Sunteți pe pagina 1din 396

Hydrogen Bonding:

A Theoretical Perspective

STEVE SCHEINER

Oxford University Press


HYDROGEN BONDING
TOPICS IN PHYSICAL CHEMISTRY
A Series of Advanced Textbooks and Monographs

Series Editor, Donald G. Truhlar

F. Iachello and R. D. Levine, Algebraic Theory of Molecules


P. Bernath, Spectra of Atoms and Molecules
J. Cioslowski, Electronic Structure Calculations on Fullerenes
and Their Derivatives
E. R. Bernstein, Chemical Reactions in Clusters
J. Simons and J. Nichols, Quantum Mechanics in Chemistry
G. A. Jeffrey, An Introduction to Hydrogen Bonding
S. Scheiner, Hydrogen Bonding: A Theoretical Perspective
HYDROGEN BONDING
A Theoretical Perspective

STEVE SCHEINER

New York Oxford


Oxford University Press
1997
Oxford University Press
Oxford New York
Athens Auckland Bangkok Bogota Bombay Buenos Aires
Calcutta Cape Town Dar es Salaam Delhi Florence Hong Kong
Istanbul Karachi Kuala Lumpur Madras Madrid Melbourne
Mexico City Nairobi Paris Singapore Taipei Tokyo Toronto Warsaw
and associated companies in
Berlin Ibadan

Copyright © 1997 by Oxford University Press, Inc.


Published by Oxford University Press, Inc.,
198 Madison Avenue, New York, New York 10016
Oxford is a registered trademark of Oxford University Press.
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,
without the prior permission of Oxford University Press.

Library of Congress Cataloging-in-Publication Data


Scheiner, Steve.
Hydrogen bonding : a theoretical perspective / Steve
Scheiner.
p. cm. — (Topics in physical chemistry)
Includes bibliographical references and index.
ISBN 0-19-509011-X
1. Hydrogen bonding. 2. Quantum chemistry. I. Title.
II. Series.
QD461.S36 1997
541.2'26—dc21 96-48698

9 8 7 6 5 4 3 2 1
Printed in the United States of America
on acid-free paper
To my mother:

You will not be forgotten


This page intentionally left blank
Preface

hy a theoretical perspective of hydrogen bonding? After all, there have been numer-
W ous texts, monographs, and compilations written about hydrogen bonds over the
1-7
years . Much of this literature has taken the viewpoint of the crystallographer or spec-
troscopist, with the emphasis placed on structural aspects of the H-bonded complexes in
their equilibrium geometries or their modes of internal vibration. Quantum chemical cal-
culations offer a rich source of supplementary information concerning this phenomenon.
First, much of the literature that has accumulated over the years concerning H-bonds has
been gathered in solvents of various types. One is then presented with the problem of sep-
arating the intrinsic properties of the complex under investigation from the perturbations
incurred by interactions with the solvent medium. In contrast, in vacuo investigation of sys-
tems, in isolation from surroundings, is a definite strength of computational methods, as
they are free of complicating solvent effects.
While spectroscopic data can provide details of the equilibrium geometries of H-bonded
complexes, it has proven difficult to extract the energetics of the interaction. Growing so-
phistication of computer hardware and more efficient algorithms have dramatically en-
hanced the level of accuracy that can be expected from the calculations, which can address
the energetics directly. Theoretical approaches offer an additional dividend: dissection of
the interaction energy into various physically meaningful components, which can provide
insights into the fundamental nature of the interaction.
Whereas experimental data are largely relevant to the global minimum in the potential
energy surface, computational methods can map entire domains of this surface. One can lo-
cate secondary minima and stationary points of higher order in addition to the minima. It is
also feasible to identify interconversion pathways from one minimum to another, along with
magnitudes and shapes of energy barriers along these paths.
Quantum chemical methods have capabilities that are not limited to energetics. It is a
straightforward matter to examine in detail the electronic redistributions that accompany
the formation of the H-bond. The precise nature of the displacements of the nuclei that com-
viii Preface

prise the normal vibrational modes can be elucidated with an accuracy that eludes analysis
of purely experimental data. This information presents the possibility of genuine under-
standing of the perturbations in vibrational spectra that accompany formation of a H-bond.
Thus, quantum chemistry has a great deal to offer the field of hydrogen bonding. And
indeed, the body of pertinent calculated data has been growing at a rapid rate. This book is
intended to digest this vast amount of information and organize it into a form that is un-
derstandable to a general reader who is interested in hydrogen bonding but has little in the
way of formal training in theoretical calculations. A brief introduction is presented so that
the reader might obtain some appreciation of the basic ideas behind the computation of var-
ious properties, and to prepare for the jargon one is likely to encounter in the literature. This
explanation is intentionally brief and simplified: the reader is referred to many fine reviews,
monographs, and texts if interested in more detail about these methods or the underlying
theory. A glossary of common abbreviations is included to aid the reader in recalling the
definition of each term as it appears.
While quantum chemical studies of H-bonded systems date back to the 1960s, this book
concentrates on calculations made principally since 1980 or so. These results are more re-
liable in a number of respects, particularly from the standpoint of higher levels of theory
and the characterization of stationary points on the potential energy surface. (In fact, an ear-
lier text on the subject of H-bonding3 had compiled a listing of theoretical studies in the
1960-73 timeframe for those interested in some of the earlier work.) The group of articles
selected is not an exhaustive one; rather, I have culled those that were considered best able
to illustrate a given point, taking into account also their level of accuracy.
The emphasis in this book is on ab initio calculations. Semiempirical methods are not
designed or parametrized to treat intermolecular interactions well. Indeed, the original for-
mulations of some semiempirical methods did not predict H-bonds to exist at all. There have
been attempts to patch them up over the years so as to permit a modicum of attractive po-
tential where a H-bond is expected. However, one cannot depend on the reliability of such
approaches. Another important distinction between ab initio and semiempirical methods is
that the former will, at least in principle, approach reality as the basis set is enlarged and as
the treatment of electron correlation is made more complete. It is thus possible to estimate
how close one is to this asymptote, even if enormous basis sets with high orders of corre-
lation are not feasible for a given chemical system, by monitoring the results of the calcu-
lations as the level of ab initio theory is improved, one step at a time. If the data are stable
to additional improvements of the method, a certain measure of confidence may be attached
to the results. Such is not the case with semiempirical methods wherein a single result is
obtained, with no real way to improve upon it or test its reliability against a higher level of
comparable theory.
A new type of method has been gaining popularity very rapidly. Density functional the-
ory (DFT) bypasses the conventional concept of individual molecular orbitals and instead
optimizes the total electron density, including all electrons. Based originally on some con-
cepts from solid-state physics, the method scales to a lower order with respect to the num-
ber of atoms or electrons, as compared to conventional ab initio theory where the compu-
tational effort rises roughly as the fourth or higher power of N. Consequently, this new
approach has enormous potential to treat systems that are far too large for ab initio meth-
ods to examine. The DFT approach is maturing quickly; it seems that major new develop-
ments appear in the literature on a monthly basis. Along with these enhancements in the
methodology have come comparisons with data generated by the older and more reliable
techniques. Results for hydrogen bonds have been mixed so far; there is still some question
Preface ix

as to which kinds of functionals are most appropriate. The next few years will likely wit-
ness improvements to the point where DFT calculations become competitive with conven-
tional ab initio methods in terms of accuracy. However, because this method is relatively
new, and has received only limited testing to this point, I have for the most part avoided any
extensive discussion of the DFT results here. It is not unlikely that shortly after this book
appears, the situation will have changed and the time will be right for an entire text devoted
to application of density functional methods to H-bonded systems.
The organization of this text is as follows. Chapter 1 presents the reader with a capsule
summary of quantum chemical methods. The intent is not to make the reader an expert in
various theoretical approaches but rather to provide a basic understanding of the techniques,
their strengths, and limitations. This chapter introduces and explains much of the jargon and
includes a glossary of abbreviations that a reader is likely to encounter in the original liter-
ature. Chapter 1 also provides a tentative definition of a hydrogen bond, and how quantum
chemical calculations can address the various contributing factors. A simple example is pro-
vided to illustrate the central points for later reference with more complicated systems. This
chapter also delves into some detail on the most common sources of error encountered in
computations of this sort.
Chapter 2 surveys the field of H-bonds that have been studied to date, focusing on
smaller molecules for which the calculations are most definitive. These small molecules
serve as models of the functional groups that occur in larger systems as well. This chapter
focuses on the energetics of various combinations of partners, and the details of their equi-
librium geometries. Emphasis is placed on systematic relationships between the properties
of the constituent molecules and the nature of each H-bond. Also discussed are the pertur-
bations that occur in each molecule as the H-bond is formed.
One way in which the fundamental forces responsible for the formation of a H-bond can
be probed is by examining of the force field that restores the equilibrium geometry after
small geometrical distortions. This field is directly manifested by the normal vibrational
modes that exhibit themselves in the vibrational spectrum of the complex. Chapter 3 is
hence devoted to a discussion of the vibrational spectra of H-bonded complexes, and what
can be learned from their calculation by quantum chemical methods. While the vibrational
frequencies are directly related to the forces on the various atoms, the intensities offer a
window into the electronic redistributions that accompany the displacement of each atom
away from its equilibrium position, so vibrational intensities are also examined in some de-
tail. Of particular interest are relationships between the vibrational spectra and the energetic
and geometric properties of these complexes.
The vibrational spectrum provides information about the potential energy surface in the
immediate vicinity of the minimum. Chapter 4 broadens the scope by considering wider
swaths of the surface. Large deviations from the equilibrium geometry are examined, some
of which take the system to a secondary minimum on the surface. This chapter discusses
paths between various minima that pass through stationary points of higher order, and pro-
vides a broad picture of the general topology of the entire surface.
Hydrogen bonding is particularly important in condensed phases where it can signifi-
cantly affect such properties as boiling point or crystal structure. H-bonds seldom occur in
isolation in condensed phases but are commonly part of a chain of molecules held together
by such interactions. The effect of one H-bond on another is the subject of Chapter 5, deal-
ing with cooperativity phenomena. The source of this effect is probed, in terms of electronic
redistributions and various contributors to the full energetic interaction. The magnitude of
cooperativity is considered, as a function of the number of contiguous H-bonds, and the
x Preface

asymptotic limit of an infinitely long chain of H-bonds. Another central question is whether
it is worth the energetic expense of bending the H-bonds in a chain, so as to enable the two
ends to approach one another to form an additional H-bond and a cyclic structure of the en-
tire chain.
The classification of any given interaction as a H-bond is not always a trivial matter.
There are many situations which have certain characteristics in common with a H-bond but
others are lacking. Chapter 6 considers a number of interactions whose designation as an
"official" H-bond could be called into question. Some of these situations include the pos-
sibility of the C—H group acting as a proton donor or an electronegative atom in a cova-
lent bond of only small polarity serving as a suitable proton acceptor. Also examined is the
question as to whether an interaction in which one or both of the partners bears an electric
charge should be considered as a true H-bond. Interactions of this type comprise some of
the strongest H-bonds known. The characteristics of the proton transfer potentials in these
ionic H-bonds are particularly interesting, sometimes containing two wells while only a sin-
gle minimum is present in other cases. The chapter explores the relationships between the
strength of the H-bond and the intrinsic acidity and basicity of the subunits. A detour is taken
to explore the intriguing question as to whether there is a catalytic advantage to one of the
two lone pairs of an oxygen atom in the carboxylate group.

References
1. G. C. Pimentel and A. L. McClellan. The Hydrogen Bond. Freeman: San Francisco, 1960.
2. S. N. Vinogradov and R. H. Linnell. Hydrogen Bonding. Van Nostrand-Reinhold: New York,
1971.
3. M. D. Joesten and L. J. Schaad. Hydrogen Bonding. Marcel Dekker: New York, 1974.
4. P. Schuster, G. Zundel, and C. Sandorfy, Eds. The Hydrogen Bond: Recent Developments in The-
ory and Experiments. North-Holland Publishing Co.: Amsterdam, 1976.
5. P. Schuster, Ed. Hydrogen Bonds. Vol. 120. Springer-Verlag: Berlin, 1984.
6. G. A. Jeffrey and W. Saenger. Hydrogen Bonding in Biological Structures. Springer-Verlag:
Berlin, 1991.
7. D. A. Smith, Ed. Modeling the Hydrogen Bond. Vol. 569. American Chemical Society: Wash-
ington, D.C., 1994.
Contents

Abbreviations xvii

I QUANTUM CHEMICAL FRAMEWORK


1.1 Quantum Chemical Techniques 3
1.1.1 Basis Sets 4
1.1.2 Electron Correlation 7
1.1.3 Geometries 10
1.2 Definition of a Hydrogen Bond 11
1.2.1 Geometry 12
1.2.2 Energetics 13
1.2.3 Electronic Redistributions 13
1.2.4 Spectroscopic Observations 13
1.3 Quantum Chemical Characterization of Hydrogen Bonds 14
1.3.1 H-bond Geometries 15
1.3.2 Thermodynamic Quantities 15
1.3.3 Electronic Redistributions 18
1.3.4 Spectroscopic Observations 18
1.4 A Simple Example 19
1.5 Sources of Error 22
1.6 Basis Set Superposition 23
1.6.1 Secondary Superposition 24
1.6.2 Important Properties of Superposition Error 25
1.6.3 Historical Perspective 25
1.7 Energy Decomposition 28
1.7.1 Kitaura-Morokuma Scheme 32
xii Contents

1.7.2 Alternate Schemes 34


1.7.3 Perturbation Schemes 37

2 GEOMETRIES AND ENERGETICS


2.1 XH ZH3 53
2.1.1 BSSE 56
2.1.2 Substituent Effects 60
2.2 XH YH2 61
2.2.1 Comparative Aspects 62
2.2.2 Angular Features 64
2.2.3 Alternate Complexes and Geometries 66
2.2.4 Energy Components 67
2.3 HYH ZH3 69
2.3.1 Substituents 71
2.4 XH XH 71
2.5 HYH YH2 77
2.5.1 Binding Energy of Water Dimer 78
2.5.2 Complexes Containing H2S 79
2.5.3 Substituent Effects 81
2.6 (ZH3)2 84
2.7 Carbonyl Group 89
2.7.1 Substituent Effects 93
2.8 Carboxylic Acid 94
2.8.1 Carboxylic Acid Dimers 99
2.9 Nitrile 101
2.10 Imine 103
2.11 Amide 105
2.11.1 Interaction with Carboxylic Acid and Ester 112
2.12 Nucleic Acid Base Pairs 113
2.13 H-Bonds versus D-Bonds 118
2.13.1 Water Molecules 120
2.14 Summary 121

3 VIBRATIONAL SPECTRA
3.1 Method of Calculation 139
3.2 Accuracy Considerations 141
3.3 (HX)2 143
3.4 H3Z HX 148
3.4.1 Analysis of Intensities 150
3.4.2 Anharmonicity 152
3.4.3 Other Properties 154
3.4.4 Relationship between H-Bond Strength and Spectra 155
3.5 H2Y...HX 156
3.5.1 Alkyl Substituents 159
3.5.2 Other Properties 159
Contents xiii

3.6 H2Y...HYH 160


3.6.1 Polarizability 162
3.6.2 Comparison between (H2O)2 and (H2S)2 163
3.6.3 Effects of Electron Correlation and Matrices 166
3.6.4 Substituent Effects 168
3.6.5 NMR spectra 171
3.7 Expected Accuracies 171
3.7.1 HF Dimer 171
3.7.2 Water Dimer 173
3.8 HYH ... NH 3 175
3.9 (NH3)2 177
3.10 Carbonyl Oxygen 179
3.10.1 Relationship between E and v 180
3.10.2 Formaldehyde + Water 181
3.10.3 Formaldehyde + HX 182
3.11 Imine 184
3.12 Nitrile 185
3.12.1 Correlation and Anharmonicity 186
3.12.2 HCN as Proton Donor 191
3.12.3 HCN Dimer 192
3.13 Amide 195
3.14 Summary 197

4 EXTENDED REGIONS OF POTENTIAL ENERGY SURFACE


4.1 Ammonia Dimer 208
4.2 H 2 O ... HX 209
4.3 (HX)2 209
4.3.1 Anisotropies of Energy Components 210
4.3.2 Interconversion Pathways 212
4.3.3 HC1 Dimer 213
4.4 Water Dimer 215
4.4.1 Characterization of Possible Minima and Stationary Points 215
4.4.2 Components of the Interaction Energy 220
4.5 Carbonyl Group 223
4.6 Amines 225
4.7 Summary 226

5 COOPERATIVE PHENOMENA
5.1 HCN Chains 232
5.1.1 Geometries 232
5.1.2 Energetics 234
5.1.3 Dipole Moments 235
5.1.4 Vibrational Spectra 235
5.1.5 Quadrupole Coupling Constants 239
5.1.6 Cyclic Chains 240
xiv Contents

5.2 HCCH Aggregates 240


5.2.1 Trimers 241
5.2.2 Tetramers and Pentramers 242
5.3 Hydrogen Halides 245
5.3.1 Open versus Cyclic Trimers 245
5.3.2 Three-Body Interaction Energies 246
5.3.3 Larger Oligomers 248
5.4 Water 252
5.4.1 Extended Open Chains 253
5.4.2 Branching Clusters 257
5.4.3 Cyclic Oligomers 257
5.4.4 Identification of True Minima 262
5.4.5 Substituent Effects 270
5.5 Mixed Systems 272
5.5.1 Geometries 273
5.5.2 Energetics 274
5.5.3 Vibrational Spectra 275
5.5.4 Effects of Electron Correlation 278
5.5.5 Other Mixed Trimers 280
5.6 Summary 282

6 WEAK INTERACTIONS, IONIC H-BONDS, AND ION PAIRS


6.1 Weak Acceptors 292
6.1.1 Dihalogens 292
6.1.2 CO 294
6.1.3 CO2 295
6.1.4 NNO 296
6.1.5 SO2 297
6.1.6 CC12 298
6.2 C—H as Proton Donor 298
6.2.1 Alkynes 299
6.2.2 Alkanes 302
6.2.3 Metal Atoms as Acceptors 306
6.2.4 Hydride as Proton Acceptor 307
6.3 Symmetric Ionic Hydrogen Bonds 308
6.3.1 Hydrogen Bihalides 308
6.3.2 Comparison with Other Anionic H-bonds 310
6.3.3 Cationic H-bonds 316
6.3.4 Comparisons between Cations and Anions 318
6.3.5 Alkyl Substituents 319
6.3.6 Other Considerations 320
6.4 Asymmetric Ionic Systems 321
6.4.1 General Principles 321
6.4.2 Test of Quantitative Relationships 322
6.5 Syn-Anti Competition in Carboxylate 326
6.5.1 Ab Initio Calculations 326
Contents xv

6.5.2 Experimental Findings 328


6.5.3 Carboxylic Group 328
6.5.4 Solvent Effects 328
6.5.5 Resolution of the Question 329
6.6 Neutral versus Ion Pairs 330
6.6.1 Amine-Hydrogen Halide 330
6.6.2 Carboxyl/Carboxylate Equilibrium 335
6.6.3 Experimental Confirmation 337
6.6.4 Long Chains 339
6.7 Summary 341
6.7.1 Low Polarity of Acceptor 341
6.7.2 C-H Donors 342
6.7.3 Ionic H-Bonds 343
6.7.4 Neutral Versus Ion Pairs 345

Index of Complexes 365


Subject Index 371
This page intentionally left blank
Abbreviations

ACPF Approximate Coupled-Pair Functional approach to compute electron correla-


tion
ANO Atomic Natural Orbitals. A basis set constructed from maximization of the
occupancy numbers of the natural orbitals of a given atom from a CI calcula-
tion.
APT Atomic Polar Tensor. An analytic means of considering the effect of atomic
motion upon the dipole moment of a given system.
BSSE Basis Set Superposition Error. The error incurred in a computation of the in-
teraction energy when the basis functions of one monomer artificially im-
prove the basis set of its partner (and vice versa), thereby lowering the en-
ergy by the variation principle.
CASSCF Complete Active Space Self-Consistent Field. An MCSCF calculation which
includes all excitations of a given type from a chosen set of reference molec-
ular orbitals.
CC Coupled Cluster. A means of including dynamic electron correlation that in-
cludes higher-order excitations.
CC Counterpoise Correction. A means of correcting BSSE.
cc Correlation Consistent, referring to a specific class of basis sets.
CCD Coupled Cluster including Double excitations.
CCSD Coupled Cluster including Single and Double excitations.
CEPA Coupled Electron Pair Approximation for including dynamic correlations.
CF Charge Flux. Loss or gain of electron density as an atom is displaced.
CPF Coupled Pair Functional procedure for including dynamic electron correla-
tion.
CI Configuration Interaction. A means of including electron correlation by mix-
ing in to the Hartree-Fock wave function, configurations generated by exci-
tations from occupied to virtual MOs.
xviii Abbreviations

CISD Configuration Interaction using Single and Double excitations.


CT Charge Transfer. The component of the interaction energy resulting from ex-
citation of the electrons of one subunit into the vacant MOs of its partner.
DISP Dispersion energy. An interaction resembling London forces, present only in
post-SCF calculations.
DZ Double-Zeta basis set. Similar to minimal basis set except that each orbital
consists of a pair: an inner and outer function.
DZP Double-Zeta Polarized basis set. Like DZ but also including polarization
functions.
ECP Effective Core Potential. Also known as pseudopotentials. A procedure for
considering only the valence electrons explicitly; used mainly with large
atoms.
EFG Electric Field Gradient. The rapidity with which the electric field generated
by a given molecular system is changing, usually evaluated at the position of
a nucleus.
ES Electrostatic energy. The Coulombic interaction between the static charge
clouds on two molecular entities.
EX Exchange energy. Part of the interaction energy between static charge clouds
of two subunits, resulting from Pauli exchange between them. Similar to
steric repulsion for molecular interactions.
GIAO Gauge-Including Atomic Orbitals. A class of orbitals which are designed to
permit computation of chemical shift tensors in NMR spectra.
GTO Gaussian-type orbital. Functions which differ from hydrogen-like orbitals in
that the r dependence is exp(— r2).
HF Hartree-Fock. Calculations based on the Hartree-Fock approximation of each
electron moving in the time-averaged field of the others. No dynamic elec-
tron correlation is included.
IEPA Independent Electron Pair Approximation. A means of including dynamic
electron correlation where the total correlation energy is partitioned into a
sum of contributions from each occupied pair of spin orbitals.
IGLO Individual Gauge for Localized Orbitals
KM Kitaura-Morokuma means of partitioning the total interaction energy of a
given complex.
LCAO Linear Combination of Atomic Oribtals. Usually refers to the practice of
constructing each molecular orbital in terms of functions centered on each
atom.
LCCM Linearized Coupled Cluster Technique
MBPT Many Body Perturbation Theory. A means of including electron correlation,
similar to MP.
MBS Minimal Basis Set. One orbital is used to represent each of the orbitals of
each shell that is full or partially filled. Examples: Is for H or He; 1s, 2s,
2px, 2py, and 2pz for Li-Ne.
MCSCF Multi-Configuration Self-Consistent Field. A means of variationally mini-
mizing the energy of several electron configurations of a given system simul-
taneously, so as to provide a better description of its electronic structure.
MINI A type of minimal basis set, the most common being MINI-1.
MO Molecular Orbital.
MPn nth-order M011er-Plesset theory. Means of including electron correlation.
Abbreviations xix

NBO Natural Bond Orbital. Orbitals resulting from a sort of localization scheme
that resembles the traditional concepts of 2-center bonds and lone electron
pairs.
NEDA Natural Energy Decomposition Analysis. A means of decomposing the total
energy using natural bond orbitals.
NPAD Normalized Proton Affinity Difference. A measure of the relative proton
affinities of the two partners in a H-bond, indicating the likelihood of observ-
ing an ion pair.
NQCC Nuclear Quadrupole Coupling Constant
PES Potential Energy Surface.
POL Polarization energy. The component of the interaction energy that results
when the electric field of one subunit perturbs the electron density on its
partner. Also abbreviated as PL.
QCISD Quadratic Configuration Interaction using Single and Double excitations.
SAPT Symmetry-Adapted Perturbation Theory. A means of partitioning the interac-
tion energy into various components.
SCEP Self-Consistent Electron Pair. A correlation technique that considers elec-
trons two at a time.
SCF Self-Consistent Field. Commonly used synonymously with Hartree-Fock
(HF). No dynamic correlation included.
STO Slater-Type Orbital. Functions which loosely resemble hydrogen-like or-
bitals, especially insofar as the dependence is exp(— r).
ZPVE Zero-Point Vibrational Energy. The vibrational energy contained by a mol-
ecule or complex at 0° K.
+ Indication that the basis set includes very diffuse functions.
// Double-slash indicates a distinction between the level of theory at which a
geometry was optimized and the level for which the energy was computed.
For example, MP2/6-31G(2d,2p)//SCF/6-31G* indicates a MP2/6-
31G(2d,2p) calculation of a structure optimized at the SCF/6-31G* level.
This page intentionally left blank
HYDROGEN BONDING
This page intentionally left blank
I

Quantum Chemical Framework

I. I Quantum Chemical Techniques

The following represents a capsule summary of the nature of ab initio quantum chemical
calculations, intended to provide the reader with the minimal background necessary to com-
prehend the theoretical literature of H-bonds and to evaluate the quality and reliability of a
given calculation. For more details about quantum chemistry in general or the specific meth-
ods, the reader is referred to any of a number of fine texts and review articles that have been
written on the subject1-8.
Quantum chemical methods are based on the time-independent Schrodinger equation

where 9€ represents the Hamiltonian of the system. The Hamiltonian is a quantum me-
chanical description, in terms of operators, of the kinetic energy of the particles, as well as
the interactions between all electrons and nuclei. is a wave function and represents the
"trajectories" of the particles. Due to the quantum nature of electrons, and the Heisenberg
Uncertainty Principle, classical trajectories are inappropriate, so the paths are described in-
stead in terms of probabilities of the particles being at any given point in space. Equation
(1.1) is an eigenvalue problem, with E representing the energy of the system.
Most quantum chemical calculations invoke the Born-Oppenheimer approximation
which distinguishes between the electrons and the much heavier nuclei. Consequently, it is
a good approximation to treat the nuclei as fixed in space, with the electrons moving in the
electric field generated by them. Equation (1.1) cannot be solved exactly because each elec-
tron repels all of the others, leading to a many-body problem. The usual method adopted to
circumvent this difficulty has been the Hartree-Fock approximation, which in essence re-
duces the problem to a single particle by time-averaging the motion of all electrons other
than the one in question. In other words, electron 1 is considered to move in the field of the

3
4 Hydrogen Bonding

electron cloud associated with the probability distribution of all other electrons. The same
idea is applied to electron 2 which moves in the time-averaged field of electron 1 plus all
the others, and so on. Of course, solution of the 1 -electron Hartree-Fock equation for each
electron changes its probability density, thereby altering the field it sets up for the other
electrons. Consequently, the equations are solved iteratively, until the 1-particle wave func-
tions, and the fields generated therefrom, no longer change appreciably from one cycle to
the next. Because of the generation of this "self-consistent field," the SCF abbreviation is
sometimes used synonymously with HF, that is, Hartree-Fock.
This Hartree-Fock approximation6,7 neglects a very important phenomenon. Since the
electrons are constantly aware of each others' presence, via their electrostatic repulsion,
they would tend to correlate their motions so as to avoid one another as much as possible.
One can take as an analogy a pair of quarreling roommates that wander through their apart-
ment from one room to the next in such a way as to avoid contact with each other. When
one is in the kitchen, the other will be in the bedroom, and so on. This phenomenon will
lower the energy of the system by reducing the time that the electrons of like charge spend
close to each other. Not only does correlation lower the energy of the system, but it also af-
fects the overall electron density of the system.

