'2..:..... A. .5. yr}; .1 $112.1. PE. . . z. 1.2.; I. Y $33.41.! .. .5 1:. ....v....9... . >533... Q .3 . \ Ivo. ti V I it]. ... 1.1:... '1'. I! f... .Y.... 1,):. I. 9. I. a. .. 4.. 1.1!.9 . 3‘99... ‘ ( 2.1)..- f o 1.6... .v i . .lv .,~ .n. i; ..., -3. . . . .I .. Egifiwwfifif .13.»... & mi; .3... axis x... $.52? as. . - . .. .Jd. .n 7.9.. . iv. . .v. . . .v... v.1: ., 3:13.? 3'1? . 2.... ..T. .1 .1417..- MICHIGAN STATE UNIVERSITY Ll RABIES l HHHH mmWW WIN/1W . l 3 1293 00910 9988 I! This is to certify that the dissertation entitled LINEAR RESPONSE FUNCTIONS OF AN INTERACTING FERMI GAS AT T = 0 presented by Raj asinghe Nimalakirthi has been accepted towards fulfillment of the requirements for Ph . D. degree in Chemistry fat: 0.4% Major professor Date March 302 1993 MS U i: an Affirmative Action/Equal Opportunity Institution 0-12771 gr LIBRARY Michigan State ; University PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. DATE DUE DATE DUE DATE DUE l J i ll [J L m 7] MSU Is An Affirmative AetiorVEqueI Opportunity Institution swan”! LINEAR RESPONSE FUNCTIONS OF AN INTERACTING FERMI GAS AT T=O By Rajasinghe Nimalakirthi A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1993 ABSTRACT LINEAR RESPONSE FUNCTIONS OF AN INTERACTING FERMI GAS ATT=O By R. Nimalakirthi The response of an interacting Fermi gas in its ground state to weak external electromagnetic fields is analyzed in this thesis. The system’s response to arbitrary scalar and vector potentials has been Studied within the random phase approximation. The electrical response is characterized via the nonlocal polarizability density denoted by a(r, r’; to), which gives the polarization P(r, (0) induced at point r in a system by a perturbing electric field E(r’, to) acting at the point r’, within linear response. A homogeneous electron gas at zero temperature is selected as a well characterized system, for the purpose of determining the nonlocal polarizability density and thus gaining information about the nature and functional form of a(r, r’; to). The longitudinal component (in It space) of the nonlocal polarizability density a(r, r’; to) is connected to the dielectric function £(k,m), and this connection is used to obtain results at two levels of approximation. Results from the Thomas-Fermi ('I'F) form and the random phase approximation (RPA) for 8(k,tr)) are compared. At TF level, the nonlocal polarizability density is evaluated analytically, while within the RPA asymptotic analytical results are obtained. The RPA results are qualitatively distinct from the TF results, which diverge as Ir - r’ I approaches zero. Within the RPA, there are two long-range components in OLO‘, r’; O): the first is a monotonically decreasing component that arises from the net charge screening in the electron gas, and varies as I r - r’ I’3. The second is an oscillatory component with terms of order Ir — r’ I’“ (n 2 3). The latter is associated with Friedel oscillations in the electron density as found in Langer and Vosoko’s study of the screening of an impurity charge. The results indicate the possibility of long-range, intramolecular terms in the nonlocal polarizability densities of individual molecules. For molecular systems, it is shown that the change in nonlocal polarizability density due to an infinitesimal shift in nuclear position is determined by the hyperpolarizability density. The same hyperpolarizability density describes the electronic charge distribution’s nonlinear response to external fields. A method is provided to obtain the asymptotic form of the dynamic charge density susceptibility, as a function of space and time variables, for a homogeneous electron gas treated within the RPA. It is shown that the calculation reduces to a single quadrature over frequency. Explicit expressions for current density susceptibilities as a function of transferred momentum and frequency are obtained within the RPA in a gauge in which the scalar potential is zero. TO MY PARENTS iv ACKNOWLEDGMENTS My sincere thanks go to Professor Katharine C. Hunt for her guidance, encouragement and support throughout this Study. I would also like to thank to my committee members Professor R. I. Cukier, Professor S. D. Mahanti and Professor P. M. Duxbury for their valuable suggestions. TABLE OF CONTENTS LIST OF FIGURES Chapter 1. 2. 5” INTRODUCTION THE GENERAL THEORY OF LINEAR RESPONSE a. Quantum Mechanical Response Theory b. Double-Time Green Functions and the Time Correlation Functions COLLECTIVE OSCILLATIONS AND THE RANDOM PHASE APPROXIMATION a. Equivalence of Self-Consistent Field Approach and Collective Approach based on RPA THE LINEAR RESPONSE TO SCALAR AND VECTOR POTENTIALS IN THE RPA CHARGE DENSITY SUSCEPTIBILITY a. Relation between the Nonlocal Polarizability Density and the Dielectric Constant b. Asymptotic results for the Nonlocal Polarizability Density in the RPA c. Fourier transform of Dynamic Charge Density Susceptibility CURRENT DENSITY SUSCEPTIBILITY a. Transverse Current Density Susceptibility b. Longitudinal Current Density Susceptibility viii Page 11 15 19 23 33 34 38 49 50 52 7. CHANGES IN ELECTRONIC POLARIZABILI'TY DENSI'TIES DUE TO SHIFTS IN NUCLEAR POSITION 8. SUMMARY AND CONCLUSIONS APPENDIX A APPENDIX B REFERENCES 54 61 63 68 69 LIST OF FIGURES Figure 1. Asymptotic form of the nonlocal polarizability density in the random phase approximation, from Eq. (23). The figure shows “RP A a (21:)3 3/1: kF'3 103 aLzz(r, r’; 0) as a function of kFx a kF (r - r’)x and sz a kF (r - r’)z, with rs = 4. The range of plotted am, A values is [-2.0, 2.0]. Friedel oscillations are evident in the plot. Figure A. 1. Contour in the complex k plane, used for evaluation of the nonlocal polarizability density in the RPA. Branch cuts are shown as striped lines; the symbol x marks the branch cut origins at the points i2kpiin, apole at k=0, and apole neari kTF' viii 1. INTRODUCTION Within linear response theory [1, 2], the properties of a perturbed system are expressed in terms of its equilibrium properties. References 1 and 2 treat the response of a system to an external force by a simple perturbation method, assuming that the external perturbation can be expressed as an additional term in the Hamiltonian and that it does not drive the system far from equilibrium. Kubo [2] has shown that the complex susceptibility and complex conductivity, and hence the electric polarization and induced currents, can be rigorously expressed in terms of equilibrium time correlation functions of the associated dynamical variables. For example, it was shown by Kubo that the conductivity tensor for a given frequency of the applied field can be rigorously expressed in terms of electric current components fluctuating spontaneously in the equilibrium state. There has been considerable interest in evaluating transport coefficients given by Kubo’s formulas. One of the important tools of quantum field theory is the Green function, which is convenient for the study of the properties of interacting quantized fields. These tools turn out to be useful also in statistical mechanics, in cases where one can sum the same type of perturbation theory diagrams. The application of Green functions is fruitful in the quantum theory of fields when combined with spectral representations. Spectral representations for the time-correlation functions and for retarded Green functions were first established and used in statistical mechanics in the theory of fluctuations and in the statistical mechanics of irreversible processes, beginning with a paper by Callen and Welton [1]. Retarded and advanced Green functions, and their simplest applications to the theory of irreversible processes are discussed at greater length in Zubarev’s paper [4]. They are very convenient for application in statistical mechanics as they can be analytically continued in the complex plane. Linear response functions are most simply expressed in terms of retarded double time Green functions. Because of the long range of the Coulomb force, the interactions in a collection of electrons involve many particles simultaneously. It is well known that an electron gas of high density can undergo organized oscillations resembling sound waves. These oscillations, the so—called “plasma oscillations” represent the effect of the long-range correlations of electron positions brought about by Coulomb interactions. A description in terms of these organized oscillations therefore provides a natural way of heating the long-range electron interactions, and leads to greater insight into the dynamical behavior of the electron gas. We can distinguish between two kinds of response of the electrons to the field. In one of these, the phase difference between the particle response and the field producing it is independent of the position of the particle. This is the response which contributes to the organized behavior of the system. In the other the phase difference between the field and the response depends on the position of the particle. Because of the general random location of the particles, this second response tends to average out to zero when we consider a large number of electrons, and we can neglect the contribution arising from this. This procedure is the “random phase approximation” (RPA). The collective description is similar to a complete perturbation theory treatment in that the perturbation is applied to the collective particle motion. The behavior of the electrons in a dense electron gas can be analyzed in terms of their density fluctuations. These density fluctuations may be split into two components. One component is associated with the organized oscillations of the system as a whole. The other is associated with the random thermal motion of the individual electrons and