. L... .. . .. , .., 2...... J. 3.... 1.2.... . .gfigfi. . z. 9.. .52.“... .. ~ 3% 7...? . I :... . .p .3. a a u .q z .1 . 2' $me , am :l. 1.1.0? .I 15.5.3!!le . tutti-.01.. J. J c. . er‘at. .I magma, a; A. . : , z z .1 2 $65.. E s. -30.... s; 1.5;: .L‘ . flux? 5 usol . , .. .. .. .. I ..._. Hm... L, . :3. a 9...... V a... m“. . .. Dru; I _ r.” 3.1.. .... .5; , I}... . 5 in... Saki. .2 , i}? .tVt it 1.... 3H”?! 62‘... .1; It: s ‘I t (gin 2.?th ‘. K. .3: .2 ‘1“ all s .3 .23... g...) to: .7... .X tax).....!t.u¢.:!.\| 93.1....“ .21“ .I . .I Mi... 32...}...3. . .gbe... .1 E Ad- lit! .9 r," ,. . .. I. .1933? t... xt}.€. L37, )1. 3.5.1.1 3.5.1.1 . t. n v LIBRARY Michigan State University This is to certify that the dissertation entitled NONLOCAL DIELECTRIC MODEL FOR INTERMOLECULAR INTERACTIONS AT SECOND AND HIGHER ORDERS presented by ANIRBAN MANDAL has been accepted towards fulfillment of the requirements for the Doctoral degree in Chemistry ”(wig 0 4M— ' Major Professor’s Signature 10/ 74 /0‘I Date MSU is an Affinnative Action/Equal Opportunity Employer PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or- before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 5/08 KzlProjIAccaPreleIRC/Dateoue.indd NONLOCAL DIELECTRIC MODEL FOR INTERMOLECULAR INTERACTIONS AT SECOND AND HIGHER ORDERS By Anirban Mandal A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Chemistry 2009 ABSTRACT NONLOCAL DIELECTRIC MODEL FOR INTERMOLECULAR INTERACTIONS AT SECOND AND HIGHER-ORDERS By Anirban Manda] The nonlocal dielectric function of a molecule determines the effective potential at a certain point due to an applied external potential at a different point, within the molecule. The effective potential at point r is determined by the nonlocal dielectric firnction sv(r,r';co) within linear response and by the nonlocal dielectric function sq (r, r', r"; to, (0') within nonlinear response to the lowest order. The nonlocal dielectric functions ev(r,r';w) and sq (r, r’,r';co, (0') depend on the charge-density susceptibilities x(r,r';m) and xq (r,r',r"; a), (0') of the molecule, respectively. This work shows that for a group of interacting molecules with weak or negligible charge overlap, the nonlocal dielectric model gives the interaction energies and forces at second and higher-orders, in agreement with the results from quantum mechanical perturbation theory. The dielectric model accounts for screening due to electronic charge redistribution in the interacting molecules; it accounts for polarization and fluctuations in the charge densities that act as sources of the external potentials. The model applies within the Born-Oppenheimer approximation. At second order, both two-body pairwise additive and three-body non-additive induction energies appear. We prove that the two-body induction energy is determined from changes in the static Coulomb interactions within each molecule, caused by a neighboring molecule that acts as the dielectric medium, whereas the three-body induction energy is determined by the changes in the static Coulomb interactions between two molecules, due to the presence of a third molecules which acts as the dielectric medium. Dispersion energy is pairwise additive at second order and results from changes in the intramolecular exchange-correlation energy caused by the dielectric screening due to a neighboring molecule. The interaction energies at second order include linear screening only, while the induction and dispersion forces on nuclei result from both linear and nonlinear screenings. We show that induction forces result from nonlinear screening of the potential due to permanent charge distribution of the neighboring molecule and linear screening of the potential due to induced polarization of the neighbor, while dispersion forces result from nonlinear screening of the fluctuating charge distribution of the neighboring molecule and linear screening of the dynamic reaction field from the neighbor. The linear screening present in the dispersion force is described by transition susceptibility of the molecule. Moreover, the dispersion force includes effects which don’t have a dielectric interpretation. At third and fourth order of molecular interactions, the induction and the induction-dispersion energies show both linear and nonlinear screening, while the dispersion energy includes linear screening only. Depending on the nature of interaction, the induction energy can be classified into different categories, each category showing a different screening effect. Screening in the dispersion energy can be described either by the dielectric function of a single molecule, or by the dielectric functions of two or three molecules. To my parents iv ACKNOWLEDGEMENTS I would like to express my sincere gratitude to my advisor Professor Kathy Hunt for her invaluable guidance and enduring patience throughout the course of research at graduate school. I am deeply indebted to her for her help, support, teaching and encouragement during my stay at MSU. I would like to thank my second reader Professor James Harrison for helping me with the computational methods and for encouraging me to continue research in theoretical chemistry. I would also like to thank my committee members Professor Piotr Piecuch and Professor James McCusker for their great advice and teaching. I take this opportunity to thank Dr. D. J. Gearhart, Dr. Ruth L. Jacobsen, Dr. Maricris Lodriguito, Afra Panahi, Mariam Sayadi, Jeff Gour, and Jesse Lutz for their help and company. I thank Janet Haun and Zeynep Altinsel for their help and kindness during my hard times. I am grateful to my friends Anindya and Paramita for their help, support and warm company. Finally, I would like to take the immense pleasure to thank my parents, my uncles, my aunts and my late grandma for their love and support. Without their help, I would not be able to come this far. Table of Contents Chapter 1: Introduction ................................................................................................................... 1 Chapter 2: Nonlocal dielectric functions on the nanoscale: Screening of two- body induction and dispersion energies at second order ............................................. 35 2.1: Dielectric screening and the induction energy 35 2.1: Dielectric screening and the dispersion energy 38 Chapter 3: Dielectric screening of second-order induction and dispersion forces on nuclei of interacting molecules ................................................................... 43 3.1: Dielectric screening and the second-order induction forces on nuclei .................................................................................................................... 43 3.2: Dielectric screening and the second-order dispersion forces on nuclei .................................................................................................................... 50 Chapter 4: Dielectric screening and the three-body nonadditive interactions at second order ............................................................................................................ 75 4.1: Dielectric screening of the three-body induction energy at second order .............................................................................................................. 75 4.2: Dielectric screening of second-order three-body forces on nuclei of interacting molecules ............................................................................................ 79 Chapter 5: Dielectric screening of three-body and four-body interactions at third and fourth orders ............................................................................................. 85 5.1: Dielectric screening of the induction energy at third order .............................................................................................................. 87 5.2: Dielectric screening and the nonadditive dispersion energy ...................................................................................................... 98 5.2A: Dielectric screening of nonadditive three-body dispersion energy at third order ............................................................................................... 98 5.23: Dielectric screening of nonadditive four-body dispersion energy at fourth order ........................................................................................... 105 5.3: Dielectric screening and induction-dispersion energy at third and fourth orders ........................................................................................... 111 Chapter 6: Summary and conclusions ......................................................................................... 117 Appendices .................................................................................................................................. 134 References ................................................................................................................................... 145 vi Chapter 1: Introduction . . . 1 . . A nonlocal dielectric functlon, 8V(l',l";(1)) characterlzes the screening of an applied scalar potential (Pex (r';w), due to electronic charge redistribution within a molecule. Within linear response, the effective potential (peff (run) at a point r within a molecule is related to an applied potential (Pex (r';o)) of frequency a) by (Peff = 80 [dr's'v1(r,r';w)+ (0 l3(r')G(-60) 50) I0) ]- (1 -3) In Eq. (1.3), 0) denotes the ground state of the isolated molecule, {5(r) is the charge- density operator, G((:)) is the resolvent operator, G(w)=(1 - 500)(Ho-Eo-hw)'1(1-soo), (1.4) 500 denotes the ground-state projection operator |0> <0 , H0 is the unperturbed Hamiltonian of the molecule, and E0 is the unperturbed ground-state energy. For a molecule with fixed nuclear positions, x(r,r’;w) determines the change in the electronic charge density Ap(r,(o) within linear response to a frequency-dependent external potential ‘Pex (r,o)) , via Ap(r, w) = Idr’ x(r,r’;0)) ‘Pex (r', (o). (1.5) The effective potential (peff (r; 0)) within the molecule is given by (pefi‘ (r, (0) = (pex(r,(1))+(41te0)_1 Idr'Ir—r'|_1Ap(r',o)), (1.6) without restriction to linear response in calculating Ap(r,o)). Thus, (peff(r,(o) has a source term containing the sum of the external charge density Pex (r’, (1)) that generates ‘Pex (r',a)) and the change in the molecular charge density Ap(r,o)) induced by (Pex (r', co) . Hence (peff (r, 0)) satisfies the Poisson equation by V2 (peff (r, (o) = —[pex (r, (1))+Ap(r, (13)] / 80. (1.7) The linear response theory suffices to describe the intermolecular interactions at first order. At second and higher orders, we must include nonlinear response of the molecule to the applied potentials. Including the lowest-order nonlinear response to (Pex (r', (0) , we obtain Ap(r,(o) = Idr' x(r,r';o)) ‘Pex (r', (o) + (1 / 2) £000 dco Idr' dr" C(r,r', r"; (o — (o',o)') x (Pex (r', (0 — 03') (Pex (r", 05') , (1.8) where the quadratic charge-density susceptibility C(r, r’, r”; (o — co',w') satisfies the relation C(rar'J'm, w') = S 0‘31“"; 60, 60')[<0 50‘) G(wo)fio (I‘")G(0)) 15(1") + o>]. (1.9) b. We do not assume that damping is negligible in general (particularly near resonance), but in cases where damping is negligible, C(r,r',r';co,w") is symmetric under permutation of all of the variable pairs:(r,—(oo.), (r',(o), and (r",co"). The method used by Orr and Ward3 to derive multipole polarizabilities yields Eq. (1.9), when applied to charge-density operators: cf. Eq. (43b) of ref. 3. Eq. (1.9) is also consistent with the results for multipole hyperpolarizabilities derived by Bishop, in Eq. (41) of ref. (4). From Eq. (1.8), a quadratic dielectric fimction can be defined by so 861031", r”; a), (11’) = (41rso )_1 jdr'" Ir —r"'|_1C(r",r',r",o),w'). (1.10) Thus, within nonlinear response to the lowest order, the effective potential within a molecule is obtained as cpeff(r,w) = so [dr'831(r,r';co) is the wavefunction of the interacting system. This particular separation of the 98-100 Hamiltonian is known as the polarization approximation. At first order, the 100.10] interaction energy for a pair of molecules is purely electrostatic and depends on the permanent charge densities (and hence on the permanent moments) of the molecules. A perturbation analysis for the first-order interaction energy AEU) of a pair of molecules A and B with fixed nuclei and negligible charge overlap yields —1 AB“) =(41tso)—1 jdrdr' p0A(r)p(1)3(r')|r—r' , (1.12) where p0A (r) and p5 (r’) are the permanent charge distributions in molecule A and B 102.103 respectively. Following Longuet-Higgins, the charge-density operator (3A (r) for molecule A can be defined as (3A(r) = Ze6(r—rj)+ZZI6(r—RI), (1.13) j I - where the sum over j is for all the electrons assigned to molecule A, with position operators rj, and the sum over I is for all the nuclei in A with charges ZI and positions RI. For a pair of molecules with fixed nuclear configurations and negligible charge overlap, the first-order interaction energy can be expressed using a multipole expansion]04 in terms of the charge q, permanent dipole no, permanent quadrupole (90, and higher moments of each molecule (1)_AB_AB_ A .8 AE —q (p0 “OaSOa (1'3)®0a330a8+'” _ B A B A B IA (1):)3 is the potential at the origin RA of molecule A due to the permanent, unperturbed charge distribution in molecule B. 303a and 3,0118 are the field and the field gradients at RA respectively, due to the unperturbed charge distribution in B. The Einstein summation convention over repeated Greek indices has been used in Eq. (1.14). Since the interaction at first order depends on the unperturbed charge distributions (and hence on the permanent moments) of the molecules, no screening effect is apparent in the first-order interaction energy. The first-order force on a particular nucleus I in molecule A can be obtained from Eq. (1 . 12) using an energy-based approach, employing the relation 171(1) =—vI AB“), (1.15) where V1 denotes differentiation with respect to the coordinate of nucleus 1, RI. Electrostatic forces on nuclei can also be derived using the Hellmann-Feynman 105J(X5 theorem and Sternheimer—type shielding tensors.107 The forces on a nucleus in an atom in presence of an external field have been related to the dipole shielding factors by Epstein108 and Sambe,109 where the dipole shielding factor 7 has been defined as y = 1 - o, with o the fraction of the external field actually felt by the nucleus. For diatomic or polyatomic molecules, the dipole shielding factor is a second-rank tensor represented by yao, and relates the effective field at the nucleus to the external field by”0 3‘11 = (Sao —yao)So. Calculations on the dipole shielding factors have been done by lll-l2 several groups. 3 70113 was related to a molecular property in the works of Lazzeretti 10 124,125 109 and Zanasi, Sambe, and Epstein.108 It was shown that Yell} is related to the derivative of the molecular dipole moment with respect to the nuclear coordinate by 21(6aB—yflfi)=ahg/0RL. (1.16) Fowler and Buckingham]26 extended the shielding factor to include non-uniform applied fields and non-linear terms. In terms of their formulation, the net electric field on nucleus I in molecule A in presence of an external field is a}, = 3%?” +(5aB ailing +(1/2)(pI 01137 3134736 +... +(1/3)[(3/2)(R[13—RBA)8W+(3/2)(Rf),—RYA)5ao—(Ra—Ré)éoy e ”asy]3§y+(“3)§nsye :53 375+“ (1.17) where we have kept only the terms which depend linearly and quadratically on the fields and the field gradients. 36 is the external field with gradient 3'6, applied at RA . 7313 is the Stemheimer shielding tensor, V3107 is the field-gradient shielding tensor, and (p1 (1’57 #1375 are the nonlinear shielding tensors. 