Molecular Interactions - Concepts and Methods
Molecular Interactions - Concepts and Methods
Molecular Interactions
David A. Micha
University of Florida, Gainesville, FL, USA
This edition first published 2020
© 2020 John Wiley & Sons, Inc.
All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or
transmitted, in any form or by any means, electronic, mechanical, photocopying, recording or
otherwise, except as permitted by law. Advice on how to obtain permission to reuse material from
this title is available at http://www.wiley.com/go/permissions.
The right of David A. Micha to be identified as the author of this work has been asserted in
accordance with law.
Registered Office
John Wiley & Sons, Inc., 111 River Street, Hoboken, NJ 07030, USA
Editorial Office
111 River Street, Hoboken, NJ 07030, USA
For details of our global editorial offices, customer services, and more information about Wiley
products visit us at www.wiley.com.
Wiley also publishes its books in a variety of electronic formats and by print-on-demand. Some
content that appears in standard print versions of this book may not be available in other formats.
Limit of Liability/Disclaimer of Warranty
In view of ongoing research, equipment modifications, changes in governmental regulations, and the
constant flow of information relating to the use of experimental reagents, equipment, and devices,
the reader is urged to review and evaluate the information provided in the package insert or
instructions for each chemical, piece of equipment, reagent, or device for, among other things, any
changes in the instructions or indication of usage and for added warnings and precautions. While the
publisher and authors have used their best efforts in preparing this work, they make no
representations or warranties with respect to the accuracy or completeness of the contents of this
work and specifically disclaim all warranties, including without limitation any implied warranties
of merchantability or fitness for a particular purpose. No warranty may be created or extended
by sales representatives, written sales materials or promotional statements for this work. The fact
that an organization, website, or product is referred to in this work as a citation and/or potential
source of further information does not mean that the publisher and authors endorse the information
or services the organization, website, or product may provide or recommendations it may make.
This work is sold with the understanding that the publisher is not engaged in rendering professional
services. The advice and strategies contained herein may not be suitable for your situation. You
should consult with a specialist where appropriate. Further, readers should be aware that websites
listed in this work may have changed or disappeared between when this work was written and when
it is read. Neither the publisher nor authors shall be liable for any loss of profit or any other
commercial damages, including but not limited to special, incidental, consequential, or other
damages.
Library of Congress Cataloging-in-Publication Data applied for
ISBN: 9780470290743
Cover image: © Photobank gallery/Shutterstock
Cover design by Wiley
Set in 10/12pt Warnock by SPi Global, Pondicherry, India
Printed in the United states of America
10 9 8 7 6 5 4 3 2 1
v
Contents
Preface xi
1 Fundamental Concepts 1
1.1 Molecular Interactions in Nature 2
1.2 Potential Energies for Molecular Interactions 4
1.2.1 The Concept of a Molecular Potential Energy 4
1.2.2 Theoretical Classification of Interaction Potentials 6
1.2.2.1 Small Distances 7
1.2.2.2 Intermediate Distances 8
1.2.2.3 Large Distances 8
1.2.2.4 Very Large Distances 8
1.3 Quantal Treatment and Examples of Molecular Interactions 9
1.4 Long-Range Interactions and Electrical Properties of Molecules 21
1.4.1 Electric Dipole of Molecules 21
1.4.2 Electric Polarizabilities of Molecules 22
1.4.3 Interaction Potentials from Multipoles 23
1.5 Thermodynamic Averages and Intermolecular Forces 24
1.5.1 Properties and Free Energies 24
1.5.2 Polarization in Condensed Matter 25
1.5.3 Pair Distributions and Potential of Mean-Force 26
1.6 Molecular Dynamics and Intermolecular Forces 27
1.6.1 Collisional Cross Sections 27
1.6.2 Spectroscopy of van der Waals Complexes
and of Condensed Matter 28
1.7 Experimental Determination and Applications of Interaction Potential
Energies 29
1.7.1 Thermodynamics Properties 30
1.7.2 Spectroscopy and Diffraction Properties 30
1.7.3 Molecular Beam and Energy Deposition Properties 30
1.7.4 Applications of Intermolecular Forces 31
References 31
vi Contents
2 Molecular Properties 35
2.1 Electric Multipoles of Molecules 35
2.1.1 Potential Energy of a Distribution of Charges 35
2.1.2 Cartesian Multipoles 36
2.1.3 Spherical Multipoles 37
2.1.4 Charge Distributions for an Extended System 38
2.2 Energy of a Molecule in an Electric Field 40
2.2.1 Quantal Perturbation Treatment 40
2.2.2 Static Polarizabilities 41
2.3 Dynamical Polarizabilities 43
2.3.1 General Perturbation 43
2.3.2 Periodic Perturbation Field 47
2.4 Susceptibility of an Extended Molecule 49
2.5 Changes of Reference Frame 52
2.6 Multipole Integrals from Symmetry 54
2.7 Approximations and Bounds for Polarizabilities 57
2.7.1 Physical Models 57
2.7.2 Closure Approximation and Sum Rules 58
2.7.3 Upper and Lower Bounds 59
References 60
Index 373
xi
Preface
The temperatures and densities of the chemical elements present in our natural
world lead to the formation of molecules, which interact to create complex
many-atom systems. Many materials and the components of living organisms
are made up of aggregates of atomic and molecular units. Fluids, molecular
and atomic solids, polymers, and proteins are examples of those aggregates.
The properties of these objects can be described from first principles of quan-
tum mechanics and statistical mechanics, after their composition in terms of
electrons and atomic nuclei have been specified.
Molecular encounters are governed by intermolecular forces, which can be
measured and calculated. This book presents concepts and methods needed
to obtain energies of a molecular system, as they change with interatomic dis-
tances to provide intermolecular forces. Once the interaction forces in a molec-
ular system are known, the equations of motion of classical or quantum
mechanics, implemented with physical boundary conditions, can be used to
obtain thermodynamic equilibrium and nonequilibrium properties from first
principles, and the response of the system to external factors, such as light.
Interactions of molecular species can be described by their electromagnetic
properties and chemical reactivity. The reverse is also true. Information about
molecular interaction energies can be derived from measurements of thermo-
dynamic, kinetic, and electromagnetic properties of matter, and from experi-
ments specifically devised to extract interaction energies, such as crossed
molecular beam and photodissociation experiments. This information has been
incorporated into the selection of concepts and methods for the book.
The advent of quantum theory and its applications during the first half of the
twentieth century has led to many of the quantitative concepts about molecular
structure and properties employed today, and how to calculate them, as well as
ways to describe their interaction energies. Molecular dynamics and statistical
mechanics have also provided thermodynamical and kinetics values of proper-
ties to be compared with experimental measurements. Comparisons and tests of
theoretical and experimental results require extensive computational work.
Computational methods, software and utilities, and the ever-increasing speed
xii Preface
and data storage power of computers have been essential for these purposes.
Some of these methods are covered in the following chapters.
Intermolecular forces are essential in many applications of molecular and
materials properties to technologies contributing to the needs of society. To
illustrate the enormous impact of the subject, some of their subjects (and their
applications) are storage of hydrogen in solids (fuel cells), storage and transport
of ions in solids (batteries), synthesis of thermally stable and conducting sur-
faces (solar energy devices), delivery of compounds through biological cell
membranes (pharmacology), catalysis and photocatalysis in electrochemical
cells (sustainable fuel production), atmospheric reactions (environmental
sciences), efficient fuel combustion (transportation and energy), and solvation
and lubricants (machinery). Furthermore, as the quantitative tools of chemistry
and physics in this book have become more useful and common in biology,
pharmaceutics, and medicine, its contents should also be of interest in these
new areas of applications.
This book is based on the author’s lectures given over many years at the Uni-
versity of Florida in Chemical Physics courses taken by graduate and advanced
undergraduate students in Chemistry, Physics, Chemical Engineering, and
Materials Sciences, with working knowledge of quantum and statistical
mechanics as usually covered in Physical Chemistry or Modern Physics courses.
The lectures have emphasized concepts and methods used in calculations with
realistic models, to be compared with empirical data. They have been expanded
to cover recent developments with advanced theoretical methods, as presented
by researchers at conferences and particularly at many Sanibel Symposia on the
quantum treatments of atoms, molecules, and materials, which this book’s
author has helped to co-organize.
The text contains hundreds of citations, but these are only a small portion of
the tens of thousands of relevant publications on the subject of molecular
interactions. To compensate, much can be done to gain additional knowledge,
including physical and chemical data, by searching for information in the
World Wide Web. The concepts and methods described here are meant to also
provide a broad vocabulary as needed to search the Web. To keep the book
length within reasonable bounds, its contents have been restricted to cover
the energetics of molecular interactions and the potential energy surfaces that
are generated as interatomic distances are varied. A short introduction has been
given about how these interactions affect the spectroscopy, dynamics, and
kinetics of molecular interactions, but these subjects are not covered in detail
here and have been left for possible future treatments. References have been
limited to fundamental publications, monographs, reviews, and comprehensive
reports. Apologies are advanced here to the authors of many relevant publica-
tions that could not be mentioned due to length limitations.
Chapters in this book have been organized to cover first the concepts and
methods for simpler systems, and next to cover more advanced subjects for
Preface xiii
complex systems. Each chapter begins with qualitative aspects before these are
treated in quantitative fashion. The first four chapters include the well-
established concepts and quantitative aspects of long-range (electrostatic,
induction, and dispersion) forces and how they extend to intermediate and
short ranges, for ground and excited states. They are followed by chapters deal-
ing with recent developments including electronically nonadiabatic interac-
tions, correlated many-electron treatments, generalized density functional
theory, decomposition and embedding of molecular fragments for large sys-
tems, and very recent developments using artificial intelligence with network
training for many-atom systems. The first four chapters and the first sections
of the following four chapters can provide an introductory course on molecular
interactions, for students with a knowledge of the fundamentals of quantum and
statistical mechanics and of molecular structure. A more advanced course can
be based on material starting with the fourth chapter and continuing with the
following ones.
This author thanks his science teachers at the Physics Institute, Universidad
de Cuyo, Bariloche (Argentina), presently Instituto Balseiro, and at the Quan-
tum Chemistry Group, Uppsala University (Sweden), and particularly Dr. Jose
Balseiro in Bariloche for his introduction to quantum theory and Dr. Per-Olov
Lowdin in Uppsala, for his mentoring in quantum chemistry. Much was learned
from colleagues and researchers at the University of Wisconsin (Madison) and
at the University of California, San Diego, and at institutions visited during sab-
baticals at Gothenburg University in Sweden, Harvard University, Max Planck
Institutes for Stroemungsforschung and for Astrophysics in Germany, Imperial
College in London, England, Institute of Theoretical Physics at the University of
California Santa Barbara, JILA at the University of Colorado in Boulder, the
Weizmann Institute in Rehovot, Israel, Florida State University, the Ecole Nor-
mal Superieur, Paris, France, and the Institute for Mathematics and its Applica-
tions, University of Minnesota, Minneapolis. This author is grateful also to the
many graduate and undergraduate students who did research under his direc-
tion and to postdoctoral associates and visiting scientists working with him at
the University of Florida. They are too numerous to be listed here but their
names can be found at the website https://people.clas.ufl.edu/Micha.
This author is greatly appreciative of the financial support by the Foundations
that provided essential funds for his research and collaborations: the Swedish
International Development Agency, the Alfred P. Sloan Foundation, the
National Science Foundation of the USA for many years of Principal Investiga-
tor research and also for support of US-Latin American workshops, NASA, the
Alexander von Humboldt Foundation Program of U.S. Senior Scientists Awards
twice, and the Dreyfus Foundation.
Thanks are also due to reviewers at the early stages when this book was being
planned, particularly Bernard Kirtman, George Schatz, and Victor Batista. Sev-
eral of the author’s research colleagues have been kind enough to read and
xiv Preface
David A. Micha,
December 2018 Gainesville, FL, USA
1
Fundamental Concepts
CONTENTS
The link between electrons and nuclei and potential energy functions related
to molecular structure is provided by quantum chemistry, and thermodynam-
ical properties follow by using statistical mechanics. Steady-state (or transport)
properties as well as nonequilibrium properties, including reactivity, can be
obtained from molecular dynamics, as shown in the block diagram illustrated
in Figure 1.1. Spectroscopic properties follow from the electrodynamics of
molecular systems. Transport in gases follow from molecular collision cross
sections.
4 1 Fundamental Concepts
Electrons, nuclei,
and photons
Figure 1.1 Properties derived from interacting electrons, nuclei, and photons.
pe pn
ve = vn =
me mn
and to a first approximation nuclei can be assumed to be at rest while electrons
move around them. Fixing the nuclear positions, the molecular energies become
functions of the nuclear coordinates and provide the potential energies for the
nuclear motions; hence, they can be referred to as molecular potential energies.
This is the Born–Oppenheimer picture of molecular structure [14, 15]. These
authors showed, using quantum mechanical perturbation theory, that for bound
molecular states, the potential energy correction due to nuclear motion goes as
(me/mn)1/2, while the correction to molecular wavefunctions goes as (me/mn)1/4,
and therefore are relatively small and acceptable for most applications.
To be more specific, we can compare electronic and nuclear kinetic energies;
for the electronic kinetic energy Kel we find, assuming hydrogenic states for an
electron moving in a framework of atomic cores with an effective (screened)
positive charge, and since then Kel ≈ − Eel for the bound states,
me v2e 1 Zscr
2
e2 1 Zscr
2
Kel = ≈ = 27 2 eV
2 2 n2 aB 2 n2
where e, aB, Zscre, and n are the electron charge in esu units, the atomic Bohr
radius, the screened nuclear charge around which the electron orbits, and the
principal quantum number of the orbital, respectively. The energy units conver-
sion is that 1 au(E) = 27.2116 eV, and 1 eV = 1.60218 × 10−19 J = 96.485 kJ mol−1
for future reference in this work.
The magnitudes of nuclear kinetic energies depend on whether the nuclei are
bound or not. Letting vn ≤ ve/10 be our criterion for separation of electronic and
nuclear motions, we find for the nuclear kinetic energy,
2
mn v2n mn ve 10 1 mn 1 e2 Zscr
2
Kn = ≤ =
2 2 100 me 2 aB n2
where mn ≈2000mn A in terms of an average atomic mass number A. This
expression shows that as the electronic state is excited, and n increases, the
upper energy bound for motion separation decreases, and it becomes more dif-
ficult to satisfy the inequality; the separation of electronic and nuclear motions
is more doubtful for electronically excited states.
For ground and low excited electronic states, we can separately estimate
nuclear kinetic energies for molecular systems with bound and unbound nuclei.
A) Bound nuclei
The nuclei in molecules vibrate and rotate. We can estimate the vibrational
and rotational kinetic energies by means of
1 2
Kvib ≈hνv nv , νv = k μn 2π
Krot ≈ℏ l l + 1
2
2I
in a system where the vibrational frequencies are of magnitude νv (given in
terms of a vibrational force constant k and a reduced mass μn), I is a
6 1 Fundamental Concepts
moment of inertia, and the vibrational and rotational quantum numbers are
nv and l, respectively. These terms would typically add to about 1.0 eV and
would be much smaller than the required upper bound to the nuclear
kinetic energy. However, the bound can be exceeded if the vibrational–
rotational levels are very high, signifying less accuracy in the separation
of electronic and nuclear motions.
B) Unbound nuclei
For colliding atoms and molecules, nuclear momenta include the center of
mass components acquired in the collision, which can be very large for
example in atomic beams. Potential energy functions can then be defined
only provided as before, in eV units, Kn ≤ 272 eV A Zscr
2
n2 For a hydrogen
atom, a potential energy function can safely be defined for kinetic energies
below 272 eV, while for deuterium (A = 2) the upper limit would be 544 eV.
The potentials can also be defined above these values, but one must consider
the possibility that electronic and nuclear motions will be coupled.
can consider the types of contributions to the potential energy, its dependence
with distance R, and the way it can be calculated with theoretical methods.
Table 1.2 summarizes this information.
The calculation of molecular interaction potential energies requires a quan-
tum mechanical treatment starting from the structure of the whole collection of
electrons and nuclei. The Schroedinger equation of motion of the whole system
can be solved to obtain stationary state solutions ΦΓ(X; Q) where the collection
of nuclear position variables Q is considered, to begin with, as fixed at chosen
values. The index Γ is a collection of quantum numbers that also labels the
energy EΓ(Q) of the steady state. These steady states can be calculated with a
variety of methods, including variational and perturbation treatments of the
equations of motion with given boundary conditions [16–19].
At very large distances, the electromagnetic contributions are very small and
may become less important than magnetic (relativistic) contributions coming
from spin couplings, and these must be accounted for.
Effects resulting from the electron spin coupled to its orbital angular momen-
tum are in principle present at all distances, particularly for open-shell systems.
They can be treated by means of a quantal perturbation treatment for the states
ΦΓ(X; Q) due to spin-orbit energy couplings, which lead to splittings of inter-
action potential energies.
Combining short- and long-range functional forms for two neutral and
isotropic species gives model potentials, with two common ones being
n 6
6 Rm n Rm
V R = Vm −
n−6 R n −6 R
the exponential-6 potential, with parameters α, Vm, and Rm. The Lennard–Jones
potential function is usually chosen to have n = 12, with Vm = ϵ, and is also writ-
ten using instead the distance R0 = Rm/21/6, where the potential energy is zero so
that V(R) = 4ϵ[(R0/R)12 − [(R0/R)6].
E Q = Φ Q HQ Φ Q = dX Φ X; Q ∗ H Q Φ X; Q
where a normalization Φ(Q) Φ(Q) = 1 has been used. This energy may be
decomposed into contributions from the several terms in the Hamiltonian oper-
ator and is the so-called Born–Oppenheimer or adiabatic potential energy.
Other properties (such as the electric dipole), represented by an operator AQ ,
are also obtained for fixed nuclei as averages,
A Q = Φ Q AQ Φ Q
An accurate treatment of the electronic state must describe the electrons as
moving in a self-consistent field shaped by the antisymmetry properties of the
wavefunction, as given usually by the Hartree–Fock formulation to incorporate
electronic exchange energies [16], and must in addition contain corrections for
the correlation energy of electronic motions, spin-orbit coupling, and magnetic
(relativistic) effects, so that
E Q = Eel Q + Enn Q
Eel Q = EHF + Ecorr + ESO + Emagn
where Enn(Q) is the nuclear–nuclear repulsion energy, and all the terms depend
on the nuclear coordinates.
The electronic state also provides information on electronic density functions
of electronic positions, ρ r ; these can be decomposed into atomic core ρc and
valence ρv electron terms, with the first leading to nuclear screening and to
core–core interactions. Electron exchange leads to a “Pauli pressure,” which
explains the distortion and repulsion of atomic cores as they overlap.
The energy contributions can be described for the case of H2+ in ground and
excited electronic states. Here, we have
K el + Vne φμ r; R = εμ R φμ r; R
e2
Eμ R = εμ R +
4πε0 R
1.3 Quantal Treatment and Examples of Molecular Interactions 11
where εμ(R) and φμ r; R are electronic energy and orbital function at internu-
clear distance R, and we have added the electronic energy of the electron orbital
to the nuclear Coulomb repulsion energy to obtain the total adiabatic potential
energy Eμ(R). Figure 1.2a and b, from [20], shows the electronic orbital
energy εμ(R) and the diatomic potential energy Eμ(R) = Uμ(R) versus the inter-
atomic distance R in atomic units of energy (Hartree energy), with 1 au(E) =
4.35979 × 10−18 J, and length (Bohr radius), with 1 au(L) = 5.2918 × 10−11m
= 52.918 pm. The electronic binding energies are negative and increasing as
the nuclei move apart, with the limits at R = 0 and at infinite R given by the elec-
tronic energies for He+ and H + H+, respectively, and are labeled by molecular
point group symmetry symbols. After adding the positive repulsive energy of the
nuclei in Figure 1.2b, the lowest (or ground) total potential energy Eg(R) shows a
minimum that defines the static equilibrium distance Re and displays a net pos-
itive repulsion energy at short distances.
Similarly, for H2 in ground and excited electronic states, energies depend only
on the internuclear distance R. Figure 1.3 shows the total, electronic plus n–n
repulsion contributions for the ground state Σg+ . Here, the separated atoms
(S.A.) limit is H(1s) + H(1s), while the electronic united atom (U.A.) limit is
He(1s2) and has a lower electronic energy. The contributions are again found
to add to a potential with a minimum, attractive at long distances and repulsive
at short distances. The contributions to the excited Σu+ electronic state add up
instead to a purely repulsive potential and here the united atom limit is
He(1s2p), with a larger electronic energy [22].
A different situation arises for LiF, where potential curves show avoided
crossings between states 11Σ+ and 21Σ+ as seen in Figure 1.4. The electronic
ground state and five electronically excited states have been obtained with a
many-electron treatment including electronic correlation energies. Electron
transfer occurs as the distance R increases, and couples the state 11Σ+ going
asymptotically to Li + F with the state 21Σ+ going to Li+ + F−. A full treatment
of the atomic interactions here requires that one also specifies the interstate
coupling energy due to atomic displacements, leading to transitions between
the two nearly degenerate potential energies.
Potential energies for more than two atoms require special treatments to
account for energy functions of many atomic position variables. Here, we give
some general examples and postpone details to the following chapters. Conti-
nuing with polyatomic intermolecular potentials, consider a molecular system
AB breaking up into fragments A and B. The potential energies are functions of
the nuclear position vectors. The molecular energy (electrons plus fixed nuclei)
is E AB Q = EelAB Q + Enn AB
Q , where the first term is the sum of electron
kinetic energies plus electron–electron repulsion and electron–nuclear attrac-
tion, and the second term is the nuclear–nuclear repulsion energy. The nuclear
position variables are collected in Q = R, q , with R the relative position vector
(a)
0
3σu
–1–
8 1πg 3σg 1
3s, 3p – 2
– to – –
9 2σg 8
(3d) 2σu
1σu to – 0.5
E (hartree)
–1.0
1σg
–1.5
1s –2.0
0 1 2 3 4 5 6 7
He+ R (bohr) H + H+
(b)
0.3
0.2
0.1
3σu
1πg
0
2σu
2σg
U (hartree)
–0.1
–0.2
1πu 3σg
–0.3
–0.4
1σu = σu*1s
–0.5
1σg = σg1s
–0.6
0 2 4 6 8 10 12 14
R (bohr)
Figure 1.2 (a) Electronic orbital energy εμ(R) for ground and excited states of an electron in
the H2+ molecular ion versus the distance R between the two protons, labeled by symmetry
symbols. The limits at R = 0 and at infinite R are the electronic energies for He+ and H + H+,
respectively. Source: adapted from Slater [20]. (b) Total potential energy Eμ(R) = U(R) of ground
and excited states of H2+ versus interatomic distance R. Source: adapted from Slater [20].
1.3 Quantal Treatment and Examples of Molecular Interactions 13
–0.5
–0.6 a3 Σ+g
Σ1 Σ+g
–0.7 B1 Σ +u
U (hartree)
–0.8
–0.9
b3 Σ +u
–1.0
–1.1 X1 Σ +g
–1.2
0 1 2 3 4 5 6 7 8 9 10
R (bohr)
Figure 1.3 Ground and excited total potential energies for several states of H2 versus the
internuclear distance R. The limiting electronic energies for infinite R are ground and excited
H + H energies. Source: from Kolos and Wolniewicz [21]. Reproduced with permission of AIP
Publishing.
between the centers of mass of fragments A and B, and with q = (qA,qB) their
internal and orientation coordinates. We can work in a reference frame with
its z-axis passing through the centers of mass of A and B, so that the interaction
potential energy is given by
VAB R,q = E AB R,q −E A q A − E B q B
where R is the distance between centers of mass, which becomes infinitely large
as the pair breaks up.
Specific examples can be given for polyatomic intramolecular and intermo-
lecular potentials. Starting with intramolecular potentials, we describe cases
where only two nuclear coordinates or degrees of freedom Q1 and Q2 are chan-
ging, which leads to potential energy surfaces (PESs) U(Q1, Q2) in a three-
dimensional depiction. The energy can alternatively be represented by contours
of equal energy, or equipotential isocontours, in the (Q1, Q2) plane.
14 1 Fundamental Concepts
1.00
31Σ+
0.75
2 1Π Li–(1S) + F+(1D)
11∆
0.50
E (hartree)
0.25
Figure 1.4 Potential energies E versus internuclear distance R for ground and excited states
of LiF. The electronic ground state and five electronically excited states have been obtained
with a many-electron treatment including electronic correlation energies. Source: from Yagi
and Takatsuka [23]. Reproduced with permission of AIP Publishing.
r BC
r A,BC r AB r AB
Figure 1.5 Typical PESs versus atomic positions for a triatomic ABC system with a fixed bond
angle between AB and BC. (a) Isocontours for the interaction energy of a noble gas atom A
interacting with a diatomic BC; (b) isocontours for the reaction A + BC AB + C, showing the
location of an activation energy barrier; (c) isocontours for a stable compound ABC in a
potential well, connected to dissociation valleys. Source: from Figure 2.3 of Smith [7].
Reproduced with permission of Elsevier.
Figure 1.5 shows typical energy isocontours of PESs: (i) with a single valley, as
for He + H2, versus the center of mass relative and internal coordinates R = rA, BC
and rBC; (ii) showing a rearrangement, e.g. H + H2 H2 + H, with a saddle point
between two valleys in a PES versus bond distances; and (iii) a well for a stable
species such as H2O connected by a saddle point to valleys for dissociations [9].
1.3 Quantal Treatment and Examples of Molecular Interactions 15
3.0
2.0
yH (Å)
S
1.0
C N
A B
0 • •
–3.0 –2.0 –1.0 0.0 1.0 2.0 3
xH (Å)
Figure 1.6 HCN PES isocontours for the HCN HNC isomerization, with H changing position
on a x–y plane and C and N fixed on the x-axis. The locations of minima are shown as A and B,
for the two isomers, and a saddle point as S. Source: from Hirst [10]. Reproduced with
permission of Taylor & Francis.
Figure 1.6 is an alternative description and shows the HCN HNC isomer-
ization by displaying the interaction energy of the hydrogen atom with the CN
fragment as the atom is displaced between minima near C and N.
Plots of potential energies versus torsion angles for macromolecular systems
also offer examples of surfaces with several minima. For example Figure 1.7
shows several minima of the potential energy of a polypeptide chain versus
the consecutive dihedral angles ψ and φ. In all these cases, figures are only show-
ing the lowest PES. The PESs of electronically excited states would have differ-
ent shapes.
Figures 1.8 and 1.9 give specific examples of dissociation in HCN H + CN
and N + CH, versus bond distances, and of atomic rearrangement from reac-
tants to products in F + H2 FH + H.
More than one PES may be involved in electron transfer or in electronically
excited systems, involving two position variables Q1 and Q2, or more. Two
potential energy functions Vg(Q1, Q2) and Ve(Q1, Q2) for ground and excited
electronic states, drawn in a 3D picture of V versus (Q1, Q2) with isocontours
of constant energy, typically show intersections where couplings due to atomic
displacements can be large [26]. Depending on how the electronic states couple
through those displacements, the surfaces can intersect along a line forming a
seam, or at a point giving a conical intersection. Seams are features of interac-
tion potentials involving electron transfer in collisions between ions and neutral
molecules, such as found in collisions of H+ + H2 [27] and of H2+ + Ar [28]. Con-
ical intersections appear in electron exchange between molecules [29], in iso-
merization [30], and in photoinduced chemistry [31].
16 1 Fundamental Concepts
–180°
–180° 𝜙 180°
3.0
14
RCN (Å)
2.0
12
10
v1 8
v3
1 2 3 4 5 6
1.0
Figure 1.8 HCN PES versus bond distances C–H and C–N for the dissociation HCN H + CN.
Contour 1 is for the energy – 13.75 eV, and other contours are drawn at intervals of 1 eV.
Source: from Hirst [10]. Reproduced with permission of Taylor & Francis.
1.3 Quantal Treatment and Examples of Molecular Interactions 17
rHH (nm)
0.2
+27
0.1 +5
+1.6
Figure 1.10 shows the intersection of two potentials for collinear Ar(1S) + H2+
( Σ) and Ar+ (2P)+ H2 (1Σ) and higher potentials versus the H–H distance when
2
Ar is far removed, and also potentials for ArH+ and ArH as a H is far away. The
potentials for H2+ and H2 versus the H–H distance are seen to cross, but when
Ar is brought in this becomes a seam between two PESs, and their coupling
leads to a linear avoided crossing of the PESs versus the Ar–H distances.
A similar situation arises for H+ + H2 H2 + H+, or H2+ + H, involving
two PESs, versus bond distances, but here the lowest one shows an energy well
for H3+ . The new feature in both cases is that the PESs have an avoided crossing
or “seam” running along the entrance valley, corresponding to electron transfer
as an approaching atom moves in.
Figure 1.11 shows two PESs interacting to give conical intersections in the
isomerization of the C2H2 ethylene molecule. The picture at the top corre-
sponds to ground and excited electronic states for isomerization into HC2H3,
ethylidene, while the one at the bottom shows the ground and excited PESs
for the pyramidalization intermediate C2H2∗ as obtained in [32]. Here, the vari-
ables are g and h, themselves functions of all the atomic positions, with g mea-
suring the strength of the coupling of the states by the atomic displacements,
and h giving the magnitude of energy splitting between the two PESs [31].
Similar treatments and images apply to interactions of atoms and molecules
with an extended system, such as a solid surface or a cavity in a liquid.
Figure 1.12 shows the PES of He approaching the surface LiF(001) of the solid,
with a face-center-cubic (or fcc) structure for a lattice of F− and Li+ ions.
The surface is perpendicular to a z-axis with the origin of a surface plane
18 1 Fundamental Concepts
6 6
2Σ 1Σ
3Σ u 1Π
u
4 4 3Σ
3Π
2Σ
Ar++H+H
Energy (eV)
2 2
1Σ
0 g 0 Ar+H++H
2Σ
g
–2 –2
1Σ
2 4 6 2 4 6 8
RH–H (bohr) RAr–H (bohr)
Figure 1.10 Potential energies for collinear ArH2+ showing their change with atom–atom
distances, while the third atom is far away, with energies given relative to Ar + H+ + H. The
intersection in H–H curves leads to the appearance of a seam between potential energy
surfaces as the Ar atom is brought in. Source: from Baer and Beswick [28]. Reproduced with
permission of American Physical Society.
(x, y) at a F− ion location, a cell cube side of 2.84 Å and a Li+ at the center of the
face. The isocontours reflect the periodicity of the surface and show deep wells
on the x–z plane above each ion, for two chosen y values corresponding to lines
along the negative and positive ions. A similar figure can be drawn for an atom
approaching a molecule adsorbed on a solid surface.
All these PESs display special points where forces are null, defined by
∂E
=0 Q0j = Q 0
∂Qj
for at least one set {Q0j }, which locates maxima, minima, or saddle points; these
can be distinguished by calculating the second derivatives ∂ 2E/(∂Qj ∂Qk) at the
special points. The dynamics of motion on PESs are closely related to their
topography as defined by the geometry of the special points. For example, min-
ima are connected by paths of steepest ascent and descent along which one finds
activation barrier energies, which in turn account for reaction rate magnitudes.
A different conceptual approach is needed to describe the energetics of react-
ing molecules in a medium such as a liquid solvent or a solid matrix, and in par-
ticular electron transfer in liquid solutions [34, 35]. Sets r n of many atomic
1.3 Quantal Treatment and Examples of Molecular Interactions 19
Ethylidene
0.4
0.3
Energy (eV)
0.2
0.1
0.06
0.03
0.0 0.00
or
ct
–0.03
ve
0.03
0.00 –0.06
g
–0.03
h vecto –0.06
r
1.25 Pyramidalization
1.00
Energy (eV)
0.75
0.50
0.25
0.06
0.03
0.00 0.00
r
to
c
–0.03
ve
0.03
0.00 –0.06
g
–0.03
h vecto –0.06
r
Figure 1.11 Conical intersections in the isomerization of the C2H2 ethylene molecule. The
picture at the top corresponds to ground and excited electronic states for isomerization into
HC2H3, ethylidene, while the one at the bottom shows the ground and excited PESs for the
pyramidalization intermediate C2H2∗ as obtained in [32]. Source: from Ben-Nun and Martinez
[32]. Reproduced with permission of Elsevier.
position variables are involved and one must instead work with collective struc-
ture variables Y r n which account for many atom rearrangements for given
thermodynamical constrains such as temperature T, density ρ, and pressure p.
Potential energies VΓ r n ,Γ = g,e, must be replaced by thermodynamical
internal energies UΓ(Y, T, ρ) or Gibbs free energies GΓ(Y, T, p).
20 1 Fundamental Concepts
3.0
b –2.0
–3.0
Z(Å)
–4.0
2.0 –5.0
–6.0
–7.0
–8.0
–9.0
–10.0
1.0 –11.0
–7.0 –12.0 –7.0
–5.0 –5.0
0 0
10.0 10.0
0
0 2.84 X(Å) 5.68
3.0
c –2.0
Z(Å) –3.0
–4.0
2.0
–5.0
–6.0
–7.0
–5.0 –5.0
1.0
0 0
10.0 10.0
20.0 20.0
30.0 30.0
0
0 2.84 X(Å) 5.68
Figure 1.12 He approaching the surface LiF(001) of the fcc solid. The surface is perpendicular
to a z-axis with the origin of a surface plane (x, y) at a F− ion location, a cell cube side of 2.84 Å
and a Li+ at the center of the face. The isocontours in meV values reflect the periodicity of the
surface and shows deep wells on the x–z plane above each ion, for two chosen y values of
0.00 Å (upper panel) and 1.42 Å (lower panel) corresponding to lines along the ions. Source:
from Wolken [33]. Reproduced with permission of Springer Nature.
Figure 1.13 shows sketches of two Gibbs free energy surfaces for ground and
excited electronic states Gg(Y, T, p) and Ge(Y, T, p) of reactants donor–acceptor
pair D–A undergoing charge transfer into products pair D+–A−, such as a
Fe2+/Fe3+ mixed valence compound, in a liquid solution, with avoided conical
1.4 Long-Range Interactions and Electrical Properties of Molecules 21
Excited state
Free energy
Transition hv
state P
TS
Ground state R
Figure 1.13 Gibbs free energy surfaces (FESs) for ground and excited electronic states in a
system containing reactants R donor and acceptor compounds D and A in a liquid solution
and the products P electron transfer ions D+ and A−. (a) Contours of FESs versus collective
variables Y1 and Y2 chosen to be average liquid cavity radii around fragments D and A,
showing minimum energy paths. (b) Free energy along the Ys reaction path coordinate
through a transition state between the two minima. Source: from Schapiro et al. [36].
Reproduced with permission of Royal Society of Chemistry.
intersections similar to the ones mentioned above now shown instead for free
energies. The variables chosen here are average liquid cavity radii RD = Y1 and
RA = Y2 around donor and acceptor fragments. The ground surface has two
wells, for ground electronic states of the D–A and D+–A− arrangements, con-
nected by a path Ys at the surface through a saddle point. The reaction energy is
ΔGR and the splitting between electronically adiabatic free energy surfaces is
given by 2GR0 . The upper surface displays a conical shape with a single minimum
for the average radii of cavities around the two fragments interacting when the
electronic state is excited.
Molecules usually have permanent electric dipole vectors D, except when they
have specific symmetries, such as inversion symmetry, which make the
dipole null. Dipole physical units are Cm in the SI system, or the Debye, with
1 D = 10−18 cgs-esu = 3.335 × 10−30 Cm. Molecular dipoles are anisotropic and
22 1 Fundamental Concepts
αω = 2 ℏ n 0
Dn0 2 ωn0 ωn0 2 −ω2
0 0
with ℏωn0 = En −E0 . As ω increases, this function goes through infinite values
at the transition frequencies and shows regions of nearly constant values in
between for each type of motion: rotational, vibrational, and electronic, corre-
sponding to the range of frequencies for radio, microwave, infrared, visible, and
UV waves. A more detailed treatment accounts for the lifetime of excited states
and gives finite polarizability values at transition frequencies, as shown in the
following chapter.
m n
m, n
2m A− 2 R = fAB ωA ,ωB
n A B
B VAB
4πε0 Rm + n + 1
Induction energies are obtained from the interaction of an induced multipole
with a permanent multipole. For the (induced dipole)–dipole interaction,
B
this involves the induced dipole DA = αA A and the electric field
B
A = −DB 4πε0 R3 ) created at A by dipole B. The interaction energy is built
B
increasing the field by increments d A as the polarized species is brought
B
closer to the permanent multipole so that in increments dVD D = − DA d A ,
which after integration over field values gives, using the polarization volume
of A,
2
B
αA A αv, A DB 2
VD D R = − =−
2 2πε0 R6
Dispersion energies originate in (induced multipole)–(induced multipole)
interactions, when charge fluctuations create transient dipole. For the interac-
tion between transient dipoles D A = αA A and D B = αB B , where each field is
created by the transient dipole at the other molecule, the energy builds up to
2 2
B A
αA A αB B CDD
VDD R = − − =−
2 2 R6
which is the van der Waals attractive long-range interaction, with CDD > 0. This
coefficient can be obtained from a quantum mechanical treatment and can be
approximated in terms of the polarizabilities and ionization potentials of species
A and B, as shown in following chapters.
FT,V , = n
F n wn T , V ,
i∂
F = Φn F Φn = dQ Φn Q F Q, − Φn Q
n
∂Q
When the operator is the energy of the system, the above procedure leads to
expressions for the free energy and internal energy. For a closed system of
molecules per unit volume, at given temperature and with volume V, the Helm-
holtz free energy ,V , T is obtained from the thermal partition function
Qth ,V ,T = n exp − En kB T by means of [37]
= −kB T ln Qth ,V ,T
from which the internal energy follows as ,V ,T = −T ∂ ∂T V .
Changing variables from volume in to pressure p gives the Gibbs free energy
,p,T .
approximately given by the Lorentz formula loc = + 3ε0 . For polar mole-
cules one must also obtain the average dipole per unit volume, related to the
permanent molecular dipole. This gives D T = D2 loc 3kB T for the thermal
average of the permanent molecular dipole D, to which one must add the
induced dipole. The average permanent dipole is the limit, for small fields ,
of a more general Langevin formula for the average [16].
In a dielectric (a fluid which does not conduct electricity), the presence of
polarization alters the Coulomb interaction between charges, by changing
the electric permittivity from ε0 to ε = ε0εr. The relative permittivity εr is given
by the Debye equation
εr − 1 ρ m
=
εr + 2 M
NA D2
m = α+
2ε0 3kB T
where NA is the Avogadro number, ρ is the mass density, M the molar mass, and
m is the molar polarization. This formula can be used also for dynamical pert-
mitivities when the field oscillates with a constant frequency, and also gives the
electrical susceptibility insofar as χ el = εr − 1. When the permanent dipole is
null, the equation gives a way to measure the polarizability, extracting it from
optical measurements of the refractive index nr(ω) = c/cmed(ω) = εr(ω)1/2, where
c is the speed of light in vacuum and cmed in the fluid medium. The average
polarization m ω depends on the intermolecular forces in the fluid and pro-
vides information about them.
VN = v
1 ≤ j < k ≤ N jk
in which case the internal energy of the fluid can be obtained from a pair cor-
relation function g2(R, Ωj, Ωk) for molecules j and k at a distance R and with
orientations given by angles Ωj and Ωk. Introducing a thermal average g 2 R
over a distribution of orientations (Ωj, Ωk; T), the average total potential energy
Epot = pot ,V ,T is a function of the temperature and density which can be
given in terms of the pair correlation function g 2 R; T , by [37]
pot N2 3
, V ,T = d R v R g 2 R; T ,
2V
It is also useful to introduce a statistical pair potential of mean force
w2 R; T , by means of g 2 R; T , = exp −w2 R; T , kB T . It can be
1.6 Molecular Dynamics and Intermolecular Forces 27
Calculations of these cross sections as functions of ΩP, vrel and the channel
states, α, starting with the interaction potential energies and the structures of
reactants and products, can be compared to experimental measurements to
provide very detailed information on potential energy functions and
parameters.
These calculations can be very demanding of computational resources. In
some cases, when transitions among rotational, vibrational, and electronic
states of the molecules have been averaged, calculations can be done with a clas-
sical mechanics treatment using equations of motion for classical positions and
momentum variables Q(t), PQ(t) of the constituent atoms. More accurately, the
state-to-state cross sections must be obtained from quantal treatments which
may involve time-dependent wavepackets or time-independent (stationary)
wavefunctions. The stationary wavefunction ΦE for a total energy
α
E = Erel + Eint is a function of nuclear an electronic position variables. Its asymp-
totic form as the distance R between the products C and D becomes large can be
written as ΦE ≈ Φ(0) + Φ(sc) for R ∞, with Φ(0) giving the incoming flux Jα0 and
the scattered wave Φ(sc) giving scattering amplitudes fβα(ΩP, vrel) from which the
L 2
differential cross section Iβα ΩP ,vrel = pβ pα fβα ΩP , vrel can be calculated
[40]. The wavefunction ΦE satisfies the time-independent Schroedinger equa-
tion with a Hamiltonian operator which contains the potential energy of the
interacting reactant and product species, and therefore is shaped by the inter-
molecular forces.
A related subject is transport rates of molecular momentum (viscosity),
kinetic energy (thermal conductivity), and mass (diffusion) in real gases and
liquids. They are linked to intermolecular forces through collision integrals
which can be calculated in many realistic situations, particularly when thermal
averages allow use of classical mechanics and deflection functions [1, 8, 33].
tunneling effects, due to small energy variations among minima, maxima, and
barriers in the intermolecular potentials.
The treatment of large amplitude motions in van der Waals complexes can be
done separating the nuclear position variables into a set of position values {uj} =
u which give small displacements from equilibrium positions for rigid structural
regions of the complex, and a remaining set {qk} = q of position variables which
can range over large distances. The complex dynamics and related spectra are
treated starting from a Hamiltonian containing the potential energy V(u, q) and
kinetic energy terms Ku, Kq, and Kuq. Terms containing u are expanded in its
powers and treated as done for standard molecular spectra, while Kq and Kuq
can be expressed in terms of a mass tensor G = Gkk q,q and generalized
momenta p = − iℏG − 2 ∂ ∂q G2 conjugate to q, with G = det[G]. Examples of
1 1
applications can be found for van der Waals complexes like ArCH4 and ArC6H6
and hydrogen bonded complexes like (H2O)2 and (HCl)2 [42, 43]. This approach
is also applicable to the study of transient (or collision-induced) complexes
detected through light absorption–emission in a collision region. The subject
of van der Waals complexes continues to be quite active and has been recently
reviewed in [44].
The mentioned treatment can be generalized to include scattering boundary
conditions applicable to photodissociation of van der Waals complexes, which
can be considered to be half-collisions and can be described as mentioned above
with channel amplitudes for collisional phenomena [45].
Intermolecular forces participate in molecular spectra in condensed phases
such as real gases, liquids, and molecular solids. They shape absorption–
emission lines, Raman spectra, and electron transfer spectra through solvation
effects. These spectra can be calculated and compared with experimental mea-
surements to obtain information about the functional form and parameters of
involved interaction potential energies [46–48]. Some of these aspects are
considered in Chapter 7 on interactions of two many-atom systems.
and for state-to-state collisional transitions, insofar it deals with isolated sys-
tems of one or two molecules, but it is limited by the necessity to produce
and detect beams of interacting species in known internal states. Spectroscopic
methods are even more accurate, but they mostly provide information around
equilibrium conformations. This is changing with the recent introduction of
time-resolved spectroscopy methods. Following are lists of some relevant meas-
urable properties.
• Molecular spectra of weakly bound (van der Waals) atomic and molecular
complexes formed in beams.
• Collision-induced spectra in gases versus light wavelengths at given temper-
ature T and pressure p.
• Line shapes in condensed matter versus light wavelengths at given tempera-
ture T and pressure p.
• Relaxation times of quantum state populations in condensed matter versus
temperature T and pressure p.
• Electron, neutron, and light diffraction intensities in condensed matter versus
temperature T and pressure p.
• Time-resolved spectral intensities versus light wavelengths in pulsed laser
(pumping–probing) experiments.
• Cross sections for elastic, inelastic, and reactive collisions in crossed molec-
ular beams.
• Energy loss cross sections for beams traversing fluids or solids.
References 31
References
1 Hirschfelder, J.O., Curtis, C.F., and Byron Bird, R. (1954). Molecular Theory of
Gases and Liquids. New York: Wiley.
2 Hirschfelder, J.O. (ed.) (1967). Intermolecular Forces (Adv. Chem. Phys. vol. 12).
New York: Wiley.
3 Margenau, H. and Kestner, N.R. (1971). Intermolecular Forces, 2e. Oxford,
England: Pergamon Press.
4 Dainton, F.S. (ed.) (1965). Faraday Society Discussions (Intermolecular Forces
vol. 40). London, England: Royal Society of Chemistry.
5 Lawley, K.P. (ed.) (1980). Potential Energy Surfaces (Adv. Chem. Phys. vol. 42).
New York: Wiley.
6 Pullman, B. (ed.) (1982). Intermolecular Forces. Dordrecht, Holland: Reidel
Publishing Company.
7 Smith, I.W.M. (1980). Kinetics and Dynamics of Elementary Gas Reactions.
London, Great Britain: Butterworth.
8 Maitland, G.C., Rigby, M., Brain Smith, E., and Wakeham, W.A. (1981).
Intermolecular Forces. Oxford, England: Oxford University Press.
9 Murrell, J.N., Carter, S., Farantos, S.C. et al. (1984). Molecular Potential Energy
Functions. New York: Wiley.
32 1 Fundamental Concepts
10 Hirst, D.M. (1985). Potential Energy Surfaces. London, England: Taylor &
Francis.
11 Stone, A.J. (2013). The Theory of Intermolecular Forces, 2e. Oxford, England:
Oxford University Press.
12 Israelachvili, J. (2011). Intermolecular and Surface Forces, 3e. New York: Elsevier.
13 Kaplan, I.G. (2006). Intermolecular Interactions, 81. New York: Wiley.
14 Born, M. and Oppenheimer, J.R. (1927). Zur Quantentheorie der Molekeln. Ann.
Phys. 84: 457.
15 Born, M. and Huang, K. (1954). Dynamical Theory of Crystal Lattices, Appendix
VIII. Oxford, England: Oxford University Press.
16 Atkins, P.W. and Friedman, R.S. (1997). Molecular Quantum Mechanics.
Oxford, England: Oxford University Press.
17 Schatz, G.C. and Ratner, M. (1993). Quantum Mechanics in Chemistry.
Englewood, New Jersey: Prentice-Hall.
18 Cohen-Tanoudji, C., Diu, B., and Laloe, F. (1977). Quantum Mechanics, vol. 1
and 2. New York: Wiley.
19 Messiah, A. (1962). Quantum Mechanics, vol. 1 and 2. Amsterdam: North-
Holland.
20 Slater, J.C. (1963). Quantum Theory of Molecules and Solids, vol. 1. New York:
McGraw-Hill.
21 Kolos, W. and Wolniewicz, L. (1965). Potential‐energy curves for the X Σg ,
1 +
Molecular Properties
CONTENTS
where summations over I display in order from the left: total charge C = ICI ,
total dipole Cartesian component Dξ = IξICI , total quadrupole, and so on. We
use the relation Ce = e/(4πε0)1/2 between the electron charge Ce and its symbol e
in esu units, with ε0 the vacuum permittivity.
These expansions can be rewritten in terms of the electric field vector com-
ponent ξ = − ∂ϕ ∂ξ, and its space derivatives, and are valid whether the electric
field is internal to a molecular system, or externally applied. Simplifications
arise when the electric field is external and the molecule is not large so that
the field can be assumed to be constant over its extent. Then the quadrupolar
and higher terms can be omitted and a quantal treatment of the molecular
properties can be derived from a quantized Hamiltonian operator
H MF = ϕ 0 I CI − ξ ξ I ξI CI , where operators are shown with car-
ats [1–4].
Dξ = I
C I ξI
and its units are Coulomb × meter (or Cm) in the SI system or e.a0 in atomic
units; another usual unit is 1 Debye = 1 D = 10−18 esu. The relation among them
is 1 au(dip) = 2.5418 D = 8.478 × 10−30 Cm. The quadrupole tensor component
along (ξ, η) is defined by
Qξη = I
CI 3 ξI ηI − rI2 δξη
where δξη equals one for ξ = η and zero otherwise, and rI2 = x2I + y2I + zI2 ,
constructed so that the trace ξQξξ = 0. An alternative definition of the
quadrupole in the literature is Θξη = Qξη/2.
In general, the Cartesian components ξ, η, … ζ of the n-th electric multipole
are given by
n
n −1 ∂ ∂ ∂ 1
Mξ, η, …ζ = CI rI2n + 1
n I ∂ζ ∂η ∂ξ r I
which is traceless when adding over a pair of identical indices, as follows from
the relation ∇2(1/r) = ξ ∂ 2 r−1/∂ξ2 = 0.
1 2
4π
Qlm = CI rIl Ylm ϑI ,φI
2l + 1 I
written in terms of the spherical harmonic function Ylm of position angles, with
well-known transformation properties under rotation. From Y00 ϑI , φI =
1 4π one finds that Q00 is the total charge. Dipole and quadrupole multipoles
follow from l = 1 and l = 2, respectively, and are related to Cartesian compo-
nents, so that, for example Q11 = Dx + iDy 2 = Q∗1, −1 and Q10 = Dz. The
potential energy of interaction takes the form VMF = lm(−1)mQlmFl, −m, where
the coefficients are related to the electric field and its derivatives. See [4, 6, 7] for
details.
38 2 Molecular Properties
for large molecules; (b) decomposition into fragments densities for molecules in
liquid solutions and at surfaces; (c) plane-wave decompositions for extended
systems with many atoms; and (d) decompositions into finite elements for con-
tinuum models. These are described in some detail as follows:
a) The charge density operator can be represented by a matrix in a basis set of
atomic orbitals, and matrix elements can be classified as atomic- or bond-
like. If an electronic structure can be accurately described by a basis set of
atomic orbitals χ νn centered at nucleus n and with atomic quantum numbers
ν, then a related basis set of localized orthonormal orbitals φνn can be intro-
duced using the overlap matrix Δ with elements Δμm, ν, n = χ μm χ νn and the
m, n
symmetric transformation orbital φνn = m, μ χ μm Δ−1 2 , which is
μν
large at location n and contains small contributions of nearby atomic orbi-
tals. It provides a partial completeness relation I = mμ φμm φμm . With
the identity I operating to right and left of the charge density operator, this
becomes a sum over single atom and atom-pair terms,
2
c el r = m, μ
φμm φμm r φμm
∗
+ mμ nν
φμm φμm r φνn r φνn
The expectation value cαel r for a given electronic state α can be interpreted
as an atomic charge- and bond-order sum, and terms can be given physical
meaning [8].
b) Molecular fragments are identified by a subset of nuclear positions, usually
chosen with well-known structure, each with its own multipoles and polar-
izations constructed to reproduce the fragments density, and the total charge
density can be given for the combined fragments in the form of a sum over all
the distributed multipoles and polarizations [4]. In each fragment, the multi-
poles and polarizations are located at chosen points a and are expressed in
terms of localized functions. These are conveniently written using Gaussian
exponentials, which can be readily translated between locations, of the form
1 2
4π
La lm r = ral Ylm ϑa , φa exp −ζra2
2l + 1
where (ra, ϑa, φa) are angular variables for r − a.
c) The density can be expanded in plane waves for a collection of vectors
selected to form a grid in a reciprocal space. This is convenient when the
charge densities are smoothly varying, which excludes ion core electrons.
It is usually done enclosing the molecule in a volume Ω = L1L2L3 with
40 2 Molecular Properties
and the operator coefficients are used in interactions. This expansion is com-
mon in treatments of solid-state properties and is particularly useful for poly-
mers, surfaces, and solids with periodic atomic structure [9].
d) The density can be numerically represented by a decomposition of a contin-
uous distribution of charges into three-dimensional finite elements. Each
element is a solid tetrahedron P3 = a0 + a1x + a2y + a3z with coefficients
fixed by density values at the four corners. A set of tetrahedra can cover a
general distribution of the density and can be used in the treatment of
two interaction densities. Finite elements have been introduced in treat-
ments of optical properties of atomic clusters [10] and are convenient also
for biomolecular distributions.
Molecular interactions can be described in terms of these density decomposi-
tions. They are all approximations, and their accuracy must be tested by system-
atic improvement of the decomposition procedures.
is perturbed.
Solutions can be generated using operator methods, or in more detail intro-
ducing an expansion in the basis set of unperturbed states. This expansion gives
physically meaningful expressions involving matrix elements of the multipole or
charge density operators. The ground and excited states must be generated and
the treatment is accurate provided enough excitations can be considered.
A general and compact perturbation treatment can also be done introducing
resolvent operators and a basis set partitioning method. This also facilitates treat-
ment of perturbation of a degenerate energy eigenvalue, with several unperturbed
states of this same energy [6, 11–13]. An operator-based treatment is needed
when the molecule is very large and its relevant unperturbed states are not all
known. In this case, it is yet possible to account for property values by expanding
in chosen basis sets, or with an expansion in a set of operator amplitudes.
and equating coefficients of each power one finds for n = p + q ≥ 1 that, after
0∗
multiplication by Φk and integration over its variables,
1 0 1 0 1 0 1 0 0 0 0
Ek = Φk H Φk , Φk = l k
Φk H Φl Φl Ek − El
1 2
2 0 0 0 0
Ek = l k
Φk H Φl Ek − El
0 1
Dξ a
= Dξ , a + α
η ξη, a η
+ β
η, ζ ξηζ , a η ζ+
2
∂ 2 Ea
αξη, a = −
∂ ξ∂ η 0
∂ 3 Ea
βξηζ, a = −
∂ ξ∂ η∂ ζ 0
0 0
Φa0 Dξ Φl Φl Dη Φa0 +c c
αξη, a = l a 0 0
El − Ea
which is positive valued when a is the ground state, of lowest energy. Here c.c.
stands for complex conjugate of the preceding form. The hyperpolarizability
requires calculation of the third-order correction to the energy and can be found
from recursion relations.