1.1.1 Basis Sets


Most quantum chemical treatments describe each molecular orbital as a linear combination
of atomic orbitals9,10. In this so-called LCAO approximation, each atom has assigned to it
certain functions that resemble the standard s, p, d, and so forth atomic orbitals that are cen-
tered about the nucleus. There are certain important differences, however. Whereas the hy-
drogen-like orbitals die off as exp(— r), where r is the distance from the nucleus and a
constant, the integrals that must be incorporated into the Hartree-Fock matrix using this
form of the orbital are difficult to evaluate. These "Slater-type orbitals" (STOs) are usually
replaced by a small number of Gaussian functions, where exp(— r) is replaced by
exp(— r2). The quadratic dependence of r in the exponent greatly simplifies the form of the
integrals, particularly those that involve several atomic centers simultaneously. So much
simpler, in fact, that it is computationally more efficient to evaluate a large number of inte-
grals involving Gaussians than a much smaller number of STO integrals. Moreover, a se-
ries of Gaussians with progressively larger values of orbital exponent a can fairly closely
reproduce a Slater-type function. As a result, most modem quantum chemical calculations
are performed using basis sets composed exclusively of Gaussian functions.
The collections of orbitals that are applied to calculations are referred to as "basis sets"
and typically fall into one of several categories. The smallest employs one orbital to repre-
sent each of the orbitals of each shell that is full or partially filled. A minimal basis set for
a first-row atom like Li or F would thus contain 1 s, 2s,2p x, 2p , and 2pz orbitals, whereas
H or He would be described by a single 1 s orbital, as indicated in the first row of Table 1.1.
Perhaps the most commonly used basis set of minimal type is STO-3G11. The name refers
to the fact that each Slater-type orbital of the minimal basis set is replaced by a triad of
Gaussian functions. (This triad is called a "contraction," and the three functions are referred
to as "primitives.") Other minimal basis sets of this type, albeit less widely used, are STO-
4G and STO-6G.
There are a number of ways that a minimal basis set can be improved. One approach is
to provide more flexibility by doubling the number of functions. A "double- " basis set is
similar to minimal, except that each atomic orbital is "split" into two. The flexibility of such
Quantum Chemical Framework 5

Table 1 . 1 Some of the most common types of basis sets, and the orbitals contained therein.

Common name Abbreviation Common example H 1st Row atoms

Unpolarized
minimal MBS STO-nG 1s ls,2s,2pa
split valence sv 3-21G,4-31G 1si, 1s0 ls,2si,2pi,2so,2po
double-zeta DZ ls
i, ls o Isi,lso,2si,2pi,2so,2po
triple valence 6-311G ls i ,ls m ,ls o ls,2si,2pi,2sm,2pm,2so,2po
Polarized
split valence, SVP 6-31G** ls i , ls o ,P 1s,2si,2pi,2so,2po,da
polarized
double-zeta, DZP lsi,lso,p lsi,lso,2si,2pi,2so,2po,d
polarized
double-zeta, DZ2P 1s
i,
1s
o,Pi,Po
1si,lso,2si,2pi,2so,2po,di,do
doubly polarized
containing diffuse + 6-31 + G** 1s i ,ls o ls,2si,2pi,2so,2po,d,sp
functions
a
Each p-set consists of 3 separate functions: p x ,p y ,p z : similarly, d refers to a set of 5 functions.
b
Diffuse set of s,px,py,pz, all with small exponent.

a "DZ" basis permits each orbital to expand or contract in size to conform to the environ-
ment in which the atom finds itself. Even greater flexibility is provided by a triple- or TZ
basis set. One line of reasoning has been that while splitting the valence shell is certainly
worthwhile, there is little to be gained by doing the same thing to the inner shell, since these
electrons will be little affected by changes in the bonding environment around the atom.
Basis sets have evolved that are similar in spirit to DZ or TZ but split only the valence shells.
Such "split-valence" basis sets are exemplified by 3-21G12, 6-31G13, or 6-311G14. The 6
in the latter case refers to the number of Gaussian primitives used to describe the inner shell,
1s orbital. The 3 and 1 of 6-31G indicate the number of primitives that model the inner and
outer valence orbitals: 3 Gaussians are used for the inner and 1 for the outer. 6-311G is sim-
ilar except that a third set of functions are added, by a single Gaussian, to split the valence
shell three ways, as indicated in Table 1.1.
The inner and outer s orbitals of an atom in a double- basis set are both spherical, that
is, isotropic. So while the presence of two of them permits the orbital to expand, it can do
so isotropically only, with no stretching in any one direction over another. Such "polariza-
tion" in a given direction would be useful in many situations. Consider for example the
O—H bond in water. The direction of the O—H bond is clearly unique; a stretching of the
H 1s orbital in this direction would permit a better description of the bond. Mixing the s or-
bital with even a small amount of a px orbital, collinear with the O—H bond axis, would
enable the former to stretch in the x direction, polarizing the orbital along the bond. This
stretching is illustrated graphically in Fig. l.la where the s-orbital is indicated by the cir-
cle and the px as a smaller orbital. For this reason, when added to hydrogen, p-orbitals are
referred to as "polarization functions". In an analogous manner, the p-orbitals of C or O, for
example, can be polarized by a small amount of a d-orbital of appropriate symmetry, as in-
dicated in Fig. 1.1b. Polarization functions on such atoms hence correspond to orbitals of
d symmetry. Rather than adding simply a px function, a full set of all three p-orbitals are
used to polarize the hydrogen basis set. A full set of five d-functions are likewise used, al-
6 Hydrogen Bonding

Figure I.I Schematic representation of the manner in


which (a) a p-orbital can polarize one of s-type and
how (b) a p-function can be polarized by a d-function.
The position of the nucleus is indicated by the dark
dot.

though in some cases it is convenient to use a set of six even though the sixth corresponds
roughly to an orbital of s-type symmetry.
There are various conventions used to indicate when polarization functions have been
added to a basis set. The addition of a "P" is commonly used, as for example when DZP
refers to a polarized double- basis set. Of course, this single letter does not clarify whether
all atoms have had polarization functions added, or only a subset. In most cases, the P des-
ignation should be taken to indicate polarization functions on all atoms. Another indication
that has been used over the years is an asterisk. A single asterisk, as in 6-31G* indicates po-
larization functions (d-type) have been added to non-hydrogen atoms; a second asterisk
would inform the reader of p-functions on hydrogen as well. Whereas the P designation car-
ries no information as to the values of the exponents used for the polarization functions, the
asterisks refer to specific values, in the context of well defined basis sets, for example,
6-31G**. It has become more common with faster computers to use multiple sets of polar-
ization functions. For example, one might wish to include both a "tight" and "diffuse" set
of d-functions, with large and small orbital exponents, respectively, on oxygen atoms. This
double polarization set can be indicated in various ways, for example, as DZ2P. Since the
asterisk convention cannot indicate this easily, it has become increasingly common to aban-
don these asterisks altogether and to indicate the numbers of polarization functions in paren-
theses. As an example, 6-31G** could be equally described as 6-31G(d,p) where the d and
p polarization functions on non-hydrogen and hydrogen atoms come respectively before
and after the comma. Doubling the d-functions, but leaving as a single set the p-functions
on hydrogen, would then be simply indicated as 6-31G(2d,p). This sort of notation readily
lends itself to the representation of orbitals of even higher angular momentum, as for ex-
ample 6-311 G(3df,2pd).
Another sort of function which has found a good deal of use in basis sets is of the same
symmetry as those mentioned above, for example, s or p. However, it is given a particularly
small orbital exponent, imparting to it a large and diffuse nature. Such diffuse orbitals are
particularly useful for describing anions, as they permit the overload of electrons to better
avoid one another as they take advantage of the large expanse of space over which this or-
bital extends. It has become common to indicate the presence of such functions by a + sym-
bol. For example, 6-31 +G* includes a diffuse sp-shell on non-hydrogen atoms; a second
+ as in 6-31 + +G* indicates diffuse functions on H as well15. Test calculations have sug-
gested that such functions on hydrogen are less important than on heavy atoms, at least for
ground states16.
While the vast majority of calculations make use of basis functions centered on the
atomic nuclei, it is sometimes worthwhile to add other functions which have their origin
along the bond axis between a given pair of nuclei. Such "bond functions" 1 7 - 2 1 can provide
Quantum Chemical Framework 7

rapid convergence of various properties with respect to number of functions in the basis set.
There are, on the other hand, certain hidden dangers that accompany the use of such func-
tions and of which the user must be wary22.
The choice of orbital exponents can be an important one. It is common for the exponents
of the s and orbitals to be established by optimization of the energy of the atom. One may
impose certain restrictions on the various exponents. For example, even-tempered basis sets
are those in which the progression of orbital exponents follows a geometric sequence from
one to the next, that is, .23 This prescription calls for only two parameters in an ar-
bitrarily long list of primitive Gaussian functions. A generalization which extends the opti-
mization to four parameters has been designated as "well-tempered"24,25. The polarization
function exponents (d, f, etc.) can also be optimized, or alternately chosen so as to yield
good reproduction of properties other than the total energy.
With the increasingly common use of electron correlation, there has been some rethink-
ing of these prescriptions to select orbital exponents. "Correlation-consistent" basis sets
have been designed26, sometimes abbreviated with the prefix "cc", specifically to be
amenable to calculations involving electron correlation. The smallest of these is cc-pVDZ
(correlation-consistent, polarized valence double zeta), and is capable of yielding well over
half of the total correlation energy in atoms such as Ne and B; cc-pVQZ can attain 95-99%.
These sets can be augmented (indicated by the prefix "aug") by additional functions opti-
mized for atomic anions, so as to describe diffuse electronic distributions27. Testing shows
these basis sets can be superior to more standard types in certain applications28. Also de-
signed to incorporate correlation effects into the construction of orbitals is the atomic nat-
ural orbital (ANO) basis29 which maximizes the occupation numbers from a CI calculation
of the individual atom, starting from an uncontracted basis.
As a general notation, it is common to enclose within square brackets the number of
functions of various types. For example, [6s5p2d/4s2p] indicates that first-row atoms are
represented by 6s functions, 5 sets of 3 (px, py, and pz) p-functions, and two separate sets
of five d-functions. The basis set for hydrogen is indicated after the slash: 4s and 2 sets of
p-functions. Because the order is always the same; s, followed by p, d, etc. it is not un-
common to leave out the letters entirely: [652/42]. In this lexicon, the 6-31G** basis set
could be represented as [321/21]. If the system were to contain atoms beyond the first row
of the periodic table, they would be indicated to the left of the first set of numbers, as in
[second-row/first-row/H].
As the atom becomes larger, the number of basis functions needed to describe it increases
as well. However, since one is most interested in the valence shell where most of the "ac-
tion" occurs, the increasingly larger number of "inactive" or core functions become more
and more of a nuisance. One cannot simply omit them as the valence orbitals would then
collapse into smaller core orbitals (which are of much lower energy). One solution is de-
velopment of "core pseudopotentials" or effective core potentials (ECP) which eliminate
the need to include core functions explicitly, yet keep the valence functions from optimiz-
ing themselves into core orbitals9,30-32. Such pseudopotentials are commonly used in ele-
ments of the lower rows of the periodic table, like Br or I.

1.1.2 Electron Correlation


There are a number of ways in which one may begin to correct the Hartree-Fock wave func-
tion so as to include electron correlation6,7,33,34. The simplest in concept is configuration
interaction (CI) which takes the Hartree-Fock solution as a starting point, or reference con-
8 Hydrogen Bonding

figuration35,36. Other configurations are generated by permitting the excitation of one elec-
tron from the subset of occupied molecular orbitals to the subset of unoccupied or "virtual"
MOs. The complete list of "singly excited configurations" is generated by considering all
such possible excitations, subject to the restriction of preservation of the spin state of the
ground state under study. The list is then extended to double excitations, again accounting
for all possible combinations of excitations of two electrons from the occupied to the vir-
tual MOs. A full-CI list is generated by progressing in this manner to include triple, quadru-
ple, and higher excitations, until all N electrons have been excited. The correlated wave
function is then expressed as a linear combination of the reference, Hartree-Fock configu-
ration, plus small amounts of all of the possible excitations. Variational treatment of this
trial wave function leads to the correlation energy by adjusting the relative amount that each
particular configuration contributes to the final correlated wave function.
Unfortunately, the number of configurations generated by all possible excitations of even
relatively small systems quickly becomes astronomical and beyond the reach of any com-
puters. For this reason, the list is commonly terminated at some point. One of the more com-
mon points of termination is after the inclusion of all single and double excitations. This
CISD treatment typically captures the bulk of the correlation energy.
One notable problem with termination of the full CI expansion has been referred to as
the size-consistency problem37. What this means is that the same treatment of a complex is
fundamentally different than that of the subunits of which it is composed. Consider for ex-
ample the CISD level. Such a treatment of the first subunit would permit double excitations
within it; similarly for the other subunit. A consistent theory should hence permit quadruple
excitations within the complex, as this would account for simultaneous double excitations
in each of the subunits. But CISD terminates the excitation list at doubles in the complex.
Truncated CI treatments are therefore fundamentally poorly disposed to handle molecular
interactions such as hydrogen bonds.
There have been a number of correction algorithms formulated over the years to help
improve this size consistency problem but they do not entirely resolve it38-40. Another
means of introducing size consistency is by a quadratic approximation, QCISD41. The ap-
proach achieves this size consistency by sacrificing its variational character. It can be con-
sidered as a simplified approximate form of CCSD (see below); the method may cease to
remain size consistent on going to higher levels of substitution42.
Certain other types of procedures are size consistent. Coupled pair theories43-45 suffer
from a different shortcoming: they are not variational. What this means is that it is possible
in principle to obtain an energy lower than the true energy of the system in question. In the
independent electron pair approximation (IEPA), the total correlation energy is partitioned
into a sum of contributions from each occupied pair of spin orbitals. A different correlation
wave function is constructed for each pair, letting their electrons be excited into the virtual
MOs of the reference configuration. The total correlation energy then corresponds to the
sum of all pair energies.
When the IEPA approach is extended to incorporate coupling between different pairs, it
becomes a coupled-pair theory. In terms of excitations from the original Hartree-Fock de-
terminant, the correlation energy depends directly upon the double excitations, but their
contributions involve quadruple excitations in an indirect way, and the latter are linked to
hextuple excitations, and so on. The coupled-cluster approximation affords a means of ex-
pressing these relationships in a closed set of equations46,47. Most applications of coupled-
cluster theory include only double excitations, and are designated CCD48,49. More general
versions of the theory that include also single and higher excitations are commonly abbre-
Quantum Chemical Framework 9

viated as CCSD50. Coupled-cluster is highly demanding of computer resources so various


approximations to it have been suggested. The linear coupled-cluster approximation (L-
CCA) sets certain products equal to zero, and is equivalent to doubly-excited many-body
perturbation theory, carried to infinite order (see below). If instead of ignoring all the prod-
uct terms set equal to zero in L-CCA, certain of them are retained, one arrives at the cou-
pled electron pair approximation (CEPA).
The type of correlated method that has enjoyed the most widespread application to H-
borided systems is many-body perturbation theory34, also commonly referred to as M011er-
Plesset (MP) perturbation theory51-53. This approach considers the true Hamiltonian as a
sum of its Hartree-Fock part plus an operator corresponding to electron correlation. In other
words, the unperturbed Hamiltonian consists of the interaction of the electrons with the nu-
clei, plus their kinetic energy, to which is added the Hartree-Fock potential: the interaction
of each electron with the "time-averaged" field generated by the others. The perturbation
thus becomes the difference between the correct interelectronic repulsion operator, with its
instantaneous correlation between electrons, and the latter Hartree-Fock potential. In this
formalism, the Hartree-Fock energy is equal to the sum of the zeroth and first-order per-
turbation energy corrections.
The first correction to the Hartree-Fock energy appears as the second-order perturbation
energy. This term can be represented as a sum over double excitations from the reference
configuration:

where i refers to the orbital energy of molecular orbital i. The sum extends over all a and
b which are occupied MOs and r and s which are vacant. The integral in the numerator
makes use of physics notation and indicates a combination of Coulomb and exchange inte-
grals over these particular MOs6. Similar, albeit increasingly more complex, equations can
be derived for higher orders of perturbation theory. The energy, after correction by Eq (1.2),
is commonly referred to as MP2. The MP3 level involves additional terms, but remains re-
stricted to double substitutions from the reference configuration. At the fourth order, there
are contributions from single, triple, and quadruple excitations, as well as doubles. Since
the expressions involving the triple excitations are the most difficult, the full MP4 is some-
times simplified by neglecting them, leading to what is denoted MP4SDQ.
Due to the computational efficiency of MP theory, correlation calculations have come
within reach of many computational chemists. For example, the time necessary to carry out
a full MP3 calculation is comparable to that of a single iteration of CID. Another strong ad-
vantage is that MP theory is size consistent, making it a good choice for molecular interac-
tions of various sorts. Moreover, as will be illustrated in greater detail later, a large number
of calculations over the years have indicated that MP2 provides results in excellent agree-
ment with the much more computationally demanding MP4. For these reasons, the litera-
ture of correlated calculations of hydrogen bonds is largely dominated by M011er-Plesset
theory.
In certain cases, a single determinant does not offer an adequate representation of the
electronic structure. In such cases, it is useful to perform a Multi-Configuration SCF cal-
culation (MCSCF) wherein a number of different configurations are chosen as important,
and their adjustable parameters (orbital coefficients, etc.) are variationally optimized54,55.
10 Hydrogen Bonding

This procedure suffers from a high degree of arbitrariness in the choice of just which con-
figurations are deemed important. The calculation can be made somewhat more objective
by including all excitations between a subset of occupied MOs and a subset of vacant or-
bitals. (These excitations are subject to certain restrictions as to multiplicity or order of ex-
citation.) The orbitals chosen for the excitations are referred to as the "active space", and
the method is dubbed Complete Active Space Self Consistent Field (CASSCF)56,57. Both
MCSCF and CASSCF provide a certain fraction of the correlation energy, relative to a
single configuration, Hartree-Fock, calculation.

1.1.3 Geometries
Before 1980, geometry optimizations were not very well automated. Locating the minimum
on the potential energy surface of even a fairly simple molecule could be a tedious chore.
One geometrical parameter was usually optimized at a time, independent of the others.
Since the value used for one parameter affects the optimized values of the others, the entire
set of geometrical parameters had to be cycled through a second and sometimes a third time,
before the geometry could be considered converged. Many of these optimizations were only
partial in the sense that it was common to make certain assumptions about the geometry.
For example, one might decide in advance that all C—H bonds of a methyl group would be
of the same length or the geometry of an ethyl group might not be optimized at all, with cer-
tain standard bondlengths or angles being assumed.
The development in the early 1980s of the means to evaluate derivatives of the energy
with respect to nuclear motion, and implementation of gradient algorithms, changed the face
of ab initio calculations58-60. It became possible to optimize all parameters simultaneously,
searching for the minimum on a multidimensional potential energy surface. From that point
in time, complete optimizations became the norm in the literature. Since the optimizations
make use of a Hessian matrix consisting of the second derivative of the energy with respect
to the various geometrical parameters, it became straightforward to determine if the opti-
mization path had proceeded to a true minimum on the potential energy surface, or to a
higher-order stationary point such as a transition state. (A true minimum is distinguished by
the Hessian having all positive eigenvalues.) It should be understood, however, that the op-
timization procedures take one to a minimum on the surface, not necessarily to the mini-
mum. In other words, there is no guarantee that the minimum one locates is the global min-
imum; it might equally well be a secondary minimum.
A shorthand has developed by which quantum chemists communicate the nature of their
calculation. The level of computation is indicated on the left of a slash, with the basis set to
the right of the slash. Thus, MP2/6-31G* would indicate a MP2 calculation with a 6-31G*
basis set. It is common to optimize the geometry at one level of theory, and then to apply
a higher level to compute the energy at that particular molecular structure. Such calcula-
tions are indicated by a double-slash. As an example, if one were to optimize the geometry
at the SCF level with a 6-31G* basis set, and then to compute the energy of this structure
at the MP2 level with the larger 6-31G(2d,2p) set, this might be indicated as: MP2/6-
31G(2d,2p)//SCF/6-31G*.
These geometry optimization procedures paid an additional dividend. With a true mini-
mum in hand, it becomes possible to compute the vibrational spectrum of a given system.
A straightforward formulation allows one to extract the normal vibrational modes of the
system, complete with the corresponding frequencies, but with the important proviso that
Quantum Chemical Framework II

these calculations be made from the standpoint of a minimum on the surface. Attempts to
compute vibrational spectral features at a nonstationary point are meaningless. Hence, the
precise determination of minima opened the door to important comparisons with experi-
mental spectral data, comparisons that were not possible prior to that time.
Whereas optimization of the geometry of a single molecule is relatively straightforward,
the same procedure can encounter difficulties for a molecular complex, particularly if
weakly bound. The first problem here is that the force constants for motions of one mol-
ecule relative to the other are quite a bit smaller than those for stretches or bends wholly
within one molecule. One tactic to circumvent this difference is to perform a geometry op-
timization using frozen subunits. That is, the internal geometry of each partner can be taken
as fixed, and the optimization carried out over the intermolecular parameters. The final re-
sult is thus not fully optimized but the earlier restraint can then be released and the geom-
etry now optimized over all parameters, intramolecular as well as intermolecular.
A second issue concerns the landscape of the potential energy surface. As will be de-
tailed in a number of examples, the surface of many H-bonded complexes is rather flat.
However, the surface is far from featureless, containing a number of different local minima,
connected by stationary points of various orders. Depending on the particular optimization
algorithm or step size chosen, it is possible for the optimization procedure to jump over a
local minimum, missing it entirely. One must also be careful that the minimum identified
is indeed the global minimum by scanning large regions of the surface for others that might
be deeper. And due to the flatness of the potential, it is not uncommon for what appears at
first sight to be a minimum to be seen on closer inspection to be a stationary point of higher
order.
In summary, then, ab initio methods represent a powerful means of extracting informa-
tion about chemical systems. When flexible polarized basis sets are used, in conjunction
with electron correlation, results of high accuracy are attainable. It is possible to locate min-
ima on the potential energy surface as well as transient entities, such as transition states.
One limitation of these methods is the size of the system that may be studied. The amount
of computational resources required rises quickly with the number of atoms or electrons.
Another caveat is that ab initio methods typically investigate a static situation: that is, the
energy, energy derivatives, and electronic structure of one given arrangement of nuclei are
calculated at a time. The ab initio methods can be supplemented by other theoretical ap-
proaches in order to simulate dynamic processes.

1.2 Definition of a Hydrogen Bond

Since its first suggestion many years ago61-65, the hydrogen bond has continued to fasci-
nate chemists. This interaction is intimately involved in the structure and properties of wa-
ter in its various phases, and of molecules in aqueous solution. In addition to the traditional
role of the H-bond as a structural element in large molecules such as proteins and nucleic
acids66-70, a series of such bonds appear to be vital to the functioning of a number of en-
zymes71-73. There are some indications that H-bonds play an even more important role in
biological electron transfer across long distances than much stronger covalent bonds74. The
principles of H-bonding have been taken as a means to design new materials capable of self-
assembly into well-ordered crystal structures75- 77, for molecular recognition of organic
molecules 78-80 , organic analogs of zeolites with supramolecular cavities and continuous
12 Hydrogen Bonding

channels81-84, for self-assembly of spherical, helical and cylindrical structures85-90. H-


bonds offer an avenue for stereocontrol of certain reactions91 and for understanding the
structure of monolayers92,93.
A Lewis structure of a H-bond violates the octet principle of striving toward two elec-
trons around each hydrogen. Much weaker than a conventional covalent bond, the H-bond
is stronger than the van der Waals forces that bind together nonpolar molecules. While
Coulombic attraction between polar molecules certainly contributes toward the interactive
force, the H-bond is nonetheless considered to be more than a simple electrostatic interac-
tion.
The classic picture of a H-bond begins with a pair of molecules, both in their ground
electronic state and both with a closed shell62,94. One molecule, AH, is designated the pro-
ton donor with the pertinent hydrogen covalently bound to an electronegative atom like O
or F. The acceptor molecule, B, contains an electronegative atom with at least one lone pair
of electrons. As the two molecules approach, the hydrogen atom forms a sort of "bridge"
between them. The lone pair of the acceptor atom is pulled toward the bridging proton to
form a weak bond, designated by the dotted line in the simple AH ... B notation. It is not nec-
essary for AH and B to be separate molecules; intramolecular H-bonds are also acceptable,
so long as it is possible for the proton-donating and accepting groups to attain the proper
positioning.
There are several consequences of this interaction that are commonly taken as criteria
for formation of a H-bond95-97.