39m is the field that acts on nucleus I in absence of any external perturbation. Using an analogy with the expansion of the induced dipole moment in powers the external fields and field gradients and the Hellmann- Feynman theorem, they connected the higher-order nuclear shielding tensors to the derivatives of molecular multipole moments and polarizabilities with respect to the nuclear coordinates: lel 1 wag, = chm/GR , (1.18a) 11 zI [(3/2)R'13 so,Y +(3/2)R§ 50113 41}, 567 +vI ]= 59" MRI}, (1.18b) (1137 B7 21 s1 - 6A /6R1 (1 18¢) (11378 _ I375 01 ’ ' a is the dipole polarizability and A is the dipole-quadrupole polarizability. However, neither in their work nor in the work by Lazzeretti and Zanasi were the derivatives of the molecular moments and the polarizabilities expressed in terms of molecular response tensors. Derivatives of the molecular dipole moment and polarizability density with respect to the nuclear coordinate were evaluated by Hunt,127 and later generalized by Hunt, Liang, Nimalakirthi, and Harris.128 It was proved127 that the Stemheimer shielding tensor is related to the nonlocal polarizability density a(r, r') '27’128‘60'62’6’129’130 by 7313 = - Idrdr'TaY(RI,r)aYB(r,r'). (1.19) Within the framework of the nonlocal polarizability density a(r,r';0), the first-order forces on nuclei were derived by Hunt and Liang.'31 Specifically they showed that although the first-order interaction energy depends only on the permanent charge distributions in the molecules, the first-order force on nucleus I in molecule A depends on the electronic polarization induced in A by B: —l A135“) = zI jdr'pO (r')(R}Jl —r;,) RI —r' +21 IdrTao(RI,r)APg\(l)(r). (1.20) APéA(1)(r) is the first-order polarization in A induced by the field from B and TaB (RI,r) is the dipole propagator. AP; (1) (r) depends on the nonlocal polarizability 12 density of A. The results in their work were obtained using an equation derived by Hunt which shows that the infinitesimal motion of a nucleus within a molecule changes the Coulomb field, and the response of the electronic charge density to that change in the field is determined by the same nonlocal polarizability density that determines the induced polarization of a molecule in presence of a static external field. In a later work, 132 Liang and Hunt showed that both the energy-based theory using Eq. (1.15) and the electrostatic-force theory based on Helhnann—Feynman theorem and the shielding tensors yield identical result for the first-order forces on the nuclei. In that work, they explicitly connected the shielding tensors and the nonlocal polarizability densities. The first order forces on nucleus I due to interaction with B were showed to be Pg“) = 210303 —Y(IXB)3gB+(l/3)[(3/2)(Rll3 —RBA)80W +(3/2)(R§, —RYA)5a[3 4R}, 419mm Him 136?” +..., (1.21) where 3% is the static external field at RA due to the permanent charge distribution in B, and V3157 = Idrdr’Ta5(RI,r)a58(r,r')[(3/ 2) (113 430579. +(3/2)(r1¥-RYA)Sog—(ré—R§‘)8oy]. (1.22) Thus the force obtained using the shielding-tensor approach was related to the force derived using the nonlocal polarizability density. However, the polarizability-based approach was not connected to a dielectric framework. The intramolecular nonlocal dielectric model was first applied to intermolecular interactions by Jenkins and Hunt.1 In their work, it was shown that the force on nucleus I 13 in molecule A depends not on the “bare” Coulomb potential due to the unperturbed charge distribution in B, but on an effective potential. The effective potential on nucleus I is determined by the nonlocal dielectric function sv(r,r';0) of molecule A due to redistribution of the electronic charge cloud within A. Using a susceptibility-based approach, they proved that the first order force 171(1) on nucleus I is 11 I . -1 . B . F( ) =.2 soa/6r[_[dr sv’A(r,r;O)(pex(r;0)] r=RI’ (1.23) where the notation [6/ arf(r)] | r = RI means that the derivative is first evaluated with respect to the coordinate r and then r is set equal to RI .Thus, the first-order force on the nucleus was connected to the nonlocal dielectric model. At second order, the interaction energy consists of a sum of induction and dispersion terms. The induction term AEigc)l is classical and appears from static, linear response of one molecule to the potential from the unperturbed charge distribution of the neighboring molecule. The dispersion term AEgzigp is purely quantum mechanical and depends on correlations between the spontaneous charge-density fluctuations within the 133 interacting molecules. Both induction and dispersion energy are obtained from time- independent perturbation theory with the second-order correction to the energy given by AEQ) = —Z Z/(Em +En —Eo). (1.24) A B . . - " AB . m and n are the excrted states in molecules A and B respectively, V 15 the interaction Hamiltonian, and Em, En, and E0 are the unperturbed energies of excited state 14 and ground state respectively. The sum of the terms from Eq. (1.24) with excited states confined to either molecule A or B yields the induction energy, while the sum of the terms with excited states on both molecules gives the dispersion energy. The induction energy depends on the permanent charge distributions and the molecular polarizabilities of the interacting molecules. In terms of the permanent 104 moments of the molecules, B —(1/3)AA sBas -A(1/6)c 3B 3B 2 as? ) d=-(1/2)aa1330a5 33 1111730 33?), 0.1375 30ers eye B 0A“ A 1A A s'A (1.25) SIA IA _(1/6)CB3075_(1/15)Ea67630030B75+ 01878 300113 where 3" is the gradient of the field-gradient, C is the quadrupole polarizability, and E is the dipole-octopole polarizability.104 Within a nonlocal response model, the induction energy is described by the nonlocal polarizability density and the static polarizations of the molecules.l3l The induction energy can also be expressed in terms of nonlocal charge-density susceptibilities and the static external potentials from the unperturbed charge distributions of the interacting molecules. We give a detailed description of that in chapter 2. Since the induction energy depends not only on the permanent moments of the interacting molecules, but also on the induced polarization in one molecule due to permanent moments in the other, it is apparent that induction energy can be described in terms of intramolecular screening. However, the induction energy has not been connected to a dielectric screening model previously. 15 The induction force at second order can be calculated using Eq. (1.15), where the induction energy is given by Eq. (1.25) or in terms of the static polarizations and the nonlocal polarizability densities of the interacting molecules. Within the nonlocal polarizability density model, the second-order induction force on nucleus I in molecule A depends on the derivatives of the static nonlocal polarizability density aaAB (r, r';0) and the field SOAa (r) at B due to permanent moments of A. Unlike the first-order force which depends on the permanent charge distribution in B and the first-order induced polarization of A, induction force at second order depends on the first-order induced polarization in B and the second-order induced polarization in Am: F5311! = [(1erTaB(RI,r)[APéA(2)(i-)+APIB(1)(i-)], (1.26) where the second-order induced polarization is determined by the nonlocal 130,134 A 13 afiy(r,r',r';0,0). Induction force on nucleus I can hyperpolarizability density, also be determined using Stemheimer-type shielding tensors by132 1(2)_I _I B 11 .B AFind,a — Z (5018 yaB)SRB+(l/3)Z vaBysRBy +(1/3)zI [(3/2)(Ré —RBA)80W +(3/2)(R§—RYA)5,,B —(R(11- Rémm] 11,337 +(1 / 2) ZI¢<£137308133€Y , II B ,B +(1/3)z 5111311530133075 +... (1.27) 16 where 333 and SQRBBY are the reaction field and its gradient from the first-order polarization of B, induced by the unperturbed charge-distribution in A. Thus, the second- order induction force on nucleus I depends on linear screening of the reaction field and its gradients due to the first-order induced polarization of B and nonlinear screening of the field and its gradients due to the unperturbed charge distribution in B. The two different approaches used to calculate the induction force are connected to each other by Eqs. (1.19), (1.22), and the relations132 between the nonlinear shielding tensors ¢€£BY , 5:375 and the nonlocal hyperpolarizability density, ¢ 0, the dispersion energy goes to zero, while in the limit r -> 00, it yields the well known multipole expansion series. Derivations and calculations of the dispersion energy in terms of approximate charge density susceptibilities have been given in Ref. 146 - 152. 153,154 Within the density functional theory (DF T), dispersion energy is contained within the exchange-correlation energy. Density functional theory has been employed to calculate van der Waals interaction energies by Langreth and Perdew,155 Harris and 156 Griffin, Gunnarsson and Lundqvist.157 Anderson et al. 158 evaluated the frequency- dependent polarizabilities for different atoms and ions and the van der Waals C6 constants for several atomic and molecular interactions by using a modified effective density neff, originally used by Rapcewicz and Ashcroft.159 Several research groupsmo'163 have used the method of a coupling constant 2, that turns on the electron-electron interactions. Kohn, Meir, and Makarov164 (see also chapter 4) have employed DFT to calculate van der Waals interaction energy between small and large intersystem distances. They have approximated the density distribution n(r) by the local density approximation 154 (LDA) or by the generalized gradient approximations (GGA).165 The Coulomb interaction energy was separated into short and long-range parts, and the long-range contributions to the interaction energy were expressed using the adiabatic connection 21 formula. Finally, the expression was transformed to the time domain and was applied to the calculations of the asymptotic van der Waals interaction between H-He and He-He. In their work, the van der Waals energy was obtained as Evdw = —C6 /R6, (1.32) where ll 22 xA (t1)xB (t2) C =3/n dt dt . 1.33 6 ( >13" 113° 2 ”Hz ( ) x22 is the 2 component of the density response to a perturbation in the z direction. A seamless van der Waals density functional has been formulated and applied to the interaction between two self-consistent jellium metal slabs by Dobson and co-workers.166 In their work, the correlation energy (EC) has been determined by the adiabatic connection fluctuation-dissipation formula (ACFD), which relates EC to the Kubo density-density response function st , the electron-electron interaction VCoul , and the Kohn-Sham density-density response function XKS- XKS depends on the average ground-state electron density. MS is determined from XKS’ the exchange-correlation kernel fchS , and a modified electron-electron interaction AVCoula by the Dyson-like screening equation. A related derivation and calculation have been done in later work by Dobson and Wang!“ For a detailed review of van der Waals studies using conventional density functionals, we refer the reader to ref. 186. In more recent work, Hunt168 has derived the electronic energy as a functional of the average electronic charge density and the average of the gradient of the charge-density fluctuations with respect to an external 22 potential. The functional is nonlocal. However, it is distinct from the other nonlocal . . 163.164.170-181 functronals berng used. Jenkins and Hunt182 have derived a model where the correlation between the polarization fluctuations has been connected to a nonlocal dielectric function sd (r,r';io)). The nonlocal dielectric function sd (r,r';o)) determines the dielectric displacement D(r,o)) within a molecule, due to an applied external field E(r’,(o) at r'. For a translationally invariant system, sd (r, r’; (1)) depends solely on r—r' , and hence can be represented as the spatial Fourier transform 5d (k; or). For these systems, 3d (k; or) is connected to the potential screening function sv(k;(o). On the intramolecular scale, the two dielectric functions are quite different due to the inhomogeneity of the intramolecular environment. The dispersion force on nucleus I in molecule A can be calculated by taking the negative gradient of the dispersion energy with respect to RI , the coordinate of nucleus 1. Within the nonlocal polarizability density model, the dispersion force on I depends on the derivative of the dynamic nonlocal polarizability density of A and the derivative of the correlation of the polarization fluctuations within A, with respect to RI. The derivative of the frequency-dependent nonlocal polarizability density with respect to RI is related to 128 the frequency-dependent hyperpolarizability. Thus, this component of the dispersion force results from the interaction between the nucleus and the nonlinear polarization of A induced by the polarization fluctuations in B. If we denote this component by Ms?) , 132 then 23 AFAS‘}; = 21 Eodm [drTaB(RI,r)APéA(2)(r,m). (1.34) Thus, the first component of the dispersion force resembles the first component of the induction force from Eq. (1.26), with the difference that in the case of the induction force the polarization in A is induced by the fields from the static polarizations in B, while for dispersion force the polarization is induced by the fields from polarization fluctuations in B. This part of the dispersion force also corresponds to the atomic dipole force in the work by Nakatsuji and Koga,135 although the dispersion forces and dispersion induced dipole were not considered explicitly in their work. The second component of the dispersion force depends on the derivative of the correlation between the polarization fluctuations within A with respect to RI. The correlation between the polarization fluctuations at r and 1" depends on the imaginary part of the nonlocal polarizability density by the fluctuation-dissipation theorem137 (1 / 2) (SP5: (1', (1)) 511? (r', to') +59%;A (r', m') 5P5,A (r, 05)) = (h / 27:) 019‘; (r, r'; 0)) 8(a) + 03') coth(hco/2kT) . (1.35) The infinitesimal shift of nucleus I within molecule A changes the static Coulomb field that modifies the above correlation. This is similar to the field-induced fluctuation 183,184 correlations studied earlier. Due to the change of the nuclear Coulomb field, the correlation depends on the imaginary part of the frequency-dependent hyperpolarizability A! “BY (1', I", 1’30, 0). Thus the magnitude of the correlation is changed. As pointed density, [3 out by Liang and Hunt,132 the change in the static external field may also introduce new 24 types of correlations in molecule A. In chapter 3 of this work, we show that the Coulomb field from nucleus actually brings in new correlation within the molecule. One thing to note here is that at second order, both the induction and the dispersion force depends on interaction of nucleus I with the polarization induced in molecule A. Within the context of dispersion force for interacting atoms, this is known as the Feynman’s “conjecture”.105 Quoting from Feynman’s work on the electrostatic description of forces between interacting atomslos, “It is not the interaction of these dipoles that leads to van der Waals’ force, but rather the attraction of each nucleus for the distorted charge distribution of its own electrons that gives the attractive 1 / R7 force.” 186 for the Feynman’s conjecture was first proved by Hirschfelder and Eliason special case of two hydrogen atoms, both in the Is state. It was also addressed by Nakatsuji and Koga135 in their work on electrostatic force theory. Hellmann-Feynman forces on the nuclei of two interacting He atoms have been calculated in later work by Allen and Tozer.187 The first general proof of Feynman’s conjecture was given by Hunt.188 In that work it was shown that the dispersion force on nucleus I in molecule A results from the interaction of I with the dispersion-induced change in the polarization of A: APSE) = zI jdrAP$(i-)T¢S(R1,r), (1.36) where AP? (r) depends on the nonlocal hyperpolarizability density of A and the nonlocal polarizability density of B, at imaginary frequencies. In the later work by Liang and 25 HuntI32 it was mentioned that the dispersion force on nucleus I depends entirely on the hyperpolarizability of molecule A; and that it does contain a component that stems directly from the polarization of B. In chapter 3, we prove that a part of the second component of the dispersion force actually depends on the linear response of A to the field from the induced polarization of B, as determined by the transition- 189‘190 of A. Thus, this part of the dispersion force is similar to the second polarizability component of the induction force in Eq. (1.26). For three or more interacting molecules, nonadditivity appears at second and higher orders. At second order, nonadditivity appears only in the induction energy. At third and higher orders, the intermolecular interactions consist of induction energy, 253,258,261-264 dispersion energy and induction-dispersion energy . All of the three types of interactions show nonadditivities. The first calculation of the three-body interaction energy was done by Axilrod and Tellerm and by Muto,192 who calculated the long- range triple-dipole energy of three interacting atoms with spherical charge distributions using third-order perturbation theory. The interaction energy was found to be 1 + 3 cos(91 ) cos(02 ) cos(0 3 ) AEDDD = C9 3 3 3 (1.37) r r r 12 23 31 112,123,131 are the sides and 01,02,03 are the angles of the triangle formed by the atoms, C9 z (9/16)Vol3 , where V is the atomic ionization potential and a is the 193 to study the polarizability. The triple-dipole interaction was first applied by Axilrod preferred lattice structures of the rare gases. Long-range many-body interactions are more dominant in condensed phases than in molecular clusters.194 The importance of many 26 body effects has been investigated in studies of the thermodynamic properties of 1 97-200 fluids,195’196 calculations of third virial coefficients , and studies of molecular 193,201-205 crystal structures. F orrnisano et al.206 have related the small-k behavior of the static structure factor S(k) in a noble gas fluid to the two (London)- and three (Axilrod- Teller triple-dipole)-body potentials. More recently, Jakse and co-workers207 have used a potential energy function based on the two-body potential of Aziz and Slarnanzog’209 and the triple-dipole Axilrod-Teller potential to study the structural and thermodynamic properties of liquid krypton. Evidence for many-body effects has been found in the 193,201,202,204,205 preferred structures and binding energies210 of rare gas crystals. In recent work, Donchev211 has studied the role of dispersion forces for cubic lattices using a coupled fluctuated dipole model (CFDM), where the particles have been treated as three- dimensional harmonic oscillators, coupled by the dipole-dipole potential. Deviations from pairwise properties have also been noted in studies of structure, dynamics, light scattering, IR and far-IR absorption spectrazn'220 of van der Waals 216-225 226-235 193,201-205 clusters, liquids, and solids. Computations and experiments on the spectroscopic properties of van der Waals trimers have specially inspired research in the field of many-body effects. Studies in this field include rotational spectra of ArzHF,236 6 237 238 9 240 l 242 Al'zDF,23 ArzHCl, ArzDCl, (HCN)3,23 ArzHCN, Ar2COz,24 Ar20CS, 243 243 NezKr, and Ne2Xe2. Far-IR intermolecular vibrations have been observed in 216 217 245 ArzHCl, ArgDCl, and in (1120),.218 Vibrations in the mid-IR in ArnHF,244 DF3, (HCN)3,,246 and (H20)3247 also show nonadditive effects. The role of nonadditive effects 27 in the properties of water trimer and liquid water has been analyzed by Gregory and 248 . 