43, 491 (1980). G. Adam and J. Katriel, Chem. Phys. Lett. 31, 451 (1973). M. Udagawa, K. Nakamura, M. Murakami and K. Ohbayashi, Phys. Rev. 3 34, 1563 (1986). U. Balucani and R. Vallauri, M01. Phys. 33, 1099, 1115 (1979). P. A. Madden, Mol. Phys. 33, 365 (1978). W. M. Gelbart, Adv. Chem. Phys. 33, 1 (1974). B. J. Alder, J. J. Weis and H. L. Strauss, Phys. Rev.~A Z, 281 (1973). A. D. Buckingham and M. J. Stephen, Trans. Faraday Soc. (1957). 9.5.3. 884 J. J. Perez, J. H. R. Clarke and A. Rinchliffe, Chem Phys. Lett. 134, 583 (1984). H. Margenau and J. Stamper, Adv. Quantum Chem. 3, 129 (1967). R. Block, R. Roel and G. Ter Maten, Chem Phys. Lett. 11, 425 (1971). S. F. O’Shea and W. J. Meath, Mol. Phys. 31, 515 (1976). J. N. Murrell, A. J. C. Varandas and M. F. Guest, Mol. Phys. 31, 1129 (1976). ’ A. V. Zaitsevskii, A. V. Nemukhin and N. F. Stepanov, M01. Phys. 41, 377 (1980). K. Kaufmann, M. Jungen and H. J. Werner, J. Phys. Chem. 33, 3806 (1983) . A. C. Roach and P. J. Kuntz, J. Chem. Phys. 34, 822 (1986). T. Koga, J. Chem. Phys. 34, 1636 (1986), and refs. therein. [206] [207] [208] [209] [210] [211] [212] [213] [214] [215] [216] [217] [218] [219] [220] [221] [222] [223] [224] 129 C. E. Hecht, Physica 33, 1159 (1959). J. V. Dugan Jr., and R. D. Etters, J. Chem. Phys. 33, 6171 (1973). R. D. Etters, J. V. Dugan Jr., and R. W. Palmer, J. Chem. Phys. 33, 313 (1975). M. D. Miller, L. H. Nosanow and L. J. Parish, Phys. Rev. Lett. 33, 581 (1975). W. C. Stwalley, Phys. Rev. Lett. 33, 1628 (1976). W. C. Stwalley and L. H. Nosanow, Phys. Rev. Lett. 33, 910 (1976). Crampton, T. J. Greytak, D. Kleppner, W. D. Phillips, Smith and A. Weinrib, Phys. Rev. Lett. 43, 1039 (1979). >.u: S. D. W. N. Hardy, A. J. Berlinsky and L. A. Whitehead, Phys. Rev. Lett. 43, 1042 (1979). Hardy, M. Morrow, R. Jochemsen, B. W. Statt, P. R. Kubik, Marsolais and A. J. Berlinsky, Phys. Rev. Lett. 43, 453 0). 5‘? W. R. (19 no I. F. Silvera and J. T. M. Walraven, Phys. Rev. Lett. 44, 164 (1980). R. W. Cline, D. A. Smith, T. J. Greytak and D. Kleppner, Phys. Rev. Lett. 43, 2117 (1980). R. W. Cline, T. J. Greytak and D. Kleppner, Phys. Rev. Lett. 43, 1195 (1981). J. T. M. Walraven, I. F. Silvera and A. P. M. Mattbey, Phys. Rev. Lett. 43, 449 (1980). C. Lhuillier and F. Laloe, J. Physique 43, 197 (1982). C. Lhuillier and F. Laloe, J. Physique 43, 225 (1982). J. P. Bouchaud and C. Lhuillier, J. Physique 43,1101 (1985). Yu. Kagan, I. A. Vartanyants and G. V. Shlyapnikov, Zh. 3333i)Teor. Fiz. 81, 1113 (1981). (Sov. Phys. JETP 34, 590 R. Sprik, J. T. M. Walraven and I. F. Silvera, Phys. Rev. Lett. 31, 479, 942 (1983). H. F. Hess, D. A. Bell, G. P. Kochanski, R. W. Cline, D. Kleppner and T. J. Greytak, Phys. Rev. Lett. 31, 483 (1983). [225] [226] [227] [228] [229] [230] [231] [232] [233] [234] [235] [236] [237] [238] [239] [240] [241] [242] [243] [244] [245] 130 H. F. Hess, D. A. Bell, G. P. Kochanski, D. Kleppner and T. J. Greytak, Phys. Rev. Lett. 33, 1520 (1984). R. V. E. Lovelace, C. Mehanian, T. J. Tommila and D. M. Lee, Nature 318, 33 (1985). L. Jansen, Phys. Rev. 162, 33 (1967). J. Juan6s i Timoneda and K. L. C. Hunt, J. Chem. Phys. 34, 3954 (1986). F: Byers Brown, Chem Phys. Lett. 3, 105 (1968). S. T. Epstein and R. E. Johnson, Chem. Phys. Lett. 1, 602 (1968). L. Jansen, Phys. Rev. 133, 1798 (1962). W. Kolos and A. Les, Int. J. Quantum Chem. 3, 1101 (1972). H. Margenau, Phys. Rev. 33, 1000 (1939). P. R. Certain, J. O. Hirschfelder, W. KoIos and L. Wolniewicz, J. Chem. Phys. 43, 24 (1968). K. Szalewicz and B. Jeziorski, Mol. Phys. 33, 191 (1979). E. E. Nikitin and 8. Ya. Umanskii. Theory of Slow Atomic Collisions. Springer Series in Chemical Physics 33. (Springer Verlag. New York, 1984), and refs. therein. 