1.2.1 Geometry
The attractive interaction generated by the formation of the H-bond draws the two groups
closer together than would be the case in the absence of such a bond. For this reason, the
distance separating the nonhydrogen atoms, R(A ... B), is typically shorter than the sum of
van der Waals radii of the A and B atoms. Indeed, it is typically assumed that there is a strong
correlation between the shortness of this interatomic separation and the strength of the H-
bond98. Concomitant with the formation of the bond is a certain amount of stretching of the
covalent A—H bond; the amount of this stretch, r, is usually closely correlated with the
strength, as well as the length R(A ... B) of the H-bond94,99-101.
There is also an expectation of certain directionality to the H-bond. The bridging hy-
drogen will tend toward the line connecting the A and B atoms. The same is presumed for
the lone pair of the Y atom. Taking FH ... NH 3 as a first and simple example, the bridging
proton lies directly along the F...N axis. The C3 symmetry axis of NH3, coincident with the
single lone pair of the N atom, also lies along this line. The situation becomes a little more
complicated when the acceptor has more than one lone pair. A simple description of the
electronic structure of the carbonyl oxygen of, say H2CO, places two sp2-hybridized lone
pairs at 120° angles from each other and from the C=O bond. Formation of a H-bond with
HF should therefore occupy one of these two lone pairs, situating both the H and F atoms
along a line approximately 120° from the C=O bond. Another source of ambiguity would
arise if there were two hydrogens on the donor molecule that were each capable of partici-
pating in a H-bond. In most cases, one would expect a standard linear H-bond to form, uti-
lizing one of these protons. Another possibility, and one to be discussed in greater detail be-
low, would have the two protons both participating, and both oriented toward the acceptor,
but neither of them lying directly along the A...B axis. More complicated situations can be
envisioned when the donor has two protons and the acceptor has more than one lone pair.
Quantum Chemical Framework I3

1.2.2 Energetics
The strength of the H-bond is typically measured in terms of the interaction energy between
the two molecules involved. Such an interaction energy is more difficult to define in the case
of an intramolecular bond. In such a case, one can consider the difference in energy between
the H-bonded geometry and another configuration in which the bridging hydrogen is not per-
mitted to come within H-bonding proximity of the proton acceptor. For example, a simple
180° rotation around another bond axis can swing the proton away from the acceptor.
In the gas phase, which most quantum calculations mimic, the H-bond energy is typi-
cally in the range of 2 to 15 kcal/mol. This interaction is weaker than most covalent bonds
by about one order of magnitude, but stronger than nondirectional "noncovalent" forces
which tend to be less than 2 kcal/mol. Nonetheless, this range is meant only as a general
rule of thumb rather than as a strict threshold. One should not consider the range as exclu-
sively the province of H-bonds, nor should interactions be discounted as H-bonds merely
for lying outside this range.
Another energetic aspect of H-bonds is not only the total interaction energy, but its ori-
gin as well. For example, certain means of partitioning the interaction energy attribute the
bulk of the stabilization to electrostatic forces between the charge distributions of the two
subunits. Within this context, an interaction of the proper magnitude, but with minimal
Coulombic contribution, might not be categorized as a H-bond.

1.2.3 Electronic Redistributions


Unlike the formation of covalent bonds which involve massive shifts of electron density,
the rearrangements that occur as a consequence of a H-bonding interaction are more
subtle. The electron distributions within each subunit remain largely intact as the H-bond
is formed. Nonetheless, there are indeed shifts of electron density that do occur. While rela-
tively small in magnitude, these shifts tend to be characteristic of H-bonds, and can be taken
as a sort of fingerprint for formation of such a bond.
For example, there is an overall shift of electron density from the proton acceptor mol-
ecule to the donor. (It is for this reason that the proton donor is sometimes referred to as the
electron acceptor, and vice versa.) This density is drawn not only from the lone pair partic-
ipating in the H-bond but from the entire molecule. Rather than residing on the bridging
proton, the density bypasses this center and becomes distributed throughout the donor mol-
ecule. Indeed, the total density associated with the central hydrogen undergoes a decrease
as the bond is formed. The above patterns are drawn from objective analysis of the spatial
distribution of electron density maps. In contrast, any attempt to quantify the amount of
charge transferred from one molecule to the other relies upon some arbitrary partitioning of
the region between them as belonging to a specific molecule, and is consequently subject
to some degree of arbitrariness. There is also a good deal of sensitivity to basis set in as-
signing charge to various atoms. Bearing in mind these caveats, typical results indicate that
some 0.01 to 0.03 electrons are transferred from the proton acceptor to the donor molecule
upon H-bond formation.

1.2.4 Spcctroscopic Observations


One of the more striking consequences of the formation of a H-bond appears in the vibra-
tional spectrum. The band which corresponds to the stretch of the A—H bond shifts to lower
14 Hydrogen Bonding

frequency, is intensified, and undergoes a concomitant broadening. In a quantitative sense,


it has been possible to establish a nice correlation between the amount of the red shift and
the strength of the H-bond94,102,103. It would also appear that the magnitude of the inter-
molecular H-bond stretching frequency is directly related to the strength of the H-bond.
NMR spectra also have diagnostic utility 104 . The electron density shifts which arise from
the H-bonding result in perturbations of the proton shielding tensor, deshielding the bridg-
ing hydrogen105-107. The isotropic shielding and anisotropies tend to correlate with the
length of the H-bond 108-1l0 as do the peak volumes in the solid state heteronuclear corre-
lation spectra111. Indeed, chemical shifts can help ascertain the secondary or tertiary struc-
ture of proteins, discriminating between -helices and -sheets112; this relationship has
been tentatively attributed to the different H-bond lengths in the two structures and the
aforementioned relationship113. It has been suggested that the single or double-well char-
acter of a proton-transfer potential is signaled by whether the chemical shift of a deuterium
is larger or smaller than that of protium 114,115 . The 13C anisotropic chemical shift tensors
of the carboxyl carbon are sensitive to H-bonding, and have been seen to correlate with vi-
brational frequencies116. Proton magnetic resonance measurements have been used to probe
the pK of catalytic residues and proton positions upon binding an inhibitor to an enzyme117.
NMR measurements of the solid state can lead to surprisingly detailed characterization of
molecular geometry, including the nature of the proton transfer potential101. 15N chemical
shifts have been interpreted to question a view of the catalytic mechanism of serine pro-
teases derived from X-ray diffraction data118 and to study model systems pertinent to charge
relay chains119,120.
The 1 JNC, nuclear spin-spin coupling constant appears to have diagnostic value for the H-
bonding of amide groups in that it increases when H-bonding occurs through the O and is
lowered when this interaction involves the NH group121,122. In conjunction with the proton
chemical shift, this parameter can also be used to estimate the relative strengths of H-bonds
within a macromolecule such as a protein, or to distinguish between various secondary struc-
ture motifs123. NMR measurements can be used also to monitor the dynamics of H-bonds.
An example is a 2H and 13C study of cooperative rearrangements among the four OH groups
that form a ring in a crystalline sample124. Quadrupolar coupling constants can be of use as
well. For example, a study of polypeptides by solid-state 17O NMR spectroscopy125,126 re-
vealed a correlation between this parameter and the length of hydrogen bonds.
It is also possible to use electron resonance techniques to determine characteristics of a
H-bond in proteins with good accuracy, as in a recent ENDOR examination of the heme
binding pocket of fluorometmyoglobin127. Another promising technique is electron spec-
troscopy for chemical analysis (ESCA) which can monitor hydrogen bonds near the surface
of a material128. The difference in binding energies of atoms in the proton donor and ac-
ceptor offers a measure of the strength of the H-bond.
While there have been few experimental measurements of the dynamics of association
involved in formation of a H-bond, or the opposite dissociation, in the gas phase, there are
modern approaches that show promise in this regard. For example, resonant photoacoustic
spectroscopy has been used to examine possible dissociation pathways of carboxylic acid
dimers129-131.

1.3 Quantum Chemical Characterization of Hydrogen Bonds

Theoretical methods have been used to obtain insights into the nature of the hydrogen bond
from their very inception. While this volume focuses on work published since 1980, the in-
Quantum Chemical Framework 15

terested reader might gain a historical perspective from inspection of a number of review
articles which summarized earlier understanding of the H-bond phenomenon132-141.
In this section we discuss the manner in which quantum chemical calculations evaluate
the various properties that are important to H-bonds.

1.3.1 H-bond Geometries


The equilibrium geometry of any given molecular entity represents the bottom of the global
minimum in the potential energy surface. Quantum chemical methods can efficiently eval-
uate this set of nuclear coordinates. It must be understood however, that the equilibrium
structure at one level of theory will typically differ from that at another level. There is a sub-
stantial literature that deals, for example, with the effect of basis set upon molecular geom-
etry.
The various facets of the equilibrium geometry can be conveniently divided into inter-
molecular properties, for example the distance separating the two molecules, and in-
tramolecular parameters, representing the bond lengths and angles within each subunit.
Since the overall interaction energy in a H-bond is typically less than 15 kcal/mol, it is not
surprising that the energy of the system is usually not very sensitive to the former inter-
molecular geometric parameters. Due to the flatness of the energy profile, the minimum in
this potential can be shifted a good deal by minor changes such as a change in basis set. It
is therefore common to find intermolecular aspects of H-bonds to be sensitive to the par-
ticular theoretical approach. Bending and stretching potentials for intramolecular geomet-
rical parameters, on the other hand, are much sharper so there is less sensitivity to level of
theory. Of greatest interest with regard to these intramolecular geometries are the perturba-
tions induced in each subunit by the formation of the H-bond.
Whereas the quantum chemical calculations provide a straightforward picture of the
geometry at the bottom of the minimum, experimental observations pertain instead to a dy-
namic average. A diatomic molecule furnishes the simplest example of this difference where
Re refers to the interatomic distance at the bottom of the well and the vibrational average
of the ground level is denoted by Ro. Due to the weak nature of H-bonds, equilibrium and
dynamically averaged quantities can differ by significant amounts. A recent study142, for
example, predicts that the third-order anharmonicity within the water dimer might alter the
interoxygen distance by as much as 0.13 A.

1.3.2 Thermodynamic Quantities


There are several different energetic quantities that are relevant to quantum chemical eval-
uation of H-bond strength. The interaction of a pair of molecules, A and B, with one another
to form an A...B complex can be represented by the reaction

where the connecting dot notation is used to indicate the specific interaction within the com-
plex. The total energy change of the process in Equation (1.3) is commonly taken as E and
is defined as:

For most complexation reactions, the complex is more stable than the isolated species so
the process is exothermic and E is negative. All species are normally taken in their fully
optimized geometries. It is important to note that the process of combining with one another
16 Hydrogen Bonding

typically causes certain changes in the internal geometries of both A and B. For this reason,
the complexation energy E has folded into it the energetic consequences of such internal
geometry changes, sometimes referred to as "nuclear relaxation energy." The latter term is
conceptually distinct from electronic redistribution energetics that accompany the combi-
nation of A and B.
A typical calculation of a molecular interaction thus involves the geometry optimization
of three entities: the reactants A and B, and the complex A ... B. The subtraction of electronic
energies in Equation (1.4) yields Eelec, the electronic contribution to the interaction en-
ergy. This term includes the internuclear repulsive energy within the molecule. Other con-
tributions arise from translational, rotational, and vibrational motions of the nuclei143. Mak-
ing the usual assumption of ideal gas behavior, the translational partition function of a given
molecule in a container of volume V at temperature T is144

where m is the mass of the molecule, h is Planck's constant, and k is the Boltzmann con-
stant. Hence, each of the three species involved in Reaction (1.3) contains 3/2 RT of trans-
lational energy per mole so

The rotational partition function is dependent upon the equilibrium geometry. Assum-
ing separation of rotational and vibrational motions,

holds for temperatures significantly above absolute zero. Ia, Ib, and Ic represent the three
principal moments of inertia of the molecule or complex, determined by the nuclear masses
and their positions in space. The symmetry number a refers to the number of indistin-
guishable orientations one can obtain by rotating the molecule in space; is larger for more
symmetric molecules. Equation (1.7) is simplified somewhat for a linear molecule

since there is only a single moment of inertia, I. The rotational energy is different for the
nonlinear and linear cases:

A given molecule has 3n-6 normal modes of vibration (3n-5 if linear), where n is the
number of atoms. Each mode i has a characteristic vibrational frequency vi and a residual
energy even at absolute zero temperature. The total zero-point vibrational energy is thus:

As the temperature rises above 0° K. higher vibrational levels begin to be populated and
the additional vibrational energy is:
Quantum Chemical Framework 17

The full thermodynamic interaction energy E is equal to the sum of terms:

where all E terms on the right refer to the differences between the product A...B and re-
actants, A plus B.
The reader is cautioned that " E" is very commonly used in the computational literature
when what is actually meant is Eelec. That is, many articles refer to the electronic contri-
bution to the interaction energy as though it were the full thermodynamic E. In this book,
we will attempt to clearly differentiate between E and Eelec.
Perhaps a more proper designation for the electronic portion of the interaction would be
De, the dissociation energy from the equilibrium geometry. Do would refer to this same
quantity, after correction for zero-point vibrational energies. Another proviso concerns the
signs of these quantities. For a stable complex, E is negative, signifying its formation to
be exothermic, while De is taken as positive since it refers to the energy required to disso-
ciate the complex. In the literature, H-bond energies are usually discussed as positive quan-
tities. Other terms for this quantity are: complexation, dissociation, and interaction energy.
The enthalpy of formation of A...B from its constituent molecules is equal to E with a
pV correction that yields

since two molecules are combining to produce a single complex in Reaction (1.3).
The statistical expressions for entropy may be used to derive equations most useful in
studying molecular interactions of the type in Reaction (1.3). As there is typically a large
energy separation between the ground and excited electronic states of adducts and their
complex, it is valid to take the electronic entropy as zero. The translational entropy can be
written as:

where m refers to the mass of the particular system and P is the pressure. Substituting in the
values of the physical constants, the molar entropy can be written as

where the mass is expressed in amu, temperature in °K, and pressure in atmospheres. Ap-
plying Equation (1.15) to reaction (1.3) at 1 atrn pressure yields a molar entropy change of

The rotational entropy of each species is


18 Hydrogen Bonding

where qrot is defined in Equation (1.7). The change in molar rotational entropy can be de-
rived from Equation ( 1 . 17) to depend upon the principal moments of inertia of reactants and
products, along with the temperature. The entropy associated with each vibrational normal
mode is a function of both temperature and vibrational frequency vi.

It is only necessary to sum the contributions made by each mode to the entropy of the var-
ious reactants and products to obtain the vibrational contribution to the total entropy change
of the reaction. The latter total is then

The Gibbs free energy of Reaction (1.3), G, then follows simply from

1.3.3 Electronic Redistributions


There are a number of ways of monitoring the distribution of electron density in any mo-
lecular entity. The total density can be computed at a number of points in space and pre-
sented as a contour map or some three-dimensional representation. Shifts are easily exam-
ined by density difference maps which plot the difference in density between two different
configurations. For example, the density shifts caused by H-bond formation can be taken
as the difference between the complex on one hand, and the sum of the densities of the two
noninteracting subunits on the other, with the two species placed in identical positions in
either case. Comparisons with x-ray diffraction data have verified the validity of this ap-
proach145. Also, the total density of the complex itself can be examined for the presence of
critical points that indicate H-bonding interactions146-148.
It is also feasible to examine individual molecular orbitals, again via plots over inter-
esting regions of space. The data may be compressed into tabular form by assigning den-
sity to various atomic centers. This sort of treatment conforms to notions of atomic charges.
While any sort of partitioning of the density of this sort suffers from arbitrariness, it can of-
fer useful insights as long as the treatment is consistent from one configuration to the next.
In other words, while the atomic charges themselves may not have much meaning, the
changes undergone during the H-bond formation have more validity.
Rather than examine the electronic distributions directly, another approach focuses upon
the electrostatic potential in the region surrounding the molecule, which is a direct result of
the charge arrangement149. Correlations can be drawn between H-bonding abilities of mol-
ecular entities and the potential at certain selected points in space150,151.

1.3.4 Spectroscopic Observations


It is entirely feasible to compute the force field for nuclear displacements from equilibrium
by quantum chemical means, leading directly to evaluation of vibrational spectra. In fact,
the frequencies of the normal modes are required in order to evaluate the zero-point vibra-
tional energies mentioned earlier so as to compute the enthalpy of formation of a H-bond.
These theoretical spectra can be compared to available segments of experimental spectra152.
Quantum Chemical Framework 19

It must be remembered however that the bulk of experimental data are gathered in solvent
or other condensed media whereas the theoretical results pertain to the gas phase.
It is possible to evaluate elements of NMR spectra by theoretical means as well. Gauge-
Including Atomic Orbital (GIAO) procedures have been developed153 and applied to a va-
riety of H-bonded systems154-158. An alternate technique has been dubbed Individual
Gauge for Localized Orbitals (IGLO)159,160. While these permit investigation at the SCF
level, correlated work is possible as well, as in the GIAO-MP2 method 161-163 , a coupled-
cluster GIAO-CCSD(T) procedure164, or by a multiconfiguration generalization165. Ab ini-
tio calculations have reached a stage of maturity where quadrupolar coupling constants can
be computed with surprising accuracy166. For example, comparison of experimental mea-
surements with calculated coupling parameters have enabled an elucidation of the manner
in which aggregates of formamide change their basic structure as the temperature of the liq-
uid is changed167.

1.4 A Simple Example

The former analysis may perhaps be best understood by way of an example. Let us consider
first the water dimer in its most stable orientation. The geometries were optimized with a
modest basis set at the SCF level, but the results are illustrative nonetheless, and more ac-
curate calculations would not influence the example in any important respect. Figure 1.2a
illustrates the geometry of the monomer and the directions of the three principal moments
of inertia; the dimer is depicted in Fig. 1.2b. Table 1.2 reports the total mass of each species
in the first row, followed by the translational energy at 25° C, 3/2 RT. Since there are two
reactants and only one product, and all with the same translational energy, Etrans for the
reaction is —3/2 RT as listed in the second row of Table 1.3. The moments of inertia are
strongly affected by dimerization. The greater distances of the atoms from the center of mass
of the dimer, as compared to the smaller monomer, account for the much increased values
of Ib and Ic. Nonetheless, the total rotational energies of the monomer and dimer are iden-
tical, according to Eq (1.9). Hence, Erot amounts to —3/2 RT as was true for Etrans.
The water monomer has three normal modes of vibration. The first three rows of the vi-
brational frequency entries in Table 1.2 report the frequency of the monomer, followed by
the corresponding frequencies of the two molecules as they occur within the dimer. One
may note the perturbations caused by the interaction to the internal modes of each water
molecule as they differ by up to 100 cm - 1 . The next three lines list the "new" intramolec-
ular vibrational modes that are present in the dimer but not in the monomer. These typically

Figure 1.2 Structures of the water (a) monomer and (b) dimer, illustrating orientations of
the three principal moments of inertia. The origin is at the center of mass in each case.
20 Hydrogen Bonding

Table 1.2 Properties of water monomer and dimer at 298° K and


1 atm pressure, calculated with 4-31G basis set.

Monomer Dimer

M, amu 18 36
Etrans, kcal/mol 0.89 0.89
Ia,au 1.842 7.492
Ib, au 4.429 269.28
Ic, au 6.271 270.73
a 2 1
Erot, kcal/mol 0.89 0.89
v i ,cm -1 1743 1751,1789
3958 3870,3970
4109 4073,4116
130, 167
174,210
416,761
EZPVE, kcal/mol 14.02 30.63
Evib,them , kcal/mol
vib, therm' 0.001 1.78
Strans, calmor-1 deg-1 34.61 36.68
S rot,cal mol-1 deg-1 10.34 20.93
Svib, cal mol-1 deg-1 0.004 11.02

correspond to intermolecular stretching, wagging, and so forth. Perhaps more important,


they are also of much lower frequency than the intramolecular modes. For this reason, they
add only a relatively small amount to the zero-point vibrational energy. That is, the value
of 30.6 kcal/mol is only slightly larger than twice the EZPVE of the water monomer. Con-
sequently, EZPVE is only 2.6 kcal/mol for the dimerization reaction, as seen in the fourth
row of Table 1.3. On the other hand, it is the low-frequency vibrations which can more eas-
ily be populated as the temperature begins to climb. This population of the low-frequency

Table 1.3 Thermodynamic parameters of water


dimerization. E, H amd G in kcal m o l - 1 ;
S in cal mol-1 deg - 1 .

Eelec -8.23

Etrans -0.89
Erot -0.89
EZPVE +2.59
E vib,therm +1.76
E -5.66
AH -6.25
Strans -32.54
Srot 0.26
S vib 11.01
S -21.28
G 0.09
Quantum Chemical Framework 21

modes of the dimer is manifested by the larger value of E vib,therm in the next line of the table
and leads to the value of 1.76 kcal/mol for Evib in Table 1.3.
The exchange of six librational modes in the pair of isolated monomers for the same
number of low-frequency vibrational modes in the dimer is characteristic of the H-bonding
process. The system loses 3/2 RT of translational energy by combining two molecules into
a single complex, and the same amount of rotational energy as a pair of rotating species
form the dimer. This loss is compensated by the gain in zero-point energy from the new vi-
brational modes in the complex. In the water dimer in Table 1.2 for example, the six new
intermolecular vibrations account for 2.65 kcal/mol of zero-point energy, as compared to
the 3RT lost in librations (1.78 kcal/mol at 298° K). In addition, the low-frequency modes
are rapidly populated as the temperature climbs, adding more energy to the complex and
making the value of AE less negative. Taking the water dimer as an example again, the pres-
ence of these low-frequency modes adds some 1.8 kcal/mol to the complex in Evib,therm.
The various contributions to the entropy of the monomer and dimer are listed in the last
three rows of Table 1.2. The translational entropies may be seen in Eq (1.15) to vary as the
logarithm of the mass so the monomer and dimer values are similar. The loss of transla-
tional degrees of freedom upon dimerization is hence responsible for the negative value of
Strans in Table 1.3. The much larger moments of inertia of the dimer lead to a higher rota-
tional entropy, by Eqs (1.7) and (1.17). This effect is due to the closer spacing of the rota-
tional energy levels and their greater accessibility to thermal population. Although three de-
grees of rotational freedom are lost upon dimerization, a compensation occurs due to the
greater rotational entropy of the dimer arising from its higher moments of inertia; Srot is
therefore close to zero in this case.
The high frequencies of the vibrational modes in the water monomer make occupation
of any levels other than the ground state quite small. The dominance of the single com-
plexion, all molecules in their ground state, is associated with the near-zero vibrational en-
tropy. In contrast, just as the presence of low-frequency vibrations in the dimer permits
Evib,therm to climb with temperature, the population of these levels also provides for more
ways of rearranging the quanta of vibrational energy and hence to a much larger vibrational
entropy. The net result of these three contributing factors is that the full AS is negative by
some 21 eu. The overwhelming factor in this loss of entropy arises from the translational
component.
Comparison of the SCF energies of the water dimer with the pair of monomers yields an
electronic contribution to the H-bond energy of — 8.23 kcal/mol, as indicated in the first row
of Table 1.3. (Note that negative values indicate greater stability of the dimer and hence
more binding energy.) This value is lessened by 2.59 kcal/mol as a result of the greater zero-
point vibrational energy in the dimer than the monomers. Translation and rotation each add
0.9 kcal/mol to the binding energy, but the extra energy resulting from the population of the
intermolecular vibrational modes drops the interaction energy by 1.76 kcal/mol. The cal-
culated value of E is thus —5.66 kcal/mol. The enthalpy of binding is more negative by
RT. When combined with the negative value of AS, the Gibbs free energy change accom-
panying dimerization is close to zero. That is, the small dimerization energy is approxi-
mately canceled by the loss of entropy which accompanies the complexation.
There has been some discussion in the literature as to whether a H-bond is stronger than
a D-bond. That is, how does the isotopic substitution of a protium nucleus by a twice-as-
rnassive deuterium affect the energetics of binding. Since the electronic part of the interac-
tion energy is based upon the Born-Oppenheimcr approximation which places the nuclei at
rest, Eelec is unaffected by any isotopic substitution, including this one. Indeed, the po-
22 Hydrogen Bonding

tential energy surface upon which the nuclei move is independent of the atomic masses.
The translational and rotational terms would be affected to a small extent. The largest
change occurs in the vibrational terms. Doubling the mass of an atom would significantly
perturb the effective mass of any vibration involving that atom and so would change the
associated frequency. The effects would first be seen in the zero-point vibrational energies,
EZPVE, and then in Evib,therm as the temperature climbs. This effect is described in de-
tail below.

1.5 Sources of Error

There are a number of sources of error in the various contributors to the thermodynamics
of the formation of H-bonds. First are the assumptions of ideal gas behavior which permit
one to write the simple expressions in Eqs (1.5)—(1.17). These assumptions also include the
full separability of vibrational and rotational motion. The vibrational terms are commonly
calculated within the framework of the harmonic approximation which precludes coupling
between various modes and third or higher order terms in the dependence of the energy of
the molecule upon the nuclear deformations. The translational and rotational energies are
probably computed rather accurately. The total mass of any system is simple and stands
apart from any quantum calculations. While the rotational partition functions are sensitive
to the moments of inertia, the rotational energy at room temperature is virtually indepen-
dent of these properties. Moreover, equilibrium geometries can be calculated to relatively
good precision with even moderate levels of theory, certainly accurate enough to obtain ex-
cellent approximations to the correct moments of inertia.
More sensitive to the level of theory is the vibrational component of the interaction en-
ergy. In the first place, the harmonic frequencies typically require rather high levels of the-
ory for accurate evaluation. It has become part of conventional wisdom, for example, that
these frequencies are routinely overestimated by 10% or so at the Hartree-Fock level, even
with excellent basis sets. A second consideration arises from the weak nature of the H-bond-
ing interaction itself. Whereas the harmonic approximation may be quite reasonable for the
individual monomers, the high-amplitude intermolecular modes are subject to significant
anharmonic effects. On the other hand, some of the errors made in the computation of vi-
brational frequencies in the separate monomers are likely to be canceled by errors of like
magnitude in the complex. Errors of up to 1 kcal/mol might be expected in the combination
of zero-point vibrational and thermal population energies under normal circumstances. The
most effective means to reduce this error would be a more detailed analysis of the vibra-
tion-rotational motion of the complex that includes anharmonicity.
By far the largest source of error in calculating the energetics of hydrogen bonding arises
in the electronic term. As exemplified by Eq (1.4), each contribution is computed as the
difference in energy between the complex on one hand and the sum of monomers on the
other. One can note in Table 1.2 that E tends to be of similar magnitude to its two con-
stituent terms for translational, rotational, and vibrational energies. Such is not the case,
however, for electronic energies. These quantities represent the energy released upon form-
ing a given molecule from an assortment of isolated nuclei and individual electrons and are
hence very large in magnitude. Nearly the same energy is released whether these compo-
nents are assembled into a pair of isolated monomers or into a H-bonded complex; the dif-
ference between these two options (representing Eelec) is many times smaller than the en-
ergy of assembly in either case. Taking the water dimer as an example again, the total
Quantum Chemical Framework 23

electronic energy of this dimer is on the order of —10 5 kcal/mol. To be more precise, as-
sembly into the pair of water monomers releases 95,266.86 kcal/mol and into the complex
95,275.08 kcal/mol. The difference, 8.23 kcal/mol, represents only 0.009% of the total. As
a consequence, even very small errors in either of the large quantities, errors as small as
several thousandths of a percent, will produce very large errors in their difference, Eelec.
It is understood that there are few means of calculation of the electronic energy of any
system that are capable of obtaining the true result within 0.01%. Indeed, many ab initio
calculations, especially those that ignore correlation, will be in error by many, many times
this amount. Calculation of even remotely reasonable values for interaction energies are
thus dependent upon large-scale cancellation between very large errors. For example, if one
calculates the energies at the Hartree-Fock level, it is implicitly assumed that the correla-
tion energy in the dimer, typically several hundred kcal/mol, will be nearly identical to the
total correlation energy of the pair of isolated monomers. Similar types of cancellation are
the presumption for calculating interaction energies with less than complete basis sets.