shows no collective behavior. This split up of the density fluctuations corresponds to an effective separation of the Coulomb interaction into long-range and short-range parts; the separation occurs at roughly the Debye length. Bohm and Pines [5] used the above split-up of the density fluctuations to study the collective response of the electron gas to the field of an individual charged particle moving with a specified velocity v0. When Va is less than the mean thermal speed of the gas, they found that the collective response is just such as to screen out the field of the specified particle within a distance of order of the Debye length. The self consistent field (SCF) method, in which a many-electron system is described by a time-dependent interaction of a single electron with a self-consistent electromagnetic field, has been shown to be equivalent to the dielectric approach of Noziéres and Pines [9] to the many electron problem, in work by Ehrenreich and Cohen [10]. Ehrenreich and Cohen’s paper establishes the relationship among the equation of motion approach, the quantum kinetic equations, the calculation of the dielectric function, and Landau’s Fermi liquid theory. The simplicity and ease of interpretation of the SCF method commends it for problems such as the calculation of the dielectric constant and the response of the system to a general external perturbation. By studying the system’s response to an arbitrary scalar potential and a vector potential we can evaluate the induced particle density and current density respectively. By using the facts that the transverse vector potential excites only a transverse current and is not screened by the longitudinal Coulomb interaction we separate the vector equation into its transverse and longitudinal parts. By using the equation of continuity, we obtain a result for the longitudinal current self-consistently. A knowledge of the RPA dielectric constant at arbitrary frequency enables one to calculate a number of interesting properties of the electron gas. It also predicts correctly a number of properties of the electron gas such as the plasmon frequency. In the RPA, since Im §RPA(q,co) vanishes for the plasmon modes, there will be no damping of the plasmons. The plasmon frequency is determined by the dispersion relation Re §RPA(q,o)) = 0 [11]. At small values of q, Im §RPA(q,to) is proportional to to at small values of to. The proportionality to to is an important feature of Im §RPA(q,to). The long-wavelength plasmon will now be damped. The linear dependence of Im gRPA(q,(o) on to must also occur for the exact dielectric function. The physical property under consideration is the rate at which electron—hole pairs are made in the electron gas [12]. A hole is a state which has an electron removed from the filled Fermi sea. An initial electron of momentum p and energy SP is excited by a perturbation with (q,(o). Thus it is excited to a new state with momentum p+ q and energy £P+q = ep+ to. The electron can only be scattered into states which are previously unoccupied, so 8 must be above the occupied Fermi sea. Thus the basic process P“! takes an electron from below to above the Fermi level. Thus the excitation processes make electron-hole pairs. By simple arguments, it can be shown that the rate of making electron-hole pairs is proportional to a), at small values of to. At small values of a) we can expect damping to be small. The calculation of the induced screening charge density (about an impurity) in the RPA has been carried out by Langer and Vosko [13]. In contrast to the Thomas-Fermi calculation, the screening density at the origin is finite in the RPA. In the RPA, the screening density does not go to zero exponentially, but rather oscillates at large values of r. The second feature had appeared in earlier calculations of Friedel [15]. These oscillations come about as a result of a logarithmic singularity in §RPA(q,0). At q = 2P1: its first derivative becomes infinite. The physical origin of the singularity is not difficult to trace. When q < 2pF, one may excite an electron from one part of the Fermi surface to the another; the electron hole excitation spectrum begins at zero energy. When q > 2pF, one must instead supply energy to excite an electron; the corresponding excitation spectrum begins at some finite energy. Friedel oscillations are an exact microscopic property of normal electron liquids; they may be considered as a direct reflection of the sharpness of the Fermi surface [11, 16]. This thesis focuses on susceptibilities of the electron gas and molecular systems, specifically the charge-density susceptibility, the current-density susceptibility, the nonlocal polarizability density, and the hyperpolarizability density. The nonlocal polarizability density is a linear response tensor that gives the polarization P(r, or) induced at a point r in a molecule by a perturbing electric field E(r’, to) that acts at the point r’ [18-23]: The nonlocal polarizability density was introduced by Maaskant and Oosterhoff in a study of optical rotation in condensed media [18]. In their work, o(r, r’; or) is expressed as a difference of two components that are individually divergent as (u —-) O, and each transition matrix element in the expression involves an operator specified only as an infinite series [8]. For response to fields that are derivable from a scalar potential, a(r, r’; to) can be recast in a computationally tractable form, via a connection [22] between its spatial Fourier transform and that of the charge- density susceptibility a(r, r’; to) [24-29]. The function a(r, r’; 0)) determines the change in charge density at point r in response to a perturbing scalar potential acting at r’, within linear response. This connection gives the longitudinal component of the polarizability density aL(r, r’; (o) [22]. In chapter 2, we briefly review the general theory of linear response. Collective oscillations and the RPA are discussed in chapter 3. Chapter 4 contains a discussion of the linear response to scalar and vector potentials acting as perturbations. There we obtain the transverse and longitudinal current density susceptibilities within the RPA in gauge- nvariant form. It is shown that the longitudinal current density is screened within the RPA. The new results of this thesis are contained in chapters 5, 6, and 7. In chapter 5 we give a generalization of the Langer and Vosko result by evaluating the charge density susceptibility and the nonlocal polarizability density, for a scalar potential of arbitrary spatial variation. We establish a connection between the static, longitudinal component of the nonlocal polarizability density in position space and the dielectric function E(k, 0), and then use the connection to obtain results at two levels of approximation to 80‘, 0): we compare the results from the Thomas-Fermi (TF) [11,12] and the RPA approximation. Within the TF approximation, we obtain analytical results, while within the RPA, we obtain asymptotic analytical results. In chapter 5 we also describe the extension of the approach used in previous sections to evaluate the dynamic charge-density susceptibility. We develop a method to evaluate the asymptotic form of dynamic charge density susceptibility within the RPA. In chapter 6 we study the response to a field of arbitrary polarization. It is most convenient to choose a gauge such that the vector potential alone represents the applied electromagnetic fields, and study the induced current. A given longitudinal electric field produces the same current, whether it is by a vector potential or by a scalar potential; thus the induced polarization current is gauge invariant. Explicit expressions for transverse and longitudinal current density susceptibilities are obtained in (q,m) space within the RPA. In chapter 7, we develop a general result for another linear response tensor, the polarizability density. The results are obtained for an isolated molecule in its ground electronic state. We prove by direct perturbation theory that when a nucleus in a molecule shifts infinitesimally, the resulting change in polarizability density is determined by the same hyperpolarizability density that fixes the response to external fields. This is a general quantum mechanical result. It is rather straightforward, but previously unanticipated. It has applications in the analysis of interpretation of intensities of vibrational Raman bands. Chapter 8 contains a brief summary and conclusions. Mathematical identities needed in the calculations are presented in two appendices. 2. THE GENERAL THEORY OF LINEAR RESPONSE The formalism of linear response dreary is straightforward. We consider a system that in the distant past (t—>-oo) was prepared in a State of equilibrium. Let us suppose that in the distant past an external force was switched on adiabatically. As a result, the system is no longer in stable equilibrium. On the other hand, internal affairs in the system should not be much affected by a small force, and the lowest- order response that moves the system to a new stable state compatible with the imposed constraints can be connected to local equilibrium fluctuations. We must, first of all, estimate when an external force can be considered small. In the most general case, we can argue that the energy transmitted to a particle by the external force over a characteristic distance in the system has to be small compared to its average energy in local equilibrium [la]. This criterion is, in fact, very conservative in a situation where the external force is switched on adiabatically. “Adiabatically” here means slow on the time scale of the regression of local fluctuations. In this case the system only has to adjust at any instant of time by an infinitesimal amount to the new external constraints. If we follow the time evolution of the system via the statistical operator, this implies that first-order, time-dependent perturbation theory should be adequate [1a]. Such an approach leads to constitutive relations that are linear in the applied external forces [1a]. It has therefore received the name “linear response theory”. The use of first-order time dependent perturbation theory in the study of transport phenomena in many-body systems was advocated in an early paper by Callen and Welton [ I] . 