The static polarizability is related to the transition dipole for each state-to-
state transition, and the largest contributions to the polarizability come from
large transition matrix elements and small excitation energies. They can be cal-
culated separating nuclear and electronic dipole operators and using the Born–
Oppenheimer factorization of states.
The Qξη a Cartesian components of the total quadrupole are similarly
obtained differentiating the total energy with respect to the field gradient tensor
components ξη = −∂ 2 ϕ ∂ξ∂η, to obtain the permanent molecular quadrupole
plus its quadrupolar polarization components.
An alternative description is based on the expansion of the molecule-field
interaction in terms of the spherical components of the dipole operator and
of the electric field. Derivatives of the energy with respect to the spherical com-
ponents of the electric field give the total spherical components of the dipole, as
sums of a permanent value plus a value derived from the polarizability and con-
tributions from hyperpolarizabilities [5]. Using spherical components, it is pos-
sible to give compact expressions for higher multipoles [4].
Molecular multipoles can be calculated as quantal averages so that the total
n n
molecular n-pole average is Mξ, η, …ζ = Φa M ξ, η, …ζ Φa , a function of the
a
applied electric field. Permanent n-pole and polarizabilities follow from the per-
turbation expansion keeping up to second-order terms.
0 1
H + λH t Ψ t = iℏ∂Ψ ∂t
−1
t † , its
0 0
This operator is unitary, and its inverse satisfies U t =U
0
adjoint. Writing Ψ t = U t Ψ I t and solving the full equation of motion
for the interaction picture states Ψ(I)(t), which are constant in the absence of
a perturbation so that Ψ(I, 0)(t) = Ψ(I)(0) and have the initial value Ψ(I)(0) = Ψ(0),
gives the integral equation
t
i
t †H
0 1 0
Ψ I t =Ψ I 0 − λ dt U t U t ΨI t
ℏ 0
2.3 Dynamical Polarizabilities 45
and one has in the interaction picture to n-th order the recursive relation
t
i
t †H
0 1 0
Ψ I, n t = − dt U t U t Ψ I , n− 1 t
ℏ 0
Replacing results in the average total dipole, one finds the results for
permanent and induced dipoles. Writing for the dipole the ex-
0 1 2
pansion Dz t a = Dz, a t + λ Dz, a t + λ2 Dz, a t 2 + … one finds that
0
Dz, a t = Ψa0 t Dz Ψa0 t , and that
This first correction gives the induced dipole (or polarization) Pz, a t =
1
λDz, a t as
t † Dz Ψa0 t
0 0
Pz, a t = λ Ψa0 t Dz U t ΨaI , 1 t + λ ΨaI , 1 t U
Ψt = k
k exp −i Ek t ℏ ck t
dck
iℏ = k H t l −Ek δkl exp iωkl t cl t
dt l
46 2 Molecular Properties
where ωkl = (Ek − El)/ℏ is a transition frequency, and the coupled equations
must be solved with the initial conditions ck(0) = δka for each initial state a.
n
Expanding in powers of the field strength with ck t = n ≥ 0 λ n ck t it follows
0 0 n
that ck t = ck 0 = δka and therefore ck 0 = 0 for n ≥ 1. The changes of
higher-order terms for n ≥ 1 satisfy the iterative equations
n
d ck 1 n −1
iℏ = k H t l exp iωkl t cl t
dt l
1 t 1
which gives ck t = 0 dt k H t a exp iωka t , linear in the field and
from an integral to be obtained for specific fields.
The average dipole takes values obtained from the coefficient. To the lowest
0
order, Dz, a t = a Dz a which is usually null by molecular symmetry of the
state a. To the next order
1 1 ∗
Dz1, a t = k
a Dz k ck t exp − iωka t + ck t k Dz a exp iωka t
is linear in the field strength and provides the polarizability. The induced dipole
or polarization is given by
t
Pz, a t = k
dt a Dz k k Dz a z t exp iωka t − t + c c
0
which can be integrated for a given time-dependent field.
More generally, the polarization component Pξ, a(t) arises in response to an
applied field η t present from time t = 0 at all times 0 ≤ t ≤ t, and can be
expressed in terms of delayed response functions for phenomena both linear
and nonlinear in the applied field, as [15]
t
1
Pξ, a t = η
dt1 αξη, a t,t1 η t1
0 t t1
2
+ η, ζ
dt1 dt2 αξηζ, a t,t1 ,t2 η t1 ζ t2 +
0 0
written here in general using tensor components of response functions. To sim-
plify, consider only the linear response in the first term, usually sufficient for stud-
ies of intermolecular interactions. For a time-independent unperturbed state a, it
is a function of the difference t = t − t1 and it is convenient to introduce the
+ 1
retarded susceptibility αξη, a t = θ t αξη, a t ,0 , where θ(t ) is the step function
null at negative times. Furthermore, the lower integration limit can be replaced
by −∞ insofar as the electric field is null for negative times, and changing the
integration variable from t1 to t gives the useful expression for the linear term,
∞
+
Pξ, a t = η
dt αξη, a t η t −t
−∞
2.3 Dynamical Polarizabilities 47
which is a convolution form and has a Fourier transform from time to fre-
+
quency, giving P ξ, a ω = η α ξη, a ω η ω in terms of the transforms of the
three expressions, using the definitions
∞ ∞
dω
f ω = dt exp iωt f t , f t = exp −iωt f ω
−∞ − ∞ 2π
+
for a function f(t). The dynamical polarizability α ξη, a ω can be obtained from a
perturbation calculation of the average polarization. The retarded susceptibility
is in general
+ i
αξη, a t = − Ψ0 0 Dξ t Dη 0 − Dη 0 Dξ t Ψa0 0 θ t
ℏ a
where the time-dependent dipole density operator Dξ t = U t † Dξ U t
0 0
Dz(t; ωL) a = Dz, a(0) + Pz, a(t; ωL), the sum of the permanent dipole plus a polar-
ization changing with the same frequency and given in general by [15]
t
Pz, a t; ωL = dt1 χ zz1, a t,t1 ; ωL z t1 ; ω L
0
t t1
+ dt1 dt2 χ zzz
2
, a t,t1 ,t2 ; ωL z t1 ; ω L z t2 ; ω L +
0 0
which displays a dipolar polarizability in the term linear with the field and a first
hyperpolarizability in the quadratic term, as functions of the field frequency. To
obtain explicit expressions for them, we perform the integrations over time.
One can introduce a specific electric field along the z-direction in the form
z t; ωL = 0 cos ωL t 1−exp − t τ and integrate the previous expression
1
for ck t which give for t τ > 0
ωka + ωL t ωka − ωL t
1 0 ei −1 ei −1
λck t = k Dz a +
2ℏ ωka + ωL ωka − ωL
where k < a signifies adding over states of lower energy, and ρL is a density of
light per unit frequency. With the same time-dependent coefficient changes
2.4 Susceptibility of an Extended Molecule 49
used in the dipole average, one finds its previous stationary term replaced by a
d
decaying dipole Dz, a t = a Dz a exp − γ a t , and that the transition fre-
quency is replaced by ωka + iγ ka/2 with γ ka = γ k + γ a the sum of initial and final
decay rates, in the first-order coefficients. With this replacement, the dynamical
polarizabilities becomes
2
2 ωka k Dz a
αzz ωL a=
ℏ k a ωka − ωL 2 + γ 2ka 4
which avoids the function’s singularity at the transition frequencies and gives
instead peaks near resonance ωL = ωka excitations.
The present treatment can be readily generalized to the case where the field
and dipole components are not in the same direction. For a general orientation
of the field, the dynamical polarizabilities become tensors and the total dipole is
0
Dξ t a
= Dξ, a + Pξ, a t; ωL
Pξ, a t; ωL = α
η ξη, a
ωL 0η cos ωL t
1 ωka a Dξ k k Dη a
αξη, a ωL =
ℏ k a ωka −ωL 2
+ γ 2ka 4
with similar expressions for the quadrupolar and higher dynamical polarizabil-
ity tensors, replacing the dipole operator with a higher electrical multipole.
More generally, the decay rates originate in the interactions of the molecule
with its medium, with light, and with other molecules, and the rates can be
obtained extending the expansion basis set to include states with a continuum
of energies, which describes the photons and the molecules in its medium.
Solving the larger set of coupled equations for the expansion coefficients, the
interactions are found to give in addition an energy shift ΔEka and new transition
frequencies ωka = ωka + ΔEka ℏ [6].
An alternative treatment of decay can be developed quantizing the electro-
magnetic field, which introduces photons instead of the field and can describe
the interaction of the molecule with photons, including stimulated absorption-
emission and also spontaneous emission [14, 17].
which is a convolution form and has a Fourier transform from time to frequency
+
giving c a r,ω = η d 3 r1 χ a r,r 1 ; ω ϕ r 1 ,ω in terms of the transforms of
the three expressions.
+
The dynamical susceptibility χ a ω can be obtained from a perturbation
calculation of the average polarization. Proceeding as before for the dipolar
response, now the charge polarization is found to be
i t
ca r, t = − dt d 3 r Ψa0 0 c r, t c r , t − c r , t c r, t Ψa0 0 ϕ r , t
ℏ 0
multiplying at two different time. The bracket is in fact a function of only the
difference τ = t − t because the time evolution is done by a time-independent
Hamiltonian, and this provides a result with a real valued retarded response
function
i
χ cc+, a r, r ,τ = − Ψ0 0 c r, τ c r ,0 − c r ,0 c r,τ Ψa0 0
ℏ a
+
for τ > 0 and χ cc τ = 0 for τ < 0. The lower integral limit can be replaced with
−∞, and the integration variable can be changed to τ giving
∞
ca r,t = dτ d 3 r χ cc+, a r, r ,τ ϕ r ,t − τ
−∞
+
and the two sides can be Fourier transformed into c a ω = χ cc, a ω ϕ ω , which
are complex valued functions of frequency.
+
The response function χ cc, a t can be calculated in a variety of ways, using
expansions of operators in a known basis set, or generating numerically a solu-
tion over time. It can also be constructed semiempirically to incorporate known
features of the response. The response functions can be calculated introducing a
convenient basis set { μ } of states to expand operators and the reference unper-
turbed state Ψa0 , in which case they can be expressed in terms of the amplitudes
μ c r,t μ = cμμ r,t . The response can alternatively be obtained from states
52 2 Molecular Properties
+
αξη, a τ = d 3 r d 3 r ξ η χ cc+, a r,r ,τ
which are moment integrals of the susceptibility with respect to positions. Sim-
ilar expressions can be obtained for higher multipolar polarizabilities from
higher moment integrals. In an extended system, the integrals over space vari-
ables can be calculated with coarse-grained numerical procedures that can be
adapted to the physical distribution of electronic densities and localized polar-
izabilities in atomic or bond components.
while keeping the axes parallel, so that the origin position R CMN =RCMN + L, and
charge positions relate instead by r I =r I − L. This gives for the average dipole
component ξ in the BF frame,
Dξ = Dξ − CLξ
which shows it has changed due to the translation, if the total charge C = IcI is
different from zero. For the quadrupole Qξη = I CI 3ξI ηI − rI2 δξη one finds
which changes with frame translation if either total charge or total dipole is not
zero. The conversions between frames when the molecule is moved, instead of
moving the frame, follow by changing L into − L in the above equations.
Changes of the n-th multipole involve the n-th power of the Lξ components,
and therefore its changes can be large, and they must be considered together
with all others.
A rotation of the BF while keeping the origins at the same location can be
done moving the frame by the set of Euler angles Ω = (α, β, γ) with respect
to , by angle α around the z-axis of (a right-handed) , followed by a rotation
–β around an intermediate y -axis, and after this a γ rotation around the final z -
axis [20]. For a column 3 × 1 matrix of elements (x, y, z) transformed into a new
column with (x , y , z ), the 3 × 3 transformation matrix A α,β,γ =
Rz γ Ry β Rz α , a product of three axial rotation matrices, with components
Aξη, can be used to transform dipole and higher multipole components and also
polarizations. From Dξ = η Aξ η α,β,γ Dη and ξ = η Aξ η α,β,γ η for the
electric field, using the inverse of the matrix A, one finds for the polarizability
tensor components in the rotated frame,
αξ η = ξ, η
Aξ η αηζ A −1 ζη
which is a function of the Euler angles. The permanent quadrupole has the same
angle dependence. If instead one rotates the molecule keeping the same refer-
ence frame, then the transformations of multipoles and polarizabilities are sim-
ilar but involve the inverse rotation and Euler angles. The usual notation in this
case is (α, β, γ) = (φ, ϑ, χ) for the orientation of the main axes of a rotated
rigid body.
The spherical components of multipoles are simply transformed under a
rotation operation R α,β,γ by means of the Wigner rotational matrices
l
Dm m α,β,γ , so that the transformed multipole is [4, 7]
l
R α,β,γ Qlm = m
Qlm Dm m α,β,γ
54 2 Molecular Properties
groups and their matrix representations provide powerful and general methods
for analyzing multipole and polarization components.
For each element G in a group of symmetry movements, the equation for sta-
el
tionary states H Φ X; Q = E Q Φ X; Q is also satisfied by Φ(GX; GQ), which
el el
means that the electronic Hamiltonian satisfies OG H = H OG , and
OG Φ X; Q can be expressed as a combination of stationary eigenstates of
the Hamiltonian. Stationary states Φμ with the same eigenenergy Er(Q) of
degeneracy lr form a basis set for an irreducible representation (or irrep) Γ(r)
of the symmetry operator by means of matrices of the same order as the energy
degeneracy. A generated set of irreducible matrix representations Γμνr G can be
r
used to construct symmetry-adapted electronic states Φλ with projection
operators so they undergo prescribed changes for given movements [5]. The
projection operators can be simply constructed from the characters
χ r G = μ Γμμr G as
lr
G ∗ OG
r
P = χ r
h G
r r
and project the adapted state Φλ = P Φλ from a given state Φλ. Products of
r
irreps generated by basis set products Φλ Φμs can be decomposed into sums
p
shown schematically as Γ r ⨂ Γ s = p ars Γ p with coefficients derivable from
the related characters [5, 22]. This is done using Tables of characters and the
orthogonality properties of the characters, Gχ (r)(G)∗χ (s)(G) = hδrs.
Symmetry considerations allow determination of what permanent multipole
or polarization components must be null, and if a transition matrix elements
between states must vanish indicating a selection rule. The integrals involved
in quantum brackets are numbers I = dXdQ f(X, Q) clearly invariant under
symmetry operations. Therefore, f(GX, GQ) must give the same integral value
or this must be zero. Consequently, nonzero integrals must involve fully invar-
iant integrands, and these can be identified using symmetry. The identification
of null tensor components is frequently obvious from simple symmetry consid-
erations, especially when the electronic state is unchanged by symmetry
movements.
A homonuclear molecule such as H2 has a center of inversion and as a
results all three components of its average dipole in its ground electronic state,
which is nondegenerate and invariant, must be null because they are identical
to their opposite when the inversion movement is applied. The average dipole
in the H2O molecule must be located along the z-axis through O and the mid-
point between H’s because the perpendicular components are equal to their
opposite upon a 180 rotation around the z-axis, and so on, for other
molecules.
56 2 Molecular Properties
A1 1 1 1 1 z; x2, y2, z2
A2 1 1 –1 –1 Rz; xy
B1 1 –1 1 –1 x, Ry; xz
B2 1 –1 –1 1 y, Rx; yz
2.7 Approximations and Bounds for Polarizabilities 57
1 − iωτD
α zz+, a ω = D2z a 1 + ω2 τD 2
Other physical choices for the dipole time correlation lead to dynamical sus-
ceptibilities with Lorentzian or Gaussian distributions of frequencies. For a dis-
tribution of charges undergoing an oscillation of frequency ωR, and a dipole
+
correlation relaxing over time τR, the choice αzz, a t = D2z a exp − t τR
cos ωR t leads to a Lorentzian susceptibility as above but with ω replaced by
ω − ωR. Another instance involves random changes over time in the charge dis-
tribution and is given by a Gaussian dipole correlation function like
+
αzz, a τ = D2z a exp −t 2 τG 2 from which a Gaussian distribution of frequen-
cies is found for the dynamical susceptibility [23].
58 2 Molecular Properties
and for a general field the orientation average is αλ = 1 3 αxx, λ + αyy, λ + αzz, λ
given by
el
e2 ℏ2 fκλ
αλ = κ λΔ 2
me κλ
el 2
el
where the averaged oscillator strength fκλ = 2me 3e2 ℏ2 Δκλ κ D λ Q
contains the electronic dipole vector. A compact expression is obtained defining
a weighted excitation energy Δ by means of
2 2
el el
κ λ
κ D λ Δκλ = κ λ
κ D λ Δ
Q Q
where the numerator to the right, using the closure (or completeness) relation
κ λ κ κ = I − λ λ of the electronic basis set, is simply the dipole standard
2 2
el el
deviation value ΔD 2 = λ D λ − λ D λ . This gives
Q Q
from the sum rule for the oscillator strength [5] which gives el
κ α fκα = Nel∗ ,
2.7 Approximations and Bounds for Polarizabilities 59
Table 2.2 Charge C, electric dipole D, and mean polarizability α of selected atoms and
molecules in SI units.
References
1 Jackson, J.D. (1975). Classical Electrodynamics. New York: Wiley.
2 Landau, L.D. and Lifshitz, E. (1975). Classical Theory of Fields, 4e. Oxford,
England: Pergamon Press.
3 Hirschfelder, J.O., Curtis, C.F., and Bird, R.B. (1954). Molecular Theory of Gases
and Liquids. New York: Wiley.
4 Stone, A.J. (2013). The Theory of Intermolecular Forces, 2e. Oxford, England:
Oxford University Press.
5 Atkins, P.W. and Friedman, R.S. (1997). Molecular Quantum Mechanics.
Oxford, England: Oxford University Press.
6 Cohen-Tanoudji, C., Diu, B., and Laloe, F. (1977). Quantum Mechanics, vol. 2.
New York: Wiley-Interscience.
7 Zare, R.N. (1988). Angular Momentum. New York: Wiley.
8 McWeeny, R. (1989). Methods of Molecular Quantum Mechanics, 2e. San Diego,
CA: Academic Press.
9 Martin, R.M. (2004). Electronic Structure: Basic Theory and Practical Methods.
Cambridge, England: Cambridge University Press.
10 Kelly, K.L., Coronado, E., Zhao, L.-L., and Schatz, G.C. (2003). Optical properties
of metal nanoparticles: influence of size, shape, and dielectric environment.
J. Phys. Chem. B 107: 668–677.
11 Messiah, A. (1962). Quantum Mechanics, vol. 2. Amsterdam: North-Holland.
12 Lowdin, P.O. (1962). Studies in perturbation theory. IV. Projection operator
formalism. J. Math. Phys. 3: 969.
References 61
CONTENTS
where Ca and Ce are the electric charges of nucleus a and the electron, in the SI units.
The electrostatic energy operator must be added to the Hamiltonian opera-
tors of A and B to describe their interaction, given in general for R > Rmin by
int
H AB = H A + H B + H AB R
int −1 CI CJ
H AB R = 4πε0 I J
R −r IA +r JB
3.1 Long Range Interaction Energies from Perturbation Theory 65
A → B
R
→ →
RA RB
int −1 −1
H AB R = 4πε0 d 3 r d 3 s cA r r − s cB s
States and energies of the whole system satisfy the eigenvalue equation
H AB ΨκAB = EκAB ΨκAB , with energies and states dependent on the intermolec-
ular distance R. This can be solved with a perturbation expansion to a conven-
ient order, assuming that the interaction is small, introducing the internal
electronic states of A and B, and the zeroth order electronic states
Ψκ0 = jA kB = jA kB of the unperturbed (non-interacting) pair AB with
A B
zeroth-order energies Eκ0 = Ej + Ek . Here, it is assumed that insofar as A
and B are electronically bound states, their wavefunctions of electronic
66 3 Quantitative Treatment of Intermolecular Forces
int
The matrix element jA kB H AB jA kB can be re-expressed as a sum of pro-
ducts of factors relating separately to A and B and has the meaning of an inter-
action where a transition in A from state jA to jA is coupled to a transition in B
from state kB to kB .
For two molecules A and B enclosed by spheres of radius aA and aB , the inter-
action potential energy is small at distances larger than a Rmin and a suitable
expansion parameter is λ = (aA + aB)/Rmin. Expanding the energy and wavefunc-
tion in powers of λ the Hamiltonian equation is
0 1
H + λH n≥0
λ n Ψκn = p≥0
λ p E κp q≥0
λ q Ψκq
0 1 1
where H = H A + H B and λH = H AB R . Equating on both sides the factors
multiplying λp, a recursion procedure follows as
0
H −Eκ0 Ψκ0 = 0
0 1
H −Eκ0 Ψκ1 + H Ψκ0 = Eκ1 Ψκ0
0 1 p− 1 q
H −Eκ0 Ψκp + H Ψκp −1 = Eκp Ψκ0 + E
q=1 κ
Ψκp− q
for p ≥ 2, and with the energy to p-th order given by the recursion relation
1 p −1 q
Eκp = Ψκ0 H Ψκp− 1 − E
q=1 κ
Ψκ0 Ψκp− q
where the normalization Ψκ0 Ψκ0 = 1 has been chosen. Higher-order wave-
function terms can alternatively be constructed so that Ψκ0 Ψκp −q = 0 for p
− q ≥ 1, which imposes the intermediate normalization Ψκ0 Ψκ = 1. Addi-
tional relations can be used to show that knowledge of Ψκp is sufficient to obtain
1
the Eκ2p + 1 energy [3]. In particular, Eκ1 = Ψκ0 H Ψκ0 ,
1 1
Eκ2 = Ψκ0 H Ψκ1 , and Eκ3 = Ψκ1 H −Eκ1 Ψκ1 .
3.1 Long Range Interaction Energies from Perturbation Theory 67
1 1
E0, 0 R = 0A ,0B H 0A , 0B
with A in its ground state interacting with B also in its ground state. This is the clas-
els 1 int
sical electrostatic interaction energy E0, 0 R = λE0, 0 R = 0A , 0B H AB 0A ,0B .
To second-order
1 2
0A ,0B H jA ,kB
2
E0, 0 R = j, k 0, 0 0 0
E0, 0 − Ej, k
and the double sum over j and k, which must not be both ground states, can be
separated into two expressions: one where only j or k is the ground state, called
an induction potential energy; and a second double sum where both j and k dif-
fer from the ground state, called a dispersion potential energy. The induction
energy is therefore
2
int
0A ,0B H AB jA ,0B
ind
E0, 0 R =− j 0 A A
Ej − E0
2
int
0A ,0B H AB 0A , kB
− k 0 B B
Ek − E0
and has the meaning of B in its ground state interacting with A in a transition
state for the first term, and the reverse in the second term. For the dispersion
energy one has a double sum where A in a state transition interacts with B also
in a state transition, in accordance with
2
int
0A ,0B H AB jA , k B
dsp
E0, 0 R =− j 0 k 0 A A B B
Ej − E0 + Ek − E0
This can be made more specific by introducing details of the AB pair inter-
action. Here, the denominators have been changed to show them as always pos-
itive, with the minus sign in front of expressions indicating that these are
negative quantities corresponding to attraction between A and B when these
are in their ground states. This is not necessarily the case for the perturbation
energies of their excited states.
68 3 Quantitative Treatment of Intermolecular Forces
A A∗
energy Eκ0 = Ej + Ek and the first-order resonance interaction energy is for
R > Rmin,
EAA∗ = 2 −1
± int int
jA , kA∗ H AB jA ,kA∗ + kA ,jA∗ H AB kA , jA∗
int
± 2 jA , kA∗ H AB kA ,jA∗
so that the first two terms are again electrostatic interaction energies from the
charge distributions of A and A∗ in states j or k, but the last term is an addition
involving transitions between states j and k for each species. This involves an
energy of interaction between two electronic transition densities. The last term
contains state transfer integrals which obey different selection rules and lead to
longer range interactions, absent when there is no state coherence.
This subject is closely related to the theory of molecular excitons [5] and
intermolecular electronic energy transfer [6].
their size, or not. In the first case, we deal with long-range (LR) interactions, and
we can express the energy as an expansion in inverse powers of R and in terms of
the quantal electronic states of the isolated compounds. The short-range inter-
actions must, however, be treated differently starting from the electronic struc-
ture of the whole pair and its quantal states, allowing for electronic
rearrangement. Indicating the electrostatic, or Coulomb, interaction of the
int
two sets of charges as HAB R , its long-range form can be written as
LR int
HAB R = fd R HAB R , with fd(R) a function changing from zero at short
distances into 1.0 at long distances with a transition around a minimum dis-
tance Rmin and a transition region of width aA + aB, the sum of radii of spheres
enclosing the charge distributions of A and B. The total Hamiltonian is then
int LR
H AB = H A + H B + HAB R 1 −fd R + HAB R , and one can concentrate on
the LR part to begin with.
We consider here situations where the two charge distributions do not over-
lap, with R > Rmin, a minimum distance, large, and with rI R and sJ R. We
work in the CMN-SF frame, take the origin of coordinates at the CM of A and
−1 −1
expand the inverse distance s− r = R +sB −r A between generic field and
charge locations in a power series for small rA/R , where R = R +sB , with com-
ponents Rξ = X ,Y ,Z . This is done using that ∂ R −r A ∂ξA = Rξ R
ξA = 0
1 ∂ 1 ∂2 1
= 1− ξ
ξ A
+ ξ η
ξ, η A A
+…
R −r A ∂Rξ 2 ∂Rξ ∂Rη R
−1
−1
and an expression for ϕels R = 4πε0 I CI R −r IA like
1 A ∂ 1 A ∂2 1
ϕels R = C A
− D
ξ ξ
+ Q
ξ, η 6 ξη ∂R ∂R
+…
4πε0 ∂Rξ ξ η R
C= I
CI , Dξ = I
C I ξI
Qξη = I
CI 3 ξI ηI − δξη rI2
for compound A. These can also be given in terms of spherical multipole func-
tions of spherical coordinates, which are useful in the description of two inter-
acting few-atom systems.
70 3 Quantitative Treatment of Intermolecular Forces
1 ∂ 1 ∂2 1 ∂ 1 2 ∂2 1
= 1 + z1 + z12 2 + … = 1 − z1 + z1 2 + …
R1 ∂z1 2 ∂ z1 R1 ∂Z 2 ∂Z R
and similarly for charge 2; further using that Z/R = cos(Θ) we find
0 1 2
ϕels R = ϕels R + ϕels R + ϕels R
0 C
ϕels R =
4πε0 R
1 Dcos Θ
ϕels R =−
4πε0 R2
2 Q 2 3cos2 Θ −1 2
ϕels R =
4πε0 R3
where we have introduced the total charge C = C1 + C2 (a 20-pole), dipole D =
C1z1 + C2z2 (a 21-pole), and quadrupole Q = 2 C1 z12 + C2 z22 (a 22-pole) of the
system. This shows the electrostatic potentials created by charge, dipole, and
quadrupole distributions. Higher-order terms include octupoles, and higher
2n-poles. The electric potential of a 2n-pole is found to vary with large R as
R−(n + 1).
The electrostatic field vector = −∂ϕels ∂ R has spherical components (for
the above charges on the z-axis) given by
∂ϕels 1 ∂ϕels
ℇR = − , ℇΘ = − , ℇΦ = 0
∂R R ∂Θ
energy of a system B of charges in that potential [7, 8]. We consider the system B
with two charges 3 and 4 located at positions R3 = R +r 3 and R4 = R +r 4 in the
potential of charges 1 and 2 (system A). The interaction energy function is now
int A A
HAB = C3 ϕels R3 + C4 ϕels R4
This can be further expanded now for small rj/R, j = 3, 4, which brings in the
multipoles of B, and we introduce CB = C3 + C4 and DB = C3r 3 + C4r 4 , and a sim-
ilar notation for the A multipoles. Using a vector notation and noticing that the
interaction of a 2m-pole of A with a 2n-pole of B gives a term with an R−(m + n + 1)
dependence, we find for R > Rmin
−1
R−
int m+n+1
HAB R = 4πε0 F AB
m, n m, n
AB
F0, 0 = CA CB
AB AB
F0, 1 = − CADB n , F1, 0 = −CBDA n
AB
F1, 1 =DA DB − 3 DA n DB n
and tξηζ = − 15nξnηnζ + 3(nξδηζ + nηδξζ + nζδξη). In general, the tensor with n
indices is
∂χ R − 1
n
tξη…χ = R n + 1 ∂ ξ ∂ η
for higher derivatives, with the notation ∂ ξ = ∂/∂Rξ [9]. These tensors depend on
the orientation angles of the relative position vector R =RB −RA and their
sign changes by (−1)n if this direction is reversed. The electrostatic potential
is then
A 1 CA A tξ 1 A tξη
ϕels R = + ξ
Dξ + ξ, η 6
Qξη +…
4πε0 R R2 R3
which is in turn differentiated n times and multiplied times the 2n-th multipole
of B to obtain the interaction energy operator as [9, 10]
A B
−1 C C tξ B A tξ
H AB LR
= 4πε0 + C A
D − Dξ 2 C B
R ξ R2 ξ R
1 A tξη B A tξη B
+ C Q + Qξη 3 C
6 ξ, η R3 ξη R
A tξη B
− Dξ D
ξ, η R3 η
1 A tξηζ B A tξηζ B
− Dξ Q + Qξη 4 Dζ + fd R
6 ξ, η, ζ R4 ηζ R
LR −1 Tμ,mν, n R
H AB = 4πε0 Mm A
nν μ
Mn B
mμ Rm + n + 1 ν
where μ designates the components of the 2m multipole, with three values for
the dipole, six for the quadrupole, and so on.
In a reference frame with the Z-axis along the intermolecular relative position,
the perturbation energies depend on R and are given as before in terms of the
1 1
contains a sum of R−(m + n + 1) terms, and
0
integrals Ψκ0 H Ψλ . Here, H
the energy term Eκp R is itself an expansion in powers of R−1, of the form
Eκp R = n≥n p
Cnp R − n
where n(p) is the lowest power of R−1 which appears to order p in the energy
expansion, and increases as p increases.
3.2 Long Range Interaction Energies from Permanent and Induced Multipoles 73
For the ground state of the pair, the energy varies as the relative distance
decreases from infinity, and the intermolecular potential energy evolving from
the ground state g = (0A, 0B) of the pair is
Vg AB R = EgAB R −EgAB ∞
AB A B
where the asymptotic term is simply Eg ∞ = E0 + E0 , the sum of energies
of the isolated molecules in their ground state. This intermolecular potential
energy is therefore a double perturbation expansion of form
Vg AB R = p≥1
λp n≥n p
Cnp R −n
and the lower limits n(p) to the summations must be found by inspection of the
asymptotic expansion series in powers of R−1. The coefficient of R−n must be
obtained from all orders in λ relevant to a desired accuracy.
An extension of the perturbation treatment is needed when the states of A or
B are degenerate [3, 4, 11], as is frequently the case for excited electronic states.
The expansions in inverse powers of R can be expected to converge only
asymptotically, particularly when applied to atoms or compounds with open
electronic shells, as has been noticed in a survey of several examples [12].
AB
The terms An/Rn making the series expansion of Vg R are frequently such
that the convergence criterion based on the quotient of two adjacent terms gives
An + 1/(AnR) ∞ for n ∞ indicating series divergence. However, the sum
of the first N terms in the expansion usually gives a good approximation to the
AB
exact potential energy so that Vg R − N n = 0 An R
n
0 as N ∞ for R lar-
ger than a critical value dependent on N, indicating an asymptotic or semicon-
vergent series. This justifies using a sum of inverse power terms to calculate the
potential energy for large distances.
B. In the ground state (jA, kB) = (0A, 0B). Therefore, each term is of the form
−1
Cm + n + 1 R −
els m+n+1
Vmels
, n R = 4πε0
els
where Cm + n + 1 is a constant obtained from the permanent multipoles in the
ground states.
To second order, the potential energy operator appears in two factors,
between the ground and excited states of the (A, B) pair, and again back from
excited states to the ground state. Depending on the nature of the excitation, we
74 3 Quantitative Treatment of Intermolecular Forces
tions into and out of states of B∗ usually requires that n = n , so that the induc-
tion potential energy is usually
ind −2 ind − m + p + 2n + 2
Vm, p, n R = − 4πε0 Cm + p + 2n +2 R
with the leading term coming from (induced dipole)–(induced dipole) interac-
tions going as R−6. The three types of interaction processes, electrostatic, induc-
tion, and dispersion, are diagrammed in Figure 3.2.
Going to higher powers in the perturbation expansion in λ of the intermolec-
ular potential energy gives additional terms combining permanent and induced
multipole interactions. The total potential energy can be written as V(R) =
V(els)(R) + V(ind)(R) + V(dsp)(R) + V(pol)(R), an asymptotic expansion in powers
of R−1 with a varying radius of convergence, where a polarization term from
higher-order perturbations in λ combines induction and dispersion
interactions.
When the two species are the same, and one is electronically excited, there is
in addition a resonance potential energy V(res)(R) resulting from the quantal
3.2 Long Range Interaction Energies from Permanent and Induced Multipoles 75
Induction
2m
2n Excitation
A 1
R m+n+1 B
2mʹ
2nʹ De-excitation
A 1
R mʹ+nʹ+1 B
Dispersion
2m 2n Excitation
1
A R m+n+1 B
−1
A −1
The electrostatic potential ϕels R = 4πε0 I CI R−r IA can be expanded
in the spherical harmonics for R and r IA as
A −1 1 rIA l 4π ∗
ϕels R = 4πε0 C
I I
Y
m lm
ϑAI , φAI Ylm ϑR ,φR
lR R 2l + 1
which gives in terms of multipoles
1
A −1 4π 2 1 ∗
ϕels R = 4πε0 Qlm r IA Ylm ϑR ,φR
l, m 2l + 1 Rl
d3 k
cA r = C δ r −r IA =
I I
exp − ik r QA k
2π 3
3.2 Long Range Interaction Energies from Permanent and Induced Multipoles 77
int −1 d3 k
H AB R = 4πε0 exp ik R QA − k QB k
2π 2 k 2
to be incorporated into the first- and second-order perturbation expansions
given above for the interaction energies. Here, we concentrate on the dispersion
component to second order, which contains denominators like
A A B B
Ej −E0 + Ek − E0 = a + b, and use the Casimir–Polder integral (for positive
a, b)
∞
2 du a b 1
=
π 0 a2 + u2 b2 + u2 a+b
to write for the dispersion energy in Section 3.1.1, with ω = iυ
dsp −2 d3 k d3 k
E0, 0 R = − 4πε0 exp i k − k R
2π 2 k 2 2π 2 k 2
∞
ℏ
dυ χ A k,k ; iυ χ B − k, −k ; iυ
2π 0
2 ωj
χ A k, k ; iυ = 0 QA − k j j QA k 0
ℏ jω 2
j + υ2
A A
with ℏωj = Ej − E0 , and similarly for B. This gives the dispersion energy as a
sum over grid points in k and k spaces and avoids an expansion in inverse
powers of R. The factors χ S are dynamical susceptibilities evaluated at imagi-
nary-valued frequencies for each species S.
Further, using the plane wave decomposition into spherical waves,
∞ l
exp i K R = l=0 m = −l
i l 4π Ylm ϑK ,φK ∗ Ylm ϑR , φR jl KR
where the spherical Bessel function jl(x) satisfies asymptotic conditions jl(x) ≈
Alx−l − 1 sin(x) for x ∞ and jl(x) ≈ Blxl for x 0, it is possible to provide a
detailed analysis of the two limits for large and small R [17]. It is found that
for large R one recovers the inverse R multipole expansion previously generated,
and that now the present treatment gives finite values also for small R, so that
the dispersion energy for all R can be written as
78 3 Quantitative Treatment of Intermolecular Forces
−2
R −2 m
dsp dsp dsp
Vm , n R = − 4πε0 f2 m +n +1 R C2 m + n +1
+ n +1
with f(R) ≈ 1 for R ∞ and f(R) ≈ 0 for R 0 [17]. A detailed analysis here
shows that for R 0 one finds Vm, n R ≈ −AR2 m −n + BR2 m −n + 2 . Therefore,
dsp
where the second line gives the static dipolar polarizability of B, and the excited
states are kB = nB1P(ML = 0). The dispersion potential energy follows from
AB A B A B
F 11 = D + D −B + D −A D + −2 D0 D0 , and its matrix elements between
0A0B and jAkB . Noticing that selection rules restrict the three terms to be
nonzero only one at a time, the dispersion energy is given by three σ = ± , 0 com-
ponents like
2
1 DσA 0j
B
D −σ 0k C6
dsp
Vσ dsp = − =−
4πε0 2 R6 j 0 k 0 A
ϵj0 + ϵk0
B 4πε0 2 R6
A A A
where DσA 0j
= 0A DσA jA and the ϵj0 = Ej − E0 are excitation energies,
dsp dsp
as V dsp = V + + V −dsp + 4 V0 . Here, the excited states for A mediated by
A A
D0 and D ± are instead jA = nA 1P(ML = 0, ± 1) and similarly for B.
For a pair of neutral atoms such as He+Ne, the induction energy disappears
and only the dispersion terms remain. The (induced dipole)–(induced quadru-
pole) dispersion energy goes as R−8.
B
with α B = α + 2α⊥ 3, an average static polarizability for the HF diatomic,
while the dispersion potential becomes
dsp dsp
dsp
C6 + 2 C6⊥ 1 α − α⊥ B
VA −D = − 1+ P2 cosθ
4πε0 2 6
R 3 α
in terms of parallel and perpendicular dispersion coefficients involving corre-
sponding transition dipoles of HF. Additional terms must appear if interactions
involving permanent and transient quadrupoles were included, containing
higher powers of R−1 and of cos(θ). The interactions of a neutral atom with a
heteropolar diatomic are obtained letting C(A) = 0 above, while the interaction
of a charged atom with a homopolar molecule follows from D(B) = 0.
The charge C, dipole D, and mean polarizability α of Li+, Ne, H2, HF, H2O, and
other selected species have been given in SI units in Table 2.2 of Chapter 2, and
can be used to calculate electrostatic, induction, and dispersion energies, how-
ever keeping in mind that multipole and polarizability values depend on the
chosen origin of coordinates for the charge and polarization distribution, which
refer here to the molecular center of mass.
∞
1 1 A iu B iu
Vσ dsp = − du ασ, 0 α
4πε0 2 R6 2π 0 ℏ −σ, 0 ℏ
82 3 Quantitative Treatment of Intermolecular Forces
where
2
A
iu ϵj0 DσA 0j
A
ασ , 0 =2
ℏ j 0 A
2
ϵj0 + u2
is a positive valued function decreasing and vanishing as u goes from zero to
u ∞. It follows that the dispersion potential energy can be constructed from
the pair polarizabilities evaluated at imaginary-value frequencies ω = iu/ℏ.
The dispersion energy written for Cartesian or spherical dipoles can also be
re-expressed as shown, and the procedure can also be applied to dispersion
interactions involving quadrupoles or higher multipoles of A or B, introducing
the corresponding multipolar polarizabilities.
An alternative expression for the orientation averaged polarizability is
A, el
A iu c2e ℏ2 fj0 A
α0 = = β0 u
ℏ me j 0 A
2
ϵj0 + u2
A B
AB 3 α0 0 α0 0
C6 = ∗ 1 2 ∗ 1 2
2 A A B B
α0 0 Nel + α0 0 Nel
which can be calculated from static polarizabilities and the known number of
valence electrons for A and B.
A slightly modified version of this procedure provides useful combination
AA
rules [21, 22]. The parameter u A appears in C 6 and it can be extracted as
AA A 2
u A = 4 3 C6 α0 0 which involves two measurable quantities. Its
values are given in [22] for atoms and small molecules and are (in au’s) for
He, Ar, Na, H2, and CH4 equal to 1.02, 0.706, 0.0774, 0.610, and 0.649, respec-
tively, with larger values of 1.71 and 1.401 for He+ and Na+, all relating to elec-
A
tronic excitation energies. Using the form of u A in β0 u , and performing the
integration over u for the A–B pair gives the combination rule
AA BB
AB 2 C6 C6
C6 = AA BB
α B α A C6 + α A α B C6
which allows calculation of the dispersion coefficient for two different interact-
ing compounds A and B, given the coefficients for the two identical pairs A–A
and B–B, and has been found to be accurate within a few percent for many com-
binations of atoms [22].
0 0 0 0
Replacing in the resolvent Ek − E0 with the smaller E1 −E0 in all the quo-
tients, so they all become larger, and using the completeness relation
k ≥ 0 k k = I, the identity operator, and the complementary projector
P = I − 0 0 it follows that
P
0 ≥ R0 ≥ −
0 0
E1 − E0
N N k k N ∞ k k
R0 = − k =1 0 0
≥ R0 ≥ R0 − 0 0
Ek − E0 k =N +1
EN + 1 − E0
1 N 1 2
0 H R0 H 0 ≥ E0
N 0 0
≥ 0 H 1
R0 H 1
0 − 0 H 1
I− N
H 1
0 EN + 1 −E0
AB
N
with = N k = 1 k k a projection operator on the known subspace of N
states excluding the ground state k = 0. A calculation can now proceed involving
integrals of states only within this subspace. Induction and dispersion energies
can be obtained as done above, by letting the states k contain only one ground
state of the A–B pair, to generate the induction interaction of one permanent
multipole and a transient one, or by excluding the two ground states, to generate
the dispersion interaction of two transient multipoles.
Calculations can be done quite generally introducing a new basis set suitable for
N N
expansion of the R0 and operators [23, 24]. Using a notation with the 1 × N
row matrix of chosen states f = [ f1 , … , fN ], and constructing the projector
N −1
= f f f f where f is a N × 1 column matrix, a positive-valued
1 2 N 1 2
operator A is found to satisfy the inequality g A g ≥ g A A g .
1 1 2
This can be used with g = H 0 , A = − R0 , and − R0 f = h , all
orthogonal to 0 , to obtain the upper bound
1 0 −1 1
2 0
E0 ≤ 0 H h h E0 − H h h H 0
−1 1
0
0 0 0
Using instead A = R0 −P E0 − E1 and A2 f = H − E1 k , with k
in the orthogonal complement space of P, one obtains a lower bound from [23]
1 1
0 H PH 0
2
g A g = E0 − 0 0
E0 − E1
1 0 0 0 0
≥ 0 H H − E1 k k H −E0
0 0 −1 0 0 1
× H − E1 k k H − E1 H 0
The calculation of bounds to the induction energy in the pair A–B in its
ground state can be done selecting the functions in h and k to generate
the interaction of a permanent dipole with a transition dipole, with 0 = 0A,
0B , hj = 0A, hBj or hk = hAk, 0B , the hAk orthogonal to 0A , and similarly
for B and for the kj . The dispersion energy calculations must involve states hj =
hAj, hBj where both components are orthogonal to the ground states of A and B.
86 3 Quantitative Treatment of Intermolecular Forces
1 0
Φ XA , XB = Φ XA ,XB f XA ,XB H AB XA ,XB where f(XA, XB) is a sum of
factors with powers of electronic coordinates in A and B, it has been possible
DD
to obtain this way highly accurate values of dipole–dipole C6 and dipole–
DQ
quadrupole C8 dispersion energies for interacting atoms with one and two
electrons, including H + H, H + He, and He + He [13]. Their collected values
are shown in Table 3.1, in au’s.
ε
unocc
occ
A B A B A B A B
Electrostatic Polarization Electron Electronic
interactions (induction and exchange excitation
dispersion) and transfer
minimum as a function of the parameters, for the ground state or for each state
of lowest energy with a given symmetry given by fixed Q. The optimized state
Φ(AB) can be taken as the reference or zeroth order state to be further corrected
by combining it with a basis set of orthonormal states to be used in a perturba-
tion expansion.
A qualitative discussion can be based on electronic orbitals and their charge
distributions. The several phenomena, appearing at short and long distances for
closed–closed, closed–open, and open–open shell interactions are schemati-
cally illustrated in Figure 3.3.
For each species A or B, electronic energy levels ε of orbitals are found to be
occupied as shown by the dashed bands in the figure, or to be unoccupied. Elec-
trostatic interactions are always present and obtained from electronic charge
densities, polarization (induction and dispersion) occurs at both long and short
distances, electron exchange appears as the distance is reduced and orbitals start
to overlap, and electronic excitation and charge transfer occur more likely at
short distances due to electron correlations and the Pauli exclusion of electrons
from occupied orbitals.
A great deal about interactions at intermediate and short distances can be
understood and estimated from the relationship between energies and elec-
tronic densities established by the Hohenberg–Kohn theorem [28]. Given a
set of nuclei a (or ion cores) at fixed positions Ra , electrons are distributed
among them and show unique spikes at their locations. This means that the
external potential energy Vext r provided by the nuclei uniquely creates an
electron number distribution ρ r per unit volume. The theorem states that
the reverse is also verified for the nondegenerate electronic ground state and
lowest energy electronic state of each symmetry, and that there is a unique func-
tional ρ r which reaches a minimum when a trial density function is varied
to reach the correct one. The density derives from a many-electron
3.5 Electron Exchange and Penetration Effects at Reduced Distances 89
ρ r ; Q = VNN Q + d 3 r ρ r vNe r; Q + F ρ r ; Q
m n
δ r 1 −r m δ r 2 −r n
= m, n
δ r 1 −r m δ r 2 −r n −δ r 1 −r 2 m
δ r 1 −r m
where the double sum in the last line includes m = n. This gives
m n
δ r 1 −r m δ r 2 −r n = ρ 1 r 1 ρ 1 r 2 − δ r 1 −r 2 ρ 1 r 1
information about the inhomogeneous electron gas [29, 30] and by parametri-
zations designed to reproduce the energetics of sets of molecules [31]. The so-
called DFT approach has been widely applied to molecular and solid-state
phenomena. Separating nuclear position variables as Q = R,QA ,QB and doing
calculations for varying R, the most recent functionals give quite acceptable
results for ground electronic state equilibrium distances Re and dissociation
energies De = E0(∞) − E0(Re) in many applications involving intermediate and
short distances.
However, in applications of intermolecular forces to the A–B pair, it is neces-
sary that as R ∞, one must find ρ AB r; R ,QA ,QB ≈ρ A r; QA + ρ B r; QB ,
and also that the functional must satisfy the limiting form
have included diatomic and polyatomic species for which the dispersion energy
terms have been added at large distances [35–37] Calculation of interaction
energies at very short ranges also requires adding distortion densities for very
short distances [38].
For a pair of species, such as Ne + Na, where one has closed shells and the
other has open shells, it is necessary to allow for a partial redistribution of
the electronic density of the open shells and to work with spin densities to allow
for the unpaired electrons. If these are in species B, a possible choice of a
∗
variational density is ρ AB r = ρ A r + ρ B r where the second term can bet-
2
ter be treated within the KS formalism by choosing ρ B, η r =
B
j cjη φj r ,a
sum including orbitals at species B with electron spin η, forming hybrid atomic
orbitals, or more general distortion orbitals. If B is a molecule, coefficients of
molecular orbitals can be optimized by minimizing its total energy functional.
For two species with open shells, such as Na + Na, the KS choice is suitable for
intermediate and short distances, and the functional must be minimized with
AB
respect to variations of KS spin-orbitals φjη with occupation numbers νjη
2
AB
that extend over the two species, as ρ AB r = jη νjη φjη r with
AB
φjη r = μ cμ, jη χ μ r a linear combination of atomic orbitals χ μ centered at
the nuclei. This is best done in the case of chemical bonding by introducing spin
densities to account for open shells of separated species [39].
The procedure for pairs of atoms was implemented in early work using a den-
A B
sity ρ AB r = ρHF r + ρHF r with the atomic densities from Hartree–Fock
calculations, and with a simple Thomas–Fermi–Dirac functional for
F ρ r ;R [40]. Results for the short-range interaction energy of many (Z =
2–36) atom pairs could be well fitted at short distances with the Born–Mayer
function V(R) = A exp(−bR), and it was found that reasonable accuracy, of the
order of 1%, could be obtained for different atoms A and B from the combina-
tion rule V(AB) = (V(AA)V(BB))1/2. This approach was substantially extended with
the GK procedure to include both short and intermediate distances and to
obtain the location and depth of van der Waals potential minima. Using
Hartree–Fock atomic densities, and kinetic energy, exchange, and correlation
energy functionals for the inhomogeneous electron gas provided by
Thomas–Fermi, Dirac, and a local energy density expansion, respectively
[33], remarkably good results could be obtained by comparison with ab initio
calculations and experiment. This is shown in Figure 3.4 for Ar + Ar, and in
Table 3.2 where GK values for Rm and ϵm are compared to experimental ones.
For Ne–Ne it gives Rm = 2.99 Å = 299 pm and ϵm = 56 × 10−16erg = 56 × 10−23 J
for the minimum location and well energy, compared with experimental values
of Rm = 303 pm and ϵm = 63 × 10−23 J.
92 3 Quantitative Treatment of Intermolecular Forces
10
x
x
x
x
1
x
10–1
eV
10–2
V(R)
10–3
0
10–16 erg
–100
–200
Figure 3.4 The Ar–Ar potential energy versus interatomic distance: full line GK calculations
[33]; dashed line from gas experiments, and dotted line from beam scattering experiments.
Source: from Gordon and Kim [33].
Table 3.2 Calculated and measured values of potential energy parameters (from [33]) in
SI units.
Similar accuracy is obtained for other pairs of atoms. The treatment was later
improved with a more accurate functional correcting for the electronic self-
energy and was applied to molecular species with similar encouraging results,
also for interaction potential anisotropies [35].
At present, there is no satisfactory density functional treatment of intermo-
lecular forces to cover all ranges with a consistent formulation, and it seems
best to construct the exchange-correlation potential energy in the form
c
μ = a, b μ
Hμν − ϵ el Sμν = 0
el
with ν = a, b, Sμν = χ μ χ ν the overlap integral, and Hμν = χ μ h χ ν an inte-
gral for the electronic Hamiltonian matrix element. The AOs satisfy certain
combination rules: (i) They combine only when they have the same spatial sym-
metry, since otherwise Sμν = Hμν = 0; (ii) they combine more strongly (with com-
parable coefficient values) if Haa Hbb as seen from (ca/cb)2 = (Hbb − ϵ(el)Sbb)/
(Haa − ϵ(el)Saa); (iii) they combine more strongly when there is a larger overlap
between atomic charge distributions at a and b, which increases the magnitudes
of coupling terms Hab and Sab.
Solving for the coefficients requires that the determinant of the matrix they
multiply must be zero, or (α − ϵ)(β − ϵ) − (γ − ϵS) = 0 in the notation α = Haa,
β = Hbb, γ = Hab, ϵ = ϵ(el), S = Sab, and with Saa = Sbb = 1. This gives energy roots [47]
1 2
ϵ ± = − A ± A2 − B ,
− α + β + 2γS αβ −γ 2
A= , B =
1 − S2 1 −S 2
el el
and energy eigenvalues ϵ1 = ϵ − ≤ ϵ2 = ϵ + . The corresponding linear combina-
tion coefficients follow from
el el
cb γ − ϵj S α− ϵj
=− el
=− el
ca j β − Ej γ − Ej S
3.5 Electron Exchange and Penetration Effects at Reduced Distances 95
and the normalization condition c2a + c2b + 2ca cb S = 1 for each j. Furthermore, the
MOs can be labeled by the axial rotation quantum number mz = 0, ± 1, ± 2, …, as
being of symmetry λ = mz and type σ, π, δ, …. The coefficients change with R
and in the limit R = 0 the MO becomes a united atom AO of nuclear charge
number Za + Zb and atomic quantum numbers (nlm), with m = mz. Further-
more, the energy functions of R for MOs of the same symmetry must not cross
as R changes, except by numerical coincidence, because any electronic coupling
el
term U in the Hamiltonian must then lead to coupling integrals
el
φj U φj different from zero and the energy roots of a new eigenvalue
equation must be different, indicating their repulsion. All this leads to a MO
energy correlation diagram as R changes from separated atoms at large values
to the united atom, as shown in Figure 3.5.
3dδ
3d
3dπ σ3sA 3sA
3dσ
2s 2sσ
σ1sB 1sB
σ1sA 1sA
1s 1sσ
United atom Separated atoms (ZA > ZB)
(R= 0) (R → ∞ )
R→
Figure 3.5 Energy-level correlation diagram (not to scale) for the one-electron heteronuclear
diatomic molecules. Source: Figure 4.1, page 96 of Bransden and McDowell [48]. Reproduced
with permission of Oxford University Press.
96 3 Quantitative Treatment of Intermolecular Forces
In H + H+, or for two identical ion cores, the molecular reflection symmetry
across a plane perpendicular to the internuclear axis leads to ca = ± cb for real
valued MOs, and the normalized LCAOs are φg , u r = χ a r ± χ b r
2 1 ± Sab 1 2 , with the gerade and ungerade forms for the + and − signs.