1.6 Basis Set Superposition

With regard to basis set, there is another and more subtle hazard to the computation of in-
teraction energy by the supermolecular approach. It is obvious that one must use the same
basis set in calculating the energy of the pair of isolated monomers as for the complex. For
example, the properties of the monomers and dimer of water were computed using the
4-31G basis throughout in the foregoing analysis (Tables 1.2 and 1.3). Each contributor to
the interaction energy was obtained by

where the subscript refers to the specific basis set. If one were to apply a smaller basis set
like 3-21G or STO-3G to the individual monomers, but retain 4-31G for the complex, it
would introduce an obvious inconsistency, wiping out any possibility of the aforementioned
cancellation necessary for reasonable results. So it is agreed that the same basis set must be
used for the complex as for its constituent subunits.
But the situation is not as simple as it sounds, as noted by Kestner in 1968168. The ba-
sis set for each monomer consists of functions centered on each of its atoms. The basis set
of the dimer is larger in the sense that there are present functions centered on all atoms of
both monomers. One may represent this fact by using a subscript to indicate the atoms cov-
ered by the basis set:

The larger basis set of the dimer provides additional flexibility. The electrons of monomer
A are free to partially occupy the orbitals provided along with molecule B, and vice versa.
Such freedom is not provided to these same electrons when the monomer is calculated in
its own basis set, with B completely absent. The availability of the extra orbitals will lower
the energy of each monomer, within the context of the dimer, by the variation principle
which states that each additional degree of flexibility provided to the electrons permits a
lowering of the energy. As a result, the complex undergoes an artificial stabilization due
solely to its larger basis set, in comparison to the smaller sets of the monomers. This spuri-
ous stabilization of the complex, in excess of any genuine interaction energy, is commonly
referred to as basis set superposition error (BSSE).
24 Hydrogen Bonding

How may this error be avoided? The simplest means is to calculate the energy of all
species within the same set of basis functions. Since any computation of A...B must surely
place functions on both A and B, the same must be true for each monomer.

The difference between this formulation and Eq (1.22) is that the basis set used to calculate
the energy of A includes not only its own functions but those of B as well (and similarly for
B in the same large basis set). The calculation of E(A)A...B is much like that of the full com-
plex except that the nuclei and electrons of B are deleted. The extra functions of B included
in the calculation of A are sometimes referred to as "ghost" orbitals and the procedure out-
lined in Eq (1.23) is commonly denoted "functional counterpoise" after the originators of
the suggestion169. The m and d superscripts on E in Eqs (1.22) and (1.23), respectively,
refer to whether the energies of the monomers are computed within the monomer or dimer-
centered basis set. The difference between these two means of calculating the interaction
energy is one measure of the superposition error.

1.6.1 Secondary Superposition


While the computation of the energy of each monomer must be performed within the con-
text of the basis set of the entire complex for the sake of consistency in order to avoid BSSE,
there is another consideration engendered by this approach. This issue was first raised by a
number of research groups in the late 1970s and early 1980s170~174 and is related to the
change in properties of each monomer associated with the ghost orbitals of its partner. Con-
sider as a simple example a spherically symmetric atom like Ar. An atom-centered basis set
would correctly reflect that Ar has no dipole (or higher) moment. Suppose now that an ad-
ditional species is added to the system, another Ar atom for example. Within the context of
the basis set of the pair, the spherical symmetry of the first Ar atom is lost; consequently
each atom has associated with it a nonzero permanent dipole moment. The interaction en-
ergy hence contains a dipole-dipole interaction that is not present in the real dimer. Similar
arguments can be extended to higher multipole moments or to elements of the polarizabil-
ity tensor. These properties are different in the original basis set of a single atom as com-
pared to that of the dimer.
This line of reasoning is valid also for more complicated systems like H-bonded com-
plexes of molecules like water. Even though each HOH molecule does indeed possess a
nonzero moment, addition of the basis functions of its partner would introduce a change in
this, or higher, moments. The above changes in the calculated properties of each monomer,
caused by the addition of the partner functions to the basis set, together with the perturba-
tions in the interaction energy associated with them, are frequently referred to as secondary
BSSE or as basis set extension effects. Not only are these effects more difficult to remove
than primary BSSE, but it is not entirely clear whether they should in fact be corrected. For
example, early on, Karlstrom and Sadlej170 argued that these effects can be beneficial in
that the properties of each monomer are improved by the enlargement of the basis set, as
would occur if the additional functions were centered on the molecule itself rather than its
partner. However, later work indicated that secondary BSSE typically represents another
artifact that deteriorates the quality of the calculation. Latajka and Scheiner17S took as a
model the interaction between a Li ' cation and a neutral molecule of water and showed that
Quantum Chemical Framework 25

the secondary BSSE can be quite large, comparable in magnitude to the primary effect. They
suggested a crude means of correcting the extension effect and found no ensuing overcor-
rection.

1.6.2 Important Properties of Superposition Error


There are several salient facts to bear in mind:
1. In general, BSSE is reduced as the basis set becomes larger and more flexible. How-
ever, there is no strict correspondence, and the BSSE can in fact become larger with
certain additions to the basis set. Minimal basis sets are particularly prone to large
BSSE, as are those like 3-21G with a poor description of the inner shells.
2. The BSSE rises rapidly as the two molecules approach one another; angular de-
pendence is more complex.
3. The origin of BSSE makes it difficult to incorporate counterpoise corrections di-
rectly into gradients of the potential energy.
4. Whereas SCF BSSE can be reduced to negligible proportions with large basis sets,
the superposition error at correlated levels goes down much more slowly, persist-
ing at large values, even with very flexible bases.

1.6.3 Historical Perspective


Understanding of the issues involved in superposition errors has evolved slowly. Conse-
quently, the reader is liable to encounter a number of different attitudes and means of han-
dling the problem over the years in the literature. This section is intended to provide some
perspective on the problem so that the reader will be able to critically assess the impact that
superposition error might have on a given set of calculations in the literature.

1.6.3.1 Early Attitudes


Although many researchers ackowledged that there was some inconsistency in using a
larger basis set for the complex than for the monomers, there was initial reluctance to ac-
cept the counterpoise procedure as a valid means to correct the problem when it was first
introduced. The chief source for this skepticism lay in the numerical results. Many of Ihe
early calculations of H-bonded complexes relied on fairly small and inflexible basis sets. It
is now known that bases of this type tend toward potentials that are much less attractive
than the true potential. Limitations of the era also prohibited application of correlation in
most cases, eliminating a major attractive component. As a result, the H-bond attractions
corresponding to these treatments are much too weak. It was only the superposition errors
that were hidden in the calculations that permitted the final results to be even slightly at-
tractive. In other words, if one does not analyze the results carefully, one can easily be mis-
led since the spurious attractive nature of the BSSE can compensate in some sense for the
unsatisfactory character of the calculations themselves.
For example, an STO-3G calculation of the water dimer, if left uncorrected for basis set
superposition, can yield an interaction energy not very different from experimental expec-
tations. But when the superposition error is removed, the remaining potential, the true
HF/STO-3G potential is not attractive at all, but indicates the water molecules would repel
one another and not form a H-bonded complex176. Rather than recognize this observation
as a legitimate failing as an error due to the STO-3G basis set or to the absence of correla-
26 Hydrogen Bonding

tion, it was tempting to attribute the repulsive potential to the counterpoise correction. And
if one takes the experimental data point as the ultimate goal of a calculation, then retaining
the superposition error appeared to offer a superior means of achieving that end.
But such an approach is shortsighted. One must understand that precise reproduction of
an experimental result is coincidental at best when using a crude method with known weak-
nesses. After all, a minimal basis set without correlation does not offer a realistic version of
a molecular system, so why should one take an experimental H-bond energy as a criterion
as to whether counterpoise is an appropriate correction? By removing the BSSE, one ap-
proaches the true picture of a given theoretical model, in this case the HF/STO-3G version
of the water dimer. Even with larger and more flexible basis sets, one should not necessar-
ily expect the calculations to duplicate experiment. Such an expectation might lead one to
erroneously conclude that counterpoise corrections do not improve the accuracy of the cal-
culations177.
Nonetheless, the early disagreement of counterpoise corrected H-bond potentials with
experiment spawned a number of variants of the technique which reduced the BSSE cor-
rection and left the potential more attractive than if the full error were removed. Some of
these methods justified themselves on the grounds that the electrons of one molecule should
not expand into the orbital space of the partner molecule that is already occupied by elec-
trons 178-181 . Hence, damping factors were introduced or more formal means of permitting
the electrons of molecule A to partially occupy only the vacant MOs of molecule B, and
vice versa176,178,182. Another technique proposed employing a perturbing charge field gen-
erated by the partner183,184. An alternate treatment has been proposed to evaluate the BSSE
which is based on an exact perturbation solution185. More recently, it has been demonstrated
on formal and numerical grounds that the Pauli exchange principle itself prevents the elec-
trons of A from expanding into the occupied space of B 186-191 . Consequently, it is the full
counterpoise correction, as originally proposed, that should be applied to the problem of
molecular interactions.
It is now widely accepted that the counterpoise correction should be applied to the
Hartree-Fock part of the potential182,190,193-195. (Indeed, recent work has suggested that
Slater basis functions might provide a realistic alternative to the more standard Gaussians
in certain cases, provided counterpoise corrections are made195.) Computer technology has
made this acceptance an easy pill to swallow since the BSSE can be made negligibly small
by application of large and flexible basis sets which can be handled by modern worksta-
tions. It is hence not even necessary in many cases to do the actual correction. The small-
ness of this error is particularly fortunate when optimizing geometries since the gradient
procedures that search potential energy surfaces for stationary points do not incorporate
BSSE corrections directly into their algorithms. Nevertheless, even with large basis sets,
superposition errors can introduce noticeable errors into equilibrium geometries of many
molecules196,197.
There are alternatives to the most commonly used Boys-Bernardi counterpoise scheme.
One approach that shows promise, for example, is a chemical Hamiltonian approach
(CHA), pioneered by Mayer 198-201 , which attempts to isolate the superposition error di-
rectly in the Hamiltonian operator. The Schrodinger equation that is solved is hence a mod-
ified one, which yields a wave function that is hopefully free of superposition error. In the
case of (HF)2, it was found that this approach mimics rather closely the results of the stan-
dard counterpoise scheme for a scries of small to moderate sized basis sets200. Later calcu-
lations202 extended these tests to other small H-bonded systems as well, again limiting their
testing to basis sets no larger than 6-31G**. A recent test203 has extended the method's use-
Quantum Chemical Framework 27

fulness. The authors find for a series of H-bonded complexes that the difference between
the Boys-Bernardi and CHA-corrected interaction energies diminishes as the core, that is,
sp-part, of the basis set improves. They recommend the use of either correction scheme as
superior to simply trying to eliminate the superposition error by improving the basis set.
However, this approach has been criticized on the grounds of its inconsistency with the
results of symmetry-adapted perturbation theory204 and because this Hamiltonian is non-
Hermitian185. A similar idea has been proposed to construct a projection operator to remove
the BSSE when applied to a wave function, with results comparable to the standard coun-
terpoise procedure when tested on He and H2205. It might be noted, however, that this ap-
proach also has its detractors who identify inconsistencies with symmetry-adapted pertur-
bation theory204.

1.6.3.2 Correlated Levels


The problem is less tractable at correlated levels where the BSSE persists at uncomfortably
large magnitudes even with very large basis sets206. Most workers now agree that the full
counterpoise error must be removed from correlated calculations of molecular interaction
potentials190,197,204,207-210. For one thing, the counterpoise-corrected interaction energy
equates nicely to the sum of perturbation theory terms, each of which is formally free of su-
perposition error211. Nonetheless, there remains some lingering controversy as to the ap-
propriateness of invoking this correction181,185,192,212.
As an example, calculations examined the fluctuations that occur in the interaction en-
ergy of the water dimer as small perturbations are introduced into the basis set213. These
perturbations included addition of a second set of d-functions on O, minor adjustments in
the polarization function exponent, or the number of primitive gaussians in the contraction.
The raw interaction energies for the water dimer with R(OO) = 3.0 A are illustrated by the
broken curves in Fig. 1.3 as the basis set was altered. Note the large fluctuations at the MP2
as well as SCF levels. For example, the MP2 interaction energies vary between —5.5 and
—8.5 kcal/mol. After counterpoise correction, on the other hand, the data is much more con-
sistent from one basis set to the next, as illustrated by the solid curves.
More recent work has confirmed these conclusions: The counterpoise method leads to
an accurate description of the correlated interaction energies in the HF dimer, with the pro-
viso that the basis set is capable of properly describing the physical forces involved209. Even
with an insufficiently flexible set, the counterpoise-corrected results are more stable with
respect to basis set than uncorrected data.
Davidson and Chakravorty214 have recently compiled an exhaustive listing of earlier
means of treating superposition error and proposed an alternative means of looking at the
problem, referred to as a complete basis set. They suggest that counterpoise corrections
should perhaps be supplemented by what they refer to as monomer and dimer nonadditivity
corrections in order to obtain the correct interaction energy at any level. These new terms
would contain within them secondary BSSE. In the water and HF dimer cases they consid-
ered, they found that the latter nonadditivity corrections are of the same sign as counterpoise
corrections at the SCF level so that they move the interaction energy in the correct direction.
On the other hand, the dimer nonadditivity correction is of opposite sign to counterpoise at
the MP2 level, which might explain why inclusion of counterpoise corrections at this level
can move the calculated interaction energy away from an experimental result.
A reexamination of this analysis led others to suggest that large nonadditivity correc-
tions result from a poor choice of basis set and do not indicate any conceptual weakness in
28 Hydrogen Bonding

Figure 1.3 Interaction energies computed for the water


dirtier213 with R(OO) equal to 3.0 A. Uncorrected values are
connected by broken lines and counterpoise corrected (cc) in-
teraction energies illustrated by solid curves.

the counterpoise procedure itself211. The authors present the notion that basis sets that yield
small counterpoise corrections are not necessarily best adapted to investigate molecular in-
teractions. A basis set designed to best reproduce the important components of the interac-
tion energy would be a better choice. While sets of the latter type may lead to a nonnegli-
gible BSSE, counterpoise correction would yield superior results. The original authors215
argue that the difference between the true dissociation energy and its counterpoise-corrected
equivalent is a nonadditive correction for basis set incompleteness.
In terms of analyzing the contemporary literature, it is probably advisable to consider as
overly attractive any interaction energies that have not been corrected for basis set super-
position error. It is emphasized that "overly attractive" refers in this context to the correct
result with a given theoretical model, not to the experimental value as a reference. Results
computed at the SCF level with a counterpoise correction can probably be taken as the most
accuracy one is likely to achieve with a particular basis set without correlation. There re-
mains some difference of opinion concerning correlated results, but counterpoise correc-
tions should probably be taken as more valid than those with no such corrections at all.
There is also some lingering question as to the precise details of properly correcting BSSE
when there are more than two identifiable units involved in the interaction 201,216-218 .

1.7 Energy Decomposition

One of the underlying questions about hydrogen bonds is just what physical forces hold the
two partner molecules together. That is, what is the fundamental nature of the H-bond? As
Quantum Chemical Framework 29

the two molecules approach one another, a number of physical phenomena can be imagined
that might be involved in the forces between them 219-221 .
From a strictly electrostatic point of view, the two molecules that eventually form a H-
bond each have associated with them an electronic distribution which produces an electric
field in the surrounding space. The interaction of the static fields of the two molecules cor-
responds to a purely Coulombic force which may be attractive or repulsive, depending upon
the orientations of the two molecules. The energetic consequence of this interaction is com-
monly denoted as the electrostatic energy. This electrostatic interaction corresponds, then,
to the classical Coulomb force between two charge distributions at long separations between
the two subunits. At such distances, it is also useful to write the full electrostatic interaction
energy as a multipole series. That is, one may construct the charge distribution of each neu-
tral molecule as the sum of a dipole moment vector, quadrupole moment tensor, and so on
to higher and higher orders. The interaction between the dipoles of the two molecules be-
haves as R-3 where R is the distance between centers of charge of the two. Dipole-quadru-
pole interactions die off as R - 4 , quadrupole-quadrupole as R - 5 , and so on, up to infinite
orders. At long distances, the higher order terms are anticipated to become quite small as
this series converges. Indeed, for very long separations, it is the dipole-dipole interaction
which dictates the preferred angular orientation of the approach of one molecule toward the
other. Should one of the molecules be charged, the ion-dipole term is of lowest order and
the series will die off as R - 2 .
The situation becomes less clearcut as the two molecules approach within H-bonding dis-
tance of one another. In the first place, the multipole series loses its utility since the higher
order terms become progressively larger and the series does not converge properly. It may
hence be misleading to consider the dipole-dipole term as dominant or indeed, as even an
important contributor. Even ignoring the multipole analysis, the full electrostatic interaction
becomes more difficult to define unambiguously. That is, when far apart, it is a simple mat-
ter to assign any "piece" of electron density to one molecule or the other, based simply on
whether the region of interest is close to A or close to B. But as the two molecules approach
one another, this dividing border becomes more vague. There is significant density in the re-
gion midway between the two molecules and it is not at all clear whether the electrons here
are part of the distribution pattern of A or of B. The energetic consequence of this density
overlap is commonly lumped under the rubric of "penetration" terms in the electrostatic en-
ergy, defined properly as the difference between the full electrostatic interaction energy and
the infinite summation of the multipole expansion. It should also be recalled at this point that
the electrostatic interaction assumes the charge density patterns of either molecule are com-
pletely unaffected by the presence of the field generated by its partner. That is, the density is
"frozen" in the configuration adopted when the two molecules are fully isolated.
The contributions of the various terms in the multipole expression to the full electrosta-
tic interaction energy may be illustrated for the water dimer in Fig. 1.4222. Each curve in
Fig. 1.4 represents the cumulative sum up to the indicated term. For example, the R-5 curve
represents the sum of the R - 3 , R - 4 , and R-5 terms. For intermolecular separations ex-
ceeding 5 A, the first term in the series, R - 3 , which contains primarily the dipole-dipole in-
teraction, nicely mimics the full coulombic energy. As the two molecules approach closer
together, higher-order terms become necessary. At the van der Waals minimum of 3 A, even
taking the series up through sixth or seventh order significantly underestimates the full term.
It is hence apparent that penetration effects arc important for this uncharged H-bond in its
equilibrium configuration.
A second factor in the interaction between the two molecules is not classical in origin
but arises instead from the requirement that the full wave function of any system, includ-
30 Hydrogen Bonding

Figure 1.4 Distance dependence of the multipole series of the


electrostatic interaction energy, truncated at various orders, for
the water dimer. Data from222. The values for the series trun-
cated at R-7 differ only very slighty from the R-5 series and
so are not shown explicitly.

ing a H-bonded complex, be antisymmetric with respect to interchange of any two elec-
trons. The interchange between pairs of electrons, both of which are contained within one
molecule, has already been taken into account in generating the wave function of that mol-
ecule. What is of interest here is the interchange of one electron from molecule A with one
from B. Because of its mathematical description, the energetic consequence of permitting
the latter exchanges into the supermolecular wave function is termed "exchange" energy.
In the case of interactions between closed shell subunits, as is the situation in H-bonded
complexes, this term is repulsive in nature (hence the descriptive as "exchange repulsion")
and can be linked in some sense with the classic picture of "steric repulsion" between charge
clouds. It must nevertheless be remembered that this force is quantum mechanical in origin
and, like electrostatic energy, is calculated within the context of electron densities of each
subunit that are unaltered by the presence of the partner. The sum of the aforementioned
electrostatic and exchange energies is sometimes classified as the Heitler-London energy.
Of course, an integral ingredient in H-bonding arises from the ability of each molecule
to perturb the electron density of the other. For example, as A approaches B, its electric field
induces redistributions of electron density within B and vice versa. Such redistributions are
driven by the energetics of the situation and are of course stabilizing. The net stabilization
achieved by the system as a result of the density redistributions can be classified as "in-
duction" or "deformation" energy, owing to its origin.
There have been attempts in the literature to further partition this deformation energy
into smaller pieces. One means of breaking down this term rests on the ability to divide elec-
Quantum Chemical Framework 3I

trons and space into that belonging to A and that of B. The electronic redistribution of A can
now be accomplished in one of two ways. The A electrons can move around from their orig-
inal locations around A to other parts of space, previously unoccupied, but still defined as
"A space." This term is commonly referred to as "polarization" energy as it fits with the
usual definition of polarization of a given molecule. The other possibility is for the A elec-
trons to invade "B space" which leads to the concept of "charge transfer" from A to B and
its energetic consequence. However, the distinction between polarization and charge trans-
fer rests on the division of space into that belonging to A or B, which is arbitrary at best,
and becomes even more so as the two subunits merge into a single complex. For that rea-
son, there are many that argue against any separation of the induction energy into these two
components as they claim it leads to spurious judgments as to the fundamental nature of the
interaction under study223,224.
The previous components of the interaction energy can be derived in the independent
particle approximation and so appear within the context of Hartree-Fock level calculations.
Nevertheless, inclusion of instantaneous correlation will affect these properties. Taking the
electrostatic interaction as an example, the magnitude of this term, when computed at the
SCF level, will of course be dependent on the SCF electron distributions. The correlated
density will be different in certain respects, accounting for a different correlated electro-
static energy. The difference between the latter two quantities can be denoted by the corre-
lation correction to the electrostatic energy.
There is another physical phenomenon which appears at the correlated level which is
completely absent in Hartree-Fock calculations. The transient fluctuations in electron den-
sity of one molecule which cause a momentary polarization of the other are typically re-
ferred to as London forces. Such forces can be associated with the excitation of one or more
electrons in molecule A from occupied to vacant molecular orbitals (polarization of A), cou-
pled with a like excitation of electrons in B within the B MOs. Such multiple excitations
appear in correlated calculations; their energetic consequence is typically labeled as "dis-
persion" energy. Dispersion first appears in double excitations where one electron is excited
within A and one within B, but higher order excitations are also possible. As a result, all the
dispersion is not encompassed by correlated calculations which terminate with double ex-
citations, but there are higher-order pieces of dispersion present at all levels of excitation.
Although dispersion is not necessarily a dominating contributor to H-bonds, this force must
be considered to achieve quantitative accuracy. Moreover, dispersion can be particularly
important to geometries that are of competitive stability to H-bonds, for example in the case
of stacked versus H-bonded DNA base pairs225.
Having listed the above components of the interaction energy, it is worthwhile to un-
derscore their arbitrariness. The total energy of a system corresponds to a real physical ob-
servable, so the interaction energy, defined as the difference between energies, is likewise
real. But the various components do not correspond to a quantum mechanical operator and
are only as real as the arbitrary definition associated with them. As an example, the elec-
trostatic energy arising within the Hartree-Fock formalism is similarly limited by the inde-
pendent particle approximation. It is necessary to apply correlation corrections to high or-
der to even approach the physical picture of this phenomenon. It is also questionable
whether one should consider "exchange" as a separate entity since its existence is intimately
connected with the quantum mechanical mandate of antisymmetry of the wave function.
The notion of first precluding and then "permitting" electron exchange between subunits,
so as to extract exchange energy, is no more real than is the ability to turn off and on this
antisymmetry principle.
32 Hydrogen Bonding

Despite the arbitrariness of definition, the decompositions of the interaction energies of


H-bonds have provided some intriguing and useful insights into the fundamental aspects of
this phenomenon. Some of these will be detailed in the ensuing chapters. Nonetheless, the
reader is cautioned that this arbitrariness has also spawned a number of different schemes
of decomposing the energy into various segments in the literature. One must be particularly
careful in comparing these sets of data since it is not uncommon for two different schemes
to use the same name for components that are derived in different ways. Most schemes fall
into one of two general categories. The various components can be computed directly via
a perturbational scheme219,221,226 or a supermolecule approach can be taken wherein the
terms are calculated as a difference between large quantities220.

1.7.1 Kitaura-Morokuma Scheme


Let us take as an example what is probably the most frequently used means of decompos-
ing the interaction energy of various complexes, including H-bonds. In the KM scheme, Ki-
taura and Morokuma220,227 first compute the wave functions of the two isolated subunits
of the complex, A° and B°. They take this pair of wave functions as the starting point for
a Hartree-Fock calculation of the complex.

Note that the electrons of A are antisymmetrized within A°, and similarly for B electrons
in B°. However, no exchange of A and B electrons is permitted in A° B°. The zeroth it-
eration of the SCF procedure yields an energy which differs from the total energy of the pair
of isolated subunits by an amount taken to be EES since it permits the field of each monomer
to interact with the electron density, i°, of the partner, without perturbing that density.
The exchange energy is extracted by again beginning the SCF procedure but this time
permitting interchange of electrons between A and B, indicated by the AB operator below.