8 The following review is patterned on a paper of Kubo [2], where a rather general framework was laid out for the calculation of the systems response to mechanical forces within a Hamiltonian formalism [1a]. Linear response theory makes an assumption that the applied external force must be such that it can be incorporated as part of the total Hamiltonian of the system. This, of course, means that the results of linear response dreary are as general as can be, and are, as with most general results in physics, not directly useful for practical calculations [la]. However, linear response theory is eminently suited to prove general features of nonequilibrium physics, such as positivity of transport coefficients, validity of the Onsager reciprocity [3] relations, sum rules, and dispersion relations. To do practical calculations based on linear response theory, one can either employ equilibrium Greens function techniques, use kinetic equations, or resort to semi-empirical models with experimental data as partial input. The latter approach is often rewarding since linear response theory expresses transport coefficients in terms of time-dependent correlation functions, which are quite often available fiom experiments [1a]. However in the work presented in this thesis, the former method is followed. a. Quantum Mechanical Resmnse Theog Let us consider an isolated system, the Hamiltonian of which is denoted by H. The dynamical motion of the system determined by H is called the ‘ ‘natural motion” of the system. We suppose that an external field F(t) is applied to the system, the effect of which is represented by the perturbation term, 10 H’ (t) = -AF(t) (2.1) The motion of the system is perturbed by the external force, but the perturbation is small if the force is weak. The response is observed through the change AB(t) of a physical quantity B. The problem is now to express AB(t) in terms of the natural motion of the system. The initial ensemble which represents statistically the initial state of the system is Specified by the density matrix p satisfying [H, p] = O [2]. The motion of the ensemble under the perturbation (2.1) is represented by p’(t), which obeys the equation, BOYD/at = 1/ il'f [H + H'O). MD] (2.2) With the initial condition p’(-oo ) = p, we expand p’(t) as p’(t) = p + A 90) (2.3) In the linear approximation, the solution for Ap(t) is [2] Ap(t) = -1/iIi IL“ exp (-i(t-t’)H/it) [A, p] exp(i(t—t')H/lt) F(t’)dt’. (2.4) It is convenient to introduce the operator ax operating on another operator b by the following definition x a b = [a, b] , (2.5) for which one finds the rule [2] x a ~a ea b=e be . (2'6) With this notation, equations (2.2) and (2.4) can be written as [2] ante/aw lfifi (HX+H"‘(t» #0) (2.7) 11 Ap(t) = -1/m It,m exp (-i(t-t’)HX/ll) [A, p] F(t') dt’. (2.8) The response AB(t) of the quantity B is statistically [2] AB(t) = Tr Ap(t)B = 4/15 Tr IL” [A, p] B(t—t’) F(t')dt’ (2.9) where B(t) is the Heisenberg representation of B following the equation dB(t)/dt = 1/il'r[B(t), H] . (2.10) The response function is now [2] vBAa) =4th Tr IA. 9er = llih Tr p [A, B(t)]. (2.11) The above Kubo formula [2] is valid at finite temperatures, because a trace over the thermal distribution is taken. At zero temperature this is equivalent to taking the expectation value in the ground state. b. Double-Time Green Functions and Time Correl_ation Function_s The Green functions in statistical mechanics are the appropriate generalization of the concept of correlation functions. They are just as intimately connected with the evaluation of observed quantifies, and they have well-known advantages when equations are formulated and solved [4]. Following Zubarev [4] we define double-time retarded and advanced Green functions Gr(t, t’) and Ga(t, t’) as follows: Gr(t, o = -i 9(t-t’)<[A(t). B (2°12) Ga(t, t') = +i 9(t'-t)<[A(t), B(t')]> (2.13) 12 where <.....> indicates that one should average over a thermal ensemble at finite temperature, and at zero temperature it simply means averaging over the Heisenberg ground state of the interacting system. A(t) and B(t) are the Heisenberg representation of the operators A and B, expressed in terms of a product of particle creation and annihilation operators (or a product of quantized field functions). The l t > 0 theta function is defined by 9(1) = O K o (For convenience, a system of units in which h=1 is used from here onward.) Finally, [A, B] indicates the commutator or anti-commutator: -l Fermion [A,B]=AB-nBA,n= +1 Boson The sign of t] is determined by the problem. Generally speaking, A and B are neither Bose nor Fermi operators [4]. When the time arguments are the same, t = t’, the Green functions are not defined, because of the discontinuous factor 9(t-t’). We note that in the case of statistical equilibrium the Green functions Gr(t, t') and Ga(t, t’) depend on t and t’ only through (t-t’) [4]. For the time being we have introduced the Green functions purely formally, by analogy with the quantum theory of fields. We shall satisfy ourselves now by concrete examples that they are very conveniently applied in quantum statistics to problems concerning a system of a large number of interacting particles. One can choose for A and B operators of different kinds: for instance, Fermi or Bose operators and their products, density operators or current operators. The choice of the operators A and B 13 is determined by the conditions of the problem [4]. The time correlation funtions used in statistical mechanics are averages over the statistical ensemble of the product of operators in the Heisenberg representation of the kind (pBAa, t’) = and (pAB(t, t’) = (2.14) In the case of statistical equilibrium the time correlation functions depend, as do the Green functions, only on t - t’. But they are defined also when the times are the same, t = t’. In quantum mechanics we need to take the symmetrized product of operators. In subsequent work we restrict our attention to the zero temperature problem, and we are interested in two types of external perturbations, namely those due to scalar potentials and those due to vector potentials. We can write down the perturbing Hamiltonians for electrons exposed to a scalar potential perturbation or a vector potential perturbation respectively as H’(t) = -e 1.13: p(r,t) q)(r,t) (2.15) H’(t) = —l/c 1:13: J(r,t) - A(r,t). (2.16) The induced particle density and paramagnetic term in the induced current density in the linear approximation are 5< p(r.t)> = ie IL... dt’1d3r’ <[p(r.t). pen] > «m (2.17) 8< J(r.t)> = i/c IL... dt’ ld3r’ <[J(r.t). J(r’.t’)]> - A(I’.t’) (2-18) Now in analogy with Eq. (1.11) we define retarded density and current correlation functions as l4 a(r t, r’ t') = -i 9(t-t') <[p(r,t), p(r’,t')] > (2.19) m t. r’ t’) = -i 0(t-t’) <[J(r.t). J(r’.t’)]>. (2.20) Then Eq. (2.17) and Eq. (2.18) may be written as 5< p(r,t)> = -e I+°°dr I d3r’ a(r t, r’ t') (p(r',t’) (2.21) 8 = -l/c1+°°dt’1d3r’ x(r t, 1’ t') - A(r’,t'). (2.22) where causal behavior is enforced by the retarded nature of or and x respectively. Eq. (2.21) and Eq. (2.22) typify a general result that the linear response of an operator to an external perturbation is expressible as the space-time integral of a suitable retarded correlation function. [ Note: Eq. (2.18) does not give us the total current, as there is an additional term proportional to -e2/mc A(r,t); see reference 11, Eq. (52.12). ] In subsequent work we wish to evaluate a and x for a specific model system, namely the interacting electron gas with a uniform and exactly opposite charged back- ground, so that the whole system is neutral. 3. COLLECTIVE OSCILLATIONS AND THE RANDOM PHASE APPROXIMATION In this section we give a short review of Pines and Bohm [5] theory of plasma oscillations, and wish to develop a detailed physical picture of the behavior of the electrons in a dense electron gas. We are concerned with the organization produced by the Coulomb interactions. From the work of Pines and Bohm: ‘ ‘In a dense electron gas, the particles interact strongly because of the long range of the Coulomb force; in fact each particle interacts simultaneously with all the other particles. As a result the equations of motion become extremely difficult to solve. The usual perturbation based on the assumption of interaction between particles breaks down.... A collective description provides a far better starting point for a solution than a description in terms of individual particles [5]”. ‘ ‘Instead of following the motion of the individual particles, we describe the gas in terms of the Fourier components of the electron density at each point in space. These Fourier components are proportional to the density fluctuations in the electron gas. We find that the density fluctuations can be split into two parts. One part represents an organized oscillation with the characteristic plasma frequency, and is clearly associated with the collective behavior of the system. The other part is associated with the motion of the individual particles... For wavelengths greater than a certain critical length 1D (the Debye length). the fluctuations are primarily collective... For wavelengths smaller than AD, however, the