The nlm hydrogenic orbital for an electron at position r at a nucleus a with
nuclear charge Za is χ a, nlm r = r anlm = Rnl ra Ylm ϑa ,φa . For the 1s orbital
−1
χ a, 1s0 r = r a1s = Za a0 3 2 2 exp −2Za ra a0 2π 1 2 with a0 the Bohr
radius, of energy E1s = − Za2 e2 2a0 , when using Gaussian units [49]. Here,
Sab = S decreases exponentially for large R as exp(−R/a0). An energy correlation
diagram from the separated atoms to the united atom for homonuclear dia-
tomics is shown in Figure 3.6.
The total energy is the sum of nuclear repulsion plus electronic energy,
el
Eg R = c2e 4πε0 R + Eg , with
σu3s
3s
σg3s
3dδg
3d 3dπg
σg σu2p
3pπu
3p πg2p 2p
3pσu
3sσg πu2p
3s
σg2s
σu2s
2s
2pπu
2p
2pσu
σu2s
2s 2sσg
σu1s
1s
σg1s
1s 1sσg
Figure 3.6 Energy-level correlation diagram (not to scale) for the one-electron homonuclear
diatomic molecules. Source: Figure 4.3, page 98 of Bransden and McDowell [48]. Reproduced
with permission of Oxford University Press.
3.5 Electron Exchange and Penetration Effects at Reduced Distances 97
j+k
Egel R = ϵ1s −
1+S
2
Ce2 χ r Ce2 χ r χb r
jR = d3 r a , k R = d3 r a
4πε0 r −Rb 4πε0 r −Ra
where j is called the Coulomb integral and describes the attraction between the
electron in orbital a and the nucleus at b, a term expected from classical elec-
trostatic. The integral k, however, is purely quantal and accounts for the inter-
action of the overlap of charges centered at a and b with the nucleus at a, a
quantal penetration effect. For large distances, one has j R ≈Ce2 4πε0 R plus
a term exponentially decreasing as exp(−2R/a0), while k(R) ≈ 0 exponentially
as exp(−R/a0), so that the results from pure electrostatic are recovered at
large R.
The choice of atomic orbitals with a nuclear charge number Za = 1.0 gives a
potential energy V0(R) = E0(R) − E1s repulsive at short distances and attractive at
large ones with a minimum at Re = 2.50 au and a dissociation energy De = E0(∞)
− E0(Re) = 0.065 au, to be compared with the accurate values of De = 0.1026 au =
2.79 eV at Re = 2.00 au. This is qualitatively correct and can be improved mini-
mizing the energy functional φ with respect to parameters in a trial function.
A flexible trial function must describe the polarization of atomic orbitals, and
the effective nuclear charge at short distance keeping in mind that as the inter-
nuclear distance decreases the charge changes from Za = 1.0 au at large dis-
tances to the united atom limit at R = 0 of value Za = 2.0 au. This can be
done allowing for variable charge numbers Z in exponents to get the energy
optimized. Polarization of H by H+ can be treated adding to the 1s0 orbital a
portion of the 2pz orbital, pointing from one nucleus to the other, to form a
hybrid orbital χ a r = χ a, 1s0 r + λχ a, 2pz r , with
3 2 −1
Za Za ra Za ra
χ a, 2pz r = exp − 2π 1 2
cos ϑa
2a0 a0 2a0
and λ treated as a variational parameter which depends on R. Varying para-
meters for each R one finds optimal values Z1s(R), Z2p(R), and λ(R) [50] and a
minimum at Re = 2.01 au with De = 0.1004 au = 2.73 eV when Z1s = 1.246,
Z2p = 2.965, and λ = 0.138, a much better result, showing the importance of
introducing effective nuclear (or ion core) charges and orbital hybridization
as functions of a decreasing distance.
A connection can be made with the interaction energy at large distances.
Keeping the variation parameter λ but letting Z1s = Z2p = 1.0 at large distances,
the minimum of λ provides the analytical form of the parameter and the
energy, which is found to agree with the long-range charge-induced dipole
interaction energy going as R−4 with a transition dipole and a polarizability
obtained from the integral 1s0 z 2pz . The long-range dependence on R−1,
98 3 Quantitative Treatment of Intermolecular Forces
atomic Hamiltonian and its 1s AO, and H ab 1,2 for the interatomic Coulomb
interactions. The corresponding electronic energy is given by the variational
function for this state, as
Φ CI x1 ,x2 = AG 1σ g α 1 1σ g β 2 + AU 1σ u α 1 1σ u β 2
with coefficients to be varied, leads to exactly the same states and results
obtained from the covalent-ionic treatment [47]. Induction and dispersion
interactions can be described by introducing AO hybrids, or alternatively by
adding configurations constructed with an atomic basis set containing polari-
zation atomic orbitals.
The two-electron heteronuclear diatomic He + H+ can be treated using MOs
constructed with He 1s and H 1s AOs, with the simplest wavefunction of form
ΦRHF(1, 2) = 1σα(1)1σβ(2)|. Long-range induction and dispersion interactions
can again be incorporated by means of basis sets of AOs containing at least the
2pz type in hybrid AOs, or by means of CIs. An accurate treatment here involves
3.5 Electron Exchange and Penetration Effects at Reduced Distances 101
S = 0 states [64]. It was pointed out that the magnitude of a three-body disper-
sion interaction as obtained from third-order perturbation theory is of magni-
−1
tude Iat α3at R3AB R3BC R3CA where Iat and αat are atomic ionization and dipolar
polarization magnitudes, and RLM is the distance between atoms L and M. The
angular dependence of the energy on the interior angles of the triangle formed
by the atoms is 3 cos(γ A) cos(γ B) cos(γ C) + 1, with γ L the angle at the location of
atom L. This factor changes sign depending on the shape of the triangle and
indicates that the three-atom dispersion energy can be either positive or nega-
tive. Perturbation theory shows that the dispersion energy is pairwise additive to
second order, and that a three-atom interaction term appears first in the energy
function to third order. This can be expressed in terms of dipolar polarizabilities
for each atom. However, when species contain permanent multipoles, induction
energies appear, and it is found that pairwise additivity is lost even to second
order [12].
More generally A, B, and C can be molecules, and their pair interactions may
contain not only monopole and dipole operators but also higher order multi-
poles. It is convenient to introduce pair interactions in terms of charge distribu-
tions, and to give a more general treatment using the resolvent operator. The
simplest case involves three species sufficiently separated so that their proper-
ties are not affected by electron exchange. This allows allocation of charges I of
electrons and nuclei to each separate species L = A, B, C.
The Hamiltonian is now, with species labeled by indices L and M,
−1
int −1
H LM RLM = 4πε0 d 3 r d3 s cL r L RLM +r L −sM cM s M
where the interaction is a sum of pairs, and the charge density operators are
functions of positions with respect to the center of mass of each species L or
M. The equation for three-body states can be solved expressing the three-body
resolvent operator in terms of two-body operators in a way similar to what has
been done for three-body transition operators [65–67]. In a compact notation,
arrangement a = 0 is A + B + C with all species noninteracting, and arrange-
ments a = 1, 2, and 3 refer to cases where A, B, or C are, respectively, noninter-
acting while the other two atoms interact. The zeroth-order Hamiltonian is
0
H = H A + H B + H C , and interaction Hamiltonians in each arrangement are
0 int 1
V = 0 and V a , so that the perturbation Hamiltonian is H ABC = λH = a V a .
Perturbation theory expressions give wavefunctions and energies to second and
third order. For each of the three species starting in its ground state 0L , the
0 0 A B C
unperturbed state is ΨG = 0A 0B 0C with energy EG = E0 + E0 + E0 .
3.7 Interactions in Three-Body and Many-Body Systems 105
The first-order energy and state satisfy, for the given G state, λE 1
=
int
Ψ 0
aV a Ψ 0
= a Ea and
0
H −E 0
λΨ 1 = a
Eaint −V a Ψ 0
1
and second- and third-order energies are obtained from E 2
= Ψ0 H Ψ1 ,
1
and E 3
= Ψ1 H −E 1
Ψ 1 , and involve only Ψ(1). This satisfies an
inhomogeneous differential equation that can be solved by inspection in terms
of two-body solutions as in
Ψ1 = a
Ψa1
0
H −E 0
λΨa1 = Eaint − V a Ψ 0
where each arrangement term in the wavefunction involves excited states of two
1
species and the ground state of the third one so that Ψ1 is a combination of
states 0AjBjC and similarly for the other two arrangements. It follows from this
and from the orthonormality of the states jL , L = A, B, C, that one can separate
in each arrangement the induction interaction integrals, where one of the spe-
cies remains in its state and the other is excited, from the dispersion interaction
integrals where both are excited. This gives after analysis of all integrals of the
1
type Ψ 0 V a Ψb , containing products with arrangement interactions V a
and V b [1]
2 2 2
E ABC = Eind ABC + Edsp ABC ,
2 2 2 2
Edsp ABC = Edsp AB + Edsp BC + Edsp CA
and indicates that the dispersion interaction of the three species is additive to
the second order, while the induction interaction is intrinsically a three-body
function.
The analysis can be extended to third order where one has that
E 3
= Ψ1 a
V a − Eaint Ψ1
1
contains 27 integrals of the type Ψb V a Ψc1 involving products with the
three arrangement interactions V 1 ,V 2 , and V 3 . When the arrangement
interactions are expanded in the basis set jAjBjC , many of the integrals are zero
and the remaining ones can be classified as giving three-body induction or dis-
persion energies. In particular, the dispersion energy arises from the interaction
of density fluctuations of the three species. They undergo a sequence of
106 3 Quantitative Treatment of Intermolecular Forces
transitions from the ground state for each pair first to excited states which trans-
fer energy in a second interaction and next return to the ground states in a third
interaction, all involving transition charge densities jL cL r L jL . A special
case detailed in the literature considers the dipole–dipole interaction energy
operator
dd −1 L tξη
H LM RLM = − 4πε0 Dξ DM
ξ, η R3LM η
which when entered into the third-order dispersion energy leads to [1]
3 dsp 1 + 3 cos γ A cos γ B cos γ C
Edsp ABC = C9 ABC
4πε0 3 R3AB R3BC R3CA
dsp
with a factor C9 ABC which can be expressed in terms of the dynamical
polarizabilities of the three species, as [13, 68]
∞
dsp 3ℏ
C9 ABC = du α A iu α B iu α C iu
π 0
3
Some calculated values of Edsp for H and He systems, and discussions
of the role of nonadditivity in condensed matter are given in the literature [2,
13, 69].
References
1 Margenau, H. and Kestner, N.R. (1971). Intermolecular Forces, 2e. Oxford,
England: Pergamon Press.
2 Maitland, G.C., Rigby, M., Brain Smith, E., and Wakeham, W.A. (1981).
Intermolecular Forces. Oxford, England: Oxford University Press.
3 Dalgarno, A. (1961). Stationary perturbation theory. In: Quantum Theory, vol. 1
(ed. D. Bates), 171. New York: Academic Press.
4 Hirschfelder, J.O., Byers Brown, W., and Epstein, S.T. (1964). Recent
developments in perturbation theory. Adv. Quantum Chem. 1: 255.
5 Davydov, A.S. (1971). Theory of Molecular Excitons. New York: Plenum.
6 Foerster, T. (1948). Intermolecular energy migration and fluorescence. Ann.
Phys. 2: 55.
7 Jackson, J.D. (1975). Classical Electrodynamics. New York: Wiley.
8 Landau, L.D. and Lifshitz, E. (1975). Classical Theory of Fields, 4e. Oxford,
England: Pergamon Press.
9 Stone, A.J. (1996). The Theory of Intermolecular Forces. Oxford, England: Oxford
University Press.
10 Hirschfelder, J.O., Curtis, C.F., and Bird, R.B. (1954). Molecular Theory of Gases
and Liquids. New York: Wiley.
11 Cohen-Tanoudji, C., Diu, B., and Laloe, F. (1977). Quantum Mechanics, vol. 2.
New York: Wiley-Interscience.
12 Kaplan, I.G. (2006). Intermolecular Interactions, 81. New York: Wiley.
13 Dalgarno, A. and Davison, W.D. (1966). The calculation of van der Waals
interactions. Adv. At. Mol. Phys. 2: 1.
14 Zare, R.N. (1988). Angular Momentum. New York: Wiley.
15 Edmonds, A.R. (1960). Angular Momentum in Quantum Mechanics. Princeton
NJ: Princeton University Press.
16 Linder, B. and Rabenold, D.A. (1972). Unified treatment of van der Waals forces
between two molecules of arbitrary sizes and electron delocalization. Adv.
Quantum Chem. 6: 203.
17 Koide, A. (1976). A new expansion for dispersion forces and its applications.
J. Phys. B: Atomic Mol. Phys. 9: 3173.
18 Buckingham, A.D. (1967). Permanent and induced molecular moments and
long-range intermolecular forces. Adv. Chem. Phys. 12: 107.
19 Casimir, H.B.G. and Polder, D. (1948). The influence of retardation on the
London-van der Waals forces. Phys. Rev. 73: 360.
20 Mavroyannis, C. and Stephen, M.J. (1962). Dispersion forces. Mol. Phys. 5: 629.
21 Tang, K.T. (1969). Dynamical polarizabilities and van der Waals coefficients.
Phys. Rev. 177: 108.
22 Kramer, H.L. and Herschbach, D.R. (1970). Combination rules for van der Waals
force constants. J. Chem. Phys. 53: 2792.
108 3 Quantitative Treatment of Intermolecular Forces
23 Linder, P. and Lowdin, P.O. (1968). Upper and lower bounds in second order
perturbation theory and the unsold approximation. Int. J. Quantum Chem.
Symp. Series 2 S: 161.
24 Goscinski, O. (1968). Upper and lower bounds to polarizabilities and van der
Waals forces. Int. J. Quantum Chem. 2: 761.
25 Langhoff, P., Gordon, R.G., and Karplus, M. (1971). Comparison of dispersion
force bounding methods with applications to anisotropic interactions. J. Chem.
Phys. 55: 2126.
26 Slater, J.C. and Kirkwood, J.G. (1931). van der Waals forces in gases. Phys. Rev.
37: 682.
27 Dalgarno, A. (1967). New methods for calculating long-range intermolecular
forces. In: Advances in Chemical Physics, 143. New York: Wiley.
28 Hohenberg, P. and Kohn, W. (1964). Inhomogeneous electron gas. Phys. Rev. B
136: 864.
29 Lundqvist, S. and March, N.H. (eds.) (1983). Theory of the Inhomogeneous
Electron Gas. New York: Plenum Press.
30 Trickey, S.B. (Special Editor) (1990). Advances in Quantum Chemistry: Density
Functional Theory of many-Fermion Systems. San Diego, CA, USA:
Academic Press.
31 Parr, R.G. and Yang, W. (1989). Density Functional Theory of Atoms and
Molecules. Oxford, England: Oxford University Press.
32 Kohn, W. and Sham, L.J. (1965). Quantum density oscillations on an
inhomogeneous electron gas. Phys. Rev. A 137: 1697.
33 Gordon, R.G. and Kim, Y.S. (1972). Theory for the forces between closed shell
atoms and molecules. J. Chem. Phys. 56: 3122.
34 Kim, Y.S. and Gordon, R.G. (1974). Unified theory for the intermolecular forces
between closed shell atoms and ions. J. Chem. Phys. 61 (1).
35 Parker, G.A., Snow, R.L., and Pack, R.T. (1976). Intermolecular potential surfaces
from electron gas methods. I. Angle and distance dependence of the He + CO2
and Ar + CO2 interactions. J. Chem. Phys. 64: 1668.
36 Clugston, M.J. (1978). The calculation of intermolecular forces. A critical
examination of the Gordon-Kim model. Adv. Phys. 27: 893.
37 Cohen, J.S. and Pack, R.T. (1974). Modified statistical method for intermolecular
potentials: combining rules for higher van der Waals coefficients. J. Chem. Phys.
61: 2372.
38 Smith, F.T. (1972). Atomic distortion and the combining rule for repulsive
potentials. Phys. Rev. A 5: 1708.
39 Gunnarsson, O. and Lundqvist, B.I. (1976). Exchange and correlation in atoms,
molecules, and solids by the spin density functional formalism. Phys. Rev. B
13: 4274.
40 Gaydaenko, V.I. and Nikulin, V.K. (1970). Born-Mayer interaction potential for
atoms with Z=2 to Z=36. Chem. Phys. Lett. 7: 360.
References 109
41 Knowles, P.J. and Meath, W.J. (1986). Non-expanded dispersion and induction
energies, and damping functions for molecular interactions with application to
HF-He. Mol. Phys. 59: 965.
42 Tang, K.T. and Toennies, J.P. (1984). An improved simple model for the van der
Waals potential based on universal damping functions for the dispersion
coefficients. J. Chem. Phys. 80: 3726.
43 Dobson, J.F. and Gould, T. (2012). Calculation of dispersion energies. J. Phys.
Condens. Matter 24: 073201.
44 Berland, K., Cooper, V.R., Lee, K. et al. (2015). van der Waals forces in density
functional theory: a review of the vdW-DF method. Rep. Prog. Phys. 78: 066501.
45 Grimme, S., Hansen, A., and Brandenburg, J.G.C. (2016). Dispersion corrected
mean-field electronic structure methods. Chem. Rev. 116: 5105.
46 Zhao, Y. and Truhlar, D.G. (2008). The M06 suite of energy functionals:
systematic testing of four M06 functionals and 12 other functionals. Theor.
Chem. Accounts 120: 215.
47 Levine, I.N. (2000). Quantum Chemistry, 5ee. New Jersey, USA: Prentice-
Hall Inc.
48 Bransden, B.H. and McDowell, M.R.C. (1992). Charge Exchange in the Theory of
Ion-Atom Collisions. Oxford, England: Clarendon Press.
49 Atkins, P.W. and Friedman, R.S. (1997). Molecular Quantum Mechanics.
Oxford, England: Oxford University Press.
50 Weinhold, F. (1971). Dickinson energy of H2+. J. Chem. Phys. 54: 530.
51 Szabo, A. and Ostlund, N.S. (1982). Modern Quantum Chemistry. New York:
Macmillan.
52 Hehre, W.J., Radom, L., Schleyer, P.v., and Pople, J.A. (1986). Ab initio Molecular
Orbital Theory. New York: Wiley.
53 Bartlett, R.J. (1981). Many-body perturbation theory and coupled cluster theory
for electronic correlation in molecules. Annu. Rev. Phys. Chem. 32: 359.
54 Jeziorski, B., Moszynski, R., and Szalewicz, K. (1994). Perturbation theory
approach to intermolecular potential energy surfaces of van der Waals
complexes. Chem. Rev. 94: 1887.
55 Messiah, A. (1962). Quantum Mechanics, vol. 2. Amsterdam: North-Holland.
56 Bethe, H.A. and Salpeter, E.E. (1957). Quantum Mechanics of One- and Two-
electron Atoms. Berlin: Springer-verlag.
57 Hirschfelder, J.O. and Meath, W.J. (1967). The nature of intermolecular forces.
Adv. Chem. Phys. 12: 3.
58 Herzberg, G. (1950). Molecular Spectra and Molecular Structure I. Spectra of
Diatomic Molecules, 2e. Princeton, NJ: Van Nostrand.
59 Chang, T.Y. (1967). Moderately long range intermolecular forces. Rev. Mod.
Phys. 39: 911.
60 Power, E.A. (1967). Very long-range (retardation effect) intermolecular forces.
In: Intermolecular Forces (Advan. Chem. Phys. vol. 12), 167. New York: Wiley.
110 3 Quantitative Treatment of Intermolecular Forces
61 Levin, F.S. and Micha, D.A. (eds.) (1993). Long-Range Casimir Forces: Theory and
Recent Experiments on Atomic Systems. New York: Plenum Press.
62 McLachlan, A.D. (1963). Retardation dispersion forces between molecules. Proc.
Royal Soc. A 271: 387.
63 Salam, A. (2011). Molecular quantum electrodynamics of radiation-induced
intermolecular forces. Adv. Quantum Chem. 62: 1.
64 Axilrod, B.M. and Teller, E. (1943). Interaction of the Van der Waals type
between three atoms. J. Chem. Phys. 11: 299.
65 Micha, D.A. (1972). Collision dynamics of three interacting atoms: the Faddeev
equations. J. Chem. Phys. 57: 2184.
66 Micha, D.A. (1985). Rearrangement in molecular collisions: a many-body
approach. In: Theory of Chemical Reaction Dynamics, vol. II, Chap. 3 (ed.
M. Baer). Boca Raton, Florida: CRC Press.
67 Micha, D.A. (2017). Quantum partitioning methods for few-atom and many-
atom dynamics. In: Advances in Quantum Chemistry, vol. 74, 107. London,
England: Elsevier.
68 Salam, A. (2016). Non-Relativistic Quantum Electrodynamics of the van der
Waals Dispersion Interaction. Cham, Switzerland: Springer.
69 Kihara, T. (1958). Intermolecular forces. Adv. Chem. Phys. 1: 267.
70 Kaplan, I.G. (1999). Role of electron correlation in nonadditive forces and
ab initio model potentials for small metal clusters. In: Advances in Quantum
Chemistry, vol. 31, 137. San Diego, CA: Academic Press.
71 McQuarrie, D.A. (1973). Statistical Mechanics. New York: Harper & Row.
111
CONTENTS
and its Cartesian axes fixed in space, a CMN-SF frame, decreases the required
number of degrees of freedom. Overall, translation can eliminate three degrees
of freedom, and in addition three (for a nonlinear system) or two (for a linear
system) variables can also be eliminated with overall rotational invariance so that
the rigid structure requires only Ndf = 3Nat − 6 variables, or Ndf = 3Nat − 5 for a
linear system. They can be chosen as atom–atom distances, of which there are
Ndf (Ndf − 1)/2. For a two atom system, this is the distance R between centers
of mass, and for a nonlinear three atom system, A–B–C one can choose
atom–atom distances RAB, RBC, RCA, or sometimes two bond distances RAB,
RBC and the common bond angle αABC. Generally, for a system with an interac-
tion energy invariant under overall rotations and described by a single potential
energy function (which must go to zero as any atom is removed to infinity) this
potential energy is a V({RIJ}) function which depends on the interatomic distances
between all atom pairs IJ. As distances are varied the energy generates a potential
energy surface (PES) in a space of Ndf (Ndf − 1)/2 degrees of freedom, with peaks
and valleys which can be pictured in three dimensions or as isocontours in two
dimensions, by freezing all but two variables at a time.
In addition to this distances dependence, the potential energy may also
depend on electron spin variables if the system has open electronic shells.
It may also be multivalued for a given atomic structure if more than one PES
is needed to account for electronic rearrangement due to electronic charge
transfer or excitation. This requires consideration of couplings between PESs
and is treated in a following chapter.
The potential can be decomposed in a variety of ways, some with more phys-
ical content than others. A number of criteria can be imposed on the form of
V({RIJ}) to make it useful in the treatment of energetics and dynamics properties.
They are the following:
V Rlm = l<m
V 2
Rlm + l<m<n
V 3
Rlm , Rmn ,Rnl + …
from which a general force field can be generated. The two-atom terms can be
parametrized as shown in the next section, while the three atom terms can be
added as corrections which are dampened as interatomic distances increase
beyond some boundary values R0lm , and have adjustable parameters at short
distances. Labeling the pair (lm) = μ = 1, 2, 3, …, and with Rlm = rμ, one such
function is
1
V 3
rμ ,rν ,rπ = μ
1 −tanh γ μ rμ − rμ0 b0 + b
μ μ
rμ − rμ0
2
+ b
μ < ν μν
rμ − rμ0 rν − rν0 + …
where the factor with a tanh goes to zero outside the boundary rμ0 values, and the
coefficients b can be chosen to shape the region within boundaries.
Alternative forms including also four-atom terms can be written in terms of
bond distances, bond angles, and dihedral (torsion) angles, as well as distances
between nonbonding atoms, and are the basis for treatments of the molecular
dynamics of polymers and biomolecules.
be given in terms of their values V = 0, Vm, Vi, … at special locations R0, Rm, Ri, …
at the zero-value or core, minimum, and inflection points, as shown in
Figure 4.1, with magnitudes and functional forms dependent on their physical
origin.
We deal here first with pairs where one or both atoms have a closed electronic
shell and are not charged, so that they do not form a chemical bond, but may
form a van der Waals complex with an interaction potential energy VvdW(R).
The potential minimum energy Vm = − ϵ and location Rm are chosen here as
units of energy and length, to introduce the reduced variables x = R/Rm and
f(x) = V(R)/ϵ. Analytical forms usually employed are the Lennard Jones (n,6),
Exp-6(α), also called the Buckingham or Slater potential, and the Morse(a)
potential. They all have three parameters, or only one in reduced variables.
Their form and range of the remaining parameter are
−1
fLJ x; n = n− 6 6x −n −nx −6 , 8 ≤ n ≤ 20
fE6 x; α = 6 exp − α x−1 − αx − 6 , 8 ≤ α ≤ 20
fM x; a = exp −2a x −1 − 2 exp − a x −1 , 4 ≤ a ≤ 8
From these, we can find the curvatures at the minimum to be, respectively,
d2 f α− 7
κ= = 6n, = 6α , = 2a2
dx2 x=1 α− 6
A +B
R0 Rm Ri R
Vm = –ε
AB
116 4 Model Potential Functions
2CAA CBB
CAB =
αB αA CAA + αA αB CBB
where αA is the static polarizability of atom A. For the short range we can use the
Born–Mayer potential
giving AAB = (AAAABB)1/2 and bAB = (bAA + bBB)/2. The parameter bAA derives
from the electronic charge distribution of the outer shell and the size of the
atomic core of inner shells; if the former is given (in a hydrogenic description)
by the effective nuclear charge ζ A, it follows that bAA = b0AA + c ζA , with the first
term given by the A core radius.
with N = 2n + 1. More elaborate models for very short distances have also been
introduced [13]. Interactions VAB(R) for heteronuclear systems follow from
combination rules there.
potential inside its inflection point using Rd = Ri. Choosing Rd = R0, and λ R0
the decomposition
S L
V R = gs R V R + g s R V R = V R +V R
separates short-range and long-range components, with the second term usu-
ally small at all distances and suitable for perturbative treatments of thermody-
namical and transport properties. This separation is shown in Figure 4.2.
Another use for these functions is in the switching from ionic to covalent
potential energy functions as distance increases, with a crossing at a location
Rd = Rx for example in alkali–halide interactions, where one can use
V R = Vion R gs R + Vcov R g s R
The damping functions, however, have large derivatives around Rd that can
affect the calculation of forces in studies of molecular dynamics, and their
change with R must be locally balanced in the transition region before deriva-
tives with respect to position variables are taken, to avoid artificial forces. Fur-
thermore, if the damping function is combined with a long-range interaction
going as R−n, then care must be taken to compensate for, or exclude, the diver-
gence of the interaction as R 0. This can be done, for example using instead
A+B
R0 R
VLR
AB
gS
1.0
R
0.0
4.3 Atom–Molecule and Molecule–Molecule Potentials 119
n n
fd R; an , bn = 1 − exp −an R− bn R2
which leads to a constant value at the origin for the potential energy containing
R−n, and has two parameters as needed to fix the transition location and its
width. An alternative proposed form contains only the Born–Mayer parameter
b and can be used if the origin is excluded [15].
Piecewise potentials can be put together from exponentials, Morse, and
van der Waals potentials at short, intermediate, and long-range distances,
respectively, and joining them with spline functions chosen to assure a smooth
function and smooth derivative; this is the so-called Exponential–Spline–
Morse–Spline–van der Waals (ESMSV) potential [16, 17]. The spline function
between points xj and xj + 1 is
k
3 x −xj
S x = a
k =0 k xj + 1 −xj
introduced with its bounds chosen to bracket the zero and inflection points and
can be used to construct very accurate potential functions and forces over a wide
range of distances. The four parameters are fixed by imposing continuity of the
function and its derivative at the two ends of its range. The spline can also be
used to smoothly join avoided crossing potentials, for example from an outer
covalent potential to an inner ionic one as in NaCl.
(a) (b)
B B
Rab
R
θ r
A A
Rac
C C
Figure 4.3 (a) Center-of-mass distance variables in the triatomic A + BC. (b) Variables for a
description in terms of atom-pair potential energies, for a chosen distance between B and C.
V R,r, θ = l
Vl R, r Pl cosθ
and the expansion functions can be further decomposed into short-range and
long-range terms as
S L
Vl R,r = Vl R,r + Vl R,r
S
Vl R,r = Al r exp − bl r R
R R−n
L n
Vl R, r = vBC r δl0 + n Cl , n r fd
Only a few Legendre polynomials are needed for weak anisotropy and, if M is a
homonuclear BB molecule, symmetry eliminates all odd polynomials from the
expansion. The A, b, and C parameters can be expanded in powers of the dis-
placement ρ = r − re from the equilibrium bond length re. For each angle, it is
possible to show potential energies as isocontours of V versus R and r. Isocon-
tours of potential energy surfaces describing the atomic interactions are shown
in Figure 4.4. The top panel is the same as Figure 1.5 in Chapter 1: Part (a),
including also the internal diatomic potential, vBC (r) which must be added
to the long-range interaction for l = 0 (Figure 4.4). Part (b) shows isocontours
for a reaction A + BC AB + C with an activation barrier, such as H + H2
H2 + H, while Part (c) shows isocontours for a stable intermediate ABC such
as H2O with a fixed bond angle, and varying atom–atom distances Rab and
Rbc. The bottom panel is the same as Figure 1.6, and shows the potential energy
of H approaching CN in HCN, in a (x,y) plane with the coordinate origin at the
midpoint between C and N.
An alternative parametrization avoids the expansion in polynomials and
instead introduces parameters dependent on r and θ, for each value of the latter.
4.3 Atom–Molecule and Molecule–Molecule Potentials 121
r BC
r A,BC r AB r AB
3.0
2.0
yH (Å)
S
1.0
C N
A B
0 • •
–3.0 –2.0 –1.0 0.0 1.0 2.0 3
xH (Å)
Figure 4.4 Isocontours for three-atom systems. Top panel: generic isocontours for (a) an inert
gas A interacting with a diatomic BC, (b) the reaction A + BC AB + C with an activation
barrier, and (c) a stable intermediate ABC with a fixed bond angle, from [18]. Bottom panel:
potential energy of H approaching CN in HCN, in a (x,y) plane with the coordinate origin at the
midpoint between C and N, from [19]. The locations of minima are shown as A and B, for the
two isomers, and a saddle point is shown as S.
For example the three atoms may be on a line or in a “T” conformation, with
different parameters. Then we have
S L
V R,r,θ = V R,r,θ + V R,r,θ
V S
R,r, θ = A r,θ exp − b r,θ R
R R −n
n
V L
R,r,θ = vBC r + n Cn r, θ fd
This may be described as a method of variable parameters, where the atom–
atom functional forms are used, but the parameters are allowed to change with
internal coordinates and orientation of the molecules.
122 4 Model Potential Functions
V S
R,r,θ = C exp − α0 R + α0 ρR A θ + B θ ρ
V L
R,r,θ = −R − 6 C6 θ fd R
with the first term giving the short-range repulsive interaction between the
CMN of A and B, and with the summation containing locations and values
of charges chosen to reproduce molecular charge densities and their interaction
and suitable only for large distances Rab. This information can be extracted from
an electronic structure calculation of the AB pair, which would provide a charge
and bond order matrix from which charges and locations can be defined [22].
The set of charges at each molecule can alternatively be replaced by a set of
distributed multipoles, which then give the total potential energy in terms of
multipole–multipole interactions [5]. Figure 4.5 shows the TIP5P model for
water [23], with four charges a = 1–4 representing hydrogens with positive point
4.3 Atom–Molecule and Molecule–Molecule Potentials 123
0.70 Å
109.47° 0.9572 Å
+
– 104.52°
Figure 4.5 The H2O monomer TIP5P charge model used for H2O–H2O interactions.
charges of 0.241e and lone pairs with balancing negative values, in addition to
the oxygen atom which interacts with other oxygen atoms only through a pair
potential.
The water–water interaction can then be constructed from an LJ(12,6) poten-
tial energy function between the two oxygen atoms with parameters σ 0 = 3 12 A
for its core radius and εOO = 0.16 kcal mol−1 for its well depth, plus point charge
interactions as shown in Figure 4.5 for the hydrogen atoms and lone pairs, and is
of the form
12 6
σ0 σ0 ca cb
VWW = 4εOO − +
ROO ROO a, b
4πε0 Rab
ε r, θ = ε
l l
r Pl cos θ
and similarly for the other parameters for small anisotropies. For example with
the Exp-6(α) parametrization, the short-range exponential parameter α can be
made to vary with the BC distance and its orientation relative to A, and can be
expanded in Legendre polynomials if the anisotropy is small. For interacting
molecules A and B, this generalizes to VAB(R,QA,ΓA,QB,ΓB) = V[R; ε(QA,ΓA,
QB,ΓB), Rm(QA,ΓA,QB,ΓB), γ(QA,ΓA,QB,ΓB)].
m 12 m 10
RHB RHB
VAHB = εHB 5 −6 fang cosαAHB
RHB RHB
V Ra ,Rb ,Rc = v
i = b, c AI
Rai ,ωi
S L
vAI Rai , ωi = vAI Rai ,ωi + vAI Rai , ωi fd, i Rai
where Rai = Ra −Ri , and the long-range term has been dampened at short
interatomic distances.
AAB = d 1 d 2 χ a 2 ∗ χ b 1 ∗ H AB χ a 1 χ b 2
where the + sign corresponds to the singlet, and the VB Coulomb integral QAB
and exchange integral AAB are shown in terms of the atomic orbitals while
ΔAB = Sab2
is the square of the integral of the orbitals overlap. Reverse relations
give expressions for the integrals in terms of the singlet and triplet energies,
which are known from accurate calculations or from spectroscopic measure-
ments. The singlet and triplet energy functions can be parametrized as Morse
fM(x; a) = exp[−2a(x − 1)] − 2 exp[−a(x − 1)] and anti-Morse fAM(x; b) = exp[−2b
(x − 1)] + 2 exp[−b(x − 1)] potential functions respectively, with adjustable
parameters. The overlap integral follows from the form of atomic orbitals,
1 2
which for 1s orbitals of the form χ a r = ζ 3a πa30 exp − ζa r −Ra a0 gives
2
Sab(Rab) = [1 + (ηRab/a0) + ((ηRab/a0) /3)] exp(−ηRab/a0) and can be used as an
adjustable function of the parameter η for each atom pair, to shape the PES.
126 4 Model Potential Functions
For the three-atom system ABC with three active electrons described with
three atomic orbitals, a London-like interaction potential energy function suit-
able for treatments of reactive atom-diatom systems undergoing a rearrange-
ment such as A + BC AB + C can be constructed from a generalized VB
description [22, 24]. Two three-electron VB wavefunctions 2ΦI and 2ΦII can
be formed to describe the doublet states corresponding to two independent
bondings A─B C and A B─C. The lowest PES is obtained from the Ham-
iltonian eigenvalues in this two-state description and can be given a simple form
provided orbital overlaps are neglected. It is however made more flexible, so it
can be applied to a variety of triatomics, by reintroducing overlap terms para-
metrized to reproduce desired features such as the location of a surface saddle
point at an energy barrier for reaction, in which case it is the so-called London–
Eyring–Polanyi–Sato (LEPS) function [3, 25, 26]
−1
VABC = 1 + ΔABC QAB + QBC + QCA − 2 − 1 2
AAB − ABC 2
2 2 1 2
+ ABC −ACA + ACA − AAB
and the integrals can be obtained from known singlet and triplet potential ener-
gies for each pair, while the triatomic overlap ΔABC can be used as an adjustable
parameter to obtain a desired activation barrier in a rearrangement A + BC
AB + C. In the Polanyi version [27], the term ΔABC is omitted and replaced with
three distance dependent parameters ΔIJ = SIJ2 introduced in the expressions for
the singlet and triplet energies of each IJ pair, from which the VB Coulomb and
exchange integrals have been obtained, to allow for desired energy barrier and
added attraction or repulsion between reactants or products. This gives
VIJ + 3 VIJ + ΔIJ 1 VIJ − 3 VIJ
1
QIJ =
2
in terms of singlet and triplet potential energy functions of R and an overlap inte-
gral, and with a similar expression for AIJ replacing 3VIJ with −3VIJ. A generic iso-
contour for a reaction with an activation barrier is found in the top Figure 4.4a,
while Figure 4.6 shows the case of the collinear F+ H2 FH + H reaction with an
activation barrier in the entrance channel. An analogous PES expression can be
written for systems with four active electrons [22, 26] such as H2 + D2.
Spin-dependent pair potentials may instead be used to write the interaction
of open shell atoms in terms of spin operators, which can be parametrized
using pair interaction functions for specific multiplet states. These
potentials can be derived from electronic structure states in a VB formulation
[30]. For three monovalent atoms the pair potential operators are V IK RIJ =
c s
VIJ RIJ + VIJ RIJ S I S J where S I is the spin operator of atom I with two spin
states. For three monovalent atoms, the spin operator V AB + V BC + V CA can be
diagonalized in a basis set of two doublet three-atom spin states to recover the
LEPS expression.
4.4 Interactions in Extended (Many-Atom) Systems 127
2
–35
–30
RH–H(Å)
–25
–20
1 –15
–10 +5
–5 +1.6
0 +1.6
1 2
RH–F(Å)
Figure 4.6 Energy isocontours for collinear FH2 versus atom–atom distances, showing an
early activation barrier as F approaches H2. Source: from Bender et al. [28], almost identical to
the figure from Bender et al. [29].
In metals, such as Cu(s), some electrons are localized into ion cores and others
are delocalized over the whole crystal with the exclusion of the ion cores, and
treatments can be based on the concept of embedded atoms [31]. This is also
helpful in the treatment of unsaturated hydrocarbon chains with delocalized
electrons.
The cohesive energy of a molecular crystal with Nat atoms at rest in a sample
volume Ω, such as crystals composed of identical noble gas atoms, can be writ-
ten as a sum of pair potentials v(Rij) between atoms i and j in a lattice, with suit-
able parameters which must depend on the atomic density ρsld = Nat/Ω. The
solid potential energy is then approximated by
1
Vsld = v Rij
2 i j
with pair potential parameters vm(ρsld) for the well depth, and l(ρsld) for the near-
est neighbors distance, at the minimum of energy between them, as functions of
the solid density. It can be conveniently re-expressed in terms of a relative atom–
atom distance pij = Rij/l, so that the lattice sum for a long range R−n interaction
would be (1/2) i j(Rij)−n = l−nαn, with a lattice sum αn = (1/2) i j(pij)−n
characteristic of the lattice symmetry and independent of density. A starting
point for calculation of the cohesive energy is a sum of pair potentials for the
isolated pair, with ρsld = 0.
For the LJ(12,6) pair potential written in terms of its energy minimum vm|
and the location R0 of its zero energy value, and with a close packed (fcc) lattice
structure at equilibrium, the potential energy for identical noble gas atoms is
12 6
1 R0 R0
Vlatt = Nat 4 vm α12 fcc − α6 fcc
2 l l
The lattice summations have been evaluated and are α12(fcc) = 12.13188 and
α6(fcc) = 14.45392, not far from the number 12 of nearest neighbors. The equi-
librium distance leq is obtained setting the force dVlatt/dl = 0 to get leq/R0 = 1.09,
indicating that the internuclear distance has increased going from an isolated to
an embedded atom pair. The measured values for Ne, Ar, Kr, and Xe are 1.14,
1.11, 1.10, and 1.09, in quite good agreement. The remaining discrepancy can be
corrected by including in the cohesive energy the zero-point energy of vibra-
tional motion of each atom as it oscillates around its equilibrium position [31].
Ionic crystals, such as NaCl(s) with the fcc lattice symmetry, contain short-
range repulsions and long-range Coulomb electrostatic attractions. They are
treated introducing a diatomic ion, like Na+Cl−, in a unit cell of a lattice, and
adding over atom pair interactions. The solid cohesive energy is
Z + Z − c2e
Vlatt = Ndiat v SR l − α1 latt
4πε0 εr l
4.4 Interactions in Extended (Many-Atom) Systems 129
where l is the distance between adjacent positive and negative ions, εr is the
dielectric constant of the medium, the charge numbers Z are positive, and
α1 = αM is the Madelung lattice constant. It is obtained from two interpene-
trating fcc lattices for Na+ and Cl−. Taking one of the negative ions as a refer-
ence center in the solid, one can write the constant as a sum of terms for its
interaction with positive and negative ions, and the sum can be done partition-
ing the lattice around the reference ion into nearly neutral groups of ionic
charges in cubes of increasing size until convergence. The result for the fcc
lattice is 1.747558. For comparison, the value for the bcc lattice, such as in
the CsCl(s) crystal, is 1.762670 [31]. The short-range term can be written
as a Born–Mayer exponential repulsion A exp(−α l) and the equilibrium value
leq of the interionic distance can be found from the location of the zero-value
force. Replacing this into the potential energy one finds the Born–Mayer equi-
librium form
Z + Z − c2e 1
Vlatt = −Ndiat 1− αM latt
4πε0 εr leq αleq
This shows that cohesion increases for more highly charged ions and for more
steeply repulsive short-range interactions.
More accurately, the cohesive energy must also contain three- and many-
atom terms such as l < m < nV(LR)(Rlm, Rmn, Rnl ), to account for nonadditivity
of long-range interactions. The many-atom effects can be included in a variety
of ways which depend on the nature of the structure. For solid crystalline Si, a
useful functional form has been given by Tersoff [32], which incorporates a
many-atom factor in the long-range interaction. The total energy can be
written as
V Rab = f
l<m c
Rlm Aexp −λ1 Rlm − B Rab exp −λ2 Rlm
The first term in the square parenthesis describes the repulsion energy
between atoms l and m, and the second term gives a modified attraction energy
where the function B({Rab}) depends on all the atom–atom distances needed to
describe the environment of the (lm) pair. This attraction factor is constructed
as a monotonically decreasing function of the number of bonds of l and m with
other atoms. It also depends on the strength of the bonds to those two atoms
competing with the (lm) bond, and on the angles between the (lm) bond and the
other bonds. The cut-off function fc(R) is chosen to equal 1.0 at short distances
and to smoothly fall to 0.0 at large distances. Results for various phases of solid
structures can be fitted with suitable parameters this way.
The zero-point vibrational motion energy of the lattice atoms must be added
even at the temperature of 0 K. If the temperature is increased, then parameters
like the atom–atom distance in the expressions for Vlatt depend on the density
130 4 Model Potential Functions
pot
ρsld and temperature in β, and become thermal potential energies Ulatt to
which one must add the thermal kinetic energy of vibrating atoms in a lattice,
to obtain the internal energy Ulatt, or cohesive energy, of the solid.
The solid dielectric properties alter the strength of the electrostatic and polar-
ization interactions, decreasing them by the magnitude of the dielectric con-
stant εr of the medium and affecting the values of the Vlatt potential energy.
A general treatment of dielectric effects can be based on the dynamical suscep-
tibility of the solid, by considering the response of the whole system to an
applied external electric field ext [31]. The local electric field lcl = εr ε0 ext at
the position of an atom in the solid can be constructed adding fields coming
from the surfaces of the solid with an inner idealized cavity, plus the field of
the atoms inside it, as
where out is the depolarization field from charges at the outer surface of the
solid, inn is the field from the charges at the inner surface of the cavity, and
atm is created by all the polarized atoms in the cavity. The sum
ext + out = appl , also called the Maxwell field, depends on the shape of the outer
surface and gives the total applied field. The field from the inner surface charges
is the so-called Lorentz field which for a spherical surface can be obtained from
the polarization vector P = j ρjd j = j ρj αj lcl of the sample, where ρj is the
number of atoms of type j per unit volume, with atomic induced dipole d j
and polarizability αj, and equals inn = P 3ε0 [31]. The last term atm depends
on the distribution of atoms in the cavity and can be seen by symmetry that it
equals zero for a cubic lattice. This gives lcl = appl +P 3ε0 in SI units, with
the dielectric susceptibility χ = εr − 1 introduced by P = ε0 χ appl . This assumes
an isotropic medium or a cubic crystal structure. More accurately, one must use
a susceptibility tensor χ ξη and extend the treatment using Cartesian compo-
nents in Pξ = ε0 η χ ξη appl, η . Considering a single component and isotropy, with
P= j ρj αj appl + P 3ε0 , and solving for χ = P ε0 appl one obtains the
relation [31]
εr − 1 1
= ρα
j j j
εr + 2 3ε0
which provides the dielectric constant needed in the solid potential energy Vlatt.
4.4 Interactions in Extended (Many-Atom) Systems 131
ρ n R1 ,…,RN = N N − n P n R1 ,…,RN
where the factor to the right comes from integration of the whole probability
over the position variables of N–n particles. In particular, when n = 1 we have
−1
for liquid volume Ωliq that Ωliq d 3 R1 ρ 1 R1 = N Ωliq = ρ, while ρ 2 R1 ,R2
describes the distribution of atom pairs and ρ(3) gives the triplets distribution.
Liquid structure correlation functions g n
R1 , …,RN are defined by ρ(n) = ρng(n).
In an isotropic liquid, these functions are invariant under reference frame rota-
tions and depend only on relative distances, so that, for example the relative dis-
tribution of atoms at a distance R = R1 −R2 | around a fixed one at position
R1 = 0 averaged over orientations is ρ 2 R ρ = ρg 2 R , which leads to
g(2)(∞) ≈ 1.
Interaction energies in liquids and in liquid solutions can be obtained from
thermal averages and are constructed from statistical pair distribution functions
g 2 = g R; ρ,β , dependent on the liquid number density ρ and an inverse tem-
perature β = 1/(kBT), with ρg(R; ρ, β)R2dR counting the number of atoms found
in a spherical shell of width dR. The simplest dependence, only on R, is for
closed shell atoms, and this must be generalized to include variables for orien-
tation angles of molecules, or for anisotropies of open shell atoms in covalent
bonding. The function g(R; ρ, β) can be measured with X-ray diffraction meth-
ods, or calculated using molecular dynamics.
For a closed thermodynamical system described by a statistical canonical
ensemble with N 1 atoms in a volume Ω, and density ρ = N/Ω, with up to
n atoms interacting with a potential energy Vn R1 ,…,RN , the n-ple distribution
function is given by [33]
−1
g n
R1 ,…,RN ; ρ,β = Ω N ZN Ω, β d 3 RN + 1 …d 3 RN exp −βVn R1 , …,RN
The pair distribution g(R; ρ, β) is small for small R where repulsion takes place,
followed by a large peak around the distance Rm for the pair potential energy
minimum, and shows subsequent minima and maxima corresponding to
atomic shells, as distance increases [33]. A pair distribution function is shown
in Figure 4.7 obtained from a molecular dynamics simulation with the Lennard–
Jones pair potential function with parameters ε and R0, as a function of the
reduced distance R/R0 and for the reduced temperatures T ∗ = kBT/ε and den-
sities ρ∗ = R03ρ, with R0 the core radius.
The potential energy of interaction in the simplest case, when the density is
low and many-atom effects can be ignored, is given as a sum of pair v(R) poten-
tial energies as
∞
pot 1 3 N
Uliq ρ,β = d R1 d3 R2 v R12 ρ 2 R1 ,R2 = ρ dR 4πR2 v R g R; ρ,β
2 2 0
kin
To this one must add the thermal kinetic energy Uliq ρ,β = 3NkB T 2 =
3N 2 β to obtain the thermodynamical internal energy
∞
3N N
Uliq ρ,β = + ρ dR 4πR2 v R g R; ρ,β
2β 2 0
or cohesive energy, at the given density and temperature. Different choices for
the atom pair potential have been used in the calculations of correlation
functions.
pot
A perturbation theory for calculating Uliq ρ,β can be based on the assump-
tion of additive pair potentials and separation of short-range and long-range
S L
contributions in v(R) = v(S)(R) + v(L)(R), so that VN = VN + VN , with the
0.5
0.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
R/R0
4.4 Interactions in Extended (Many-Atom) Systems 133
0
short-range term used to construct a reference potential energy Uliq ρ,β =
S
Uliq ρ,β , and the long-range part treated as a perturbation to obtain a correc-
L
tion Uliq ρ,β = Uliq ρ,β . The short-range summations over pairs can be usu-
ally limited to the average number zNN of nearest neighbors, while the long-
range part is treated as a small perturbation [33]. Restricting the nearest neigh-
bors to a narrow distance around the maximum of the pair distribution func-
tion, which coincides with the minimum of the pair potential, provides a
S
rough estimate of the short-range potential energy as Uliq ρ, β = Nv Rm zNN 2.
Contributions from the LR interactions can be described by means of the
Fourier components of the new function h R = g R − 1, given by the structure
factor
−3 2
h k = ρ 2π d 3 R exp ik R g R −1
from which the pair distribution function is obtained by the inverse Fourier
transform, as
−3 2
g R = 1 + 2π d 3 k exp −ik R h k ρ
L
Replacing this in the integrals over R for Uliq ρ,β we find the long-range
potential energy of the liquid solution expressed in terms of the Fourier trans-
L
form v k of the long-range part of the pair interaction potentials, and of the
structure factor, as
∞
L N
Uliq = d3 k v L
− k h k −ρ dR 4πR2 v L R
2 0
with
−3 2
v L
k = 2π d 3 R exp ik R v L R
For the present potential and distribution functions dependent on only the
radial variable, the transforms are
∞
L −3 2 −1
v k = 2π k dR 4πR sin kR v L R
0
∞
− 3 2 −1
h k = ρ 2π k dR 4πR sin kR g R −1
0
134 4 Model Potential Functions
and the coefficients can be related to partial potentials of mean force for mole-
cules with fixed orientation. Expansions in terms of the Wigner rotational func-
l
tions Dmm Γj [21] can instead be done for nonlinear molecules.
An alternative for interacting polyatomics is to represent each one with a set
of potential centers a and b, possibly anisotropic, for A and B, respectively, and
to introduce center-pair distribution functions so that, for example for interact-
ing H2O molecules one would define pair distributions gHH, gOO, and gHO and
use them to construct the internal energy of liquid water.
As in the case for solids, the long-range electrostatic and polarization inter-
action energies depend on the dielectric constant of the liquid. This can be
obtained from the polarization P liq = j ρjd j = j ρj αj lcl given by the sum of
induced atomic dipoles per unit volume in an external electric field. In the sim-
ple case of a liquid of average density ρ with a single atomic species of induced
dipole d R , the local polarization is P liq R = d R ρ R , where
ρ R = jδ R−Rj is the local density of atoms. Its thermal average is usually
a smooth function over space within the liquid, which introduces the liquid pair
distribution function in ρ R = ρg R; ρ, β , giving
in terms of the local atomic polarizability α R . From this, one can obtain the
dielectric constant εr of the liquid as done before for a solid, given by
εr − 1 1
= d 3 Rρ α R g R; ρ,β
εr + 2 3ε0
which however must be generalized for anisotropic species requiring a polariz-
ability tensor, and for polar molecules with a permanent dipole [34].
The treatment is simpler when there are no chemical bondings between solute
and solvent species, and one can use already constructed pair and triplet inter-
action energies. The liquid solution total interaction energy takes the form
Coul pol rep Coul pol rep
Vsol = VMS + VMS + VMS + VSS + VSS + VSS
which includes Coulomb (or electrostatic), polarization (induction and disper-
sion), and repulsion (electron penetration and exchange) contributions between
int
the molecule and solvent species through the pair potential energies vMS , with
int
int = Coul, pol, rep, and solvent–solvent interaction potential functions vSS ,
plus possibly contributions from triplet interactions. The interactions among sol-
vent species have been described in the previous section. Here we concentrate on
Coul pol rep
the total interaction potential energy Vslt = VMS + VMS + VMS acting on the
solute molecule M.
It helps to distinguish short-range (SR) and long-range (LR) interactions, with
SR sums containing only nearby molecules, while sums of LR terms are averaged
over large distances, are smoother, and can be expanded in a Fourier series or
averaged over space. To proceed, it is convenient to separately treat short-range
rep
interactions in VMS as differently from the other, long-range interactions.
Short-range interactions between solute or solvent molecules are important
only between nearest neighbors or next nearest neighbors. We assume to sim-
plify first that the molecules are isotropic on average. The short-range (repul-
SR rep
sion) interaction energy Vslt = Vslt for the solute is a sum of short-range pair
potentials for atoms in the solvent S, which can be written as
SR SR
Vslt = v
n S MS
RM −Rn
However, the SR potential extends only a short distance away from the origin
and involve only the regions of pair distributions coming mostly from nearest
neighbors, and next-nearest-neighbors shells to a smaller extent. Restricting the
nearest-neighbors to a narrow distance around the maximum of the pair distri-
m
bution functions, which coincide with the minima RMS of the pair potential, a
rough estimate of the short-range potential energy near thermal equilibrium at
the solute molecule M is
SR SR m NN
Vslt ≈ vMS RMS zMS
eql
NN
where zMS is the number of nearest neighbor species S. The SR term can also
be written as
SR SR
Vslt = d 3 R vMS RM − R ρ S R
4.5 Interaction Energies in a Liquid Solution and in Physisorption 137
−1
ρS R = ZN Ω,β d 3N R exp − βVN R1 , …,RN δ R −Rn
n
to obtain
SR SR
V slt RM = ρS d 3 R vMS RM − R gMS R; ρS , β
in terms of the pair distribution function gMS for the M and S species insofar
ρS R = ρS gMS R; ρS , β , with ρS the number of solvent molecules per unit vol-
ume, that can be obtained from a potential of mean force. More generally the
pair potential function of R must be replaced by a function containing internal
distances and orientation angle variables for the M and S molecules, and statis-
tical averages can be performed over all internal variables.