The energy associated with 2 differs from that of 1 by an amount defined here as the ex-
change energy, EEX, owing to the electron exchange within the 2 wave function.
Starting with 1r and enforcing the restriction of no exchange between electrons in A
and B, convergence of the iterative SCF procedure permits the electrons in each subunit to
relax in the presence of the field of the partner. The extra stabilization gained as a result of
this relaxation is associated with the polarization, and is labeled EPL. A similar SCF relax-
ation, but now allowing the full antisymmelrization of all electrons, yields an energy which
is lower than the latter one by an amount which is taken to correspond to charge transfer
between A and B, and is hence denoted ECT. The latter four terms do not encompass all of
the effects associated with complexation. Any remaining effects, in addition to those that
result from the artificial separation described above, are collected into a last "mixing" term,
E MIX . Moreover, there have been attempts to further partition some of the above terms into
smaller pieces as well as other mixing terms. Note that the Kitaura-Morokuma scheme lim-
its itself to SCF calculations so includes no dispersion, nor any correlation corrections to
the aforementioned terms.
One of the problems of this technique lies in the mixing term which provides no phys-
ical insights. Furthermore, this term grows to uncomfortably large proportions when the in-
teraction strengthens223. It was also found that the various components of the interaction
energy are even more sensitive to basis set choice than is the total interaction energy. The
Quantum Chemical Framework 33

separation of deformation energy into charge transfer and polarization is purely artificial;
indeed, the latter two effects are indistinguishable in the limit of a complete basis set.
Some of these points can be illustrated with simple examples. Table 1.4 lists the com-
ponents of the interaction energy of the water dinner computed by the Morokuma-Kitaura
scheme for the water dimer, taken in its experimental geometry, with an interoxygen sepa-
ration of 2.98 A227. The electrostatic term is clearly highly sensitive to the basis set chosen.
Going from minimal STO-3G to split-valence 4-31G more than doubles this attractive
force; adding d-functions reduces EES by 1.4 kcal/mol. The exchange energy is more sta-
ble but the other terms are again rather erratic. The charge transfer energy is notable in that
it is quite large for the minimal basis set. Much of this contribution can be attributed to ba-
sis set superposition error which is expected to contaminate this term the most. (Efforts have
been made more recently to correct the various components for BSSE194,228-232.)
If one of the two species is charged, the ion can be expected to be much more effective
at polarizing its partner. Furthermore, the presence of ion-dipole, ion-quadrupole, and so
forth terms will enhance the electrostatic interaction. These presumptions are confirmed by
the comparison of the data in Table 1.5 for the ionic complex between (NH4)+ and NH3224,
with the neutral dimer in Table 1.4. The data were computed using a very large basis set,
including d and functions on nitrogen. A range of different intermolecular separations is
considered to illustrate the distance-dependence of the various components. Note from the
bottom lines of the two tables that the ionic (H 3 NH ... NH 3 ) + complex is bound more
strongly by several-fold. Whereas the electrostatic component in the neutral dimer is less
than 10 kcal/mol, this term exceeds 20 kcal/mol at the same separation (3.0 A) and is greater
than 30 kcal/mol at the equilibrium separation of the ionic complex (2.75 A). One can also
see the rapid growth of the exchange repulsion as the two subunits approach one another,
as it is the principal factor preventing their collapse into one another. While the polariza-
tion energy in the neutral water dimer is less than 1 kcal/mol, the presence of the ion in
(H 3 NH ... NH 3 ) + greatly enlarges this component, even at distances as far as 3.25 A. The
charge transfer term is similarly enhanced in the ionic complex.
But the data in Table 1.5 also point out one of the prime weaknesses of decomposition
methods like Kitaura-Morokuma. Note that as the two subunits come closer than their equi-
librium separation, the polarization energy blows up beyond all reasonable proportions.
This very large negative value is due to the "merging together" of the basis sets of the two
molecules, particularly for a basis as extended as the one being used here. The problem
arises because the polarization energy, as defined in such a scheme, fails to observe the Pauli
exclusion principle and electrons from one subunit begin to occupy space already taken by

Table 1.4 Morokuma-Kitaura components of SCF interaction


energy of water dimer.a
STO-3G 4-31G 6-31G*

EES -4.2 -8.9 -7.5


EEX 4.0 4.2 4.3
EPL -0.1 -0.5 -0.5
ECT -4.8 -2.1 -1.8
EMIX
0.1 -0.3 -0.1
AE -5.1 -7.7 -5.6
a
All values in kcal/mol 227
34 Hydrogen Bonding

Table 1.5 Morokuma-Kitaura components of SCF interaction


energy of H 3 NH +... NH 3 . a
R(A)

3.25 3.0 2.75 2.50

EES -18.2 -23.5 -31.6 -44.8


EEX 5.1 11.5 25.8 57.3
EPL -8.2 -14.6 -28.5 -180.9
ECT -3.4 -6.0 -12.3 -30.1
EMIX 5.2 10.8 24.8 183.7
E -19.5 -21.7 -21.9 -16.2
a
All values in kcal/mol. Data224 calculated using large basis set (S + f).

the electrons of the partner. Another symptom of the problem is the very large magnitude
of the mixing term, whose positive values seem to compensate for the overly attractive po-
larization (and charge transfer energies).
There have been numerous schemes proposed in the literature to circumvent some of
these difficulties. One worth mention sets up a "Pauli blockade" which prevents any of the
intermediate wave functions from violating the Pauli principle233. It effectively combines
into a single term the KM polarization, charge transfer, and mix terms. There is also the
problem that basis set superposition error contaminates each of the components, making a
physical interpretation difficult. Various means have been devised over the years to cir-
cumvent this problem193,194,234.

1.7.2 Alternate Schemes


A reduced variational space method235, related to the KM procedure, has been developed
in which the orbitals of one fragment are optimized in the field of the frozen orbitals of its
partner. Truncation of the variational space by deletion of unoccupied orbitals of one part-
ner or the other is the pathway to evaluation of polarization, charge-transfer, and BSSE
terms. When applied to the water dimer235, the Coulomb and exchange sum dominates the
interaction but charge transfer and polarization terms are needed for proper angular depen-
dence.
A quite different scheme has been proposed by Weinhold et al.236-238 which is based on
natural bond orbitals. After an initial transformation of the atomic orbital basis set into nat-
ural atomic orbitals which optimize the occupancy, one obtains a set of core plus valence
orbitals with high occupancy, and another set of residuals with low occupancy. The natural
bond orbitals are derived from the formation of an optimal orthonormal set of directed hy-
brids which translate to a set of localized orbitals that correspond roughly to the traditional
concepts of chemical bonds and lone pairs. In this framework, the total energy of the dimer
is partitioned into two principal components. E arises when all unoccupied bond orbitals
are deleted and is associated with the electrostatic interactions, plus dominant effects of the
Pauli principle (that is, steric repulsions). The other component is denoted E and is rep-
resentative of charge-transfer delocalization. While E * is many times smaller than E
the contributions of these two terms to the interaction energy, indicated by the A, are much
more comparable in magnitude:
Quantum Chemical Framework 35

E can be related to the Heitler-London energy, the sum of electrostatic and exchange in-
teractions, while the SCF deformation energy, containing both intramolecular polarization
and intermolecular charge transfer, corresponds roughly to E
Analysis of their wave functions for the water dimer revealed that the bulk of the charge
transfer consisted of density shifting from the lone pair of the proton acceptor to the anti-
bond between the oxygen and bridging hydrogen of the donor molecule. This work em-
phasized the importance of delocalization of electron density into the unoccupied orbitals
in stabilizing the water dimer. The authors estimate that 5.4 kcal/mol arises from the spe-
cific donation from the aforementioned lone pair to O— H antibond. Their version of the
Heitler-London energy (electrostatic plus exchange) is repulsive, in contrast to most other
decomposition treatments wherein the attractive Coulombic force is stronger than the ex-
change repulsion. One of the more intriguing findings of the Reed-Weinhold treatment is
that relatively large energetic stabilizations result from only very small amounts of charge
being transferred, generally less than 0.01 electrons.
Reed et al.236 compared their results for a number of systems with the KM scheme which
attributes much of the H-bond attraction to electrostatic energy. The authors attribute the
distinction to the operational definition of charge transfer. They claim that the KM electro-
static term contains contributions from determinants which are better described as donor-
acceptor in nature, leaving to the charge transfer energy only that portion of the acceptor
orbital which is Schmidt orthogonalized to the partner's donor orbital. On the other hand,
Olszewski et al. echo the KM contention of the dominance of electrostatics in their studies
of H-bonding of various pairs of simple hydrides233. One may conclude that the distinction
between the two treatments is largely a semantic one which underscores the arbitrary na-
ture of any means of partitioning the interaction energy.
A more modern version of this general scheme based on natural bond orbitals involves
a decomposition of the SCF part of the interaction energy into electrostatic, charge trans-
fer, and deformation terms239. While these terms are similar in name to the KM compo-
nents, there are significant differences in formulation. The electrostatic term, for example,
contains an exchange contribution as it enforces antisymmetry of the appropriate wave
function. It, moreover, is evaluated from fragment wave functions deformed by the inter-
action, so contains induction/polarization energy. The deformation energy is repulsive as it
comprises the distortions of the monomer electron clouds so as to maintain orthogonality
of their wave functions. Their analysis of the water dimer led Glendening and Streit-
wieser239 to attribute strong contributions to the binding from both electrostatics and charge
transfer, with sizable repulsive forces arising from the deformation. The results are exhib-
ited in Table 1.6 where they may be compared with the more commonly used Morokuma-
Kitaura procedures. Figure 1 .5 illustrates the dependence of each of these components upon
the particulars of the basis set. Their behavior can be compared to that of the total interac-
tion energy, indicated by the solid line. Most of these terms are reasonably basis set insen-
sitive, that is no more sensitive than the total E itself.
Still another form of decomposition is based on first transforming the canonical molecu-
lar orbitals into localized MOs of a different sort. The so-called localized charge distribu-
tions240 also contain contributions, not necessarily integral, from the various nuclei. The to-
tal energy of a H-bonded complex like the water dimer is partitioned into kinetic and potential
energy terms. The drive for a given charge distribution to spread out so as to lower its kinetic
energy is balanced against the "suction" of the charge into a region of low potential energy.
Table 1.6 Comparison between Morokuma-Kitaura
and natural energy decomposition analysis
(NEDA)239. All values in kcal/mol, calculated with
4-31G basis set.

KM NEDA

total E -7.8 -7.8


CT -2.4 -13.3
ES - 10.5 -17.8
POL -0.6
EX 6.2
MIX -0.5
DEF 24.8
BSSE -1.1

Figure 1.5 Values of natural energy decomposition analysis com-


ponents of water dimer for various basis sets, from239. Basis sets
are as follows: (1) STO-3G, (2) 4-31G, (3)6-31G*, (4) 6-31G**,
(5)6-31+G**, (6) 6-31 ++G**, (7)6-31++ G(2d,p), (8) 6-
31 + +G(2d,2p), (9) cc-pVDZ, (10) aug-cc-pVDZ, (11) cc-
pVTZ, (12) aug-cc~pVTZ.
Quantum Chemical Framework 37

Another, and generally older, philosophy partitions the total energy into contributions
that are associated with one or two atomic centers241,242. A strength of this approach is the
ability to focus upon bond strengths via the latter term. This technique has been applied to
analysis of chemical phenomena such as the Cope rearrangement243, internal rotation244,
and the nature of aromaticity245,246.

1.7.3 Perturbation Schemes


Still another means of partitioning the interaction energy is based on standard Rayleigh-
Schrodinger perturbation theory. It takes as its starting point a wave function which is the
simple product of those of the isolated molecules and uses the interaction between the two
molecules as the perturbation222,247-250. A difficulty inherent in this approach is enforcing
the proper symmetry into the wave function. The MSMA expansion, due to Murrell and
Shaw251, and Musher and Amos252, incorporates this symmetry only upon the wave func-
tion used as a starting point for the iterative process. The technique is commonly referred
to as symmetry-adapted perturbation theory (SAPT). By treating the interaction as a per-
turbation, the interaction energy itself can be written as an infinite series of terms of higher
and higher order. The first-order terms are similar in nature to the electrostatic and exchange
energies in the other partitioning approaches. For example, the perturbational definition of
exchange energy differs from the aforementioned MO description by terms of fourth order
in the overlap integrals222. To a good approximation, the first-order exchange energy is pro-
portional to the square of the overlap between the two subunits at their equilibrium separa-
tion. Accurate evaluation of this term for long distances requires a good representation of
the tails of the valence orbitals.
Second-order terms include first an induction energy which corresponds roughly to the
sum of polarization and charge transfer in the KM scheme. Dispersion energy makes its first
appearance at second order. Also present here are two manifestations of electron exchange
in the form of induction-exchange and dispersion-exchange. Their sum is commonly re-
ferred to as polarization exchange energy. Whereas empirical expressions can be used to
approximate dispersion or induction, the latter two exchange-related phenomena are more
difficult to approximate. It should be emphasized here that various components occur re-
peatedly at progressively higher orders of perturbation theory. Dispersion, for example, is
not limited to second order but is also present in higher order terms. It is hoped in most cases
that the terms beyond second order are small enough to be safely ignored, but this is not al-
ways the case.
A prime advantage of the perturbational approach is that the individual terms are eval-
uated explicitly, rather than as the difference between very much larger quantities as is true
of supermolecule approaches. These SAPT terms are free of basis set superposition error.
More important, each term corresponds to a well-defined physical phenomenon, which per-
mits an insightful analysis. Each term can be evaluated using a different basis set, most ap-
propriate for that particular component. For example, dispersion requires orbitals of high
angular momentum whereas electrostatics can usually be derived with only a moderate ba-
sis set.
It is possible to express the second-order induction energy in terms of the multipole mo-
ments of any small molecules involved and their static polarizability tensors253 but further
simplification is difficult for a pair of polyatomic subunits. A similar analysis permits the
dispersion to be placed within the context of dynamic polarizabilities. In the case of a pair
38 Hydrogen Bonding

of spherically symmetric atoms, the dispersion reduces to a series in even powers of 1/R,
beginning with R - 6 , with the polarizabilities buried within the coefficients of the series, re-
ferred to as van der Waals constants222. While the expansion is generally divergent, damp-
ing factors may be invoked to permit obtaining useful coefficients based on experimental
data or accurate ab initio computations.
Some of the correspondence between the KM and perturbational values of the various
components can be seen in Table 1.7. The sum of polarization, charge transfer, and mixing
energies is roughly comparable to the induction energy, IND, which first appears at second
order in perturbation theory. The values in Table 1.7 indicate this correspondence is fairly
good, but not exact. It is useful to point out also that when combined together in this man-
ner, the unphysical enlargements of the polarization and mixing terms pointed out earlier
cancel one another to provide a reasonable net induction energy.
Chalasinski and Szczesniak254 have provided a means of decomposing the correlation
contribution to the interaction energy into four separate terms. Their philosophy takes the
electron exchange operator as a second perturbation in the spirit of many-body perturbation
theory, with molecular interaction as the first perturbation in their intermolecular M011er-
Plesset perturbation theory (IMPPT). At the level of second order of the correlation opera-
tor, they obtain a number of separate terms. The first is the dispersion energy, disp (20), cor-
rect through second order of correlation. ES(12) refers to the effect of correlation upon the
Hartree-Fock electrostatic energy. The remaining terms represent the change in the defor-
mation and exchange energies, relative to their SCF values. The third row of Table 1.7 re-
peats the SCF electrostatic energy of the H 3 NH +... NH 3 system, and is followed by the cor-
rection to this term that occurs when correlation is included to the wave function. It is
apparent that these correction terms are of fairly small magnitude and of opposite sign to
the Hartree-Fock level Coulomb energy. The last row lists the dispersion energy computed
for this ionic system and shows it to be a negative quantity that grows quickly as the two
species are brought toward one another.
Cybulski et al.255 furnish an example of the sensitivity of the various perturbation com-
ponents of the H-bond energy to the choice of basis set. In their study of the dimer of HF,
6-31G** refers to a standard split-valence set, with polarization functions. GD is similar in
character but was designed to specifically address the dispersion energy more accurately.
The S2 set was proposed by Sadlej to produce reliable dipole moments and polarizabilities
of the monomers, augmented by extended polarization functions ( on F; d on H). Well-

Table 1.7 Comparison between Kitaura-Morokuma and


perturbational components of the interaction energy of
H 3 NH +... NH 3. a
R(A)

3.25 3.0 2.75 2.50

E
PL+ ECT+ EM IX -5.6 -8.8 -14.8 -27.2
IND -5.0 -8.3 -14.9 -28.7
EES -18.2 -23.5 -31.6 -44.8
(12)
ES 0.5 0.6 0.8 1.0
(20)
disp -1.7 -2.9 -5.0 -8.9
a
All values in kcal/mol. Data 224, calculated using large basis set (S + f).
Quantum Chemical Framework 39

tempered sets were tested as well; referred to as WTS2 when polarized as in the S2 set. The
first row of Table 1.8 illustrates good consistency of the electrostatic term at the SCF level.
The exception is 6-31G** which was not formulated with good monomer properties as a
prime goal. One may expect poorer results with smaller basis sets as the electrostatic term
is fairly sensitive in this regard. The other SCF terms are much less sensitive so moderate-
sized basis sets would generally be appropriate, provided polarization functions are in-
cluded. The authors expressed surprise at the insensitivity of the deformation energy and
postulated that the "ghost orbitals" of the partner help to make up for deficiencies in the de-
scription of each subunit. They also point out the similarity between the perturbational
ind(20) and variational EdefSCF quantities that address the same property from different
perspectives. As a result of the relative constancy of exchange and induction, the full SCF
interaction energy mirrors the sensitivity to basis set of the electrostatic term alone.
As in the SCF case, the correlation correction to the electrostatics, ES(12), indicates a
poor result with 6-31G** but better consistency with the other three basis sets. This term is
repulsive, which is consonant with the general trend in H-bonded systems and is attributed
to a correlation-induced reduction in monomer multipole moments. Indeed, it is common
to find that correlation reduces charge separation within a variety of molecules256. The dis-
persion term is probably most difficult to saturate with sufficient diffuse polarization func-
tions, so exhibits a continued rise in magnitude as the basis set is enlarged. The/functions
included in S2 and WTS2 account for the most negative values there. Exchange-correlation
and deformation-correlation effects are repulsive at the second order, and do not show very
much sensitivity to basis set. On the other hand, they are not insignificant so should be in-
cluded wherever possible. The authors were finally emphatic in pointing out that their re-
sults are much less meaningful if basis set superposition errors are left uncorrected.
It might be noted finally that a perturbational approach offers the possibility of a rigor-
ous definition of nonadditive terms within clusters257. Such unambiguous definitions are
useful in understanding cooperativity, that is, the manner in which one molecule can influ-
ence the interaction between two others.

Table 1.8 Perturbation components to interaction energy of HF


dimer at equilibrium geometry with various basis sets.a
6-31G** GD S2 WTS2

SCF level

ES(10) -7.51 -6.36 -6.19 -6.22


HL
exch 4.39 4.16 4.14 4.14
ind(20) -1.6a9 -1.79 -1.84 -1.86
EdefSCF -1.58 -1.72 -1.81 -1.83
ESCF -4.70 -3.92 -3.86 -3.92

MP2 level
(I2)
ES 0.08 0.55 0.54 0.49
(20) -0.90 -1.17 -1.41 -1.43
disp
exchange/deformation 0.76 0.62 0.59 0.69
E(2) -0.06 0.01 -0.28 -0.26
a
All values in kcal/mol255.
40 Hydrogen Bonding

References
1. Schaefer, H. F., Molecular electronic structure theory: 1972-1975, Annu. Rev. Phys. Chem. 27,
261-290(1976).
2. Schaefer, H. F., ed. Methods of Electronic Structure Theory; vol. 3. Plenum, New York (1977).
3. Ballhausen, C. J. and Gray, H. B., Molecular Electronic Structures; Benjamin/Cummings, Lon-
don (1980).
4. Carsky, P. and Urban, ML, Ab Initio Calculations; Springer-Verlag, Berlin; vol. 16 (1980).
5. Diercksen, G. H. F. and Wilson, S., eds.; Methods in Computational Molecular Physics; Reidel,
Dordrecht; vol. 113(1982).
6. Szabo, A. and Ostlund, N. S., Modern Quantum Chemistry: Introduction to Advanced Electronic
Structure Theory; MacMillan, New York (1982).
7. Hehre, W. J., Radora, L., Schleyer, P. v. R., and Pople, J. A., Ab Initio Molecular Orbital The-
ory; Wiley, New York (1986).
8. Hehre, W. J., Practical Strategies for Electronic Structure Calculations; Wavefunction, Irvine,
CA(1995).
9. Davidson, E. R. and Feller, D., Basis set selection for molecular calculations, Chem. Rev. 86,
681-696(1986).
10. Dunning, T. H. J. and Hay, P. J., In Methods of Electronic Structure Theory; H. F. Schaefer, ed.;
Plenum, New York; vol. 3; pp 1-27. (1977).
11. Hehre, W. J., Stewart, R. F., and Pople, J. A., Self-consistent molecular-orbital methods. I. Use
of Gaussian expansions of Slater-type atomic orbitals, J. Chem. Phys. 51, 2657-2664 (1969).
12. Gordon, M. S.,Binkley, J. S., Pople, J. A., Pietro, W. J., and Hehre, W. J., Self-consistent molec-
ular-orbital methods. 22. Small split-valence basis sets for second-row elements, J. Am. Chem.
Soc. 104,2797-2803(1982).
13. Ditchfield, R., Hehre, W. J., and Pople, J. A., Self-consistent molecular-orbital methods. IX. An
extended Gaussian-type basis for molecular-orbital studies of organic molecules, J. Chem. Phys.
54,724-728(1971).
14. Krishnan, R., Binkley, J. S., Seeger, R., and Pople, J. A., Self-consistent molecular-orbital meth-
ods. XX. A basis set for correlated wave functions., J. Chem. Phys. 72, 650-654 (1980).
15. Frisch, M. J., Pople, J. A., and Binkley, J. S., Self-consistent molecular orbital methods. 25. Sup-
plementary functions for Gaussian basis sets, J. Chem. Phys. 80, 3265—3269 (1984).
16. Glukhovtsev, M. N., Should the standard basis sets be augmented with diffuse functions on hy-
drogens to provide a reasonable description of the lowest Rydberg state of hydrogen-containing
molecules?, J. Mol. Struct. (Theochem) 357, 237-242 (1995).
17. Frost, A. A., Prentice, B. H., and Rouse, R. A., A simple floating localized orbital model of mo-
lecular structure, J. Am. Chem. Soc. 89, 3064-3065 (1967).
18. Burton, P. G., The computation of intermolecular forces with Gaussian basis functions. Illus-
tration with He 2 , J. Chem. Phys. 67, 4696-4700 (1977).
19. Gutowski, M., Verbeek, J., van Lenthe, J. H., and Chalasinski, G., The impact of higher polar-
ization functions on second-order dispersion energy. Partial wave expansion and damping phe-
nomenon for He2, Chem. Phys. 1ll, 271-283 (1987).
20. Tao, F.-M., The use of midbond functions for ab initio calculations of the asymmetric potentials
of He-Ne and He-Ar, J. Chem. Phys. 98, 3049-3059 (1993).
21. Tao, F.-M., On the use of bond functions in molecular calculations, J. Chem. Phys. 98,
2481-2483(1993).
22. Burcl, R., Chalasinski, G., Bukowski, R., and Szczesniak, M. M., On the role of bond functions
in interaction energy calculations: Ar . . . HCl, Ar ... H 2 O, (HF)2, i. Chem. Phys. 103, 1498-1507
(1995).
23. Bardo, R. D. and Ruedenberg, K., Even-tempered atomic orbitals. VI. Optimal orbital exponents
and optimal contractions of Gaussian primitives for hydrogen, carbon, and oxygen in molecules,
.T. Chem. Phys. 60, 918-931 (1974).
Quantum Chemical Framework 41

24. Huzinaga, S., Klobukowski, M., and Tatewaki, H., The well-tempered GTF basis sets and their
applications in the SCF calculations on N2, CO, Na2, and P2, Can. J. Chem. 63, 1812-1828
(1985).
25. Matsuoka, O. and Huzinaga, S., Relativistic well-tempered Gaussian basis sets, Chem. Phys.
Lett. 140, 567-571 (1987).
26. Dunning, T. H. J., Gaussian basis sets for use in correlated molecular calculations. I. The atoms
boron through neon and hydrogen, J. Chem. Phys,, 90, 1007-1023 (1989).
27. Kendall, R. A., Dunning, T. H. J., and Harrison, R. J., Electron affinities of the first-row atoms
revisited. Systematic basis sets and wave functions, J. Chem. Phys. 96, 6796-6806 (1992).
28. Bauschlicher, C. W. J. and Partridge, H., A comparison of correlation-consistent and Pople-type
basis sets, Chem. Phys. Lett. 245, 158-164 (1995).
29. Almlof, J., Helgaker, T., and Taylor, P. R., Gaussian basis sets for high-quality ab initio calcu-
lations, J. Phys. Chem. 92, 3029-3033 (1988).
30. Christiansen, P. A., Lee, Y. S., and Pitzer, K. S., Improved ab initio effective core potentials for
molecular calculations, J. Chem. Phys. 71, 4445-4450 (1979).
31. Stevens, W. J., Basch, H., and Krauss, M., Compact effective potentials and efficient shared-
exponent basis sets for the first- and second-row atoms, J. Chem. Phys. 81, 6026-6033 (1984).
32. Topiol, S. and Osman, R., On the use of minimal valence basis sets with the careless Hartree-
Fock effective potential, J. Chem. Phys. 73, 5191-5196 (1980).
33. Kutzelnigg, W., Electron correlation and electron pair theories, Top. Curr. Chem. 41, 31—73
(1973).
34. Bartlett, R. J.. Many-body perturbation theory and coupled cluster theory for elctron correla-
tion in molecules, Annu. Rev. Phys. Chem. 32, 3 5 9 - 0 1 (1981).
35. Foster, J. M. and Boys, S. F, Canonical configurational interaction procedure, Rev. Mod. Phys.
32, 300-302 (1960).
36. Shavitt, I., In Methods of Electronic Structure Theory; H. F. Schaefer, ed.; vol. 3. Plenum: New
York, pp 189-275. (1977).
37. Shavitt, I., In Advanced Theories and Computational Approaches to the Electronic Structure of
Molecules; C. E. Dykstra, ed.; Reidel: Dordrecht (1984). pp 185.
38. Simons, J., Size extensivity correction for complete active space multiconfiguration self-consis-
tent-field configuration interaction energies, J. Phys. Chem. 93, 626-627 (1989).
39. Langhoff, S. R. and Davidson, E. R., Configuration interaction calculations on the nitrogen mol-
ecule, Int. J. Quantum Chem. 8, 61-72 (1974).
40. Pople, J. A., Seeger, R., and Krishnan, R., Variational configuration interaction methods and
comparison with perturbation theory, Int. J. Quantum Chem., Quantum Chem. Symp. 1l,
149-163(1977).
41. Pople, J. A., Head-Gordon, M., and Raghavachari, K., Quadratic configuration interaction. A
general technique for determining electron correlation energies, J. Chem. Phys. 87, 5968-5975
(1987).
42. Paldus, J., Cizek, J., and Jeziorski, B., Coupled-cluster approach or quadratic configuration in-
teraction?, J. Chem. Phys. 90, 4356-4362 (1989).
43. Ahlrichs, R., Many body perturbation calculations and coupled electron pair models, Comp.
Phys. Comm. 17, 31-45 (1979).
44. Ahlrichs, R., In Methods in Computational Molecular Physics; Diercksen, G. H. F. and Wil-
son, S., eds.; Reidel, Dordrecht; pp 209-226. (1983).
45. Carsky, P., Hubac, I., and Staemmler, V., Correlation energies in open shell systems. Compari-
son of CEPA, P N O - C I , and perturbation treatments based on the restricted Roothaan-Hartree-
Fock formalism, Theor. Chim. Acta 60, 445-450 (1982).
46. Cizek, J. and Paldus, J., Coupled-cluster approach, Physica Scripta 21, 251 (1980).
47. Cizek, J., On the correlation problem in atomic and molecular systems. Calculation of wave-
function components in Ursell-type expansion uisng quantum-field theoretical methods, J.
Chem. Phys. 45, 4256-4266 (1966).
42 Hydrogen Bonding