249 . . . . Clary, and by L1 et al. Collrsron—rnduced absorption spectra of compressed gases and liquids show evidence of nonadditive three-body dipoles,250 since the pairwise- additive dipole is not sufficient to explain the observed intensities of the transitions. Reddy, Xiang, and Varghese251 detected an absorption between 12 300 and 12 700 cm"1 for compressed H2. This absorption corresponds to the v = 0 —> 1 transition on all three of the molecules in an H2'"H2’"H2 complex. This particular transition is known as a triple transition252 and it is forbidden with pairwise additive dipoles. Early theoretical studies of nonadditive interactions were done using the point- multipole form. The Axilrod-Teller-Muto triple-dipole formulation was extended by Stogryn253 to evaluate the three-body dispersion energy of molecules with arbitrary symmetry. The energy denominator from third-order perturbation theory was separated using Buckingham’s method of evaluating the second-order dispersion energy and the dispersion energy was obtained in terms of the polarizability tensors, the mean polarizabilities of the molecules, and the dipole propagators. The method was applied to calculate the cohesive energies of molecular crystals. A similar method was used to calculate the third virial coefficients of C02 and N2, where the spherically symmetric component of the two-body potential was described by Lennard-Jones (12-6) or (18-6) potentials. Martin254 derived the three-body dipole moment of three spherical atoms from the fourth-order perturbed energy with the perturbing Hamiltonian expanded in terms of the spherical multipole moments of the interacting atoms. For H3, coefficients of the leading terms were calculated by diagonalizing the unperturbed H-atom matrix with s, p, 28 and d basis sets. The numerical values of these coefficients were then estimated for He3 . A similar method was followed by Gray and L0255 to evaluate the long range part of the three-body dipole moment of three interacting atoms and to estimate the density required to observe the collision induced infi'ared absorption in rare gases experimentally. The 256,257 multipole-moment expansion method has been used by Bruch and co-workers to study the three-body dipoles of interacting spherical atoms. In later work, Stogryn258 did a systematic analysis of the third-order perturbation energy for a system of N asymmetric molecules where the perturbing Hamiltonian was defined in terms of the multipole moment tensors, as used earlier by Kielich.259 The overall third-order energy was separated into induction, dispersion and induction-dispersion energies. The induction energy was further separated into two components: one component (W3) is linear in the hyperpolarizability and the other one (WA) is bilinear in the polarizabilities. The dispersion energy was separated into three parts: One involves the polarizabilities of the interacting molecules at imaginary frequencies (WD), one (WBA) contains the polarizability of one molecule and the hyperpolarizability of the other, and the last part (WCD) extends to asymmetric molecules of the result found by Chan and Dalgarno.260 Nonadditivity of the induction energy at second order was derived explicitly by Piecuch.261 In his work, the Rayleigh-Schrodinger perturbation theory was used to express the second-order correction to the energy in a system of N interacting molecules. The pairwise nonadditive induction energy was obtained as 29 A=_ E Z <1Vijlpj>‘ (138) . - E(pj) ’ ' 1,1,1?! Pirgj (kiln) where Stogryn’s notation258 has been used. i, j, k are the interacting molecules, )and lp) are respectively the unperturbed ground and excited states, and e (p j) = Epa —Ega . The interaction Hamiltonian Vij was then expressed in terms of the multipole expansion using a spherical tensor technique, to obtain N p I I! _ li+lj _li_lj—1 _lj—lk —1 A _ —41r. Z Z (—1) Rij Rjk 1,],k.=.1 Ii [j [j [k (109,1) 1j Lik Li,”- 21,421}- “2 2194.211c “2U ] :1 [1k 3" 2,1, 21k J L116 L119 J J J l,- +l'j l} +lk Lilg' X[[[Q1. -D’i(o:1)®Q, .n’k(m‘1)]L. ®a1 .{l'-1'-).D’f(n:1))L, , z 1 k k 1k 1 J J _] 1k) ®[Y1,- +1; (Bij)®Yl;- +1k (8 jlt)]1.,.kj 18- (1.39) Rij is the vector pointing from i to j and (Rijsfiij) are the spherical components of Rij in the global coordinate system fixed in space. Q11. is the spherical multipole moment of i, (oi,coj,o)k are the Euler angles describing orientations of local coordinate systems fixed in molecules i, j, k with respect to the local coordinate system fixed in space. D1 (or) is the matrix that represents a rotation to in the (2j + 1)-dimensional irreducible representation of the SO(3) group. aj denotes the irreducible spherical polarizability of 30 atom j and LikaLilg' are tensors which couple the molecular moments and the polarizabilities. Eq.( 1.39) gives the three-body induction energy at second order, where the dependence on orientation is simplified as far as possible and the overall energy is completely separated into the spherical multipole moments and irreducible spherical polarizabilities of the isolated atoms. In three successive later works, the same method was used by Piecuch to derive . . . 262 . . . 263 . . . . . 264 the 1nductron energres, dlspersron energres, and the isotropic interactlon energres at the first three orders of perturbation theory. In those works, the interaction energies at first, second and third orders were explicitly derived in terms of the spherical tensor formalism and the physical significance of the third-order interaction energies fiom Stogryn’s work258 were explained very clearly. The overall third-order interaction energy was given by «as 83:22.rare-Esta:wire where a is the polarizability tensor and B is the hyperpolarizability tensor. In Eq. (1.40) a right Q denotes the field from the permanent multipole moment of one molecule and a left Q denotes the permanent multipole moment of another molecule. The induction energies were farther separated into two-body, three-body and four-body interactions. The induction and dispersion energies from Eq. (1.40) correspond to the interaction energies in Stogryn’s work by, WB 5138);:1522, WA =E822g, WD Z-Eledésp , wBA=— _E(3l)3gsp, and WCD a Eggdisl’. 31 A complete description of the three-body interactions must include short range 222-225 0 interactions. These interactions can be included either by ab initio calculations r by using exchange-perturbation methods such as symmetry-adapted perturbation theory 265-271 (SAP'I) or intermolecular perturbation theory/Moller-Plesset perturbation theory 272-275 (IMPT/MPPT). A symmetry-adapted Rayleigh-Schrodinger perturbation method was used by Moszynski et al.271 to calculate the nonadditive three-body interaction energies of van der Waals trimers. The three-body terms from the polarization and exchange effects were separated and the polarization terms were evaluated using the linear and the quadratic polarization propagators.276 The three-body polarization terms were obtained as - . . (210) (210) . (1,1,1) . . 1nductron. Bind (B (— A,C), Ei n d (A <— B, B <— C), Bind (A (— B,C (— B), dispersion: E(3.’3)(3,3); and induction-dispersion: E (2 l 0) dlsp ' 1nd —disp‘ Nonadditive effects have been calculated by a IMPT/MPPT method developed by Chalasir’rki et al.272'275 Nonadditive dispersion interactions have been described within a reaction-field . 2.277 . 201 . , approach by Linder and Hoemschemyer, and by Langbern. In then works, a nonlocal response theory has been used to obtain the nonadditive dispersion energies in terms of frequency-dependent polarizabilities or susceptibilities. Li and Hunt278 have used a nonlocal polarizability density model to evaluate the three-body polarizations and three-body forces on nuclei of interacting molecules. The three-body polarizations and forces were determined from the three-body interaction energies at the third order. The overall third-order interaction energy was derived as a sum of three-body dispersion 32 energy, three-body induction energy, and three-body induction-dispersion energy. The induction energy was farther separated into hyperpolarization, static reaction field and third-body reaction field. It was also shown that the energies obtained using the nonlocal polarizability density method correspond the third-order interaction energies obtained by 261-264 Stogryn,258 Piecuch, and Moszynski et (11.271 In the present work, we have used the nonlocal response model used by Li and Hunt to treat the three-body and four-body interaction energies. However we have worked within a charge-density susceptibility based model to connect the interaction energies explicitly to a nonlocal dielectric model. In a later work, Li and Hunt”9 evaluated the nonadditive three-body dipoles of inert gas trimers and H2--~H2---H2 using the model based on nonlocal polarizability and hyperpolarizability. In chapter 2 of this work, we derive the two-body induction and dispersion energies at second order within the nonlocal dielectric model. These interaction energies have not been described previously in terms of intramolecular screening. Chapter 3 shows the dielectric screening present in the second-order induction and dispersion forces. We also derive a new fluctuation-correlation and the physical origin of the terms present in the second-order dispersion forces. In chapter 4 we extend the nonlocal dielectric model to derive the three-body induction energy at second order. We prove that at second order, the nonadditive three-body induction energy results from dielectric screening, where a particular molecule acts as the nonlocal dielectric medium to screen the electrostatic interaction between the other two molecules. Finally in chapter 5, we use the nonlocal dielectric model to derive the nonadditive three-body and four-body interactions and third and fourth orders. We specifically describe the induction energy at 33 third order and the dispersion and induction-dispersion energies at third and fourth orders. Chapter 6 includes a brief summary and conclusions. 34 Chapter 2: Nonlocal dielectric functions on the nanoscale: Screening of two-body induction and dispersion energies at second order 2.1 Dielectric screening and the induction energy Within quantum perturbation theory for intermolecular interactions, the energy changes due to static reaction fields determine the induction energy. The permanent charge distribution of each molecule sets up a field that polarizes the neighboring molecule; in turn, this produces a reaction field that acts back on the first molecule, shifting its energy. Thus the induction energy depends on the static, nonlocal polarizability densities ag‘B (r, r';0) and (12B (r, r';0) of interacting molecules, as shown in earlier work;131 AEind=—(1/2)jdrdr'aa B(r, r'- ,0)3B a(r)3B OB(r —(1/2) jdrdr'oaB(r-,r';0)30Aa(r)sA B(r), (2.1.1) to second order in the intermolecular interaction. In Eq. (2.1.1), 30130 (r) denotes the or component of the field acting on A due to the unperturbed, static charge distribution pg (r) of molecule B, and similarly for 38‘“ (r). The Einstein convention of summation over repeated Greek subscripts is followed in Eq. (2.1.1) and below. The result for AEind includes higher-multipole polarization, as well as the dipole-induced dipole interactions, because 01(r,r';03) is defined to allow for the distribution of polarization within the molecule, 35 o(r,r';to) = <0|f’(r)G((D)f’(r') 0)+(0|r(r')G(-o)r(r)|0). (2.1.2) In Eq. (2.1.2), P(r) is the polarization operator and G((t)) is the resolvent operator defined in Eq. (1.4). The polarization operator P(r) is related to the charge-density operator fi(r) by V-P(r) = —p(r), (2.1.3) and hence, from Eqs. (1.3), (2.1.2), and (2.1.3), VV':01(r,r';or) = — x(r,r';o)). (2.1.4) From Eqs. (1.5) and (2.1.2) -— (2.1.4), a(r,r';c0) functions as an integral kernel, to give the polarization P(r, (1)) at point r in a molecule by an applied field S(r'mr) acting at r’ . The field 3’8 (r) due to the unperturbed charge distribution pOA (r) of molecule - - - A _ ~A _ A __ A rs related to the electrostatrc potentral (00 (r,co — 0) by JO (r)——V(p0 (r,co—0). Below, we use the notation (06" (r) for the potential (p8 (r,(0 = 0) . Integration by parts in Eq. (2.1.1) gives the induction energy in terms of the charge-density susceptibilities of A and B, and the potentials (08‘ (r) and (pg)3 (r) , AEind =1/21drdr'x 0-. no 61,30»me +1/2 jdrdr'x (r,r';0)oOA(r)tp(A(r'). (2.1.5) With the potentials expressed in terms of the permanent charge densities of the two molecules, “1132,3001 AEind =1/2(1+5oAB)jdrdr'x (r,r';0)[(4ne0)"1jdr"r—r" 36 x[(41tso)_1 jdr'" r'—r"|_1p(l)3(r'")], (2.1.6) where 50 AB interchanges the molecule labels A and B. From Eqs. (1.2) and (2.1.6), with a change of the labels on the integration variables, it follows that the induction energy is accurately expressed in the dielectric model, by , _ /24 ‘1 ddrdnB '1 n, 17—!“1 B I 413m —1 ( nso) (1+sOAB){ j r r r 90(r)[808V,A(r,r .0)1|r rl pom — Idrdr'p(1)3(r)|r —-r' _1pOB(r')} , (2.1.7) where sv’ A (r, r'; 0) is the static, nonlocal dielectric function of molecule A. The first term in Eq. (2.1.7) gives the static Coulomb energy of the unperturbed charge distribution pg (r) of molecule B in presence of molecule A, which acts as a dielectric medium to screen the interactions within molecule B. The screening is nonlocal, since 8;} A (r, r';0) 9 depends on both r and r'. The second term in Eq. (2.1.7) is the single-molecule, static Coulomb energy of the permanent charge distribution of molecule B in absence of A. The operator 50 AB generates the corresponding terms that depend on the static Coulomb energy of A. Thus the induction energy depends on the difference between the dielectrically screened and unscreened interactions of the permanent charge distributions within each molecule. 37 2.2 Dielectric screening and the dispersion energy The dispersion energy AEd results from spontaneous, quantum-mechanical fluctuations of the charge density in each of the interacting molecules; these give rise to fluctuating fields acting on the neighboring molecules, polarizing it, and thus producing a dynamic reaction field, which acts on the original field source, shifting its energy. To derive the dispersion energy within the dielectric framework, we start from a standard expression for AEd , which is obtained both from time-independent perturbation theory and from reaction-field theory,2‘201 —1 AEd = — (h / 27t)_2 (47rso )—2 £0 dco Idr dr' dr' dr" x (r, r'; 1(1)) lr' — r" —1 X xB(r",r"';iw)lr" — r| (2.2.1) From Eq. (2.1.4) and integration by parts, Eq. (2.2.1) is equivalent to an expression for [Ed in terms of the nonlocal polarizability densities of A and B,129 AEd = — (h / 21:) go do) Idr dr'dr' dr'" aaB (r, r';ico) ToY (r', r" xaB (r" r"'i(0)T (r" r) (2 2 2) Y5 a 9 6a 9 9 ' ' where the tensor T(r,r') is the dipole propagator, Tao (r, r') = (47t80)—1Va Vo |r — r'|_1. (2.2.3) For molecules A and B interacting at long range, Eq. (2.2.2) reduces to the well-known 188 form, AEd = —(h/21t)TB.Y(R)T5a(R) If doo0AB(ior)oB5(ior), (2.2.4) 38 to leading order in R", where R is the vector from A to B and R = I R | . However, Eqs. (2.2.1) and (2.2.2) also include the effects of higher-multipole fluctuation correlations (beyond the dipole); so these equations give accurate results for the dispersion energies of nonoverlapping molecules, when the molecules can not be adequately approximated as point-polarizable multipoles - e. g. for large molecules in configurations such that typical intramolecular distances may exceed the shortest intermolecular distances. In order to show the connection to charge-density fluctuations at real frequencies explicitly, we reverse the steps of the derivation by Linder and Rabenold (Ref. 2), Eqs. (61) - (\67). We assume that the temperature is sufficiently low that the susceptibility densities xA (r, r';ico) and x3 (r',r"';i(0) change little over an interval of A0) = 27tikT/h on the imaginary axis; then the integral in Eq. (2.2.1) can be approximated by the discrete sum, w! AEd = — kT (41rso)-2 Z Idr dr’ dr' dr" x (r, r'; 21rinkT/h) n=0 —1 r'—r' x 1130",r"';27rinkT/h)lr""-r1—1 , (2.2-5) where the prime on the summation indicates that the n = 0 term is multiplied by 1/2. Equivalently, —1 r'—r" AE =(ih/41I)(47t8 )‘2 do) drdr'dr"dr'"xA(r,r';co) d 0 xxB(r",r";co)|r"—r|—1coth(h(D/2kT), (2.2.6) as shown by evaluating the frequency integral in Eq. (2.2.6) in the upper co half-plane, around a closed contour C that runs along the real 60 axis from o) = - W to co = - s (s > 0), clockwise around the small semicircle c0 = s exp(i0) from 0 = 7: to 0 = 0, along the real 39 axis from (0 = s to (.0 = W, and then counterclockwise around the large semicircle o) = W exp(i0) from 0 = 0 to 0 = it. In the limit as W -—> 00, the integral around the large semicircle vanishes, since both xA (r, r';o)) and xB (r',r"';co) fall off as (0—2 for large (0. The poles of the susceptibilities are located in the lower complex half-plane, by causality. Therefore, the only poles within the contour C are those of the hyperbolic cotangent function, and Eq. (2.2.6) is equivalent to Eq. (2.2.5) by the residue theorem. The susceptibilities have both real and imaginary parts, denoted by x' and x" , x(r,r';o)) = x'(r,r';m)+ix'(r,r’;(o). (2.2.7) The real part x'(r,r';or) is an even function of frequency, while the imaginary part x"(r,r';a)) is odd in 0). Since coth(hco/2kT) is odd in 00, Eq. (2.2.6) is also equivalent to AEd = -(h / 410(4480)‘2 (1 +59A13) 1: do) [or dr' dr" dr" xA" (r, r'; (1))lr' _ .