fluctuations are primarily associated with individual particle motion [5]”. “As any electron moves through the assembly, the other electrons are pushed 15 16 away from it by the Coulomb repulsion. Each particle is thus surrounded by a cloud of extent "Dv in which there is a deficiency of electrons, which is responsible for screening the field of the particle in question [5] ”. ‘ ‘The splitup of the density fluctuation into collective and individual-particle components may be viewed in the following way. The collective part includes the effects of the long range of the Coulomb force which lead to the simultaneous interactions of many particles. The individual-particle component represents the density fluctuations arising from the randomly moving individual particles plus their, comoving electron clouds, and thus includes the effects of the residual short-range screened Coulomb force [5] ”. ‘ ‘Certain examples of collective behavior in an electron gas are well known from the study of gaseous discharges. These are the organized oscillations of the system as a whole, the plasma oscillations. The ions in a metal are also susceptible to a collective description, and, in interaction with electrons, they give rise to sound waves, whose properties canbe calculated with the collective method. In this way, one can obtain an improved treatment of the so-called lattice-electron interaction, which is important in the dreary of electrical conductivity [5]”. “We begin by a study of the way in which the interactions in an assembly of electrons bring about organized behavior and collective oscillations. We shall consider an aggregate of approximately free electrons embedded in a medium of fixed positive charges whose average density is equal to that of the electrons [5] ”. For the purpose of these calculations this distribution of charge can be regarded as uniformly smeared out, throughout the entire system. Hence, it merely serves to neutralize the 17 net electron charge. Each electron in the assembly is acted on by the sum of the forces arising from all of the other electrons plus that resulting from the smeared-out positive charge [5]. The potential energy of interaction between the ith and jth electrons, e2] 1 xi - xj l , may be expanded as a Fourier series in a box of unit volume with periodic boundary conditions, which gives [5] e2/lxi - le =4min1/1c2)e.xp[irr(xi . x9]. (3.1) k The equation of motion of the i‘h electron is given by [5] dzxi ldt2=-(41te2i/m) Z’(k/k2)CXP[ik'(Xi-xj)l (3.2) is where the prime denotes a sum in which It = 0 is excluded. (The term with k = 0 takes into account the uniform background of positive charge, and hence the overall charge neutrality of the system.) ‘ ‘The range of the Coulomb potential is so great drat many-body collisions are important... Under drese conditions, the electrons move together in organized fashion, and one finds the well-known phenomenon of “plasma” oscillations of the system as a whole [5]. The particle density in our box of unit volume is given by p(x)=):.6|p+q> + 1 n(p + q) - n(p) 1V(q.t). (3.15) 21 where (pl VI p+q) = V(q,t). The potential V consists of an external potential V0 plus the screening potential Vs, which is related to the induced change in electron density [10]. An = Tr {5(rrc - x)p(1)} = Zq exp -iq-x 2p, (p’l p(1)| p’+q) (3.16) by Poisson’s equation: V2 Vs = -41t An e2. (3-17) Here 8(x’e - x) is the charge density operator, xe being the position operator and x referring to a specific point in space [10]. We drus find [10] mm = qup. . (3.18) where Vq = 41te2/q2. By substituting the above expression giving V8 for V in Eq. (3.15), we obtain the Liouville-Poisson equation determining (pl pm I p-t-q) in the absence of an external perturbation: i 8(pl p“)|p+q)/8t=(ep-8M) (plpmlrmr) + v.1 :10) + q) - n(p) 1 2,. (p'l p“) I p’+q>. (3.19) ‘ ‘In solving problems by the SCF method, however, one can usually avoid the explicit expression of V8 in terms of pm widrin the equation of motion by means of an assumption about the time dependence of V(q,t). To illustrate this point, we calculate the frequency and wave-number dependence of the longitudinal dielectric constant §(q,to). We imagine drat the external potential V0(q,t) acts on the system with time dependence exp (itat+nt), where t’)-90* corresponds to an adiabatic turning on of the perturbation. This potential polarizes the system [10]”. From 22 electrodynamics we have the relation P(q. t) = (U41!) [£01.00 - 1] E(q,t). (3.20) The polarization P(q,t) is related to the induced change in electron density by V-P = eAn or -iqP(q.t) = eAn(q.t) (3.21) and the electric field E(q,t) is given by eE(q,t) = -qu(q,t). (3.22) “Equation (3.15) is readily solved for (pl pa” p-t-q) by assuming that (p 1 pa” p+q) and Vs(q,t) have the same time dependence as V0(q,t). The induced change in electron density An(q,t) may then be calculated from Eq. (3.16) and §(q,o)) deduced from the field equations (3.20, 21, 22). We find [10]” §(q.m)=1-11m VQX n(q.t) (4.2) Hv ’(t) = —l/c):.‘l J(-q,t ) - A(q,t) (4.3) where the particle density operator and the current density operator have the forms p(q.t) = 2 p c“ M0) cp(t) (4.4) J(q.t)=(-e/m)2,(p+ q/2)c+,.q(t)c, — (8pm ep) + [ S(p+ q) £(p)] M Cr) + [ n(p) - n(p + q) ltvq [1-G(q)} - e (P(q,t) + (e/mC) (p + q/Z) ° A(q.t)]. (4.15) and in the case of the current: <[I-I, (-e/m)(p + q/Z) C+P+q cp]> = (elmI - 8p) (~e/m)(p + q/2) + [3(1) + q) - E(p)] (~e/m)(p + qfl) <61”+q C]? + [ n(p) - n(p + q) 11vq (-e/m)(p + q/2)1 10(q)} -1/c (elm)2(p + mm: + an) . A(qn + (elm) e (P + q/Z) (P(q,t)l . (4.16) where E(p) = {fivq n(p + q) is the exchange energy and the Hubbard approximation as defined in Ref. (12) has been used to introduce the function G(q). Since the density and current respectively satisfy the equations of motion: i d< 0+9“! cp>ldt = co (4.17) id(-e/m)(p + q/2) (C+ C >/dt = oX-e/m)(p + q/2) p+q°p = I n(p) - n(p+q) ] [vq{1-G(q)} - 8 Wm) + (e/mc) (p + 11/2) - A(q.t) ] (4. l9) and 28 {to + 89- ‘m’ 8(1)) - £(p+q)l (-e/m)(p + (In) = [n(p) - n(p+q)] [vq(-e/m)(p + q/2)1 1-G(q)} +e(e/m)

by p(q,(o) exp(-iort) and <1 (q,t)> by J (q,to) exp(-iaJt) to obtain equations for the induced particle density and current density respectively: p(q.m) = E(q,to) vqi l-G(q)}p(q,co) - e E(qm) 2kF because one is passing into a region in which the momentum transfer can no longer take an electron from one part of the Fermi surface to anodrer [11]. This disturbance is therefore effective at rather large distances. Since the polarization of the system satisfies the equation, P(MD) = (1/41t)[8(k.(0) - l] ' E(k.(0). (5.1) the charge density susceptibility is intimately related to the complex polarizability. In the subsequent sections of this chapter we evaluate the charge density susceptibility and the nonlocal polarizability density of the interacting electron gas as linear responses to an arbitrary perturbing scalar potential. 33 34 a. Relation between the nonlocal mlarizabilig densig and the dielectric constant I l7| The nonlocal polarizability density is a linear response tensor that gives the polarization P(r, (1)) induced at a point r in a molecule by a perturbing electric field E(r’, (1)) that acts at the point 1" [18-23]: P(r, m) = I dr’ a(r, r’; to) - E(r’, m) . (5.2) Thus a(r, r’; (1)) represents the distribution of polarizability widrin a molecule or an extended quantum mechanical system. The purpose of this work is to gain information about the nature and functional form of the static polarizability density, by determining this property for a homogeneous electron gas at T = 0. The induced polarization P(r, to) in Eq. (5.7) is related to the change in charge density 8p(r, (a) due to E(r’, (a) by [22] V - P(r, co) = - 6p(r, to) , (53) so a(r, r’; or) provides a complete description of the response to perturbing fields that are derivable from scalar potentials. The polarization can be interpreted as the density of the induced dipole moment, but it also includes quadrupolar and higher-order response [23]. The nonlocal polarizability density was introduced by Maaskant and Oosterhoff in a study of optical rotation in condensed media [18]. In their work, a(r, r’; to) is expressed as a difference of two components that are individually divergent as a) —-) 0, and each transition matrix element in the expression involves an operator specified only as an infinite series [18]. For response to fields that are derivable from a scalar potential, a(r, r’; (a) can be recast in a computationally tractable form, via a connection [22] between its spatial Fourier transform and that of the charge-density susceptibility a(r, r’; co) [24-29]. 