48. Pople, J. A., Krishnan, R., Schlegel, H. B., and Binkley, J. S., Electron correlation theories and
their application to the study of simple reaction potential surfaces, Int. J. Quantum Chem. 14,
545-560 (1978).
49. Bartlett, R. J. and Purvis, G. D., Many-body perturbation theory, coupled-pair many-electron
theory, and the importance of quadruple excitation for the correlation problem, Int. J. Quantum
Chem. 14,561-581 (1978).
50. Purvis, G. D. and Bartlett, R. J., A full coupled-clusters single and double model: The inclusion
of disconnected triples, J. Chem. Phys. 76, 1910-1918 (1982).
51. Binkley, J. S. and Pople, J. A., M0ller-Plesset theory for atomic ground state energies. Int. J.
Quantum Chem. 9, 229-236 (1975).
52. Pople, J. A., Binkley, J. S., and Seeger, R., Theoretical models incorporating electron correla-
tion, Int. J. Quantum Chem., Quantum Chem. Symp. 10, 1-19 (1976).
53. Krishnan, R. and Pople, J. A., Approximate fourth-order perturbation theory of the electron cor-
relation energy, Int. J. Quantum Chem. 14, 91-100 (1978).
54. Wahl, A. C. and Das, G., In Methods of Electronic Structure Theory; Schaefer, H. F, ed.; vol. 3.
Plenum, New York (1977) pp 51-78.
55. Shepard, R., In Ab Initio Methods in Quantum Chemistry; Lawley, K. P., ed.; vol. 2. Wiley, New
York (1987) pp 63-200.
56. Roos, B. O., Taylor, P. R., and Siegbahn, P. E. M., A complete active space SCF method (CASSCF)
using a density matrix formulated super-Cl approach, Chem. Phys. 48, 157-173 (1980).
57. Roos, B. O., In Ab Initio Methods in Quantum Chemistry; Lawley, K. P., ed.; vol. 2. Wiley, New
York (1987) pp 399-445.
58. Pulay, P., Ab initio calculation of force constants and equilibrium geometries in polyatomic mol-
ecules. I. Theory, Mol. Phys. 17, 197-204 (1969).
59. Schlegel, H. B., Wolfe, S., and Bernardi, R, Ab initio computation of force constants. I. The sec-
ond and third period hydrides, J. Chem. Phys. 63, 3632-3638 (1975).
60. Yamaguchi, Y, Osamura, Y, Goddard, J. D., and Schaefer, H. F, A New Dimension to Quantum
Chemistry: Analytic Derivative Methods in Ab Initio Molecular Electronic Structure Theory;
Oxford University, New York (1994).
61. Moore, T. S. and Winmill, T. F, The state of amines in aqueous solution, J. Chem. Soc. 101,
1635-1676(1912).
62. Latimer, W. M. and Rodebush, W. H., Polarity and ionization from the standpoint of the Lewis
theory of valence, J. Am. Chem. Soc. 42, 1419-1433 (1920).
63. Huggins, M. L., 50 years of hydrogen bond theory, Angew. Chem., Int. Ed. Engl. 10, 147-152
(1971).
64. Huggins, M. L., Hydrogen bridges in organic compounds, J. Org. Chem. 1, 407-456 (1936).
65. Pauling, L., The nature of the chemical bond; Cornell University Press, Ithaca, NY (1939).
66. Cantor, C. R. and Schimmel, P. R., Biophysical Chemistry; vol. 1. Freeman, San Francisco (1980).
67. Baker, E. N. and Hubbard, R. E., Hydrogen bonding in globular proteins, Prog. Biophys. Molec.
Biol. 44, 97-179(1984).
68. Dado, G. P. and Gellman, S. H., Intramolecular hydrogen bonding in derivatives of -alanine
and -amino butyric acid: Model studies for the folding of unnatural polypeptide backbones,
J. Am. Chem. Soc. 116, 1054-1062 (1994).
69. Byrne, M. P., Manuel, R. L., Lowe, L. G., and Stites, W. E., Energetic contribution of side chain
hydrogen bonding to the stability of staphylococcal nuclease, Biochem. 34, 13949-13960 (1995).
70. Shiber, M. C., Lavelle, L., Fossella, J. A., and Fresco, J. R., -DNA, a single-stranded secondary
structure stabilized by ionic and hydrogen bonds: d(A+-G)n, Biochem. 34,14293-14299 (1995).
71. Kiefer, L. L., Paterno, S. A., and Fierke, C. A., Hydrogen bond network in the metal binding site
of carbonic anhydrase enhances zinc affinity and catalytic efficiency, J. Am. Chem. Soc. 117,
6831-6837(1995).
72. Lesburg, C. A. and Christianson, D. W., X-ray crystallographic studies of engineered hydrogen
bond networks in a protein-zinc binding site, J. Am. Chem. Soc. 117, 6838-6844 (1995).
Quantum Chemical Framework 43

73. Harel, M., Quinn, D. M., Nair, H. K., Silman, I., and Sussman, J. L., The x-ray structure of a
transition state analog complex reveals the molecular origins of the catalytic power and sub-
strate specificity of acetylcholinesterase, J. Am. Chem. Soc. 118, 2340-2346 (1996).
74. de Rege, P. J. F, Williams, S. A., and Therien, M. J., Direct evaluation of electronic coupling
mediated by hydrogen bonds: Implications for biological electron transfer, Science 269,
1409-1413 (1995).
75. MacDonald, J. C. and Whitesides, G. M., Solid-state structures of hydrogen-bonded tapes based
on cyclic secondary diamides, Chem. Rev. 94, 2383-2420 (1994).
76. Hanessian, S., Gomtsyan, A., Simard, M., and Roelens, S., Molecular recognition and self-
assembly by "weak" hydrogen bonding: Unprecedented supramolecular helicate structures
from diamine/diol motifs, J. Am. Chem. Soc. 116, 4495-4496 (1994).
77. Mak, T. C. W., Yip, W. H., and Li, Q., Novel hydrogen-bonded host lattices of urea and the elu-
sive allophanate ion, J. Am. Chem. Soc. 117, 11995-11996 (1995).
78. Etter, M. C., Hydrogen bonds as design elements in organic chemistry, J. Phys. Chem. 95,
4601-4610(1991).
79. Etter, M. C., Encoding and decoding hydrogen-bond patterns of organic compounds, Acc. Chem.
Res. 23, 120-126(1990).
80. Feng, W. Y. and Lifshitz, C., Influence of multiple hydrogen bonding on reactivity: Ion/molecule
reactions of proton-bound 12-crown-4 ether and its mixed clusters with ammonia and methanol,
J. Am. Chem. Soc. 117, 11548-11554(1995).
81. Endo, K., Sawaki, T., Koyanagi, M., Kobayashi, K., Masuda, H., and Aoyama, Y, Guest-bind-
ing properties of organic crystals having an extensive hydrogen-bonded network: An orthogonal
anthracene-bis(resorcinol) derivative as a functional organic analog of zeolites, J. Am. Chem.
Soc. 117,8341-8352(1995).
82. Miki, K., Masui, A., Kasai, N., Miyata, M., Shibakami, M., and Takemoto, K., New channel-type
inclusion compound of steroidal bile acid. Structure of a 1:1 complex between cholic acid and
acetophenone, J. Am. Chem. Soc. 110, 6594-6596 (1988).
83. Simard, M., Su, D., and Wuest, J. D., Use of hydrogen bonds to control molecular aggregation.
Self-assembly of three-dimensional networks with large chambers, J. Am. Chem. Soc. 113,
4696-4698(1991).
84. Byrn, M. P., Curtis, C. J., Goldberg, I., Hsiou, Y, Khan, S. I., Sawin, P. A., Tendick, S. K., and
Strouse, C. E., Porphyrin sponges: Structural systematics of the host lattice, J. Am. Chem. Soc.
113,6549-6557(1991).
85. Zimmerman, S. C., Zeng, R, Reichert, D. E. C., and Kolotuchin, S. V., Self-assembling den-
drimers, Science 271, 1095-1098 (1996).
86. Hanessian, S., Simard, M., and Roelens, S., Molecular recognition and self-assembly by non-
amidic hydrogen bonding. An exceptional "assembler" of neutral and charged supramolecular
structures, J. Am. Chem. Soc. 117, 7630-7645 (1995).
87. Huyghues-Despointes, B. M. P., Klingler, T. M., and Baldwin, R. L., Measuring the strength of
side-chain hydrogen bonds in peptide helices: The Gin . Asp (i,i+4) interaction, Biochem. 34,
13267-13271(1995).
88. Clark, T. D. and Ghadiri, M. R., Supramolecular design by covalent capture. Design of a pep-
tide cylinder via hydrogen-bond-promoted intermolecular olefin metathesis, J. Am. Chem. Soc.
117, 12364-12365(1995).
89. Shimizu, K. D. and Rebek, J. J., Synthesis and assembly of self-complementary calix[4]arenes,
Proc. Nat. Acad. Sci., USA 92, 12403-12407 (199.5).
90. Hartgerink, J. D., Granja, J. R., Milligan, R. A., and Ghadiri, M. R., Self-assembling peptide
nanotubes, J. Am. Chem. Soc. 118, 43-50 (1996).
91. Hanessian, S., Yang, H., and Schaum, R., Hydrogen-bonding as a stereocontrolling element in
free-radical C-allylalion reactions: Vicinal, proximal, and remote asymmetric induction in the
amino acid series, J. Am. Chem. Soc. 118, 2507-2508 (1996).
92. Engquist, I., Lestelius, M., and Liedberg, B., Hydrogen bond interactions between self-assem-
44 Hydrogen Bonding

bled monolayers and adsorbed water molecules and its implications for cluster formation, J.
Phys. Chem. 99, 14198-14200 (1995).
93. Nuzzo, R. G., Zegarski, B. R., Korenic, E. M., and Dubois, L. H., Infrared studies of water ad-
sorption on model organic surfaces, J. Phys. Chem. 96, 1355-1361 (1992).
94. Pimentel, G. C. and McClellan, A. L., The Hydrogen Bond; Freeman, San Francisco (1960).
95. Joesten, M. D. and Schaad, L. J., Hydrogen Bonding; Marcel Dekker, New York (1974).
96. Jeffrey, G. A. and Saenger, W., Hydrogen Bonding in Biological Structures; Springer-Verlag,
Berlin (1991).
97. Sennikov, P. G., Weak H-bond by second-row (PH 3 , H2S) and third-row (AsH 3 , H2Se) hydrides,
J. Phys. Chem. 98, 4973-4981 (1994).
98. Gilli, P., Bertolasi, V., Ferretti, V., and Gilli, G., Covalent nature of the strong homonuclear hy-
drogen bond. Study of the O—H . . . O system by crystal structure correlation methods, J. Am.
Chem. Soc. 116, 909-915 (1994).
99. Sokolov, N. D. and Savel'ev, V. A., Dynamics of the hydrogen bond: Two-dimensional model
and isotope effects, Chem. Phys. 22, 383-399 (1977).
100. Olovsson, I. and Jonsson, P.-G., In The Hydrogen Bond, Recent Developments in Theory and Ex-
periments; Schuster, P., Zundel, G., and Sandorfy, C., eds.; vol. 2. North-Holland Publishing Co.,
Amsterdam (1976). pp 393-456.
101. Panich, A. M., NMR study of the F — H . . . F hydrogen bond. Relation between hydrogen atom po-
sition and F-H . . . F bond length, Chem. Phys. 196, 511-519 (1995).
102. Badger, R. M. and Bauer, S. H., Spectroscopic studies of the hydrogen bond. II. The shifts of the
O—H vibrational frequency in the formation of the hydrogen bond, J. Chem. Phys. 5, 839-851
(1939).
103. Laurence, C., Berthelot, M., Helbert, M., and Sraidi, K., First measurement of the hydrogen bond
basicity of monomeric water, phenols, and weakly basic alcohols, J. Phys. Chem. 93, 3799—3802
(1989).
104. Tucker, E. E. and Lippert, E., In The Hydrogen Bond. Recent Developments in Theory and Ex-
periments; Schuster, P., Zundel, G., and Sandorfy, C., eds.; vol. 2. North-Holland Publishing Co.,
Amsterdam (1976) pp 791-830.
105. Berglund, B. and Vaughan, R. W., Correlation between proton chemical shift tensors, deuterium
quadmpole couplings, and bond distances for hydrogen bonds in solids, J. Chem. Phys. 73,
2037-2043(1980).
106. Jeffrey, G. A. and Yeon, Y, The correlation between hydrogen bond lengths and proton chem-
ical shifts in crystals, Acta Cryst. B42, 410-413 (1986).
107. Ando, S.,Ando, I., Shoji, A., and Ozzaki,T., lntermolecularhydrogen-bonding effect on I3CNMR
chemical shifts of glycine residue carbonyl cations of peptides in the solid state, J. Am. Chem.
Soc. 110, 3380-3386(1988).
108. Roller, H., Lobo, R. F., Burkett, S. L., and Davis, M. E., SiO - . . . HOSi hydrogen bonds in as-syn-
thesized high-silica zeolites, J. Phys. Chem. 99, 12588-12596 (1995).
109. Eckert, H., Yesinowski, J. P., Silver, L. A., and Stolper, E. M., Water in silicate glasses: Quan-
titation and structural studies by 1H solid echo and MAS-NMR methods, J. Phys. Chem. 92,
2055-2064(1988).
110. Kaliaperumal, R., Sears, R. E. J., Ni, Q. W., and Furst, J. E., Proton chemical shifts in some hy-
drogen bonded solids and a correlation with bond lengths, J. Chem. Phys. 91,7387-7391 (1989).
111. Gu, Z., Ridenour, C. F., Bronnimann, C. E., Iwashita, T., and McDermott, A., Hydrogen bond-
ing and distance studies of amino acids and peptides using solid state 2D 1 H — 1 3 C heteronu-
clear correlation spectra, J. Am. Chem. Soc. 118, 822-829 (1996).
112. Wishart, D. S., Sykes, B. D., and Richards, F. M., Relationship between nuclear magnetic reso-
nance chmical shift and protein secondary structure, J. Mol. Biol. 222, 311—333 (1991).
113. Zhou, N. E., Zhu, B.-Y, Sykcs, B. D., and Hodges, R. S., Relationship between amide proton
chemical shifts and hydrogen bonding inamphipalhic a-helical peptides, J. Am. Chem. Soc. 114,
4320-4326(1992).
Quantum Chemical Framework 45

114. Altman, L. J., Lauggani, D., Gunnarsson, G., and Wennerstrom, H., Proton, deuterium, and tri-
tium nuclear magnetic resonance of intramolecular hydrogen bonds. Isotope effects and the
shape of the potential energy function, J. Am. Chem. Soc. 100, 8264-8266 (1978).
115. Vener, M. V., Model study of the primary H/D isotope effects on the NMR chemical shift in strong
hydrogen-bonded systems, Chem. Phys. 166, 311-316 (1992).
116. Gu, Z., Zambrano, R., and McDermott, A., Hydrogen bonding of carboxyl groups in solid-state
amino acids and peptides: Comparison of carbon chemical shielding, infrared frequencies, and
structures, J. Am. Chem. Soc. 116, 6368-6372 (1994).
117. Zhong, S., Haghjoo, K., Kettner, C., and Jordan, F., Proton magnetic resonance studies of the
active center histidine of chymotrypsin complexes to peptideboronic acids: Solvent accessibility
to the N and N sites can differentiate slow-binding and rapidly reversible inhibitors, J. Am.
Chem. Soc. 117, 7048-7055 (1995).
118. Bachovchin, W. W., I5N NMR spectroscopy of hydrogen-bonding interactions in the active site of
serine proteases: Evidence for a moving histidine mechanism, Biochem. 25, 7751-7759 (1986).
119. Golubev, N. S., Smirnov, S. N., Gindin, V. A., Denisov, G. S., Benedict, H., and Limbach,
H.-H., Formation of charge relay chains between acetic acid and pyridine observed by low-
temperature nuclear magnetic resonance, J. Am. Chem. Soc. 116, 12055—12056 (1994).
120. Le-Thanh, H. and Vocelle, D., Nuclear magnetic resonance study of complexes formed between
conjugated Schiff bases and carboxylic acids, Can. J. Chem. 72, 2220-2224 (1994).
121. Berger, S., The conformational dependence of15N13C spin spin coupling constants, Tetrahedron
34,3133-3136(1978).
122. Walter, J. A. and Wright, L. C., I3C— 15N spin-spin coupling constants of amides: H-bonding ef-
fects in a cyclohexapeptide, Tetrahedron Lett. 41, 3909-3912 (1979).
123. Juranic, N., Ilich, P. K., and Macura, S., Hydrogen bonding networks in proteins as revealed by
the amide 'JNC, coupling constant, J. Am. Chem. Soc. 117, 405-410 (1995).
124. Aliev, A. E., Harris, K. D. M., Shannon, I. J., Glidewell, C., Zakaria, C. M., and Schofield, P. A.,
Solid-state 2H and 13C NMR studies of hydrogen-bond dynamics in ferrocene-1,1'-diylbis
(diphenylmethanol), J. Phys. Chem. 99, 12008-12015 (1995).
125. Kuroki, S., Takahashi, A., Ando, L, Shoji, A., and Ozaki, T., Hydrogen-bonding structural study
of solid peptides and polypeptides containing a glycine residue by 17O NMR spectroscopy,
J. Mol. Struct. 323, 197-208 (1995).
126. Kuroki, S., Ando, S., and Ando, I., An MO study of nuclear quadrupolar coupling constant and
nuclear shielding of the carbonyl oxygen in solid peptides with hydrogen bonds, Chem. Phys.
195, 107-116(1995).
127. Fann, Y.-C., Ong, J.-L., Nocek, J. M., and Hoffman, B. M., 19F and 1 , 2 HENDOR study of distal-
pocket N( )-H . . . F hydrogen bonding influorometmyoglobin, J. Am. Chem. Soc. 117, 6109-6116
(1995).
128. Wozniak, K., He, H., Klinowski, J., Jones, W., and Barr, T. L., ESCA, solid-state NMR, and
X-ray diffraction monitor the hydrogen bonding in a complex of 1,8-bis(dimethylamino)naph-
thalene with 1,2-dichloromaleic acid, J. Phys. Chem. 99, 14667-14677 (1995).
129. Winkler, A. and Hess, P., Study of the energetics and dynamics of hydrogen bond formation in
aliphatic carboxylic acid vapors by resonant photoacoustic spectroscopy, J. Am. Chem. Soc.
116,9233-9240(1994).
130. Winkler, A., Mehl, J. B., and Hess, P., Chemical relaxation of H bonds in formic acid vapor stud-
ied by resonant photoacoustic spectroscopy, J. Chem. Phys. 100, 2717-2727 (1994).
131. Sauren, H., Winkler, A., and Hess, P., Kinetics and energetics of hydrogen bond dissociation in
isolated acetic acid-d 1 and -d4 and trifluoroacetic acid dimers, Chem. Phys. Lett. 239, 313—319
(1995).
132. Murthy, A. S. N. and Rao, C. N. R., Recent theoretical studies of the hydrogen bond, J. Mol.
Struct. 6, 253-282(1970).
133. Kollman, P. A. and Allen, L. C., The theory of the hydrogen bond, Chem. Rev. 72, 283-303
(1972).
46 Hydrogen Bonding

134. Kollman, P., McKelvey, J., Johansson, A., and Rothenberg, S., Theoretical studies of hydrogen-
bonded dimers. Complexes involving HF, H2O, NH3, HCl, H2S, PH3, HCN, HNC, HCP, CH2NH,
H2CS, H2CO, CH4, CF3H, C2H2, C4H4, C 6 H 6 , F-, and H3O+, J. Am. Chem. Soc. 97, 955-965
(1975).
135. Allen, L. C., A simple model of hydrogen bonding, J. Am. Chem. Soc. 97, 6921-6940 (1975).
136. Dill, J. D., Allen, L. C., Topp, W. C., and Pople, J. A., A systematic study of the nine hydrogen-
bonded dimers involving NH3, OH2, and HF, J. Am. Chem. Soc. 97, 7220-7226 (1975).
137. Kollman, P., A general analysis of noncovalent intermolecular interactions, J. Am. Chem. Soc.
99, 4875-4894 (1977).
138. Umeyama, H. and Morokuma, K., The origin of hydrogen bonding. An energy decomposition
study, J. Am. Chem. Soc. 99, 1316-1332 (1977).
139. Morokuma, K., Why do molecules interact? The origin of electron donor-acceptor complexes,
hydrogen bonding, and proton affinity, Ace. Chem. Res. 10, 294-300 (1977).
140. Pople, J. A., Intermolecular binding, Faraday Discuss. Chem. Soc. 73, 7-17 (1982).
141. Hobza, P. and Zahradnik, R., Van der Waals molecules: Quantum chemistry, physical properties,
and reactivity, Int. J. Quantum Chem. 23, 325-338 (1983).
142. Astrand, P.-O., Karlstrom, G., Engdahl, A., and Nelander, B., Novel model for calculating the
intermolecular part of the infrared spectrum for molecular complexes, J. Chem. Phys. 102,
3534-3554 (1995).
143. Del Bene, J. E., Mettee, H. D., Frisch, M. J., Luke, B. T., and Pople, J. A., Ab initio computation
of the enthalpies of some gas-phase hydration reactions, J. Phys. Chem. 87, 3279-3282 (1983).
144. Levine, I. N., Physical Chemistry; 3rd ed.; McGraw-Hill: New York, 1988.
145. Espinosa, E., Lecomte, C., Ghermani, N. E., Devemy, J., Rohmer, M. M., Benard, M., and
Molins, E., Hydrogen bonds: First quantitative agreement between electrostatic potential cal-
culations from experimental X-(X + N) and theoretical ab initio SCF models, J. Am. Chem. Soc.
118,2501-2502(1996).
146. Koch, U. and Popelier, P. L. A., Characterization of C—H—O hydrogen bonds on the basis of
the charge density, J. Phys. Chem. 99, 9747-9754 (1995).
147. Carroll, M. T. and Bader, R. F. W., An analysis of the hydrogen bond in BASE-HF complexes us-
ing the theory of atoms in molecules, Mol. Phys. 65, 695-722 (1988).
148. Boyd, R. J., Hydrogen bonding between nitrites and hydrogen halides and the topologicalprop-
erties of molecular charge distributions, Chem. Phys. Lett. 129, 62-65 (1986).
149. Politzer, P. and Truhlar, D. G., eds. Chemical Applications of Atomic and Molecular Electrosta-
tic Potentials; Plenum, New York (1981).
150. Murray, J. S., Ranganathan, S., and Politzer, P., Correlation betwen the solvent hydrogen bond
acceptor parameter fl and the calculated electrostatic potential, J. Org. Chem. 56, 3734-3737
(1991).
151. Murray, J. S. and Politzer, P., Correlation betwen the solvent hydrogen-bond-donating parame-
ter a and the calculated molecular surface electrostatic potential, J. Org. Chem. 56, 6715-6717
(1991).
152. Hagelin, H., Murray, J. S., Brinck, T., Berthelot, M., and Politzer, P., Family-independent rela-
tionships between computed molecular surface quantities and solute hydrogen bond acidity/ba-
sicity and solute-induced methanol O—H infrared frequency shifts, Can. J. Chem. 73, 483-488
(1995).
153. Ditchfield, R., Self-consistent perturbation theory of diamagnetism. L A gauge-invariant LCAO
method for N.M.R. chemical shifts, Mol. Phys. 27, 789-807 (1974).
154. Ditchfield, R., Theoretical studies of magnetic shielding in H2O and (H2O)2, J. Chem. Phys. 65,
3123-3133(1976).
155. Ditchfield, R., GIAO studies of magnetic shielding in FHF- and HF, Chem. Phys. Lett. 40,
53-56(1976).
156. Ditchfield, R. and McKinncy, R. E., Molecular orbital studies of the NMR hydrogen bond shift,
Chem. Phys. 13, 187-194 (1976).
Quantum Chemical Framework 47