-~ ‘1 x XB (r", r"; (0) Ir" - r|_1 coth(hco/2kT) . (2.2.8) As above, 50 AB perrnutes the labels A and B. l37,l38,2 By the fluctuation-dissipation theorem, the imaginary part of the charge- density susceptibility is related to the spectrum of charge-density fluctuations by — (h / 4n) xA" (r, r'; or) coth(ho)/2kT) = (1 / 2n) 1:” d(t —t') exp[—io(t-t)]<5pA(r,t)6pA(r',t')> , (2.2.9) 00 + where <8pA(r,t)8pA(r',t')> s <5pA(r, t) opA(r', t')+opA(r',t') opA (r, t)>. (2.2.10) + Therefore, 40 AEd = (1/8r)(4rtr.0)‘2 (1+ goAB) [:0 do jdr dr'dr'dr" Eda — t')exp[—io(t — t')] "'1 XB (r", rm; (1)) lrm _ rl‘l x <8pA (r, t) SpA (r', t')> |r' — r” + —2 I II M I A A II =1/4 4 1 d d d d dt—t 5 ,t5 ,t ()(7580) (+60AB)jrrrrE:( )(ptr)p(r)>+ I x [r —r"‘1xB(r',r";t-t') “‘1. (2.2.11) r" — r| Next, we use the relation between x(r, r';t - t') and 8;} (r, r';t - t’) in the time domain, ‘1 x(r",r';t-t'), (2.2.12) so 8;,1 (r, r';t — t') = S(r — r') 6(t — t') + (47rso)_l Idr" r — r” along with the Born symmetry [xB(r,r';t — t') = x3 (r',r;t — t')] , to obtain AEd = (1 / 4)(4rtso)‘1 (1 + goAB) Idr dr'dr" [:0 d(t - t’) + —1 I H ><|r —r so 833 (r, r’;t — t') 1‘1 . (2.2.13) _(1 / 4) (41t80)—1 (1+50 AB) jdrdr' |r—r + The first term shown explicitly in Eq. (2.2.13) gives the Coulomb energy associated with interactions of the fluctuating charge densities 5pA (r,t) and SpA (r',t’) , in the presence of molecule B, which acts as a dielectric medium with the nonlocal screening function 333 (r, r",t— t') , introduced by Jenkins and Hunt.1 A charge-density fluctuation at r',t' sets up a potential at r', t' (in the Coulomb gauge, with retardation neglected). Molecule B gives a screened potential at r,t , via 83,3 (r, r",t—t') , and the screened potential affects the energy of a charge-density fluctuation at r,t. The response is integrated over 41 all “time lags” t— t' , but s;1B(r,r',t —t') = 0 if t— t'<0 , so the response is causal. The second term in Eq. (2.2.13) gives the Coulomb energy of associated with the charge- density fluctuations in molecule A, in the absence of molecule B. The operator 50 AB generates the corresponding term, in which molecule A acts as a dielectric medium for fluctuating charge interactions in B. The Coulomb energy of interaction between the charge-density fluctuations in the same molecule equals the intramolecular exchange-correlation energy in density functional theory (after the self-energy has been removed). With Eq. (2.2.13), this implies that - for molecules with weak or negligible charge overlap — the dispersion energy is equal to the screening-induced change in the intramolecular exchange-correlation energy, summed for the two molecules. A different dielectric function sd (r, r';i00) at imaginary frequencies183 is directly related to the correlation of the polarization fluctuations at r and r'. On intramolecular scale, sd (r, r';60) is distinct from sv(r,r';(o). Thus, sv(r,r';io)) does not relate directly to the charge density fluctuations. However, Eq. (2.2.13) proves that sv(r,r';t —t') and hence sv (r, r';(o) is directly related to the screening of the correlations of the intramolecular charge density fluctuations. Since the correlations between the permanent charge density and the fluctuating charge density vanish for each molecule, there is no net Coulomb energy associated with the interactions between po (r) and 8p(r’, t) . Hence, in the region of negligible overlap, Eqs. (2.1.7) and (2.2.13) give the energy to second order, in a dielectric framework; and the results are consistent with quantum perturbation theory. 42 Chapter 3: Dielectric screening of second-order induction and dispersion forces on nuclei of interacting molecules In this chapter, we express the second-order induction and dispersion forces on the nuclei of interacting molecules within the dielectric framework. The force on nucleus K in molecule A is determined by the negative gradient of the interaction energy of the molecule with respect to the coordinate RK of nucleus K. FK =—VK AE, (3.1) where VK denotes differentiation with respect to RK. In the following sections we use Eq. (3.1) to derive the induction and dispersion forces on the nuclei. Throughout the derivations, we use the Bom—Oppenheimer approximation: The forces on the nuclei are determined as functions of the nuclear coordinates, fixed within individual calculations but not restricted to the equilibrium configuration. 3.1 Dielectric screening and the second-order induction forces on nuclei The induction force Filrid on nucleus K in molecule A is given by the derivative of the induction energy with respect to coordinate RK of nucleus K, FIE = —VK AEind- From Eq. (2.1.1) and the Born symmetry280 of the polarizability density 01A (r r" 0) = 01A (r' r'O) we obtain GB 9 9 GB 9 9 9 FiEd = 1/ 2 Idr dr'aaaAB (r, r';0) / 6RK 3861 (r) 8013“,) 43 + jdrdr’aaB(r,r';0)630Aa(r)laRK sgBu'). (3.1.1) The derivative of the polarizability density of molecule A with respect to the nuclear coordinate depends on the nonlocal hyperpolarizability density B(r,r',r";co,0) of A,127’128 ado/AB (r, r';o)/aR1Y< = zK Idr'Bé‘BSU, r',r";(t), 0) T57 (r", RK), (3.1.2) where T(r”,RK) is the dipole propagator defined in Eq. (2.2.3) and the hYPCIPOIarizability density is given byl30,134 Momma)133(r")G+ (3.2.2) The derivative of the real part of xA (r, r';:o) with respect to RK is connected to the real part of quadratic charge-density susceptibility via the relation axA' (r, r'; o) / aRK = -zK (47:.s0)‘1 50 . -3 r1v _ RKI x Idriv Re CA (r, r',riv;m, O) (RK —riv) (3.2.3) Eq. (3.2.3) follows from the contraction of Eq. (3.1.2) with Va VB, and integration by parts. From Eqs. (3.2.2) and (3.2.3), with 8o(r,t) used to denote the potential acting on A, due to the fluctuations 6pB(r,t) in the charge density of molecule B (and neglecting retardation effects), K _ K -1 , . A Fd(l)—(1/87:)Z (471280) foodo Idrdr dr Reg (r,r,r,o,0) —3 x(RK —r") r"—RKI x E; d(t — t') exp[ — i a) (t — t') ] <8th (r', t) SoB (r, t')>+ . (32.4) The quantum mechanical average of S(pB(r',t) vanishes; however, the average of the product 8oB(r',t) SoB (r,t') is nonvanishing, because the charge-density fluctuations that give rise to the potentials are correlated. The Fourier transform of the correlation function <5ch (r',t) SoB (r,t')> in Eq. + (3.2.4)is B r r B n _ r B r B - r . n <5(p (r ,(o)5(p (r,o) )>+ — foodt Eodt <6q> (r,t)6:p (r,t')>+ exp(ro)t+1o) t'). (3.2.5) Since <8¢B (r',t) SoB (r,t’)> is a function only of the time interval t— t' , after changing + the variables of integration in Eq. (3.2.5) to t— t' and t' , we obtain 51 B I I B II _ _ I B I _ I B I: <5cp (r,o))6(p (no) )>+ — food“ t') Eodt <6cp (r,t t)5(p (r,t 0)>+ xexp[i(o'(t—t')+i((o'+ 60")t'] = 27: <8 8(or'+(o'). (3.2.6) + In Eq. (3.2.4), we express <6qu (r',t) 5(pB(r,t')> as the inverse Fourier transform of its + Fourier transform, to obtain Ffa) = (1 / 87:) (l / 27:)2 ZK (47:so )-1 1:0 do) Idr dr' dr' Re CA (r, r', r"; (o, 0) -—3 x(RK -r") r"-RKI x fwda—t') exp[-iw(t-t')] x E; dd [:0 do)" <8th (r', 00') 6(0B(r, 00">+ exp(— i 01' t — i 0)” t') , (3.2.7) which is identical to F+ (3.2.8) Next, we evaluate the t — t' integral: 52 Fle) = (1 / 81:)(1/27t) 2K (47:80)_1 x E; do) [:0 do)’ E000 do)" [dr' dr' dr' Re CA (r, r', 1'"; 03,0) -3 x(RK -—r") r"—RK‘ x 8(a) + 03') exp [- i ((1)’ + (0") t' ] <8q>B (r', (0') S(pB (r, (1)")>+ . (3.2.9) After evaluating the co' integral, use Eq. (3.2.6) for <6 , to obtain + F31?” = (1 /81t) zK (mm—1 x E; dco ED do)” Idr dr' dr' Re CA (r, r', r"; o), 0) —3 x(RK —r") r'—RKI x exp[—i (0)” —w) t' ] <5ch(r', —o)) 5¢B(r, t' =0)> 8(0)" —co) . (3.2.10) + Then we evaluate the a)" integral, which gives F3140) = (1 /81t) zK (47mo)_1 x fwdm Idrdr'dr” Re§A(r,r',r";w,0) —3 x(RK —r") r'— RKI <6 (3.2.11) +. The real part of the quadratic charge-density susceptibility, Re CA (r, r', r';a), 0) , has the permutation symmetry, 53 Re CA (r, r', r'; 0), 0) = Re CA (r', r, r'; —(o, (1)) , (3.2.12) and it is even in a); so Re §A(r,r',r';(o,0) = Re §A(r', r, r'; 0), —a)). (3.2.13) Therefore FED = (1 / 81:)ZK (mm-1 x fooda) jdrdr'dr'RegA(r",r,r';m.—m) -3 ><(RK —r") r' — RK| <8ch(r',—(o)5(pB(r,t’=O)> . (3.2.14) + We insert an integration over (0' , using a delta function, FED = (1 /87t)ZK (mm-1 x foo dco foo dw' Idr dr' dr' Re CA (r", r, r'; (10', —o)) 6((o'-o)) —3 x(RK -r") r"-RK| <5ch(r',—w)8(pB(r,t'=O)> . (3.2.15) + From Eq. (3.2.6), 2n <6 5(0)’—(o) = <6 , (3.2.16) + + from which we obtain FK d(l) =(1/167t2)ZK(41t80)—1 x foodo) Eodm' Idrdr'dr' Re CA (r',r,r';o)',—(o) 54 -3 x (RK — r") r" — RKI <8 , (3.2.17) + or equivalently, FK = (1 /161t2)ZK (mm—1 d(l) x I: do) [:0 do)’ Idr dr' dr" Re CA (1", r, r'; 03’, (o) x ( VK r' — RK [—3 )<8+. (3.2.18) In a form that makes the dielectric screening interpretation clear, F360) = —(1 / 161t2) ZK 20 V” [:0 do) Eoda)’ Idr dr' dr' 8211’ A (r', r, r'; a), co') x<5Jr r' = RK . (3.2.19) Eq. (3.2.19) shows that the first term in the dispersion force on nucleus K in molecule A results from the nonlinear dielectric screening of the correlated fluctuations in potential due to molecule B. This component of the force is analogous to the component of induction force that results from quadratic screening of the static potential due to molecule B. The remaining component of the dispersion force on a nucleus in molecule A is given by the second term in Eq. (3.2.1), denoted by Fdé2) , with —1 = (h / 4112)(47t80)_2 VK foo do) Idr dr’ dr' dr'" xA" (r, r'; w) r' — r" K Fd(2) xxB'(r”,r";a))|r"'—r|—1coth(hw/2kT). (3.2.20) 55 VKXA” In appendix A, we derive a relation that connects (r, r' ,w) to the 1maginary part of the quadratic charge-density susceptibility, CA"(r, r', r'; (o, 0) , vK xA"(r,r';0~)) = (47:80)-1 Jldrn 2K —1 II r"—RK| §A(r,r',r',m,0), (3.2.21) To show the dielectric screening present in F31?” , we take Eq. (3.2.21) and separate the terms with n = j and n ¢ j. We obtain 1211(2): (it/h 2)(h/47t)(47t80)_3 jdrdr'dr'dr"dr“’ 2K r"—RKl x12 2 < ‘ ">j>w355(wno—w) n¢0j¢0,j¢n +2 2 (0|P(")|J >J<|f>(r")ln>ln '>j><1‘">0>w3:,a n¢0j¢0j¢n +2 2 (OIPWIHX ">110" 30> n¢0j¢0j¢n x Re [(033%) -iF ° / 2 — (0)4 ]5(o)n0 — (o) J -2 Z ln >ljf>< ">0>w3})6 n¢0j¢0,j¢n -Z Z <0|fi(r'>ln>|0> n¢0j¢0,j¢n 56 xRe[(m3%) +iFj /2+m)’1]5(mn0 + (1)) -ZZ< n¢0 j¢0,i¢n -Z Z <0Ipl>< >ln>I0> >j>j<|p|0)w3},6(cono+w) n¢0 j¢0,j¢n x Re [(033%) + iFj /2 + m)—1]8(mn0 + (n) + Z < >(Dnlo ammo -co) +2 (0|13|nn>< |P(r' n¢0 )nn>< n) (n | p(r')-p00(r') l n) (n | (S(r) | O>co;110 8((on0 + w) -Z<“') n¢0 O) (1)510 S(wno + (o) ' Z (0113031“>|0> + Z (015(r>|n>("')-Poo ") n>0°>|0> n¢0 >< Re [(00110 —co)"1 6(wno —w) )n> 0> - Z <0|fi(r>|n>|0> n¢0 x Re 1(wno +w)"1 S(wno +co)} . . —1 xIr—r"|‘1xB(r",r",m) r" —r' coth(h(o/2kT). (3.2.22) The first eight terms in Eq. (3.2.22) can be described as correlation between the charge- density fluctuations and the susceptibility fluctuations in molecule A. To show this correlation explicitly, we define a transition susceptibility of molecule A, 1300', (01;r', (1)2) following first-order transition hyperpolarizability defined by Hanna, Yuratich, and Cotter in Eq. (2.19) of Ref. (281). In particular, we need only the real part of the transition susceptibility; assuming for simplicity that the states of A are real, we have élj>Re(coJ-o—iFj/2—m2)" Rexn0(r,co1;r',(o2) =(1/h) Z[(n j¢0 + (n|p(r')|j)= Z Homo-">1j>ln>w3$ j¢OJ¢n +Re(o)j0-iFj/2—w)'1]. (3.2.24) Thus, the first eight terms in Eq. (3.2.22) give the correlation between a charge-density fluctuation at r and a susceptibility fluctuation at r' (and vice versa), within molecule A. To illustrate the quantum mechanical nature of this fluctuation, we consider the 58 correlation between a susceptibility fluctuation at r with frequency —(0 and a charge- density fluctuation at r' with frequency (1). In the limit T—+0, the fluctuation-correlation is given by (1 / 2)( 6x(r', 0;r. — (0)5962 co) >+ =(1/2)< 6x(r", O;r, —m)5p(r', 0)) + apl (r', (0)5)(1' (r, — co;r”, 0)> . (3.2.25) Using the facts that 8xl(r,w;r",0) = 8x(r,—o);r",0) , (3.2.26) and 5pl(r,m) = 5p(r,—co), (3.2.27) we obtain (1 / 2)< 5x(r"‘,0;r, -o)) 6p(r', (1)) )+ = (1 / 2) [ mu (1", 0; r, —w) pno (r') 5(wno + w) + 9011 (r') xno (r, w; r", 0) 5(w0n + (0)] = (1 / 2) “Mn (r', O; r, —co) PnO (r') S(wno + (o) + POn (r') XnO (r, (o; r", O) 5(w-mn0)] = (l / 2) ”On (r', O; r, —m) PnO (r') S(mno + w) + POn (r') XnO (r, a); r", 0) S(mno — 03)] . (3.2.28) From Eq. (3.2.28), we can describe the first eight terms in Fd%2) (which we denote as Fézm below, for convenience) as . —1 1&2” = (1 / 4) (mm—3 Eodco jdr dr' dr'dr’" drdr'v zK VK r" — RKI x [( 8x(r",0;r, —(1)) 5p(r', (a) )3 +< 5x(r',0;r', —(1)) 8p(r, (1)) >3 ] 59 . . —l xlr—r"|—1xB(r",rw;w) er —r' (3.2.29) Eq. (3.2.29) can also be represented as F(Ii<(2),1 = (1 / 4) (4ne0 )_3 E; dco' f; d0) Idr dr' dr" dr‘" drdrIv —1 sz VK r'-—RKl 8((o+o)') x[( 5x(r",0;r, (0') 5p(r', 03) >5 +< 8x(r",0;r', (0') 8p(r, (1)) >3 ] . . —l xlr —r'"|_1xB(r",rw;o3) rlv —r' (3.2.30) Eq. (3.2.30) proves the fact that when a nucleus shifts in the molecule, the change in the nuclear Coulomb field due to the position shift brings in new correlations within the molecule. When nucleus K in molecule A shifts infinitesimally, it changes the static Coulomb field, given by 30.6") .—. (4ne0)‘1 2K Tag (r”,RK)5R[I3< . (3.2.31) Previously, Liang and Hunt'32 noted that the change in the nuclear Coulomb field may introduce new types of fluctuation correlations in the molecules, as well as altering the magnitude of the correlations. Eq. (3.2.29) establishes the fact that the shift in the position of nucleus K does bring new fluctuation correlations within the molecule, namely the correlation between charge-density fluctuation and susceptibility fluctuation. In absence of any external field, the charge-density fluctuations are correlated by the imaginary part of the charge-density susceptibility, x"(r,r',(o). The nuclear Coulomb field alters that fluctuation, brings in new fluctuations, and also introduces a new 60 correlation function, §"(r,r',r';co,0). From Eqs. (3.2.21), (3.2.22), and (3.2.30), we can relate the correlation between the charge-density fluctuation and the susceptibility fluctuation as (1 / 2)[< 8x(r',0;r,w') 5p(r',w) >3 +< 5x(r",0;r,co') 5p(r, (0) >2 ] = (h / 211:) 63:11 0(r,r’,r";co,0) 8((o+ (0') coth(ho)/2kT), (3.2.32) where 63:21 O(r,r",r";(1),0) means that only the terms with j i n,0 of CA. (r, r', r';tn, 0) determine the correlation. Next, we show the dielectric screening present in Fcll((2)1' The spontaneous charge-density fluctuation in molecule A gives rise to a perturbing potential that acts on molecule B, shifting the charge density in B and therefore producing a fluctuating reaction potential 5680, m) that acts on A: 8ch (r, (n) = (47:30)_2 Idr.’ dr' dr'" |r — r' —1 x3 (r', r"; w) r" — r"'|—1 5pA (r'", (1)). (3.2.33) Using Eq. (3.2.32), we can simplify Eq. (3.2.29) to —1 Fd((2),1 = (1 / 4) (411280 )—1 foo do) Eodm’ Idr dr" ZK VK r” — RK| x 6(0) + co')< SxA (r", O; r, (0') S(pB (r, m)> + —1 —1 I I a K K n K +(l/4)(41t80) foodm Eodco Idr dr Z V r —R l x 5(0) + co')< SXA (r", 0; r', (0') S(pB (r', m)> . (3.2.34) + 61 The fluctuations in the susceptibility SXA (r”, 0; r, to) correspond to fluctuations in the dielectric function 68;1A(r”,0;r,co), with the same relation as in Eq. (1.2); thus we obtain FdK(2)1 — —(1/4)a/ar' Eldon) Eodco' jdr" x 50<83 3',"A(r ,O;r, (0') S(pB (r, (o)> r"_15((o+co') +r.=RK —(1/4)a/ar' Eodo) 2K < 8 (3.2.35) + r! = RK Thus, FcIl((2)1 in the second component of the dispersion force on a nucleus in molecule A comes from the dielectrically screened dispersion potential, due to the change in the charge density of B induced by the spontaneous fluctuations in A. In this case, the average of the dispersion potential from molecule B vanishes; however, the fluctuations in the dielectric screening function are correlated with the fluctuations that give rise to the dispersion potential. Hence, the screened field vanishes, but the screening effect does not. The remaining terms in the second component of the dispersion force can not be described in the dielectric framework, because they do not stem from the field induced fluctuation correlations described above. These remaining terms are not related to a nonlocal response function of molecule A. In the next part we explain the physical origin of all the terms present in 1736(2) using time-dependent perturbation theory. We use the fact that a charge-density fluctuation in A at r',t' creates a potential in B at r'v,t' , which induces a shift in the charge density in B at r'",t. The induced shift in the charge 62 density in B creates a reaction potential on A at r, t. This reaction potential acts as an external time-dependent perturbation and perturbs the ground and the excited states of molecule A. That results in field induced transitions and perturbed transition charge density in molecule A. Below we show a systematic analysis of these effects and relate them to the second component of the dispersion force. The interaction between the charge-density of A, and the reaction potential from B is given by the Hamiltonian 121(1)“) = jdr p(r) -6(pB(r,t), (3.2.36) where {n(r) denotes the charge-density operator for molecule A. Using standard time- dependent perturbation theory, the ground and excited state wave functions of A to first order in the applied potential are given by lw1(t)) = In)CXP[-iwn t1 +(1/ih) fwdt' Idr{Z(j|p(r)|n)8(pB(r;t')exp[iwjn t']} |j>exp[-i(oj t] , l (3.2.37) and |wr(t)>=l0>eXP[-iwot] +(1/ih) foodt' Idr{zexp[-i wj t], 1 (3.2.38) 63 respectively. The average value of the charge density at r",t between the perturbed ground and excited states in molecule A is 8p(r",t) = (wf(t)|p(r") wi(t)>+cc , where cc means the complex conjugate of the expression. From Eqs. (3.2.3 7) and (3.2.38), 6pm» = (Olav) n)exp[i(1)0n t] +(1/ih) fwdt' jdr{2(o|8(r') j j> j x 8¢B(r;t')exp[im0j t']}exp[i wjn t]. (3.2.39) In Eq. (3.2.39), the first term represents the unperturbed charge-density fluctuations between the ground and the excited states at r’,t. The average of this unperturbed fluctuation vanishes. The second and the third term are the transition charge densities between the ground and the excited state, perturbed by the reaction potential. We call these terms the “first-order transition charge density”, Spgln) (r", t) between the ground state [0) and the excited state In). Using Eq. (3.2.29) along with a Fourier transform of the potential to the frequency domain, we obtain 8pgln)(r";t) = (1/2ih) fiat.) foodt' jer(o|,3(r') l j> x {8ch (r; (o) exp[i (mjn + (n) t' ] + 863* (r; (1)) exp{ i ((Djn — (o) t' ] } exp[i cooj t] 64 —(1/2ih)£:da)foodt'Ier J x{((1)jn +(0)’1exp[i0)t]5(pB(r;(o) + ((0 jn - (0)'1 exp [-i o) t] 663* (r; (0)} exp [i wOn t] +(1/2h) 1:08..) Idrz(o|,s(r)| 001800111) J x {((ooj +03)"1 exp[i (0t]8(pB (r; (0) + (we) -8)-1 exp 1-i cot] 843*0; m) } expfiwOn 1] = (-1/2h) fwdmfidco’ jer(0|;3(r')|j)(j|§(r)|n) j X{((°jn +(0)'1exp[io)t]6(pB(r;0)) + ((0jn -0))'1exp[-imt]5(pB*(r;(o)}exp[ i0)’t]8((00n —(1)') +(1/2h)food(0f;d(0' Ier8((0-o)no). (3.2.41) From Eqs. (3.2.40) and (3.2.41), the first order transition charge-density at r',t in molecule A is given by 1 I! _ I I M . 