8. T. Epstein, in The Force Concept in Chemistry, edited by B. M. Deb (Van Nostrand Reinhold, New York, 1981), Chap. 1. M. E. Rose, Elementary Theory of Angular Momentum. J. Wiley & Sons. New York, 1957. A. Messiah, Quantum Mechanics. North-Holland Publishing Co.; New York, 1963. M. Weissbluth, Atoms and Molecules. Academic Press; New York, 1980. C. G. Gray and K. E. Gubbins, Theory of Molecular Fluids. Clarendon Press; Oxford, 1984. E. Lombardi, C. Almirante and L. Jansen, Chem. Phys. Lett. 133, 1 (1984). W. Kolos and L. Wolniewicz, J. Chem. Phys. 43, 1426 (1967). W. KoIos and L. Wolniewicz, J. Chem. Phys. 43, 2429 (1965). R. F. Stewart, J. Chem. Phys. 33, 431 (1970). [269] [270] [271] [272] [273] [274] [275] [276] [277] [278] [279] [280] [281] [282] [283] [284] [285] [286] [287] [288] [289] 131 M. Krauss, W. J. Stevens and D. B. Neumann, Chem. Phys. Lett. 31, 500 (1980). M. E. Rosenkrantz, 186§ACS Meeting 1983, Phys. Abstract 87. M. E. Rosenkrantz, 187Q ACS Meeting 1984, Phys. Abstract 56. M. E. Rosenkrantz, R. M. Regan and D. D. Konowalow, J. Phys. Chem. 33, 2804 (1985). P. J. Knowles and W. J. Meath, Chem. Phys. Lett. 134, 164 (1986). K. T. Tang and J. P. Toennies, J. Chem. Phys. 33, 1496 (1977); 3Q, 3726 (1984). W. J. A. Maaskant (1964). and L. J. Oosterhoff, Mol. Phys. 3, 319 x. L. c. Hunt, J. Chem. Phys. zg, 6149 (1983). N. Jacobi and Gy. Csanak, Chem. Phys. Lett. 33, 367 (1975). M. Battezzati and V. Magnasco, J. CHsm. Phys. 33, 2924 (1977). A. Koide, T. R. Proctor, A. R. A11natt and W. J. Meath, M01. Phys. 33, 491 (1986). P. W. Langhoff, Chem. Phys. Lett. 33, 33 (1973). B. Linder, K. F. Lee, P. Malinowski and A. C. Tanner, Chem. Phys. 33, 353 (1980). P. Malinowski, A. c. Tanner, K. F. Lee and B. Linder, Chem. Phys. 33, 423 (1981). S. Wozniak, K. F. Raof and B. Linder, J. Chem. Phys. 33, 220 (1983). R. LeSar, J. Phys. Chem. 33, 4272 (1984). R. LeSar and M. S. Shaw, J. Chem. Phys. 34, 5479 (1986). J. Hepburn, G. Scoles and R. Penco, Chem. Phys. Lett. 33, 451 (1975), and references therein. W. Byers Brown and D. M. Whisnant. Chem. Phys. Lett. 3, 329 (1970). S. F. Abdulnur, Chem. Phys. Lett. 33, 355 (1973). K. L. C. Hunt, Chem. Phys. Lett. 23, 336 (1980). [290] [291] [292] [293] [294] [295] [296] [297] [298] [299] [300] [301] [302] [303] [304] [305] 132 K. L. C. Hunt and J. E. Bohr, J. Chem. Phys. 33, 5198 (1985). M. Giraud and L. Galatry, Chem. Phys. Lett. 33, 18 (1980). B. Linder and R. A. Kromhout, J. Chem. Phys. 34, 2753 (1986). K. L. C. Hunt, J. Chem. Phys. 33, 393 (1984). K. L. C. Hunt and J. E. Bohr, J. Chem. Phys. 34, 6141 (1986). B. Linder and D. A. Rabenold, Adv. Quantum Chem. 3, 203 (1972). R. D. Richtmyer, "Principles of Advanced Mathematical Physics", vol. I.(Springer Verlag. New York, 1978). J. Ducuing, in Proceedings of the International School of Physics "Enrico Fermi". Course XLII. Edited by R. J. Glauber. Academic Press. New York, 1969. P. N. Butcher and T. P. McLean, Proc. Phys. Soc. 31, 219 (1963). S. Kielich, Acta Phys. Polon. 33, 875 (1966) D. C. Hanna, M. A. Yuratich and D. Cotter, "Nonlinear Optics of Free Atoms and Molecules". (Springer Verlag. Berlin,1979). N. Bloembergen, "Nonlinear Optics". (W. A. Benjamin Inc. Reading (MA), 1965). B. J. Orr and J. F. Ward, Mol. Phys. 33, 513 (1971). A. D. Buckingham, in ”Nonlinear Behaviour of Molecules, Atoms and Ions in Electric, Magnetic or Electromagnetic Fields". Edited by L. Neel. (Proceedings of the 3lst International Meeting of the Societé de Chimie Physique, 1978). (Elsevier Scientific Publishing Co. Amsterdam, 1979). We focus our attention on the susceptibility at nonresonant frequencies. Therefore, we drop all the terms in Eq. (11) which involve Dirac distributions since we consider in this section the susceptibility and hyperpolarizability densities at purely imaginary frequencies. K. L. C. Hunt and J. Juanés i Timoneda, to be published.