157. Rohlfing, C. M., Allen, L. C., and Ditchfield, R., Proton chemical shift tensors in neutral and
ionic hydrogen bonds, Chem. Phys. Lett. 86, 380-383 (1982).
158. Chesnut, D. B. and Phung, C. G., Functional counterpoise corrections for the NMR chemical
shift in a model dimeric water system, Chem. Phys. 147, 91-97 (1990).
159. Kutzelnigg, W., Theory of magnetic susceptibilities and NMR chemical shifts in terms of local-
ized quantities., Isr. J. Chem. 19, 193-200 (1980).
160. Schindler, M. and Kutzelnigg, W., Theory of magnetic susceptibilities and NMR chemical shifts
in terms of localized quantities. II. Application to some simple molecules, J. Chem. Phys. 76,
1919-1933(1982).
161. Gauss, J., Calculation of NMR chemical shifts at second-order many-body perturbation theory
using gauge-including atomic orbitals, Chem. Phys. Lett. 191, 614-620 (1992).
162. Gauss, J., Effects of electron correlation in the calculation of nuclear magnetic resonance chem-
ical shifts, J. Chem. Phys. 99, 3629-3643 (1993).
163. Buhl, M., Thiel, W., Fleischer, U., and Kutzelnigg, W., Ab initio computation of 77Se NMR chem-
ical shifts with the IGLO-SCF, the GIAO-SCF, and the GIAO-MP2 methods, J. Phys. Chem. 99,
4000-4007 (1995).
164. Gauss, J. and Stanton, J. R, Perturbative treatment of triple excitations in coupled-cluster cal-
culations of nuclear magnetic shielding constants, J. Chem. Phys. 104, 2574—2583 (1996).
165. van Wiillen, C. and Kutzelnigg, W., Calculation of nuclear magnetic resonance shieldings and
magnetic susceptibilities using multiconfiguration Hartree-Fock wave functions and local gauge
origins, J. Chem. Phys. 104, 2330-2340 (1996).
166. Eggenberger, R., Gerber, S., Huber, H., Searles, D., and Welker, M., The use of molecular dy-
namics simulations with ab initio SCF calculations for the determination of the oxygen-17
quadrupole coupling constant in liquidwater, Mol. Phys. 80, 1177-1182 (1993).
167. Ludwig, R., Weinhold, R, and Parrar, T. C., Temperature dependence of hydrogen bonding in
neat, liquidformamide, J. Chem. Phys. 103, 3636-3642 (1995).
168. Kestner, N. R., He-He interaction in the SCF-MO approximation, J. Chem. Phys. 48, 252-257
(1968).
169. Boys, S. F and Bernardi, R, The calculation of small molecular interactions by the differences
of separate total energies. Some procedures with reduced errors, Mol. Phys. 19, 553-566
(1970).
170. Karlstrom, G. and Sadlej, A. J., Basis set superposition effects on properties of interacting sys-
tems. Dipole moments and polarizabilities, Theor. Chim. Acta 61, 1—9 (1982).
171. Dacre, P. D., A calculation of the helium pair polarizability including correlation effects, Mol.
Phys. 36, 541-551 (1978).
172. Dacre, P. D., On the pair polarizability of helium, Mol. Phys. 45, 17-32 (1982).
173. Fowler, P. W. and Buckingham, A. D., The long range model of intermolecular forces. An SCF
study of NeHF, Mol. Phys. 50, 1349-1361 (1983).
174. Kurdi, L., Kochanski, E., and Diercksen, G. H. R, Determination of the basis set superposition
error with "DZP" basis sets in SCF calculations: CO + H2, NH3 + H2, H2 + H2, Chem. Phys.
92, 287-294 (1985).
175. Latajka, Z. and Scheiner, S., Primary and secondary basis set superposition error at the SCF
andMP2 levels: H3N..Li+ and H 2 O .. Li + , J. Chem. Phys. 87, 1194-1204 (1987).
176. Johansson, A., Kollman, P., and Rothenberg, S.,An application of the functional Boys-Bernardi
counterpoise method to molecular potential surfaces, Theor. Chim. Acta 29, 167-172 (1973).
177. Schwenke, D. W. and Truhlar, D. G., Systematic study of basis set superposition errors in the
calculated interaction energy of two HF molecules, J. Chem. Phys. 82, 2418-2426 (1985).
178. Daudey, J. P., Claverie, P., and Malrieu, J. P., Perturbation ab initio calculations of intermolec-
ular energies. I. Method, Int. J. Quantum Chem. 8, 1—15 (1974).
179. Collins, J. R. and Gallup, G. A., The full versus the virtual counterpoise correction for basis
set superposition error in self-consistent field calculations, Chem. Phys. Lett. 123, 56-61
(1986).
48 Hydrogen Bonding

180. Yang, J. and Kestner, N. R., Accuracy of counterpoise corrections in second-order intermolec-
ularpotential calculations. 2. Various molecular dimers, J. Phys. Chera. 95, 9221-9230(1991).
181. Tolosa, S., Espinosa, J., and Olivares del Valle, F. J., Computation of spectroscopic properties of
van der Waals systems from post-SCF ab initio potentials including the EICP alternative coun-
terpoise technique, J. Comput. Chem. 5, 611-619 (1991).
182. Alagona, G., Ghio, G., Cammi, R., and Tomasi, J., The effect of "full" and "limited" counter-
poise corrections with different basis sets on the energy and the equilibrium distance of hydro-
gen bonded dimers, Int. J. Quantum Chem. 32, 207-226 (1987).
183. Loushin, S. K., Liu, S., and Dykstra, C. E., Improved counterpoise corrections for the ab initio
calculation of hydrogen bonding interactions, J. Chem. Phys. 84, 2720-2725 (1986).
184. Loushin, S. K. and Dykstra, C. E., Polarization counterpoise corrections to correlated hydro gen
bond interactions, J. Comput. Chem. 8, 81-83 (1987).
185. Sadlej, A. J., Exact perturbation treatment of the basis set superposition correction, J. Chem.
Phys. 95, 6705-6711 (1991).
186. Ostlund, N. S. and Merrifield, D. L., Ghost orbitals and the basis set extension effects, Chem.
Phys. Lett. 39, 612-614 (1976).
187. Gutowski, M., van Lenthe, J. H., Verbeek, J., van Duijneveldt, F. B., and Chalasinski, G., The
basis set superposition error in correlated electronic structure calculations, Chem. Phys. Lett.
124, 370-375 (1986).
188. Gutowski, M., van Duijneveldt, F. B., Chalasinski, G., and Piela, L., Does the Boys and Bernardi
function counterpoise method actually overcorrect the basis set superposition error?, Chem.
Phys. Lett. 129, 325-330 (1986).
189. Gutowski, M., van Duijneveldt, F. B., Chalasinski, G., and Piela, L., Proper correction for the
basis set superposition error in SCF calculations of intermolecular interactions, Mol. Phys. 61,
233-247 (1987).
190. van Duijneveldt, F. B., van Duijneveldt-van de Rigdt, J. G. C. M., and van Lenthe, J. H., State
of the art in counterpoise theory, Chem. Rev. 94, 1873-1885 (1994).
191. Sordo, J. A., Sordo, T. L., Fernandez, G. M., Gomperts, R., Chin, S., and Clementi, E., A sys-
tematic study on the basis set superposition error in the calculation of interaction energies of
systems of biological interest, J. Chem. Phys. 90, 6361-6370 (1989).
192. Cook, D. B., Sordo, J. A., and Sordo, T. L., Some comments on the counterpoise correction for
the basis set superposition error at the correlated level, Int. J. Quantum Chem. 48, 375—384
(1993).
193. Alagona, G., Cammi, R., Ghio, C., and Tomasi, J., Noncovalent interactions of medium strength.
A revised interpretation and examples of its applications, Int. J. Quantum Chem. 35, 223-239
(1989).
194. Sokalski, W. A., Roszak, S., and Pecul, K.,An efficient procedure for decomposition of the SCF
interaction energy into components with reduced basis set dependence, Chem. Phys. Lett. 153,
153-159(1988).
195. Alagona, G. and Ghio, C., Basis set superposition errors for Slater vs. gaussian basis functions
in H-bond interactions, J. Mol. Struct. (Theochem) 330, 77-83 (1995).
196. Sellers, H. and Almlof, J., On the accuracy in ab initio force constant calculations with respect
to basis set, J. Phys. Chem. 93, 5136-5139 (1989).
197. Nowek, A. and Leszczynski, J., Ab initio study on the stability and properties of XYCO . . . HZ com-
plexes. HI. A comparative study of basis set and electron correlation effects for H 2 CO ... HCl,
J. Chem. Phys. 104, 1441-1451 (1996).
198. Mayer, I., Towards a "chemical" Hamiltonian, Int. J. Quantum Chem. 23, 341-363 (1983).
199. Mayer, I., On the non-additivity of the basis set superposition error and how to prevent its ap-
pearance, Theor. Chim. Acta 72, 207-210 (1987).
200. Mayer, I. and Vibok, A., A BSSE-free SCF algorithm for intermolecular interactions, Int. J.
Quantum Chem. 40, 139-148 (1991).
201. Mayer, I., Vibok, A., Halasz, G., and Valiron, P., A BSSE-free SCF algorithm for intermolecular
Quantum Chemical Framework 49

interactions. III. Generalization for three-body systems and for using bond functions, Int. J.
Quantum Chem. 57, 1049-1055 (1996).
202. Vibok, A. and Mayer, I., A BSSE-free SCF algorithm for intermolecular interactions. II. Sample
calculations on hydrogen-bonded complexes, Int. J. Quantum Chem. 43, 801-811 (1992).
203. Pye, C. C, Poirier, R. A., Yu, D., and Surjan, P. R., Ab initio Hartree-Fock calculations of the in-
teraction energy of the bimolecular complexes, J. Mol. Struct. (Theochem) 307, 239-259 (1994).
204. Gutowski, M. and Chalasinski, G., Critical evaluation of some computational approaches to the
problem of basis set superposition error, J. Chem. Phys. 98, 5540-5554 (1993).
205. Wind, P. and Heully, J. L., Reduction of the basis set superposition error at the correlation level,
Chem. Phys. Lett. 230, 35-40 (1994).
206. Wells, B. H. and Wilson, S., van der Wauls interaction potentials. Basis set superposition effects
in electron correlation calculations, Mol. Phys. 50, 1295-1309 (1983).
207. Tao, F.-M., A new approach to the efficient basis set for accurate molecular calculations: Ap-
plications to diatomic molecules, J. Chem. Phys. 100, 3645-3650 (1994).
208. Tao, F.-M., The counterpoise method and bond functions in molecular dissociation energy cal-
culations, Chem. Phys. Lett. 206, 560-564 (1993).
209. Novoa, I. J., Planas, M., and Whangbo, M.-H., A numerical evaluation of the counterpoise
method on hydrogen bond complexes using near complete basis sets, Chem. Phys. Lett. 225,
240-246(1994).
210. Cybulski, S. M. and Chalasinski, G., Perturbation analysis of the supermolecule interaction en-
ergy and the basis set superposition error, Chem. Phys. Lett. 197, 591-598 (1992).
211. Gutowski, M., Szczesniak, M. M., and Chalasinski, G., Comment on "A possible definition of
basis set superposition error", Chem. Phys. Lett. 241, 140-145 (1995).
212. Chalasinski, G. and Gutowski, M., Dimer centred basis set in the calculations of the first-order
interaction energy with CI wave/unction. The He dimer, Mol. Phys. 54, 1173-1184 (1985).
213. Szczesniak, M. M. and Scheiner, S., Correction of the basis set superposition error in SCF and
MP2 interaction energies. The water dimer, J. Chem. Phys. 84, 6328-6335 (1986).
214. Davidson, E. R. and Chakravorty, S. J., A possible definition of basis set superposition error,
Chem. Phys. Lett. 217, 48-54 (1994).
215. Davidson, E. R. and Chakravorty, S. J., Reply to comment on "A possible definition of basis set
superposition error", Chem. Phys. Lett. 241, 146-148 (1995).
216. Latajka, Z., Scheiner, S., and Chalasinski, G., Basis set superposition error in proton transfer
potentials, Chem. Phys. Lett. 196, 384-389 (1992).
217. Turi, L. and Dannenberg, J. J. Correcting for basis set superposition error in aggregates con-
taining more than two molecules: Ambiguities in the calculation of the counterpoise correction,
J. Phys. Chem. 97, 2488-2490 (1993).
218. Abkowicz, A. J., Latajka, Z., Scheiner, S., and Chalasinski, G., Site-site function and successive
reaction counterpoise calculation of basis set superposition error for proton transfer, J. Mol.
Struct. (Theochem) 342, 153-159 (1995).
219. Buckingham, A. D., In Intermolecular Interactions: From Diatomics to Biopolymers; B. Pull-
man, ed.; Wiley: New York, 1978; pp 1-67.
220. Morokuma, K. and Kitaura, K., In Molecular Interactions; H. Ratajczak and W. J. Orville-
Thomas, eds.; Wiley: New York, 1980; vol. 1; pp 21-87.
221. Chalasinski, G. and Gutowski, M., Weak interactions between small systems. Models of study-
ing the nature of intermolecular forces and challenging problems for ab initio calculations,
Chem. Rev. 88, 943-962 (1988).
222. Jeziorski, B. and Kolos, W., In Molecular Interactions; Ratajczak, H. and Orville-Thomas,
W. J., eds.; vol. 3. Wiley, New York (1982). pp 1-46.
223. Frey, R. F. and Davidson, E. R., Energy partitioning of the self-consistent field interaction en-
ergy of ScCO, J. Chem. Phys. 90, 5555-5562 (1989).
224. Cybulski, S. M. and Scheiner, S., Comparison of Morokuma and perturbation theory approaches
to decomposition of interaction energy. (NH4) + —NH3, Chem. Phys. Lett. 166, 57-64 (1990).
50 Hydrogen Bonding

225. Hobza, P., Sponer, J., and Polasek, M., H-bonded and stacked DNA base pairs: cytosine dimer.
An ab initio second-order M0ller-Plesset study, J. Am. Chem. Soc. 117, 792-798 (1995).
226. Claverie, P., In Intermolecular Interactions: From Diatomics to Biopolymers; Pullman, B., ed.;
Wiley, New York (1978) pp 69-305.
227. Morokuma, K. and Kitaura, K., In Chemical Applications of Atomic and Molecular Electrosta-
tic Potentials; Politzer, P. and Truhlar, D. G., eds.; Plenum, New York (1981) pp 215-242.
228. Sokalski, W. A., Roszak, S., Hariharan, P. C., and Kaufman, J. J. Improved SCF interaction en-
ergy decomposition scheme corrected for basis set superposition effect, Int. J. Auantum Chem
23, 847-854 (1983).
229. del Valle, F. J. O., Tolosa, S., and Espinosa, J., Basis set superposition effects in electronic popu-
lations calculatedon hydrogen bonded systems, J. Mol. Struct. (Theochem) 120,277-283 (1985).
230. Bonaccorsi, R., Cammi, R., and Tomasi, J., Counterpoise corrections to the components of bi-
molecular energy interactions: An examination of three methods of decomposition, Int. J. Quan-
tum Chem. 29, 373-378 (1986).
231. Cammi, R., Del Valle, F. J. O., and Tomasi, J., Decomposition of the interaction energy with
counterpoise corrections to the basis set superposition error for dimers in solution. Method and
application to the hydrogen fluoride dimer, Chem. Phys. 122, 63-74 (1988).
232. Alagona, G., Ghio, C., Latajka, Z., and Tomasi, J., Basis set superposition errors and counter-
poise corrections for some basis sets evaluated for a few X-...M dimers, J. Phys. Chem. 94,
2267-2273 (1990).
233. Olszewski, K. A., Gutowski, M., and Piela, L., Interpretation of the hydrogen-bond energy at the
Hartree-Fock level for pairs of HF, H2O, and NH3 molecules, L Phys. Chem. 94, 5710-5714
(1990).
234. Alagona, G., Ghio, C., Cammi, R., and Tomasi, J., The decomposition of the SCF interaction en-
ergy in hydrogen bonded dimers corrected for basis set superposition errors: An examination of
the basis set dependence, Int. J. Quantum Chem. 32, 227-248 (1987).
235. Stevens, W. J. and Fink, W. H., Frozen fragment reduced variational space analysis of hydrogen
bonding interactions. Application to the water dimer, Chem. Phys. Lett. 139, 15-22 (1987).
236. Reed, A. E., Weinhold, F., Curtiss, L. A., and Pochatko, D. J., Natural bond orbital analysis of
molecular interactions: Theoretical studies of binary complexes of HF, H2O, NH3, N2, O2, F2,
CO and CO2 with HF, H2O, and NH3, J. Chem. Phys. 84, 5687-5705 (1986).
237. Reed, A. E., Curtiss, L. A., and Weinhold, F., Intermolecular interactions from a natural bond
orbital, donor-acceptor-viewpoint, Chem. Rev. 88, 899-926 (1988).
238. Reed, A. E., and Weinhold, F., Natural bond orbital analysis of near Hartree-Fock water dimer,
J. Chem. Phys. 78, 4066-4073 (1983).
239. Glendening, E. D. and Streiwieser, A., Natural energy decomposition analysis: An energy par-
titioning procedure for molecular interactions with application to weak hydrogen bonding,
strong ionic, and moderate donor-acceptor interactions, J. Chem. Phys. 100, 2900-2909 (1994).
240. Jensen, J. H. and Gordon, M. S., Ab initio localized charge distributions: Theory and a detailed
analysis of the water dimer-hydrogen bond, J. Phys. Chem. 99, 8091-8097 (1995).
241. Fischer, H. and Kollmar, H., Energy partitioning with the CNDO method, Theor. Chim. Acta 16,
163-174(1970).
242. Kollmar, H., Partitioning scheme for the ab initio SCF energy, Theor. Chim. Acta 50, 235-262
(1978).
243. Dewar, M. J. S. and Lo, D. H., Ground states of -bonded molecules. XIV. Application of energy
partitioning to the MINDO/2 method and a study of the Cope rearrangement, J. Am. Chem. Soc.
93,7201-7207(1971).
244. Gordon, M. S., A molecular orbital study of internal rotation, J. Am. Chem. Soc. 91, 3122-3130
(1969).
245. Ichikawa, H. and Ebisawa, Y., Hartree-Fock MO theoretical approach to aromaticity. Interpre-
tation of Hiickel resonance energy in terms of kinetic energy of electrons, J. Am. Chem. Soc.
107, 1161-1165 (1985).
Quantum Chemical Framework 5I

246. Hiberty, P. C., Shaik, S. S., Lefour, J.-M., and Ohanessian, G., Is the delocalizedn-system of ben-
zene a stable electronic system, J. Org. Chem. 50,4657-4659 (1985).
247. Jeziorski, B. and van Hemert, M., Variation-perturbation treatment of the hydrogen bond be-
tween water molecules, Mol. Phys. 31, 713-729 (1976).
248. Arrighini, P., Intermolecular forces and their evaluation by perturbation theory, vol. 25.
Springer-Verlag, Berlin (1981).
249. Szalewicz, K., Jeziorski, B., and Rybak, S., Perturbation theory calculations of intermolecular
interaction energies, Int. J. Quantum Chem. QBS18, 23-36 (1991).
250. Williams, H. L., Mas, E. M., Szalewicz, K., and Jeziorski, B., On the effectiveness of
monomer-, dimer-, and bond-centered basis functions in calculations of intermolecular interac-
tion energies, J. Chem. Phys. 103, 7374-7391 (1995).
251. Murrell, J. N. and Shaw, G., Intermolecular forces in the region of small orbital overlap, J. Chem.
Phys. 46, 1768-1772 (1967).
252. Musher, J. I. and Amos, A. T., Theory of weak atomic and molecular interactions, Phys. Rev.
164,31-43(1967).
253. Dalgarno, A., Adv. Phys. 12, 143 (1962).
254. Chalasinski, G. and Szczesniak, M. M., On the connection between the supermolecular Mft/ller-
Plesset treatment of the interaction energy and the perturbation theory of intermolecular forces,
Mol. Phys. 63, 205-224 (1988).
255. Cybulski, S. M., Chalasinski, G., and Moszynski, R., On decomposition of second-order M0ller-
Plesset supermolecular interaction energy and basis set effects, J. Chem. Phys. 92, 4357-4363
(1990).
256. Carpenter, J. E., McGrath, M. P., and Hehre, W. J., Effect of electron correlation on atomic elec-
tron populations, J. Am. Chem. Soc. 1 l l , 6154-6156 (1989).
257. Moszynski, R., Wormer, P. E. S., Jeziorski, B., and van der Avoird, A., Symmetry-adapted per-
turbation theory of nonadditive three-body interactions in van der Waals molecules. I. General
theory, J. Chem. Phys. 103, 8058-8074 (1995).
2

Geometries and Energetics

n this chapter, we focus our attention on the equilibrium geometries of various H-bonds,
I and how the formation of the complex alters the internal structure of each subunit. The
energetics of hydrogen bonding are also stressed. Comparison is made to experimental in-
formation where available. Whereas many geometries have been evaluated to high preci-
sion, energetic data in the gas phase, to which the calculations directly pertain, have been
harder to obtain. One of the handicaps against experimental evaluation of H-bond energies
in the gas phase has been the difficulty in accurate evaluation of equilibrium constants for
formation of complexes involving a pair of neutral species. Advances in methodology, us-
ing Fourier transform IR spectrometry1, promise to alleviate this problem in the future. Pre-
liminary results indicate a close equivalence between the equilibrium constants of forma-
tion in the gas phase and those obtained in inert solvent like CC14.
The hydrides AHn provide a good forum by which to extract the hydrogen bonding char-
acteristics of the A atom, both as proton donor and acceptor. A nomenclature is introduced
here so as to systematize the presentation. X is used to represent the halide atoms F, Cl, and
so on so the hydrogen halides are referred to as HX in the general case. H2Y corresponds
to H2O, H2S, and so on while NH3 and its congeners in lower rows of the periodic 54 are
represented by ZH3.
This chapter is organized by system type. The complexes pairing HX with ZH3 are
simplest in that it is obvious, due to their differences in acidity, which molecule will act as
proton donor and which as acceptor. There is little ambiguity about the geometry of the
H-bond since HX has a single proton and ZH3 only one lone pair. When HX is paired with
YH2, there are two lone pairs on the latter so guessing the relative orientaion becomes
less trivial. The other heterogeneous pairing, between YH2 and ZH3, is most complicated
in that the acidities can be similar enough that one could imagine cither molecule acting
as the donor in certain situations. There are also a fair number of possibilities in terms of
numbers of protons and/or lone pairs so that the nature of the geometry is not intuitively
obvious.

52
Geometries and Energetics 53

Following the foregoing discussion of heterogeneous pairs, the homogeneous pairs are
considered wherein both molecules of the dimer are the same, or at least of the same type,
for example H2Y. The pairing of two HX molecules forces one to act as proton acceptor,
despite the poor basicity of the halide atoms. While the X—H . . . X arrangement will clearly
tend toward 180°, it is not entirely clear from first principles how the acceptor molecule will
align itself. That is, the dipole-dipole interaction would clearly favor a fully linear
X—H . . . X—H even though there are no lone pairs on the X atom directly opposite the H—X
bond. The situation is more complicated in dimers of H2Y, where one could imagine cyclic
and bifurcated arrangements where more than one H-bond could be formed at the same
time. Since NH3 is so weakly acidic, it is not obvious that the ammonia dimer would form
a H-bond at all. This has indeed been a point of debate, as discussed in this chapter.
The discussion extends beyond the simple hydrides and delves into some of the functional
groups as well. The carbonyl oxygen is interesting in that it contains a pair of equivalent lone
pairs, displaced to either side of the C=O bond. But it is not clear whether a proton donor
would prefer to interact with one of these lone pairs or with the electron density directly
opposite the C=O bond. It is also of interest to compare the proton accepting ability of the
carbonyl and hydroxyl oxygens. Combining the C==O with a —OH on a single entity yields
the carboxyl group. Its acidity makes it a potent proton donor, but it is interesting to examine
the proton accepting ability of the C=O group here and on the simpler aldehyde or ketone.
Of interest also are alternate bonding schemes for the nitrogen atom. The C=N double
bond in imines is analogous to the carbonyl oxygen; it is interesting to examine whether the
triple bond in nitriles hampers the nitrogen's ability to accept a proton in a H-bond. But of
perhaps greater interest is the acidity of the C—H group in HC N; the triple bond endows
the C with strong proton donating potential. The chapter concludes with a discussion of the
amide group which combines a C=O and N—H on the same species. Of especial impor-
tance is the competition for H-bonding between the amide and water, due to its relevance
to protein structure. Also discussed is the ability of the "larger" groups like carboxyl and
amide to establish more than one H-bond in simple dimers.
The emphasis in this chapter is on the fundamental properties of these complexes. Of
particular interest are the trends in geometry and energetics: Are there clear patterns in terms
of H-bond strength on going from one type of complex to the next? How are the properties
affected by going down a column in the periodic table? What sort of correlations might ex-
ist between various features of the geometries and the energetics?
The final sections of this chapter discuss two interesting points. One of the limitations
of ab initio methods is the rapid increase in computer resource demands as the size of the
system grows. The accelerating pace of improvements in computer hardware and code de-
velopment has permitted these methods to be extended up to the range of nucleic acid base
pairs. This extended range is demonstrated in this chapter, where it is shown that the com-
puted results are in excellent agreement with experiment. The last section addresses a fun-
damental question dealing with isotopic substitution. Is the D-bond stronger or weaker than
the H-bond? That is, if the normal protium of a H-bond is replaced by its heavier deuterium
isotope, how does this affect the properties of the interaction, especially the energetics?

2.1 XH...ZH3

The simplest type of H-bond would be one in which the proton donor molecule contained
only one hydrogen that could participate and the acceptor only one lone pair capable of in-
54 Hydrogen Bonding

Figure 2.1 Equilibrium geometry of XH...ZH3.

teracting with the bridging hydrogen. These respective criteria are met by hydrogen halides
like HF or HC1 and by molecules of the ammonia family, ZH3, where Z is nitrogen or any
other atom below it in the periodic table. In accord with the aforementioned expectations,
the equilibrium geometry adopted by these complexes has the H—X axis coincident with
the line joining X and Z, placing the bridging hydrogen directly along the H-bond axis as
illustrated in Fig. 2.1. The entire complex belongs to the C3v point group.
The interaction energies calculated for this series where X=F,Cl,Br and Z=N,P,As2 are
listed in Table 2.1. These are uncorrelated values with a moderate sized basis set so should
not be taken as definitive. Nonetheless, the data illustrate the important trends in going
down a column of the periodic table. Greater electronegativity in the proton donor X atom
makes for a more polar X—H bond which creates a stronger electrostatic pull on the lone
pair of the acceptor. The more acidic nature of HX can also act to better release the proton
toward the acceptor, again favoring a stronger H-bond. These expectations are confirmed
by the larger H-bond energies as one reads from right to left in Table 2.1. With regard to the
proton acceptor molecule, ZH3, the enlargement from N to P greatly diminishes the H-bond
energy. This is not surprising as the low electronegativity of P makes it a poor candidate for
proton acceptor. The combination of the high acidity of HF and the strong basicity of NH3
makes the H 3 N ... HF complex the most strongly bound of this series and indeed, of any com-
plex composed of a pair of simple hydrides.
It is a general observation that correlation adds to the H-bond energy in most complexes.
Some representative data are reported in Table 2.2 for pairs of NH3 or PH3 with HF or
HC13,4. Comparison of the binding energies, computed at the SCF and MP2 levels in the
first and second rows, respectively, reveals the correlation-induced strengthening of the H-
bond. This effect is proportionately greater for the weaker complexes containing PH3 where
it can account for as much of a contribution as the entire SCF interaction itself. The distance
between the nonhydrogen atoms, optimized with and without correlation, is listed in the
next two rows of Table 2.2. The correlation-induced strengthening is also reflected in a
small shortening of the H-bond. R(Z..X)MP2, is reduced by anywhere from 0.03 A, relative
to R(Z..X)SCF, for the most strongly bound H 3 N ... HF complex to a maximum of 0.36 A for
the weakest H3P...HC1.