8.581301): —(1/2h)(41t30) 2 fiodw [:dm jdrdr dr dr" j>{((0jn +0))_1exp[i((0+0)') t]8((0n0 -0))6((00n —0)') XZ<0lf><§jl< ><|jp(r)|n n>{((0jn+(0) 1exp[i(()1)+(0')—t]5((0no (0)5((0+(0') +((0jn - (0)'1exp [i ((0 — (0') t] 8((0n0 + (0) 8(0) - (0')} (n xlr_rIII|-1xB(rlI,riv; rlv _rI ) O) +(1/2h)(4n.e.0)‘2 Eodo) Edd Idrdr'dr’”driv n>{((00j +(0 )_1exp[i( (0+(0' )t]8((0n0 —c0)5((0+(0') x2018» 1001 w j +((00j — (0)'1Exp[i ((0 — (0') t] 8((0n0 + (0) 8((0 - (0')} (n rlv _ rI ) 0). (3.2.43) X lr_rlfl‘_1 XB (rm, riv; Eq. (3.2.43) shows that although the reaction potential acting on A is time-dependent, the perturbed first-order charge density in A is time—independent. From Eq. (3.2.43), we collect the terms with j i n, 0. Then integrating over (0' and rearranging the terms, we obtain 8p(1n)(r" ;—t) — —[(1/2h)(47r80)_2 l: d(0 jdrdr' dr'dr'v x Z [{(O A "Hill IP(')| “(03-0) +<01P(r)| JXJ l6(r")ln) j¢0,n x((0j0 —(0)'1}5((0n0 —(0) +{< ")J'J'>< |13(r)In 10°35 +<01P(")| >< |P(r')ln> 67 +(0|13‘2 1:04P ldrdr'dr"driv [(PIPP'”) 00113009830 x { 8(6),,0 - (0) + S(wno + (0)} -(0 P(r') 1‘) (”I110 x { 5((0n0 — (0) + S(mno + (0)} ]Ir — r'"I_1 xB(r", riv — (n I p(r') I rw —l‘ 0 ;m) = _(1/2}‘1)(47:g0)_2 fiodm Idr dr'dr" driv [<0lf3(l’") n)(nl13(r)-Poo(r)ln) x { 5((0n0 —- (0) + S(wno + (0)} Ir - r"'I—1 1B (r', rIv ; (0) iv - xr —r' I0) due to interaction with an external time-dependent Coulomb potential is given by c“) (t) = —(1/2h) 1:066 Idrdr'dr'"driv {(0 In)-—>IO) P(rJI n>(w0n + w)" 0)(01)()n — (0)-l >< eXP[i((n I130“) x exp [i ((00n - (0) t] (sz (r, (0)} . (3.2.48) Using the reaction potential 8680', (0) in Eq. (3.2.48), we obtain CIPI->I O>(t) = —(1/2h)(47t80)_2 [lam Idrdr'dr" driv {<0Ip(r)In)(t). (3.2.50) n¢0 These four terms are not connected to the dielectric model, since they do not originate from the response of A to the perturbing potential. Eqs. (3.2.45), (3.2.46), and (3.2.50) show the physical significance of the second component of dispersion force. We have proved the origin of the terms using perturbation theory. Although the external perturbation is time-dependent, the dispersion force does not show time-dependent behavior, which is the exact same result obtained using reaction field theory?"129 Finally in this chapter we discuss the apparent disagreement between the signs of the terms from perturbation theory and the terms we obtain from the derivative of the imaginary part of the charge-density susceptibility. The hyperbolic cotangent function in the dispersion energy appears from the fact that when we consider the charge-density fluctuations at a finite temperature T, we need to use the canonical distribution of the 138,282 molecular eigenstates. The ratio of the spectrum of the charge-density fluctuations and the imaginary part of the charge-density susceptibility yields the hyperbolic cotangent function in the dispersion energy. In the formulation described above, we have 71 used zero-temperature fluctuations. In the limit T -—> 0, coth(h(0/2kT) —> [0((0) -0( - (0)] , where 0((0) is the Heaviside step function.278 Thus in the limit T -> 0, Eq. (3.2.22) simplifies to . —1 Fd%2)_ _ (n/hz)(h/4n)(4na0)‘3 Idrdr'dr"dr"drw zK VK r'—RKI x{ Z Z .. 1') j>1><1p1n>1o> n¢0 j¢0,j¢n x Re [(003%) — iFj / 2 — 0))_1]8((0n0 — (0) +2 2 <0|Pln>< ">00” )0) n¢0j¢0j¢n x Re[((03I) —iFj /2 -(0)_1]6((0n0 —(0) -Z Z < 100180111X W>w388 n¢0j¢0,j¢n -2 2 <0” '>|n>lO> n¢0j¢0,j¢n xRe[((63I) +il"j /2+(D)_1]5((Dn0 +0)) -2 Z < ‘ ">1><1°113(r')|n>63},S(mno+0) n¢0j¢0,j¢n 72. -Z Z <0|P n¢0 j¢0,i¢n j)(jI13(r”) n) xRe[((63I) + i1“ 1- /2 + (0)-1]5((0n0 + (0) n) (n I p(r)-p00 (r) I n) (n I p(r') I 0) (0510 8((0nO — 0)) + Z <0113 natO + Z (011mln>-poo1n> n>-poo0186101411..56.0+0» n¢0 - Z <01131n>-poo(r>1n>|0> x Re [((0n0 + (6)1] 5(6),,0 +0)) - Z <0|P(r')|n>C)][ zK(4ne0) 2 Idrdr"dr"'APll,3 x xA (r', r"; 0) VaK r"—RKI +ZK(47t80) 1IdrAPB B(r)VKVB r— RKI—l ]. (4.210) The operator C ( B —> C ) in Eq. (4.2.10) means replacing the molecular label B by C in the expression that follows. The potential acting on molecule A is the sum of the potential due to the unperturbed charge distributions of B and C, and the potential due to the shift in the charge density of B, induced by the potential due to the unperturbed charge distribution in C (and similarly the potential due to the shift in the charge density of C, caused by the potential from B). Thus from Eq. (4.2.10), the three-body induction force on nucleus K depends quadratically on the potentials from the permanent charge distributions in B and C and linearly on the potential, A0)? (r) , due to the first-order shift in the charge distribution in B, ApI3 (r’) (and similarly for C). Using A(pI3 (r) from Eq. (3.1.10) in Eq. (4.2. 10) and by repeated use of divergence theorem, we obtain FK(2,3) in d,a = ZK (47mg).1 Idr dr' dr"[VI3 V7 V3 BIPKY 6 (r, r',r"; 0, 0)] 636056-68 r”—RKI 82 —1 —[1+C(B—+C)][ZK (41(80)—1 Idrdr'x (r,r';0)VK r—RKI xA¢P(r',0) +sz§[A(plB(RK,0)]], (4.2.11) where we have used the Born symmetry of the charge-density susceptibility xA (r,r';0) with respect to an interchange of its arguments. Next, we use the relation between the nonlocal hyperpolarizability density [3&5 (r, r', r"; 0,0) and the quadratic charge-density susceptibility C(r,r',r';0,0) from Eq. (3.1 . 12) and a relabeling of the integration variables, to obtain FK(2,3) = ind, a — ZK (47t80 )-1 Idr dr' dr” CA (r, r', r'; O, O) x (pg)3 (r) (0% (r') (3 / are, Ir" — rI—1 r=RK —[1+C(B—>C)][ZK (mm—1 Idr'dr"x (r',r;O) r—RKI A(pB(r',0) —zK VaK[A(pI3(RK,0)]] (4.2.12) Using Eqs. (1.1), (1.10), and (4.2.12), the second-order three-body induction force on nucleus K in molecule A is given by 1K(2,3) z 1nd,a v,A K —[1+C(B—>C)]ZK 806/6ra[_Idr'8_ (r,r';0)A(pI3(r',0)] r 83 _ ZK 80 ('3 / (3ra[ Idr' dr' 8:1,1A (r, r', r”; 0, 0)] xeguzoxpgozon RK. , _ (4.2.13) Equation (30) shows that the second-order three-body induction force on a nucleus in molecule A results from screening of the potentials from neighboring molecules B and C; the first-order potential due to the induced shift in the charge density in B (or in C) is screened linearly within A , while the unperturbed potentials from B and C are screened quadratically. At second order, the eflective three-body potential within molecule A is given by (p:f§2,3) (r, 0) = [1 + C ( B —) C ) ] 80 Idr' 8;,1A (r, r'; 0) A(pI3 (r', 0) + 80 Idr'dr" a; 1 (r, r', r";0, 0) (0(1)3 (r’, 0) (pg (r', 0) . (4.2.14) Thus, using Eqs. (4.2.13) and (4.2.14), the second-order three-body induction force on nucleus K is K(2,3) _ Find . — A(2,3) —zK a/ar[quff (r,0)] (4.2.15) r = RK . Eqs. (4.2.13) and (4.2.15) prove the fact that the three-body induction force on nucleus K in molecule A at second order can be exactly described by the dielectric screening model. 84 Chapter 5: Dielectric screening of three-body and four-body interactions at third and fourth orders In this chapter, we derive the three-body and four-body intermolecular interaction energies at third and fourth order within the dielectric model. We show that the results are in agreement with the results from quantum perturbation theory. In chapter 4, we have showed that at second order, the three-body induction energy results from dielectric screening of the Coulomb interactions between the permanent charge densities of two molecules and the screening arises due to the presence of a third molecule which acts as the dielectric medium. In the present chapter, we prove that at third order the induction energy results from either intermolecular or intramolecular screening, depending on the type of interaction. Moreover, nonlinearity appears in the induction energy at third order, resulting quadratic response and nonlinear screening. At second order, the interaction energies show linear screening only. Nonadditivity in dispersion energy first appears at third order. In chapter 2, we have proved that the second-order dispersion energy results due to the dielectric screening of the intramolecular exchange-correlation energy. We show that at third and fourth orders, the dispersion energy still appears due to screening-induced change in the intramolecular exchange-correlation energy. However, at third and fourth order, the dielectric medium consists of two and three molecules respectively and that brings many- body effects in the screening function. We derive the many-body dielectric functions from the many-body susceptibilities of the interacting molecules and we describe the screening of the dispersion energy at third and fourth orders in terms of these dielectric 85 functions. We also prove that the dispersion energy at third and fourth orders results from screening of the dispersion energy at second and third orders respectively, and that screening appears due to the presence of a third or a fourth molecule which acts as the dielectric medium. The third category of interaction that appears at third and fourth orders is the induction-dispersion. It results form the perturbation of the dispersion energy by a static external field. The external field perturbs the response function of the molecules and brings in new type of fluctuation correlations. Dispersion energy shows linear screening only, but the perturbation by an external field produces nonlinear response, and hence nonlinear screening in the induction-dispersion energy. At third order, induction- dispersion energy includes nonlinear screening only. At fourth order, both linear and nonlinear screenings appear. We prove that the screenings present in the interaction energies at third and fourth order are described by the nonlocal dielectric functions introduced in chapters 2 and 3. 86 5.1 Dielectric screening of the induction energy at third order In this section, we show that at third order, the three- and four-body nonadditive induction energies are described within the nonlocal dielectric model. We work within the third order of perturbation theory and express the nonadditive induction energies in terms of the static nonlocal charge-density susceptibilities of the interacting molecules. Then we relate them to the nonlocal dielectric functions 8V(r,r';0) and 8q(r,r’, r'; 0,0) introduced earlier, in order to show the dielectric screening. Depending upon the type of interaction (and hence the molecular excitation pattern), both linear and nonlinear responses contribute to the induction energy at third order. In presence of a dielectric medium, the interaction energy of two test charges is screened. The shift in the interaction energy caused by a nonlocal dielectric medium with linear screening is described by Eq. (4.1.1) in the last chapter, and we have proved that Eq. (4.1.1) accurately describes the dielectric screening in the induction energy at second order. In the present section, we prove that the same equation still applies for the third- order induction energy. However, at second order the interacting test charges correspond only to the unperturbed, permanent charge distributions of the molecules. At third order, p(r) in Eq. (4.1.1) can be either the permanent charge distribution of a molecule, or the induced shift in the charge density of one molecule cause by the permanent charge density of another molecule. Depending on whether the two interacting charge distributions are permanent or induced shifts caused by an applied potential, the interaction can be categorized as a particular many-body type (i.e. three-body, four-body etc.). 87 Nonlinear screening appears at the third-order induction energy. In presence of a nonlinear dielectric medium characterized by the quadratic dielectric firnction sq (r, r', r'; 0, 0) , the interaction energy of three charge distributions is given by AB = (41(80)—2 Idrdr' dr' dr‘" drivp(r)[41t8q (r, 1", l‘";0I 0) l—1 x r" — l’Iv r'-r"'I_1 p(r'") p(riv). (5.1.1) At third order, the charge distributions in Eq. (5.1.1) correspond to the permanent charge densities of the interacting molecules. Unlike the screening caused by the linear response of the dielectric medium, nonlinear screening does not stem from the screening of a lower-order interaction. This is because, at different orders of perturbation theory the interactions arising purely due to the nonlinear response are characterized by response functions of different orders and they can not be interrelated to each other. From intermolecular perturbation theory, the third-order energy for a cluster of molecules AmBmCmD of arbitrary symmetry, interacting at long range is given by (Em — E0)(En — E0) AE(3) = Z Z m¢0n¢0 , (5.1.2) where Im) and In) are the excited states of the molecules. Vis the interaction Hamiltonian and for a pair of molecule A and B, vAB = (47:80)_1 Idr dr' 6A6) 1330") r—r'l‘l. (5.13) 88 0 In Eq. (5.1.2), the operator \7 = V—(OIVIO). Following Li and Hunt, we separate the third-order induction energy AESZi into the hyperpolarization energy AElr3y)p , the static (3) reaction potential energy AEsrp , and the third-body reaction potential energy AE(3) tbp' Hyperpolarization energy results from Eq. (5.1.2), with the excited states Im), In) confined to one molecule and m, n ;15 0. For molecule A, the three-body and four- body hyperpolarization energies at third order are given by AEQJS’” = (1 / 2) (1 + SOBC) Idr dr' dr' CA (r, r', r'; 0, 0) ([353 (r) ((3%): (r') ((363 (r") +(1/2) (1+SOBD) Idr dr' dr"(A (r. r'.r";o.0) ((1,361 086') 480") + (1 / 2) (1 + (ecu) [dr dr' dr' CA (r, r'Ir"; 0, 0) (pg 0‘) 98 (r') (19(1)) (r"), (5.1.4) and ABSJSA) = Idr dr' dr' CA (r, r’, r'; 0, 0) (p(r)3 (r) (p((): (r') (p5) (r") , (5.1.5) respectively. In Eqs. (5.1.4) and (5.1.5), CA (r,r',r"g0,0) is the quadratic charge-density susceptibility of molecule A. The operator @BC perrnutes the labels B and C in the expression that follows. The hyperpolarization energy of A in Eqs. (5.1.4) and (5.1.5) can be interpreted as due to reaction field effects, where the potentials from the permanent charge distributions of two neighboring molecules create a nonlinear shift in the charge density at r in A. The induced shift in the charge density interacts with the potential due to the permanent charge distribution of the third molecule, thus resulting an overall 89 energy shift of A. To connect the hyperpolarization energy to the dielectric model, we expand the potentials in Eqs. (5.1.4) and (5.1.5), and then use Eq. (1 .10), to obtain AEQSS’” = (1/2) (1+goBC)(41t30)_2 Idrdr'dr'dr"drwp0 B(r) rII _ rrv Pg (rlv) x [ 80 86.1.4. (r, r', r'; 0, 0)] Ir' — r"'I_1 pg (r'") —2 I II III ivp B I II +(1/2)(1+pBD)(4n80) Idrdr dr dr dr Oq(r)[808 1A(r, r ,r ,,0 0)] rv p5) (rrv) x r'—r"'I_1p(I))(r") r"—r +(1/2) (1 +50CD)(41t80)—2 [drdr'dr'dr' dr'v pC(r) [e0 sq 1A(,r r', r"; 0, 0)] o -1 o I "—1 M N r —r I p80 r —rIv p80”), (5.1.6) and A 4 — I l M . - I N AEhyg’ ) = (41w0) 2 Idrdr dr dr drwpg(r)[80 eq’1A(r,r,r ;0,0)] 0 _1 o I "—1 M " xIr —r I p0C(r )r —rIv p80”). (5.1.7) If we define the two-body and the three-body effective potentials at r due to nonlinear screening by (p(2 )(r) and (p(3 )(r) respectively, then A(3,3)_ 2,AC AEhyp ( )(r ) —(1/2)IC(B—+C)1]drp0(r)

D)] Idrp(l)3(r)(pgi~AD)(r), (5.1.8) and AEQJSA) = Idr pg(r)cpg%fACD)(r). (5.1.9) The effective potentials in Eqs. (5.18) and (5.1.9) are given by (pngfAC)(r) = Idr'dr'[80 8211, A (r, r',r";0,0)](p8j (r’) (08 (r") , (5.1.10) and (péifACD) (r) = Idr'dr'[80 8211,14 (r, r', r";0, 0)](pg: (r') (pg (r'). (5.1.11) The operator C (B —+ C) in Eq. (5.1.8) replaces the labels B by C in the expression that follows it. Eqs. (5.1.6) — (5.1.9) prove that the hyperpolarization energy at third order is accurately described within the dielectric framework, where one molecule acts as the nonlinear dielectric medium to screen the interaction between the permanent charge distributions of other two or three molecules. The results are consistent with Eq. (5.1.1). The net three-body and four-body hyperpolarization energy at third is obtained by (3,3) hyp (3,4) summing the AE and AEhyp for A, B, C, and D. Static reaction-potential effects correspond to the dynamic effects in dispersion interaction, with the difference that the reaction-potential is produced in response to the permanent charge-density in this case, rather than the charge-density fluctuation. Static reaction potential energy results from linear screening and is obtained from Eq. (5.1.2) with m ;E 0 in one molecule and n It 0 in the other molecule. For molecule A, 91 A(3, BC) A(3, BD) A(3, CD) AE(3 3)— —AES srp, A + AES +AES = (1 / 2) (471180)—3 [ Idr dr' dr' dr‘" driv drv pA ' —1 xB (r', I"; 0) —l xr"—r"'I 1)(C("',r r —rv p0A(rv) + Idr dr'dr' dr'" driv drv pA 1 x8 (r', r';0) II 1 XD 1v v"1 A v xr — r’"I (r’”, r r —r pO (r ) +Idrdr'dr"dr"'driv drv pA — '_1xC(r',r";0) 1 XD v_1 A v xr"— r"'I (r, r —r pO (r ), (5.1.12) where each term accounts for two different polarization routes. For example, AEngE) accounts for the polarization routes A—IB—>C—>A and A—IC—>B—>A. Summing AE,(3 3) AE,srp A C(3, 3) D(3, 3) from Eq. (5. 1.12) with AEB(3’ 3), ABS and AES gives the total A13(3 3) A(3, BC) static reaction-potential energyAE at third order. AES can be viewed as the induction energy of molecule B, in the presence of the unscreened external potential fi'om molecule A and the screening potential from the shift in the charge density induced in C by A. Alternatively, it can be interpreted as the unscreened interaction between the first order shifts in the charge densities of B and C, caused by the permanent charge distribution in A. To show the dielectric screening present in the static reaction potential energy, we take AEB(3’BC) from Eq. (5.1.12), the second-order two-body induction SFP 92 energy of B from Eq. (2.1.5), and the relation between the nonlocal dielectric function and the charge-density susceptibility from Eq. (1.2), to obtain AEQQ’BC) = (1/ 2) (4Tt80 )—2 Idrdr'dr" dr'" driv pOA(r)Ir—r'_1xB(r',r”;0) '1 II iv iv III _1 A III ><[80 8v,C(r ,r )] r —r pO (r) A A —(1/2) [drdr'x (r,r';0)