The remaining long-range interactions due to electrostatic and polarization
terms are again given by sums of long-range pair potentials suitably cut-off
LR Coul pol LR
at short distances, vMS = vMS + vMS , which contribute to an energy Vslt .
The long-range term becomes
LR LR LR
Vslt = v
n Mn
RM −Rn = d 3 R vMS RM − R ρS R
This integrand is usually a smooth function over space within the solid. It can
be thermally averaged to obtain
LR LR
V slt RM = ρS d 3 R vMS RM − R gMS R; ρ, β
Here, the radial integral is nearly zero for values smaller than the radius RMS,0,
where the LR potential is large but the pair distribution function vanishes.
Beyond this radius the integrand is a smooth function of R which can be
approximated by an analytical expression. It can be obtained for an attractive
polarization potential of form vMS R = −CMS R −s , with s = 6 for a dispersion
LR s
the order of the sum of van der Waals radii for M and S. With the origin of coor-
dinates at the CMN of M and for an isotropic solvent, introducing the relative
position vector r =RM − R we find
∞
LR LR LR
V slt = ρS d3 r vMS r gMS RM − r ; ρ, β ≈ρS 4πr 2 dr vMS r
pol c
RMS
c −s + 3
s
4πρS CMS RMS
≈
−s + 3
giving a negative (attractive) solute potential energy which is larger for a smaller
solute cavity radius. A similar approximation for the solvent density is obtained
c
assuming that solvent atoms are packed with hard sphere radius RSS so that
3 −1
c
ρS = 4π RSS 3 , the inverse of the volume per solvent species.
A more accurate treatment can be derived using what is known about the pair
distribution function and related structure factor, from measurements or from
molecular dynamics calculations. Introducing the Fourier transform of the LR
pair potential and its inverse transform,
LR −3 2 LR
v MS k = 2π d 3 R vMS R exp ik R ,
LR −3 2 LR
vMS R = 2π d 3 k v MS k exp − ik R
LR −3 2 LR
Vslt = 2π d 3 k v MS k exp − ik RM n
exp ik Rn
Further, considering that the transform of the thermally averaged number den-
−3 2
sity ρS R of the solvent is ρ S k = 2π n exp ik Rn we find for the
th
LR part,
LR LR
V slt = d 3 k v MS k ρ S k exp − ik RM
A general way to evaluate this involves the introduction of the Fourier transform
hMS k of the M–S pair correlation function hMS R = gMS R − 1, that is a
structure factor measurable by X-ray scattering. In detail,
−3 2
hMS k = ρS 2π d3 R exp ik R gMS R − 1
4.5 Interaction Energies in a Liquid Solution and in Physisorption 139
from which the pair distribution function is obtained by the inverse Fourier
transform, as
−3 2
gMS R = 1 + 2π d 3 k exp − ik R hMS k ρS
Replacing this in the integrals over R, we find the solute long-range potential
energy expressed in terms of the Fourier transforms of the long-range part of
the pair interaction potential and of the structure factors, as
LR LR LR
V slt RM = d 3 k v MS k hMS k exp −ik RM −ρS d 3 R vMS RM − R
Further simplification follows from observing that for the present isotropic
potentials and distributions, the potential Fourier transforms and the
present structure factors depend only on the wavevector magnitude k and
are given by
∞
L − 3 2 −1 L
v MS k = 2π k dR 4πR sin kR vMS R
0
∞
−3 2 −1
hMS k = ρS 2π k dR 4πR sin kR gMS R − 1
0
L
which allows calculation of Vslt from one-dimensional integrals over k. The
structure factor hMS(k) can also be calculated from pair potentials by means
of perturbation theory [33].
The extension to nonspherical, rotating, and vibrating M and S molecules
requires reintroduction of internal orientation angles and atomic displacement
variables in pair potentials and in pair distribution functions, and statistical
averages over all internal variables of M and S components of the liquid solution.
The mean interaction wMM(R) between two solute molecules in the solution
differs from the bare interaction pair potential vMM(R) due to medium effects
and is a potential of mean force which depends on the dielectric constant of
the solvent. A general treatment can be based on the susceptibility function
of the solvent, by considering the response of the whole system to an applied
external electric field ext [36].
z-axis perpendicular to the solid surface and pointing into a vacuum. The solid
may contain several types of atoms, and may have a crystalline structure or be
amorphous, such as a silicon solid with its surface dangling bonds saturated with
hydrogen (Si(s):H), or titanium dioxide (TiO2(s)). In the absence of chemical
bonding between the adsorbate and the surface atoms, it is usually accurate
to construct the total adsorption potential energy Vads as a sum of pair potential
energies vAn RA −Rn , so that for fixed positions of atoms in the solid,
Vads RA = n vAn RA −Rn . The treatment is then similar to the one for a sol-
ute in a solvent. However, for metal and semiconductor surfaces with deloca-
lized electrons, the pair potentials vAn must be constructed to account for
embedding effects as treated in Chapter 8, and are also functionals of the elec-
tronic densities at the surfaces. What follows here applies to nonbonding adsor-
bates and localized surface electrons.
Separating short range (SR) and long range (LR) terms in the pair potential
SR LR
one can write Vads = Vads + Vads to derive approximations. The short range
potential can be obtained from the nearest neighbors (NTs) of the adsorbed atom,
so that
SR SR
Vads RA ≈ v
n NT An
RA −Rn
which can be evaluated for a known location of the adsorbate on the surface, and
from the atomic structure of the surface and inner layers. This can be improved
by adding next-NTs.
The long range term becomes
LR LR
Vads RA = v
n An
RA −Rn
which is usually a smooth function over space within the solid. For a solid con-
taining atoms of type J = B, C, … it can be written in terms of the number den-
J
sities ρsld R = n Jδ R −Rn of the solid as
LR LR J
Vads RA = d 3 R v
J AJ
RA − R ρsld R
Introducing the Fourier transform of the LR pair potential and its inverse
transform,
LR −3 2 LR
v AJ k = 2π d 3 R vAJ R exp ik R ,
LR −3 2 LR
vAJ R = 2π d 3 k v AJ k exp − ik R
4.5 Interaction Energies in a Liquid Solution and in Physisorption 141
we find
LR −3 2 LR
Vads RA = 2π d 3 k exp −ik RA v
J AJ
k n J
exp ik Rn
Further considering that the transform of the thermally averaged number den-
J −3 2
sity of the solid is ρ sld k = 2π n J exp ik Rn we get for the ther-
th
mally averaged LR part,
LR LR J
V ads RA = d 3 k exp − ik RA v
J AJ
k ρ sld k
Special cases follow when the atoms form a crystalline lattice, and when one can
assume that on the average the solid has a constant density which vanishes past
the surface.
The treatment simplifies when all atoms in the solid are the same, with vAJ = vA.
The total energy operator for a solid in a volume Ω is then
Vads RA = Ωd
3
R vA RA − R ρsld R and this can be statistically averaged over
a thermal distribution of atom locations for a solid at temperature T giving a
density ρ R; T = ρsld g R; ρsld ,β in terms of the pair distribution function of
atoms in the solid relative to the origin of coordinates, and an average potential
energy of adsorption
which can be further treated separating SR and LR terms. For the SR term one
S S
again has Vads RA ≈ n NT vA RA −Rn , while for the LR we can estimate the
potential energy on the adsorbate by assuming that the solid is crystalline, or
homogeneous with a smooth surface at distance D from the adsorbed atom.
Consider the case where the solid surface is on a (xy)-plane with atoms
labeled by integers n = (nx, ny, nz) in a cubic crystalline lattice generated by
translations of a unit cell with axes along orthogonal unit vectors ux ,uy ,uz
and edges of length ax, ay, az, with one atom per cell at locations Xn = X0 + nxax
and Yn = Y0 + nyay with − ∞ < nξ < ∞. The locations along the z-direction
are Zn = Z0 − nzaz with 0 ≤ nz < ∞. The density is now a lattice periodic
s
function ρsld R = ρlatt R = ρlatt R + t n , with a surface periodic translation
s
t n = nx axux + ny ayuy . The condition that the density must be invariant under
translations along x- and y-directions by steps ax and ay means that now an
expansion in related wavevector components must only contain values
142 4 Model Potential Functions
x
kx = 2π mx ax = gmx and similarly for ky, with integers − ∞ < mξ < ∞ [31]. The
density is given by
s +∞
ρlatt R = m
exp ig ms R dk z exp ikz Z ρ latt g ms ,kz
−∞
s s
where g m = 2π mxux ax + myuy ay and R = xux + yuy are surface vectors. The
density ρ latt in the reciprocal space of k vectors can be extracted by taking its
inverse transform and using it in the adsorbate expression
L s L
Vads RA = m
exp ig ms RA dkz v A g ms ,kz ρ latt g ms ,kz exp −ikz ZA
s s
with the lattice density obtained for R = t n as
−3 2 −3 2
ρ latt g ms ,kz = 2π n
exp ig ms t ns exp ikz Zn = 2π n
exp ikz Zn
The integral and summations over lattice points can be evaluated given the
Fourier transform of the pair interaction potential.
For the solid treated as a homogeneous structure, and a pair interaction
vA R = − CA R − s between the adsorbed atom and an atom within a solid ele-
L s
∞ ∞ s
s s r 2πρsld CA
V A D = − 2πρsld CA dr dzA =−
0 D r 2 + zA2
s 2 s− 2 s −3 D s−3
This gives, for the van der Waals dispersion potential with s = 6, the well-
known D−3 dependence on the distance from the adsorbate to the surface.
This result must be modified when the adsorbate is surrounded by a medium
(such as a liquid) outside the solid, to account for the dielectric constant of that
medium [36]. Here, we can assume that the dielectric effects have been incor-
s
porated into the constant CA .
4.6 Interaction Energies in Large Molecules and in Chemisorption 143
0 0 0
VM = ΨM H M ΨM =V B
+V NB
with the first term a function of internal bond variables describing bonding frag-
ments, and with the second term a function of atomic group or fragment posi-
tions for interacting nonbonding fragments. In detail, the bonding potential
energy can be given in terms of the bond distances dab between atoms a and
b, bending angle ϑabc, torsion angle γ ab, cd between two planes with a common
edge, and buckling angle γ a, bcd between a plane and a line to a vertex. Provided
displacements from equilibrium values are small, changes in lengths or angles
can be used as variables, and deformation energies can be expanded in power
series of displacements from equilibrium. We write
B
V = Vstr + Vbnd + Vtor + Vbck
where Vstr is the stretch component involving sums over two atoms, Vbnd is the
bending components for sets of three atoms, Vtor the torsion component with
sums over four atoms, and Vbck the buckling component also for four atoms in
146 4 Model Potential Functions
1 eq
2
Vstr = a<b2
Kab dab − dab
1 eq
2
Vbnd = K
a < b < c 2 abc
ϑabc − ϑabc
1 eq
Vtor = a<b<c<d 2
Kab, cd 1+ cos n γ ab, cd − δab, cd
1 eq
2
Vbck = a<b<c<d 2
Ka, bcd γ a, bcd − γ a, bcd
where the summations involve only neighbor atoms. To this one must add
energy terms for hydrogen bonding, if present, taken to involve atoms
a, b (for H), and c, and expressed as a function of a variable Sac =
d cb −d ab = dab dbc cos ϑabc of the form
1 2
vaHc = KH Sac −Saceq + KH Sac − Saceq
2
where εr is the dielectric constant of the medium. The Coulomb and induction
terms have been given as sums of pair interactions between point charges and
polarizabilities located at fragments and at the relative distance Rij. They
can also be cut-off at short distances using as shown suitable damping functions
fd(Rij), previously introduced. The present force fields V (B) + V (NB) and similar
variations have been extensively parametrized on the basis of electronic struc-
ture calculations and empirical information [40–44]. The energy landscapes
generated by these force fields can be shown as energy isocontours versus
two bond variables, keeping the other ones constant. Figure 4.8 shows the def-
inition of torsion angles φ and ψ for a polypeptide strand, and potential energy
isocontours versus those angles.
(a) (b)
180°
–180°
–180° 𝜙 180°
Figure 4.8 (a) Planes showing the definition of torsion angles ψ and φ between two peptide
units in a polypeptide. (b) Potential energy isocontours for the polypeptide glycyl residue energy
versus the torsion angles phi around the C─N bond and psi around the C─C bond
(a Ramachandran diagram). Ovals lead to energy minima. Source: from Hovmoeller [45] and [46].
148 4 Model Potential Functions
More accurately, the LR terms must be obtained from the perturbed wave-
0
function ΨM instead of ΨM , to account for interfragment interactions, and must
also include electric multipole interactions between permanent and induced
charges, which arise from intramolecular and intermolecular polarization of
atoms and bonds. For the isolated molecule M, the resulting more accurate
interactions are so-called polarizable force fields [47–51]. In addition, the total
potential energy for the M–S liquid solution must account for the interaction of
the molecule M with solvent molecules.
The M–S interaction requires knowledge of the wavefunction of M and cal-
Coul pol rep
culation of the expectation value ΨM H MS + H MS + H MS ΨM , where
the Hamiltonian operators depend on the locations and charge distribution of
each solvent species interacting with M. These interactions perturb M and
involve repulsion, polarization, and Coulomb (electrostatic) terms as shown.
To first order in the M–S interaction potentials and assuming that the solvent
species have unchanged charge distributions, the expectation value can be
0
obtained from the wavefunction ΨM of the isolated M with its own permanent
and induced charge distribution. Higher-order terms in a perturbation expan-
0 1 2
sion ΨM = ΨM + ΨM + ΨM + … involve excited states of M and describe solute
polarization, which include additional induction and dispersion terms, to first
and second order in the M–S interaction. Further changes in potential energies
arise from polarization of the solvent species.
Thermal equilibrium properties of the molecular system can be obtained
from internal energies derived from the above potential energies plus the system
kinetic energy at a given temperature. More detailed properties, such as isomer-
ization rates, follow from molecular dynamics calculations involving force fields
given by gradients of the potential functions with respect to atomic displace-
ments [52, 53].
V Q = E emb
a a
G Q
The global variables can be chosen so that fittings to known values are phys-
ically invariant under collective translation and rotation of the rigid many-
atom system, and can also include environmental symmetries such as they
arise from the presence of identical atoms there. For example global (symme-
try) variables X1, … , X8 for the H2/Pd(100) adsorbate have been chosen as
functions of the six diatomic position variables (three for the center-of-mass
of H2 and three for its distance to the surface and the diatomic orientation
angle) with X4, X5, X7, X8 taken as Fourier spatial components oscillating with
the reciprocal lattice vector periodicity of the substrate surface [59]. Symmetry
adapted global variables have also been constructed using permutational sym-
metry of identical atoms in treatments of reaction PES for H + H2 and Cl +
H2 [60].
More generally, global symmetry functions for each embedded atom a have
been defined in terms of the position coordinates of neighbor atoms within a
rad ang
finite sphere of chosen radius Rc. They are radial Ga and angle Ga variables
dependent on two or three atom positions and given by [61]
sph
Garad = j a
exp − ηR2aj fc Raj
sph
Gaang = 21− ζ j, k a
1+ cosθajk exp − η R2aj + R2ak + R2jk fc Raj fc Rak fc Rjk
with the sphere cut-off function fc(R) = [1 + cos(πR/Rc)]/2 for R ≤ Rc and zero
past Rc, angle θajk between the bonds aj and ak, and parameters η and ζ to
be found.
150 4 Model Potential Functions
xa1, m = bm1 + μ
Gaμ wμ01
,m
01
Here, the weights wμ, m are parameters to be found, connecting the global var-
iable in the input layer to node m in the hidden layer. This partial energy com-
ponent is accepted with an activation (or acceptance) probability biased by the
1
magnitude of bm . The bias parameter is introduced to allow for an acceptance
somewhere between zero and one for the output of node m, given by a sigmoid-
1 1 1
shape (activation) function f xa, m = ya, m varying from zero at minus infinity
1
to one at plus infinity, with bm setting its value in the slope region and its
acceptance probability. This function can be chosen, for example as [1 + tanh
(x)]/2.
When only one hidden layer is used in the NN, the final stage constructs the
total energy of each embedded atom with
Eaemb = y1
m a, m
Gaμ ; bm1 , wμ01
,m
as a sum of the outputs from all the nodes of the hidden layer. The parameters
1 01
bm and wμ, m in the acceptance function follow from a NN training that opti-
emb
mizes the fitting of a set of reference PES values Ea, ref for given input atomic
coordinates, using usually the mean absolute deviation (MAD)
1 emb emb
MAD = Ea, ref −Ea, NN
Nat a
Bias
Forces Fa, ξ at atom a along the ξ-direction in the many-atom structure can be
obtained as needed in molecular dynamics calculations from
emb
∂V N M ∂ Ea ∂Gα
Fa, ξ = − =−
∂Ra, ξ a=1 α = 1 ∂G
α ∂R a, ξ
where N is the number of atoms and M the number of global symmetry func-
tions. The fraction in the first factor is given by the NN architecture and the
second factor follows from the choice of the global variables.
The procedure given here for interpolation of PESs allows for accurate fit-
tings, improvement if additional calculated points are available for training of
the NN, and is quite general for ground electronic states. It allows for descrip-
tion of bonding and nonbonding (weak or van der Waals) interactions and for
systems undergoing reactions with broken and formed bonds. And it can pro-
vide results with the accuracy of the training set (such as DFT or CCSD) at the
same computational cost as force-field calculations. The AI-NN procedure,
however, requires preliminary calculations of PESs with standard treatments
to be used for training and validation so that the AI-NN procedure is comple-
mentary to standard treatments of intermolecular forces using molecular prop-
erties or many-electron methods.
Given calculated PESs for a large class of compounds and interacting pairs of
molecules, an AI-NN procedure can also be used to calculate new compounds
and their interaction, provided they contain the same embedded atoms [62, 63].
This however must be done with caution and extensive validation, and some of
these aspects are treated in Chapter 7.
References
1 Hirschfelder, J.O., Curtis, C.F., and Byron Bird, R. (1954). Molecular Theory of
Gases and Liquids. New York: Wiley.
2 Margenau, H. and Kestner, N.R. (1971). Intermolecular Forces, 2e. Oxford,
England: Pergamon Press.
3 Murrell, J.N., Carter, S., Farantos, S.C. et al. (1984). Molecular Potential Energy
Functions. New York: Wiley.
4 Schatz, G.C. (1989). The analytical representation of electronic potential energy
surfaces. Rev. Mod. Phys. 61: 669.
5 Stone, A.J. (2013). The Theory of Intermolecular Forces, 2e. Oxford, England:
Oxford University Press.
6 Kaplan, I.G. (2006). Intermolecular Interactions, 81. New York: Wiley.
7 Herzberg, G. (1950). Molecular Spectra and Molecular Structure I. Spectra of
Diatomic Molecules, 2e. Princeton NJ: Van Nostrand.
8 Lawley, K.P. (ed.) (1975). Molecular Scattering (Advan. Chem. Phys. vol. 30).
New York: Wiley.
References 153
9 Maitland, G.C., Rigby, M., Brain Smith, E., and Wakeham, W.A. (1981).
Intermolecular Forces. Oxford, England: Oxford University Press.
10 Jensen, F. (1999). Introduction to Computational Chemistry. New York: Wiley.
11 Karplus, M. and Porter, R.N. (1970). Atoms and Molecules. New York: W. A.
Benjamin.
12 Kramer, H.L. and Herschbach, D.R. (1970). Combination rules for van der Waals
force constants. J. Chem. Phys. 53: 2792.
13 Smith, F.T. (1972). Atomic distortion and the combining rule for repulsive
potentials. Phys. Rev. A 5: 1708.
14 Koide, A., Meath, W.J., and Alltnatt, A.R. (1981). Second order charge overlap
effects and damping functions for isotropic atomic and molecular inetractions.
Chem. Phys. 58: 105.
15 Tang, K.T. and Toennies, J.P. (1984). An improved simple model of the van der
Waals potential based on universal damping functions for dispersion
coefficients. J. Chem. Phys. 80: 3726.
16 Siska, P.S., Parson, J.H., Schafer, T.P., and Lee, Y.T. (1971). Intermolecular
potentials from crossed beam differential elastic scattering measurements III. He
+He and Ne+Ne. J. Chem. Phys. 55: 5762.
17 Buck, U. (1975). Elastic scattering. Adv. Chem. Phys. 30: 313.
18 Smith, I.W.M. (1980). Kinetics and Dynamics of Elementary Gas Reactions.
London, Great Britain: Butterworth.
19 Hirst, D.M. (1985). Potential Energy Surfaces. London, England: Taylor & Francis.
20 Gordon, M.D. and Secrest, D. (1970). Helium-atom-hydrogen-molecule
potential energy surface employing the LCAO-MO-SCF and CI methods.
J. Chem. Phys. 52: 120.
21 Edmonds, A.R. (1960). Angular Momentum in Quantum Mechanics. Princeton
NJ, USA: Princeton University Press.
22 Levine, I.N. (2000). Quantum Chemistry, 5e. New Jersey: Prentice-Hall.
23 Mahoney, M.W. and Jorgensen, W.L. (2000). A five-site model for liquid water
and the reproduction of the density anomaly by rigid, non-polarizable potential
functions. J. Chem. Phys. 112: 8910.
24 Karplus, M. (1970). Potential energy surfaces. In: Molecular Beams and Reaction
Kinetics (ed. C. Schlier), 320. New York: Academic Press.
25 Sato, S. (1955). Potential energy surface of the system of three atoms. J. Chem.
Phys. 23: 2465.
26 Truhlar, D.G., Steckler, R., and Gordon, M.S. (1987). Potential energy surfaces
for polyatomic reaction dynamics. Chem. Rev. 87: 217.
27 Polanyi, J.C. (1972). Some concepts in reaction dynamics. Acc. Chem. Res. 5: 161.
28 Bender, C.F., Pearson, P.K., O’Neil, S.V., and Schaeffer, H.F. III (1972). Potential
energy surface including electron correlation for F + H2 FH + H. I.
Preliminary surface. J. Chem. Phys. 56: 4626.
29 Bender, C.F., O’Neil, S.V., Pearson, P.K., and Schaeffer, H.F. III (1972). Potential
energy surface including electron correlation for F + H2 FH + H: refined linear
surface. Science 176: 1412.
154 4 Model Potential Functions
30 Micha, D.A. (1985). General theory of reactive scattering. In: Theory of Chemical
Reaction Dynamics, vol. II (ed. M. Baer), 181. Boca Raton, USA: CRC Press.
31 Kittel, C. (2005). Introduction to Solid State Physics, 8e. Hoboken, NJ: Wiley.
32 Tersoff, J. (1986). New empirical model for the structural properties of silicon.
Phys. Rev. Lett. 56: 632.
33 McQuarrie, D.A. (2000). Statistical Mechanics. Sausalito CA: University
Science Books.
34 Hansen, J.P. and McDonald, I.R. (1986). Theory of Simple Liquids, 2e. Orlando
FL, USA: Academic Press.
35 Gordon, M.S., Fedorov, D.G., Pruitt, S.R., and Slipchenko, L.V. (2011).
Fragmentation methods: a route to accurate calculations on large systems.
Chem. Rev. 112: 632.
36 Israelachvili, J. (1992). Intermolecular and Surface Forces. San Diego CA:
Academic Press.
37 Jeziorski, B., Moszynski, R., and Szalewicz, K. (1994). Perturbation theory
approach to intermolecular potential energy surfaces of van der Waals
complexes. Chem. Rev. 94: 1887.
38 Friesner, R.A. and Guallar, V. (2005). Ab initio quantum chemical and mixed
quantum mechanics/molecular mechanics (QM/MM) methods for enzymatic
catalysis. Annu. Rev. Phys. Chem. 56: 389.
39 Albaugh, A., Boateng, H.A., Bradshaw, R.T. et al. (2016). Advanced potential
energy surfaces for molecular simulation. J. Phys. Chem. B (Feature Article)
120: 9811.
40 Case, D.A., Cheatham, T.E.I., Darden, T. et al. (2005). The AMBER biomolecular
simulation programs. J. Comput. Chem. 26: 1668.
41 Brooks, B., Brooks, C.L., Mackerell, A.D. et al. (2009). CHARMM: the
biomolecular simulation program. J. Comput. Chem. 30: 1545.
42 Arnautova, Y.A., Jagielska, A., and Scheraga, H.A. (2006). A new force field
(ECEPP-05) for peptides, proteins, and organic molecules. J. Phys. Chem. B
110: 5025.
43 Van Der Spoel, D., Lindahl, E., Hess, B. et al. (2005). GROMACS: fast, flexible,
and free. J. Comput. Chem. 26: 1701.
44 Allinger, N.L., Chen, K., and Lii, J.-H. (1996). An improved force field (MM4) for
saturated hydrocarbonds. J. Comput. Chem. 17: 642.
45 Hovmöller, S., Zhou, T., and Ohlson, T. (2002). Conformations of amino acids in
proteins. Acta Cryst D58: 768.
46 Atkins, P. and De Paula, J. (2010). Physical Chemistry, 9e. New York: W. H.
Freeman and Co.
47 Kaminski, G.A., Stern, H.A., Berne, B.J. et al. (2002). Development of a
polarizable force field for proteins via ab initio quantum chemistry: first
generation model and gas phase test. J. Comput. Chem. 16: 1515.
48 Warshel, A., Kato, M., and Pisliakov, A.V. (2007). Polarizable force fields: history,
test cases, and prospects. J. Chem. Theory Comput. 3: 2034.
References 155
49 Shi, Y., Xia, Z., Zhang, J. et al. (2013). Polarizable atomic multipole based
AMOEBA force field for proteins. J. Chem. Theory Comput. 9: 4046.
50 Baker, C.M. (2015). Polarizable force fields for molecular dynamics simulations
of biomolecules. WIREs Comput. Mol. Sci 5: 241.
51 Xie, W., Song, L., Truhlar, D.G., and Gao, J. (2008). The variational explicit
polarization potential and analytical first derivative of energy: towards a next
generation force field. J. Chem. Phys. 128: 234108.
52 Brooks, C.L.I., Karplus, M., and Montgomery-Pettitt, B. (1988). Proteins:
A Theoretical Perspective of Dynamics, Structure, and Thermodynamics. New
York: Wiley-Interscience.
53 Warshel, A. (1991). Computer Modeling of Chemical Reactions in Enzymes and
Solutions. New York: Wiley.
54 Blank, T.B., Brown, S.D., Calhoun, A.W., and Doren, D.J. (1995). Neural network
models of potential energy surfaces. J. Chem. Phys. 103: 4129.
55 Brown, D.F.R., Gibbs, M.N., and Clary, D.C. (1996). Combining ab initio
computations, neural networks, and diffusion Monte carlo: an efficient method
to treat weakly bound molecules. J. Chem. Phys. 105: 7597.
56 Raff, L.M., Malshe, M., Hagan, M. et al. (2005). Ab initio potential energy
surfaces for complex, multichannel systems using modified novelty sampling
and feedforward neural networks. J. Chem. Phys. 122: 084104–084101.
57 Behler, J. (2011). Neural network potential energy surfaces in chemistry: a tool
for large scale simulations. Phys. Chem. Chem. Phys. 13: 17930.
58 Behler, J. (2016). Perspective: machine learning potentials for atomistic
simulations. J. Chem. Phys. 145: 170901–170901.
59 Lorenz, S., Scheffler, M., and Gross, A. (2006). Description of surface chemical
reactions using a neural network representation of the potential-energy surface.
Phys. Rev. B 73: 115431.
60 Jiang, B. and Guo, H. (2013). Permutation invariant polynomial neural network
approach to fitting potential energy surfaces. J. Chem. Phys. 139: 054112.
61 Behler, J. and Parrinello, M. (2007). Generalized neural network representation
of high-dimensional potential-energy surfaces. Phys. Rev. Lett. 98: 146401–
146401.
62 Smith, J.S., Isayev, O., and Roitberg, A.E. (2017). ANI-1: an extensible neural
network potential with DFT accuracy at force-field computational cost. Chem.
Sci. 8: 3192.
63 Ramakrishnan, R., Dral, P.O., Rupp, M., and von Lilienfeld, O.A. (2015). Big data
meets quantum chemistry approximations: the delta-machine learning
approach. J. Chem. Theory Comput. 11: 2087.
157
Intermolecular States
CONTENTS
maRa + mbRb
RCMN =
ma + mb
R =Rb −Ra
and we can translate the reference frame keeping axes parallel so that in a new ref-
erence frame CMN called the center-of-mass of the nuclei, space-fixed (CMN-
SF) frame, we have RCMN = 0 and the (x, y, z) axes pointing along the previous
unit vectors. A further convenient transformation, when we describe the prop-
erties of a molecule in an equilibrium conformation, is to reorient the axes to
coincide with the principal axes of inertia of the nuclear framework, along new
unit vectors n x , ny , nz . This can be done specifying three Euler angles (α, β, γ) = Γ
of the new frame called now the center-of-mass of the nuclei, body-fixed (CMN-
BF) frame, designated as CMN . New coordinates (x , y , z ) are related to the old
ones by a 3 × 3 rotation matrix A α,β,γ = Rz γ Ry β Rz α , a function of the
Euler angles constructed from three axial rotations [1]. The three mentioned
reference systems are shown in Figure 5.1.
In the case of intermolecular potentials, one usually finds two molecules A
and B, each with its own CMN-BF reference frame I , I = A, B, with origins
at their centers of mass and oriented by Euler angles ΓI with respect to an overall
CMN-SF frame located along the line joining the centers of mass of A and B.
This is shown in Figure 5.2.
The remaining nuclear coordinates qI are internal ones, such as bond dis-
tances, bond angles, and torsion angles. Therefore, the whole system of Nnu
nuclei is described by the set of 3Nnu variables Q = R, ΓA ,qA , ΓB ,qB .
5.1 Molecular Energies for Fixed Nuclear Positions 159
→
Ra
𝒮L
𝒮L
2
Pr = ζi
dX i
Φ X; Q r i =r
and note that the Coulomb interaction between nuclei and electrons
1 Nnu Cα Ce
H ne = v ri , v ri = α=1
i
4πε0 Rα −r i
has an expectation value
H ne = d 3 r ρ r v r
5.1 Molecular Energies for Fixed Nuclear Positions 161
E Q = Φ Q HQ Φ Q < Φ Q HQ Φ Q
E Q < ΦQ H Q − H nu + H nu ΦQ
Reversing primed and unprimed quantities, the last three terms only change
sign and when we add the two expressions they cancel and we find
E Q +E Q <E Q +E Q
which is a contradiction resulting from the assumption that two different
energies could arise from the same density. This establishes the one-to-one
correspondence between the energy functional and the density. As a corollary,
we find that if E ρ r; Q is the ground state energy functional of the electronic
density for a given nuclear conformation, then minimizing this functional leads
to an upper bound to the exact ground state energy, which is yet a function of
the nuclear position variables.
The potential energy function for species A and B at relative position R is
obtained as
VAB Q = E ρ r; Q −E ρ r; Q R ∞
makes use of a total density written as a sum of one-electron terms, and the
Gordon–Kim procedure [5], giving the total density as a sum of the densities
of molecular components of the whole system. Many details of the DFT
treatments can be found in reference [6].
E ρ = d3 r ρ r v r + F ρ + Hnn
Fρ = ΦQ K e + H ee ΦQ ρ
= VCoul + G ρ
1 3 3 Ce 2
VCoul = d rd r ρ r ρ r −δ r − r ρ r
2 4πε0 r −r
G ρ = Gkin ρ + Gxc ρ
where VCoul is the classical Coulomb energy of interaction among all the elec-
trons, a Hartree-like (purely electrostatic) energy, with the second term in VCoul
introduced to prevent electron–electron self-interactions. The two terms in G[ρ]
come from the electronic kinetic energy, and combined electronic exchange and
correlation energies, with the second term containing the difference between the
exact energy of the electron system and the Hartree energy. Once this functional is
constructed, one can proceed to minimize it for variations of the electron density
that conserve the number of electrons so that d 3 r ρ r = Nel . The variation of the
constrained functional E ρ − μ d 3 r ρ r , where μ has the meaning of a chemical
potential at absolute zero temperature, leads to the equation
δE
−μ = 0
δρ r
The minimum for E[ρ] provides an upper bound to the ground state elec-
tronic energy and the optimal density ρ r; μ .
In the Kohn–Sham (K–S) procedure, the density is written as a sum of terms
containing the spin-orbitals ψ j from each electron, for spin up and down vari-
ables ζ = (+, −),
Nel 2
−
ρr = ζ j=1
ψ j r, ζ =ρ +
r +ρ r
Φ KS X; Q = C j
ψ j r j , ζj
Ks ρ = Φ KS Q K e Φ KS Q
ρ
F ρ = VCoul ρ + Ks ρ + Exc ρ
The last term is an exchange-correlation functional (which also includes a
contribution to correct the kinetic energy functional). This is parametrized
and used in calculations. Improvements follow using separate functionals for
up and down spins, needed to describe bonding between open shell species,
and using functionals of both the density and its gradient, to account for inho-
mogeneous charge distributions.
The K–S spin-orbitals are obtained solving a one-electron eigenvalue
equation
ℏ2 2
− ∇ + v KS r,ζ ψ j r,ζ = ϵj ψ j r,ζ
2me
Consider two nuclei a and b on the z-axis, with relative position R, surrounded
by electrons in a state Φk with energy Ek. This can be a ground or excited
electronic state, and here we assume it is noninteracting with other states.
We find the force on nucleus a by differentiation of the expectation value of
the energy operator with respect to its position coordinate Za. We begin with
the Schrödinger equation and
∂ ∗
dX Φk X;R H R − Ek R Φk X;R = 0
∂Za
Expanding this expression, displaying the Coulomb terms in the Schrödinger
equation and identifying the force on a, we have with r 1a =r 1 −Ra ,
∂Ek Ca Cb Zab Ce ρ r 1 Ca z1a
Fa, z = − =− + d 3 r1
∂Za 4πε0 R3 3
4πε0 r1a
This is the electrostatic force theorem, formally identical to the corresponding
classical mechanical expression, but now constructed from a quantal electronic
density. The first term to the right is negative and shows the repulsive force on a
from the nucleus b; the integrand in the second term is the Coulomb force
between an electronic distribution at position r 1 and nucleus a, with positive
or negative integrand values depending on the location of the element of volume
d3r1 along the z-direction. If the element of volume of electrons is found
between two parallel planes perpendicular to the axis through the nuclei, then
it gives a positive force increment contributing to bonding. Far outside these
planes, the volume will oppose bonding. The first region is called the bonding
region, the second one is the antibonding region. They are separated by a sur-
face where the electronic force densities in Fμ, z = d 3 rfμ, z r at each nucleus μ =
a, b are equal, or
fa, z r = fb, z r
Φ R, λ = Φ R, λ H R Φ R, λ Φ R, λ Φ R, λ
optimized to obtain a minimal energy when the parameters are varied and cal-
culated at the minimum for each distance R. This gives an energy Emin R =
Φ R,λmin from which the force at nucleus a follows from the gradient as
Φ CI x1 , x2 = J
AJ ΦJ x1 , x2
with all terms of the same space and spin symmetry. The energy follows from
the variational functional Φ CI = E ζ μ ,cμj ,AJ and forces are given by
with λ = {ζ μ, cμj, AJ} the set of variational parameters. This expression shows how
the forces are affected by changing AO exponents, MO coefficients, and coeffi-
cients of CIs when these parameters have not been optimized by minimization
of the energy. If instead they have been optimized, then the last three summa-
tions are zero and the force is given by the gradient of the energy calculated at
the optimized parameter values.
∂E ∂H
= Ψλ Ψλ
∂λ ∂λ
kin Coul
We consider a special case where the Hamiltonian is H el = H el + H el , a
sum of electronic kinetic and Coulomb electron–electron and electron–nucleus
Coul
energies in H el = H ee + H en , for fixed nuclear positions and excluding the
Coulomb interaction among nuclei. The corresponding electronic eigenstate
is Φel(R) with eigenenergy Eel(R) for a relative position vector R = Rnz along
the z-axis. A parameter λ is introduced to scale the position coordinates of elec-
trons, but not of the nuclei, writing λr j =sj for electron j and changing integra-
tion variables in expectation values to the set sj . The electronic energy
kin Coul
becomes Eel λ = λ2 H el R + λ H el λR R and differentiation with respect
to the parameter λ, using that
∂ 1 ∂ 1 ∂ λR
=
∂λ s− λRnz ∂ λR s−λRnz ∂λ
in the expectation value for the Coulomb Hamiltonian, gives for λ = 1 a virial
theorem valid for any fixed internuclear distance,
kin Coul
2 H el + H el R = − R∂Eel ∂R
R R
where the Hellmann–Feynman theorem has been used for the right-hand side.
kin Coul
This can be combined with H el R + H el R R = Eel R to obtain expecta-
tion values of separate electronic kinetic and Coulomb energies,
kin
H el = −R∂Eel ∂R −Eel R
R
Coul
H el R = 2Eel R + R∂Eel ∂R
R
showing how they depend on the internuclear distance. At equilibrium for the
minimum of the electronic energy, where ∂Eel/∂R = 0, one recovers the virial
equality for Coulomb interactions,
kin Coul
2 H el = − H el Rmin
min
relating the electronic kinetic and Coulomb energies for equilibrium nuclear
positions. The expectation values for kinetic and Coulomb energies obtained
from the virial relations, in the case of variable distances, are more accurate than
their values when calculated from variationally optimized state averages. This
follows because the errors in the wavefunction obtained in a variational calcu-
lation are of first order with respect to the exact wavefunction, but errors are
168 5 Intermolecular States
only of second order for the optimized variational energy that appears in the
virial relations.
The derivation for a diatomic system can easily be generalized to a polyatomic
system where atoms are located at positions Ra , a = 1 to N, in which case the
virial relation is
kin Coul
2 H el + H el R =− R
a a
∂Eel ∂Ra
R R
with R = Ra the set of all atomic positions, and with electronic energy gradi-
ents calculated for each atomic displacement.
When the exact wavefunction is not known (as usual), then it is necessary
to instead differentiate the previously introduced functional Φ R, λ .
Assuming that the approximate wavefunction has been normalized to
Φ R,λ Φ R, λ = 1, one finds that
∂ ∂H ∂Φ ∂Φ
= Φλ Φλ + H λ Φλ + Φλ H λ
∂λ ∂λ ∂λ ∂λ
which can again be analyzed in terms of scaled kinetic and Coulomb compo-
nents but with results depending now on how the states Φ depend on the
parameter λ by construction.
for the energy change due to the interaction of A and B, with E(1) = E and
E(0) = E0.
However, when the state Ψ(λ) is only approximately known as Φ(λ) from a
variational or perturbative procedure, the relevant derivative ∂ ∂λ contains
additional terms and the integral form becomes
1
∂Φ ∂Φ
Δ = dλ Φ λ H AB Φ λ + H λ Φλ + Φλ H λ
0 ∂λ ∂λ
5.3 Molecular States for Moving Nuclei 169
HΨ = EΨ
where H, E, and Ψ are the Hamiltonian operator, energy, and state of the whole
system of electrons and nuclei. The state Ψ must be normalized to describe a
bound system, or must satisfy asymptotic boundary conditions for an unbound
system at large R.
The Hamiltonian operator can be separated into the kinetic energy operator
of the relative nuclear motion plus the Hamiltonian H R for fixed nuclei, with the
latter including the electronic kinetic energy, the Coulomb attraction of elec-
trons to nuclei, the Coulomb repulsion among electrons and the repulsion
among the nuclei, as
ℏ2 2
H=− ∇ + HR
2M R
H R = K e + H ee + H en + H nn
170 5 Intermolecular States
If some of the electrons are found to form rigid ion cores, then the electron-
nucleus interaction must be replaced by an electron-core interaction which is
Coulombic at large distance and repulsive at short distance due to electronic
exclusion; similarly the nucleus–nucleus interaction must be replaced by the
core–core interaction.
Many-electron states of molecules can be constructed from many-electron
atomic states, in an approach known as the valence-bond (VB) method, or from
atomic orbitals combined to form molecular orbitals, in the so-called molecular
orbital (MO) method, as we have seen with the example of H2 in the previous
Chapter 3.
The molecular wavefunction Ψ can be expanded in a basis of orthonormal
electronic states Φk which are functions of the electronic variables and depend
parametrically on the nuclear positions. Using the Dirac notation for electronic
states we write Φk X;R = X Φk R and work in a linear space of K states,
introducing an orthonormal and complete basis set of such states for fixed
nuclei, the expansion is
K
Ψ R,X = k =1
Φk X;R ψ k R
Φk R Φl R = δkl
K
k =1
Φk R Φk R =I
The coefficients ψ k are functions of the nuclear positions that describe the
nuclear motions in the electronic states k. The bracket notation here and in
what follows signifies integration and sum over all the electron variables, for
fixed nuclei as shown. The last line gives an assumed completeness relation,
with I the identity operator.
Several choices are possible for the basis set of electronic states, which lead to
different electronic representations of potential energies and nuclear ampli-
tudes. These representations are related by matrix transformations which can
be conveniently introduced to simplify treatments of molecular spectroscopy
and dynamics [15–19].
Ψ R = Φ R ψ R
where the first factor is a 1 × K row of states Φk and the second factor is a K × 1
column of amplitudes ψ k. The Schrödinger equation can now be put in a matrix
5.3 Molecular States for Moving Nuclei 171
where the second power of the square bracket signifies a scalar product of matri-
ces with vector-valued elements. This is a set of coupled differential equations
that must be solved with the appropriate boundary conditions for the ψ ampli-
tudes, for bound or unbound (decaying or scattering) states.
It is of interest to analyze the vector product of the total momentum appear-
ing in this equation. Expanding it,
2
ℏ ℏ
I ∇R + G R ψ R = −ℏ2 ∇R 2 + 2G ∇R + ∇R G + G G ψ R
i i
172 5 Intermolecular States
we find that the two terms in the square bracket to the right add up to a Her-
mitian operator but each of them is, separately, non-Hermitian. Hence, it is safer
to keep them together as a sum.
with help of the relation f ∇R2g = ∇R. (f ∇Rg) − (∇Rf ). ∇Rg and cancelling oppo-
site terms, we find the equation for the divergence of the intermolecular flux
matrix J R , with vector valued elements J kl , as
∇R J R = 0
1ℏ † 1
J R = ψ ∇R ψ − ∇R ψ † ψ + ψ † Gψ
Mi M
showing that probability is conserved, provided however that one must now
include a contribution from the electronic momentum fluctuation in the defi-
nition of the flux.
and the nuclear amplitudes satisfy corresponding matrix equations with the
momentum and Hamiltonian matrices G a , HR calculated in the adiabatic
a
C †k Δ C l = δkl
ℏ †
C l + C †k Δ ∇R C l
a
G kl = C Dk R ∇R Dl R
i k
where the first term describes the momentum coupling contributed by the dis-
placement of the basis functions, and the second term gives the contribution by
the changes of coefficients with positions. The gradients of coefficients in the
second term can be obtained by differentiation of the secular equation for
the coefficients, as
−1
a a
∇R C l = H R − El I ∇R H R − El I Cl
The equation of motion for the nuclear amplitudes in the adiabatic represen-
tation, with a diagonal matrix of adiabatic energy potentials E a R =
a
Ek R δkl , is
2
1 ℏ
I ∇R + G a
R +E a
R − IE ψ a
R =0
2M i
which shows that the amplitudes are interacting only through the momentum
couplings. Ways to generate nonadiabatic representations have been given in
the literature [12, 13, 15, 16, 20].
d d
Φk ∇R Φl = δkl
Φd R = Φa R A R;R0
with A R0 ;R0 = I and the unitarity condition A† = A−1. From the condition
defining the diabatic basis set, we have the differential equation
ℏ
∇R A R;R0 + G a
R A R;R0 = 0
i
which can be solved to obtain the transformation matrix from given adiabatic
momentum couplings. The total molecular wavefunction in the new basis set is
M d d
Ψ R,X = k =1
Φk X;R ψ k R = X Φd R ψ d
R
cos γ sin γ
A R; R0 =
−sin γ cos γ
It follows from the differential equation for A(R; R0) that the angle γ(R; R0) is a
function satisfying the differential equation dγ/dR = τ(a)(R) with the boundary
condition γ(R0; R0) = 0. Its solution
R
γ R; R0 = dR τ a R
R0
can be used to construct the Hamiltonian matrix elements in the strictly dia-
d
batic representation, where Gjk = 0 for all elements, as
d a a
H11 R = E1 R cos2 γ + E2 R sin2 γ ,
d a a
H22 R = E1 R sin2 γ + E2 R cos2 γ ,
d a a
H12 R = E1 R − E2 R sin γ cos γ
d
the crossing point, H12 R is small but the diagonal elements do not cross, and
for even shorter distances, the diabatic elements differ significantly from the
adiabatic energies.
d
We also find that the diagonal diabatic potential energies Hjj R will cross
whenever γ = ± (n/2 + 1/4)π and that this can happen more than once on the
way to small R if the adiabatic momentum couplings are large. This is not
physically meaningful and is a consequence of the definition of the strictly
diabatic representation. An alternative more physically meaningful diabatic rep-
resentation can be introduced by parametrizing the transformation angle to a
form γ(R; α) with the parameter α chosen to minimize the momentum coupling
in a new “m” representation [18]. This generates new smooth diabatic potentials
m
Hjk R and a new smaller momentum coupling τ(m)(R; α) which can be used
for molecular spectroscopy and collision treatments in the new representation.
Figure 5.3 shows an example constructed from Morse type potentials.
In any case, the new states are related to adiabatic ones through the transfor-
mation matrix as
d a a
Φ1 R = Φ1 R cos γ − Φ2 R sin γ
d a a
Φ2 R = Φ1 R sin γ + Φ2 R cos γ
(a)
–0.11
–0.12
–0.13
V (a.u.)
–0.14
–0.15
–0.16
1.8 2.0 2.2 2.4
R (a.u.)
(b)
0.005 10.0
τa
0.004 8.0
0.003 6.0
0.002 4.0
τa (a.u.)
V (a.u.)
0.001 2.0
0.0 0.0
vd12
–0.001
vm
12
–0.002
–0.003
1.6 2.0 2.4 2.8
R (a.u.)
Figure 5.3 (a) Adiabatic (full line), diabatic (dash line) and minimum coupling (dash-dot line)
potential energies for two states and (b) their couplings in the three representations. Source:
from Ref. [18]. Reproduced with permission of John Wiley & Sons.
178 5 Intermolecular States
E (hartree)
Na H 1Σ+
162.25
Na++ H–
Na(4p) + H(1s)
0.30 Na(3d) + H(1s)
Na(4s) + H(1s)
Na(3p) + H(1s)
162.35
0.40
Na(3s) + H(1s)
162.45
2 5 10 20 50 100 R(bohr)
Figure 5.4 NaH potential energies showing multiple pairs of avoided crossings. Source: from
Ref. [23]. Reproduced with permission of John Wiley & Sons.
a a a
with Φk a real-valued normalized function for which Φk ∇R Φk = 0.
When this is replaced in the Schrödinger equation one finds
5.4 Electronic Representations 179
ℏ2 a a a
− ∇R 2 + Ek R + Wk R − E ψ k R = 0
2M
with
a ℏ2 a a ℏ2 a a
Wk R =− Φk R ∇2R Φk R = ∇R Φk R ∇ R Φk R
2M 2M
a repulsive potential energy addition, with the last equality obtained by partial
integration. Comparisons of the magnitudes of these approximations for the
energy of the ground molecular state of hydrogen molecule isotopes can be
found in [24]. There one finds that the ground state energy E0 of the X 1 Σg+
molecular term for H2, in cm–1 for E0/(hc) is 36 112.2, 36 118.0, and 36 114.7
for the fixed-nuclei, adiabatic, and nonadiabatic approximations, to be com-
pared with the experimental value of 36 113.6 cm–1. These differences are typ-
ical of corrections to equilibrium energy values. Considering that 1.0 cm–1 =
1.9864 × 10–23 J in the SI units and 1.0 cm–1 = 4.55634 × 10−6 au, it is found that
the nonadiabatic corrections are meaningful in very high precision spectros-
copy, but not essential for most work on intermolecular forces around equilib-
rium structures. The nonadiabatic couplings are however essential in many
treatments of photoinduced or collision-induced electronic rearrangement.
Another nonadiabatic approximation useful in molecular spectroscopy, for a
molecular system where internal variables remain near their equilibrium values
Qeq in the initial electronic state, is the so-called Condon approximation where a
C
basis set of electronic functions Φk X; Qeq is used in an expansion of the
molecular state. Here all the expansion functions have distances kept constant
at the equilibrium values Qeq of internal coordinates, and the variation of ener-
gies with displacements is described expanding the electronic Hamiltonian as a
function of displacements around the equilibrium conformation. This generates
a vibronic Hamiltonian displaying couplings of electronic and nuclear degrees
of freedom in molecular dynamics.
Given the set of coordinates Q = {Qν}, the total molecular wavefunction is
expanded in the basis set of Condon states as
K C C
Ψ Q, X = k =1
Φk X; Qeq ψ k Q,
C
calculation of the coefficient functions ψ k , starting with an expansion of the
C
equation for Ψ(Q, X) in the basis set Φk X; Qeq . This leads to equations
μ
containing force matrix elements Fkl multiplying the qμ displacements and cur-
μν
vature matrix elements Gkl multiplying the products qμqν.
For a simple case of a diatomic, this amounts to the expansion
K C C
Ψ R, X = k =1
Φk X;Req ψ k R
General statements can be made about the location and nature of interaction
energies at these special points. To simplify, we consider first the case with only
one internal degree of freedom, for two atoms at the internuclear distance R,
and where only two states k = 1, 2 are involved. Examples are the lowest two1Σ
states of LiF or of NaH, where ionic electronic distributions give states of lower
energy around their ground equilibrium distance, but covalent distributions
give lower energy states at larger distances.
A special point Rc of a crossing or avoided crossing is located by finding where
the difference H11(R) − H22(R) shows a minimum as R is varied. Writing R =
c
Rc + x and Hkl Rc = Hkl , matrix elements can be expanded around x = 0 as
c ∂Hkl 1 2 ∂ 2 Hkl
Hkl x = Hkl + x + x +
∂Rc 2 ∂Rc 2
c c
with H11 = H22 if there is a crossing when working with the original basis set.
c
A nondiagonal Hamiltonian matrix element H12 appears if the representation is
nonadiabatic. Near the special point, an electronically adiabatic representation
a
is generated by the linear combinations Φk = Φ1 C1k + Φ2 C2k with the coeffi-
cients obtained by diagonalizing the Hamiltonian matrix at each R so that
a a
the new elements are Hkl R = δkl Ek x . These two energies are given by
the well-known roots of the second-order equation det[H − EI] = 0 where 2 ×
2 matrices are shown in boldface, as
a 1 1 2 2 1 2
E1, 2 x = H11 x + H22 x ± H11 x −H22 x + 4 H12 x
2 2
and become equal only when the radical L(x) = [H11(x) − H22(x)]2 + 4|H12(x)|2 = 0,
or when both H11(x) − H22(x) = 0 and H12(x) = 0. These are two equations in x to
be satisfied by a single solution root x0, and there is no real valued answer, except if
in a particular system one of the two equations is already obeyed for all x, such as
when the states 1 and 2 are degenerate and H11(x) = H22(x). For two different
diagonal element functions of x and in the absence of accidental vanishing of their
a
coupling H12, the two adiabatic energies E1, 2 x are different for all R = Rc + x
around x = 0. This is the noncrossing rule for adiabatic potential energies.
In this case, the potentials show an avoided crossing and the difference
a a
between the new potential energies, ΔE x = E1 x − E2 x , has a minimum
at the gap located at xm, where ∂(ΔE)/∂x = 0, equal to
2 2 1 2
ΔEm = ΔHm + 4 Wm
with ΔHm = H11(xm) − H22(xm) and Wm = H12(xm). Introducing derivatives
c c a
fkl = ∂Hkl ∂Rc and gkl = ∂ 2 Hkl ∂R2c , the energies E1, 2 x can be expanded up
to second order in x and are found to be opposing parabolas glancing at xm.
5.5 Electronic Rearrangement for Changing Conformations 183
The momentum coupling between the two adiabatic states is a special case of the
a
previous relation described for approximate adiabatic states, here Φk R =
Φ R C k R in matrix notation with Ck a 2 × 1 column matrix, giving
ℏ †
C Φ R ∇R Φ R C 2 + C †1 Φ R Φ R
a
G12 = ∇R C 2
i 1
−1
a a a
∇R C 2 = E1 − E2 ∇R H R − E2 I C2
The denominator in the second line has a minimum value ΔEa at the avoided
crossing, showing that this is where the momentum coupling will be large.
Conditions under which H12(x) ≡ 0 give situations with crossings and can be
found from considerations of state symmetry. This happens if: (i) The two
original states have different electronic spin quantum numbers; or (ii) the
two original states belong to different irreducible representations of their point
group symmetry. The groups are ∞ h for homonuclear diatomics and C∞v for
heteronuclear diatomics. When the symmetry species (or labels) of the diatomic
states are different, then H12(x) ≡ 0, and if the Ek(R) = Hkk(x) are equal at a real
valued x0 in the original representation, it follows that a crossing of adiabatic
a
potential energies E1, 2 may occur at a distance R = Rc + x0.