35 The function a(r, r’; (1)) determines the change in charge density at point r in response to a perturbing scalar potential acting at r’, within linear response. This connection gives the longitudinal component of the polarizability density aL(r, r’; or) [22]. For small molecules, the spatial Fourier transforms of a(r, r"; or) (and drerefore GLGt, k’;a))) have been computed via pseudo-state methods [28], a stationary principle [25], and an Unsold-type approximation [27; see also Refs. 26 and 29]. The point—atom- polarizability approximation (PAPA) [20, 30], truncated at the dipole terms, is equivalent to approximating a(r, r’; (a) as a weighted sum of products of Dirac delta functions located at the nuclei. Modifications of the PAPA model to allow for distributed point multipoles [30, 31] can improve its approximation to the continuous polarizability distribution of the actual function 60', r’; or) and drus extend the range of validity of the PAPA model. In this section (cf. [17]), we analyze the static, longitudinal polarizability density aL(r, r’; a) = 0), for a homogeneous electron gas at zero temperature. Our intent is to gain a better understanding of o(r, r’; a) = 0), and to provide information for the development of approximations to molecular polarizability densities. In this section, we derive an equation for the longitudinal component of the nonlocal polarizability density aLaB(r, r’; 0) via its connection to the charge-density susceptibility 01(r, r’; 0). In wave-vector space [22], 6L(k, k’; or) = k k’ n(k, 11'; to) / (k k’) , (5.4) where k and k’ are unit vectors in the directions k and lt’, and a(k, k’; to) = 1dr Idr' a(r, r’; to) e" i" - ' elk" " . (5.5) In sum-over-states form, the charge-density susceptibility satisfies [24] 36 01(k.k’;m)= 2/112' (snow |p(-k) ln)(n | p(k’) IO)/(arn02-to2). (5.6) n In Eq. (5.6), the sum runs over the excited states 11, (Ono denotes the transition frequency (E,1 — Eo)/h, p(k) is the Fourier transform of the electronic charge density operator, 90‘) = - e 2 exp (i k - rj) , (5.7) .1 and rj is the coordinate for the jth electron. Because of the translational invariance of the homogeneous electron gas, a(r, r’; 0)) depends only on r - r’, and therefore a(k, k’; or) takes the form n(k, k’; or) = (211)36 (k, to) so" - k) . (5.8) The susceptibility a(k, (1)) gives the change in the k, a) component of the charge density 8(p(k, 01)) due to an external scalar potential 8¢°x(k, (a): 5(p(k. (0)) = a(k. (o) 5¢°"(k. 03) . (5.9) Linear response theory relates the susceptibility a(k, or) to the retarded density-density correlation function DR(k, to) [l 1]. In terms of the function TIR(k, to) defined by 11%, to) = 11-1 13%., (a): 6 = (:2 nR(k.m)6¢“(k.m) . (5.10) In diagrammatic dreary [l l], I'IR(k, (a) is the "retarded polarization part." From Eqs.(5.4), and (5.8)-(5.10), aL(k, k’; or) = (21:)3 e2 k k’ 11R(k, (n) 8(k’ - k) / (k k’) . (5.11) 37 Thus in coordinate space, the Static longitudinal polarizability density is “Lab“, 1"; 0) = (210—3 e2 I dk ei ‘ ' ("'3 ka k5 nR(k, 0) 162 . (5.12) We simplify Eq. (5.12) using [11] nR(k. 0) c2 1-2 = (410-1 [E(k. or‘ —11 . (5.13) and the Rayleigh expansion [32] ei " ' 0"") =2? if (21+ 1) j,(1cx) P[(cos y) , (5.14) [=0 where I r — r’ l= x, j [(kx) is the [th spherical Bessel function, and yis the angle between k and (r - r’). In terms of the orientation angles 01., (pk for the vector k and the angles 0, (p for (r — r’), the addition theorem for spherical harmonics gives [33] I P[(cosy) = 41: (21+ 1)‘12 Y,m(ek, (pk) Y,m*(e, (p) . (5.15) m=-[ For an isotropic, homogeneous electron gas, after specializing the result from Eqs. (5.12)-(15) to eLnu, r’; 0), we obtain 0Lu(r. r'; 0) = 1/3 (2n)'3lo dk 1.2 joacx) 180:. 0)"1 - 11 — 213 (210-310 dk k2 j2(kx) [e(k, 0)"l - 1] P2(cos 9) , (5. 16) where j0(z) = 2’1 sin z and j2(z) = (3 z'3 - 2") sin 2 — 3 2‘2 cos 2. The analytical results in Sec. g are based on Eq. (5.16) for aLzz(r, r"; 0). Other components of the polarizability tensor may be computed analogously. 38 b. Asymptotic results for the ngnlgal mlarizabilig densig in the RPA | l7| In the random phase approximation (RPA) [5, 10], the dielectric function is [8,11,12,16] E(k, 0) = 1 + 4 (:1 rs 1.1.2 (rtk2)‘1 g(k/kF) . (5.17) In Eq. (5.17), at = (4/911:)1/3. The density of the electron gas is characterized by the dimensionless ratio rs = r0 [80, where a0 is the Bohr radius and the volume per particle V/N satisfies V/N = 411: I3 r03. The wavenumber kF of the highest occupied state in the electron gas at T = 0 is given by It]; = l/(aro); kF is designated the Fermi wavenumber [l l]. The function g(z) is [8, 11, 12, 16] g(z) = 1/2 - (22)'1 (1- 2%) In I (1 - 212) / (1 + 212) I . (5.18) We find the form of aLzz(r, r’; 0) for large I r — r’ 1 following the medrod of Ref. 11. Since g(z) is an even function of z, QLZZO', r’; 0) = (210-3 (6ix)‘1l_°°dkk eikx [e(k, 0)"l - 1] - 2/3 (210'3 10L, dk k2 ((2i)'1 13 (too-3 — (loo-11 — an (new) eikx [e(k, orl - 1] P2(cos e) , (5.19) where ‘0 denotes the principal value of the integral (see the Appendix). After rewriting g(k/kF) in the form [1 l] g(k/kp) = 1/2 - (2 k n(p)-1 (k1.2 — 1.2/4) xrimn_,0+1/21n{[(k-sz)2+n2]/[(k+2kF)2+n2]} ,(5.20) 39 we evaluate the integrals in Eq. (5.19) by complex contour integration around the contour shown in Appendix A. In the upper half plane (in k), the integrands have simple poles near ik-I-F where kn; is the Thomas-Fermi wavenumber given by kn: = («It's/101,2 k... In the second integral in Eq. (5.24), the function 3 1:2 (too-3 eikx [E(k, 0)’l - 1] has an additional pole at k = 0. Bodr integrands have branch points (in the function g) at r: 2kF :1: in, with branch cuts originating at these points, as shown in Figure Al in Appendix A, which also gives the details of the calculation. Results for integrals around the branch cuts C(1) and C(2) in the upper half plane (of k) are: 1%) dkk“ eikx [e(k, 0)“1 — 1] = 2M1 1: g kpn‘l x'2 exp (2ikFx) (4 + 1:)“2 x {-i+2[§(C-3/2+ln4kFx)-8+(2n+3)(1+§/4)]/[kFx(4+§)]+... } (5.21) and Ice) dk k" eikx [e(k, or1 - 1] = (—1)“+ll 2ml 1: g kph-1 x'2 exp (~21kFx) (4 + §)-2 x {—i-2[§(C-3/2+ln4kFx)-8+(2n+3)(1 +§/4)]/[kpx(4+§)]+ ...} (5.22) where n 2 —l, E, is defined by é a kHz/(2kg) E 2 (1 rs it”, and C is Euler’s constant (C = 0.577216). The result obtained for shun, r’; 0) is 41-220; r’; 0) = (210-3 411/3 x-3 g (4 + §)-2 {cos(2kFx) + 2 (kpxrl (4 + §)-1 x sin(2kFx) [3 - g(c - 1/4) - a 1n(4kpx)] } + (2a)"3 811/3 x’3 t; (4 + 0'2 P2(cos e) {cos(2kFx) - 2 (kFx)'1 g x (4 + §)'1 sin(2kFx) [c + 1/2 + ln(4kFx)] } + (210-3 n x'3 P2(cos e) , (5.23) where x = Ir — r’ I . Fig. l [ 17] shows the asymptotic form of aLzz(r, r’; 0) from Eq. (5.23), with rs = 4. The error due to terms omitted from Eq. (5.23) is of the order (a x—5 + b x‘5 1n kFx). In Eqs. (5.19) and (5.23), aLzz(r, r’; 0) has two components which transform differently under rotations, one as Po and the other as P2. We have not found the integrals needed for the P2 component in any earlier work. However, the integrals needed for the P0 component have previously been used to find the change in the charge density 5(p(x)) induced in an electron gas when a point charge Ze is inserted at an origin a distance x away [11, 12, 13]: 6(p(x))= (210-3 Ze I d3k exp(ik-x) [e(k, 0)’l - 1] = Ze/(4rr2 in]: k eib‘ [E(k, 0)‘l - 1] dk . (5.24) Thus, comparison of asymptotic results for 5(p(x)) serves as a check on the P0 component of Eq. (5.23). Specializing Eqs. (5.21) and (5.22) to n = 1, we obtain 8(p(x)> = 2e 1:“ 2t (4 + of x‘3 {cos(2kpx) - 2 (4 + o‘1 (14.x)-1 x sin(2kFx) [g (c - 1/4) - 3 + g In 4kFx] + } , (5.25) 41 in agreement with the result of Fetter and Walecka [l 1] for large x. We note that Eqs. (5.21) and (5.22) can also be used to obtain the asymptotic form of the charge-density susceptibility 01(r, r’; 0) in the RPA: 01(r, r’; a): (210-3 (zip-IL” k3 eikx [15(k,0)'l - 1] dk = (210-3 161: 1.1.2 x'3 g (4 + §)-2 {cos(2kpx) — 2 (4 + §)-1 (kFx)"l x sin(2kFx) [ g (c + 3/4) + 1 + g In 4kFx] + } , (5.26) with x defined as above. The susceptibility a(r, r"; 0) gives the change in charge density induced by a scalar potential of arbitrary spatial variation. This contrasts with the function 6(p(x)) previously computed in Refs. 11 and 13, since 8(p(x)) is the response specific to a single added point-charge impurity. The asymptotic form of aLu(r, r’; 0) contains oscillatory terms and a non- oscillatory term (the last). Madrematically, the oscillatory, long-range terms in aLu(r, r’; 0), 8(p(x)), and a(r, r’; 0) are produced by integrations along the branch cuts in 801, 0), widrin the RPA. The oscillations in oLn(r, r’; 0) are evident in Figure 1 [17]. The non-oscillatory, long-range (x'3) term in aLzz(r, r’; 0) has a different mathematical and physical origin. It stems from the singularity of the integrand in the P2-component at k = 0, and an identical term is present even at Thomas-Fermi level. In the Thomas-Fermi approximation, the dielectric function is [12] owns, 0) = 1 + RTFZAZ . (5.27) When Eq. (5.27) is substituted into Eq. (5.16) for aLzz(r, r’; 0), the resulting integrands have simple poles at k = 0 and k = a i kTF° Hence by complex contour integration, we 42 I I. . 6 .‘ ... I f O 0 ’0... of" ( t 5”]: I O \\ ‘ \~ Figure l. Asymptotic form of the nonlocal polarizability density in the random phase approximation, from Eq. (23). The figure shows “RP A a (211:)3 3/71: kF_3 103 oLu(r, r'; 0) as a function of kF" a kF (r — r’)x and sz .=. 1‘1: (r - r')z, widr rs = 4. The range of planed oRPA values is [-2.0, 2.0]. Friedel oscillations are evident in the plot. 43 obtain “L'I'FZZO'. 