B—>C—>B and B—->C—>B—IA. AEIBSIABC = (1/ 2) (47:80 )"3 Idr dr' dr' dr'" driv drv pg (r) Ir — r' _1 . . —1 X xC (rI’ rII; O) rII __ r"II—1 XB (r", rlv ; O) rlv _ rv p8 (rV) . (5.1.14) AEI13)§)ABC can be described as the interaction between the permanent charge density of B and the first-order shift in the charge density in B induced by A, in presence of C which acts as the dielectric medium. The first-order shift in the charge density of B induced by the potential due to the permanent charge distribution in A is given by ApIB(r) =(41rgo)-1 Idr'x (r,r';0)Ir—r’—lp(‘:‘(r'). (5.1.15) From Eqs. (5.1.14) and (5.1.15), 3,3 — I n III I "1 AEfbijBC = (1/2)(41t80) 2 Idrdr dr dr pg(r)Ir—rI xxC(I-',r'; 0) r" —r"'|‘1 Ap1B(r"') . (5.1.16) Energy shift of B due to direct intramolecular interaction between its permanent charge density and the first-order induced shift in the charge density is ‘lAplB(r'). (5.1.17) AEB = (1/2)(4m.~0)‘1 Idrdr'p(l)3(r)Ir—r' From Eqs. (5.1.16), (5.1.17), and (1.2), along with a change of the integration variables, we obtain 94 -1 ' APIB (r) r"—r' £8133,ch = (1 / 2) (mm—1 Idr dr' dr' pg(r)[eoe;}c (r, r";0)] —(1 / 2) (411.90)“1 Idrdr'p(l)3(r)Ir—r’|_1Ap?(r'). (5.1.18) The first term in Eq. (5.1.18) is the intramolecular interaction between the permanent charge density of B at r and the first-order induced shift in the charge density at r' , in presence of molecule C which acts as a dielectric medium to screen the interaction. Thus, within the dielectric model the three-body terms present in the third-body potential energy depend on the difference between the dielectrically screened and the unscreened Coulomb interactions between the permanent charge density and the first-order shift in the charge density within a molecule. This result can be compared to the second-order two-body induction energy described in chapter 2, where we showed that the two-body induction energy at second-order depends on the difference between the screened and the unscreened Coulomb interactions between the permanent charge densities within a molecule. Thus, the three-body terms in the third-body potential energy show the similar screening effect, but at the next order. Finally in this section, we derive the four-body effects in the third-body potential energy within the dielectric framework. Interactions present in the four-body terms in the third-body potential energy at third order are purely intermolecular. For example, the interaction energy associated with the polarization route B—>D—>C—>A is given by “5%:th = (1 / 4) (41:30 )_3 Idr dr' dr” dr'" driv drv pOA (r) Ir — r' -1 xc (r', r”; 0) . . -1 r"—I-"'I"1)(12(r"',r";0)r'V—rv 656V). (5.1.19) X 95 Eq. (5.1.19) can be viewed as the induction energy of C, in the direct potential from the permanent charge density in A and the screening potential from the induced shift in the charge density in D, caused by the potential from B. Alternatively, it can be described as the energy due to the intermolecular interaction between the first-order shift in the charge density of C induced by A, and the first-order shift in the charge density in D induced by B. Note that this energy corresponds to the term E82323 introduced by Piecuch,262 with i It j It k. To connect AESS‘IBDC to the dielectric model, we take Eq. (5.1.19) along with all the polarization routes with C and D in the excited states, the second-order three-body induction energy of C from Eq. (4.1.2), and the relation between the nonlocal dielectric function and the charge-density susceptibility from Eq. (1 .2), to obtain ’1 XC (r',r'";0) 4 - I N M . II 131383;),th = (47:80) 2 Idr dr dr dr drIv pfj‘(r)Ir —r . . —l ><[80 e;{o(r".r“’ ;0)1 r'v —r' p80) — (41:20 )72 Idr dr’dr"dr"’ p0 (r) Ir — r" ”1 xc (r', r'";0) xIr" -— t-'|'1 of? (r'). (5.1.20) In Eq. (5.1.20), the first term is the induction energy of C due its interaction with A and B, in presence of D which acts as the dielectric medium to screen the interaction. The second term is the unscreened three-body induction energy of C at second order. Thus, the four-body effects in the third-body potential energy are obtained as the difference between the screened and the unscreened second-order three-body induction energy. The second-order three body induction energy itself results from screening of the Classical 96 electrostatic interactions between two molecules. Hence, the four-body terms in the third- body potential energy are described as the screening effects present in the same type of interaction, but at the next higher order. 97 5.2 Dielectric screening and the nonadditive dispersion energy In this section, we derive the nonadditive dispersion energy within the dielectric model. We focus on the dispersion energy of a particular molecule and relate the change in the correlation between the intramolecular charge density fluctuations of that molecule with the nonlocal dielectric functions of other molecules. In section 5.2 A, we show the screening present in the three-body dispersion energy. In section 5.2 B, we derive the nonadditive four-body dispersion energy and show the screening present in the four-body dispersion energy of a particular molecule. The direct effects of overlap damping are included in the expression of the dispersion energy, but not modifications due to exchange or orbital distortion. 5.2A Dielectric screening of nonadditive three-body dispersion energy at third order Previously, Li and Hunt278 have developed a theory for the nonadditive three- body dispersion energy, based on the correlations in the fluctuating polarization of interacting molecules A, B, and C. The three-body dispersion energy derived in their work is given by AEE13’3) = —-r1 Eodm Idr dr'dr" dr'" driv drv Tr[T(rv,riv)0lC (riv,r";i(0) II III II B II I, - I A V, ' xT(r ,r )01 (r ,r,1(0)T(r,r)01 (r,r ,1(0)]. (5.2A.1) In Eq. (5.2A.1), a(r,r';i(t)) denotes the nonlocal polarizability density generalized to imaginary frequencies, and T(r,r') is the dipole propagator defined in Eq. (2.2.3). Tr means the trace of the expression that follows. The result in Eq. (5.2A.1) is derived after 98 summing the three-body dispersion energies of A, B, and C, and using the Kramers- Kronig relation between the real and the imaginary parts of the nonlocal polarizability density. In the present section we use a susceptibility based approach that we used in section 2.2 to Show the screening present in the intramolecular charge density fluctuations due to the two-body dispersion interaction. Following the same approach that we used in section 2.2, we show that the average energy shifi of molecule A at third order due to the correlation between the charge density fluctuations at points r and r’, in presence of molecules B and C is given by A _ I I M . " I AEd (3’3) = —(h/47t)(47t80) 3(1+@BC)E:Od0) J'drdr dr dr drIv drvxA (r,r;oo) -1 —1 IV f 0 r —r —1xB(r",r";w) r'"—r xC(riv,rv;co) rv —r X x coth(h(o / 2kT) , (5.2A.2) where xA"(r,r';w) is the imaginary part of the charge-density susceptibility of A, defined in chapter 2. Charge-density fluctuations in A at time t' induces a shift in the charge density in B at time t" which eventually induces a shift in the charge density in C at time t, thus creating a reaction potential on A at time t. Using the fluctuation- dissipation theorem from Eq. (2.2.9), we obtain A 3,3 — I II III ' I AEd( )= (1/4)(41t80) 3(1+goBC)[dz-dr dr dr dr" dr" fwda—t) v—l r—r x flaw—t')<5pA(r,t)SpA(r',t')> + 99 . . -1 xxC(rV,rlv;t --t") r” —r"' xB(r",r”;t"-t') r"-—r' _1 . (5.2A.3) From Eq. (5.2A.3), the two-body dispersion energy of A from Eq. (2.2.11), and the relation between the nonlocal dielectric function and the charge—density susceptibility in the time domain from Eq. (2.2.12), AE§‘(3’3) = (1 / 4) mam-2 (1 + gOBC) jdr dr' dr" dr'" driv J: d(t - t') j: d(t" — t') iv _1 r—r x<6pA (r,t)éipA (r’,t')> 30 8;1C (riv,r"';t—t") + 9 X xB (rm, rII; tII _ tr) ru _ rIl—l _(1/4)(4u.~20)‘2 [1+C(B —) C)] Idrdr'dr'dr’" fde—t') >< <59A(r. t)59A(r’, t')> 1" - r" ‘1 xB(r'Ir"';t - t') r'" - fl"1 , + (5.2A.4) where the operator C (B —> C) replaces the label B by C in the expression that follows. The first term in Eq. (5.2A.4) gives the screened two-body dispersion energy of A due to interaction with B (or C), in presence of C (or B), which acts as the dielectric medium with the nonlocal dielectric screening function a;lc (riv,r'";t—t"). A charge density fluctuation in A at r',t’ sets up a potential in B at r',t' (in the Coulomb gauge, with retardation neglected) which induces a shift in the charge density in B at r',t’. The reaction potential in A at r,t due to the shifi in the charge density in B is screened by the presence of C via its nonlocal dielectric function e;1C(rw,r"';t—t”), and the screened 100 potential affects the two-body dispersion energy of A. The second term gives the second- order two-body dispersion energy of A in presence of molecules B and C. Thus, the three-body dispersion energy at third-order depends on the difference between the dielectrically screened and the unscreened two-body dispersion energy at second order. In an alternate way, we can describe the three-body dispersion energy of A in terms of an effective, two-body susceptibility of B and C. Following Kohn, Meir, and Makarov,I64 if the long-range interaction between B and C acts as a small perturbation, this two—body susceptibility is given by xBC(r,r';co) = 9» Jdn drz xB (r,r1;co) 4 xC(r2,r';co), (5.2A.5) n80 ln - 12| where it is a coupling constant that “turns on” the long-range interaction between B and C. If B and C are non-interacting, the overall susceptibility is given by XBC = xB(r,r1;w) + xC(r2,r';w). (5.2A.6) 283 Previously, Li and Hunt have showed that for a pair of interacting centrosymmetric linear molecules A and B, the overall polarizability in presence of an external field Se is given by A B lim :56 —> o ——a(““ + ”(1) 63E _ A B ind — aaB +aaB “Mali , (5.2A.7) where (19B and (12B are the polarizabilities of the isolated molecules and Aug}? is the collision-induced electronic polarizability of the pair. Using a self-consistent solution of a set of equations that relate the induced dipole moment to the local field, they showed that, at first order 101 Aug}? = (1 + goAB )aég, TYMR) (1533 . (5.2A.8) In Eq. (5.2A.8), (13b is the dipole polarizability of molecule A, R is the vector from an origin at the center of symmetry of molecule A to the origin at the center of molecule B, TaB (R) is the dipole propagator given by, TaB (R) = VaVB (R_1) , and 50 AB permutes the labels A and B in the expression that follows. The collision-induced electronic polarizability defined in Eq. (5.2A.8) determines the first-order dipole-induced-dipole. Here we use the same method to derive the two-body susceptibility for a pair of interacting molecules A and B, in presence of a fluctuating external potential (Pext (r; (0). Within linear response, the shifts in charge densities of A and B are related to the applied potential by the equations ApA(r;co) = J'dr'x (r,r';w)(pé},p(r’;(o). (5.2A.9) ApB(r;(o) = jdr'x (r, r';o))(pgpp(r'; 0)). (5.2A.10) The applied potential at A is related to the external potential (Pext (r; m) and to the potential due to the shift in the charge density of B, ApB (r; 0)) by ApB(r';m) A I r;o) = r;o) + dr (Papp( ) (Pext( ) I 4 O|r_rI = + xIr—rn‘"1xBC(rfi,rM;t_tI) _1. r'" —r| (5.2A.15) Finally, using Eqs. (5.2A.15), (2.2.11), and (2.2.12), we obtain AEdA(3’3) = (1 / 4) (41:80)_1 jdr dr' dr' food (t — t')< 5pA(r, t) 8pA(r', t') >+ x rI_rII—1 so 8;",ch (r", r; t — t') 103 1‘1 . (5.2A.16) _(1/4)(47t80)—1Idrdr'<6pA(r,t)5pA(r',t')> |r—r + The first term in Eq. (5.2A.16) gives the Coulomb energy associated with interactions between the fluctuating charge densities SpA (r, t) and SpA (r',t') in presence of the pair of molecules B and C, which together act as the dielectric medium with the nonlocal screening function .9;ch (r’,r;t—t'). The second term is the Coulomb energy of direct intramolecular interactions of the charge-density fluctuations in molecule A, in absence of B and C. Thus, Eq. (5.2A.16) proves that the three-body dispersion energy of A results from the difference between dielectrically screened and unscreened interactions between its intramolecular charge-density fluctuations, where the dielectric screening is characterized by a two-body screening function, defined in Eq. (5.2A.14). 104 5.2B Dielectric screening of nonadditive four-body dispersion energy at fourth order In this section, we describe the nonadditive four-body dispersion energy at fourth order Within the nonlocal dielectric model. We consider the dispersion energy of molecule A in a cluster of interacting molecules A---B---C°-°D, and derive the dispersion energy using the same susceptibility base approach that we used in chapter 2 and in the previous section. In appendix B, we derive a new equation which gives the four-body dispersion energy of the A---B°-~C---D cluster at fourth order. We show that A 4,4 3h " II— AEd( ) = —-;(41t80)_4 Soda) Idr...drv“ xA r le (r',r";ic0) rIII _ l,lv iv r;v V] X xC(r V“ —r - (5.23.1) XXD (rVi,rVii;im) r In terms of real frequencies, the four-body dispersion energy of A is given by AEdA(4’4)= —(h/41t)(41c30)4(1+5oBCD) Idr...drVii [:0de x ",(r r' 00,) . —l X |r’ _ rII "' rIII _rIV x C(rivr V ;(D) x rv —rVi — xD (rVi,rVii;co) rVii —r— coth (hail 2kT). (5.23.2) In Eq. (5.2B.2), the operator (QBCD permutes the labels B, C, and D. Using the fluctuation-dissipation theorem from Eq. (2.2.9), 105 AEdA(44)" _(1/4)(4n30) —4(.1+5oBCD)Idr .drVii Eodfl- 1W)J:Od(tt .—1 rvii xfood(t'— t'A)<5p (r, t)5pA (r', t'r-)> + Vll’r V XC (rv’riv;tIII _ tII) TV] —I‘ xx D(r iv _1 r —r"' II r —r’|_1. >< xB(r’”.r";t" -t') (5.23.3) From Eqs. (5.28.3), (2.2.12), and the three-body dispersion energy of A from Eq. (5.2A.3), we show that 4 I ”I N A932?4 )— —(1/4)(47t80) 3..(1+5oBCD)jdr .drv'food(t—t)J:od(t —t) xfo d(t"—t') 8 A(r t)8 A(r' t') r—rVi _1 00 P 9 P 9 + X8 -1 Vi V, _ III C V iv, III_ II 08v,D(r ,r ,t t))( (r ,r ,t t) . —1 B _1 X rIV _ rIII x (rm, rII; tn _ tI) rII _ rIl A(3,3) _AEB CD . (5.2B.4) In Eq. (5.2B.4), the first term gives the screened three-body dispersion energy of A in presence of molecule D (or B or C) which acts as the dielectric medium with the nonlocal screening function 5v,D (rVi ,rvgt — t’"). The second term gives the unscreened three- body dispersion energy of A at third order in the molecular cluster A-"B-"C-"D. Thus, from Eq. (5.2B.4), the four-body dispersion energy of A depends on the difference 106 between its dielectrically screened and unscreened three-body dispersion energies at third order. Finally in this section we show that the four-body dispersion interaction can also be described as screening to the direct intramolecular interactions between fluctuating charge densities, where the other three molecules provide the dielectric screening. To show this screening effect, we derive a three-body susceptibility in terms of the nonlocal charge density susceptibilities of three interacting molecules. Previously, Champagne, Li, and Hunt284 have showed that for a cluster of non-overlapping, isotropic species A, B, and C, interacting at long range, the nonadditive three-body polarizability at second order is 280151263) = (1A (13 (1C 3 ABC Ta5(RA,RB)TB5(RA,RC) + (1 / 3) s ABCCA aBaC Tay5(RA,RB), (5.23.5) where the species centers are located at RA, RB, and RC respectively; SABC denotes the sum over all permutations of the labels A, B, and C in the expression that follows it; and the propagators T aB...(o (r, r') of arbitrary rank are given by Tag...m(r,r') = vavB mvm lr—r'I—l. In Eq. (5.23.5), (1A is the dipole polarizability of the isotropic species A, and the C tensor determines the quadrupole induced by a uniform field gradient, within linear response. Here we derive the three-body susceptibility of three molecules A, B, and C, interacting at long range, in presence of a fluctuating external potential. Following Eq. (5.2A.9) — (5.2A.10), the shifts in the charge densities of A, B, and C, within linear response is given by 107 APA (r; 0)) + ApB (r; (o) + ApC (r; 0)) = Id!“ X (1’, I"; (0) (PaApp (r'; (o) + Idr' x (r, r'; w) (ngp (r'; (o) + Jdr'x (r, r';(o) (pgpp (r';o)) . (5.2B.6) The applied potential at A depends on the external potential and on the potentials due to the shifts in the charge densities of B and C. Thus, ApB(r';co> +1 r, ApC(r';w) A . _ . I (papp(r,(o) — cpext(r,(o)+ Idr 41:80lr—r’| , . (5.23.7) 41:80 lr—r The charge density shift of B, in turn, depends on the charge density shift of C (and vice versa), which finally yields I II 1 I II 0 (paApp(r;w) = (pext (r; (o)+ Idr dr —, xB(r ,r ;m) (Pext(r ;co) 41:80 lr—r | + Idr' dr' ——————,- xC (r', r'; (o) (Pext (r"; 0)) 47:80 |r — r ' 1 + 1+ dr'...drIv — B r',r";w xxC (r",riv;o)) (Pext (riv;o))+... (5.2B.8) Solution of Eqs. (5.2B.6) and (5.2B.8) gives ApA(r;co> + 413%: co) + 4pc (r; co) = [dr' x (r,r'; 0)) (Pext (r'; m) + [dr' x (r. r'; 03) ‘Pext (r'; w) + [dr’ x (r, r'; 0)) <9ext(r'; 0)) + Idr' AXAB (r, r'; (o) ‘Pext (r’; (o) + Idr' 13ch (r, r'; (o) (Pext (r'; (0) + [dr' AxCA (r, r'; (o) (Pext (r'; 00) 108 + Idr' AXABC (r, r'; w) (Pext (r'; (o) + (5.28.9) In Eq. (5.28.9), AXAB (r, r';to) denotes the two-body part of the susceptibility defined in Eq. (5.2A.13), and AxABC (r, r';(o) is the three-body part of the susceptibility, given by ABC(r,r';00) = SABC Idr'...drv xA(r,r";(o) 1' X B(r"'r iv ;(n) 41:80 r r'"| >< 1. XCO'V , r';t») , (5.23.10) 47:80 rIv —rv In time domain, the three-body susceptibility is ABC I, _ I _ II V III _ II II _ I A II, _ III (r,r,t t)-SABC Idr...dr food“ t )fwda t)x (r,r ,t t ) 1 BIIIivIIIII 1 ivr,v ,,,x (r ,r ;t —t) . x C(r ;"'t -t'). lV_rV 411280 r (5.28.11) Using the definition of the three-body susceptibility in the time domain from Eq. (5 .28.] 