If instead the two states have the same spin quantum numbers and point
group symmetry, then in general H12(x) 0. In this case, to find the root where
adiabatic energies are equal one must require that the radical L(x) = L0 + xL1 +
x2L2 = 0, giving in general a complex-valued root x0 + ix0 , and a crossing in the
complex plane of position variables.
For a complex valued coupling H12(x) = H12(x) exp[iα(x)], the two new adi-
abatic states around an avoided crossing are [25]
a iα iα
Φ1 R = Φ1 R exp − cos β + Φ2 R exp sin β
2 2
a iα iα
Φ2 R = − Φ1 R exp − sin β + Φ2 R exp cos β
2 2
with
tan 2β x = 2 H12 x H11 x − H22 x
taking a large value at the avoided crossing, leading there to a rapid change of
β(x) and rapid switch between states.
This analysis can be extended to cover interactions involving a spin-orbit
couplings in a Hamiltonian, in which case the relevant point symmetry groups
are extended groups with representations generated by the direct product of
spin dependent and space-dependent basis functions. In particular double
groups arise when orbital angular momentum states are extended to total
angular momentum states to include half-integer spin states, and their
184 5 Intermolecular States
E1, 2 x, y = M x, y ± N x,y
1
M x,y = H11 x, y + H22 x, y
2
2 2 1 2
N x,y = H11 x,y − H22 x, y + 4 H12 x,y
The condition for intersection of PESs is N(x, y) = 0, or both H11(x, y) − H22(x, y)
= 0 and H12(x, y) = 0, which may be satisfied by roots (x0, y0), corresponding
to variables Qν0 expected to be close to Qνm , ν = a, b. When this happens for
the original (diabatic set) matrix elements, the new energies E1,2(x, y) of the
adiabatic states can be expanded around the roots up to second order in the
displacements so that
E1, 2 x, y = M 0, 0 + M 1, 0 x + M 0, 1 y + M 2, 0 x2 2 + M 0, 2 y2 2
+ M 1, 1 xy N 1, 0 x + N 0, 1 y + N 2, 0 x2 2 + N 0, 2 y2 2 + N 1, 1 xy
Each of these two energy functions gives an isocontour surface for each con-
stant energy in the space of E versus x, y. Since they are quadratic forms, the
5.5 Electronic Rearrangement for Changing Conformations 185
Figure 5.5 Illustration of three cases of conical intersections between two potential energy
surfaces: (a) Bound-to-bound; (b) bound-to-unbound; (c) unbound-to-unbound.
isosurfaces are conical forms, and they touch at a degeneracy vertex, at the loca-
tion of the roots (x0, y0) in a (x, y) plane. They can be opposing cones or collateral
ones, depending on slopes along x and y. This in turn depends on what types of
bonding surfaces are interacting. Figure 5.5 illustrates conical intersections for a
triatomic system ABC, as can be expected coming from (a) isomerization, that is
a bound-to-bound structure interaction; (b) from dissociation in bound-to-
unbound structure changes; or also (c) from collisions as unbound-to-unbound
structure events.
Insight on the nature of these conical intersections can be gained constructing
simple models of the intersecting conical surfaces (or CISs) in a space E versus
(x, y), where x and y are bond distances or angles, and the other Nint − 2 internal
conformation variables are kept fixed. The intersection of two diabatic potential
energies leads to CISs when coupled, and as the other variables are changed the
conical shapes can be pictured as moving along seams in a hyperspace of Nint − 2
dimensions. Examples of models are given as follows.
a) Bound-to-bound CIS:
ω21 a 2 ω22 2 ω2 a 2 ω22 2
H11 = x+ + y , H22 = 1 x− + y − Δ, H12 = cy
2 2 2 2 2 2
This corresponds to two intersecting paraboloids and describes for exam-
ple SH2, of point group symmetry C2v, in the excited states 11A and 21A
[27], with x corresponding to the H-H distance r, and y for the distance R
from S to the center of mass of HH. The third variable is the angle between
r and R orientations, and is considered here to be chosen at a fixed value.
b) Bound-to-unbound CIS:
ω2 a 2 ω22 2 ω2
H11 = 1 x + + y , H22 = A exp − α x− x0 + 2 y2 −Δ,
2 2 2 2
x− x0 2 y −y0 2
H12 = cy exp +
λ2 μ2
where the unbound motion occurs as the x variable increases. This applies to
the dissociation of excited HNO(1Δ) into H(2S) + NO(2Π) [28], with the two
186 5 Intermolecular States
original electronic states having the same symmetry. The magnitude of the split-
ting at the vertex can be found doing perturbation theory with the vibronic per-
turbation energy operator generated by displacing internal variables from the
reference values at the minimum conformation. Models can be presented in
terms of adiabatic or nonadiabatic PESs [29]. Two adiabatic PESs are also sub-
a
ject to nonadiabatic momentum couplings G kl , with momentum components
along the x and y directions, even if all other internal variables are assumed to
be fixed.
Depending on the nature of the two variables, such as whether they describe
distortions of a bond length or of a bond angle, the conical shapes may become
flattened and in the limit of a breaking bond the degeneracy may become a line,
showing as a seam for two intersecting PESs along the direction of the breaking
bond. This is what happens, for example in the case of the PES for collinear H+
+ H2 interacting with the PES for H + H2+ [20], and of the PES for Ar+ + H2
interacting with that of Ar + H2+ as shown in Figure 5.6a for the adiabatic
potential energies and Figure 5.6b for their nonadiabatic momentum coupling
a
τ = G kl ℏ along the diatomic axis [30]. The crossing between the two lowest
PESs can also be described as two intersecting potential energy lines in the plane
of the energy E and the bond distance r in H2 and in H2+, translated along a seam
in a direction perpendicular to that plane, as the distance R between Ar+ and H2
is changing. The crossing in Figure 5.6a below 1.0 Å (or 0.1 nm) is transformed
due to the momentum coupling along R, into an avoided crossing as shown in
Figure 5.6b.
The two adiabatic PESs are also subject to nonadiabatic momentum cou-
a
plings G kl , with momentum components along the x and y directions, even
if all other internal variables are assumed to be fixed.
A special situation arises when the two original electronic states, 1 and 2, are
degenerate with the same energy H11(x, y) = H22(x, y) for all argument values.
The degeneracy can be lifted due to a coupling H12(x, y) 0. Conditions under
which this happens can be found assuming vibronic coupling and can be stud-
ied by expanding the electronic Hamiltonian operator of the system in powers
of small displacements from the reference structure. Coupling matrix ele-
ments of the resulting force operator are different from zero if the direct prod-
uct of the three representations of the point groups for the two original states
and for any of the displacements contain the totally symmetric representation.
In this case, the degeneracy can be lifted with new states and PESs found from
the solution of the two-state eigenvalue equations. This analysis has been fol-
lowed to arrive at the Jahn–Teller theorem [31], stating that in any symmetric
nonlinear molecule with electronically degenerate states, displacement of
atoms lead to symmetry distortions that decrease the original symmetry.
188 5 Intermolecular States
(a)
5
+
Ar + H 2(2Σu)
4 Ar+ + H2(3Σu)
3
Ar+ + H + H
2
1
E (eV)
Ar + H+ + H
0 Ar+ + H2(4Σg)
–1
+
–2 Ar + H 2(2Σg)
–3
0 1 2 3 4
RH–H (Å)
(b)
Energy relative to Ar + H+ + H (eV)
–1.5 20
v1 v2
Interaction (Å–1)
–2.0 10
τr(R,r)
–2.5 0
–3.0
0.5 0.6 0.7 0.8 0.9 1.0
Distance (Å)
Figure 5.6 (a) Adiabatic potential energy surfaces for collinear Ar+ + H2 interacting with Ar + H2+
along the diatomic axis, as functions of the H to H distance and for large intermolecular Ar-H
distances; (b) non-adiabatic momentum coupling between the two lowest PESs at a fixed
intermolecular distance. Source: adapted from Ref. [30]. Reproduced with permission of Elsevier.
If the energy of the two degenerate states is, to begin with, a paraboloid show-
ing a single minimum versus x and y, the theorem indicates that this form will be
distorted leading to a more complex form with two minima in the (E, x) plane as
shown in Figure 1.6. But now as a circle is traced in the (x, y) plane around the
5.5 Electronic Rearrangement for Changing Conformations 189
degeneracy point there are energy maxima and minima corresponding to lower
symmetry structures [32]. An example is provided by the radical C5H5, of symme-
try D5h to begin with, and in doubly degenerate states of label 2 E1 . The coupling
between the two states can be written in terms of the variables x = ρ cos(5φ) and
y = ρ sin(5φ) and as the angle φ varies over a circle this shows energies of struc-
tures with the lower symmetry C2v. The distortion may also arise from atomic
momentum couplings between the two original states, involving derivatives in
the x and y variables, again provided symmetry arguments indicate that the cou-
pling is not zero.
Conical intersections of PESs can be expected to appear when dealing with
photoexcited polyatomic molecules and may show up in large numbers when
the molecular system is complex, as in polymers and biomolecules. Recent
reviews have covered the role that CISs play in molecular spectroscopy and
photoinduced dynamics [33], in isomerization [34], organic photochemistry
[35], and inorganic compounds with transition metal atoms. The presence of
CISs in the primary isomerization event in vision has been experimentally con-
firmed using ultrafast optical spectroscopy in the visible and near infrared [36].
It can be expected that many CISs will be present and linked through seams in
photoinduced biomolecular dynamics.
electronic wavefunction points, where other states with l k have the same
energy Ek(Q) and interfere quantum mechanically with the state above, and also
what happens due to momentum couplings of electronic states.
Even when degeneracies and couplings are absent, the phase may alter
numerical values of PESs and the related dynamics. From the definition of
a a a
the momentum G kk Q , now for electronic state ηk Φk , one finds that it is
a
not any longer null but that instead G kk Q = ℏ∂γ k ∂Q, a hyper-momentum-
like quantity in the Q hyperspace, and that by integration along a line integral
from point Q0 to Q, the phase function factor is
Q
a i a
ηk Q = exp dQ G kk Q
ℏ Q0
H Q s Υk Q s = Ek Q s Υk Q s
Differentiating it with respect to s, one can construct a state change at position
s1 = s + Δs from
∂H Q s
∂Υk Υl ∂s Υk
= Υl 1
∂s 1
l k 1
Ek Q s1 −El Q s1
followed by a second step to s2 = s + 2Δs giving
∂H Q s ∂H Q s
∂Υk Υk ∂s Υl Υl ∂s Υk dγ k
Υk =− 1 2
=
∂s 2
l k Ek Q s1 − El Q s1 2 ds 2
observation made above for a single adiabatic state and shows the way to con-
struct geometric phase factors for many coupled states.
It can be expected that geometric phases in an adiabatic representation will
play a quantitative role in the calculation of dynamical properties of electron-
ically excited systems, such as optical spectra and collisional phenomena show-
ing electronic rearrangement. This happens in the reaction D + H2 HD + H,
for which comparison of theory and experiment have shown the existence of a
conical intersection and a need for a geometric phase in the calculation of the
reaction cross section using adiabatic potential energy surfaces [43].
References
1 Arfken, G. (1970). Mathematical Methods for Physicists, 2e. New York:
Academic Press.
2 Hohenberg, P. and Kohn, W. (1964). Inhomogeneous electron gas. Phys. Rev. B
136: 864.
3 Levy, M. (1979). Universal variational functionals of electron densities. Proc. Nat.
Acad. Sci. U. S. A 76: 6062.
4 Kohn, W. and Sham, L.J. (1965). Self-consistent equations including exchange
and correlation effects. Phys. Rev. A 140: 1133.
5 Gordon, R.G. and Kim, Y.S. (1972). Theory for the forces between closed shell
atoms and molecules. J. Chem. Phys. 56: 3122.
6 Parr, R.G. and Yang, W. (1989). Density Functional Theory of Atoms and
Molecules. Oxford, England: Oxford University Press.
7 Levine, I.N. (2000). Quantum Chemistry, 5e. Upper Saddle River: Prentice-Hall.
8 Born, M. and Huang, K. (1954). Dynamical Theory of Crystal Lattices, Appendix
VIII. Oxford, England: Oxford university Press.
9 Hirschfelder, J.O., Curtis, C.F., and Bird, R.B. (1954). Molecular Theory of Gases
and Liquids. New York: Wiley.
10 Hirschfelder, J.O. (ed.) (1967). Intermolecular Forces (Adv. Chem. Phys., vol. 12).
New York: Wiley.
11 Nikitin, E.E. (1974). Theory of Elementary Atomic and Molecular Processes in
Gases. London, England: Oxford University Press.
12 Micha, D.A. (1974). Effective Hamiltonian Methods for Molecular Collisions
(Adv. Quantum Chem., vol. 71), 231. New York: Academic Press.
13 Garrett, B.C. and Truhlar, D.G. (1981). The coupling of electronically adiabatic
states in atomic and molecular collisions. In: Theoretical Chemistry, Part A,
vol. 6 (ed. D. Henderson), 216. New York: Academic Press.
14 Baer, M. (1985). The theory of electronic nonadiabatic transitions in chemical
reactions. In: Theory of Chemical Reaction Dynamics, vol. II (ed. M. Baer), 219.
Boca Raton, FL, USA: CRC Press.
References 193
15 Smith, F.T. (1969). Diabatic and adiabatic representations for atomic collision
problems. Phys. Rev. 179: 111.
16 Baer, M. (1975). Adiabatic and diabatic representationsfor atom-molecule
collisions: treatment of the collinear arrangement. Chem. Phys. Lett. 35: 112.
17 Mead, A.C. and Truhlar, D.G. (1982). Conditions for the definition of strictly
diabatic electronic basis for molecular systems. J. Chem. Phys. 77: 6090.
18 Olson, J.A. and Micha, D.A. (1982). Electronic state representations at molecular
potential pseudocrossings. Int. J. Quantum Chem. 22: 971.
19 Micha, D.A. (1983). A self-consistent eikonal treatment of electronic transitions
in molecular collisions. J. Chem. Phys. 78: 7138.
20 Tully, J.C. (1976). Nonadiabatic processes in molecular collisions. In: Dynamics
od Molecular Collisions Part B (ed. W.H. Miller), 217. New York: Plenum Press.
21 Cohen, J.M. and Micha, D.A. (1992). Electronically diabatic atom-atom
collisions: a self-consistent eikonal approximation. J. Chem. Phys. 97: 1038.
22 Fernandez, F.M. and Micha, D.A. (1992). Time-evolution of molecular states in
electronically diabatic phenomena. J. Chem. Phys. 97: 8173.
23 Bruna, P.J. and Peyerimhoff, S.D. (1987). Excited-State Potentials (Adv. Chem.
Phys., vol. 67, Part I), 1. New York: Wiley.
24 Hirschfelder, J.O. and Meath, W.J. (1967). The nature of intermolecular forces.
Adv. Chem. Phys. 12: 3.
25 Cohen-Tanoudji, C., Diu, B., and Laloe, F. (1977). Quantum Mechanics, vol. 1,
Chap. IV. New York: Wiley.
26 Tinkham, M. (1964). Group Theory and Quantum Mechanics. New York:
McGraw-Hill.
27 Matsunaga, N. and Yarkony, D.R. (1997). Energies and derivative couplings in
the vicinity of a conical intersection II. CH2 and SH2. J. Chem. Phys. 107: 7825.
28 Herzberg, G. and Longuet-Higgins, H.C. (1963). Intersection of potential energy
surfaces in polyatomic molecules. Discuss. Faraday Soc. 35: 77.
29 Worth, D.A. and Cederbaum, L.S. (2004). Beyond Born-Oppenheimer: molecular
dynamics through a conical intersection. Annu. Rev. Phys. Chem. 55: 127.
30 Baer, M. and Beswick, J.A. (1979). Incorporation of electronically nonadiabatic
effects into bimolecular reactive systems. III. The collinear Ar + H2+ system.
Phys. Rev. A 19: 1559.
31 Jahn, H.A. and Teller, E. (1937). Stability of polyatomic molecules in degenerate
electronic states. I. Orbital degeneracy. Proc. R. Soc. A 161: 220.
32 Herzberg, G. (1966). Molecular Spectra and Molecular Structure III. Polyatomic
Molecules. New York: Van Nostrand Reinhold.
33 Domcke, W. and Yarkony, D.R. (2012). Role of conical intersections in molecular
spectroscopy and photoinduced chemical dynamics. Annu. Rev. Phys. Chem.
63: 325.
34 Levine, B.G. and Martinez, T.J. (2007). Isomerization through conical
intersections. Annu. Rev. Phys. Chem. 58: 613.
194 5 Intermolecular States
Many-Electron Treatments
CONTENTS
E ≈ EA + EB and ρ ≈ ρA + ρB
This will be called here asymptotic consistency, but is usually termed size con-
sistency in the literature. Its requirement is essential in treatments of intermo-
lecular forces. It guarantees that the potential energy of interaction at all R
satisfies V(R) = E(R) − (EA + EB) ≈ 0.
When the number of interacting identical species S is increased, as one pro-
ceeds from clusters to solids or fluids, the extensive properties (in the sense of
statistical thermodynamics) must increase linearly. For a system SN with a large
number N of the stable species S, one has E(SN) ≈ NE(S), as N ∞, where SN
may be a compound or a complex. This is size extensivity, a property needed in
treatments of many-molecule systems.
Two general methods for intermolecular forces are the so-called supermole-
cule method, where the interacting species are considered to be a single many-
electron system in a field of all the nuclei, and the many-atom method where the
total many-electron system is instead treated as a collection of atomic many-
electron systems. In both cases, one must properly treat the overall symmetry
of the many-electron state with regard to electron exchange, its total electronic
spin state, and its total point group or crystal group symmetry.
The methods described in what follows are applicable also to chemical reac-
tions such as A + B C + D, where C and D are new species, involving the rear-
rangement of nuclei. Treatments of energetics require exploration of potential
energy surfaces over wide ranges of atomic positions, to identify potential
6.1 Many-Electron States 197
A B
(b2)
A+ B–
(c1)
A B
(c2)
A+ B–
energy barriers between regions of low energy. Much can be learned about
allowed or disallowed reactions from qualitative treatments based on the shapes
of molecular orbitals and the interactions between highest occupied and lowest
unoccupied molecular orbitals (HOMOs and LUMOs) of the interacting
species [6–8]. But accurate calculations of reaction energy barriers, stable
conformations, and vibrational-rotational properties require many-electron
treatments as described in what follows.
198 6 Many-Electron Treatments
int
energy that E = ΦAB H A + H B + H AB ΦAB ≈ EA + EB and the potential
energy V(R) vanishes at large distances.
Spin and space symmetries can be imposed starting from normal many-
† †
electron operators N, with the property N N = N N, which generate symme-
try transformation and are constants of motion insofar as NH = HN. Their
eigenstates and eigenvalues in N k = νk k can be used to define projection
operators k = k k , which extract a component of correct symmetry from a
wavefunction as Φk X = X kΦ = X k k Φ X , where it is shown that
the second factor may also depend on some of the electronic variables. In appli-
cations where the constant of motion operator has only a discrete set of eigen-
values, a frequent situation, the projection operator can be conveniently
constructed as
N − νl
k = l kν
k −νl
with the product extending over only different eigenvalues. It satisfies the
orthogonality property k l = δkl k .
This is applicable to the angular momentum operator with components
M ξ , ξ = x,y, z, whether this is a component of the spin S or of the orbital angular
momentum L. For the total spin operator with components S ξ = j S ξj , of mag-
2
nitude and projection S and S z with eigenvalues S(S + 1) and MS = − S, − S + 1,
…, S, the projector SMS = S MS can be written as [10]
with h the order of the group and dk the dimension of the representation.
The orthogonality properties of characters ensures that Γk Γl =
δkl Γ k . The asymptotic form of this projector depends on how the overall
symmetry is expressed in terms of the symmetries of the Hamiltonians of A and
B. In some cases, such as for overall translation or reflection on a plane, the
group is a direct product of those of A and B, = A ⨂ B , with characters
χ k R = χ kA RA χ kB RB , which allow factorization of the projector, and a
proper asymptotic decomposition will follow. However, if the total group
involves also a complementary subgroup as in = ⨂ A ⨂ B , then a further
analysis of the asymptotic wavefunction and expectation values must be done to
establish whether cross terms vanish to give the correct behavior [3].
Many-electron wavefunctions are sometimes variationally obtained by mini-
mizing the total electronic energy but do not display the total symmetries of the
system. In such cases, it is yet possible to extract the correct symmetry by using
projection operators after the variation, in a postsymmetrization. Consider a
variational but nonsymmetrized wavefunction Φ, which would contain symme-
try-adapted components Φk as in Φ = kakΦk, assumed to be all normalized for
simplicity. The variational energy is Evar = k|ak|2Ek with factors |ak|2 ≤ 1.
Therefore postsymmetrized components are found to provide energies Ek lower
than the unsymmetrized original wavefunction. But while the original unsym-
metrized wavefunction gives an upper bound to the calculated energy, the post-
symmetrized one may not do so for the exact symmetrized energy and must be
tested for accuracy of the postsymmetrized associated energy. This can be done,
for example, by doing some calculations with prior symmetrization and other-
wise the same conditions at selected nuclear conformations, to compare ener-
gies with results from postsymmetrization.
The interaction energy of open-shell or of electronically excited species
presents additional complications because they frequently involve a number
ND of states ΦJ, J = 1 to ND, of the whole system that are degenerate or nearly
degenerate in energy, and are either eigenstates of the Hamiltonian for the
energy E(D), with H ΦJ = E D ΦJ or have very similar expectation values
ΦJ H ΦJ E D . They define a multireference, or configuration interaction
(CI), state space to be treated as a whole. In these cases, one must extend the
usual treatments to include linear combinations of the ΦJ states in the CI multi-
reference wavefunctions, or one must construct unperturbed multireference
Hamiltonians in perturbation and coupled cluster treatments.
202 6 Many-Electron Treatments
the solution of the Schrodinger equation, one finds that using the reduced resol-
vent operator
−1
RE = α + E −H
Ψ = ΩΦ D = + RE H ΦD
as can be verified applying the inverse of R E to both sides of this equation. This
multireference wave operator can be constructed using CI methods or pertur-
bation expansions, described as follows.
Electronic antisymmetry and spin and space symmetries enforced by projec-
AS
tion operators and k can all be maintained when the wave operator is
applied, provided the projector has been symmetry-adapted so that
AS AS AS AS
= and k = k . Then one finds that Ω = Ω and
Ω k = k Ω insofar as the wave operator has been constructed from the full
Hamiltonian.
The energy of the whole system is obtained as E = Φ D H ΩΦ D N D ,
and the potential energy follows from V = E − (EA + EB). For the purpose of
generating potential energy surfaces going to the correct asymptotic limits,
it is convenient, but not necessary, that the reference state is chosen as
D D
ΦD = ΦA ΦB and that the wave operator for the pair of species A and B
D D
satisfies ΩAB Φ D ≈ ΩA ΩB ΦA ΦB as R ∞, with the complementary
antisymmetrizer. In this case, the expectation value of an extensive property
would go asymptotically to the sum of its values for the two separate species,
and in particular V = E − (EA + EB) ≈ 0, with asymptotic ener-
D
gies EI = Φ D H ΩΦ D I NI ,I = A,B.
6.1 Many-Electron States 203
1 1
ΦJ A ΦJ = d 1 d 1 a 1 1, 1 ρJ 1 ;1 ,
2 2
ΦJ A ΦJ = d 1 d 1 d 2 d 2 a 2 1, 2,1 ,2 ρJ 1 , 2 ; 1, 2 ,
N N −1
d N Φ∗ J 1 ,2 , 3,…,N ΦJ 1, 2,3, …,N
2
ρJ 1 ,2 ; 1, 2 = d 3
2
1 2
and so on. Here ρJ and ρJ are one- and two-electron reduced density func-
tions for this state in the coordinate representations of reduced density opera-
1 2
tors (RDOps) ρJ and ρJ with the argument (1) a short form for (x1) and with
d 1 … = d 3 r1 ζ1 = ± … . They have been defined here with normalization
tr 1 ρ 1 = d 1 ρ 1 1; 1 = N, tr 2 ρ 2 = d 1 d 2 ρ 2 1, 2; 1, 2 = N N − 1
2, the number of different electron pairs, and so on. This is a common normal-
n
ization [14] but not universal. Transition averages ΦJ A ΦK can similarly
n
be expressed in terms of transition density matrices [14]. The electronic ρJK
reduced density operators are usually constructed in a basis set of one-electron
functions for computational purposes.
204 6 Many-Electron Treatments
ρ1 = k, k
k ρ 1 k,k k
ρ2 = k1 , k1 k2 , k2
k1 k2 ρ 2 k1 ,k2 , k1 k2 k1 k2
matrix S and, if they form in practice a complete set, then they also satisfy the
identity decomposition μ, ν χ μ S −1 μν χ ν = I. Orbitals of large systems like
solids or polymers can also be alternatively expanded in a basis set of plane
wavefunctions ϕk r for wavevectors k forming a grid in reciprocal space.
Important computational considerations in the choice of a basis set are how fast
can electronic integrals be done with the basis set and how large are the matrices
that must be stored and multiplied in calculations of properties. The size of the
basis set is important for achieving high accuracy of potential energies, and con-
sistent basis sets must be used when subtracting the energies of components A
and B from the total energy of a complex A + B, to obtain changes of energies
with structural parameters.
Numerous basis sets of AOs have been introduced in the literature and imple-
mented for computational work [3, 4, 15]. The original basis sets derived from
hydrogenic orbitals, products of radial functions times angular functions, and
have used Cartesian coordinates (x, y, z) or spherical coordinates (r, ϑ, φ) with
origin at the atomic nucleus positions. In the so-called (nlm) Slater-type orbitals
(or STOs), a radial factor Rn(r) = rn − 1 exp(−ζr) containing an exponential
parameter ζ multiplies spherical harmonics Ylm(ϑ, φ) for the angular depend-
ence. The nodeless radial functions are combined into contractions with fixed
coefficients to describe the physical radial functions. The more popular and
recent basis sets are made up of Gaussian radial functions (or GTOs), which
simplify the calculation of one- and two-electron integrals. The primitive radial
factors are of form exp(−ζr2) times a power of the radius, multiplying spherical
harmonics, or instead the exponential multiplies powers xaybzc of Cartesian
coordinates with a + b + c = l. The quantum number l corresponds to an occu-
pied l-subshell (s-, p-, d- …) or to a subshell with higher l if there is a need to
describe polarization or hybridization of shells.
Products of primitive GTOs ga, b, c(x, y, z; ζ) = exp(−ζr2)xaybzc located at dif-
ferent centers can be rewritten as combinations of GTOs centered at interme-
diate points, a fact that speeds the calculation of electron integrals. But GTOs do
not accurately describe the electronic wavefunction cusp at an atomic position,
and they decrease with r much faster than the physical exponential functions.
These shortcomings are compensated using: (i) primitives with more than one
value of ζ combined to give double-zeta (or DZ), triple- (TZ), quadruple- (QZ),
or n-zeta (n-Z) functions; (ii) contracted sums of gaussians (CGTO) where Nl
primitives of a given l are combined with fixed coefficients to form nl < Nl
CGTOs. A set such as 6-31G, suitable for compounds with first and second
row elements, contracts six primitive GTOs for inner shells and treats outer
shells with a contraction of three primitives plus another primitive with a smal-
ler exponent parameter ζ, for a more spread out orbital. Parameters are
obtained from atomic HF calculations. Hydrogen and He orbitals of 31G type
have exponentials further adjusted for molecular calculations.
6.1 Many-Electron States 207
B must be calculated with the same type of basis set. The other relates to the size
of the basis sets for the complex and for each component. Indicating with
ES[R, QS; {χ S}] an electronic energy obtained at relative distance R with the basis
set {χ S} suitable for species S = A, B, and AB, it would appear as if the potential
energies should be EAB[R, QA, QB; {χ A} {χ B}], obtained from the union of the
component basis sets, minus EA[{χ A}] and EB[{χ B}] from each separate basis set.
However, this means that at each distance R, the total energy is being calculated
with a more extensive (or more complete) basis set for the complex and less so
for the components, giving an inaccurate potential energy V (R, QA, QB). The
resulting error is called the basis set superposition error. It can be corrected
by calculating the components energy with the larger (union) basis set at the
conformations of the components for each distance R, as a counterpoise (CP)
correction energy
ΔECP(QA,QB) = EA QA ; χ A χ B + EB QB ; χ A χB
−EA χ A − EB χ B
giving a more accurate potential energy V(R, QA, QB) = V (R, QA, QB) +
ΔECP(QA, QB) [4, 18].
Basis sets suitable for extended systems frequently use instead plane wave
expansions for valence shells and delocalized orbitals, introduced together
with a description of inner shells in terms of atomic pseudopotentials [19].
−1 2
Plane waves ϕk r = Ω exp ik r for a system of volume Ω with a set of
wavevectors k forming a grid k n ,n = n1 ,n2 ,n3 , in reciprocal space, must
be dense enough to describe spatial variations of the electronic density. The
B B
upper limit Nj of the grid numbers nj, in 1 ≤ nj ≤ Nj , is increased to account
for highly oscillating orbitals in real space. This provides a basis set suitable for
calculations of interaction forces in large systems, and properties such as the
total energy of a crystalline solid as a function of the size of its unit cell volume.
Inner- and outer-shell atomic orbitals are differently treated, with inner-shell
PS
orbitals ϕl r introduced to construct atomic pseudopotential operators
PS
V l in which valence electrons move [20]. Expansions of one-electron wave-
functions in plane wave functions, and fast Fourier transformations between
grids in real and reciprocal spaces, provide one-electron valence states and ener-
gies for many-electron treatments of intermolecular forces in solids and at solid
surfaces. The same grid in reciprocal space can be used for a total system AB and
for its isolated components A and B to obtain accurate energy differences.
a
Expanding MOs ψ j r as linear combinations of AOs χ μ r , with Nbf basis
functions for atom 1 ≤ a ≤ Nat and a total number of basis functions Nbf,
means that
Nbf
ψj r = c χ
μ = 1 jμ μ
r
6.2 Supermolecule Methods 209
d 1 d 1 χ μ 1 ∗ a 1 1, 1 χ ν 1 = μ a 1 ν
d 1d 1 d 2 d 2 χ κ 1 ∗ χ μ 2 ∗ a 2 1, 2; 1 ,2 χ λ 1 χ ν 2 = κμ a 2 λν
Doubly excited CSFs Φijab , and generally n-excited ones, can be similarly
constructed. A CI wavefunction is given by the combination of the reference
CSF plus singly excited (S) ones, plus doubly (D) excited ones, and so on. This
can be described within the wave operator formalism of Section 6.1.2, starting
with the projector = Φ0 Φ0 valid for a single reference state. Single exci-
1
tation determinants add up to a wave operator term Ω , doubly excited ones
2
add to Ω and so on, leading to
1 2
Ψ= I +Ω +Ω + Φ0
1 2
Ω = i, a
Aia c†a ci , Ω = i, j, a, b
Aijab c†b c†a cj ci , …
HF
where f HF is the one-electron Fock operator and ϵj is the energy of the SO
[3]. The term h contains the electron n kinetic energy operator and the Cou-
lomb attraction energy of the electron to all the nuclei (or atomic ion cores).
The Coulomb and exchange HF operators are given by
occ 2
J n ψj n = k
d 1 v 1, n ψ k 1 ψj n
occ
K n ψj n = k
d 1 ψ k 1 ∗ v 1,n t 1n ψ k 1 ψ j n
with v(1, n) the Coulomb energy between electrons 1 and n, and t 1n the trans-
position operator exchanging the variables of electrons 1 and n in the electron
integrals: t 1n ψ k 1 ψ j n = ψ k n ψ j 1 . The integrals here are over space vari-
ables and also indicate sums over spin variables ζ = ± 1. The SO energies are
given by
HF occ
ϵj = j h j + k
jk v 1 − t jk
0.06
0.04
0.02
0.6 1.0 1.4 1.8 2.2 2.6 3.0 3.4 R (a.u.)
0.00
R RHF
–0.02
E(H2) – 2E(H) (a.u.)
UHF
–0.04
–0.06
Exact
–0.08
(Kolos–Wolniewicz)
–0.10
–0.12
–0.14
–0.16
–0.18
Figure 6.2 Results from RHF, UHF, and CI calculations of potential energy versus
interatomic distance for H2, using the basis set 6-31G∗∗. Source: from Ref. [5]. Reproduced
with permission of Dove Publishing.
higher excitations, must be solved for the lowest energy E and the associ-
ated state
Ψ = Φ0 + i, a
Aia Φia + i, j, a, b
Aijab Φijab + = ΦA
in a matrix notation with the row matrix Φ of CSFs and the column matrix A of
expansion coefficients. This can be done searching for the variational minimum
of Ψ H Ψ = A† Φ H Φ A = A† HA subject to the normalization Ψ
Ψ = A† Φ Φ A = 1 in a well-known procedure leading to the matrix eigen-
value equation (H − EI)A = 0, which needs to be solved only for the lowest
eigenvalue as a function of conformations and in particular as a function of
R, if one wants only the ground state PES.
A more flexible and accurate procedure is the multiconfiguration SCF (or
MCSCF) treatment where the CI coefficients A and also the HF coefficients
a are varied to achieve self-consistency at the lowest energy. This gives orbitals
adapted to optimize not only the reference configuration, but also useful to opti-
mize its mixture with excited ones, and leads to more realistic excited SOs for
the supermolecule.
The CASSCF expansion is a MCSCF treatment involving all the electron exci-
tations among active SOs, and guarantees that the potential energy V(R) from the
difference of total energies goes properly to zero at large distances, provided the
reference CSF Φ0 correctly describes the asymptotic states of A and B at large R.
The limitation here is that the number of configurations to be generated increases
exponentially with the number of active SOs. In fact, if Nact is the number of active
orbitals, N the number of active electrons, and S the total spin quantum number,
the number of complete active space functions to be generated is [23]
2S + 1 Nact + 1
NCAS =
Nact + 1 N 2 −S Nact + 1 − N 2 + S
Nact + 1
×
N 2+S+1 Nact + 1 − N 2 − S − 1
a very large number when Nact is larger than 10, which prevents calculations of
this type for large molecular systems or large electronic excitation.
The calculation of interaction potential energies gets more demanding when
there are state degeneracies or electronically excited states, or where a single
CSF Φ0 does not properly describe the asymptotic A and B states. In these cases,
one must use a treatment based on multireference states. Instead of a single ref-
erence CSF Φ0, one must consider a set {ΦJ, J = 1 to ND} of CSFs, introduce the
projection operator = J ΦJ ΦJ , and proceed to construct multireference
CI (or MR-CI) states, to calculate energies.
An example is provided by the treatment of H2 H + H in its ground elec-
tronic state 1 Σg+ with a minimal MR-CI containing only two configurations
6.2 Supermolecule Methods 215
0 1 HF
reconstruct the HF energy, E0 + E0 = E0 , a meaningful total energy.
The first summation in the second-order energy involving single excited
1 2
states is zero in the HF treatment, insofar as D0 H +H Dia =
a f HF i = 0, per the so-called Brillouin theorem [4].
The remaining summation in the second-order energy gives the leading cor-
relation energy involving excited SOs. However, the excited (unoccupied) SOs
generated in HF correspond formally to an electron moving in a field of N elec-
trons, instead of N-1 electrons as happens for the occupied SOs. As a result, they
usually describe electrons that are less bound and are more delocalized than
accurate excited SOs. Calculation of interaction energies may require a larger
basis set and inclusion of higher orders in the MPPT expansion, as compared
to expansions with more realistic excited SOs, or otherwise they may lead to
large errors in intermolecular forces at large distances. Furthermore perturba-
tion terms must be carefully constructed to avoid unlinked diagrams and non-
extensive results.
The asymptotic consistency requirement E ≈ EA + EB for a pair of species A
and B, essential for intermolecular forces, is not necessarily satisfied by MBPT.
This is because the reference state must satisfy D0 ≈ D0, A D0, B , a correct
asymptotic reference state made up of products of states of A and B, and in
MB
addition, the perturbation treatment must construct the wave operator ΩAB
MB MB MB
so that the energy as obtained from H AB ΩAB D0 ≈ EA + EB ΩA ΩB
D0, A D0, B gives accurate energies for A and B. This must be verified in
a treatment of the intermolecular forces.
218 6 Many-Electron Treatments
One finds, with the expansion of the exponential operator, a sum of config-
urations as in the CI method, but here their superposition is generated by the
CC excitation terms in a convenient way, which involves fewer transition ampli-
tudes than the number of coefficients in the CI. The unknowns in the CC treat-
ment are E, tia , tijab ,…, and these can be extracted by projecting (or taking
moments of ) the equation exp −T Hexp T D0 = E D0 on the Dirac bras
D0 , Dia , Dijab ,…. This gives the same number of nonlinear algebraic
equations as there are unknowns, which can be solved by iteration. An analysis
of many-electron terms in E = D0 exp − T Hexp T D0 after an expansion
in commutators of H and T shows that only linked diagrams appear, assuring
size extensivity [25].
6.2 Supermolecule Methods 219
RK = r 0 I + r a c† c
i, a i a i
+ r ab c† c† c c +
i, j, a, b ij b a j i
where the terms are assumed to apply in the normal order (of creation operator
followed by annihilation operator to the right) as done with the terms in the
excitation operator T . Since operators in RK and T are all of excitation type,
this gives the commutator RK ,T = 0. It follows from H RK Ψ0 =
EK RK Ψ0 and from Ψ0 = exp T D0 that [25]
–108.95
–109.00
–109.05
–109.10
E (Hartree)
–109.15
–109.20
FCI
CCSD(T)
CCSDT
CCSDTQ
–109.25 UHF-CCSD(T)
UHF-CCSDT
UHF-CCSDTQ
UHF-CCSDTQP
UHF-CCSDTQPH
–109.30
1.5 2.0 2.5 3.0 3.5 4.0 4.5
R (a.u.)
Figure 6.3 Potential energy curves versus interatomic distance for the ground state of the N2
molecule. Results from CC are compared to full CI using a cc-pVDZ basis set. Incorrect
behavior of CC at large distances can be corrected using a UHF reference state. Source: from
Ref. [27]. Reproduced with permission of American Physical Society.
and its application into the subspace can be done by first defining an effective
eff
Hamiltonian H operating only within the subspace with matrix elements
eff
HKJ = ΦK HΩ ΦJ = ΦK H exp T J ΦJ . The perturbation corrected
energies EJ in HΨJ = EJ ΨJ , J = 1 to ND , and excitation amplitudes like tia J in
each partial excitation operator follow from projections of the kets
eff
exp − T J H exp T J ΦJ = K
exp −T J exp T K ΦK HKJ
eff
onto the bras ΦJ , Φia , Φijab , …, and diagonalization of HKJ with a linear
transformation in the subspace.
This procedure allows consideration of dissociation and intermolecular forces
in the AB system when the subspace contains all the determinantal functions
needed to construct correct asymptotic states of A and B. For example, in the
case of H2 H + H, it is convenient to choose the subspace as made of states
with the configurations Φ1 = σ gα, σ gβ and Φ2 = σ uα, σ uβ , which are close in
energy for large internuclear distances. The MR-CC procedure can then give the
correct potential energy surface at all distances. Dissociation in other diatomic
systems including multiple bonds has been reviewed in reference [27]. Accurate
results over all interatomic distances have been obtained for F2, N2, and HF.
Results so far indicate that using a MR-CC approach, or alternatively using
an UHF reference state, provides ways to correctly treat bond dissociation. This
is an area of theoretical research yet under development.
The scaling of MBPT and CC calculations with the size Nbf of the basis set can
be described separating the MO basis set into two sets of occupied and unoc-
o u
cupied MO functions numbering Nbf ≤ N and Nbf , respectively. Excitation
operators fix the indices of occupied MOs so that transformations from the AO
basis to the MO basis involve fewer summation loops. Transforming a four
2
index (two-electron) integral like D0 H Dijab in MBPT or tijab in a CC treat-
u 2
ment, for fixed i and j, involves summations scaling only as Nbf , which must
222 6 Many-Electron Treatments
State
quality
Full CI
MCSCF
CASSCF
S
CCSDT PE
of
SAPT cy
ura
Acc
CCSD(T)
MBPT4
CCSD
CISD
MBT2 Basis
set size
DZP TZP ... ccVDZP cc-pVTZP aug-cc-pVDZP
Figure 6.4 Diagram with increasing basis set quality, containing polarization terms, going
from left to right on the horizontal axis, and increasing many-electron quality going up as
needed for the calculation of accurate potential energy surfaces (PESs).
2
o
be done Nbf times, an effort much smaller than (Nbf)4 for a general four-
index quantity. Analysis of the scaling in CCSD(T) shows it growing as
o 3 u 4
Nbf Nbf . Further analysis of scaling is found in the literature [4, 15, 25].
The quality of electronic structure calculations for intermolecular forces
depends on the type of treatment being used and the size of the basis set in
expansions. Figure 6.4 shows what treatments of electronic structure account
for asymptotic and size consistency and what basis sets are large enough to
account for electron transfer, electronic localization, and polarization.
then allowed to distort, polarize, or transfer between species. One of these treat-
ments derives from the valence-bond (or VB) method for electronic structure,
generalized to deal with electron transfer and polarization. The GVB method
has been mathematically developed, and more recently has been reintroduced
as a viable computational method [2, 33, 34]. It is appealing for studies of inter-
molecular forces because it provides a direct route to construction of reference
states with the correct asymptotic behavior.
Total wavefunctions for N electrons are written as products of a spin function
J
ΘSMS ζ1 ,…,ζN of spin variables for quantum numbers (S, MS) times a space
function of position variables FJ r 1 ,…,r N for many-electron states J, describing
single or multiple bonds with paired electrons in the case of bonding between A
and B, or describing their nonbonding interactions. The space function is con-
structed from atomic orbitals of each species and the pairing of spin and space
functions is done so that as R ∞, one finds the correct asymptotic limit fac-
torized, with the total wavefunction
J
Φ 1, …,N = F
J J
1, …, N ΘSMS 1, …,N
K L
≈ KL
CKL FK 1, …, NA ΘSA , MSA 1, …,NA FL 1, …, NB ΘSB , MSB 1, …, NB
In a simple example with one valence electron per species A and B in atomic
orbitals (AOs) χ p r 1 and χ q r 2 forming a single bond, the standard covalent
VB wavefunction, written here for a homonuclear diatomic, is the single bond
function with F(cov)(1, 2) = χ p(1)χ q(2) and
χp 1 χq 2 + χp 2 χq 1
Φ cov 1, 2 = 1 2
Θ0, 0 1,2
2
2 + 2 Sab
Θ0, 0 ζ1 ,ζ 2 = α ζ1 β ζ 2 − α ζ 2 β ζ 1 21 2
with Spq = χ p χ q ≈ 0 the overlap integral of the two nonorthogonal AOs, going
to zero at large distances. This wavefunction clearly gives the correct asymptotic
behavior.
When the interacting species dissociate instead into ions with the two elec-
trons at the same site, the two-electron state contain F(ion)(1, 2) = χ u(1)χ u(2),
u = p, q, and is
χp 1 χp 2 + χq 1 χq 2
Φ ion 1, 2 = 1 2
Θ0, 0 1,2
2
2 + 2 Spq
again giving the correct limit. An accurate treatment of the potential energy for
all distances R combines the cov and ion states.
A pair of species with multiple bonds can be similarly described with a product
of bond functions constructed with the correct total electronic antisymmetry.
224 6 Many-Electron Treatments
Generalizations can be derived using more flexible orbitals ϕμ, which are yet cen-
tered at each atom or fragment μ, but have each AO mixed with the orbitals of
other fragments, present as small components. Another generalization involves
using all the spin functions, which can be generated for N electron spins, instead
of just the spin-pairing per bond.
The generalized VB (or GVB) treatment [33] replaces the AO χ p with a
more flexible localized orbital ϕp written as a linear combination of functions
with coefficients varied to optimize the system energy. It can be chosen to be a
combination of AOs all centered at the same fragment A, to account for
hybridization and polarization, or it can be a combination of orbitals at
center A with some mixing of orbitals at B to account for electron transfer
between them. Furthermore, orbitals localized at A can be the original ones
nonorthogonal to those of B, or they can be transformed by symmetric ortho-
normalization into orbitals predominantly centered at A or B [14]. Orthonor-
malization allows for simpler expressions in matrix elements of operators, but
dilutes the physical significance of spin pairing of orbitals in the description
of bonds.
The simple spin pairing shown above in Θ0,0(1, 2) is only one of several
pairings of electron spin that are acceptable in polyatomic systems with multiple
bonds. The treatment can be made rigorous introducing a set of N-electron
J
spin functions ΘSMS 1, …,N for given quantum numbers (S, MS) and making
use of properties of the symmetric group N of all permutation operators
N of electron variables in N-variable functions [2]. Insofar as these permuta-
tions and the projection operator SMS for given S, MS commute, it follows that
J K
N ΘSMS = K ΘSMS PJK , with PJK a matrix element of an irreducible matrix rep-
resentation of the symmetric group [2]. These irreducible representations and
their characters are well known, and they can be used to construct all the needed
spin functions for each MS. The space functions FJ(1, …, N) multiplying the spin
functions in the expansion of total states Φ can be written as products of N loca-
lized orbitals ϕu, u = a, b, of a basis set chosen to account for electronic
rearrangements.
Among the possible total states, the ones relevant to intermolecular forces go
A B A B
asymptotically to products ΦK ΦL with ΦK and ΦL states of the species A
and B constructed from the same orbitals. The number of spin functions created
in the unitary group procedure above is
2S + 1 N
fSN =
N 2 + S + 1 N 2− S
and it can get quite large for a large number N of active electrons, so that this
general procedure is in practice limited to systems with few active electrons for
all the intermolecular conformations of interest in A + B.
6.3 Many-Atom Methods 225
Ψ = ΩΦ D = + RE H ΦD
−1
with R E = α + E −H , which is as written valid for any partition of
the total Hamiltonian and implies an intermediate normalization. Provided the
multireference projection operator commutes with total symmetry projection
operators , which can be done by choosing a basis set of symmetry-adapted
total functions, this wave operator can be used to generate all perturbation
orders for the wavefunction and the energy, by recurrence. Expanding the resol-
vent R E in terms of the unperturbed resolvent R0 E0 , for the noninteracting
int 1
A + B pair with total energy E0, and using the perturbation H AB = λH AB [11],
the resulting Rayleigh–Schrodinger expansion of energy and wavefunction in
powers of λ can then be implemented computationally.
226 6 Many-Electron Treatments
The direct and exchange energies have been analyzed in detail and developed
for computational work. To simplify here we consider a single reference state
A B A B
Φ0 Φ0 of energy E0 = E0 + E0 . The direct, or “pol,” term contains to first
order the classical electrostatic (or Coulomb) interaction energy and to second
order induction and dispersion energies,
1 2
Vdir AB = Velst AB ,Vdir AB = Vind A + Vind B + Vdisp AB
The corresponding terms in the exchange Vexch are gathered from perturba-
tion terms containing transpositions for each order in the perturbation as
1 2
Vexch AB = Vexch −elst AB , Vexch AB = Vexch −ind AB + Vexch −disp AB
where Vexch − elst is a correction to the classical Coulomb interaction containing
the effect of electron exchange. The second-order energies involving sums over
excited states of A and B can be rewritten in terms of their dynamical polariz-
abilities as done in the Casimir–Polder treatment of dispersion forces [37].
Higher order terms for direct and exchange energies have been similarly ana-
lyzed [35, 36].
The implementation of this SAPT to calculate intermolecular forces requires
S
in addition a procedure for correcting the component wavefunctions Φ0 and
S
energies E0 , S = A, B, when as usual these have been obtained in approximate
treatments for many-electron systems. They are intramonomer electronic cor-
relation corrections that can be incorporated by means of perturbation treat-
S
ments of each species. Starting with known eigenfunctions ΦK of an
S
approximate Hamiltonian F S with eigenenergies EK , an additional perturba-
tion treatment can be done with F = F A + F B as the zeroth-order Hamiltonian
6.3 Many-Atom Methods 227
where K is the highest affordable perturbation order for the molecular interac-
tion energy, usually K = 3, and L is the largest order in the intramolecular per-
turbation, which within a CC treatment can include some multiple excitations
of the species A and B to all orders.
The SAPT combined with HF molecular functions, in SAPT(HF) treatments,
involves some simplifications resulting from the absence of singly excited
k, 1
matrix elements, insofar as a f HF i = 0, so that Vdir R = 0 Electronic
exchange between A and B and correlation in A and B, however, distort their
HF charge distributions and add to the intramolecular corrections. The
SAPT(HF) treatment has been successfully applied to closed-shell systems
such as (H2O)N, and Ar + H2O. They have been obtained going to second order
(K = 2) in the intermolecular interaction, which scales as Nbf7 with the number of
basis functions [36, 38]. The formalism is also suitable for interactions involving
open-shell species provided unrestricted HF states are chosen as reference ones
in a SAPT(UHF) treatment.
When F is chosen from a DFT treatment, the present PT is named
SAPT(DFT). This has been found to give accurate results provided that:
(i) the second-order interaction energies are obtained from dynamical polariz-
abilities, which correctly account for the coupling of one-electron KS excita-
tions, as in time-dependent DFT, called SAPT(CKS); (ii) high-quality basis
sets are used for atomic orbitals; and (iii) density functionals are preferably of
the hybrid type containing some exact long range electron exchange energy.
Results have been published for dimers (He)2, (Ne)2, (H2O)2 and (CO2)2 [39],
and also for the benzene dimers in a variety of conformations.
Nonadditivity in three-body systems has also been considered within SAPT,
which extends to intermediate and shorter distances the arguments presented
in Chapter 3 for long-range nonadditivity of pair interactions, by incorporation
of electron exchange [40]. The nonadditive terms originate already in first-order
228 6 Many-Electron Treatments
1 2
PT due to exchange, with Enon −add ABC = Eexch ABC + Eind ABC +
2 2 3 3
Eind −exch ABC + Edsp−exch ABC + Eind ABC + Edsp ABC + where each
term is an intrinsic three-body function. Combining SAPT with DFT, it has
been possible to calculate nonadditive energy terms for trimers of He, Ar, water,
and benzene [41].
In addition to giving accurate results for a variety of AB pairs and ABC triplets,
the SAPT treatment provides a systematic procedure for analyzing electrostatic,
induction, and dispersion contributions to intermolecular forces as functions of
atomic conformations, including electron exchange, even for many-body sys-
tems. This is useful in the development of approximations where accurate elec-
tronic energies are calculated first at short and intermediate intermolecular
distances, to which long-range induction and dispersion energies are added a
posteriori to cover all distances. In these treatments, it is necessary to avoid
overcounting electronic correlation energies, which are sometimes duplicated
in initial and added terms, something that can be avoided with a consistent
SAPT treatment over all intermolecular distances.
The LSDF approach describes near-range interaction energies and also long-
range electrostatic and induction forces, but not the dispersion (or vdW) forces
that arise from electronic density fluctuations and require introduction of
dynamical susceptibilities. Efforts to extend DFT to include dispersion forces
have involved recasting the long range electron–electron interactions in terms
of charge density operators and dynamical susceptibilities for each species A
and B, and approximating the susceptibilities in a variety of ways. This provides
a treatment suitable for extended systems, which contains as special cases vdW
interactions derived from multipolar polarizabilities present in the susceptibil-
ities. We consider in what follows a case where electronic charge distributions of
A and B do not overlap, avoiding electronic exchange or transfer, and only inter-
act through Coulomb forces.
Dispersion energy functions of intermolecular and intramolecular position
variables can be combined with near-range functionals obtained from DFT
and containing all energies (electrostatic, exchange, near-range correlation,
long-range induction) except for the vdW dispersion energy. This is a very
active area of research where several DFT-plus-vdW treatments have been pro-
posed and tested by comparison with accurate results from CCSD(T) or SAPT
calculations for interaction potentials V(R, Q). Here the focus is on methods
applicable to large classes of chemical systems, described as many-atom struc-
tures A and B containing sets of atoms {a} and {b} with their centers of mass at a
relative distance R in a body-fixed reference frame, and with internal variables
Q = Ra ,Rb referred to their C-of-M.
We have seen in a previous chapter that the dispersion energy can be sepa-
rately treated and calculated from the dynamical susceptibilities of the two spe-
cies, or from their multipolar polarizabilities. The aim here is to add them to the
semilocal energies while avoiding double counting of long-range interactions
and to ascertain how important these are for development of potential energy
surfaces valid at all distances. A variety of ways to do this can be termed DFT-D
treatments, with the D label signifying that dispersion energies are to be added
to DFT results. This has led recently to treatments where the semilocal DFT
energies E(sl)(R, Q) at near-range distances (rapidly decreasing at large dis-
tances) and the nonlocal dispersion (vdW) energies Edsp(R, Q) at large distances
are combined to give total energies as
sl nl
E R,Q = E R,Q + Edsp R,Q fd R,Q
where hm contains the kinetic energy of the m-th electron, its attraction energy to
positive ion cores, and the energy coming from the repulsion among the ions,
with the latter independent of electron variables. When the one-electron term
also contains some mean-field contribution from electron–electron interactions,
2
such as a functional of electron density, this must be subtracted from λH .