1"; 0) = (210-3 111/3 P2(cos 9) k’I‘F3 x { 3 (rum-3 — exp(-kTFx) [ 3 (k-I-Fx)'3 + 3 (RTFxr2 + awn-1]} — (210-3 11/6 k-I-F3exp(-kTFx) (kTFxYl . (5.28) The long-range (x'3), non-oscillatory component of the polarizability density in Eqs. (5.23) and (5.28) reflects charge screening. For example, when a perturbing point charge + Ze is added to the electron gas, the change in net charge inside a large sphere centered on the perturber approaches — Ze, as the radius R1 of the sphere increases [11, 12]. Equation (5.2) relates the induced polarization P(r) to the induced change in charge density; drerefore as Ri increases, 1 P(r) . dsi = —l 5p(r) dvi —> Ze , (5.29) where the integrals run over the surface dSi and the volume dVi of the sphere of radius Ri° Thus, for sufficiendy large Ri, the polarization on the surface of the sphere must vary as Ri‘z. The non-oscillatory, x’3 terms in the Thomas-Fermi and RPA expressions for aLzz(r, r’; 0) give rise to the required Ri'z component of the polarization. We note that charge is conserved overall for the electron gas, so Eq. (5.11) holds. A surface charge of Ze develops at the outer boundary of the electron-gas system, on a sphere of radius RS. In the calculations, the limit Rs -) oo is taken before the limit R1 -) no. For numerical results based on the RPA and a more accurate dielectric function derived by Vashishta and Singwi [34, 35], see Ref. 17. c. Fourier Transform of gmamic Charge Densig Susceptibilig The dynamic charge-density susceptibility can be evaluated by extension of the approach used in the previous sections. The susceptibility is determined by the following integral over wave vector k and frequency a): a(x, t) = (210-4 (21x)'l L“ dk L” d0.) k3 exp(ikx) exp(—ia)t) [e(k, to)'1 —1], (5.30) where x = Ir - r’ I, t is the time elapsed between perturbation and response points, and E(k, to) is the frequency-dependent dielectric function. Widrin the RPA, the dielectric function is connected to the lowest-order polarization insertion 110(k, to) via E(k, (0) = 1 - U00.) node, 0)), (5.31) with U0(k) = 4ne2/k2. Below, we use scaled variables q and v, defined by q = k/kp and v = mar/(likpz). Then for positive v, Re 11%, v) = map/(41:2 112) {-1 + 1/(2q) [1 — (v/q - q/2)2] x In I [l + (v/q - q/2)][1-(v/q - 2/q)] I - mo 11 - (v/q + q/2121 1n 1 11 + (v/q + q/2)1/11 - (v/q + cm I). (5.32) The imaginary part takes on different functional forms, depending upon the value of q and the relation between q and v: i) Ifq>2, q2/2+q2v2q2/2-q, orifq<2andq+q2/22v2q-q2/z, then 1m 11%. v) = —mk../(4nqfiz)11-(v/q—q/2)21. (5.33) 45 ii) Ifq<2, andOSqu—q2/2, then Im r1°(q, v) = — mka/(li2 21: q). (5.34) ITO(q, v) is even in q, Re I'IO(q, v) is even in v, and Im [10(q, v) is add in v. The integral over k can be found by complex contour integration in the q plane, with the contour closed in the upper half-plane. Poles in the upper half-plane of q (with a finite displacement from the real axis) do not contribute to the asymptotic value of the susceptibility density, because the terms due to such poles decay exponentially with increasing x. There are nonvanishing, asymptotic contributions from poles on the real axis, and from integrations around the branch cuts of the log functions in Re I'IO(q,v). To locate the branch cuts, we rewrite In | [1 +(V/(1 - q/2)]/[l-(VI(1 - (1/2)]| = tunfl _,01/21n 1 1(1+v/q-q/2)2+n21/1(1-v/q +q/2)2+n2]} (5.35) and similarly for the second logarithmic term. There are four equations defining the origins of the branch cuts: a) l + v/q - q/2 = i it] (5.36) b) 1 - (v/q — q/Z) = r: it] (5.37) c) 1 + (v/q + q/2) = :1: it] (5.38) d) 1 — (v/q + q/2) = a: in. (5.39) When v < 1/2, the origins of the branch cuts lie just above and below the real axis, at points with real components given by q(")i = l :t (l + 2v)m, qwi = —l i (l + 2v)m, q(°)2t = —1 i (1 — 2v)1/2, q“):t = l i(1— 2v)1’2. Thus for small but non-zero v, there are eight branch points in the upper half plane. The branch cuts are all taken parallel to the Im q axis, widr those originating in the upper half plane running out to 1m q = + co; cuts originating in the lower half plane run to Im q = -- on. As v approaches 1/‘2, two branch cuts in the upper half plane coalesce, with real components of the origins at Re q = l; and a second pair in the upper half-plane coalesces at Re q = —1. (Similarly, there are two pair- wise coalescences in the lower half-plane). For v > 1/2, drere are four branch-cut origins just above the real axis, four just below the real axis, and two (with vanishing separation as r] —-> 0) in each quadrant of the complex q plane. We run branch cuts between the two branch points in each quadrant that are well off the real axis, and choose the remaining branch cuts as before. Then the integral to be evaluated takes the form I: f(q) dq = — ct f(q) dq - 1C2 f(q) dq - 1C3 f(q) dq - Id f(q) dq + i n 21.. Kn 1 — Ics f(q) dq -- Ice f(q) dq - Icv f(q) dq -— Ice f(q) dq 1. (5.40) asymptotically. Kn denotes the residue at the nth pole on the real axis, and we have neglect- ed contributions from residues at poles with a finite displacement into the upper half-plane, as explained above). The contours Cl-C8 run around the branch cuts (down on the right, up on the left); integrals along the contours Cl-C4 always contribute in the expression above, while integrals along CS-C8 contribute only when v < 13. First, we focus on the integral along C1, widr origin q(a)+ = l + (1 + 2v)1/2: — Cl f(q) dq = (27t)'4 (2 i x)"1 LL” do) I0 du (q(a)++ iu)3 exp[i (q(“)+ + iu) x] x exp(—i(r)t) [eR(q(3)+ + iu, iu)-1 — eL(q(a)+ + in, or)"1], (5.41) where t:R(q(“)+ + in, (0) denotes the value of the dielectric function to the right of the branch cut, and eL(q(a)+ + in, (1)), the value to the left. These differ because of the difference in 47 the phase of the logarithm across the cut. The equation above simplifies furdrer to - Cl f(q) dq = (210-4 (2 i x)_1 1L” d(1) lo du (q<3)++ iu)3 exp[i (11(3)+ + iu) x] x exp(-i(1)t) U081“). + iu) [110R(q(a)+ + iu, v) - 110L(q(a>+ + iu, v)] + [[1 - 0081(1)+ + iu) 110L(q((1>+ + iu, v) ] [ 1 - U0(q(a)+ + in) 110R(q(a>+ + iu, v) 1]. (5.42) Next, we observe that the difference I'IOR(q(“)+ + iu, v) — H0L(q(a)+ + iu, v) is linear in u, to lowest order. We obtain the leading asymptotic term by retaining the u dependence here and in exp[i (q(a)+ + iu) x], but in the remaining factors, we replace q(“)+ + in simply by the value at the origin of the branch cut q(a)+. This gives a contribution from the Cl integral equal to - 1C1 f(q) dq = (211)-4 (2i x)-1 i (88).)3 exp[i q.x 1 Uo(q.) L, (1(1) exp(-i(1)t) {[1 - U0(q(a)+) n0(q(a)+, v) ] l0 du 2143),, u) exp(-ux), (5.43) where I)(q(a)+, u) denotes the linear term in H°R(q(a)+ + iu, v) — H0L(q(a)+ + iu, v). The u integral in this expression can be evaluated analytically. Then summing the contributions from Cl-C8 (with n < 1/2) or Cl-C4 (with n > 1/2) gives the branch-cut contributions to (r(x, (1)). Contributions to (r(x, (1)) of a second type come from the poles of the integrand on the real-q axis, where E(q, (1)) = 0. These poles are also the poles of I'l(q, (1)), so they 48 occur at the excitation energies of electron-gas states that are connected to the ground state via the density operator [1 l]. The poles occur on the real q axis when 1=Uo(q)RcrI°(q.nq). (5-44) and Im 11%, ()q) = o . (5.45) The dispersion relation (5.44) has the approximate solution (see Ref. 11, Sec. 15): (2,, = 1- am [1 + 9/10 (q/q-I-p)2 + ] (5.46) where up] is the plasma frequency: up, = (4 it ne2/m)1/2 (5-47) and q": is the Thomas-Fermi wavenumber, (ITF = (6 1|: n62/8F0)1/2 . (5.48) For a specified real frequency (1), we substitute 0.) for Qq in Eq. (5.46), then solve for q. We designate the positive solution q*. If (IQ and q* satisfy the condition lflq l > hq*kp/m + hq*2/2m, then 1m 110(q“, Qq) vanishes, and there are poles on the real q axis at :hq*, for drat value of (1). The residues at these poles contribute to the asymptotic value of the integral in Eq. (5.40). The poles correspond to undamped plasma oscillations, which occur widrin the RPA. (We note drat for numerical purposes, it may also be necessary to in- clude contributions fiom poles that are slightly displaced fiom the real axis, aldrough the formal asymptotic contribution of these poles vanishes.) In this section, we have shown how to reduce the evaluation of a(x, t), asymptotic in x, to a single remaining quadrature over (1). Numerical (or analytical) investigation of the ar-integral is reserved to future work. 6. CURRENT DENSITY SUSCEPTIBILITY There has been considerable'interest in quantum derivations of the conductivity using the density matrix formalism. Kubo [2] has given a formal solution and Kohn and Luttinger [36] have shown how the Boltzmann equation appears in a certain approximation. Among the earlier work, we mention papers by Dresselhaus and Mattis [37] and Mattis and Bardeen [38], related to an anomalous skin effect in normal metals. For a fiee electron gas in the absence of a magnetic field, expressions for conductivities are given in Mattis and Dresselhaus’s paper mentioned above. In the Coulomb gauge an expression for the current density is obtained by taking into account particle-like excitations, as in Mattis and Bardeen’s paper. The response of an insulator to a weak external electromagnetic field of long wave-length was studied by Ambegaokar and Kohn [39] from a many-particle point of view. They treated the Coulomb interaction between electrons to all orders of perturbation theory and analyzed the structure of the corresponding Feynman graphs. In their analysis they defined proper and improper polarization graphs. The proper polarization graphs occur for both transverse and longitudinal fields. Improper polarization graphs occur only in the response to a longitudinal field. We remark drat there has to be a close connection of their graphical analysis to our equation of motion. Here we study the paramagnetic current density susceptibility in a gauge wherein the scalar potential (1) is zero, and the vector potential alone represents the applied electromagnetic field. Then the longitudinal electric field can be expressed in terms of the longitudinal vector potential. The gauge invariance condition gurrentees that the a longitudinal electric field gives rise to the same current, whether it is 49 50 described by a vector potential or by a scalar potential. a. Tmsvme Current Densig Susceptibilig Eq. (4.26) serves as the starting point of our evaluation of the transverse current density susceptibility. The momentum transfer q is fixed along the z direction, and by considering the x component of the induced current density and replacing the summation by integration over p, we obtain the transverse susceptibility density xxx(q/z\,(1)) as )c,.,.