1) and a change in the integration variables, the four-body dispersion energy of A in Eq. (5.28.3) can be written as AE§(4’4) = (1/ 4) (47320 )-2 Idr dr' dr' dr'" E d (t — t')< 8pA (r, t) 8pA (r', t') >+ —1 XBCD I xlr_rII II III —1 —r|. (5.23.12) Thus, from Eqs. (5.28.12), (2.2.11), and (2.2.12), we obtain A3§(4I4) = (1/4)(41I280)_1 jdr dr’dr' fiod (t — t')< 8pA(r, t)8pA(r',t') >+ x rI_rII—1 80 83,130) (r", r;t — t') 109 ‘1 (5.23.13) —(1/4)(47t80)—1 Idrdr'<5pA(r,t)8pA(r',t')> |r-r' + Eq. (5.28.13) proves that the four-body dispersion energy at fourth order results from the screening of the intramolecular exchange-correlation energy, where the dielectric screening depends on molecules 8, C, and D which together act as the dielectric medium. Thus, the many-body effects are contained within the nonlocal dielectric function. 110 5.3 Dielectric screening and induction-dispersion energy at third and fourth orders In this section, we prove that the dielectric screening model also describes the simultaneous induction-dispersion energy of a cluster of molecules. Simultaneous induction-dispersion effects appear, because the permanent charge-density of one molecule acts as the source of a static external potential (00 that perturbs the two-body or three-body dispersion interaction of two or three other molecules respectively: Each of the two or three molecules in the cluster is hyperpolarized by the simultaneous action of the static external potential and the fluctuating potentials from its partners. The static external potential also alters the correlations of the spontaneous, quantum mechanical fluctuations in the charge densities of the other interacting molecules. Within the dielectric model, the induction-dispersion interaction can be interpreted as the perturbation of the dielectric medium by the external potential. This perturbation brings in nonlinear screenings into the dielectric medium, which are of secondary importance in 183,184 the case of pure dispersion interaction. Previously, Hunt and Bohr developed a theory for the dispersion dipole of an A---B pair, based on the change in dispersion energy due to a uniform, static external field‘ Se. Li and Hunt278 applied the same analysis to the dispersion energy of the A---B pair in presence of the static external field SOC due to the permanent charge distribution of molecule C, after allowing for the nonuniformity of the field. In the present work, we relate the three-body and four-body induction-dispersion interactions to the dielectric model. We focus on the energy shift of a particular molecule due to the correlation between its intramolecular charge density fluctuations in presence 111 of the static external potential. We use the charge density susceptibility based reaction potential approach that we used in the previous section, to describe the induction- dispersion energy of the interacting molecules, where the overlap between the charge distributions of the molecules is assumed to be weak or negligible. In presence of the static external potential (pOC due to the permanent charge distribution in molecule C, the two-body dispersion energy of molecule A is given by A3303) = (1 / 4) (47t80)_2 jdr dr'dr' [:on — t')< 8pA(r,t)5pA(r',t') >+ C —1 II III I "‘1 XB(r 9r 9(p0 9t-t) ' (5.3.1) r"—r| xlrI_rII In Eq. (5.3.1), xB (r',r’”;(pg,t -t’) denotes the nonlocal charge density susceptibility of molecule 8 in time domain in presence of the static external potential (pg. If the external potential is significantly small, x((p,t — t') can be expanded in Taylor series, x(tp0C,o)) = x(t —t')+:;(t —t',0)chC +..., (5.3.2) where x(t — t') and C(t — t',0) denote the linear and the quadratic charge-density susceptibilities respectively, in the absence of the perturbing potential. Substituting Eq. (5.3.2) in Eq. (5.3.1), we obtain AES‘Q’Z) = (1 / 4)(47t80)—2 jdr dr'dr" [:Oda — t')< 6pA(r, t)8pA(r', t') >+ rI_rn—1xB(rl,rn;t_tI) “'1 X r'"-r| —2 iv I A A I I 1/4 4 d...d d - 6 , 6 I +< )(mo) Ir r I: (t t)< p (rt) p (r t)>+ 112 r"‘1((r',r, " r";—t t,0)|r"—r|_1+ . -1 xr'— _1§B (r",r " ,riv;—t t',"0)|r— r| 1rW—rv pg(rv). (5.3.4) The energy in Eq. (5.3.4) can be interpreted as the hyperpolarization energy of 8 caused by the fluctuating potentials from A and the static potential from C. The nonlinear shift in the charge density of 8 caused by the potentials due to the charge density fluctuations in A interacts with the static external potential from C, thus giving an overall energy shifl. Using Eq. (5.3.4) and the nonlinear dielectric function in the time domain, the three-body induction-dispersion energy of A is given by AB“: 3) — _(1/4)(4n.«~.0)’2 [dr...driv Ed(t—t')<59A(rat)59A(r',tI)> + XI”— "‘1[eosq1,r,B(r'r'" r'v,t— t',"'—0)]|r r|1pg(ri"). (5.3.5) Thus from Eq. (5.3.5), the induction-dispersion energy of A is described within the dielectric model as the energy shift due to the nonlinear interaction between the charge- density fluctuations in A and the permanent charge distribution in C, in presence of B which acts as the nonlinear dielectric medium. This interaction is analogous to the 113 hyperpolarization energy described in section 5.1, with the difference that the hyperpolarization energy appears due to the nonlinear interaction between permanent charge distributions, while the induction-dispersion energy is due to the nonlinear interaction between permanent charge distribution and charge-density fluctuations. Following the same line of argument, we can relate the four-body induction- dispersion energy of molecule A with the dielectric model. Here we consider the three- body dispersion energy of A in presence of 8 and C, perturbed by the external potential due to the permanent charge distribution of D. Expanding the charge density susceptibilities of 8 and C in Taylor series with respect to the external potential and keeping only the lowest order terms, the four-body induction-dispersion energy of A is givenby A44 — ' I II I AEi+(d’ )=(1/4)(41:80) 3(1+goBC)Idr...drv' f: d(t-t) f: d(t —t) A A —1 3 iv ‘1 x<5p (r,t)8p (r',t')> r'—r" x (r',r'";t"—t') r'—r + C iv v vi v _1 D vi xC (r ,r ,r ;t—t",0)r —r (90 (r ). (5.3.6) Eq. (5.3.6) can be interpreted as the hyperpolarization energy of C (or 8) caused by the fluctuating potential from A, potential due to the shift in the charge density of B (or C) caused by the fluctuating potential from A, and the static external potential due to the unperturbed charge density in D. Using Eq. (5.36) and the relation between the quadratic charge-density susceptibility and the nonlinear dielectric function, we obtain A(4,4) AEi+d = (1/4)(47re0)_3(1+gOBC) Idr...drVi Eda—t') [luv-t') 114 . -l rI _ rII _1xB(rII,rIII;tII _ tr) rIII _ rlv x< 6pA (r, t) SpA (r', t')>+ x[803;,1C(r'v,rv,rw;t—t',0)]rv—r p80"). (5.3.7) Eq. (5.3.7) proves that the four-body induction-dispersion energy of A arises due to the interaction between the charge density fluctuations at r and the reaction potential caused by the shift in the charge density of B (or C), induced by the potential from the charge density fluctuations in A atr' , in presence of C (or B) which acts as a nonlinear dielectric medium, perturbed by the potential from the permanent charge density in D. This interaction energy can also be explained as screening to the three-body induction-dispersion energy of A defined in Eq. (5.3.5). To show this screening effect, we consider the cluster of molecules A, B, and C, in presence of the external potential from the permanent charge density of D, where the dispersion energy arises due to the charge density fluctuations in A. The overall induction-dispersion energy of A in this case is . 3,3 3,3 given by the sum of the three-body terms ABE A-))B _) A) (—D and ABE A-—))C __) A) (—D , and the four body terms from Eq. (5.3.7). Using the relation between the nonlocal dielectric function and the charge-density susceptibility from Eq. (2.2.11), we can show that A(4,4) Alai+d = (1/4)(41ra0)_2(1+5oBC)Idr...drv Eda—t’) Eod(t'—t') X< 5pA(r’1)5pA(rI’tI) > |rI_rII _1[80 8B1 (r", I.III;tII_tI)] + . —I p33) XISO 8;,1C(r",r'v,rv;t-t",0)] 115 — (1/4)(41t30)_2[1+C(B —> C) jdr...dri" [:odO—t') I 0‘1 —1 II III iv, I r —r [aoeq’B(r ,r ,r ,t—t,0)] x< 5pA (r, t) 8pA(r', t') >+ xIr" 414,580“). (5.3.8) In Eq. (5.3.8), the first term gives the screened three-body induction-dispersion energy of A. Here molecule 8 (or C) acts as the dielectric medium to screen the induction- dispersion interaction, and the screening is linear. The second term gives the unscreened three-body induction-dispersion energy of A. Thus, the induction-dispersion energy at fourth order depends on the difference between the screened and the unscreened three- body induction-dispersion energy. 116 Chapter 6: Summary and conclusions The present work proves that the intermolecular interactions at second, third and fourth orders are accurately derived in terms of the nonlocal dielectric model, where the overlap between the interacting molecules is weak or negligible. Within linear response, the nonlocal dielectric function av (r, r';(o) determines the effective potential at r , when an external frequency—dependent potential (p(r',(o) acts at r'. A separate dielectric function ad(r,r';o)) relates the dielectric displacement D(r,o)) to the external field E(r',w). For translationally invariant systems, the isotropic average of ad(r,r';m) reduces to 3V(r,r';o)). Within the intramolecular environment, ad(r,r';m) and ev(r,r';(o) are different. The nonlocal dielectric function av(r,r';w) is related to the nonlocal charge-density susceptibility x(r,r';(o) by Eq. (1.2). x(r,r';w) determines the induced shift in the charge density 6p(r,co) at point r in the molecule due to an applied potential (p(r',o)) at r'. Molecular properties which are related to the charge-density 129,188 susceptibility are nonlocal polarizability density, infi'ared intensities,127 the Stemheimer electric field shielding tensor, 127 charge reorganization terms in vibrational force constants,286 and the softness kernel of density fimctional theory.287’288 Dielectric response of translationally invariant systems or systems with spatial periodicity can be described by a dielectric function which depends only on the distance between the response point r and the point r' where the external perturbation acts. For these systems a convenient choice is to use the dielectric function 8( k,o)) , which is the spatial Fourier transform of e(r—r',o)). Dielectric functions of the form e(k,oo) have 117 been used in order to study quantum many-body problems, properties of quantum dots, solvation dynamics and polarization fluctuations in liquids, and electron transfer. Dielectric models have been applied to study interactions within proteins and biomolecules. Models have been developed to probe the dielectric environment inside protein molecules and these models have been used to interpret different experimental observations such as determination of pKa shifis of inserted amino acid residues,289 dynamic shifts of the fluorescence, for the markers placed at various sites of protein,290 measurements of Stark effect on absorption bands of different chromophores,“ and determination of the apparent basicities of the different charge states of protein”:2 Inhomogeneities inside the protein molecules make it impossible to define a universal dielectric constant (or constants) for proteins. The choice of the value of protein’s dielectric constant depends on the particular property or interaction to be studied and the model used to study those properties. To give a complete description of the dielectric nature within the protein environment, it is necessary to develop a model in terms of the site-dependent dielectric constants. Extension of the dielectric model to describe interactions within the intramolecular framework was suggested in several works. Early works on light scattering by fluids and collisional polarizability anisotropy of interacting noble gas atoms used a polarizability density instead of the point dipole approach. In later works on light scattering by fluids, response functions were used that depend both on r and r'. Importance of nonlocal response has been noted in recent works on surface enhanced Raman scattering by metal nanoshells. 118 Intermolecular interactions at first order are purely electrostatic in nature. Within quantum perturbation theory, first-order intermolecular interactions are obtained as Coulomb interactions between the unperturbed charge distributions or polarizations of the molecules. When the molecules are far apart, the electrostatic interaction energy is obtained as a sum of the interaction energy of permanent multipole moments of the molecules, given by Eq. (1.14). For a pair of interacting molecules, the first-order force on nucleus I in molecule A is calculated by taking the negative gradient of the first-order interaction energy with respect to the coordinate RI of nucleus 1. Thus, it depends on the derivative of the permanent charge distribution of A with respect to RI. When using a multipole expansion, the first-order force is given in terms of the derivatives of the permanent multipole moments of A. First-order forces on nuclei can also be calculated using the electrostatic Hellmann-Feynman theorem and the Stemheimer-type shielding tensors, where the force on nucleus I is obtained as sum of the interactions between the charge on nucleus I and effective fields and field gradients at I due to molecule 8, given by Eq. (1.21). The effective field at I due to molecule 8 depends on the field from 8 due to its permanent moments and the nuclear shielding tensors of I. At first order, the effective field and the field gradient originate due to linear screening of the external field and the field gradient and are determined by the linear shielding tensors of nucleus 1. These linear shielding tensors are related to the derivatives of permanent multipole moments of molecule A with respect to RI [Eqs. (1.16), (1.18b)]. Physically, the shielding appears due to the electronic screening of the external field. Within the nonlocal polarizability density model the electronic screening is shown by the nonlocal polarizability density a(r,r';0) , where the first-order force depends on the fields at I from 119 the unperturbed charge distribution in B and the first-order induced shift in the polarization of A [Eq. (1 .20)] which is determined by a(r,r’;0) of A and the field due to unperturbed charge distribution in B. The Stemheimer—type shielding tensor 7:13 is connected to the nonlocal polarizability density “YB (r, r') by Eq. (1.19). The first-order force on nucleus 1 in molecule A was first derived within the nonlocal dielectric model by Jenkins and Hunt. A susceptibility-based approach was used to express the first-order force on I in terms of the static nonlocal charge-density susceptibility of A and the potential from the unperturbed charge-distribution in 8. Using the relation between the charge-density susceptibility and the nonlocal dielectric function from Eq. (1.2), the first- order force was expressed [Eq. (1.23)] as interaction between nucleus I and the external potential from B in presence of the intramolecular dielectric medium A which is characterized by the nonlocal dielectric function av, A (r, r'; 0). In chapter 2 of this work, we have proved that the induction and dispersion energies at second order are derived within the nonlocal dielectric model. Using quantum perturbation theory, the second-order induction energy is obtained from Eq. (1.24) with the excited states confined to either molecule A or molecule 8. At second order, the induction energy depends on the static fields due to the permanent charge distributions of the interacting molecules and the responses of the molecules to those static fields. Induction energy is given within the nonlocal polarizability model by Eq. (2.1.1). When the molecules are far apart, the induction energy can be written in terms of the permanent moments and the multipole polarizabilities of the interacting molecules [Eq. (1.25)]. 