Differentiating the eigenstate equation H λ Ψλ = Eλ Ψλ with respect to λ and
projecting on the normalized state Ψλ, we find as in Chapter 5 that
6.4 The Density Functional Approach to Intermolecular Forces 233
2
∂Eλ ∂λ = Ψλ H Ψλ , the H–F relation. Its integral form for 0 ≤ λ ≤ 1 can be
1
written considering that Eλ = 0 = Es , the sum of single-electron kinetic and
nuclear attraction energies, and E = Eλ = 1 is the exact energy, so that the integral
form of the H–F relation is
1
2
E = Es 1 + dλ Ψλ H Ψλ
0
The exact energy E = Eλ = 1 can be expressed in terms of a two-electron density
2
ρλ 1,2; 1, 2 obtained as a function of the λ-parameter values. To derive its
2
form, one can start with the expression for H written in terms of the density
operator
ρ1 r = 1≤m≤N
δ r −rm
where the electron position is being treated as an operator and one performs
integrals over the electron variables in r. The double summation in
2
Ψλ H Ψλ must be constructed while avoiding the interaction of each elec-
tron with itself. To this effect, first write
2 c2e
H = 1 2 d 3 r1 d 3 r2 δ r 1 −rm δ r 2 −rn
r 1 −r 2 m n
m n
δ r 1 −rm δ r 2 −rn = m, n
δ r 1 −rm δ r 2 −rn
−δ r 1 −r 2 m
δ r 1 −rm
m n
δ r 1 −rm δ r 2 −rn = ρ 1 r 1 ρ 1 r 2 − δ r 1 −r 2 ρ 1 r 1
to be replaced in the double integral. Its expectation value in the state Ψλ can
2
also be obtained from the two-electron density function ρλ 1, 2; 1,2 as
previously done for λ = 1, with the short-hand notation 1 = r 1 , ζ1 including
space and spin variables. For the spin-independent electron–electron interac-
tion energy, a summation over spin variables can be carried out and the
expectation value is found to contain the two-electron pair density
2 2
Pλ r 1 ,r 2 = ζ1 , ζ2 ρλ 1, 2; 1, 2 in
2 c2e 2
Ψλ H Ψλ = 1 2 d 3 r1 d 3 r2 P r 1 ,r 2
r 1 −r 2 λ
2
Pλ r 1 ,r 2 = Ψλ ρ 1 r 1 ρ 1 r 2 Ψλ
234 6 Many-Electron Treatments
− δ r 1 −r 2 Ψλ ρ 1 r 1 Ψλ
From this and the integral form of the H–F relation, it follows that
1
1 c2e 2
E = Es 1 + dλ d 3 r1 d 3 r2 Pλ r 1 ,r 2
2 0 r 1 −r 2
This can be compared to the Hartree energy that contains the classical elec-
tron–electron interaction as
1
1 1 c2e 1
EH = Es 1 + dλ d 3 r1 d 3 r2 Pλ r1 Pλ r 2
2 0 r 1 −r 2
1
where Pλ r = Ψλ ρ 1 r Ψλ . The difference E − EH = Exc is the exchange-
correlation energy of DFT, which is therefore given by
1
1 c2e 2 1 1
Exc = dλ d 3 r1 d 3 r2 Pλ r 1 ,r 2 −Pλ r 1 Pλ r2
2 0 r 1 −r 2
Further introducing the density fluctuation operators
1
Δλ ρ 1 r = ρ 1 r − Pλ r
2
and the spatial correlation ΔPλ r 1 ,r 2 , one finds the equivalent expression
1
1 c2e 2 1
Exc = dλ d 3 r1 d 3 r2 ΔPλ r 1 ,r 2 −δ r 1 −r 2 Pλ r1
2 0 r 1 −r 2
2
ΔPλ r 1 ,r 2 = Ψλ Δλ ρ 1 r 1 Δλ ρ 1 r 2 Ψλ
2
so that ΔPλ r 1 ,r 2 is the spatial correlation function of two density
fluctuations.
The next step in the proof of the ACDF theorem relates the correlation of
fluctuations of a pair of electrons to the dynamical susceptibility of the pair
2 2 1
[66]. The function Pλ, xc r 1 ,r 2 = ΔPλ r 1 ,r 2 −δ r 1 −r 2 Pλ r 1 can be re-
expressed in terms of a dynamical susceptibility χ λ, xc r 1 ,r 2 ; ω of the many-
electron system, dependent on the frequency ω, as an integral over an imagi-
nary-valued argument iω, to obtain an expression used within DFT [52, 53,
63, 67, 68].
The relation follows from response theory applied to the calculation of
changes in ρ 1 r,t induced by an external electric potential ϕext r , t starting
at time t = 0 and null before then, as it couples to the density with energy
1
H ext t = d 3 r ϕext r ,t ρ 1 r ,t and with the system initially in its ground
electronic eigenstate 0λ of the Hamiltonian operator H λ . This has been done
6.4 The Density Functional Approach to Intermolecular Forces 235
t
Δλ ρ 1 r,t = d 3 r dt χ λ r, t; r ,t ϕext r ,t
0
χ λ r,t; r ,t = − i ℏ 0λ ρ 1 r,t ρ 1 r ,t −ρ 1 r , t ρ 1 r, t 0λ
χ λ r, r ; ω = 2 ℏ ℐm 0λ Δλ ρ 1 r,ω Δλ ρ 1 r ,0 0λ
+ ∞ +
showing that χ λ r, r ; ω = − ∞ dτ χ λ r, r ; τ exp iωτ is the scaled dynam-
ical susceptibility, related to the transform of the response through
+ ∞
χ λ r, r ; ω = 0 dω χ λ r, r ; ω−ω θ ω . Therefore the dynamical
response χ λ r, r ; ω provides both total energies and also the time evolution
of the electronic density.
236 6 Many-Electron Treatments
as needed. This relation can alternatively be derived from the analytical proper-
ties of χ λ r, r ; iω in the complex frequency plane and its integration in a
closed path [66].
The final expression for exchange correlation is then
∞
1 1
c2e ℏ
Exc = dλ d 3 r1 d 3 r2 − dω χ λ r 1 ,r 2 ; iω
2 0 r 1 −r 2 π 0
1
− δ r 1 −r 2 Pλ r1
This allows calculation of Exc from the dynamical susceptibility, and if the lat-
1
ter is constructed as a functional of the electron density ρλ r = Pλ r , then a
formally exact expression has been obtained for the functional Exc ρλ r .
Applications to the calculation of intermolecular forces between species A
and B at relative distance R require a knowledge of χ λ and Exc over each distance
R and in particular for large R. In this case, it is known that the dispersion energy
Edsp(R) is given in terms of susceptibilities χ S r, r ; iω of S = A, B by a Casi-
mir–Polder expression, and Exc R; ρλ r must correctly give Edsp(R) at large
distances. An important distinction in the two asymptotic forms (the one above
and Casimir–Polder’s) is that each spatial integral in the above Exc extends over
all space including both species, while the spatial integrals in the Casimir–
Polder expression extend only into nonoverlapping regions containing A or
B. The connection can be done introducing localized basis functions around
A and B, to partition the spatial integrals in Exc. This is presented in some detail
in the next chapter on extended systems.
6.4 The Density Functional Approach to Intermolecular Forces 237
The connection of the ACDF relation with DFT was pointed out and used in
early molecular calculations [52] and reviewed [53] some time ago and has been
the basis for the vdW–DF formulation of density functional, which has provided
many results in applications to complexes involving molecules and surfaces,
adsorbates, and their potential energy changes with distances [55, 63, 69].
One such result for two interacting benzene molecules in a complex with par-
allel planar structures is shown in Figure 6.5. Here the semilocal (near-range)
energies were obtained from generalized gradient density functional as shown,
and the vdW interaction was added as described in a vdW–DF treatment.
1 1
The factorization Δλ ρ 1 r = ρ 1 r − Pλ r = Pλ r Δλ φ r , which intro-
duces a relative density fluctuation operator Δλ φ r at all distances R between
2 1 2 1
A and B, leads to the expression ΔPλ r 1 , r 2 ; R = Pλ r 1 Φλ, xc r 1 , r 2 ; R Pλ r2
2
where the two-electron function Φλ, xc
r 1 , r 2 ; R accounts for exchange correla-
tion arising from density fluctuations at each relative distance R. This factori-
zation can be used to bring Exc to a form similar to the one developed to
account for long-range vdW interactions in terms of local polarizabilities,
5
CCSD(T)
4 MP2
vdW-DF
3 GGA(revPBE)
Interaction energy (kcal mol−1)
GGA(PW91)
2
–1
–2
–3
–4
3 3.5 4 4.5
Separation (Å)
Figure 6.5 Interaction energy versus distance between the planes of two benzene molecules
in a “sandwich” conformation. Source: from Ref. [63]. Reproduced with permission of Institute
of Physics.
238 6 Many-Electron Treatments
but here we have an expression that is applicable to all distances and contains
both semilocal and nonlocal contributions to the energy. Several treatments
have been based on this expression, which can in principle provide a seamless
connection between the long-range dispersion energy given by a Casimir–
Polder integral, and semilocal DFT energies for intermediate and short ranges
[55, 65, 67]. This is presently an active area of research, and the following chap-
ters present some related aspects relevant to interaction energies between
extended systems and at solid surface.
with fractional occupation numbers wd, j of K–S orbitals, instead of the previous
occupation numbers of 1 or 0. These procedures provide average energies as
functions of interatomic distances, and they can be extended to include pertur-
bation energies arising from external fields. Properties relevant to intermolec-
ular forces, such as electric dipoles and polarizabilities, energy gradients, and
vibrational frequencies can then be obtained by taking derivatives of perturbed
energies with respect to applied fields [4].
An important concern in DFT treatments of intermolecular forces relates to the
appearance of electron self-interaction errors resulting from approximations in
density functionals. In a wavefunction treatment of a many-electron system at
the level of the Hartree–Fock approximation, it becomes clear that terms in
the HF one-electron potential energy operator, in the notation of Section 6.1,
occ ∗
v HF 1 = J 1 −K 1 = k
d 1 ψk 1 v 1 , 1 1 − t1 1 ψ k 1
HF
containing the Coulomb and exchange energy operators vCoul = J and
HF
vx = − K partly cancel so that an electron in an occupied orbital, generated
by HF or KS methods, does not interact with itself, and also that asymptotically
HF HF
for large r1, vCoul 1 ψ l 1 ≈ ψ l 1 Ne c2e r1 while vx 1 ψ l 1 ≈ −ψ l 1 c2e r1 , for
occupied ψ l, with the sum correctly containing Ne − 1 c2e r − 1 . But when the
exchange-correlation functional Exc[ρ] is approximated, that cancellation does
not occur because the one-electron DFT exchange potential vx(1) = δEx/δρ(1)
does not cancel the Coulomb self-interaction in the DFT one-electron potential
energy vCoul(1). As a consequence, the KS one-electron potential energy
v KS r = vCoul r + vx r + vc r deviates from the correct Ne − 1 c2e r − 1 behav-
ior at large distances and is unphysical near each ion core. This affects the values
of molecular potential energies for dissociation and charge transfer and proper-
ties like energy-band gaps in semiconductors and polarization interactions in
molecular crystals. Calculations of potential energy surfaces must be done using
240 6 Many-Electron Treatments
functionals that at least decrease the self-interaction error and preferably lead to
correct one-electron potential functions over all distances.
Another complication in the application of DFT to intermolecular forces
arises from the continuous dependence of the energy on the electron density
ρ r , which integrates to the total number of electrons Ne for two interaction
species A and B. When these interact and exchange electrons, as in the disso-
ciation of a compound AB into A+ + B–, the density must split at large distances
as ρ r ≈ ρ A r + ρ B r with each term integrating to the correct number of
A B
electrons, Ne and Ne . However, this is not assured in DFT, where a func-
tional of the AB pair density may lead to asymptotic densities integrating to
S
noninteger values Ne ± δNe of electrons and to unphysical charges in the frag-
ments S = A and B. This problem is related to the absence of discontinuities in
functionals as electron densities are changed, since the energy functionals
change continuously with electronic density variations, while physically one
must instead have sharp energy changes for varying integer number of electrons
in each fragment. This may lead to inaccurate calculations for electron ioniza-
tion energies and for proton affinities.
At a more formal level, the appearance of fractional numbers of electrons when
bonds are broken can be reinterpreted by means of an ensemble DFT (or EDFT),
where a fractional number N is considered to result from a weighted average of
energy functionals for two species with integer number of electrons N and N + 1.
The ensemble density is chosen as ρ = 1− ν ρN + νρN + 1 , with an interpolating
variable ν obtained from a variational procedure [76]. This approach provides a
foundation for chemical reactivity theory with a consistent use of chemical reac-
tivity indices [43] for species with physically correct integer number of
electrons.
The heavily parametrized functionals already mentioned in connection with
DFT calculations, including dispersion energies, have also been constructed to
minimize errors from electronic self-energy interactions and from continuous
variations of the number of electrons, and provide an alternative treatment giv-
ing more accurate dissociation energies, proton affinities, reaction barriers, and
vibrational constants [59, 77].
Two approaches have been developed to correct for self-interaction and to
impose the correct asymptotic limit for the electron potential energy function.
They are the introduction of hybrid functionals containing the exact exchange
potential energy, and treatments based on functionals of orbitals obtained from
an optimized effective potential (or OEP) with the correct asymptotic behavior.
Here they are briefly reviewed as they relate to molecular interactions.
Hybrid exchange-correlation functionals containing a mixture of DFT and
HF exchange functionals take the form
with three coefficients a chosen to fit sets of known results, and LSD and
GGA signifying the local spin-density and generalized gradient approximations.
They were introduced and tested for thermodynamically relevant properties
like atomization energies, ionization and proton affinity energies [46]. This
approach was further developed to avoid fitting parameters [78] and to include
a dependence of the functional on the electronic kinetic energy density [48]
again free of adjustable parameters. This is potentially more useful, for the cal-
culation of intermolecular forces, than using parametrized functionals with
parameters fit to properties of molecules around equilibrium interatomic posi-
tions, insofar as in principle parameters are physically expected to change with
intermolecular distances.
HF
The calculation of Ex can be very demanding for extended systems such as
solids and biomolecules. An efficient procedure has been developed to decrease
HF
the computational demands for Ex . The electron–electron Coulomb interac-
tion is separated for this purpose into short-range and long-range e–e dis-
tances as
1 1− erf ωr erf ωr
= +
r r r
in terms of the error function erf, which goes from zero at r = 0 to one at large
distances with a slope controlled by the parameter ω, so that the first term to the
right is short-range (SR) and the the second term is long-range (LR). The semi-
local (sl) exchange energy can be separated into ωSR and ωLR terms and can be
HF , SR HF , SR sl, ωSR
combined with Ex and a sl correlation as Exc = aEx + 1 − a Ex +
sl, ωLR sl
Ex + Ec [47]. Introducing a screened electron–electron Coulomb interac-
HF , SR
tion at short distances, with a functional Ex containing the HF exchange
only at short electronic ranges (SR) where the density is larger, while keeping the
DFT exchange at long ranges (LR) leads to computational times shorter than for
other hybrid functional treatments, and provided good energy gaps for semi-
conductors and molecular crystal properties [79].
This mixture of HF and sl exchange energies can give good results for bond
energies and lengths around equilibrium interatomic distances in solids, but it
does not assure a correct − c2e r −1 asymptotic form of the KS one-electron vx r
potential. The correct asymptotic form, or electronic long-range corrected (LC)
behavior, can be recovered modifying hybrid functionals so that the long-range
exchange energy functional is equal to the Hartree–Fock expression,
with [49–51]
asymptotically only the exact HF potential. This is important for treating inter-
molecular forces when electron transfer is present. The choices of exchange and
correlation functionals to be used in the above expression can be made based on
needs for accuracy and efficiency, and need not be of semilocal form, with the
above parameter a chosen to reproduce known sets of accurate values.
An alternative general treatment is to recast the theory for a system with N
electrons introducing a functional E ψ j r, ζ of a set of j = 1 to N spin-
orbitals ψ j describing electrons moving in an effective potential v r , with
the orbitals given as functionals of v r . This can be derived from the relation
δE δv r = 0, and optimized by energy minimization, to obtain an optimized
effective potential v OEP r . This was done in early work to calculate the
OEP for the Hartree–Fock functional, and from it energies for atoms [80,
81]. It has been the subject of recent developments and applications to mole-
cules [82], some of which are covered here as they can be applied to intermo-
lecular forces.
Given an accurate many-electron wavefunction, it is possible to calculate
from it ρ r and v r to compare their form with results from DFT functionals.
The deviation of v KS r from a physically correct one-electron potential v r
can be drastic. The OEP procedure instead gives a new v OEP r , which is more
physical and accurate, allowing calculation of more reliable PESs [83].
A general derivation of the equations for a possibly spin-dependent
v OEP r,ζ starts by choosing an explicit functional E OEP
ψ j r,ζ of spin
orbitals, constructed from a wavefunction treatment. A suitable choice is the HF
occ occ
functional E HF
ψ j r,ζ = j j h j + j<k jk v Coul 1 − t jk , in
the notation of Section 6.2 with the index j = (p, σ) containing both orbital
and spin quantum numbers. More accurately, an extension containing electron
correlation can be constructed from many-electron perturbation theory, so that
variation of the total energy gives a single-electron reference potential
vs r, ζ ≈ − ce 2 r − 1 , with the proper asymptotic form. Using the chain rule for
functional differentiation, with the notation x = r,ζ [84],
δE δE OEP δψ j x
= dx +c c =0
δvs x j δψ j x δvs x
and the differential in the second factor to the right can be obtained from the
equation
ℏ2 2
− ∇ + vs x ψ j x = εj ψ j x
2m
6.5 Spin-Orbit Couplings and Relativistic Effects in Molecular Interactions 243
when adding a perturbation δvs(x), in terms of the equation resolvent (or elec-
tronic Green function) Gj(x, x ) giving δψ j(x ) = dx Gj(x , x)δvs(x) . The first
differential factor can be expressed as δE(OEP)/δψ j(x ) = ψ j(x )∗uj(x ) with the
orbital potential energy uj(x ) known by initial construction. This leads to an
OEP
integral equation for the potential vs x which can be solved when this is
expanded in a basis set, or is taken to be a functional of the electron density
ρ(x) as in the HK formalism. In this case, it can be decomposed into its Coulomb
electronic, plus external, plus exchange-correlation terms as
δ Exc δψ j x δvs x
OEP OEP
δ Exc
vxcOEP x = = dx dx +c c
δρ x j δψ j x δvs x δρ x
eff
by a sum of one-electron terms with an effective nuclear charges Zu obtained
from a mean-field approximation and fit to atomic fine structure levels that
include atomic spin-orbit energies,
eff
α2 Nel Nat Zu c2e Nat u
H SO liu si = H SO
2 i=1 u=1 riu 3 u=1
with liu =r iu ×pi the orbital angular momentum of an electron in the field of a
nucleus, giving a sum of atomic terms. Each electronic term liu si does not com-
mute with the orbital and spin components li, ξ or si, ξ ,ξ = x,y,z, but it does com-
mute with their sum, the total electronic angular momentum components
ji, ξ = li, ξ + si, ξ for nucleus u so that its quantum numbers (j, mj) can be used
to label the q-th atomic spin-orbital located at nucleus u as q j mj u, a function
of electron position and spin variables r,ζ . For relativistic effective atomic
REP
pseudopotentials ΔUu, l r , with l an atomic orbital quantum number, chosen
to depend only on the radial distance to each nucleus u, one finds that
u REP
H SO = l ΔUu, l r l j mj l j mj u provides a useful parametrization for
potential energy calculations [100]. Reference [102] analyzes how these pseudo-
potentials are related to the detailed Breit–Pauli Hamiltonian in molecular
interactions. Insofar as the pseudopotentials are obtained from expectation
values with many-electron wavefunctions, they are found to depend on the
atomic conformation of the AB pair, with atomic positions Q = (R, Q ) appearing
REP
as parameters in ΔUu, l r; R,Q . In the limit of large R, one finds that
A B
H SO R,Q ≈ H SO QA + H SO QB as needed.
References
1 Levine, I.N. (2000). Quantum Chemistry, 5e. New York: Prentice-Hall.
2 McWeeney, R. (1989). Methods of Molecular Quantum Mechanics, 2e. London,
England: Academic Press.
3 Atkins, P.W. and Friedman, R.S. (1997). Molecular Quantum Mechanics.
Oxford, England: Oxford University Press.
4 Jensen, F. (2001). Introduction to Computational Chemistry. New York: Wiley.
5 Szabo, A. and Ostlund, N.S. (1982). Modern Quantum Chemistry. New York:
Macmillan.
6 Woodward, R.B. and Hoffmann, R. (1970). The Conservation of Orbital
Symmetry. New York: Academic Press.
7 Pearson, R.G. (1976). Symmetry Rules for Chemical Reactions. New
York: Wiley.
8 Turro, N.J. (1978). Modern Molecular Photochemistry. Menlo Park, CA:
Benjamin-Cummings.
9 Lowdin, P.O. (1962). Normal constants of motion in quantum mechanics
treated by projection techniques. Rev. Mod. Phys. 34: 520.
10 Lowdin, P.O. (1964). Angular momentum wavefunctions constructed by
projection operators. Rev. Mod. Phys. 36: 966.
11 Lowdin, P.O. (1962). Studies in perturbation theory IV. Solution of eigenvalue
problems by a projection operator formalism. J. Math. Phys. 3: 969.
12 Messiah, A. (1962). Quantum Mechanics, vol. 2. Amsterdam: North-Holland.
248 6 Many-Electron Treatments
13 Micha, D.A. (2017). Quantum partitioning methods for few-atom and many-
atom dynamics. In: Advances in Quantum Chemistry, vol. 74, 107. London
(England): Elsevier.
14 Lowdin, P.O. (1959). Correlation problem in many-electron quantum
mechanics. I. Different approaches and some current ideas. In: Advances in
Chemical Physics, vol. II, 207. New York: Interscience.
15 Hehre, W.J., Radom, L., Schleyer, P.V., and Pople, J.A. (1986). AB INITIO
Molecular Orbital Theory. New York: Wiley.
16 Claudino, D., Gargado, R., and Bartlett, R.J. (2016). Coupled-cluster based basis
sets for valence correlation calculations (Erratum in vol. 145, p. 019901-1). J.
Chem. Phys. 144: 104106–104101.
17 Claudino, D. and Bartlett, R.J. (2018). Coupled-cluster based basis sets for
valence correlation calculations. New primitives and frozen atomic natural
orbitals. J. Chem. Phys. 149: 064105–064101.
18 Chalasinski, G. and Szszesniak, M.M. (2000). State of the art and challenges of
the ab initio theory of intermolecular interactions. Chem. Rev. 100: 4227.
19 Ashcroft, N.W. and Mermin, N.D. (1976). Solid State Physics. London, England:
Thomson.
20 Martin, R.M. (2004). Electronic Structure: Basic Theory and Practical Methods.
Cambridge, England: Cambridge University Press.
21 Schaefer, H.F. (1972). The Electronic Structure of Atoms and Molecules.
Reading, MA, USA: Addison-Wesley.
22 Schaefer, H.F. (1977). Methods of Electronic Structure Theory. Volume 3: Modern
Theoretical Chemistry (eds. W.H. Miller and H.F. Schaefer). New York: Plenum.
23 Roos, B.O. (1987). The complete active space self-consistent field method and
its applications in electronic structure calculations. In: Advances in Chemical
Physics, vol. 69 (ed. K.P. Lawley), 399. New York: Wiley.
24 Bruna, P.J. and Peyerimhoff, S.D. (1987). Excited-state potentials. In: Advances
in Chemical Physics, Part I, vol. 67 (ed. K.P. Lawley), 1. New York: Wiley.
25 Shavitt, I. and Bartlett, R.J. (2009). Many-Body Methods in Chemistry and
Physics. Cambridge, England: Cambridge University Press.
26 Bartlett, R.J. (1981). Many-body perturbation theory and coupled cluster theory
for electronic correlation in molecules. Annu. Rev. Phys. Chem. 32: 359.
27 Bartlett, R.J. and Musial, M. (2007). Coupled-cluster theory in quantum
chemistry. Rev. Mod. Phys. 79: 291.
28 Stanton, J.F. and Gauss, J. (2003). A discussion of some problems associated
with the quantum mechanical treatment of open-shell molecules. In: Advances
in Chemical Physics, vol. 125, 101. Hoboken, NJ: Wiley-Interscience.
29 Hirata, S., Fan, P.D., Auer, A.A. et al. (2004). Combined coupled-cluster and
many-body perturbation theory. J. Chem. Phys. 121: 12197.
30 Krylov, A.I. (2008). Equation-of -Motion Coupled-Cluster Methods for Open-
Shell and Electronically Excited Species (Annu. Rev. Phys. Chem, vol. 59), 433.
Palo Alto, CA: Annual Reviews Inc.
References 249
48 Perdew, J.P., Tao, J., Staroverov, V.N., and Scuseria, G.E. (2004). Meta-
generalized gradient approximation: explanation of a realistic nonempirical
density functional. J. Chem. Phys. 120: 6898.
49 Iikura, H., Tsuneda, T., and Yanai, T.H.K. (2001). A long-range corrected
scheme for generalized-gradient approximation exchange functionals. J. Chem.
Phys. 115: 3540.
50 Vydrov, O.A. and Scuseria, G.E. (2006). Assessment of a long-range corrected
hybrid functional. J. Chem. Phys. 125: 234109.
51 Chai, J.-D. and Head-Gordon, M. (2008). Systematic optimization of long-
range corrected hybrid density functionals. J. Chem. Phys. 128: 084106.
52 Gunnarsson, O. and Lundqvist, B.I. (1976). Exchange and correlation in atoms,
molecules, and solids by the spin density functional formalism. Phys. Rev. B
13: 4274.
53 Jones, R.O. and Gunnarsson, O. (1989). The density functional formalism, its
applications and prospects. Rev. Mod. Phys. 61: 689.
54 Grimme, S., Hansen, A., and Brandenburg, J.G.C. (2016). Dispersion corrected
mean-field electronic structure methods. Chem. Rev. 116: 5105.
55 Hermann, J., DiStasio, R.A.J., and Tkatchenko, A. (2017). First-principles
models for van der Waals interactionsin molecules and materials: concepts,
theory, and applications. Chem. Rev. 117: 4714.
56 Sato, T. and Nakai, H. (2009). Density functional method including weak
interactions: dispersion coefficients based on the local response approximation.
J. Chem. Phys. 131: 224104.
57 Sato, T. and Nakai, H. (2010). Local response dispersion method. II.
Generalized multicenter interactions. J. Chem. Phys. 133: 194101.
58 Podeszwa, R. and Szalewicz, K. (2012). Communication: density functional
theory overcomes the failure of predicting intermolecular interaction energies.
J. Chem. Phys. 136: 161102–161101.
59 Zhao, Y. and Truhlar, D.G. (2008). The M06 suite of energy functionals
systematic testing of four M06 functionals and 12 other functionals. Theor.
Chem. Accounts 120: 215.
60 von Lilienfeld, O.A., Tavernelli, I., Rothlisberger, U., and Sebastiani, D. (2005).
Performance of atom-centered potentials for weakly bonded systems using
density functional theory. Phys. Rev. B 71: 195119.
61 Peverati, R. and Truhlar, D. (2012). Exchange-correlation functionals with good
accuracy for both structural and energetic properties while depending only on
the density and its gradient. J. Chem. Theory Comput. 8: 2310.
62 Peverati, R. and Truhlar, D.G. (2012). Screened exchange density functionals
with broad accuracy for chemistry and solid state physics. Phys. Chem. Chem.
Phys. 14: 16187.
63 Langreth, D.C., Lundqvist, B.I., Chakarova-Kaeck, S.D. et al. (2009). A density
functional for sparse matter. J. Phys. Condens. Matter 21: 084203.
References 251
64 Bortolani, V., March, N.H., and Tosi, M.P. (eds.) (1990). Interaction of Atoms
and Molecules with Solid Surfaces. New York: Plenum Press.
65 Vydrov, O.A. and Van Voohis, T. (2010). Nonlocal van der Waals density
functional: the simpler the better. J. Chem. Phys. 133: 244103–244101.
66 Nozieres, P. and Pines, D. (1999). The Theory of Quantum Liquids. Cambridge,
MA: Perseus.
67 Dobson, J. F. and T. Gould, "Calculation of dispersion energies," J. Phys.
Condens. Matter, vol. 24, p. 073201, 2012.
68 Dobson, J.F. (2012). Dispersion (van der Waals) forces and TDDFT. In:
Fundamentals of Time-Dependent Density Functional Theory (eds. M.A.L.
Marques, N.T. Maitra, F.M.S. Nogueira, et al.), 417. Berlin: Springer-Verlag.
69 Berland, K., Cooper, V.R., Lee, K. et al. (2015). van der Waals forces in density
functional theory: a review of the vdW-DF method. Rep. Prog. Phys. 78: 066501.
70 Sun, J., Ruzsinszky, A., and Perdew, J. (2015). Strongly constrained and
appropriately normed semilocal density functional. Phys. Rev. Lett. 115: 036402.
71 Staroverov, V.N., Scuseria, G.E., Tao, J., and Perdew, J.P. (2003). Comparative
assesment of a new nonempirical density functional: molecules and hydrogen
bonded complexes. J. Chem. Phys. 119: 12129.
72 Xu, X. and Goddard, W.A. (2004). Bonding properties of the water dimer: a
comparative study of density functional theories. J. Phys. Chem. A 108: 2305.
73 Taylor, D.E., Angyan, J.G., Galli, G. et al. (2016). Blind test of density functional
based methods on intermolecular interaction energies. J. Chem. Phys. 145:
124105–124101.
74 Mermin, N.D. (1965). Thermal properties of the inhomogeneous electron gas.
Phys. Rev. 137: A1441.
75 Gross, E.K.U., Oliveira, L.N., and Kohn, W. (1988). Density functional theory
for ensembles of fractionally occupied states. I. Basic formalism. Phys. Rev. A
37: 2809.
76 Cohen, M.H. and Wassermann, A. (2007). On the foundations of chemical
reactivity theory. J. Phys. Chem. A 111: 2229.
77 Zhao, Y. and Truhlar, D.G. (2008). Density functionals with broad applicability
in chemistry. Acc. Chem. Res. 41: 157.
78 Perdew, J.P., Burke, K., and Ernzenhof, M. (1996). Generalized gradient
approximation made simple. Phys. Rev. Lett. 77: 3865.
79 Heyd, J. and Scuseria, G.E. (2004). Efficient hybrid density functional
calculations in solids: assessment of the Heyd – Scuseria – Ernzerhof screened
Coulomb hybrid functional. J. Chem. Phys. 121: 1187.
80 Talman, J.D. and Shadwick, W.F. (1976). Optimized effective atomic central
potentials. Phys. Rev. A 14: 36.
81 Krieger, J.B., Li, Y., and Iafrate, G.J. (1995). Recent developments in Kohn-Sham
theory for orbital dependent exchange-correlation energy functionals. In:
Density Functional Theory (eds. E.K.U. Gross and R.M. Dreizler), 191.
New York: Plenum Press.
252 6 Many-Electron Treatments
100 Balasubramanian, K. and Pitzer, K.S. (1987). Relativistic quantum chemistry. In:
Advances in Chemical Physics. Volume 67: Ab Initio Methods in Quantum
Chemistry Part I (ed. K.P. Lawley), 287. New York: Wiley.
101 Fedorov, D.G., Koseki, S., Schmidt, M.W., and Gordon, M.S. (2003). Spin-orbit
coupling in molecules: chemistry beyond the adiabatic approximation. Int. Rev.
Phys. Chem. 22: 551.
102 Fedorov, D.G. and Gordon, M.S. (2000). A study of the relative importance of
one and two-electron contributions to spin–orbit. J. Chem. Phys. 112: 5611.
103 Hafner, J. (2008). Ab initio simulation of materials using VASP: density
functional theory and beyond. J. Comput. Chem. 29: 2044.
104 Kleinschmidt, M., Tatchen, J., and Marian, C.M. (2002). Spin-orbit coupling of
DFT/MRCI wavefunctions: method, test calculations and applications to
thiophene. J. Comput. Chem. 23: 824.
105 Herzberg, G. (1950). Molecular Spectra and Molecular Structure I. Spectra of
Diatomic Molecules, 2e. Princeton NJ: Van Nostrand.
106 Herzberg, G. (1966). Molecular Spectra and Molecular Structure III.
Polyatomic Molecules. New York: Van Nostrand Reinhold.
255
CONTENTS
densities per unit volume and Hamiltonian operators involving space integrals
over charge density operators [1–7]. This bypasses the need for expansions of
interaction energy operators in terms of electrical multipoles of high order.
Long-range interaction energies follow from perturbation theory or from
many-body treatments in terms of response functions.
The form of the interaction energy operator between two species A and B,
now considered to be extended systems containing many atoms, has been given
in Sections 2.1 and 2.4 and in Section 3.2 in terms of charge density operators
cA r and cB r , defined as functions of positions over the whole region of inter-
actions. The two species are given by the locations of the nuclei in each one,
labeled by the index a for A and by b for B. They are then defined by the col-
lection QA = r a of atomic position vectors for A and similarly QB for B, which
are parameters in the density operators. Those density expressions are used here
in applications to extended systems such as clusters, polymers, and solid sur-
faces interacting with molecules or among themselves. The expectation values
of the density operators, for given electronic states of A and B, occupy regions of
space with a variety of shapes. Some may show cavities in which a species may be
located, so that a distance R between centers of mass may not be meaningful as a
separation distance. Two examples are shown in Figure 7.1.
We consider two species A and B with density boundaries sufficiently far away
so that electron exchange is improbable and does not affect the properties of the
system and can be omitted. In this case, electrons i = 1 to NA can be assigned to
species A, and electrons j = NA + 1 to NA + NB are assigned to species B. Their
relative distance rij remains larger than some distance d and their Coulomb
interaction remains finite, insofar electrons i and j are well separated during
the interaction of the two many-electron systems. This treatment, however,
must be reconsidered when electron exchange is relevant.
(a) (b)
B(1)
) B
B(
(2
9)
B
B(8
)
A B(3) A
)
B (7 B(
4)
B(6)
B(5
)
Figure 7.1 (a) Molecule A in a solvent B containing molecules B(n); (b) A in a cavity of a large
molecule B.
7.1 Long-range Interactions of Large Molecules 257
nu el
cA r = a
Ca δ r −r a , cA r = Ce i
δ r −r i
a sum of nuclear and electronic charge densities, and similarly for species B. The
nu
nuclear charge density cA is, for fixed nuclei, simply a function of nuclear posi-
el
tions, but the electronic charge density cA is a one-electron operator in the
space of many-electron wavefunctions.
The Hamiltonian of species A and B can be written in terms of nuclear and
electronic charge densities as
e en ee
HA = K A + HA + HA
el nu
e ℏ2 en 1 c r cA r
KA = − ∇2 , H A =
i i
d3 r d3 r A
2me 4πε0 r−r
el el el
ee 1 c r cA r −δ r − r cA r
HA = d3 r d3 r A
4πε0 r−r
where the form of the electron–electron Coulomb energy accounts for the
absence of interaction of each electron with itself. A similar expression applies
to species B.
The Coulomb interaction energy operator for all charges in A interacting with
all those in B, but without electron exchange between them, is
Coul −1 −1
H AB QA ,QB = 4πε0 d 3 r d 3 s cA r; QA r − s cB s; QB
Coul nn n, e e, n ee
which can be expanded as H AB = H AB + H AB + H AB + H AB to display types
of interactions of nuclear and electronic charges. Insofar the nuclear charge dis-
tributions are dependent on nuclear positions for fixed nuclei, and the elec-
tronic distributions are operators dependent on the electronic position
operators, the second and third terms are one-electron operators dependent
on all nuclear positions while the last term is a two-electron operator.
nn
In more detail, H AB is just the Coulomb energy of interacting nuclear charge
distributions for A and B,
nn −1 nu −1 nu
HAB QA ,QB = 4πε0 d 3 r d 3 s cA r; QA r − s cB s; QB
258 7 Interactions Between Two Many-Atom Systems
ne nu nu
H AB = d 3 r cA r; QA φB r + d 3 s φA s cB s; QB
el el
1 c s 1 c r
φB r = d3 s B , φA s = d3 r A
4πε0 r−s 4πε0 r−s
where φB r is the electric field created by electronic charges of B at the location
of the nuclear charges of A, and similarly for φA s . This Hamiltonian is a one-
electron operator and its matrix elements can be obtained from one-electron
1 1
density operators ρA and ρB .
The electron–electron interaction is
ee −1 el −1 el
H AB = 4πε0 d 3 r d 3 s cA r r − s cB s
and does not depend explicitly on the nuclear positions. This Hamiltonian is a
two-electron operator, and calculation of its matrix elements between many-
electron states of A and B involve the two-electron density operator
2 1 1
ρAB = ρA ρB , derived from the separate many-electron states of A and B, insofar
they do not exchange electrons. The matrix elements do change with nuclear
distances.
These one- and two-electron expressions appear in the total Hamiltonian in a
form useful for calculations involving both ground and excited electronic states
for fixed atomic positions.
A
where ρ0 r is the ground state electronic density of A, and similarly with
B
c0 s for B. Its form is
AB −1 A −1 B
Eels QA ,QB = 4πε0 d 3 r d 3 s c0 r; QA r − s c0 s; QB
and can be further decomposed into Coulomb interaction energies of nuclei and
AB nn ne ee
electrons in A and B, Eels = E0, 0 + E0, 0 + E0, 0 .
Induction and dispersion energies given by
2
Coul
0A ,0B H AB JA ,0B
AB
Eind QA ,QB = − A A
J 0
EJ − E0
2
Coul
0A ,0B H AB 0A ,KB
− K 0 B B
EK − E0
2
Coul
0A ,0B H AB JA ,KB
AB
Edsp QA ,QB = − J 0 K 0 A A B B
EJ − E0 + EK − E0
A
contain transition values c0J r = 0A cA r JA of the charge density opera-
tors for A, and similarly for B, with
A A B B
c0J r = cnuA r δ0J + Ce ρ0J r , c0K s = cnuB s δ0K + Ce ρ0K s
where contributions from nuclei and electrons have been separated. Here
A el
Ce ρ0J r = 0A cA r JA is an electronic charge excitation, and similarly
for B, obtained from their many-electron states. The second-order energies
Eind = E0, 0 AB∗ + E0, 0 A∗ B and Edsp = E0, 0 A∗ B∗ involve integrals like
AB ind ind AB dsp
Coul −1 A −1 B
0A , 0B H AB JA ,0B = 4πε0 d 3 r d 3 sCe ρ0J r r − s c0 s
A B
many-atom systems may involve degenerate ground states all of the same
energy, requiring modified treatments for perturbation of degenerate states.
In addition, large systems may have vanishing or very small excitation energies
A A B B
EJ − E0 and EK − E0 , which lead to large perturbation energy corrections
and require alternative treatment. These problems can sometimes be avoided
introducing static and dynamical susceptibilities functions of excitation fre-
quencies. Instead of using the molecular excited states, the summations in sec-
ond-order perturbation terms can best be calculated as integrals over their
frequencies with susceptibilities obtained from a variety of electronic basis sets.
A el
where Ce ρel, JJ r = JA cA r JA is an electronic charge density fluctuation,
and similarly for B, obtained from their many-electron states. Transition den-
A
sities ρel, JJ can be obtained from one-electron transition density functions
1
ρJJ 1,1 , for states J and J of A with NA electrons. They are defined as an exten-
sion of the single-state one-electron density functions in Chapter 6, by means of
,NA ∗ ΦJ 1 , 2,
1
ρJJ 1; 1 = NA d 2 d NA ΦJ 1, 2, ,NA
with (1) standing for space and spin electron variables r 1 , ζ1 = x1 , and so that
A 1
ρel, JJ r = ρ
ζ JJ
r,ζ; r ,ζ
−1 2
= (NA!)−1/2 det[ xn j ] and ΦJ 1 ,2 , …, NA = NA det xn j con-
structed from a basis set of N orthonormal SOs, one finds
1 1 n = NA
ρJJ x1 ; x1 = d 2 d NA det j xn xn j
NA −1 n=1 2 = 2 , …, NA = NA
with expansion coefficients PJJ μ,μ determined by the form of the many-
electron states J. The basis functions ξμ(r ,ζ can be chosen as atomic spin orbi-
tals (ASOs) to form MSO-LCAO orbitals, or as delocalized functions.
Choosing a basis set of ASO μ for A, with overlaps ξμ ξμ = Sμμ , and sim-
1
ilarly for ρKK with AOs ξν(r , ζ for B, it is possible to perform a population anal-
ysis to decompose the densities into terms giving atomic populations and atom–
atom bond orders. Labeling ASOs centered at the nuclei a of A, with orbital and
spin quantum numbers (n, l, m, ms) by μ = (a, n, l, m, ms), atomic components
PJJ a of transition densities at each center a are obtained adding all terms
where μ and μ contain only a. This can be quite generally done introducing pro-
jection operators
Pa = μ a μ a
ξμ S − 1 μμ
ξμ
which satisfy Pa Pa = δaa P a and add up to the identity. Applied to the density
operator, they select its components localized at centers a in A and give
1
ρJJ = P
a JJ
a Pa + a a
P a PJJ a,a P a
A second choice for the basis functions ξμ(r ,ζ is provided by plane waves
times spin functions Ω − 1 2 exp −iqn r ϑ ζ with quantized wavevectors satis-
fying periodic boundary conditions in a large volume Ω. This is a convenient
choice when the many-atom systems are very large, when they display structural
periodicity, or when electrons are delocalized, and is further considered in what
follows.
A A B B
with a = EJ −E0 and b = EK − E0 and introduce the dynamical susceptibility
A A A A ∗
c0J r E0 − EJ c0J r
A
χ0 r , r ; ω = −2 J 0 2
A A 2
E0 − EJ − ℏω
∞ A B
A∗ B∗ ℏ χ0 r, r ; iω χ 0 s, s ; iω
Edsp =− d3 r d3 r d3 s d3 s dω 2
2π 0 4πε0 r−s r −s
A
A E0 − H A
χ0 r, r ; iω = −2 0A cA r P A 2 P A cA r 0A
A 2
E0 −H A + ℏω
which can now be calculated with an alternative many-electron basis set. Prop-
erties of this susceptibility and ways to calculate it and to provide upper and
lower bounds to it as a function of the variable u have been the subject of
detailed research [10–12]. The procedure is similar to the one followed in
Section 3.4 for molecular interactions. The static susceptibilities in the induc-
tion energy are found for ω = 0.
Two particularly simple forms are obtained from a closure approximation,
A
and from an interpolation of χ 0 iω between its values at ω = 0 and ω ∞.
A
In a closure approximation, the operator E0 − H A P A in the dynamical sus-
A A
ceptibility is replaced by E0 − Eexc P A, which contains an excitation energy
A
Eexc to be treated as an adjustable parameter chosen to reproduce the value of
264 7 Interactions Between Two Many-Atom Systems
2
A A A A A
with the frequency in the factor Fexc ω = E0 − Eexc E0 −Eexc
2
+ ℏω , which multiplies the charge–charge correlation function
A
Gcc r, r for the ground state expectation value of charge densities at two posi-
tions. This contains one- and two-electron operators averaged only over a
A A
known many-electron ground state. The choice Eexc = E1 , the first excited
state of A, gives an upper bound to this approximate susceptibility. The corre-
lation function provides the static susceptibility through
A A A A A
χ 0 r, r ; 0 = 2Gcc r, r E0 − Eexc = χ 0 r, r , with the parameter
A A
ℏω A = E0 −Eexc to be extracted from calculations or measurements.
B
With a similar approximation for χ 0 iω , it is possible to estimate the induc-
tion and dispersion energies of the AB pair. Induction energies are given by
interactions between permanent charges and expectation values of charge–
charge fluctuations, however involving just the ground states of A and B, and
information on their energies for AA and BB. By analogy with the treatment
of interaction energies for two molecules in the dipole approximation, it follows
that here we can also introduce a combination rule giving the dispersion energy
of AB in terms of the values for AA and BB. By analogy with the treatment for
two interacting molecules leading to combination rules for the dispersion
energy [13], here for two large many-atom systems, it is convenient to introduce
A A A A
the limit for ω ∞, χ 0, exc r, r ; iω ≈ 2 E0 − Eexc Gcc r, r ℏω 2 =
2 A
ω A ω χ 0 r, r ; 0 . The limits at ω = 0 and ω ∞ can be combined in
an interpolated approximate form
2 −1
A A ω
β0 r, r ; ω = χ0 r, r 1+
ωA
A A
where β0 r, r ; ω = χ 0 r, r ; iω . The parameter ω(A) can be obtained from
A∗ A∗ , containing a calculated Gcc r, r
dsp A
E0, 0 and written in terms of
B∗ B∗ , these two para-
A dsp
β0 ω . Doing the same for ω (B)
obtained from E0, 0
A B
meters and the correlation functions Gcc r, r and Gcc s,s can be used
A∗ B∗ from a combination rule, here for extended systems.
dsp
to obtain E0, 0
7.2 Energetics of a Large Molecule in a Medium 265
AB AB B
E0, slt QA ,QB = E0 QA ,QB + E0, slv QB
The terms to the right can be separately considered and are simpler to con-
struct when electrons are not exchanged between A and B, and they are also
sufficiently separated so that short-range repulsion does not distort them. This
is first done as follows.
The electrostatic, induction, and dispersion energies for A in a given elec-
tronic state 0A interacting with a species B in a general electronic state KB
els ind dsp
can be obtained from potential energy operators V AB ,V AB , and V AB arising
from the charges in A interacting with electric potentials originating in B. The
els
electrostatic potential energy operator V AB r; QA , QB involves the charge
density operator c A r; QA of A multiplying the electric scalar potential
els
φB r; QB resulting from the stationary charge distribution of B in its ground
ind
state 0B . Additional potential energy operators V AB r, r ; QA ,QB , and
dsp
V AB r, r ; QA , QB can be constructed from the charge distributions
cA r; QA and cA r ; QA at two locations linked by reaction potential operators
ind dsp
φAB r, r ; QA ,QB and φAB r, r ; QA ,QB generated by density fluctuations
when B is excited into states KB . These reaction potentials can be obtained
as shown below, or from time-dependent response theory [6].
This treatment is convenient when describing a large solute molecule A in a
medium B composed of many solvent molecules treated as molecular frag-
ments [14], or when B is a medium involving many atoms in a complex structure
described by distributed multipoles and polarizabilities [15]. The external electric
potentials generated by B can then be considered as one- and two-electron opera-
tors to be added to the Hamiltonian H A of A in a treatment requiring a many-
electron description of A but treating B only in terms of its electric multipoles
and polarizabilities. The energy of the solute A in its ground electronic state in a
medium provided by B is then obtained to first order in the potentials as
which is valid here only provided the electronic charge distributions of species A
and B in their ground and excited states do not overlap. The electric potentials
can be derived comparing the above terms with the known forms of terms in
AB A AB AB AB
E0 QA ,QB = E0 QA + Eels QA ,QB + Eind QA , QB + Edsp QA , QB . The
ext
external potentials V AB , with ext = els, ind, dsp, can also be used in a pertur-
bation treatment of states of A to obtain more accurate interaction energies
when they are large compared to internal energies in A.
From the previous ground state first-order perturbation energy, one finds
AB els
Eels QA ,QB = d3 r 0A cA r; QA 0A φB, 0 r; QB
B
els 1 c s; QB
φB, 0 r; QB = d3 s 0
4πε0 r−s
els els
V AB r; QA ,QB = cA r; QA φB, 0 r; QB
which is just the classical electrostatic interaction between the ground state
A
charge distributions c0 r = 0A cA r; QA 0A and the electrical potential
els B
φB, 0 created by c0 s = 0B cB s; QB 0B .
The second-order energies lead to forms of the desired induction and disper-
sion electric potentials generated by B, after rewriting
2
Coul Coul Coul
0A ,0B H AB JA ,KB = 0A , 0B H AB JA ,KB JA , KB H AB 0A ,0B
0 JA JA 0 KB KB
RA = − J 0 A A
, RB = − B B
,
EJ − E0 K 0
EK − E0
ind 1 cB s 0 cB s
φAB∗ , 0 r, r = 0A 0A d3 s d3 s 0B RB 0B
4πε0 2
r −s r−s
7.2 Energetics of a Large Molecule in a Medium 267
dsp 1 cB s 0 cB s
φAB, 0 r, r = d3 s d3 s 0B RAB 0B
4πε0 r −s r−s
AB dsp
Edsp QA ,QB = d 3 r d 3 r 0A V AB r, r ; QA ,QB 0A
B
and ρel, 0K s . However, it is then necessary to make sure there are no contribu-
tions from overlapping electronic states of A and B, which would lead to small
values of r − s in integrands, and inaccurate energies.
For decreasing distances between solute and solvent, one can proceed some-
what as done when damping long-range interactions for small distances to avoid
unphysical superposition of forces. Here there is no single intermolecular dis-
tance R to construct a damping function, but the problem can be avoided repla-
−1
cing r − s with a damped charge–charge interaction function such as
−1
fd r − s r−s with
exp α r − s − 1
fd r − s ; d,α = ≈0 for r − s 0
exp α r − s − exp αd
where d is the range over which the function is damped as r − s decreases, and
with α giving the slope of the transition region. This function goes to 1.0 for
large r − s d and preserves the values of the electric potentials at these large
distances. The calculation of energies with this damped function requires
numerical evaluation of the integrals over r and s.
AB ctr
E0, ctr QA ,QB = 0A V AB 0A , with these energies obtained in suitable
approximations.
A reliable treatment of exchange-repulsion and charge transfer at short dis-
tances can be obtained from a self-consistent field (or Hartree–Fock) treatment
of the electronic structure of the pair AB, insofar the short-range region of elec-
tronic interactions is dominated by Coulomb and exchange electron energies.
In the notation of Chapter 6, the H–F energy of species A is
A occ occ
E0 HF = jA
jA h jA − ϵjA + jA < kA
jA k A v 1 − t jA k A
in terms of MSOs jA and kA of A, and similarly for B. The H–F energy of the
pair AB is given as
AB occ occ
E0 HF = j
j h j −ϵj + j<k
jk v 1 −t jk
treatments of very large fragments, with sizes that can be increased to obtain
convergence in total energy values.
Similar aspects have been treated in the previous section on the interactions of
a molecule in a solution, as they relate to long-range forces but without bonding.
Here more generally one is dealing with strong coupling due to bonding
between a cluster of atoms and its environment. The QM/MM treatments ori-
ginated on work dealing with dynamics and reactions in biomolecular systems
[24–26]. and has been extended in several ways, for example, introducing mod-
els with a transition region between the QM region of the fragment and the MM
region of the environment [27]. The challenge here is to identify the boundaries
of the regions and to allow for changes of the electronic distribution in the QM
region and of the force field in the MM region, due to electronic rearrangements
or chemical reactions within the selected QM region. Recent developments have
been reviewed as they apply to molecular complexes and to biomolecules
[28–31] and are briefly summarized here.
To be consistent with our notation, we let A indicate the selected molecular
fragment (or PS), including a set of link atoms capping bonds, and let B indicate
the surrounding medium (or SS). We show with labels QM or MM that a many-
electron treatment or force field description is being done. The total energy can
be obtained starting with the MM treatment of the whole system and correcting
it with the detailed energy change for the subsystem A,
E AB
QM MM = E AB
MM + E A
QM − E A
MM
The calculation of the MM energies are done as described in Chapter 4, with a
sum of terms for bond distance, bond angles, dihedral angles, electrostatic, induc-
tion, and dispersion energies, parametrized from independent calculations. The
A
remaining term E(A)(QM) is obtained from an electronic wavefunction Φ0 cal-
culated for the state of interest from the Hamiltonian operator H A of A. This is
called a subtraction scheme or mechanical embedding and is straightforward to
implement numerically and to improve if needed by increasing the size of the PS
region. It assumes that the force field in the SS region is unchanged as the two
regions interact, and even if electronic rearrangement occurs in A.
However, physically one knows that all three energy terms depend on all the
atomic positions and are functionals of the total electronic density of the pair
AB. Therefore, charge rearrangement or chemical reactions in A (the PS), invol-
ving bond breaking or formation, affect B (the SS) by polarizing the force field
and require some self-consistent coupling between the two regions. This is done
with an additive scheme or electrostatic embedding writing instead that
AB A B AB
E =E QM + E MM + Eint QM MM
(B)
with E (MM ) obtained from a polarizable force field in B (show as MM ) and
AB
the interaction Eint (QM/MM ) obtained from the quantal expectation value of
272 7 Interactions Between Two Many-Atom Systems
AB
the energy coupling operator between A and B for many-electron states ΦJ .
Coul
This must be calculated from the full Hamiltonian H AB = H A + H B + H AB
in
our previous notation. Here the B (or SS) region is treated as a static but polar-
ized force field so that H B = E B MM is an energy value, and the AB coupling
Coul Coul
H AB = V AB is a potential energy operator that depends on the electronic
and nuclear position variables of A, but is only a parametric function of atomic
AB
positions and polarizable charge densities in B. Therefore, states ΦJ satisfying
Coul AB AB AB
H A + V AB +E B
MM ΦJ = EJ ΦJ
st , a
with known coefficients Cμν, λ , which contain by construction information
about the molecular charge distribution. One can then derive multipole ampli-
a
tudes Qlm from expansions
a a
ρr = a, lm
Qlm χ lm r
a
where χ lm is localized at a position r a in A, and rotates as a spherical tensor.