((1’z\.m)=(e/m)2 d3p n(p) _ n(p) (1) Sin 9 Cos (102 (amp-(51ml (1)+ep .q- ep (6.1) The transverse susceptibility simplifies to pF 1‘ -(e/m)2 0! p4 tip 0 I Sin3 8 d0 1 1 (21:)2 ar+eq-(pq/m)Cos 0 +in ar—eq-(pq/m)Cos 9 HT] (6.2) where eq = q2/2m, 'n is a small positive quantity, and an extra factor of 2 is introduced to account for the spin summation. By evaluating the angular integral followed by the momentum integral (Appendix B), we obtain the following results for the transverse susceptibility. 51 xxx(q.(1)) = -(e/m)2/(21c)2 2mp3p/3 — (In/(l)3 (CD-+8.I + in)2{1/2[p2p -(m/q)2(eq+a) + i102] 1n eq+§+0+iTl eq-§+(o+in + (eq+co + in)(p,.m/q) } + ((0-8,, + in)2{ 1/21p2p -(m/q)2(eq-w - i102] 1n “hi-“mi E(r-fi-(D-ifl + (sq—co — inprm/q) } +(m/q) 1/41p4p -(m/q)4(eq+co + i104] 1n “WW" aq-§+w+in + 1/4(m/q)4(eq+m + in)‘{ 2&(eq+co + in) + ZINE/(8.1m + in)]3 } +1/4lp4p —(m/q)"(eq-(o - in)"] In w‘m'i“ sq-é-(n-in +1/4 (uvq)4(eq—w - in)4{2§/(eq-(° - in) + ”315"qu ‘ “m3 } (6.3) where F; = qu lm. By taking into account the identity (B.3) the above equation can be further simplified in the limit 11—90". 52 b. n ' ' nt ni usc tibili Eq. (4.30) serves as the starting point of our evaluation of the longitudinal current density susceptibility. By considering the 2 component of the induced current density and replacing the summation by integration over p we obtain the longitudinal current susceptibility density xzz(qQ,(1)) as xu((1/z\.(o)= (elm)2 d3p n(p) n(p) [p2 Cos 29 + q2/4] (map-2pm w ' Err-(1'81) §RpA(qr(0) + (elm)2 L131) n(p) n(p) P(lCOS 9 -— + (D-I-Ep-E'Hq WEN-8p The numerator of the longitudinal susceptibility simplifies to P1: +1 NLS(q£‘,cu)=_-2_ OJ p4 dp _1_1x2dx 1 _ 1 (211)2 meq-(pqx/m)+ in m—eq-(pqx/m) + it] . PF +1 +2_q OI p3dp _1_lxdx 1 + 1 (21:)2 (meg-(memoir (D-Eq-(qu/m)+iTl + (12 P‘(qz.co) I4 (6.5) where P1(q,(o) is the RPA polarization diagram and x = Cos 9. By evaluating the angular integral followed by the momentum integral (Appendix B), we obtain the following result for the numerator of the longitudinal susceptibility denoted by NLS(qz,(1)): 53 NLS(qz.m) = —2 K2102 2mp3n/3 +(m/q)3(co+eq + in)2{ 1/2[P21= —(m/q)2Fm/q)} -(mZ/quq + in){ 1 ”(1)21: "(m/Q)2(€q+0) + i102] 1n [Wmifl] eq-émin + (eqw + in)(ppm/q)} +(m2/Q)(0)—€q +in){1/2[p2p —(m/q)2(eq-co - i102] 1n [em-é—m—in] eq—g—co—i'n + (sq—o) - in )(me/q) } + q2 P1(q,to) /4 (6.6) Here again we can make use of the identity (B.3) and take the limit 11-)0” for further simplification. ] 7. CHANGES IN ELECTRONIC POLARIZABILITY DENSITIES DUE TO SHIFTS IN NUCLEAR POSITION The nonlocal polarizability density 00‘; r’, (1)) gives the co-frequency component of the polarization induced at point r in a molecule by an external electric field F(r’, to) acting the point r’, within linear response [18-23]. This property reflects the distribution of polarizable matter within the molecule; it represents the full response to external fields derived from scalar potentials of arbitrary spatial variation. Thus'0(r; r’, 0)) is a fundamental molecular property. It has applications in theories of local fields and light scattering in condensed media [20], and in approximations for dispersion energies [22], and collision-induced polarizabilities [23, 41] of molecules interacting at intermediate range. The hyperpolarizability density B(r; r’, (0’, r”, 0)") gives the polarization induced at r by the lowest-order nonlinear response to a field of frequency of acting at r’ and a field of frequency on” acting at r”. The electronic polarization Pim(r, 0)) induced in a molecule by an external field F(r, (0) depends on the polarizability density a(r; r’, (u), the hyperpolarizability density B(r; 1", (0’, r”, to”) and higher-order nonlinear response tensors: Pind(r-, 0)) = I dr’ am I“, or) - F(r’, to) + 1/2 1603' I dr’ dr” B(r; r’, (u - to', r”, (0’) : F(r’, to - on F(r”, m’) +. . (7.1) This work focuses on the changes in the frequency-dependent molecular polarizability density when a nucleus shifts infinitesirnally. In Ref. 42 it is shown 54 55 that the derivative of the static polarizability 867043(0)/3RIY is related to the nonlinear response tensor B(r; r’, O, r”, 0). This accounts for the connection between the polarizability derivative and the quadratic electric field shielding tensor at nucleus I, noted by Buckingham and Fowler [43]. The purpose of this work is to prove that the relation between 3%th r’, 0)/<3RI7 and the nonlinear response tensors generalizes to the frequency- dependent case. From Eq. (7.1), if a molecule is placed in a static external field Fs(r), its reaction to an additional external field P(r, 0)) [44] can be characterized by the effective polarizability density f(r; r’, to; F3), given by o°(r; 1‘30); F) = a(r; r’, to; F8 = 0) + I dr” B(r; r’, to, r”, 0) - Fs(r”) +. . . (7.2) The permutation symmetry of the B hyperpolarizability density has been employed to obtain this result. A shift 5R1 in the position of nucleus I in a molecule changes the nuclear Coulomb field acting on the electrons. In this section, we prove directly that the resulting change in polarizability density is determined by the same hyperpolariza- bility density B051“; r', (0’, r”, to”) that fixes the response to external fields. Specifically, we show agents r’, 0))[6R1a = I dr” [3515“; r’, to, r”, 0) 2I Two", R1), (7.3) where ZI is the charge on nucleus I and T8a(r"' R1) is the dipole propagator, 1043011, 1') = Vm v,3 ( I R1 - r H). (7.4) 56 The proof is via time dependent perturbation theory. We first calculate the nonlinear polarization of an isolated atom in an electromagnetic field. The atom may be characterized by a complete set of unperturbed eigenstates In) satisfying a Schrbdinger wave equation with energy eigenvalues En. In the presence of the perturbing Hamiltonian H’, the new ground state is denoted by IW ). In!) is expressed as an expansion in In) with coefficients an(s), where 3 refers to the order of of perturbation: llll>=2,,,s an“) In). (7.5) The 311(8) are determined by the Schrodinger equation to satisfy 8311(s+1)(t)/at = (in)1 2", (n lH’I n’) an,(s) + (an-1 (n IHOI n) an (7.6) subject to the initial condition anw) (t=0) = ans , an“) (t=0) = o for s at o (7.7) where the subscript g refers to the unperturbed ground state. The polarization P is the expectation value of the polarization operator in the perturbed ground state. P=E mug] (cang - 0)) (cang + to) (7.13) B“: I", (01, r”, (02) = K(-m0'; (01’ (02)('fi)-2 112 2",2’ m [ Emmaz>m m’ml)n§ + (11%,? at“)... (on, + (may, (F)... ,,] (0mg + m2)(mng + (no) (a)mg + m2 )(cong — (1)1 ) (7.14) where m6 = (01 + (02. The symbol I denotes a permutation operator: for example, 112 denotes the average of all terms generated by simultaneously permuting the frequencies (01 and 58 (1)2, and the corresponding operators. The notation 2’ denotes a summation from which the ground state is excluded. The barred notation (Hm)mu is shorthand for the expression aim)“m - (H7588. The numerical coefficients K depend on the nonlinear optical process of interest. Their evaluation has been discussed in Ref. 48. A shift 5RI in the position of nucleus I in a molecule changes the nuclear Coulomb field acting on the electrons. This shift can be considered as a static perturbation, and any perturbed state can be written as lat) = In) + 2,, Im)(m|H’|n) ——(En rm) . (7.15) Now we expand all states and energy denominators in the polarizability expression, Eq. (7 . 13), to first order in the nuclear shift perturbation, and neglect the higher order terms. In addition we replace the difference between the first order cor- rection to the energy (AF.n - ABS) by ( (H’(r”))nn - (H’(r”))gg). The resulting effective polarizability density can be written as, gem r’,co)= 0(r; r’, 0)) +oP(r-; r’, to) + . .. (7.16) 0%; r’. c0)= (-B)'2 I dr”E(r”) '2n’ (P(r))gn (H"“’(I"))..g (H’(r”)),,n (cons - (0)2 -gg (00,.g - c0)2 59 + 2', 2' m <1>(r»,,,..,,.g + m8ngE..E(P(r))gl,,mn + (H“"(r’)),..,,,g,.(H’(r”)>um (0m (cong — 0)) comm)n8 - to) 1 + [mung - (0) replaced by (con8 + (.0) ][complex conjugate of the rest ] . (7.17) The last two terms in aP(r; r’, co) can be rewritten by splitting the-sum into two parts, the first part with m at g and the second part with m = g. Fifth = 21.2 'nta nng(P(r))mrnn ‘2 'n(H 4”“ 7>ngns<1’('»ss<“’("7>sn term (um (a)m8 - 0)) tong ((l)“g - 0)) (7.20) By adding the first two terms in Eq. (7.19) and Eq. (7.20) we obtain the single term Z’nZ’mfl, (H"’(r’)),,.§(P(r))gngmmn ms (was ' E) mms(‘°ns ' “’5 (D + “(film — to) replaced by ((1),.g + to) ][complex conjugate of the rest] (7.21) The expression in brackets is identically equal to B(r; r’, to, r”, O) (with all permutations counted). This proves directly that the resulting change in polarizability is determined by the same hyperpolarizability density B(r; r’, (1)1, r”, (02) that fixes the response to external fields. Thus we prove the validity of Eq. (7 .2) and Eq. (7.3). 