120 In section 2.1 of chapter 2, we derived the second-order induction energy in terms of the static charge-density susceptibilities and the potentials due to the permanent charge distributions of the interacting molecules [Eq. (2.15)]. In order to derive Eq. (2.15), we expressed the fields in terms of the potentials and used the relationship between the charge-density susceptibility and the nonlocal polarizability density from Eq. (2.14). Then using the relation between the charge-density susceptibility and the nonlocal dielectric function ev(r,r';0) in Eq. ( 1.1) we have proved that the induction energy at second order results from the difference between the dielectrically screened and the unscreened Coulomb energies due to the permanent charge distributions within a molecule, where the second molecule acts as the nonlocal dielectric medium. The result is given in Eq. (2.1.7), where the first and the second term give the screened and the unscreened Coulomb interactions respectively, within a molecule. Thus, we conclude that the two-body induction energy at second order is derived within the nonlocal dielectric model as the difference between the dielectrically screened and the unscreened intramolecular interactions between the unperturbed charge densities of the molecules. In section 2.2, we have derived the second-order dispersion energy within the dielectric framework. Dispersion energy results from the correlation between the charge- density fluctuations or polarization fluctuations within a molecule. Using a reaction field method, the dispersion energy is derived as an integral over frequency, where the integrand is factored into the nonlocal polarizability densities of the molecules at imaginary frequencies. This result is given in Eq. (1.29). Dispersion energy is also obtained from the second-order perturbation theory using Eq. (1.24) with excitations confined to both molecules. Second-order perturbation theory has been applied to 121 calculate the dispersion energy in several works. Within the density functional theory, dispersion energy is obtained as the exchange-correlation energy. In the present work in section 2.2, we have expressed the dispersion energy in Eq. (2.2.1) as an integral over frequencies with the integrand factored into the charge- density susceptibilities of molecule A and B at imaginary frequencies. Then we have used a contour-integration technique and the symmetries of the real and the imaginary parts of the susceptibility to write the dispersion energy in terms of the charge-density susceptibilities of the molecules at real frequencies [Eq. (2.2.8)]. Using the fluctuation- dissipation theorem from Eq. (2.2.9), we have expressed the dispersion energy in the time domain. Finally, from Eq. (2.2.9) and the relation between the charge-density susceptibility and the nonlocal dielectric function in the time domain from Eq. (2.2.12), we have derived the dispersion energy as the difference between the screened and the unscreened Coulomb interactions between the charge-density fluctuations within the molecules. The final result is given in Eq. (2.2.13). The first term in Eq. (2.2.13) is the screened interaction between the charge-density fluctuations in A in presence of B, which acts as the dielectric medium (and similarly for B). The second term gives the unscreened interactions between the charge-density fluctuations within the molecules. Thus, we have proved that the second-order dispersion energy results from the screening of the intramolecular charge-density fluctuations. Induction and dispersion forces on nuclei at second order are calculated by taking the negative gradients of the second-order induction and dispersion energies with respect to the nuclear coordinates. At second order, the induction force on nucleus I in molecule A depends on the first-order and the second-order induced polarizations in A, as given in 122 Eq. (1 .26). The second-order polarization in A results from the nonlinear response of A to the static fields from B and is determined by the nonlocal hyperpolarizability density A Bafir (r, r',r';0, 0) of A. The first-order polarization of A in Eq. (1.26) is induced by the first-order polarization of 8 caused by the. unperturbed polarization in A. Thus, the second order induction force on I is related both to the linear and the nonlinear response of A. It is important to note here that the induction force on nucleus I does not stem from the interaction of I with the polarization of B, but from the interaction of I with the polarizations induced in A. The induction force can also be described in terms of the nuclear shielding tensors by Eq. (1.27). The terms which depend linearly on the reaction field from 8 and its gradients, correspond to the first term of Eq. (1.26) [i.e. the first- order induced polarization in A]. Terms depending quadratically on the reaction field and its gradients are related to the second term of Eq. ( 1.26) [nonlinear polarization of A]. The nonlinear shielding tensors depend on the derivatives of the molecular polarizability densities with respect to the nuclear coordinates [Eqs. (1.18a), (1.18c)] and hence, on the nonlocal hyperpolarizability densities [Eq. (1.28a) — (1.28b)]. Eqs. (1.19), (1.22), (1.28a) and (1.28b) connect the electrostatic and the second-order induction forces calculated within the nonlocal polarizability density model to the forces calculated applying the electrostatic Helhnann-Feynman theorem. In section 3.1 of chapter 3, we have derived the second-order induction force on nucleus K in molecule A in terms of the nonlocal dielectric model. We began the derivation with the induction energy expressed in terms of the static polarizabilities of the molecules and the fields from the unperturbed charge distributions in the molecules. Within this approach, the second-order induction force on nucleus K depends on the 123 derivatives of the nonlocal polarizability density (1A (r, r';0) of A and the field 365‘ (r) at B, with respect to RK. The derivative of aA(r,r';0) with respect to RK depends on BA (r, r';0) , the nonlocal hyperpolarizability density susceptibility of A [Eq. (3.1.2)]. The field 380) in molecule 8 depends both on the electronic and the nuclear charge densities of A and the derivative of the electronic charge density with respect to RK is related to the nonlocal charge-density susceptibility xA(r,r';0) of A., given by Eq. (3.1.5). Thus, we have expressed the induction force on nucleus K in terms of the nonlocal charge-density susceptibility and the nonlocal hyperpolarizability density of A. Next, we have written the fields in terms of the potentials, used the potential from the first—order shift in the polarization of B from Eq. (3.1.10), and used the relation between the nonlocal hyperpolarizability density and the quadratic charge-density susceptibility C(r,r’,r';0,0) from Eq. (3.1.12), to obtain the second-order induction force on K within the susceptibility based approach in Eq. (3.1.13). Eq. (3.1.13) shows that the induction force on K depends on the nonlinear screening of the potentials due to the unperturbed charge distribution in 8 and linear screening of the potential from the first-order shift in the charge distribution of B. This result is consistent with the induction force obtained previously using the nonlocal polarizability density model and the nuclear shielding tensors. In order to show the nonlinear screening in the induction force, we have used the nonlinear dielectric function given in Eq. (1.10). Finally using the effective potentials in A due to linear and nonlinear screenings we have expressed the induction force on K by Eq. (3.1.17). Eq. (3.1.17) proves that the second—order induction force on nucleus K 124 depends on the intramolecular screening of the external potential acting on A. This result is similar to the first-order force derived by Jenkins and Hunt. The difference is that the first-order force includes linear screening only, while the second-order induction force results both from linear and nonlinear screenings within molecule A. The dispersion force on nucleus K can be calculated the same way, by evaluating the negative gradient of the second-order dispersion energy with respect to RK. Within the real frequency domain, the dispersion force on K in molecule A contains two different terms: one includes the derivative of the frequency-dependent polarizability a(r,r’;(o) of A with respect to RK, and the second one contains the derivative of the correlation between the polarization fluctuations within A with respect to RK. The derivative of (1A (r,r';(o) with respect to RK depends on the nonlocal hyperpolarizability density BA (r,r',r";o),0) of A. From the fluctuation-dissipation theorem, the correlation between the polarization fluctuations is related to the imaginary part of the nonlocal polarizability density, 01A" (r,r';co). Thus the derivative of the correlation is given by the derivative of 01A" (r,r';o)) with respect to RK, which is related to the imaginary part of the nonlocal hyperpolarizability density BA" (r,r',r";co,0). The first component of the dispersion force resembles the first component of the induction force, with the difference that in the case of later, the external fields are time dependent and the nonlocal hyperpolarizability density depends on frequency. The second is quite different, since it depends on the imaginary part of BA (r,r',r";(o,0) and shows no linear screening like in the case of the induction force. In earlier work, it was concluded that this part of 125 dispersion force might depend on the polarization of 8. Moreover, since the correlation between the polarization fluctuations is affected by the change in the nuclear Coulomb field, it was noted that this field might bring new correlations and could even change the magnitude of the correlation function (field induced fluctuation correlations). However, those new type of correlations were not derived explicitly. In section 3.2, we have derived the second-order dispersion force within the nonlocal dielectric model. The dispersion energy was written within the frequency domain, where it depends on the real part of the charge-density susceptibility of one molecule and the imaginary part of the charge-density susceptibility of the other [Eq. (3.2.1)]. Using Eq. (3.2.1), the first part of the dispersion force (F3609 on nucleus K has been derived in terms of the real part of the quadratic charge-density susceptibility of A and the correlation of the charge-density fluctuations in B. This part of the dispersion force appears due to the nonlinear screening of the fluctuating potentials from molecule 8. Thus it is similar to the first part of the induction force. We have showed the nonlinear screening present in the dispersion force using the frequency-dependent nonlinear dielectric function of A [Eq. (3.2.19). The second component of the dispersion force (RED) depends on the derivative of xAfl(r,r';(o) with respect to RK, and hence on CA"(r,r',r';m,0). We have expressed PEG) explicitly by expanding CA"(r,r',r';m,0) in terms of the charge-density matrix elements of the unperturbed eigenstates and the unperturbed Bohr frequencies. From that, we have separated the terms with j = n and j 36 n. The terms with j ¢ n have been written in terms of the transition susceptibility of A. Then we have showed that this part of the 126 second order dispersion force [designated by FK ] actually results from the d(2),1 correlation between the charge-density fluctuations and the susceptibility fluctuations within molecule A. This result has been given in Eq. (3.2.30), which proves the fact that when nucleus K shifts infinitesimally within molecule A, the change in the Coulomb from K modifies the correlation of the charge-density fluctuations in A and actually introduces new type of correlation, namely, the correlation between charge-density fluctuations and susceptibility fluctuations. Finally, the dielectric screening present in Ffile has been given by Eq. (3.2.3 5), which shows that F312” results from screening of the fluctuating potential from 8 within A and the screening depends on the fluctuation of the nonlocal dielectric function 8v, A(r", O; r, 0)). Terms with j = n in Fcll((2) were separated into two sets. One set of terms where r" is directly connected to RK (denoted by Fci((2) 2) and the other set where either r or r' is directly connected to RK (denoted by F(11((2) 3 ). We have used time-dependent perturbation theory to explain the physical significance of all the terms present in FCIl((2),1’ Fcll<(2),2’ and FC11<(2)I3' The external potential in the perturbation Hamiltonian is the reaction potential from molecule 8. Using the first-order perturbed wave functions for the initial and final states of molecule A, we have showed that 61(0),] and Fcll((2),2 result from the interaction between the transition charge density of A with nucleus K [ Eqs. (3.2.45) and (3.2.46)]. The difference between and F ((ono —iFn /2—o))_—1 n¢0 +(0|(3(r') n>(mn0+iFn /2+co)‘1]. (A.1) The derivatives of the states are given by V§|g> =—(1/h) Z lj)(le<11< H|0>m}i) j¢0 I K‘1 . .. I -1 r—R I Z|J>wj0 (A.2) j¢0 = —(4nsoh)‘1 [dr'zK Vii and similarly I K71 . , -1 r—R ’ 2 |m> = -(47t80h)-1 [dr' ZK VaK where damping is neglected for static perturbations. Also v§ (tong +11“n /2i(o)_1 = -(1/h)(wn0 +irn /2i(o)—2 [—(0|VaK H|o)] —1 = -(oon0 + iFn / 2 i 00-2 (47530”).1 idr' ZK V‘If r, —RKi X (“I50") - Poo (r") | n) I (A-4) 135 where we have used the Hellmann—Feynman theorem. In Eq. (A.4) we have neglected the dependence of the inverse radiative lifetime on the nuclear coordinates, treating I], as a property of the entire adiabatic electronic state. The off-diagonal elements of the derivatives of the charge-density operators themselves vanish, because the only explicit dependence of p(r) on the nuclear coordinates comes from the nuclear contribution to the charge density, and states I j) are orthogonal Bom-Oppenheimer electronic states. From Eqs. (A.2) — (A.4), the derivative in Eq. (A.1) becomes VK x(r,r;—o))— (1/h)2(411:30)_1 Idr' zK H-RKI XIXZI >lnn>I0>w,-‘o1 (wno—irn/z—wr‘ n¢0j¢0 I42 2(0 p(r)|k)(k )n)(n|,3(r')|o)m;;(mn0—irn/2—m)"1 n¢0k¢n XI 2 z <0Ip(r)In>013 ~>o)w;gl nn>< nIéI0>wgJ<>Ikk >n>wggn>I0>w}3 n¢0j¢0 'n)l>|n>|0>(r")—poo(r")In>|n>|0>(n|f>(r">-poo(r")ln>(wn0+irn / 2+w>'2]. n¢0 (A.5) Next, we separate the terms with k = 0, in the summation over k i n; this gives vK x(r,r';w) = (1/15)2 (41:2.0)‘1 Idr'ZK —1 r"—RKI >I1><1I1’1In>l031,701 «one —irn /2-w>“ n¢0j¢0 >|n> n¢0j¢0 XIX Z<0lfilk>< ‘ '> n¢0 k¢0 k¢n ') .00 A ") 0) (03.01 (wno — 11“,, / 2 — (.5)—1 ">1'0> )nn)(n|3(r)|o)m‘j’01(mn0+irn /2—m)’1 n>m3701(cono urn/2+1»)-1 n) (n I p(r') I 0) (31:111 (tong — iFn / 2 — co)—1 n)|0>< ‘ "> n¢0 x[ z z (0|(3(r)In>013, (cone—in. /2—m>‘l n¢0 k¢0 k¢n k>< * )0 137 x [ Z <0Ifi(r)In> <0l13 n) (n I p(r) I 0) (1)611 (con0+ iFn / 2 + (of—1 s(r') k)(k|13(r)|o)m;11 (wno +11“n /2 +111)—1 n) (n x[ Z <0Ip(r')In)(r”) 13(1')I0>0)6r11((0n0+irn/2+03)_1 n¢0 0) (0 x I Z (013(I)|n)(n|,s(I-) axiom) —1rn /2-o))_2 n¢0 0)< [ Z (0 l 66') n) (n l 130') I 0) (n l 130'") - 9000'") l 0) (01110 + irn / 2 + co)—2 ] n¢0 (A.6) 1th We interchange the labels n and k in the 7th and 1 terms in Eq. (A6) and combine the terms, to obtain —1 vK x(r,r';co) = (1/11)2 (471.90)‘1 Idr'zK vK r"—RKI —1 >(nlf1(r') X Z Z (0|13(r>|n>(n|f>(r'>-Poo(r'> j>0|i3(r") —1 o)m.‘1(mn0 —1rn /2—(o) jO n¢0j¢0 138 —1 )nn>< Ip(r)I0>co coj()1(0°n0+irn/7—‘*'(0) X Z Z (0|P(r">| >10 |P(r' n¢0j¢0 XX Z<0|P(r'n><>ln|P(r>-Poo(r>|1'1'><|P(r" 0>mj-01(wn0+irn/2+w)— n¢0j¢0 X Z 20) >PP>< '>0>(‘1 n¢0k¢0 X Z Z <0|P(r'>|k>0< ‘ n¢O k¢0 > n>l0>2+wr1 x((un0 —iFn /2 —m)‘1 ]. (A.7) We assume that the inverse radiative lifetimes l"n and Fk are much smaller than the transition frequencies mnk, (an- Using the identity lim 1, =PI—l—)$i715(y), (A.8) y__)()xi1y x we take the imaginary part of Eq. (A7) and the relabel the summation indices to obtain —1 VK)("'(r,r';o))=(11;/h)2(47t.s())_1 Idr'ZK r'—RKI “XXI n¢0j¢0 )j)0li‘1(r)- 1100(r>|n >01 ) 0)co;01 S(mno — (0) + Z Z IPIn>—pooI1><1I13I0>w;01 5011110 —w> n¢0j¢0 - Z 2 <0|P(r">l >10 |P< Ip(r)I0>o)j018(con0+o1) 139 - Z Z <01P(r'>|n>-poo|j>0|1> —1 0) ij S(mno — 0)) n¢0j¢0 n> + Z Z(0|P(r)lj)5(mn0—m) n¢0j¢0 xRe[(mj0—irj/2—m)‘1] + Z Z <01Pln>-poo8IJ'>0'IP("'>~Poo(r"> n)5(wno+m) n¢0j¢0 xRe[(o)jO+iFj/2+o))_1] - Z Z<0|13(r') n)(nl13(r")-Poo(r") j) (il13(r)l0)5(wno+w) n¢0j¢0 xRe[((ojO+iFj/2+o))_l]}. (A9) A comparison of Eq. (A9) and 8;"(r,r’,r";o),0) yields —1 VK x"(r,r';(o) = (471:80 )_'1 Idr" ZK VK r" — RKI (;"(r,r',r";(o,0) . (A.10) 140 Appendix B: Derivation of the four-body irreducible dispersion energy for the cluster of four interacting molecules In this section we derive the four-body irreducible dispersion energy of a cluster of interacting molecules A-"B-"C-"D using the reaction field approach. We derive the dispersion energy as an integral over frequencies, where the integrand is factored into the charge-density susceptibilities of the interacting molecules at imaginary frequencies. The charge density fluctuation SpA (r';to) in molecule A at r' produces a fluctuating potential that propagates through the surrounding molecules inducing transient changes in their charge densities. The potential which acts back on molecule A at r and time t is the reaction potential, and it’s given by r _ rv11 —1 .. . XD (rvn , l,vr ; (o) 5(pA (r;m) = (41180)—4 sBCD Idr'...drv“ _1 M —1 6pA (r'; (o) . . . —-1 XC (rv , l,rv ; co) r1v _ r XB (rm, I,II; co) Ir. _ rI (8.1) The operator SBCD denotes the sum of the terms obtained by permuting the labels 8, C, and D in the expression that follows the operator. The average energy shift of molecule A due to the interaction between the fluctuating potential S(pA (r;t) and the charge density . A . . fluctuation 8p (r,t) rs ABA = (1 / 2) Idr< 5pA(r,t)6(pA (r,t) >. (8.2) From Eqs. (8.1) and (8.2) along with the Fourier transforms of the charge-density fluctuation and the potential, we obtain 141 I'Vil ABA = (1/2)(47t80)_4 SBCD End") Eda) Idr" drVii rw —l‘ 1 . xC(rv,r'v, xx I)(r ;1'01)v —-r I I—1 xr—r 6pA(r',0))> exp[—i(0)+0)')t. (8.3) + The charge density fluctuations 8pA(r,0)') and 8pA(r',0)) are correlated by the fluctuation-dissipation theorem 1 37 (1/ 2)< SpA (r, 0)') SpA (r', 0)) + 8pA (r', 0)) SpA (r,0)')> = _ (h / 2n) XA'(r, r'; 0)) 500+ 0)')coth (710) / 2kT) . (3.4) Substituting (8.4) in (8.3) and integrating over 0)’ , we obtain .. .. —1 ABA =—(1/2)(h/27t)(47c80)—4 53CD Eodw Idr...drV“r—rV“ x D(rV“, rV';m) . —1 c . 1 3 erI—rv X (rv,r1v, r1v_rIII X (r..,rII; 0)) II I—l A" I . xr —rI x (r,r,(o)coth(h(o/2kT). (8.5) In the limit T —> 0, the hyperbolic cotangent function simplifies to [0(01) - 0(-0))], where 0(0)) is the Heaviside step function (cf. chapter 3). Thus, in the limit T —+ 0, .. ..—1 ABA = —(1 / 2) (h / 21:) (41:80 )—4 sBCD Re fiodco Idr...drv" r—rv" vii vi v v"1 C v iv 1v — 8 xx [’0 ;w)r '—r x (r .r . r -r" x (r",r';w) >O In deriving Eq. (8.8), we have used the identity given in Eq. (A.8). From Eq. (8.8) and the fact that x”(r,r';0)) is an odd firnction of frequency, we obtain x(r,r';0)) = lim (2/ It) dxxx"(r,r';x)[x2 —(0)+iE,)2 ]—1 . (8.9) §—>0 We use Eq. (8.9) in Eq. (8.6), take the sum AEA + AEB + ABC + AED in the limit I; —> 0, and then use the symmetry of the charge-density susceptibility with respect to interchange of its arguments, to obtain AE(4 4): —(3h/21t 4)(4neo)‘4 1:011 [Zodx Eody Eodz Ian. .dl'v“ {[uxyiu2 —22)“ (x2 —22)‘1 (y2 —22)“1+1xyz(x2 -u2)‘l (y2 _u2)—1(22 —u2)‘11 +1uyz< xB" (r'", r”; X) 11"’-r’|—1 xA' (r', r; n) . (B. 10) The frequency integral over u, x, y, and 2 can be separated into independent quadratures using the identity [min2 —22)‘1 (x2 -22)“ (y2 —22)“1+Ixyz