Doing the same for another molecule B, the interaction Hamiltonian energy
a
for the AB molecular pair can be written as a bilinear expression in Qlm and
b
Qpq operators physically arising from the energy of interaction of multipoles
at r a in A with multipoles at r b in B,
a a, b
H AB = a, lm b, pq
Qlm Tlm, pq Qpqb
a, b
with coupling coefficients Tlm, pq fit to calculated energies for each separate frag-
a, b
ment pair ab and containing a damping factor so that Tlm, pq ≈ 0 as r a −r b
becomes small. This has been done and tested for many systems and requires
a
using rotational and translational properties of the functions χ lm as relative
positions and orientations of A (and B) are changed. The calculations can be
accurate provided they account for changes in values of properties going from
free fragments to fragments interacting with their neighbors, but can be labo-
rious [15].
A simpler but less general procedure decomposes the one-electron density by
means of weights generated from the electronic densities of atoms in A and B
[32]. The required two items of information are the spherically averaged elec-
tronic densities ρ a r of each free atom a appearing in the molecule A, and the
electronic density ρ A r of the whole molecule. A set of atomic weight func-
tions is defined by
a
ρfree r
wa r = a
a ρfree r
274 7 Interactions Between Two Many-Atom Systems
d3 r F r = a
d 3 r F r wa r = F
a a
for a generic function F r . The free atom densities can be corrected to account
for bondings in A using knowledge of the molecular density, by defining the
a
bonded atom densities ρbnd r = wa r ρ A r . The atomic deformation density
a a a
is given by Δρbnd r = ρbnd r −ρfree r and the molecular deformation density is
a
Δρ A r = ρ A r − a ρfree r , with both functions providing insight on the
changes due to bonding.
The weights can be used to decompose molecular properties into atomic frag-
ments and to define effective values for each fragment. For example, the physical
proportionality between a fragment electronic volume and its polarizability has
led to the introduction of an effective static dipolar polarizability for atomic
fragment a as
3
a a Ad
3
r wa r ρ A r r −r a
αeff 0 = αfree 0 3
3r ρ A r r −r a
Ad
el A A
JA cAα JA = Ce d 3 r fα A r ρJJ r = Ce ρα, JJ
and similarly for B in its many-electron basis set { KB }, that appear in the inter-
action energies, can be obtained from the many-electron functions. The matrix
A B
elements of one-electron Hamiltonian terms contain only ρα, JJ and ρβ, KK , while
A B
matrix elements of two-electron terms contain products ρα, JJ ρβ, KK . When
A
using a density basis set of localized functions, a population analysis of ρα, JJ
in terms of atomic orbitals will show that many of its terms will be small and
negligible when the location of AOs are far removed from the location of fα A r .
Coul nn
The interaction energy operator can again be written as H AB = H AB +
ne ee
H AB + H AB , with the first term given by the Coulomb repulsion of nuclear
charges (or atomic ion cores) and the other two terms decomposed into con-
tributions from electronic charge components. In particular, the electronic–
nuclear interaction Hamiltonian energy is
ne −1 el AB AB el
H AB = 4πε0 c I
α, b Aα αb
sb Cb + β, a
Ca Iaβ r a cBβ
B
AB 3 fα A r AB 3
fβ s
Iαb sb = d r , Iaβ ra = d s
r −sb ra − s
AB AB
where the Iαb Cb and Ca Iaβ integrals are electric potentials generated by
nuclei of charges Ca and Cb interacting with electronic density distributions.
276 7 Interactions Between Two Many-Atom Systems
AB 1 B
Jαβ = d 3 r d 3 s fα A r f s
r−s β
with the integral J in the second line giving the interaction of two densities. This
Hamiltonian is a two-electron operator and its calculation involves two-
AB B
electron density components ραβ = ραA ρβ derived from the many-electron
states of A and B separately, insofar they do not exchange electrons. These
one- and two-electron expressions appear in the total Hamiltonian in a form
useful for calculations involving both ground and excited electronic states for
fixed atomic positions. The choices of the density basis sets are made to limit
the number of (α, β) terms needed for convergence.
Interaction energies between A and B in their ground electronic states involve
the total charge densities
A A
c0 r = cnuA r + c f A
α α, 0 α
r
A el
containing the expectation values cα, 0 = 0A cAα 0A , and also involve
dynamical susceptibilities
A A A
χ0 r, r ; iω = α
f
α α
A
r χ0 α,α ; iω fα r
1 λ2 ξ− ξa 2
uma ξ; λ = λ2 Cm exp Hm λ ξ− ξa , m = 1 to Na
2
with Hm(x) the Hermite polynomial [34], containing displacements along direc-
tions ξ = x, y, z, at each location r a , and with optimized exponent parameters λ,
which may vary with direction to account for local anisotropy. The spatial basis
a a a a
functions umx x umy y umz z = um r with (m = mx, my, mz) can be used to
expand the charge distribution as
el
cA r = a A
c el
m a
m uma r
el
where the coefficients ca m provide a coarse-graining representation of the
charge densities, convenient for further treatment of interaction energies in a
many-atom system.
el
Oscillator functions in cA at different locations a and a are not orthonormal
a, a a a
and their overlap integrals Sm,m = um um , involving different nuclear posi-
tions, must be incorporated into calculations. These overlaps are, however, short-
ranged in (a,a ) due to the localized nature of the basis functions. The expansion
a
coefficients are obtained multiplying the above equation to the left times um and
a, a
integrating over space to obtain in matrix notation with S = Sm,m ,
a, a
S −1
a
cael m = a A m i
um ri m ,m
using the property that a product of Gaussian functions gives a new Gaussian at a
shifted position, and using translation properties of the harmonic oscillator func-
tions within A. The resulting integrals are similar to the ones extensively employed
in calculations of electronic structure with Gaussian basis sets [36]. Expansions
b b b b
for B are done similarly with a basis of functions un = unx x uny y unz z loca-
lized at r b .
Interaction energies between A and B involve oscillator functions of r and s
localized around r a and r b in each structure. Their products overlap in space
at electron positions ρ and the products of Gaussian functions for A and B con-
tain the relation
2 2
exp −α ρ −r a exp − β ρ −r b
αβ 2 2
= exp − r a −r b exp − α + β ρ −r I
α+β
int nn ne ee
operator H AB = H AB + H AB + H AB . This leads to expressions for electrostatic,
induction, and dispersion energies in terms of ground state electronic charge
el a a b
densities 0A ca m 0A = c0 m = Ce ρ0,m for A and Ce ρ0,n for B. The
dynamical susceptibility for A is a double sum over locations a and a ,
A Na Na aa a
χ0 r, r ; iω = a A m=1 a A
ua
m =1 m
r χ0 m,m ; iω um r
with small terms when the distance r a − r a is large for molecular fragments far
away, so that many of them can be discarded in calculations. A similar expres-
b
sion applies to the susceptibility of B using un s .
The two expansions replaced in induction and dispersion energy integrals
over space variables contain products of summations over (a, m) and (b, n),
and here again many terms can be discarded when r a −r b is large.
cr = n
exp iqn r cn ,
3
d r 1
cn = exp − iqn r c r = CI exp −iqn r I
Ω Ω Ω I
and the coefficients cn are used in interactions energy operators. This expansion
is common in treatments of solid state properties, and is particularly useful for
polymers, surfaces, and solids with periodic atomic structure [37]. Products of
operators like c r c r are likely to contribute little to interactions when the
phase difference qn r −qn r is large, in which case the exponential functions
in the products show fast oscillations that add up to nearly zero in integrals.
The expansion in the plane waves Ω −1 2 exp − iqm r gives for the dynamical
susceptibility
A A
χ0 r, r ; iω = m m
exp iqm r χ 0 qm ,qm ; iω exp iqm r
280 7 Interactions Between Two Many-Atom Systems
These summations can be replaced in the space integrals for induction and
dispersion energies of interaction between A and B, containing also
B
χ 0 qn ,qn ; iω , and can be transformed as done in Section 3.2.5.
The charge density operator of all the charges I in species A distributed over
space can be re-expressed in terms of its Fourier components QA k as in
cA r = I
CI δ r −r IA = m
exp iqm r QA qm
2 ∞
AB Ω 1 1 ℏ A B
Edsp = − m n 2 dω χ 0 qm ,qn ; iω χ 0 −qm , −qn ; iω
4πε0 qm qn 2 2π 0
A 2 ωJ
χ0 qm ,qn ; iω = 0 QA qm J J QA q n 0
ℏ J 0ω 2
J + ω2
A A
with ℏωJ = EJ −E0 , and similarly for B. This gives the dispersion energy as a
double sum over grid points in reciprocal space, and avoids an expansion in
multipolar components of the density.
el el
with the total density given by cA r = a A ca r . Neglecting bond–bond
couplings insofar the localized functions have little overlap, we can introduce
related bond susceptibilities χ a(ω), which (for now) can be taken as isotropic
averages. Similarly chain B is broken into components B(b) at Rb with their sus-
ceptibilities χ b(ω). The bond–bond induction energies are then
a b a
AB∗ 1 3 c0 r χ 0 s, s c0 r
Eind = − d r d3 r d3 s d3 s 2
a b 2 4πε0 r−s r −s
1
ab∗
ind
= E
a b 0, 0
2
a
where c0 r is the total (nuclear plus ground state electronic) charge density at
A∗ B . The integrals can be simplified consider-
ind
bond a, and similarly for E0, 0
ing that functions in the numerator are localized at Ra and Rb . The integrand can
be expanded in powers of the displacements r −Ra and s −Rb , keeping only
terms of zero and first order corresponding to total charge and dipole. To lowest
order, induction energies result from the charge C(a) at a multiplying the
−2 2
ab∗ = − 4πε0
b ind
static dipolar polarizability α1 0 at b, with E0, 0 C a
b
α1 0 2R4ab , as shown in Chapter 3. This is zero in the present example inso-
far the net charge of bond a is zero here. The same applies to E0, 0 a∗ b . Bond–
ind
bond induction energies appear, however, when there are permanent bond
dipoles and quadrupoles of a, and bond quadrupole polarizability at b.
The dispersion energy follows from a similar analysis leading to
A∗ B∗ 1
a ∗ b∗
dsp
Edsp = E
a b 0, 0
2
∞ a b
ℏ χ 0 r, r ; iω χ 0 s, s ; iω
a ∗ b∗ = −
dsp
E0, 0 d3 r d3 r d3 s d3 s dω 2
2π 0 4πε0 r−s r −s
Here again it is possible to expand the integrand around Ra and Rb and to keep
to lowest order the bond dynamical dipole polarizabilities, which give
E0, 0 a∗ b∗ = −C6 a∗ b ∗
dsp dsp dsp
4πε0 2 R6ab with the C6 coefficient expressed
a b
in terms of dipolar bond polarizabilities α1 ω and α1 ω as shown in
Chapter 3. These polarizabilities can be chosen to be isotropic averages in
the simplest case.
In the present example, the dispersion coefficients for bonds in the two chains
AB
are equal to a value C6 and one finds for chains with Nchn = NC − 1 bonds,
7.4 Models of Hydrocarbon Chains and of Excited Dielectrics 283
AB
A∗ B∗ C6 Nbnd a
Edsp =− 2 a=1
R−6
b = 1 ab
4πε0
with R2ab = D2 + lbnd
2
a− b 2 . The two summations can be rewritten as a sum over
a = b and a sum over k = a − b , of which there are 2(Nchn − k) at distances
1 2
D2 + lbnd
2
k2 . This gives
AB
A∗ B∗ C6 Nchn Nchn −2 2 Nchn −k
Edsp =− 2 + 3
4πε0 D6 k =1 2 k2
D2 + lbnd
This allows calculation of limits when the chain length Lchn = Nchnlbnd is much
larger or much smaller than the chain-to-chain distance D. When Nchn ≥ 10, the
quotient y = k/Nchn varies almost continuously and the summation can be
replaced by the integral Nchn dy…. Doing the integration with ρ = Lchn/D as
a parameter, one finds [38]
AB
A∗ B∗ C6 Nchn ρ ρ
Edsp, chn = − + 3 atan ρ +
4πε0 2 D6 2
4 lbnd Nchn 4 1 + ρ2
When D Lchn (or ρ 1) and the chains are distant, one simply finds
dsp AB
E0, 0 = − C6 Nchn 2
4πε0 2 D6 , for pairs of bonds with energy varying as
the inverse-six distance. When instead the chains are close and D Lchn (or
ρ 1), however, the interaction of pairs does not reach far, and one finds
A∗ B∗ AB
Edsp, chn = − C6 3π Nchn 4πε0 2 8lbnd D5 , linear in the number of bonds and
AB
varying as the inverse-five distance. The value of the coefficient C6 can be
fit to theory or experiment, and has a magnitude around 0.561
kJmol−1 nm6 [38].
A similar procedure has been followed to describe the interaction of two iden-
tical cyclic compounds on parallel planes. They are taken to be circular with
aligned centers and a circumference of length Lcycl = NClbnd, at a distance D
between planes. The dispersion energy is found to be [38]
AB
A∗ B∗ C6 3π 2 ρ 2 2 1 1
Edsp, cycl = − 1+ +
4πε0 2 8 lbnd 2 D4 ρ 2 + 1
1
2 3 ρ 2 +1 ρ +1
2 2
where now ρ = L/(πD). The linear chain result is recovered at distances D small
compared with L.
More accurately, the bond can be described by anisotropic parallel and per-
a a
pendicular bond polarizabilities α ω and α⊥ ω obtained in a reference
frame attached to the bond, such that for a bond making an angle θ with a ref-
erence z-axis in a space fixed frame, the oriented bond polarizability is
284 7 Interactions Between Two Many-Atom Systems
a a a
αθ ω = cos2 θ α ω + sin2 θ α⊥ ω
a a a
with an average over orientations αAv ω = α ω + 2α⊥ ω 3. Also, a
molecular fragment can be chosen to include the C─H bonds so that the polar-
izability can be constructed along a given axis as α HC − CH =
a
α C− C + 2α C −H introducing expansions in basis sets um r located at
the midpoints of all three atom–atom bonds, and using oriented bond polariz-
abilities. Values of static polarizabilities for a large number of molecular frag-
ments have been collected [40, 41] and can be used to construct dynamical
polarizabilities with interpolation formulae as shown in Chapter 2.
−inx πD − iqn r
j ce exp −iqn r Ω j = exp j ce exp j
B Lx Ω
A
AB Ω inx πD A A −2
Eels, π D = n
exp Q0, π qn ; 0 Q0, π −qn ; 0 qn
4πε0 Lx
with wavenumbers qξ = nξπ/Lξ, nξ = 1, … Nξ, max . The total electrostatic inter-
action energy must also include interactions with nuclear charges, and the net
result is zero if the chains are neutral and far apart. Induction and dispersion
energies are obtained from susceptibilities. The dynamical susceptibility
π
χ πA qn ,qn ; iω of A can be calculated from integrals 0A QA qn ; 0 JA and
π
excitation energies Δεj j = ε π j − ε π j appearing in denominators, which
are smallest for the transition between highest occupied (HO) and lowest unoc-
cupied (LU) energy levels. Similar quantities are needed for B displaced by D
along x with the dependence on the distance D appearing in the integrals for
the Fourier components of the density.
In the box model, two chains can be located one inside a box A with bound-
aries 0 ≤ x ≤ L⊥, 0 ≤ y ≤ L⊥,and 0 ≤ z ≤ L = Lchn and the other inside a similar box
B parallel to the first with its origin displaced along the x-axis by a distance D so
that a new boundary is D ≤ x ≤ D + L⊥. In a particle-in-a-box model, the elec-
tronic states are null outside the boxes and at their edges. For A, they are
1 1 1
2 2 2 2 2 2
φπA r;k j = sin kx x sin ky y sin kz z
L⊥ L⊥ L
where kξ = jξπ/L⊥, ξ = x, y and kz = jzπ/L for integers jξ. This assumes that the
spacing of grid points is dense enough, with the total system in a box with Lξ
L⊥, L , allowing choices like Lx, y = 100L⊥ and Lz = 100L to properly describe the
interaction of the chains for varying distance D between them. The box energies are
2
ε π k j = ℏ2 2me k j , 1 ≤ jξ < ∞ , and the ground state energy has quantum
numbers j = (111). The orbitals φπB r ;k l for B are similarly obtained for the
identical displaced chain along x, with boundaries D ≤ x ≤ D + L⊥, 0 ≤ y ≤ L⊥,
1
and 0 ≤ z ≤ L and the factor (2 L⊥ 2 sin kx x− D giving the x-dependence.
7.4 Models of Hydrocarbon Chains and of Excited Dielectrics 287
The present model assumes that the pi electrons occupy orbitals and are excited
only along the z-axis, but remain in the lowest orbital along x- and y-axes, so
that here jx = jy = 1.
The one-electron integrals j ce exp − iqn r Ω j between box orbitals
r j = φπA r ;k j are products of integrals over the x, y, and z variables. Each
factor contains products of sin and cos functions, readily integrated and giving
negligible values for very large jz. For example, the z-integrals for the density
el
component QA qz are
inz πz 2 L
π inz πz π
jz exp − jz = dz sin jz z exp − sin jz z
Lz L 0 L Lz L
jz − jz jz + jz
iL 1+ exp inz π L Lz − 1 1− exp inz π L Lz −1
=− nz +
πLz 2
jz − jz − nz 2 L Lz
2
jz + jz
2
−nz 2 L Lz
2
which can be further simplified for Lz L , showing then that the matrix ele-
ments change linearly with L /Lz. Integrals over 0 ≤ x ≤ L⊥ and 0 ≤ y ≤ L⊥ have
corresponding forms, also changing linearly with L⊥/Lξ. All integrals decrease
as jξ−2 for large orbital quantum numbers, and as nξ−1 for the density operator
Fourier components. Matrix elements k Ce exp −iqn r Ω k for B
involve the same integrals over y and z as for A, while the integral over x has
new limits ≤x ≤ D + L⊥. Integrals have the same form as given, but with the addi-
tional factor exp(−inxπD/Lx) from the integration over x − D.
Ground state expectation values of electronic density components follow
π A A A
from 0A QA qn ; 0 0A = Q0, x qn ; 0 Q0, y qn ; 0 Q0, z qn ; 0 , where the two
first factors are given by the above integral with jx = jx = 1, jy = jy = 1 and
A Ce jz = Nπ 2 inz πz
Q 0, z q n ; 0 = 2 jz exp − jz
Ω jz = 1 Lz
for an even number of carbon atoms in the chain, and orbitals with
double spin occupancy. The corresponding expectation value for B
π B A
is 0B QB qn ; D 0B = Q0, π qn ; D = exp − inx πD Lx Q0, π qn ; 0 .
The dynamical susceptibility χ πA qn ,qn ; iω of A can be calculated from inte-
π π
grals 0A QA qn ; 0 JA and excitation energies Δεj j = ε π k j −ε π k j =
2
ℏ2 2me π L jz2 −j2z and similar quantities for box B displaced by D along
x, all of which are known for the particle-in-a-box model. These susceptibilities
give the right qualitative dependences with the length of chains, but their
numerical values must be corrected introducing more accurate excitation ener-
π A A
gies Δεj j + ΔεG with a HO–LU energy gap parameter ΔεG . Related induction
288 7 Interactions Between Two Many-Atom Systems
the nξ index for density components, the dispersion energy recovers in this limit
the van der Waals form corresponding to interacting dipole fluctuations, and a
D−6 dependence on the distance between the chains.
π
In the model using pi-MO-LCAOs φj r = μ cμj ξμ r , coefficients and
orbital energies can be obtained from a one-electron pi-Hamiltonian treatment
where μ = 1 to NC gives the location of a pi-AO, and orbital energies are given in
terms of Coulomb (or atom) and resonance (or bond) energy parameters α and
β, as [42]
1
2 2 jμπ
cμj = sin
NC + 1 NC + 1
jπ
εj = α + 2βcos
NC + 1
π
Matrix elements JA QA qn JA are given by sums over one-electron
atom–atom integrals μ Ce exp −iqn r Ω μ with r μ = ξμ r . These inte-
grals are large only for pi-AOs at the same or near-neighbor locations. This
2
means that we need the Fourier transform of only two densities, ξ1 r and
∗
ξ1 r ξ2 r , for a generic pair of pi AOs. Matrix elements can be obtained
from Fourier integrals μ ce exp − iqn r Ω μ = 1 Ce exp −iqn r Ω 1 =
Ce Ω Iatm qn for an atom in chain A and μ Ce exp −iqn r Ω μ + 1 = 1
Ce exp −iqn r Ω 2 = Ce Ω Ibnd qn for a bond in A, and corresponding
integrals Iatm,bnd for B which follow by displacing integrands by D along the
x-axis. These integrals have explicit forms in the Fourier variables qn for given
pi-AOs functions and their Fourier transforms.
This model can be made more accurate allowing for changes of parameters
with bond lengths. In particular, excitation energies are more accurate when
allowance is made for changes of the β parameter between short and long bond
7.4 Models of Hydrocarbon Chains and of Excited Dielectrics 289
ΦmJ = Jm n m
0n
290 7 Interactions Between Two Many-Atom Systems
for J 0 as
H aggr Φ J = Eaggr J Φ J
These states and energies characterize the properties of excitons, which are
defined as electronic excitations delocalized over all the molecules. They have
been extensively studied for molecular solids and for aggregates including
organic macromolecules [35, 43–46].
The total Hamiltonian contains one- and two-molecule terms and its matrix
elements between many-molecule states have similarities with matrix elements
between many-electron states [9]. Here we provide a simplified version assum-
ing no electron exchange between molecules and using a basis set of molecular
states satisfying the approximate orthonormalization Jm Jn = δmn δJJ which is
accurate when electronic overlaps between molecules are small. Matrix ele-
ments of H m and H mn can be obtained in the basis set { Φ0 , ΦmJ } and to sim-
plify, we consider here only one excited molecular state J = 1.
Excitonic energies can be obtained from first-order perturbation theory, as
ΔE1 = Eaggr 1 − Eaggr 0 Φm1 H aggr Φm1 − Φ0 H aggr Φ0
m
Matrix elements EJJ = Jm H m Jm of H m with J = 0, 1 are nonzero only
m m m m mn
for EJJ = EJ , E0, 1 , or E1, 0 . Matrix elements of H mn are VJJ , KK =
Jm Kn H mn Jm Kn and nonzero only for J = J = 1, K = K = 0, and K = K = 1,
J = J = 0, corresponding to molecular excitation transfers 1 0 and 0 1.
Here the Coulomb interaction involves matrix elements Jm c r; Qm Jm of
the density operators, and the density–density interaction contains not only the
dipole–dipole interaction but also all higher multipoles. A formalism for inter-
acting excitons beyond the first-order perturbation treatment can be based on a
model Hamiltonian with linear, bilinear, and quartic terms in the excitation
(and de-excitation) bosonic operators B†m = 1m 0m (and Bm = 0m 1m ) [47].
More accurately, one must solve for the eigenenergies Eaggr(J), J = 0, 1. This
can be conveniently done when the aggregate is a crystal structure or periodic
chain, taking advantage of the periodicity of the molecular lattice. For a cubic
lattice with molecules at locations Rm = ξ mξaξ ,ξ = x,y, z, with −Mξ ≤ mξ ≤
Mξ an integer, and unit cells of sides aξ , a choice of periodic boundary conditions
7.5 Density Functional Treatments for All Ranges 291
centers of mass, but also damping when atoms of A and B come too close, with a
new fd(R, Q) as a function also of interatomic distances.
The interaction of two many-atom molecules has been similarly treated add-
ing atom-pair dispersion energies with pair damping functions. This assumes
that atom-pair dispersion energies are approximately additive, which can be
shown to be correct for many-atom interactions to second order of perturbation
theory, as previously discussed. The schemes developed so far have made dif-
ferent choices for the construction of Edsp from atomic structures, as a sum over
ab
(ab) atom-pair interactions Edsp Rab ) with Rab = R +Ra −Rb multiplying
ab
instead a pair damping function fd Rab , so that
ab ab
Etot R, Q = Esl R,Q + E
a, b dsp
Rab fd Rab
1 1 2
H MBD = − μ = a, b
∇μ 2 − μ = a, b
ωμ 2uμ
2 2
1 Nb
1 2
lr
+ ω ω αaNa 0 αb
b a b
0 u t
ξ, η a, ξ ξη
ub, η
2 a
1
with atomic mass weighted displacements such as ua, ξ = Ma 2 ξa − Ra, ξ oscillat-
ing with frequencies ωa, ξ . Its eigenvalue equation H MBD Fp u = ℏΩp Fp u
MBD
is solved to obtain many-body dispersion (MBD) energies Edsp R,Q from a
df
sums over NAB = 3 NA + NB −3 eigenfrequency shifts Ωp(R, Q) − Ωp(∞, Q),
given that there are NA atoms in A and NB in B. This provides
MBD
Etot R, Q = Esl R,Q + Edsp R, Q
in a DFT-MBD treatment and has been applied usually combined with the PBE
density functional, to generate results for several benchmark sets of small and
intermediate size compounds (S66 and S22), crystals (X23 set), and large molec-
ular complexes (S12L). They have been compared with accurate results from the
CCSD(T) treatment with generally very good agreement [33].
after using that, insofar density fluctuations integrate to zero, one finds
el el
d 1 χS 1,1 ; iω = d 1 χ S 1,1 ; iω = 0
S S
The relation with DFT is done assuming that the polarizabilities are local. The
original literature by several authors has been updated in recent reviews that
also include more general formalisms [51–54].
The present equations are simplified first choosing the local form
el el
α S r, r ; iω = δ r − r γ S r; iω , and using that for electron distributions
(1) and (2), respectively, at molecules A and B far from each other, their leading
−3
dipole–dipole interaction energy goes as r 1 −r 2 , which gives
∞
dsp 3ℏ el 1 el
EAe, Be =− dω d 1 d 2 γ A r 1 ; iω γ r 2 ; iω
π r 1 −r 2
6 B
0 A B
el el
ΓS iω = d 1 γS 1; iω
S
296 7 Interactions Between Two Many-Atom Systems
nl ℏ 3 3 c4e
Elrc ρe r ,ρe r = d r d 3 r ρe r ρe r
2 2 m2e g r g r g r + g r
2
g r; r − r = ω0,VV r r − r +κ r
obtained from a model of a spherical inhomogeneous electron distribution with
an energy band gap given by
1
ω2p ℏ2 kF4 4
2
ω0, VV r = +c s
3 m2e
where ω2p = 4πρe c2e me = 4c2e 3πme kF3 is a local plasmon frequency squared,
and where κ = b ℏkF 2 m2e ωp with an adjustable parameter b, which controls
the damping of g at short distances r − r . The term containing s prevents a
el
divergence in γ S when both ω 0 and ρe 0.This choice, named VV09/
6
10, has given values for 58 CAA vdW coefficients to an accuracy within 15%,
obtained for a test set of pairs including identical closed- and open-shell atoms,
diatomics, and polyatomic molecules, with the value of the numerical parameter
c = 0.149 multiplying the s4 term adjusted to minimize standard deviations from
6
accurate values of CAA [56]. The above volume integrals over d3r and d3r
extend over the whole space containing both A and B species and have been
constructed so that the dispersion energy vanishes at short distances for a
homogeneous electron gas. An example of results obtained this way is shown
for the methane dimer in Figure 7.2 from [56].
The good accuracy of the described nonlocal density functional
nl 6
Elrc ρe r ,ρe r for CAB coefficients, and its physical behavior for all dis-
tances, allows a treatment combining the more standard semilocal exchange-
sl
correlation functionals Exc ρe r ,… , already tested for interaction energies
nl
at short and intermediate distances, with the nonlocal form Elrc to construct
intermolecular energies valid over all distances R as
nl
Exc R = Excsl ρe r ,…; R + Elrc ρe r , ρe r ; R
0.1
(CH4)2
0.0
–0.1
–0.2
ΔE (kcal mol−1)
–0.3
–0.4
–0.5 Reference
vdW-DF2
–0.6 VV10
–0.7
3.2 3.6 4.0 4.4 4.8 5.2 5.6
R(Å)
Figure 7.2 Methane dimer interaction energy versus distance R. Source: from Ref. [56].
Reproduced with permission of American Institute of Physics. Here R is the distance between
the centers of mass of the monomers. VV10 and vdW-DF2 results were obtained self-
consistently with the aug-cc-pVTZ basis set.
Section 7.2, but can be more generally done using efficient methods from DFT
for both cluster and environment [59–61]. Embedding DFT methods start with
a partition of the physical system into molecular fragments K containing
at
selected atomic cores κ = 1 to NK and assigned NK electrons, with this number
allowed to change as interfragment electronic rearrangement occurs. This num-
ber of electrons can be fractional and interpreted as resulting from a statistical
average over an ensemble of fragments with varying integer number of electrons
in each one. In a simple version, the total system is partitioned into two: a cluster
of selected atoms with K = A and an environment with K = B.
The total energy Etot can be obtained to begin with by means of a density func-
tional embedding theory (or DFET), which describes the whole system, to be
corrected by accurate calculation of the embedded cluster (or emb cls) fragment,
for example, with a correlated many-electron wavefunction (or CW), by adding
CW DFT DFT
a correction term ΔEemb cls = Eemb cls −Eemb cls to Etot . This is similar to the men-
tioned subtraction scheme in the QM/MM approach, and assumes that the
environment is not changed by atomic rearrangements in the cluster.
7.5 Density Functional Treatments for All Ranges 299
K
for a collection of interacting fragments. The functional Etot ψj r,ζ of a
K
set of j = 1 to NK spin-orbitals ψ j for each fragment K can be constructed to
describe electrons moving in an effective potential Vemb r , with the orbitals
given as functionals of Vemb r . This can be derived from the relation
δE δVemb r = 0, and optimized by energy minimization, to obtain an optimized
OEEP
embedding effective potential (OEEP) Vemb r and the corresponding PES
satisfying Etot(QA, QB) ≈ EA(QA) + EB(QA) with integer number of electrons in
each dissociation species [60].
Total
energy ET
(each a cycle of data inputs). The results have shown accuracy errors smaller
than 100 meVs in comparison with the original PES, and have shown smooth
behavior, so that molecular dynamics simulations of dissociation could be accu-
rately done [68]. This can also be done for many other PESs of interactions at
solid surfaces, relevant to a wide range of applications to the discovery of new
materials.
The majority of applications of AI-NN learning have been done to generate
the energies of a many-atom system in its ground electronic state. However, the
ML procedure can also be used for nonadiabatic processes to generate energies,
PES forces and momentum couplings for a set of coupled ground and excited
states, as needed for nonadiabatic dynamics calculations, provided enough
information is available to construct data sets to be used for training and ver-
ification. This has recently been shown to be doable for conical intersections
appearing in the photoisomerization of CH3NH [69] and for coupled ground
and excited states of the reactive triatomic LiFH [70].
Another related application of AI-NN learning of relevance to the develop-
ment of new medical drugs involves the calculation of receptor–ligand scoring
indices for small molecular ligands interacting with large biomolecular recep-
tors. Neural network scoring functions have been found to reproduce results
from state-of-the-art docking programs based on atomistic models, with NN
computational speeds up to two orders of magnitude larger than speeds for
the available scoring procedures [71].
Development of new methods and applications for AI-NN machine learning
is very active at present. These AI methods are complementary to the ones
based on atomistic models and detailed calculations of electronic structure,
because AI methods can be used only when large numbers of data points can
be collected and organized in data sets to be used for NN training and verifica-
tion. Organization of the data sets require the introduction of molecular struc-
ture descriptors based on physical and chemical considerations, a subject that
needs much attention. A good choice of descriptors can lead to the desired
results in a short time, while bad descriptor choices may yet lead to needed
values but after a much longer computational time. When the data sets are avail-
able, the AI-NN procedures are computationally very efficient. Many improve-
ments in accuracy and efficiency of AI-NN machine learning methods can be
expected in the near future.
References
1 Longuet-Higgins, H.C. (1956). The electronic states of composite systems. Proc.
R. Soc. A 235: 537.
2 Dzyaloshinskii, I.E., Lifshitz, E.M., and Pitaevskii, L.P. (1961). The general theory
of van der Waals forces. Adv. Phys. 10: 165.
304 7 Interactions Between Two Many-Atom Systems
21 Xie, W., Song, L., Truhlar, D.G., and Gao, J. (2008). Variational explicit
polarization potential and analytical first derivative of energy: towards a new
generation force field. J. Chem. Phys. 128: 234108.
22 Jensen, F. (1999). Introduction to Computational Chemistry. New York: Wiley.
23 Ufimtsev, I.S. and Martinez, T.J. (2008). Quantum chemistry on graphical
processing units. 1. Strategies for two-electron integral evaluation. J. Chem.
Theory Comp. 4: 222.
24 Warshel, A. and Levitt, M. (1976). QM/MM. J. Mol. Biol. 103: 227.
25 Bash, P.A., Field, M.J., and Karplus, M. (1987). Free energy perturbation method
for chemical reactions in the condensed phase: a dynamic approach based on a
combined quantum and molecular mechanics potential. J. Am. Chem. Soc.
109: 8092.
26 Field, M.J., Bash, P.A., and Karplus, M. (1990). A combined quantum mechanical
and molecular mechanical potential for molecular dynamics simulations.
J. Comput. Chem. 11: 700.
27 Maseras, F. and Morokuma, K.I. (1995). A new integrated Ab Initio+ molecular
mechanics geometry optimization scheme of equilibrium structures and
transition states. J. Comput. Chem. 16: 1170–1179.
28 Friesner, R.A. and Guallar, V. (2005). Ab initio quantum chemical and mixed
quantum mechanics/molecular mechanics (QM/MM) methods for enzymatic
catalysis. Annu. Rev. Phys. Chem. 56: 389.
29 Lin, H. and Truhlar, D.G. (2007). QM/MM: what have we learned, where are we,
and where do we go from here? Theor. Chem. Accounts 117: 185.
30 Senn, H.M. and Thiel, W. (2009). QM/MM methods for biomolecular systems.
Angew. Chem. Int. Ed. 48: 1198.
31 Brunk, E. and Rothlisberger, U. (2015). Mixed quantum mechanical/molecular
mechanical molecular dynamics simulations of biological systems in ground and
electronically excited states. Chem. Rev. 115: 6217.
32 Hirschfeld, F.L. (1977). Bonded-atom fragments for describing molecular charge
densities. Theor. Chim. Acta 44: 129.
33 Hermann, J., DiStasio, R.A.J., and Tkatchenko, A. (2017). First-principles models
for van der Waals interactions in molecules and materials: concepts, theory, and
applications. Chem. Rev. 117: 4714.
34 Cohen-Tanoudji, C., Diu, B., and Laloe, F. (1977). Quantum Mechanics, vol. 1,
Chap. IV. New York: Wiley.
35 May, V. and Kuhn, O. (2000). Charge and Energy Transfer Dynamics in
Molecular Systems. Berlin: Wiley-VCH.
36 Shavitt, I. (1963). The Gaussian function in calculations of statistical mechanics
and quantum mechanics. In: Methods of Computational Physics, vol. 2 (eds. B.
Alder, S. Fernbach and M. Rotenberg), 1. New York: Academic Press.
37 Martin, R.M. (2004). Electronic Structure: Basic Theory and Practical Methods.
Cambridge, England: Cambridge University Press.
306 7 Interactions Between Two Many-Atom Systems
CONTENTS
and can be treated as dielectrics, while metals (and the related semiconductors)
contain delocalized electrons and require a different treatment. Surfaces of
these solids have corresponding electronic properties, which affect the way
the molecule interacts with the surface. Furthermore, the solid can be regular
with a periodic lattice structure or amorphous. Rearrangement of atomic posi-
tions and of electronic distributions occur at a surface when this is formed in a
solid, which further affects its interaction energy with a molecule.
The potential energy of interaction of a molecule near a solid surface can be
constructed as a sum of terms for the molecule interacting with unit cells for a
periodic lattice or with atom groups for amorphous surfaces. This provides a
starting description of more complicated situations, where many molecules
may be present as the solid surface is in contact with a gas, liquid, or another
solid. At short distances between an adsorbed molecule and a surface, the inter-
action leads to physisorption if it is weak and nonbonding, while a strong inter-
action involving electronic rearrangement and bond breaking or formation
leads to chemisorption. The interaction between two adsorbed molecules is both
direct and indirect through surface effects, and their interaction strength
depends on whether they are physisorbed or chemisorbed.
A clean and regular surface of a crystal can be defined by a cut through its
three-dimensional lattice, giving a plane with atomic or molecular units in a
periodic structure. An introduction to the nomenclature and definitions for lat-
tices can be found in book chapters on the physics and chemistry of solids [1, 2]
and in more detailed treatments of clean surfaces [3] and surfaces with adsor-
bates [4]. In brief, a plane through a crystal lattice formed by periodic translation
of a unit cell with side vectors (primitive translations) a, b , c is defined by
Miller indices (h, k, l), which are the inverse of intersection distances along
the three axes, in units of primitive side lengths. Examples for a cubic lattice
are the (001) plane, cutting the c axis and parallel to the a, b plane, (011) cut-
ting b and c axes, and (111) cutting all three.
For a given planar lattice, the primitive translations a, b define five Bravais
lattice types (oblique, square, hexagonal, rectangular, and centered rectangular),
which combine with 10 crystalline point group symmetries to give 17 two-
dimensional space groups. Translation vectors between cell units on a plane
are of the form T s = ma + nb with m and n integers. Lines on the plane are
defined by two Miller indices (h, k) and the distance between two lines, when
a b = 0, is d(h, k) = [(h/a)2 + (l/b)2]−1/2.
Electronic densities in the planar lattice must satisfy periodicity conditions,
and they determine the periodic potential energies of interaction with a nearby
molecule. Periodicity can be imposed introducing the primitive reciprocal lat-
tice vectors A, B , which define a Brillouin lattice in reciprocal space, and
8.1 Interaction of a Molecule with a Solid Surface 311
V R = V Z exp iG Rs
G G
1
VG Z = d 2 Rs V R exp − iG Rs
As As
where As is the area of a unit surface cell and the VG are Fourier components of
the potential energy at each distance Z. Periodicity follows from the fact that
G T s = 2π hm + kn with the factor in the parenthesis being an integer.
The potential energy V R is composed of long-range, intermediate-, and
short-range interactions of the molecule with the surface and depends also
on their internal atomic positions (QA, QS) for the molecule and surface.
Long-range interactions can be obtained by analogy with the procedure for mol-
ecule–molecule systems, with the difference that molecule–surface interactions
may extend over a large region at the solid surface and also inside it, so that the
overall interaction results from sums of local interactions over the surface and
volume of the solid.
This strongly modifies the dependence of the potential energy on the distance
Z between a molecular center-of-mass and a surface plane. The van der Waals
interaction energy, which goes as R−6 for two molecules, becomes a function
going as Z−3 after addition over the solid’s atomic components, with a coeffi-
cient expressed in terms of the molecular and solid dynamical susceptibilities.
The induction energy between a charged molecule and induced anisotropic
dipoles extending over the solid surface similarly changes from R−4 to Z−1, with
a coefficient given in terms of the solid polarizability per unit volume. And the
interaction energy between a surface dipole layer and a charge at the molecule
becomes a constant independent of Z. This will be shown in more detail in what
follows.
Intermediate- and short-range interaction energies can be obtained from
functionals of the electronic densities of molecule and solid surface, similarly
to what is done for molecule–molecule interactions. The treatment, however,
312 8 Interaction of Molecules with Surfaces
(c)
A+BS
V ABS
Zm Z
AB+S
Ze
d(A–B) A+B+S
ABS≠
position R changes while the internal positions (QA, QS) remain close to their
values at large relative distances.
However, a more general approach is needed when there is extensive atomic
rearrangement at shorter distances, for example due to molecular dissociation
of A, which affect not only A but also the substrate S. This situation can be
described introducing new fragments A and S and their self-consistent cou-
pling, with new densities ρA and ρS , and are functions of new atomic internal
positions QA ,QS , and the relative coordinates. The fragment A includes A
and neighboring surface atoms, while S is the remaining indented surface struc-
ture. Their densities can be constructed so that the total density is ρAS = ρA + ρS ,
and the total energy of the AS system can be obtained from an additive scheme
(or electronic embedding) writing instead that the total energy for A at relative
location R over the surface is
AS
E AS
=E A
R; ρA r +E S
R; ρS r + Eint R; ρA r + ρS r
AS
which contains the interaction energy Eint between the two coupled subsys-
tems. This is given by the quantal expectation value of the energy coupling oper-
ator between charges of A and S obtained from a many-electron state Φ(AS),
which allows for atomic rearrangement, calculated from the full Hamiltonian
Coul Coul
H AS = H A + H S + H AS = H A + H S + H A S . Alternatively E(AS) can be con-
structed as an approximate functional within DFT. The interaction energy
can be given as
AS
VA, S R = ΔE A
R; ρA r + ΔE S
R; ρS r + Eint R; ρA r + ρS r
where ΔE A
=E A
−E A
and ΔE S
=E S
−E S
are fragment adsorption and
indentation energies changing with R, in an embedding treatment.
The following sections provide details on electronic states, densities, and sus-
ceptibilities for a solid surface, as needed to consider its interaction with
molecules.
φk , μ r = T
exp ik T χ μ r − T
which has the limits ρ ≈ |A0|2 = ρ+ for z − ∞ and ρ ≈ 0 for z D−, and is
zero for z > D. This shows that the density rises from z = D inwards and oscil-
lates along decreasing z away from the surface plane, with a length period
zF = kF/π which increases with the density of electrons, and with an amplitude,
which decreases as the density increases. The parameter D can now be fixed
requiring that deviations of the electronic density ρ(z; D) from the value − ρ + ,
which neutralizes the positive background, should add up to zero,
D
or − ∞ dz ρ z; D + ρ + = 0.
More details about the surface electronic structure can be obtained from a
model of a solid with a surface perpendicular to the z-axis at location z = 0, with
an electronic potential energy V(z) of constant value Vin < 0 inside a semi-
infinite region with − ∞ ≤ z ≤ 0, and value Vout = 0 for z ≥ 0 in a vacuum, so that
now electrons stay mostly within − ∞ ≤ z ≤ 0 but spill out into the vacuum. As
before, the electron density ρ r = ρ z and electron states φk r can be obtained
from the solution for a particle with momentum ℏk, but now they decay expo-
nentially into the vacuum when their energy is εk = ℏ2 k 2 2me + Vin < Vout so that
φk r = Ak z exp ik r sin k⊥ z + γ ,z ≤ 0,
This summation extends over all cell units contained within the two surface
planes (h, k, l) defining the slabs and is different for each set of plane indices.
Computational treatments and results of DFT for surfaces can be found, for
example, in Ref. [7].
The potential energy of an electron in the solid can be calculated for this elec-
trostatic potential and provides work functions W(h, k, l), which depend on the
surface indices. Taking the surface to be perpendicular to the z-axis, with an
averaged (over x- and y-variables) electronic potential energy v z , W can be
obtained as the difference of electronic energies
W = v ∞ + EN − 1 −EN = v ∞ −μ
where μ is the chemical potential of the electrons. For example, values for fcc Cu
surfaces are 4.59, 4.48, and 4.98 eV (or 442.9, 432.3 and 480.5 kJ mol−1) for sur-
face planes (100), (110), and (111), respectively. These values are obtained from
photoemission measurements, and DFT calculations show the same trends.
Accurate calculations must account for distortion of the solid lattice near the
surfaces. The work function for Si(111) is calculated to be about 4.15 eV, but
that surface shows reconstruction into Si(111) (2 × 1) and (7 × 7) at varying tem-
peratures, with a related small change in the work function for the (2 × 1) struc-
ture, but giving a work function lower by about 0.2 eV (19.3 kJ mol−1) for (7 × 7).
8.1 Interaction of a Molecule with a Solid Surface 319
εr ω −1 1
= ρ αJ ω
εr ω + 2 3ϵ0 J J
This expression can also be used to account for changes across a surface intro-
ducing a molecular density ρJ Z , which varies when averaging its value over
a surface perpendicular to the Z-axis, and gives a varying dielectric function
εr(Z, ω). It can further be simplified by introducing parametrized forms with
frequencies and intensities obtained from independent calculations or measure-
ments. Letting αJ(ω) = αJ,el(ω) + αJ,ion(ω) to include electronic and displaced ion
core terms, these polarizabilities can be written as [2]
c2e me c2J MJ
αJ , el ω = , α J , ion ω =
ωJ , 0 2 − ω2 ωJ , T 2 −ω2
where ℏωJ, 0 is the lowest electronic excitation energy of species J, while cJ, MJ,
and ℏωJ, T are the charge, mass of ion J, and the transversal optical vibrational
energy of the lattice of J ions. Typical energy values for dielectrics are ℏωJ, 0
10 eV (or 103 kJ mol−1) and ℏωJ, T 10−1 to 10−2 eV (or 10 to 1.0 kJ mol−1)
so that for photons in the visible, near-IR and near-UV spectral regions, with ω2
ωJ, 02, the electronic polarizability can frequently be taken to be a constant inde-
pendent of frequencies.
work with its spatial Fourier transform from Z to the wavenumber q within a
slab of thickness L,
1
εr q,ω = dZexp −iqZ ε r Z, ω
L L
the collective oscillation of the electrons, forming the plasmon. The electronic
dielectric function follows from
ext + ind ωp 2
ϵr, el ω = = 1−
ext ω2
which gives propagating waves for positive values, as found from the Maxwell
equations, when ω > ωp.
In more detail, the susceptibilities of conductive materials, whether semicon-
ductors like Si or TiO2, or metals like Cu or Ag, can be obtained from their
atomic structure expanding charge density operators in a basis set of stationary
many-electron states ΨK x1 ,…,xNe and deriving the dynamical susceptibility
+
χ cc, a ω , with K = a, from its expression given above and involving the time-
correlation of charge density operators, as described in Chapter 2 for dynamical
polarizabilities. A useful but more approximate treatment derives the suscepti-
bility from self-consistent field or DFT one-electron treatments with a basis
+
set of Bloch orbitals φk , j r = φλ r giving matrices χ λλ , a ω , a susceptibility
+ +
χa ω = χ a ω and ε r, a ω = εr , a ω = 1 + χ a ω .
8.1 Interaction of a Molecule with a Solid Surface 323
As explained in the treatment of solids [2, 15], the electronic dielectric func-
tion ϵ el 0, ω = ϵ0 ϵr , el ω of an isotropic metal with electronic density ρe can be
written as a sum over electronic state-to-state transitions Λ = (λλ ), for fixed
nuclei, so that
ρe ce 2 fΛ
ϵr, el ω = 1 +
ϵ0 me Λ ωΛ 2 − ω2 − iωΓΛ
where ωΛ, fΛ, and ΓΛ are transition frequencies, strengths, and rates. In more
detail, the summation must be written as a double integral over densities of band
state λ = k,j and λ = k ,j per unit energy. The quotient ρece2/(ϵ0me) = ωp2 is
the square of the plasmon frequency for the free electron gas of the given den-
sity, with typical plasmon energies ℏωp of the order of 10 eV (about 103 kJ
mol−1) for metals. At zero frequency, the static dielectric constant is then
ϵr,el(0) = 1 + ωp2 ΛfΛ/ωΛ2 = 1 + ωp2/ω02 with ω0 a frequency parameter. At large
frequencies, provided transitions have small strengths fΛ for ω2 ωΛ2, it is
1
ωp 2 ρ e ce 2 2
ϵr, el ω = 1 − , for ω > ωp =
ω2 ϵ0 me
as already found with the simple displacement model. The total dielectric func-
tion including ionic contributions can be obtained within a model of oscillating
ion charges with a restoring (transverse optical) frequency ωT and ionic plas-
mon frequency ωp, ion from ωp, ion2 = ρioncion2/(ϵ0Mion) giving a dielectric func-
tion for each type of ion in the solid metal as
at low frequencies ω < ωT, and ϵion(ω) = ϵion(∞) at the high frequencies when the
ions do not respond to a rapidly oscillating field. The constant parameter ωT2
can be re-expressed in terms of a measurable ϵion(0). The total relative dielectric
function can be written as
2
ϵr ω = ϵr , el ω ϵion ω = ϵion ω − ωp ω ω2
with ωp 2 = ϵion ω ωp 2 [15], at frequencies ω > ωp. Typical values for the static
dielectric function ϵr(0) are about 12 for Si, 11 for Al, and 6 for Cu.
The presence of a surface in a conductor alters its collective electronic oscilla-
tions, and leads to the appearance of slower collective oscillations tangential to
the surface with surface plasmon frequencies ωs = ωp/ √ 2. This follows from
consideration of the total fields inside and outside the surface created by an
oscillating charge localized at the surface, and by imposition of continuity of
the perpendicular electric displacement fields inside and out, at the surface
324 8 Interaction of Molecules with Surfaces
(a) z (b) z
–e
→
rH C
+e
d+zH r d
d
0 rD 0
x x
– – – – – – – – – –
L + + + + + + + + + +
d
–eimg d+zH Dv
–d0
+eimg
Figure 8.2 (a) Atomic nucleus and electron point charges ±e above a dielectric and their
images ±eimg inside it; (b) A charge C above a surface dipole layer of a dielectric, with a dipole
density DV per unit volume.
Figure 8.2a, leading to rearrangement of the electronic density near the surface
and to appearance of an image charge Cimg and an electric potential ϕ R , which
can be derived from C and its image charge Cimg at distance −d inside the solid.
Its magnitude is determined by continuity of the gradient of the electric poten-
tial at the surface.
A single positive point charge C at a distance d from a slab surface perpen-
dicular to the z-axis attracts electrons in the slab toward the surface, and this
creates the same effect as a charge image Cimg of opposite sign inside the slab
at a distance d from its surface [16]. The electrical potential ϕ R at a point
R =Rs + Znz in the vacuum due to the charge and its image is
1 C Cimg
ϕ vac R = +
4πϵ0 R + dnz R −dnz
The force between the charge at distance Z from the surface and its image
is Fimg = CCimg/(4πε02Z)2, which gives by integration over d ≤ Z ≤ ∞ a
charge-image potential energy
vac
Eimg = CCimg 4πϵ0 4d
Inside the solid, the electric potential ϕ(sol) has the same form as ϕ(vac) but with
ε0 replaced by εrε0.
326 8 Interaction of Molecules with Surfaces
e2 εr −1 1 1 1
H H , img r H = − − +2
4πε0 εr + 1 4d 4 d + zH 2 2d +r H
els e2 εr − 1 2
VH , img d = H H , img =− g rH g
g 4πε0 12d3 εr + 1 H
8.2 Interactions with a Dielectric Surface 327
that shows its distance dependence and trend with changing dielectric permittiv-
ity. The interaction is attractive insofar εr > 1, it decreases as εr 1 and reaches
an attraction maximum when εr ∞, in pure metals.
Along the same lines, another application of the images model can be made
for a polar molecule A with a permanent dipole DA at an angle ϑ with the surface
normal nz . From a model of two opposite charges making the dipole, and their
images one finds
els 1 εr −1 2
Vdip, img d,ϑ = − DA 1 + cos2 ϑ
4πε0 12d 3 εr + 1
for the electrostatic interaction energy. Here again the energy varies as d−3 with
distance, and reaches a maximum for a perfectly metallic solid.
A more general treatment for electrostatic, induction, and dispersion interac-
tions of a molecule near a surface can be done constructing an interaction Ham-
iltonian from the collection of all charges and their images, and calculating first-
and second-order perturbation energies of interaction. Results can be obtained
from charge distributions and the dynamical polarizabilities of the molecule
and solid. The dispersion interaction energy appears in a second-order perturba-
tion treatment and involves, similarly to the expression for two interacting mole-
cules, the polarizability αA(ω) of the species A, which is a 3 × 3 tensor, and a
dielectric function of frequency, or relative permittivity εr(ω) = εS(ω) of the
solid (or a slab). Insofar a transient dipole DA(ω) creates an image dipole
DA,img(ω) = DA(ω)(εS − 1)/(εS + 1), one expects (and finds) that such quotient
appears in the molecule–surface dispersion energy. More accurately, the treat-
ment also must account for retardation effects in the very long-range interaction
of adsorbate charges with surface charges very far away. This can be done within a
quantal dielectric response treatment to second order in the strength of the inter-
action Hamiltonian, alternatively using quantum field theory [18], the theory of
fluctuating electromagnetic fields [16], response theory [19], and a quantal per-
turbation treatment [20], all of which lead to the same results, with the response
treatment being simpler. The main relations are summarized as follows.
The atomic structure of the whole system can be described by the distance d
between the center-of-mass of the molecule and the surface plane and by the
internal atomic positions QA of the molecule and QS of the surface. The polar-
izability αA and dielectric function ϵS depend on these internal positions. The
result for the molecule–surface dispersion potential energy, omitting fixed
internal positions and with αA given by its the tensor zz-component, is [19]
∞
dsp 1 ℏ ϵS iω − 1
VA, S d =− dω αA iω
4πε0 2π 0 ϵS iω + 1
4 4ω 2ω2 dω
× 3 + 3 + 3 exp − 2
2d c 2d c2 2d c
328 8 Interaction of Molecules with Surfaces
where c is the speed of light. This expression contains retardation effects that
arise for very long-range interactions. For relatively small distances d c/ω,
compared to radiation wavelengths, the dispersion energy becomes
∞
dsp 1 ℏ ϵS iω − 1
VA, S d =− dω αA iω
4πε0 4πd3 0 ϵS iω + 1
ϵS iω −1 3F iω
=
ϵS iω + 1 2 + F iω
to be obtained from the polarizabilities of species in the dielectric solid.