8. SUMMARY AND CONCLUSIONS We have unified the asymptotic expansions of linear response functions of an interacting Fermi gas in its ground state with a uniform positive charge background, so that the whole system is neutral. The results obtained are valid within the random phase approximation. In chapter 4 we demonstrated the possibility of studying the induced current density by use of the equation-of-motion method, as a linear response to a vector potential. We separated the transverse and longitudinal current-density susceptibilities, and in accord with physical expectations, there is no screening in the transverse current density susceptibility. In chapter 5 we have given a generalization of the Langer and Vosko result by evaluating the charge density susceptibility and the nonlocal polarizability density for a scalar potential of arbitrary spatial variation. The static longitudinal component of the nonlocal polarizability density is related to the dielectric function of the electron gas. The asymptotic analysis shows that 0Lu(r, r’; 0) has two long-range components within the random phase approximation. One is associated with Friedel oscillations in the charge density, and it reflects the impossibility of constructing a smooth function with the restricted set of wavevectors for the states in the electron gas that are unoccupied at T = 0 (k > kF) [11, 12, 16]. The second long-range component is an I r — r’ I’3 non-oscillatory term due to charge screening. Because of its physical origin, the presence of this term has significant implications for calculations of aL(r, r’; 0) for individual molecules: it raises the possibility of long-range, intramolecular polarization effects. As an extension of 61 62 the above work, we have reduced the calculation of the asymptotic, dynamic charge density susceptibility (within the RPA) to a single quadrature over frequency. In chapter 6 we have obtained explicit expressions for current density susceptibilities within the RPA in a gauge where in the scalar potential is zero. In chapter 7 we obtained a connection between linear and nonlinear response tensors, via a direct perturbation theory of general validity. We proved directly that the electronic charge distribution responds to the change in Coulomb field due to a shift in the position of the nucleus via the same hyperpolarizability density that describes its response to external fields. APPENDIX A This appendix provides mathematical details on the derivation of the asymptotic form of the nonlocal polarizability density aLzz(r, r"; 0) in the RPA, as given in Eq. (5.23). It also provides auxiliary results for the derivation of Eqs. (5.25), (5.26), and (5.28), and brief comment on the numerical evaluation of the RPA and VS polarizability densities. In deriving Eq. (5.23), we start with Eq. (5.16), which we first express in terms of integrals over the full real axis from -oo to co, and then evaluate by complex contour integration. To convert Eq. (5.16) to Eq. (5.19), we have used the fact that g(z) is an even function of its argument. In Eq. (5.16), the integrand containing j2(kx) is well-behaved as k —) 0, since limk _) 0 (kx)"l sin kx = 1. In contrast (kx)’l exp(ikx) diverges as k —-) 0. The convergent integral in Eq. (5.16) can be expressed in terms of the Cauchy principal value of the integral containing (kx)-l exp(ikx), however, as in Eq. (5.19). The Cauchy principal value of the integral is designated by p; it is defined as a symmetric limit: to L, dk k2 jzasx) «cikx mo —8 N lime _, 0L” dk k2 j2(kx) oi"x F(k) +1 8 dk k2 j2(kx) eikx F(k) . (A. 1) The function g defined in Eq. (5.20) has branch points at :1: 2kF :t in. By choosing the branch cuts shown in Figure 1, we ensure that g is real on the real axis, as required [11]. Also with this choice, g is real on the imaginary axis, in the limitn —+ 0. The integrals in Eq. (5.19) are evaluated by complex contour integration, with the contour shown in Frg. A. 1. The only singularity enclosed by the contour is a simple pole, 63 lmk 1 C(2) C(1) III/IIIIIIIIIIIIIIIIIIII ---- ------ -_-_-_-_--- ---- - --- --- - rain—7 4 -2kF + in 2kF + in 2.6% A x . Rek 6* \l I‘ ‘)fivvvuvwvzznt "iZI‘f: " i1] to r 'n I 3 4uvuuuvuvuUvvwwuanmnnnnannnnnnannnnnb( Figure A.l. Contour in the complex It plane, used for evaluation of the nonlocal polarizability density in the RPA. Branch cuts are shown as striped lines; the symbol x marks the branch cut origins at the points iZkFtin, apoleatk=0, andapole neari k1?- 65 located near ik'rF (see below). The integral over the large semicircular portion of the contour vanishes in the limit as R -> no, where R is the radius of the semicircle. The integral of ku eikx [E(k, 0)"1 - 1] along the smaller semicircle around k = 0 (with radius 1:, K -—> 0) has the value in: when n = -1; for larger n, the integral along the smaller semicircle vanishes. ' Hence, by the residue theorem, to L” dk 1tn eih [e(k, 0)-1— 1] + i rt is",l + 10(1) tiltlrn eikx [6(k, 0)-1 - 1] + C(2) dk k“ eikx [e(k, or1 - 1] = 21: i Kl , (A.2) where K1 is the residue at the enclosed pole. The contours C(1) and C(2) run around the branch cuts, in the sense shown in Fig. 1. Since [8(k.0)’1-ll = -k1p2ga 0, evaluation of g at points on the imaginary axis yields g(icpt) = 1/2 + (ope-1[1+(cp02/4larctanlcpt/21 . (A12) Hence the location of the pole depends on rs. When rs = 2.0, cp = 1.056; when rs = 4.0, cp = 1.114; and when rs = 6.0, cp = 1.172. The residue K1 is K1 = —1/2 (rely-1 kn,“+1 g(iepc) exp(-ckaFx) . (A.13) Therefore, the contribution to the integral hour the pole in the upper half-plane decays exponentially with x, and for large x it is negligible relative to the error in the asymptotic expansions of the integrals along C(1) and C(2). Use of Eqs. (5.21), (5.22), and (A2) yields Eqs. (5.23), (5.25), and (5.26) in the main text. We obtain Eq. (5.28) with the use of auxiliary integrals L, k (k2+ltTF2)‘1sinltx dk = m exp (-kTFx) (A.14) L, (k2 +k1-F2)-1 cos lot dk = warn.) exp (—k-I-Fx) (A15) and Jo [(1 (k2 + kWh-1 sin lot dk = w(2k1,,2)[1- exp (—kTFx)] . (A.16) 68 APPENDIX B The integral needed to evaluate the transverse susceptibility is [47, b] 1|: J Sin3 0 d9_ = (2m?) + (1mm - (a/b)21 1n[ Mb] . a+b Cos 9 a-b (3.1) The equality needed to evaluate all resulting momentum integrals is [47, c] Idx xan In (ax+b) = [l/(m-1-1)][x""'l - (-h/a)m+1] ln (ax+b) m+1 -[1/(m+l)](-b/a)“1+1 2 (1/r)(-ax/b)r r=l (B.2) We have used the following identity in order to simplify the current susceptibility densities: ln(a+i11)=ln I a | +i tan1(n/a) (B.3) 1. REFERENCES H. B. Callen and T. A. Welton, Phys. Rev. 83, 34 (1951). 1a. H. J. Kreuzer, Nonequilibrium Thermodynamics and its Statistical Foundation 10. ll. 12. 13. 14. 15. 16. (Oxford, Clarenden Press, 1981) Ch. 9. R. Kubo, J. Phys. Soc. Japan 12, 570 (1951). L. Onsager, Phys. Rev. 37, 405 (1931). D. N. Zubarev, Sov. Phys. Usp. 3, 320 (1960). D. Pines and D. Bohm. Phys. Rev. 85, 338 (1952). J. Bardeen, Phys. Rev. 52, 688 (1937). R. A. Ferrel, Phys. Rev. 107, 450 (1957). J. Lindhard, Kgl. Danske Videnskab. Selskab, Mat-Phys. Medd 28, No. 8 (1954). P. Nozieres and D. Pines, Phys. Rev. 109, 741, 762, 1062 (1958). H. Ehrenreich and M. H. Cohen. Phys. Rev. 115, 786 ( 1959). A. L. Fetter and J. D. Walecka, Quantum Theory of Many-Particle Systems (McGraw-I-lill, New York, 1971), Ch. 5. G. D. Mahan, Many-Particle Physics (Plenum, New York, 1990), Ch. 5. J. S. Langer and S. H. Vosko, J. Phys. Chem. Solids 12, 196 (1960). J. Friedel, Advan. Phys. 3, 446 (1954). J. Friedel, Nuovo Cimento, Suppl. 2, 287 (1958). P. Nozieres and D. Pines, The Theory of Quantum Liquids, Vol. 1 (Benjamin, New York, 1966), Ch. 5. 16a. T. J. Rowland and N. Bloembergen, Phys. Rev. 119, 912 (1960). 69 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31 32 33 70 R. Nimalakirthi and K. L. C. Hunt, J. Chem. Phys. 98, 3066 (1993). W. J. A. Maaskant and L. J. Oosterhoff, Mol. Phys. 8, 319 (1964). L. M. Hafltensheid and J. Vlieger, Physica 75, 57 (1974). T. Keyes and B. M. Ladanyi, Mol. Phys. 33, 1271 (1977). J. E. Sipe and J. Van Kranendonk, Mol. Phys. 35, 1579 (1978). K. L. C. Hunt, J. Chem. Phys. 78, 6149 (1983). K. L. C. Hunt, J. Chem. Phys. 80, 393 (1984). B. Linder, Adv. Chem. Phys. 12, 225 (1967); B. Linder and D. A. Rabenold, Adv. Quantum Chem. 6, 203 (1972); T. B. MacRury and B. Linder, J. Chem. Phys. 58, 5388, 398 (1973). A. Koide, J. Phys. B 9, 3173 (1976). M. Krauss and D. B. Neumann, J. Chem. Phys. 71, 107 (1979); M. Krauss, D. B. Neumann, and W. J. Stevens, Chem. Phys. Lett. 66, 29 (1979); M. Krauss, W. J. Stevens, and D. B. Neumann, ibid. 71, 500 (1980); M. Krauss and W. J. Stevens, ibid. 85, 423 (1982). B. Linder, K. F. Lee, P. Malinowski, and A. C. Tanner, Chem Phys. 52, 353 (1980); P. Malinowski, A. C. Tanner, K. F. Lee, and B. Linder, ibid. 62, 423 (1981);S.Wozniak, K. F.Raof, and B. Linder, J. Chem. Phys. 79, 220 ( 1983). A. Koide, W. J. Meath, and A. R. Allnatt, Chem. Phys. 58, 105 (1981). M. E. Rosenkrantz and M. Krauss, Phys. Rev. A 32, 1402 (1985). J. Applequist, J. R. Carl, and K.-K. Fung, J. Am. Chem. Soc. 94, 2952 (1972); J. Applequist, Acc. Chem. Res. 10, 79 (1977); J. Chem. Phys. 83, 809 (1985). . A. J. Stone, Chem. Phys. Lett. 155, 102, 11 (1989); A. J. Stone and C. S.Tong, Chem. Phys. 137, 121 (1989). . G. Arfken, Mathematical Methods for Physicists, 2nd ed. (Academic, New York, 1970). P. 557. . C. Cohen-Tannoudji, B. Diu, and F. Laloé, Quantum Mechanics, Vol. 1 (Wiley, New York, 1977), p. 688. 71 33a. P. Vashishta and K. S. Singwi, Phys. Rev. B 6, 875 (1972). 34. P. Vashishta and K. S. Singwi, Phys. Rev. B 6, 875 (1972). 35. K. S. Singwi, M. P. Tosi, R. H. Land, and A. Sjtllander, Phys. Rev. 176, 589 (1968); K. S. Singwi, A. Sjolander, M. P. Tosi, and R. H. Land, Phys. Rev. B 1, 1044 (1970); K. S. Singwi and M. P. Tosi, Solid State Phys. 36, 177 (1981). 36. W. Kohn and J. M. Luttinger, Phys. Rev. 108, 590 (1957). 37. G. Dresselhaus and D. C. Mattis, Phys. Rev. 111, 403 (1958). 38. D. C. Mattis and J. Bardeen, Phys. Rev. 111, 412 (1958). 39. V. Ambegaokar and W. Kohn, Phys. Rev. 117, 423 ( 1960). 40. K. L. C. Hunt, Y. Q. Liang, R. Nimalakirthi, and R. A. Harris, J. Chem. Phys. 91, 5251 (1989). 41. K. L. C. Hunt and J. E. Bohr, J. Chem. Phys. 84, 6141 (1986). 42. K. L. C. Hunt, J. Chem. Phys. 90, 4909 (1989). 43. P. W. Fowler and A. D. Buckingham, Chem. Phys. 98,167 (1985) 44. A. D. Buckingham, Proc. R. Soc. London, Ser. A 267, 271 (1962); and D. M. Bishoop, Mol. Phys. 42, 1219 (1981); J. Chem. Phys. 86, 5613 (1987). 45. B. J. Orr and J. F. Ward, Mol. Phys. 20, 513 (1971). 46. F. W. J. Olver, Asymptotics and Special Functions (Academic, New York, 1974), Ch. 3. Ch. 9. 47. I. S. Gradshteyn and I. W. Ryzhik, Tables of Integrals, Series, and Products (Academic, New York, 1965), 4th Ed., (a) 4.3522. (b) 3.6445 (C) 2729.1 "llllllllllllllll