The quotient containing the solid dielectric function in the above integral
over frequencies is the one that arises in the response of a dielectric to pertur-
bation by an oscillating dipole far from the surface. It can be reinterpreted in
terms of an effective susceptibility of the solid. Introducing the susceptibility
dsp
χ S(ω) = ϵS(ω) − 1, one finds that the quotient in the integral for VA, S is equal
to χ S ω = χ S ω 2 + χ S ω a renormalized solid susceptibility, which accounts
for medium effects and gives a χ S iω smaller than the original χ S(iω). In this
notation,
∞
dsp CA, S 1 ℏ
VA, S d =− 3
,CA, S = dω αA iω χ S iω
d 4πε0 4π 0
dsp
species J in the solid dielectric, to parametrize VA, S using static polarizabilities
J
α0 0 of atoms and molecules and their frequency constants ω J .
A more detailed analysis of the coupling of electronic charge density fluctua-
tions at the surface with fluctuations in a molecule at position R, with coordi-
nates (X, Y, Z) relative to the solid surface, leads to a correction needed at short
distances. It replaces Z−3 with the form (Z − Z0)−3 where Z0 is the position of a
reference plane and depends on the surface electronic density [20]. Without
going into details, the procedure is to express the second-order interaction
energy E(2) as a sum over excited states, with state-to-state transition integrals
for the Coulomb electronic interaction Hamiltonian containing 1/r functions
written in cylindrical coordinates ρ,z there, doing a two-dimensional Fourier
transform from ρ to a wavevector q , and re-expressing the sum over excited
states as an integral over frequencies containing the dynamical susceptibilities
of molecule and solid.
∞
This leads to an expression E 2
= −2 π 0 dω αA iω FS iω;q ,Z where Z
is the distance from A to the surface location defined by the first layer of
atomic positions. Derivation of FS ω;q ,Z and its expansion in powers of
dsp
q for large Z leads to the expansion of E(2) as VA, S Z = − A Z 3
1 + 3Z0 Z+ ≈ − A Z −Z0 3 with an explicit form for Z0 in terms of sus-
ceptibilities [20]. This correction to the Z−3 dependence is relevant at short dis-
tances, but a more detailed treatment is needed to account for the transition
from long distances d to short ones. For a molecule interacting with a surface
lattice, the coefficient A depends also on the coordinates (X, Y) of molecule
A above the surface, with the form
dsp 3
VA, S R = − A X, Y Z −Z0
nd
VA, S R = β
V Ab
R −Rβ
The total potential must satisfy the condition VA, S R +T s = VA, S R , and its
Fourier components VG Z can be extracted from the sum over atomic posi-
tions β given above.
For example, the potential energy for He in empty space at location (X, Y, Z)
interacting with the surface LiF(001) of the ionic solid, with the LiF molecular
unit in a f.c.c. cell of side length a = 0.402 nm, can be constructed from
8.2 Interactions with a Dielectric Surface 331
with the function Q satisfying Q(X + a, Y) = Q(X, Y + a) = Q(X, Y). It has been
used in calculations of cross sections for low-energy elastic scattering of the
atom by the ionic surface, which requires introduction of both attractive and
repulsive potential energies along the z-direction. A useful parametrization is
given by [25]
V0 Z = Dexp α Zm −Z exp α Zm −Z − 2 0
V1 Z = −2βDexp 2α Zm − Z
2πX 2πY
Q X,Y = cos + cos
a a
with parameters a = 0.284 nm, α = 11.0 nm−1, D = 7.63 meV (0.7362 kJ/mol),
β = 0.04 to 0.10, and Zm= 0.10 nm, with the range of β values available to fit
experimental results [26]. This gives a potential well along Z, and the same form
can be used for other rare gas atoms with suitable parameters. Addition of the
nd
dispersion energy to this VHe, S R would only give a small correction, com-
pared with the attractive induction energy due to polarization of the atom by
the ions. This potential energy function has been used in calculations of reso-
nance energies and shapes, in low-energy collisions of noble gas atoms with the
ionic dielectric surface.
Better agreement with scattering experimental results were obtained for He +
LiF(001) with a semi-(ab initio) potential energy obtained from a surface-
projectile sum V He, S R of He-(ion pair p) potential energies v He, p R−R β
with each term including a long-range van der Waals attraction as well as a
ind
short-range repulsion, plus an additional term V R describing the induc-
tion attraction of a polarized He atom to the lattice of ions in the solid [27]. The
latt 2
induction term is obtained as V ind
R = − αHe 2 R , where the elec-
latt
tric field = − ∇ ϕ latt is the gradient of the electric potential ϕ latt R of
the lattice ions interacting with He, with the potential itself a sum over all lattice
charges Cb at Rβ . This lattice potential is calculated for LiF(001) as a sum over
Yukawa-type potentials for each ion, decomposed as a Fourier series, and taken
332 8 Interaction of Molecules with Surfaces
He, S
to the limit of the Coulomb potentials. The results is VHe, S R = V R +
ind
V R , or in detail
2
He, S αHe latt
VHe, S R = V
G G
Z exp iG Rs − fd Z G G
Z exp iG Rs
2
latt
with G as derived in [27] in terms of charges and their positions for the given
lattice. The induction interaction is cut-off at short distances with a damping
factor fd (Z) going from 1.0 at large distances to zero at Z = 0.
More accurately, the long-range induction and dispersion energies can be
obtained from the charge distributions and dynamical polarizabilities of the
atom and the solid surface, using methods of response theory [19].
VAS R = m, n, p
v A R − d 0 +T m, n, p
∞ ∞ ∞
1
VAS R dSx dSy dSz v A S
abc −∞ −∞ Z + d0
8.3 Continuum Models 333
For a cell species B such that the interaction with A has axial symmetry around
the z-axis, with a = b, integration variables (Sx, Sy) can be replaced by the length
Ss = (Sx2 + Sy2)1/2 and an axial angle. The interaction potential becomes
∞ ∞
2π 1 2
VAS Z dSs Ss dSz v A Ss 2 + Sz 2
a2 c 0 Z + d0
which is now only a function of the distance Z from A to the surface, and can be
calculated for a variety of A–B pair interactions where the distance of A to the
species B is given by (Ss2 + Sz2)1/2.
Three useful examples, with a = b = c, for isotropic potentials and with bars
over their parameters signifying values adjusted for medium effects, are [26]:
a) An inverse power potential energy v(A) = − W(R0/R)m, giving
∞ ∞
π W R0m
VA, S Z = − d Ss2 dSz m
a3 0 Z + d0 Ss 2 + Sz 2 2
3 m −3
2πW R0 R0
=−
m −3 m −2 a Z + d0
9 3
zm zm
VA, S Z = W −3
Z + d0 Z + d0
4πW α Z + d0
VA, S Z = 1+ exp − α Z + d0 −R0
a3 α 3 2
which is a shifted repulsion in Z with the same relative slope.
z-direction pointing into the solid, as shown in Figure 8.2b. A dipolar layer also
appears when a layer of polar molecules is adsorbed at the solid surface. For a
collection of charges CI at distances dI from the surface, contained in a polar mol-
ecule, the interaction energy can be obtained by first locating each single positive
charge C in the vacuum at location (0, 0, d) above the surface plane, interacting
with a layer of surface electric dipoles of thickness l divided into cells with dipole
density DV per unit volume. The dipoles are assumed to be perpendicular to the
surface with the dipolar plane centered at −d0 = − l/2, as shown in Figure 8.2b.
The interaction energy per unit volume for the charge at distance R = [SD2 +
dD2]1/2 from a cell dipole located on the surface at distance SD from the z-axis,
with dD = d + d0 the distance from C to the plane of the dipole layer is given,
including the cosine dD/R of the angle between dipole and charge position, by
DV dD
uC , S R = − C
4πε0 R2 R
which can be integrated over a surface area of radius LD with elements of volume
l 2πSDdSD. The integral over 0 ≤ SD ≤ LD gives a total electrostatic energy
for small dD/LD. It provides an attractive force for a positive charge and shows a
small decrease of the interaction energy as the distance d of the charge C to the
surface increases, for a fixed radius of the dipole layer. It also shows increasing
attraction at a given distance as the radius of the dipolar layer increases. This
result can be extended to a collection of charges CI at distances dI contained
in a polar molecule, adding over charges in the presence of a surface with a dipo-
lar layer. More generally, the energy can be obtained for any charge density dis-
tribution at the solid surface, solving the Poisson equation for the resulting
electrostatic potential [17].
The previous result can be applied as an example to the calculation of the elec-
trostatic interaction of a hydrogen atom with the planar surface dipole distribu-
tion. Locating the hydrogen proton with charge Cp = |ce| = + e at a distance
ZD = Z + d0 from the dipole layer, and the electron at position r e relative to
the proton, the interaction Hamiltonian of the two atomic charges and the sur-
face dipoles is
C p DV l
H H , S Z,ze = − 2π γ ZD − γ ZD + ze
4πε0
with γ(Z) = 1 − Z/(LD2 + Z2)1/2. This can be expanded up to second order in
powers of ze insofar ze/ZD 1 away from the surface, and the electrostatic
els
potential energy can be calculated from VH , S ZD = g H H , S g H to first order
8.3 Continuum Models 335
H ads = − εads c−
1≤j≤M j 1≤j<k ≤M
φ j,k cj ck
Here the first term contains the thermal kinetic energy in two dimensions,
and the last term accounts for the bonding energy of the adsorbate to the sub-
strate, with coverage density θads(ρads, β).
and total energies are obtained adding all one-electron energies for spin orbitals
with occupation number nνσ and subtracting an electron–electron residual cor-
AM
relation energy Eres after accounting for electronic correlation, to avoid dou-
ble counting of correlation energy, so that the total energy is [29, 30]
AM
Etot = n ε −EresAM
νσ νσ νσ
AM A M
from which one can obtain VAM R = Etot R −Etot R − Etot with subtrac-
A M
tion of similar expressions for Etot and Etot . This is also a function of all the
internal coordinates of atoms in A and M.
Simple models for molecules near metal surfaces help to extract physical
insight into the many aspects of surface interactions. A one-electron treatment
relies on a piecewise potential energy along the z-direction and expands the
expressions introduced in Section 8.1 for a pure metal M to include the poten-
tial energy at the location of an adsorbed or nearby molecule A, and the inter-
action of each electron with the atomic ions in A and M in the field of all the
other Ne − 1 electrons screening the ion charges. The electronic structure of
A + M can be obtained from a one-electron jellium model of the metal with
a surface perpendicular to the z-axis at location Z = 0, with an electronic poten-
tial energy vM(z) of constant value vin < 0 inside a semi-infinite region with
− ∞ ≤ z ≤ 0, and value vout = 0 for z ≥ 0 in a vacuum, plus an attractive molecular
potential vA r for an isolated A at location Z so that now orbitals of electrons
staying mostly within − ∞ ≤ z ≤ 0 spill out to overlap the orbitals of A. One must
int
add an interaction potential energy vAM r,Z to account for electron delocal-
ization between A and M.
As A approaches M to a distance Z, each screened core ionic charge Ca in A
creates an image charge Ca, img = − Ca (εr − 1)/(εr + 1) in M, and each electron in
A interacts with all the ion charges and their images, and also with its own image
charge. This creates a long-range electronic potential energy operator
a, img
vAM r,Z,QA for each charge a, which add up to give an electrostatic poten-
img
tial energy vAM r,Z, QA between the electron and all other charges. By analogy
8.4 Nonbonding Interactions at a Metal Surface 339
with the treatment of a hydrogen atom near a surface in Section 8.2.1, it is found
that the leading interaction term at large distances between a neutral molecule
A and the surface comes from the electron at location r A = r − Zuz interacting
with all charges, so that
img Cel2 εr − 1 2
vAM r,Z = − r A + zA 2
4πϵ0 16Z 3 εr + 1
AM
plus higher order terms in 1/Z, with Cel = eNe =
a Ca . Further addition of a
nr
near-range repulsive electronic potential energy vAM gives an one-electron
int img nr
interaction potential energy vAM r,Z = vAM r,Z + vAM r,Z , and a total
one-electron potential energy operator
eff int
vAM r,Z = vM z + vA r,Z + vAM r,Z
which also depends on internal positions QA.
A one-electron description of the interaction valid for all distances can be based
eff
on orbitals for electrons moving in the vAM r,Z potential energy, con-
structed from linear combinations of molecular and metal orbitals, φμ r and
φk r with energies εμ and εk , respectively, and normalizations φμ φμ = δμμ
and φk φ = δ k − k , as
k
φλ r; Z = c
μ μλ
Z φμ r + d 3 k ckλ Z φk r
where λ is a continuous orbital label, and the dependence on QA has been omit-
ted. These orbitals are obtained from solutions of the one-electron Schrodinger
equation for fixed atomic positions, with a normalization φλ φλ = δ(λ − λ )
suitable for delocalized orbitals. The one-electron energies go asymptotically
int
as Z−3, as follows from the state averages φλ vAM Z φλ of the interaction
potential. Total energies are given by the sum of all one-electron energies of
occupied orbitals corrected by subtracting the residual electronic correlation
AM
energy Eres Z so that
E AM
Z = dε ε nAM ε,Z gAM ε,Z − EresAM Z
unoccupied molecular orbitals and metal orbitals filled up to the Fermi level. At
large distances with no interaction between A and M, it is given for each electron
spin state by
where the first term is gM(ε) with kF the Fermi wavenumber, and the last two
terms give gA(ε). At a shorter distance Z, the interaction shifts and broadens
εHO by amounts ΔεHO(Z) and γHO(Z), and εLU by amounts ΔεLU(Z) and γHO(Z).
The density of states is then of the form
γHO γLU
gAM ε; Z = d 3 kδ ε −εk + 2 + 2
k ≤ kF π ε − εHO + γ2HO π ε − εLU + γ2LU
where nAM(ε; Z, QA) is the electronic energy level population in AM. It shows
that the interaction energy dependence on (Z, QA) comes from the density of
one-electron energies of A + M and also from the broadening and shift of molec-
ular levels.
This simple treatment provides some insight on the dependence of interac-
tion energies as the distance Z between A and M is varied. In particular, insofar
vAM r,Z ≈ − Q r Z −3 , the sum of one-electron energies gives an attractive
img
potential energy changing as 1/Z3, and inasmuch level shifts and broadenings
int
involve the square of the coupling matrix elements φμ vAM Z φk , it is
found that at large distances, Δελ and γλ contribute a repulsive energy changing
as 1/Z6.
8.4 Nonbonding Interactions at a Metal Surface 341
∞ dsp
1 ℏ ϵM iω − 1 C
= − A,3M
dsp
VA, M Z =− dω αA iω
4πε0 4πZ 3 0 ϵM iω + 1 Z
dsp
with CA, M a dispersion coefficient dependent on internal coordinates of mol-
ecule and solid metal. As done for the interaction of molecules with dielectric
solids, we can introduce the function αA(iω) = β(A)(ω), and also define a new
function ϵM(iω) = ηM(ω). Similarly for the solid metal, we have ω2ηM(ω) ≈ ω2
+ ωp2 for ω ∞, ηM(ω) = 1 + ωp, ion2/(ωT2 + ω2) for ω < ωT, and the static value
ηM(0) = ϵM(0) as a parameter. A suggested interpolation formula for a metal with
a single-atom composition is
ω2p, ion ωp 2
ηM ω = 1 + 1 − gd ω + 1 − gd ω
ωT 2 + ω2 ω2
342 8 Interaction of Molecules with Surfaces
with gd(ω) a damping function equal to one at large frequencies and decaying to
zero over a transition region ωT < ω < ωp. This can be used in the approximate
dispersion coefficient
∞
dsp 1 A ℏ 1 ηM ω − 1
CA, M = α0 0 dω
4πε0 4π 0 1+ ω ω A
2 ηM ω + 1
which can be calculated from the static polarizability and the known number
of valence electrons of A contained in ω A , and from the static dielectric con-
stant and plasmon frequency of the solid. This shows that the solid dielectric
function contributes to the integral mostly when ω < 2ω A , and should give rea-
sonable results for noble gases on metal surfaces provided transitions Λ = (λ λ)
there, as introduced in Section 8.1.3, have small oscillator strengths fΛ where
ω2 ωΛ2 [18].
The effect of variations of the surface electronic charge distribution
ρe r ; R ,QA ,QM , for fixed atomic positions, on the dispersion energy can be
approximately calculated introducing a plasmon frequency dependent on the
2
location r across the surface by means of ωp r = ρe r ce 2 ϵ0 me and integrat-
dsp
ing over r the functional VA, S R; ρe r to average it across the metal surface.
This can provide trends for the interaction energy dependence on the distances
over which the metal electron density ρe decays outside the metal surface. An
alternative is to introduce the dispersion dependence of ϵr on the wavevector q
of light, derived from a Boltzmann-equation treatment of the electronic fluid
and parametrized as ϵr , el q,ω = 1 −ωp 2 ω ω− iΓ − αq 2 [33], to be com-
bined with a parametrized form for the polarizability of the molecule to obtain
dsp
an energy density vA, S R, q , and to integrate this over a distribution of q
values.
More generally, the energy of physisorption of a molecule A at all distances Z
from the surface shows a distance dependence with a minimum at a well posi-
tion Zm, which results from combined long-range electrostatic, induction, and
dispersion attraction and near-range electrostatic, exchange, and correlation
dsp
forces giving repulsion. Long-range dispersion energies VA, M R,QA ,QM are
obtained from polarizabilities of A and M and go as Z−3. Near-range potential
nr
energies VA, M R,QA ,QM , containing also electrostatic and induction long-
range forces for ground electronic states, must be obtained from the combined
molecule–surface electronic charge distributions where they overlap. They are
repulsive at short range due to the Pauli exclusion effect between electrons at
the molecule and surface, imposed by the antisymmetry of joint electronic
8.4 Nonbonding Interactions at a Metal Surface 343
with fd a damping function in the Z variable and showing surface lattice peri-
odicity in the (X, Y) variables. The two terms also depend on the internal atomic
coordinates (QA, QM).
Results of near-range DFT calculations containing electrostatic, exchange,
correlation, and polarization terms can be fitted to parametrized forms, and
in the simplest case have been fitted in the literature to functions decreasing
exponentially with Z. However, DFT can provide some of the attraction forces
through polarization and a more general potential function fit should include a
potential well. Combined near-range and long-range potential energies can be
parametrized for A at position R outside the solid S, with functions such as a
near-range Morse potential with a minimum coming from the induction attrac-
tion, as
(a)
–0.09 Ar
–0.03
0.03
0.06
3 4 5 6 7 8
d (Bohr)
(b)
Ar Ar Ar
+ + +
Figure 8.3 (a) Potential energy versus distance for Ar interacting with the surface of a metal
with electronic density corresponding to a spread radius rs = 3 au (near the value for the
density of Ag). (b) Electronic density changes for three distances, with full and dashed lines
for positive and negative changes, respectively. Source: from Ref. [36]. Reproduced with
permission of American Physical Society.
The computational procedure starts with a guess for ρcls, ρenv, and Vemb r .
The fragment functionals EK ρK r ,Vemb r ,K = cls,env, are constructed for
fixed atom positions. The energy is minimized to find the densities ρK r for
Vemb
+
Vemb
the given embedding potential, and this is updated by calculating δEtot δVemb r
to search for a minimum for fixed ρK r [52]. An iteration of this sequence
DFT
leads to the final potential energy surface Etot for the total system. This must
CW DFT DFT
be corrected with the addition of ΔEemb cls = Eemb cls − Eemb cls , where Eemb cls differs
DFT
from Ecls insofar the first term accounts for the boundary effect of Vemb r
present in the Hamiltonian for the cluster. The equations for the combined
DFT and CW treatment including the embedding potential are
DFT
δ Ecls
+ Vemb r = μemb cls
δρcls r
Ncls CW
H cls + j=1
Vemb r j Φcls = Eemb cls Φcls
The embedding treatment with the subtraction scheme has been applied to
physisorption of atoms and molecules.
For example, embedding has been used to obtain the conformation and vibra-
tional spectra of CO adsorbed on a Cu surface. Binding sites and energies and
vibrational spectra of this adsorbate have been obtained for the Cu(001) [53]
and Cu(111) surfaces. In the case of CO adsorbed on Cu(111), cluster and slab
models left doubts as to whether the CO would have larger binding energy on
top of a Cu atom or in a hollow of the surface, with different conclusions
depending on the quality of the DFT functional. It was reconsidered with the
embedded cluster treatment [54], using DFT for the environment and a multi-
configuration wavefunction with single and double excitations for the cluster,
which lead to the conclusion that CO is perpendicular to the surface and sits
on top of a Cu(111) atom, as known from experiments.
8.5 Chemisorption
8.5.1 Models of Chemisorption
Chemisorption is the result of extensive electronic rearrangement when a mol-
ecule or atom interacts with a surface, leading to electron transfer or to bond
breaking and formation [4, 8, 22, 28, 55]. Treatments of chemisorption energies
350 8 Interaction of Molecules with Surfaces
depend on whether the solid surface refers to a dielectric surface S with localized
electrons, to a semiconductor surface, or to a metal (using then S = M) with
delocalized electrons.
Chemisorption at a dielectric or semiconductor surface can be accurately
described with a cluster model, which includes the adsorbate and nearby atoms
in the substrate, with boundary bonds atomically saturated. The whole many-
atom system can be treated as a supermolecule with many-electron wavefunc-
tions including electron correlation as described in Chapter 6, to allow for both
short-range and long-range interaction energies. A cluster model can also be
treated within DFT-D formulations [45, 46], allowing for dispersion energies
as done for two interacting many-atom systems in Chapter 7. This is less
demanding of computing times and can be applied to quite large clusters.
The DFT-D treatments can also be applied to models of crystalline solid sur-
faces such as slabs with extended periodic lattices. This avoids problems with
structural and bonding properties dependent on the size of the cluster models
that may converge only for very large cluster sizes.
Chemisorption on metal surfaces brings in new aspects due to the delocali-
zation and high susceptibility of electrons in the metal, and must be treated dif-
ferently. This requires models with extended lattice structures for the metal
surfaces, with their electronic properties usually described within DFT. It can
be done including dispersion energies with a version of DFT-D that, however,
must go beyond sums over atom-pair interactions to account for many-atom
effects. Some of these extensions have been covered in Chapter 7 and in reviews
[45, 46].
An alternative to the DFT-D approaches has been provided by the vdW-DF
treatment [43, 44, 49] derived from the adiabatic connection theorem described
in Chapters 6 and 7, which also includes dispersion energies in a seamless way
between short and long ranges. Its introduction of a nonlocal expression for the
electronic correlation energy has been implemented for both localized and delo-
calized electron distributions and therefore is suitable for calculations of chem-
isorption energies for atoms and molecules on dielectric, semiconductor, and
metal surfaces.
Reviews of published calculations have compared results for chemisorption
with a variety of DFT functional also adding dispersion interaction energies,
and the general conclusion is that dispersion contributions are important
and needed to obtain correct chemisorptions binding energies and structural
bond distances [44–46]. Some examples are mentioned in what follows.
Physical insight can be obtained from one-electron treatments that account
for strong coupling of orbitals in the molecule and metal and that incorporate
electronic correlation through density functionals [8, 31]. The coupling leads to
shifts of molecular levels and to their broadening, depending on the relations
between the ionization and affinity energies of the molecule, and the work func-
tion and energy band shapes of the metal. This section considers features of
8.5 Chemisorption 351
where R,QA , QS are the relative position of A with respect to S, the internal
atomic positions of A, and those of S. Electronic charge transfer is likely to affect
the atomic structure of the solid surface near the adsorbate.
As described in Section 8.1, these adsorbates can be treated calculating the
±
functional E A S as a sum over energies of the charged fragments with internal
atomic positions QA ± and QS and with densities ρA ± and ρS , plus their inter-
action energy in an embedding treatment, letting
A± S A±
E R,QA , QS ; ρA ± S r =E R,QA ± ; ρA ± r +E S
R,QS ; ρS r
±
A S
+ Eint R,QA ± ,QS ; ρA ± S r
200
Energy (meV)
Z
–200 DLPNO-CCSD(T) Ag2/Ti9O25H14
PBE-D3(BJ) Ag2/Ti9O25H14
PBE Ag2/Ti9O25H14
–400
PBE-D3(BJ) Ag2/TiO2(110) periodic
–600
2 4 6 8 10 12 14 16 18 20
Z (Å)
Figure 8.5 (a) To the left, cluster model Ag2/Ti9O25H14 of the perpendicular adsorbate and
nearby surface atoms, with boundary bonds saturated with hydrogen atoms. (b) To the right,
potential energy versus distance Z from a Ag atom to a nearby Ti surface atom, for the cluster
model calculated from CCSD, DFT and DFT-D, and also for a periodic slab with four layers, from
DFT-D. Source: from Ref. [56]. Reproduced with permission of Royal Society of Chemistry.
done for a cluster model Ag2/Ti9O25H14 [56] as shown in Figure 8.5a for the
adsorbate in a perpendicular orientation including nearby surface atoms (with
boundary bonds saturated with hydrogen atoms) with accurate many-electron
CCSD(T) wavefunctions, and also with a DFT-D treatment including long-
range dispersion energies for the same cluster. The addition of the dispersion
energy has been done as explained in Chapter 7, with a damping function.
The DFT-D treatment has also been implemented within a model of Ag2
adsorbed on a slab of the solid, with a periodic lattice. Results for potential ener-
gies are shown in Figure 8.5b for the on-top Ti location. They indicate excellent
agreement for the binding energies between CCSD(T) and PBE-D3 results, but
large errors if the dispersion energy correction is ignored.
Chemisorption also happens for closed-shell molecules such as H2, CO, N2,
H2O, CO2, and C6H6 on solid surfaces, when electronic rearrangement near the
surfaces weakens the molecular bonds of a species AB and results in dissociation
into fragments A and B. These can both bond to surface atoms to form adsor-
bates AS and BS, or one or both fragments can go free away from the surface,
depending on the bond strength or repulsion force between each fragment and
the surface atoms.
The dissociation of a molecule AB as it approaches a surface S and breaks into
AS + BS occurs on a potential energy function E ABS
R,QA , QB ,QS ; ρABS r ,
with the relative position of the center of mass of AB near the surface reaching
a value Rads where the dissociation of AB occurs. The fragments then move on a
8.5 Chemisorption 353
diss
PES VAB, S Rads ,QA ,QB ,QS = VAB, S QA ,QB ,QS , a dissociation potential
energy function obtained from
diss
VAB, S Q = E ABS
Rads ,Q; ρABS r −E AB
QA , QB ; ρAB r − E S
QS ; ρS r
with Q = (QA, QB, QS). The functional E(ABS) can be obtained for different atomic
arrangements along the dissociation path from embedding treatments, with
clusters A , B , and AB including, respectively, A, B, and AB and also neighbor
atoms at the surface, and for a surface S with an indentation large enough so
that all three clusters can be accommodated in the surface hole. At the short
distance Rads between the adsorbate AB and the surface, the energy isocontours
of E(ABS) show a minimum for the equilibrium structure of AB , and a dissoci-
ation path leading to A + B .
The energy function E ABS
R,QA , QB , QS describes more generally dissoci-
ation for varying approach distances and orientation of AB with respect to the
surface, with energy isocontours showing a valley as AB approaches the surface
along a variable distance ZAB − S and as AB breaks into A + B , which separate
along RAS − BS with an activation energy along a path into an energy valley for AS
ads
+ BS. The potential energy function VAB, S ZAB− S ,RAS −BS can be fit to a Lon-
don–Eyring–Polanyi–Sato (or LEPS) function of the two position variables
for a surface rearrangement, as derived from a valence-bond description of bond
breaking and forming and described in Chapter 4. However, one must keep in
mind that here pair bonding and repulsion-potential functions are affected by
the presence of the remaining surface atoms.
An example of molecular dissociative adsorption on a dielectric surface is
found in the calculation of the adsorption fragmentation of a benzene C6H6
molecule on a Si(001)-(2 × 1) surface obtained within the vdW-DF approach,
which shows two possible fragmentations. The dispersion energy contributions
appear to be relevant to clarification of which dissociation structures are the
most stable [43].
The rearrangement reaction A + BC AB + C on the surface S, for all species
already adsorbed on the surface, involves the reaction potential energy function
rct ABC , S A + BC , S
VABC , S Q = E Q; ρABC , S r −E Q; ρA + BC , S r
with Q = (QA, QB, QC, QS) the collection of all atomic positions in a reference
frame attached to the surface. The energies to the right must be obtained as
functions of the positions of A, B, and C on the surface S. In a treatment
using embedding in an indented surface S , the energies of fragments A , B ,
and C containing the species A, B, and C and neighbor surface atoms can
be obtained along reaction paths leading from A + BC to AB + C . Surface
atoms in the fragments must be allowed to rearrange too, but the atomic
354 8 Interaction of Molecules with Surfaces
structure of the indented surface S can be kept fixed if the fragments and
indentation are large enough.
Ce 2 − Ce Ce, img 4 Ce 2 3 εr + 2 1
Zc = =
4πε0 W −I 4πε0 4 εr + 1 W − I
where as before, εr is the static dielectric constant of the solid. At short dis-
tances, repulsion forces take over, while at distances larger than Zc, the potential
energy is made up of induction and dispersion terms. A potential can be con-
structed over all distances using a crossing (or switching) function around Zc, as
done in the chapter on model potentials.
The reverse charge transfer situation involves A− + M+ present at intermedi-
ate distances, going into A + M at large distances, for example for halide atoms
interacting with the surface. Transfer happens for increasing distances at the
− +
crossing value determined instead by the difference A − W in V A M Zc =
− Ce 2 4πε0 Zc + Aeff Zc −W ≈0, past which the electron moves back to
the metal.
A one-electron model extends the treatment described in Section 8.4 for phy-
sisorption, to allow here for strong coupling of molecular and metal orbitals.
The strong coupling leads to large changes in molecular orbital energy shift
and broadening. These can be obtained using a partitioning treatment, which
leads to energy-dependent level shifts and widths, and several different cases
for chemisorption due to electron transfer [8, 31]. The molecular and metal
orbitals φμ and φj, j = b, k, σ , for given electron spin quantum number, can
be taken to be orthonormalized to begin with, so that φμ φμ = δμμ and
φj φj = δjj . Molecular and metal orbitals are nonorthogonal with overlap
φμ φj = Sμj. A transformation of molecular orbitals from the set {φμ} to a
new basis set {φν} can, however, be done by symmetrical orthogonalization
[57], to obtain new molecular and metal orbitals so that φν φj = δνj, and to sim-
plify a partitioning treatment. In practice, the original overlap integrals are small
356 8 Interaction of Molecules with Surfaces
and the new functions φν can be physically related to the old φμ and assigned to
atomic positions. Partitioning relies on the introduction of projection operators
A = ν
φν φν , M = j
φj φj
ε + iη− f m Gm ε = I
where, for example, f AM = A f M . Solving formally for GMA from the second
line and substituting in the first line, one obtains
−1
1
GAA ε = ε + iη −f AA − f AM f MA
ε + iη− f MM
1 1
= − iπδ E
E + iη E
ΔA ε = f AM f MA
ε −f MM
ΓA ε = πf AM δ E −f MM f MA
8.5 Chemisorption 357
−1
φν GAA ε φν = ε + iη− fνν − Δν ε + iΓν ε
and its imaginary part as
1 Γν ε
gA, M ε =
π ν ε −fνν − Δν ε 2
+ Γν ε 2
2
Γν ε = π f
j νj
δ ε − fjj ,
2 ∞
fνj Γν ε
Δν ε = j ε−f
= dε
jj −∞ ε −ε
where the summation over j = b, k involves also integration over values of the
wavevector k. This density of states can be considered for each separate state φν
of energy fνν to find how the molecular energy levels have changed due to inter-
actions with the metal. The density of states function gA,M(ε) can be used to cal-
culate a potential energy function VAM R,QA ,QM similar to the one given in
Section 8.4.1 within a one-electron treatment of interaction energies at surfaces,
possibly including now the change in residual correlation energies, by means of
calculated for varying atomic positions. The partitioning treatment has also
been developed for time-dependent equations that arise in the description of
electron transfer dynamics and has been applied to calculations of their prob-
abilities in collisions of ions with metal surfaces [58, 59].
Given the forms of chemisorption shifts and widths Δν(ε) and Γν(ε) as they
follow from the location and widths of energy bands in the metal and from
the strength of the molecule–metal coupling energies fνj, it is possible to classify
the changes of A energy levels due to chemisorption as giving either sharp iso-
lated levels or shifted and broadened ones that are overlapping [8, 31]. For iso-
lated levels, it is convenient to find the roots εν of the equation ε − fνν − Δν(ε) = 0,
and to expand the density of states around each root value. Letting ε1 and ε2 be
the lower and higher limits of the range of Γν(ε), and considering first weak cou-
pling and only one root εν near fνν, the new level εν = fνν + Δν εν has no width if
εν < ε1 or εν > ε2 , and has a small width Γν(εν) if ε1 ≤ εν ≤ ε2 .
When the coupling is strong and level widths overlap, an alternative treatment
is to first combine an adsorbate orbital with one or more substrate orbitals, such
358 8 Interaction of Molecules with Surfaces
as a 3s- or 3p- orbital of Si and a d-orbital of Ni(001) for Si adsorbed on Ni, to form
surface (bonding and anti-bonding) orbitals φs, s = a, b and to calculate their ener-
gies εs. Allowing for their interaction with the remaining energy band orbitals of
the metal substrate, the partitioning procedure provides their shifts Δs and widths
Γs. This is a generalization of tight-binding models, valid for adsorbates and their
energies near metal surfaces [7, 8].
In applications, the matrix elements of f m can be parametrized at the Har-
tree–Fock level using ionization and affinity energies for the molecule and
energy bandwidths for the metal, to reach qualitative conclusions about chem-
isorption energies. Parametrizations of chemisorption energy expressions con-
taining the coupling energy between adsorbate and substrate orbitals have also
been developed in semiempirical models and have been used in treatments of
catalysis, at surfaces of simple and transition metals [23].
Some of the considered systems have been hydrogen, oxygen, and alkali atoms
adsorbed on light metals and transition metals and CO adsorbed on transition
metals [8, 60]. An example is provided by the chemisorption of O atoms on the
Ni(111) and Ni(001) surfaces. The coordination number of O with a Ni surface
atom changes from three to four between the surfaces, respectively, with a
change of the amount of electronic charge transfer between O and Ni, in a
one-electron description. This can qualitatively explain the corresponding
changes in binding energy and bond distance for the two surfaces [8, 55]. Cor-
rect qualitative conclusions on chemisorption energy trends can be obtained
with one-electron treatments, but accurate values of chemisorption energies
and potential energy changes with the relative position of the molecule near
the surface require a many-electron treatment, or a DFT treatment including
dispersion energies.
At intermediate and shorter distances, the potential energy is shaped by elec-
tron exchange and correlation, while large distances involve electrostatic,
induction, and dispersion energies. Many-electron treatments can be done
within DFT extended to include long-range dispersion energies, or more accu-
rately with atomic lattice models for the metal slab and adsorbate. The roles of
ionization and affinity electron energies were found in early DFT calculations,
with electronic density distributions for Li, Si, and Cl adsorbed on a metal with
the density of Al corresponding to a spread radius rs = 2.0 au, and the jellium
model for the metal, and are shown in Figure 8.6 [61]. Electronic rearrangement
increases that electronic density for Li and Si in the region between atom and
surface, and decreases it between Cl and the surface. Potential energies of inter-
action follow from the subtraction of energy density functionals, as previously
given, and display bonding minima and equilibrium chemisorption distances
due to the attraction effect of long-range induction forces.
The main features of the DFT results for chemisorption can be reproduced by
an effective-medium treatment of chemical binding of an atom A to a surface M
with a realistic inhomogeneous electronic distribution ρAM r [35], embedding
8.5 Chemisorption 359
Total
1a.u.
superposition
Total minus
XX
XX
XX
XX
XX
XX
XX
XX
XX
XX
XX
XX
XX
XX
XX
metal
Bare
Figure 8.6 Electron density contours for chemisorption of Li, Si, and Cl on a substrate with the
electron density of Al, corresponding to a spread radius rs = 2.0 au (or 0.03 electrons/bohr3). The
middle panels (the second row) show the total densities minus the superposition of the atom
and metal densities and the extent of electronic charge transfer, with full lines indicating a net
increase in density and dashed lines a decrease. The bottom panels show the bare metal
electron density near its surface. Source: from Ref. [61]. Reproduced with permission of Elsevier.
included in the cited work, may, however, be needed to obtain accurate values
for activation barriers.
Another example is given by the diatomic dissociation and atomic adsorption
of O2 on an aluminum surface, for which one wants to establish whether there is
an activation energy and if so what is its origin, insofar it involves in principle
electronic charge and spin rearrangements. The orientation of O2 is again fully
specified by six atomic coordinates, and the aluminum surface has different
shapes for different Miller indices. A full treatment of the energetics of the dis-
sociation would be very laborious, but insight can be extracted considering only
approaches of O2 with its axis parallel or perpendicular to the surface of Al(111)
and at locations with chosen surface symmetries such as on-top, fcc, hcp, or
bridge of nearby surface Al atoms.
Figure 8.7 shows isocontours of energy versus O – O and O – Al distances
obtained by an embedding treatment [51, 52], with a cluster including O2
and several Al atoms embedded in an environment provided by an indented
Al surface. Results are presented in Figure 8.7a for an O2 perpendicular to
the surface and for the fcc symmetry, described with the PBE density functional,
while Figure 8.7b shows the results for the cluster and a CW treatment imple-
mented with CASSCF + PT2 calculations [50], which include long-range inter-
actions. The isocontours differ, with the CW results showing an activation
(a) (b)
DFT fcc ⊥ Embedded CW fcc ⊥
5 5 5 5
2 2
E(eV)
E(eV)
4 4 4 4
LA1–O2 (Å)
LA1–O2 (Å)
–1.3 –1.3
3 3 3 3
2 2 2 2
1 1 1 1
1.2 1.6 2.0 2.4 1.2 1.6 2.0 2.4
LO–O (Å) LO–O (Å)
energy due to electronic charge rearrangement, which varies with the location
on the surface and is around 600 meV (57.891 kJ mol−1) for the on-top position
with R(O – Al) 280 pm, in accordance with experimental results. Additional
calculations of the PES for O2 on Al(111) have been used to parametrize a flex-
ible periodic LEPS potential energy function, suitable for reaction dynamics on a
periodic surface structure. It is constructed as in the LEPS formulation from
valence-bond Coulomb and exchange energies using Morse functions and
parameters, which are, however, expanded in Fourier series with the periodicity
of two-dimensional lattice distances corresponding to the surfaces’ Miller indi-
ces [66].
A surface chemical reaction such as A + BC AB + C, with A, B, and C here
taken to be atoms to simplify, starts with the two reactant species adsorbed at
positions RA, ads and RBC , ads relative to the surface, and proceeds on a PES
BC , M
−E QB ,QC ,QM ; ρBC , M r
with Q = (QA, QB, QC, QM). The first term to the right depends on positions of all
participating atoms including metal atoms, the second term is the energy of an
adsorbed atom A, and the last term is the energy of an adsorbed BC diatomic. As
in the case of dissociation, the set of all metal atom positions can be separated
into a small subset for active atoms that undergo large displacements during
reactions, and a larger subset for the remaining atoms that stay close to their
original equilibrium positions. These two subsets define a cluster and its envi-
ronment, respectively, and can be treated within embedding methods as
described.
One of the concerns about using DFT for chemisorption energies is that the
treatment can lead to fractional electronic charge transfer for each reaction frag-
ment, with a related unphysical energy dependence on a continuously varying
number of electrons, instead of physically meaningful fragments with integer
numbers of electrons. As described in Chapters 6 and 7, an alternative treatment
can be based on the introduction of an energy Etot written as a functional of elec-
tron orbitals as explained in Section 6.4, for the total system, to be obtained
within an optimized effective potential procedure for interacting adsorbate
K
and substrate fragments K. The functional Etot ψj r,ζ of a set of j = 1 to NK
K
spin orbitals ψ j for each fragment K can be constructed to describe electrons
moving in an effective potential Vemb r , optimized by energy minimization, to
OEEP
derive an optimized embedding effective potential (OEEP) Vemb r suitable
diss
for chemisorption phenomena, with potential energy functions VAB, M Q
rct
and VABC , M Q obtained as sums of energies for fragments with integer num-
bers of electrons in each one [52, 67].
8.6 Interactions with Biomolecular Surfaces 363
Water
Hydrophilic
head
Hydrophobic
tail
Water
Coul Ci Cj
vIJ Rij = fd Rij
i I, j J 4πε0 εr Rij
ind Ci αj + αi Cj
vIJ Rij =− fd Rij
i I, j J 4πε0 εr 2 Rij 4
where εr is the dielectric constant of the medium. The first term in the summa-
tion describes dispersion energies, and the Coulomb and induction terms have
been given as sums of pair interactions between point charges Ck and polariz-
abilities αk in fragments = , located at the relative distance Rij. They con-
tain a damping function fd(Rij) ≈ 0 for Rij 0 at short distances, as described in
Chapters 4 and 7. These point charges and polarizabilities are approximate
values for the actual spatial distributions of charges and polarizabilities, which
furthermore change as conformations vary. With reference to citations in [69],
well-tested force fields (FFs) are available from GROMOS, where CHn, n = 1−3,
groups of atoms are treated as a single particle, as recently improved for lipids
[72], and from the all-atoms CHARMM FF as recently improved for lipids [73],
AMBER with improvements incorporated for lipids, MARTINI with a coarse-
grained model [74], and other ongoing developments for polarizable FFs suit-
able for lipids.
The very large number of atoms to be considered in the calculation of struc-
tural and dynamical properties of a biomembrane interacting with a large
molecule presents a challenge in the computational modeling of the complex.
This can be dealt with grouping several atoms into chemically meaningful
fragments, which become basic constituents, reducing the total number of
degrees of freedom in the whole system. It requires parametrization of frag-
ment properties independently calculated with many-electron treatments.
Consistent ways to do this employ multiscale techniques based on a full
atomic description and proceed from the ground-up by coarse-graining
many-atom systems into fragments (or new particles), to describe condensed
matter and biomolecular systems with fewer interacting particles [75–79].
Optimal ways to extract parameters for the fragment structures and their
interaction, from extensive atomistic calculations, can be found using
machine learning of artificial intelligence rules from big data collected for
biomolecules [80–82].
A quantitative coarse-graining (CG) treatment of the modeling of a large
many-atom system by a smaller collection of particles can be done starting from
a mapping of the set RA = r 3NA of all the 3NA atomic position coordinates into
a set of 3NP model particle position coordinates RP = r 3NP plus NL model
parameters λ, such as point charges and polarizabilities, collected in the set
of particle variables (Λ = RP,λ). The mapping is of form
Λ = M RA
366 8 Interaction of Molecules with Surfaces
where Λ and RA are column matrices and M is a matrix of order 3NA × (3NP +
NL). This describes a general case where the parameters λ are functions of the
original atomic positions and change with the conformation of molecules.
A simpler treatment is usually developed where the particle parameters do
not change, in which case a simpler mapping is done between atomic positions
and particle positions of the form RP = MPA RA ; λ .
The construction of particles as groups of atoms must be done so that their
interactions reproduce statistical properties of the original many-atom system.
This can be enforced using the probability distribution pA RA of the atomic
ensemble, to obtain the probability distribution of particles by means of
pP Λ = dRA pA RA δ M RA − Λ
UP Λ, β = −β − 1 ln ZP β dRA pA RA ,β δ M RA − Λ
References
1 Atkins, P. and De Paula, J. (2010). Physical Chemistry, 9e. New York: W. H.
Freeman and Co.
2 Kittel, C. (2005). Introduction to Solid State Physics, 8e. Hoboken, NJ: Wiley.
3 Somorjai, G.A. (1972). Principles of Surface Chemistry. Englewood Cliffs, NJ:
Prentice-Hall.
4 Somorjai, G.A. (1994). Introduction to Surface Chemistry and Catalysis. New
York: Wiley-Interscience Publication.
5 Israelachvili, J. (2011). Intermolecular and Surface Forces, 3e. New York: Elsevier.
6 Lang, N.D. (1973). The density-functional formalism and the electronic
structure of metal surfaces. In: Solid State Physics, vol. 28 (eds. H. Ehrenreich, F.
Seitz and D. Turnbull), 225. New York: Academic Press.
7 Martin, R.M. (2004). Electronic Structure: Basic Theory and Practical Methods.
Cambridge, England: Cambridge University Press.
8 Desjonqueres, M.C. and Spanjaard, D. (1995). Concepts in Surface Physics, 2e.
Berlin, Germany: Springer-verlag.
9 Lang, N.D. (1983). Density functional approach to the electronic structure of
metal surfaces and metal-adsorbate systems. In: Theory of the Inhomogeneous
Electron Gas (eds. S. Lundqvist and N.H. March), 309. New York: Plenum Press.
10 May, V. and Kuhn, O. (2000). Charge and Energy Transfer Dynamics in
Molecular Systems. Berlin: Wiley-VCH.
11 Mukamel, S. (1995). Principles of Nonlinear Optical Spectroscopy. Oxford,
England: Oxford University Press.
12 Micha, D.A. (2015). Generalized response theory for a photoexcited many-atom
system. In: Adv. Quantum Chemistry, vol. 71, Chapter 8 (eds. J.R. Sabin and
R. Cabrera-Trujillo), 195. New York: Elsevier.
13 Yu, P. and Cardona, M. (2005). Fundamentals of Semiconductors: Physics and
Materials Properties, 3e. Berlin, Germany: Springer-Verlag.
14 Haug, H. and Koch, S.W. (2004). Quantum Theory of the Optical and Electronic
Properties of Semiconductors. World Scientific.
368 8 Interaction of Molecules with Surfaces
15 Ashcroft, N.W. and Mermin, N.D. (1976). Solid State Physics. London, England:
Thomson.
16 Landau, L.D. and Lifshitz, E.M. (1960). Electrodynamics of Continuous Media.
London: Pergamon Press.
17 Jackson, J.D. (1975). Classical Electrodynamics. New York: Wiley.
18 Mavroyanis, C. (1963). The interaction of neutral molecules with dielectric
surfaces. Mol. Phys. 6: 593.
19 McLachlan, A.D. (1964). Van der Waals forces between an atom and a surface.
Mol. Phys. 7: 381.
20 Zaremba, E. and Kohn, W. (1976). van der Waals interaction between an atom
and a solid surface. Phys. Rev. B 13: 2270.
21 Kramer, H.L. and Herschbach, D.R. (1970). Combination rules for van der Waals
force constants. J. Chem. Phys. 53: 2792.
22 Bortolani, V., March, N.H., and Tosi, M.P. (eds.) (1990). Interaction of Atoms and
Molecules with Solid Surfaces. New York: Plenum Press.
23 Greeley, J., Noerskov, J.K., and Mavrikakis, M. (2002). Electronic structure and
catalysis on metal surfaces. Annu. Rev. Phys. Chem. 53: 319.
24 Huang, P. and Carter, E.A. (2008). Advances in correlated electronic
structure methods for solids, surfaces, and nanostructures. Annu. Rev. Phys.
Chem. 59: 261.
25 Wolken, G.J. (1976). Scattering of atoms and molecules from solid surfaces. In:
Dynamics of Molecular Collisions Part A (ed. W.H. Miller), 211. New York:
Plenum Press.
26 Goodman, F.O. and Wachman, H.Y. (1976). Dynamics of Gas-Surface Scattering.
New York: Academic Press.
27 Celli, V., Eichenauer, E., Kaufhold, A., and Toennies, J.P. (1985). Pairwise
additive semi ab initio potential for the elastic scattering of He atoms from the
LiF(001) crystal surface. J. Chem. Phys. 83: 2504.
28 Zangwill, A. (1988). Physics at Surfaces. Cambridge, England: Cambridge
University Press.
29 Levine, I.N. (2000). Quantum Chemistry, 5e. Upper Saddle River: Prentice-Hall.
30 Parr, R.G. and Yang, W. (1989). Density Functional Theory of Atoms and
Molecules. Oxford, England: Oxford University Press.
31 Newns, D.M. (1969). Self-consistent model of hydrogen chemisorption. Phys.
Rev. 178: 1123.
32 Maurer, R.J., Ruiz, V.G., Camarillo-Cisneros, J. et al. (2016). Adsorption
structures and energetics of molecules on metal surfaces: bridging experiment
and theory. Prog. Surf. Sci. 91: 72.
33 Landman, U. and Kleiman, G.G. (1977). Microscopic approaches to
physisorption. In: Surface and Defect Properties of Solids, vol. 6 (eds. M.W.
Roberts and J.M. Thomas), 1. London: Chemical Society.
34 Gordon, R.G. and Kim, Y.S. (1972). Theory for the forces between closed shell
atoms and molecules. J. Chem. Phys. 56: 3122.
References 369
Index
charge density fluctuation 260, 262, computing times 209, 210, 350
291, 329 Condon approximation 176–180
charge-image potential energy 325 Condon states 179
CHARMM FF 365 conductive materials 322
CH2=[CH–CH=]kCH2 284 configuration interaction (CI) 99, 100,
chemical reactivity 2, 240, 299 113, 122, 165, 166, 198, 201,
chemisorption 310 209–215, 360
cluster model 350 configuration state function
density of states 357 (CSF) 210–215
on metals 350, 351 conical intersections 15, 17, 19, 181,
shifts 357 184–189, 192, 303
widths 357 conjugated polyenes 284
CH3(CH2)N-2CH3 281 consistent basis sets 206
CH3NH 303 continuum model 39, 144, 332–337
C4H10 on Cu(111) 347 contracted gaussians 206
C6H6 on Si(001)-(2×1) 353 CO on Cu(001), Cu(111) 349
chronological Table 2 coordinates scaling 167
CISD treatment 210 core point 115
Cl + H2 149 coronene dimer 293
closure approximation 58–59, 263 correlated embedding
CO2 22, 59, 60, 352 wavefunction 347
(CO2)2 227 correlation consistent orbitals 207
coarse-grain entropy 366 correlation diagram 95, 96
coarse-graining 144, 274, 277, correlation potential 216
365, 367 Coulomb forces 3, 4, 164,
coarse-grain mapping 365 166, 230
coherent states 68 Coulomb functional 89
cohesive energy 30, 106, 127–130, Coulomb integral 97, 99, 125, 211
132, 363 counterpoise (CP) correction 208
collisional cross sections 27–28 coupled cluster treatment 198,
combination rules 82–83, 91, 94, 201, 218
116–117, 264 coupled stretching bonds 146
short distances 91 covalent-ionic function 99
compensating phase functions 191 covalent-ionic states 223
complementary damping 118 crossing point 176, 182, 355
complete active space 214 crystalline benzene 289
complete active space self-consistent crystal structure 127, 290
field (CASSCF) 210, 214 CsCl(s) 129
complex-valued crossing 183 Cu(s) 128, 316, 322, 323, 337,
complimentary 349, 360
antisymmetrizer 199, 225 cubic crystalline lattice 141
complimentary operator 199 curvature 115
compound AB 196, 240 cyclic compounds 283
376 Index
reduced variables 115 228, 241, 265, 269, 277, 293, 311,
reference frame 6, 13, 43, 52–54, 64, 312, 331, 347, 350
76, 79, 80, 94, 111, 122, 131, 139, Si(s) 127
158–160, 169, 230, 283, 291, Si structures 151
353, 360 size consistency 101, 112, 196, 222
refractive index 26 size extensivity 196, 198, 217, 218
relativistic Hamiltonian 103 Slater determinant 99
renormalized susceptibility 328 Slater-type orbitals (STOs) 206
representation transformation 174 smooth coupling 175
resolvent operator 41, 50, 83, 104, smooth diabatic potentials 176
202, 356 solid Si 315
resonance energy 68, 74 solute pair correlation function 138
resonance interaction 68, 98 solute pair distribution function 137
resonance peak 49 solute–solvent interactions 265–268
response function 44, 46, 50, 51, 57, space-fixed (CMN-SF) frame 79, 158
235, 236, 243, 256, 319, 320 space groups 54, 310
restricted Hartree–Fock (RHF) 100, space symmetry 200
198, 211–213, 219 space symmetry projection 201
restricted open shell HF (ROHF) 212 spatial and spin symmetry 112
retardation effects 8, 102–103, spatial correlation function 234
327, 328 special points 18, 115, 181, 182
retarded susceptibility 46, 47, 51, 319 spectroscopy 28–30, 113, 116, 145,
rotation matrices 53 170, 175, 176, 179, 189, 246
rotation matrix 80, 158 spherical components 37, 38, 43, 53,
Rydberg potential 116 57, 70, 75–76, 78, 79
spherical multipoles 37, 54, 57, 69, 75
s spherical waves 77
saddle point 14, 15, 18, 21, 121, spin couplings 9
126, 313 spin density matrix 229
scalar potential 40, 191, 265 spin-dependent potential 126
scaled atomic polarizability 293 spin-orbit coupling 10, 78, 102–103,
scaled dynamical susceptibility 235 183, 196, 243–247
scaled response function 235 spin-other-orbit couplings 244
scaling in CCSD(T) 221, 222 spin-projection rising and
screened electron–electron lowering 200
interaction 241 spin recoupling 246
second-order energy 67, 83, 217 spin-space projection operator 200
selection rules 54, 55, 57, 68, 74, 76, spin states 90, 101, 126, 166, 183,
78–80, 84 196, 340
SH2 185 spline function 119
short-range (SR) 6, 7, 27, 69, 91, 101, spontaneous emission 49
103, 114, 117, 118, 120, 122, 124, spread radius rs 316, 345, 358, 359
129, 132, 133, 136, 143, 144, 146, static susceptibility 262, 264, 280
384 Index