Table 2.1 Electronic contributions to binding


energies, - Eelec, of H-bonds of type H3Z...HX,
calculated using DZP basis set at SCF level. Data in
kcal/mol2.
H3Z HF HC1 HBr
H3N 11.8 7.3 5.7
H3.P 3.7 3.3 1.6
H 3 As 3.6 1.9 1.5
Geometries and Energetics 55

Table 2.2 Calculated binding energies (— Eelec in kcal/mol) and bond lengths (A)
of H-bonded complexes3,4.

H 3 N ... HF H 3 N ... HC1 H3P...HF H3P...HCl

- ESCF 11.8 9.3 4.1 2.1


__ EMP2 6.0 4.4
15.1 11.0
R(Z..X)SCF 2.728 3.297 3.455 4.166
R(Z..X)MP2 2.693 3.144 3.291 3.802
r(HX)SCF,a 0.022 0.023 0.006 0.004
r(HX)MP2 0.028 0.040 0.012 0.011
a
Stretch of HX bond caused by formation of complex.

One typical aspect of the formation of a H-bond is the stretching of the bridging proton
away from the donor atom. This stretch can be calculated as the difference in r(XH) be-
tween the isolated HX molecule and in the complex. The values listed for r(HX) in Table
2.2 indicate some relation between the stretch and the strength of the H-bond. The stretches
calculated here range from 0.004 A for H3P...HC1 up to 0.02 A for the complexes contain-
ing NH3. The effects of correlation are particularly important for accurate assessment of the
degree of stretching; uncorrelated values can underestimate by several-fold.
One can take the H 3 N ... HF system to illustrate the potential effects of basis set super-
position error upon the calculated interaction energies. The results in Table 2.3 are taken
from Latajka and Scheiner5 where basis sets of the general split-valence type were modi-
fied in an effort to minimize this error. The first two entries in the table illustrate that the
superposition error, calculated by the counterpoise technique, is close to 1 kcal/mol at the
SCF level, and a comparable amount is added at the correlated level. The values reported
for Eelec in Table 2.3 refer to the separate SCF and MP2 contributions to the interaction
energy, uncorrected for BSSE, followed by their sum. The last three columns illustrate these
same properties following counterpoise correction. This correction reduces the SCF disso-
ciation energy from 11.8 to 10.7 kcal/mol but has an even more dramatic effect on the MP2
contribution, lowering it by a factor of five, from 1.5 to 0.3 kcal/mol. The combined effect,
illustrated by the last columns, is that counterpoise correction of the full SCF+MP2 inter-
action energy reduces it from 13.3 to 11.0 kcal/mol, a drop of 17%. The next two rows of
Table 2.3 belie a common notion in the literature that superposition error drops as the ba-

Table 2.3 Basis set superposition errors and their effect on interaction energies of H3N...HF.
Data5 in kcal/mol.

-BSSE - Eelec -- ( Eelec -- BSSE)

Basis Set SCF MP2 SCF MP2 SCF + MP2 SCF MP2 SCF + MP2

6-31G** 1.07 1.16 11.79 1.48 13.27 10.72 0.32 11.04


+ 1.21 1.21 11.85 1.71 13.56 10.64 0.51 11.15
2d 1.48 1.59 11.08 2.28 13.37 9.61 0.69 10.30
+-VP S 0.69 1.23 11.31 1.63 12.94 10.62 0.40 11.02
+-VP s (2d) s 0.35 0.79 10.42 1.66 12.08 10.07 0.87 10.94
6-311G** 2.43 1.87 12.27 1.93 14.20 9.83 0.07 9.90
56 Hydrogen Bonding

sis set is enlarged. In fact, the opposite can occur, as the next two rows attest. The + sym-
bol signifies the addition to the standard 6-31G** basis set of a set of diffuse sp-functions
on nonhydrogen centers N and F. These functions produce an increase in the BSSE at both
the SCF and MP2 levels, as do the second set of polarization functions, denoted as "2d."
Standard basis sets are typically constructed by optimizing the exponents within the con-
text of each individual atom. Incorporation of the atoms into a molecule, such as HF, would
change the requirements on the basis set. With this in mind, Latajka and Scheiner reopti-
mized the exponents of the 6-31G** basis set within the context of the individual subunits,
HF and NH35. The results of such an optimization, including the diffuse sp-set on nonhy-
drogen atoms, are reported in the next row of Table 2.3, labeled +VPS. While this approach
reduced the SCF BSSE from 1.07 to 0.69 kcal/mol, very little change was observed in the
correlated BSSE which remained at 1.2 kcal/mol. More dramatic are the improvements
when the same prescription is applied to the doubly polarized basis set. The SCF BSSE for
this + VPs(2d)s basis set is only 0.35 kcal/mol, 1/4 the value for the 2d set. The reduction
at the correlated level is significant but not as marked, dropping from 1.6 kcal/mol for the
unoptimized 2d down to 0.8 for + VPs(2d)s. Such a lowering of the correlated error is im-
portant because of the smallness of this contribution. That is, if one were to compute the in-
teraction energy of this system with the 6-31G** basis set, augmented by a second set of d-
functions, the MP2 contribution would be 2.28 kcal/mol, but fully 70% of this amount
consists of the superposition error. The BSSE contamination is less than 50% if the expo-
nents are reoptimized for the molecules, yielding the +VPs(2d)s basis set. These results
underscore the difficulty in lowering superposition errors at correlated levels, a problem
with which researchers are still wrestling.
The last row of Table 2.3 reveals the profound difficulties in using the 6-311G** basis
set where the triple split of the valence set might normally be expected to be an improve-
ment over 6-31G**. Instead, the SCF BSSE is more than doubled, and an increase of sim-
ilar magnitude occurs in the correlated superposition error. Indeed, the correlation compo-
nent is completely distorted by superposition effects: Essentially all the (1.93 kcal/mol)
stabilization predicted by MP2 with this basis set is due to the artifact of superposition. Re-
moval of this error leaves only a net stabilization of less than 0.1 kcal/mol.
Although they failed to correct their interaction energy for this very substantial error,
Sadlej and Miaskiewicz6 did compute a useful value of the zero-point vibrational energy of
the complex. They found that the complex contains 3.1 kcal/mol more of this type of en-
ergy than the sum of the isolated complexes, using the 6-311G** basis set at the SCF level.
Del Bene7 has computed the binding energy of this complex at the MP4 level, using a
6-31G type basis set, augmented by two sets of d-functions on heavy atoms and two sets of
p on hydrogen. After making the required corrections (but not accounting for superposi-
tion), she obtained a binding enthalpy at 298° K of —10.8 kcal/mol. This value is likely
overestimated by several kcal/mol due to its contamination by BSSE, especially at the cor-
related level.

2.1.1 BSSE
The H 3 N ... HF complex has also furnished a model system for investigation of the spatial
attributes of BSSE and the effects of secondary basis set superposition error. In other words,
the magnitude of the superposition error will depend on how close together the two sub-
units arc and their angular orientations. This issue was considered by Latajka and Scheiner8
who allowed a ghost center to approach a HF molecule. They found that BSSE is negligible
for separations of 3 A or greater. Closer approach leads to a rapid increase in this error,
Geometries and Energetics 57

climbing up to several kcal/mol for chemical bonding distances. This strong distance de-
pendence suggests that geometry optimizations that do not correct for the BSSE are likely
to be in error with respect to equilibrium separation. The approximate midpoint of the FH
bond acts as a sort of central point from the standpoint that the BSSE depends on the dis-
tance of the test center from this point, and is relatively independent of direction. This near
isotropy suggests that BSSE will have a negligible effect upon the angular aspects of a given
H-bonding interaction. The result also suggests that "bond functions," centered not on a nu-
cleus but rather in the space between them, may offer an efficient means of quenching su-
perposition problems in the future.
The same authors investigated secondary BSSEi by determining the effect of ghost func-
tions upon the dipole moment of their prototype molecule, HE. When these functions are
placed on the F side of the molecule, the moment increases but undergoes a decrease when
the functions are situated on the hydrogen side. This trend can most simply be explained on
the basis of partial transfer of electron density from HF to the ghost functions. When the
negative charge associated with this electronic shift is located beyond the F atom, it acts to
enhance the normal dipole of the molecule which is -F—H + . The presence of a negative
charge cloud near the hydrogen will dampen the dipole, causing the decrease noted. While
the two directions are opposite in sign, they are not equal in magnitude. The maximal ef-
fect of ghost functions upon the dipole occurs when they are placed about 1 A from the F
atom, along the H—F axis. The effects of these functions die off as they are drawn away
from the HF molecule, but persist to longer distances on the F side. One may infer that sec-
ondary basis set superposition is highly dependent on angular orientation and can easily in-
fluence the directional character of H-bonding.
On the other hand, the opposite sign of secondary BSSE on the two sides of a molecule
can act to lower the net effect of this phenomenon in the following way. Consider the com-
ing together of two molecules, preparatory to formation of a H-bond. The upper portion of
Fig. 2.2 illustrates the initial approach of HX and ZH3, and includes their dipole moments
as the arrows. As the two monomers approach one another, the orbitals of each can act as
ghost functions for the partner. This effect is indicated by the circle, and the negative charge
which shifts into these orbitals by the negative sign. As a consequence of this negative
charge, the dipole moment of the HX molecule is reduced, indicated by the shorter arrow,
relative to that in the isolated monomer. Analogously, the dipole moment of ZH3 is enhanced
since the negative charge is on the side of the molecule which is already negative. Since the
dipole of one molecule is lowered and the other raised, the net result is that the dipole-
dipole interaction is not much affected by the secondary effect.
This expectation is confirmed by Latajka and Schemer's calculations of H3N...HF,8 but
their results caution against overgeneralization. The first column of data in Table 2.4 reports
the change in the dipole moment of HF resulting from the ghost functions of NH3, placed
just as they would occur in the complex in its equilibrium geometry with R(NF) = 2.66 A
(the experimental value). The negative values are consistent with the effects just described.
One should note, however, that one basis set yields a small increase in the dipole moment
of HF. This discrepancy is likely due to the presence of very diffuse functions which may
permit a good deal of density to be drawn from the H atom, which would tend to increase
the moment of HF. The changes in the NH, moment are more erratic and are only positive
in certain cases. One should conclude that a thumbnail prediction of the effects of secondary
superposition are not always possible.
The next column of Table 2.4 lists the total dipole moment computed for the H 3 N ... HF
complex by each basis set. As mentioned, the formation of the H-bond causes a panoply of
electron density shifts, both within each monomer and some from one to the other. It is in-
58 Hydrogen Bonding

prior to close approach of monomers

effect of ghost functions

Figure 2.2 Approach of XH to ZH3, indicating effects of


ghost functions. Molecular dipole moments are indicated
by arrows. Any electron density which accumulates in
the ghost orbitals is represented by the negative sign.

formative to compare the total moment of the complex with that which would arise if these
charge shifts were prohibited, that is with the sum of the moments of the two isolated
monomers. These differences are reported in the penultimate column of Table 2.4 and show
that the formation of the H-bond leads to an enhancement of the dipole moment, relative to
two unreacting subunits, of about 1 D. The last column of the table addresses the question
of how much of this moment enhancement is due to secondary BSSE. In other words, the

Table 2.4 Secondary basis set superposition errors on the dipole moments of H3N--HF. Data8
all in D.
a
Basis set (HF) (NH3) (FH-NH3) corrb

6-31G** -0.022 +0.078 4.781 0.918 0.862


6-311G** -0.028 +0.072 4.719 0.932 0.888
6-31+G** -0.026 -0.079 4.773 0.860 0.965
dif(2d) 0.018 +0.083 4.453 1.099 0.998
+VPS -0.031 -0.098 4.758 0.838 0.967
+VP s (2d) s -0.021 -0.055 4.661 0.897 0.982
Geometries and Energetics 59

first two columns describe the change in subunit moment resulting not from any genuine
interaction, but only from the presence of the orbitals of the partner. After subtraction of
this artifact from both monomers, we are left with a corrected dipole moment enhancement,
listed in the last column. Whereas some of the corrected moment changes are greater than
their uncorrected counterpart, the opposite is true for a number of basis sets, underscoring
the difficulty in predictions of even the sign of this effect.
Latajka and Scheiner9 further elaborated on secondary superposition by considering the
effects of a set of ghost orbitals along the C3 symmetry axis of NH3, using a variety of dif-
ferent basis sets. It was found that some of these basis sets would increase the calculated
dipole moment and others would diminish this quantity. Moreover, the dependence of these
changes upon the distance of the ghost functions from the N center were rather erratic. In
some cases, the moment change would vary from positive to negative as the functions ap-
proached the NH3 molecule, while others simply pass through a maximum. The effects at
the correlated level were far from negligible. For example, change in the MP2 moment in-
duced by these ghost functions was found in one case to exceed the true MP2 contribution
to the dipole moment. The authors noted, however, that much of this erratic behavior could
be damped by including diffuse functions in the basis set of the NH3.
In contrast to the dipole moment, the polarizability of the NH3 molecule always in-
creases when ghost functions are added9. Such increases can be considered beneficial as
these same basis sets underestimate the polarizability. These increases are larger for com-
ponents along the C3 direction, where the ghost functions are placed and several times
smaller for perpendicular components. Consistent with the dipole trends, the ghost func-
tions have more of an absolute effect upon the SCF segment of the polarizability than the
correlated contribution. But again, despite its smaller value, the MP2 ghost orbital effect on
the polarizability cannot be ignored as it is competitive in magnitude with the genuine MP2
contribution.
The primary BSSE is considered as an energy term while secondary effects are usually
placed within the context of molecular properties such as dipole moment or polarizability.
In order to have some basis of comparison on the same scale, one can consider the interac-
tion between NH3 and an ion like Li+. Any artifact that changes the dipole moment of the
neutral NH3 by an amount Au. will produce an energy increment of

where R represents the intermolecular separation, due to the ion-dipole term of the electro-
static interaction. It was found9 that this estimation of the secondary basis set superposition
can be comparable to, and in many cases larger than, the primary BSSE. Whereas the pri-
mary effect is always negative, the secondary term can take either sign, belying any as-
sumption of cancellation between the two in the general case. Of course, the energetic con-
sequences of error in the dipole moment would be less severe when the partner is a neutral
molecule, as in a typical H-bond, rather than the ionic Li+. Nonetheless, the results provide
a cautionary note in calculations of this type.
Clearly the issue of how to handle secondary superposition error is a thorny one. Szczes-
niak and Scheiner10 have demonstrated that it is possible to avoid the problem with a judi-
cious choice of basis set. Focusing again on the strong interaction between NH3 and Li + ,
they demonstrated that a well-tempered basis set can not only reproduce fairly accurately
the molecular properties of the subunits, but can also yield good total SCF energies. Su-
perposition errors are also quite small. Positioning of ghost orbitals, even as close as 2 A
from the NH 3 , yields a negligible change (<0.05%) in its calculated dipole moment; the
60 Hydrogen Bonding

quadrupole moment is stable to less than 1%. Most impressive were the correlation com-
ponents of the polarizability of NH3 which were altered by basis set extension by less than
1 %. The authors conclude that, at least for a small system such as NH 3 ... Li + , it is possible
to calculate results uncontaminated by superposition errors at either primary or secondary
levels. They advise against any basis set which does not describe the core well (e.g., 6-311G)
as large primary BSSE is likely to ensue. Nor should the basis be chosen with minimiza-
tion of the superposition error as the major criterion since this prescription can lead to an
inflexible set which does not adequately handle second-order properties. On the other hand,
limited computational resources do not always allow the use of a long list of basis functions
for a H-bonded system of interest, so one should be prepared for certain compromises be-
tween accuracy and feasibility.

2.1.2 Substituent Effects


In the simple case where the ZH3 molecule belongs to the C3v point group like NH3, the
equilibrium complex with a linear HX is of the same symmetry. However, deviations can
occur if the three groups bonded to the Z atom differ, as for example, in the case of substi-
tuted amines. The loss of the C3 rotation axis in the amine slightly clouds the precise loca-
tion of the lone pair, and the bridging hydrogen may deviate by several degrees from the
X...Z internuclear axis. Latajka et al. 11 computed the equilibrium geometries of complexes
pairing HBr and HI with mono and dimethylamine. The 0 (NXH) angles may be seen from
the first rows of Table 2.5 to take the proton less than 3° from the H-bond axis in these cases.

Table 2.5 Properties of complexes of HBr and HI with substituted amines, computed with
polarized split-valence basis sets, uncorrected for BSSE. All vaues of E refer to electronic
contribution to binding energy 11 .

MeH 2 N ... HBr Me2HN...HBr Me,N...HBr

0(NBrH), degs 2.5 0.5 0.0


R(N..Br)SCF, A 3.361 3.288 3.049
R(N..Br)MP2, A 2.952 2.961 2.972
r(HBr)SCF, A 0.036 0.048 0.561
r(HBr)MP2, A 0.298 0.392 0.393
- ESCF, kcal/mol 7.65 7.90 17.24
- EMP2, kcal/mol 13.14 17.47 20.13
- EMP3, kcal/mol 10.19 14.56 17.15
- EMP4, kcal/mol 10.22 14.37 17.13

MeH2N...HI Me2HN...HI Me 3 N ... HI

(NIH), degs 1.4 0.2 0.0


R(N..I)SCF, A 3.659 3.572 3.313
R(N .. I) MP2 , A 3.210 3.212 3.219
r(HI)SCF, A 0.030 0.043 0.663
r(HI)MP2, A 0.497 0.512 0.534
- ESCF kcal/mol 5.08 5.40 21.83
- EMP2, kcal/mol 16.89 22.76 25.91
- EMp3, kcal/mol 13.73 19.45 22.75
- KMP4, kcal/mol 13.19 18.89 21.19
Geometries and Energetics 61

The next several columns illustrate the effects of progressive methyl substitution upon
the geometric and energetic aspects of the complex formed by an amine with hydrogen
halides. As more hydrogens are replaced by methyl groups, the basicity of the amine in-
creases, making it an improved proton acceptor. This change is reflected in a number of
trends. The H-bond contracts, revealed by the reduction in R(N .. X) at the SCF level, as one
moves across the appropriate row of Table 2.5. A big jump in this quantity occurs between
the di- and trimethyl amine due to a fundamental change that occurs. The greater acidity
of the trimethylamine is sufficient to cause the bridging proton to transfer across from the
HBr or HI molecule, leaving the complex as an ion pair N H + . . . - X . This transition is most
obvious in the HX stretching parameter r(HX)SCF and is also responsible for the large in-
crease in SCF binding energy when the third methyl group is added. One way of concep-
tualizing the latter increase is the added electrostatic interaction between the two ions once
the proton has transferred across to the amine. (Backskay and Craw have recently demon-
strated that the basicity of the trimethylamine is sufficient to extract the proton from
HC112.)
The MPn data in Table 2.5 underscore the difference in fundamental character that is as-
sociated with incorporating correlation into the treatment of these systems. As will be dis-
cussed in greater detail in chapters to follow, the transition from neutral to ion pair that oc-
curs with progressively higher methyl substitution of the amine is a smooth one when
described at the correlated level, as opposed to the sharp transition from one type of com-
plex to the other without correlation included. It is for this reason that the MP2 intermo-
lecular separations do not obey the same pattern as the SCF distances. Nonetheless, one can
still discern that the increased basicity of the methylated amine yields a greater interaction
energy, albeit between ions rather than between neutral molecules. The much larger corre-
lated versus SCF interaction energies are due in large measure to the aforementioned con-
version to an ion pair and the additional stabilization that arises from the ion-ion forces.

2.2 XH...YH2

When the proton acceptor molecule contains an atom like O or S, a second lone pair is pre-
sent that makes prediction of the equilibrium geometry less obvious. If one assumes an sp3
hybridization, the two lone pairs are disposed as in Fig. 2.3A, which would yield an angle
P, between the HYH bisector and the X..Y axis, in the vicinity of 125°. The alternate type
of hybridization, sp2, leaves one of the lone pairs in a p-orbital, oriented 90° from the other
lone pair. This arrangement would lead to a 180° angle. Geometry B is also favored by
certain electrostatic arguments. Specifically, it would permit the dipole moment of YHL,
collinear with the HYH bisector, to align itself with the dipole of the X—H molecule.

Figure 2.3 Dispositions of molecules and lone pairs in HX + H2Y.


62 Hydrogen Bonding

There appears to be a delicate balance between the factors favoring configurations A and
B. For example, systematic analysis of a large set of diffraction data indicates the tendencies
toward structures A and B are very nearly equal in the solid state13. It should be understood
that while hybridization is indeed a useful concept, most molecules cannot be categorized
as simply sp2 or sp3. If one insists on that language, an analysis of the wave function will
usually yield nonintegral values, for example sp2.1 or sp2.77. It is thus an oversimplification
to consider the electronic configuration of YH2 as simply one or the other of those pictured
in Fig. 2.3. The electron cloud would more accurately be visualized as a "smear" extend-
ing from the region above the Y atom to below, with fluctuations of density as one goes
around. Nor should it be thought that the geometry is controlled solely by factors of elec-
tron density of the acceptor molecule. An alternate description (see later) places more em-
phasis upon the electrostatic factors.
Another factor has to do with the linearity of the H-bond itself. The C2v symmetry of
Fig. 2.3B would place the bridging hydrogen directly along the X...Y axis, but it is plain that
the forces above and below this axis in Fig. 2.3A are not equal. It should therefore not be
surprising to see the proton deviate from the H-bond axis in such cases, with nonzero val-
ues of a. It is likely the proton would be above the line as drawn, based on the direction of
the YH2 dipole moment in this configuration.

2.2.1 Comparative Aspects


Hinchliffe published in 1984 a comprehensive theoretical comparison of hydrides of the
H 2 Y-HX type14. The calculations reported in Table 2.6 were all undertaken at the SCF
level and assumed Cs geometries, with a fully linear Y-HX arrangement. The patterns are
consistent with the aforementioned data for H3Z..HX complexes. The strength of the H-
bond diminishes with lowered electronegativity of the donor atom F > Cl > Br. H2O forms
the strongest H-bonds; the distinction between H2S and H2Se is a small one. Any of these
H2Y proton acceptors form a weaker complex than the corresponding H3Z molecule of the
same row of the periodic table. Hinchliffe also found that the complexes involving H2O
were of type B while those incorporating H2S or H2Se were of type A, with angles of
about 110°.
An additional set of SCF data emerged from calculations by Hannachi et al.15 who paired
water with each of the hydrogen halides listed in Table 2.7, optimizing the geometry using
a pseudopotential basis function of polarized split valence quality. The data echo the above
trends of a weaker H-bond as the X atom of HX comes from a lower row of the periodic
table. This progressive weakening is reflected in smaller stretches of the H—X bond. It is
also worth noting that the small nonlinearity of the H-bond and the preferred angle of the
proton acceptor water molecule are virtually independent of the nature of the HX molecule.

Table 2.6 Electronic contributions to binding energies (— Eelec in


kcal/mol) of H-bonds of type H2Y...HX, calculated using DZP
basis set at SCF level14.

H2Y HF HC1 HBr


H2O 9.0 5.4 4.3
H2S 3.5 2.4 1.6
H2Se 3.4 2.0 1.1
Geometries and Energetics 63

Table 2.7 Energetic and geometric aspects of complexes of water with HX calculated
at SCF level15.
H2O...HF H2O...HC1 H 2 O ... HBr H 2 O ... HI

- Ee|ec, kcal/mol 8.2 4.9 4.1 2.5


R(O .. X), A 2.702 3.268 3.496 3.830
Ar(XH), A 0.012 0.013 0.010 0.006
a, degs 3.8 2.4 2.5 2.6
, degs 135.9 140.1 139.4 139.0

Correlated data for a set of four of these complexes are listed along with SCF values in
Table 2.816-18. These data were collected using basis sets of double-valence quality, and
augmented with two sets of polarization functions on all atoms. The SCF energies agree
fairly well with those in Table 2.6 and reinforce the same trends. Correlation acts to
strengthen all interactions. This effect is proportionately larger as the number of second-
row atoms is increased. The energetic data seem to parallel experimental results pretty well.
For example, the MP2 electronic binding energy of H2O...HF of 9.6 kcal/mol is only slightly
smaller than an experimental determination of 10.2 kcal/mol for De, based upon absolute
intensities of rotational transitions19. If one extrapolates an enthalpy of dissociation for the
H2O...HF complex using the vibrational, rotational, and translational corrections derived
above, a calculated value of — H298 = 6.9 kcal/mol is obtained, within the uncertainty of
the experimental estimate of 6.2 ± 1 kcal/mol. A recent calculation of the H 2 S ... HF com-
plex with a very large polarized basis set comprising 169 functions, obtained a binding en-
ergy — Eelec of 4.7 kcal/mol at the MP2 level, with counterpoise correction20.
MP2-optimized intermolecular separations R(Y..X) are reduced relative to SCF dis-
tances, consistent with the strengthening role of correlation. The separations optimized at
the MP2 level furnish fairly good reproductions of experimental estimates, typically too
short by 0.005 A or better. It is intriguing to note the similarity between entries for
H2O...HC1 and H2S...HF, both of which contain one first-row and one second-row atom.

Table 2.8 Calculated energetic (- Eelec) and geometric aspects of H-bonded complexes. Data
uncorrected for BSSE16J7.
H 2 O ... HF H2O...HC1 H 2 S ... HF H2S...HC1

- ESCF, kcal/mol 7.8 4.2 3.9 2.2


- EMP2, kcal/mol 9.6 6.6 6.3 5.0
R(Y..X)SCF A 2.71 3.37 3.36 4.09
R(Y..X)MP2, A 2.65 3.19 3.20 3.75
R(Y .. X) expt,a , A 2.66 3.21 3.25 3.81
r(HX)SCF, A 0.012 0.009 0.007 0.005
r(HX)MP2, A 0.017 0.015 0.011 0.011
SCF
degs 140 140 100 101
MP2
, degs 129 130 98 92
SCF
, degs 3.1 2.8 1.3 1.5
MP2
, degs 4.5 0.9 -0.3 1.2
a
See Reference 18.
64 Hydrogen Bonding

The next two rows reveal that correlation also enhances the stretch that occurs within the
X—H bond upon formation of the H-bond.

2.2.2 Angular Features


The values of listed in Table 2.8 illustrate that these complexes all adopt a "pyramidal"
geometry, closer to A shown earlier than to "planar" configuration B. This is particularly
true of the complexes containing H2S where the angles approach 90°. Indeed, early mi-
crowave spectra of H2S...HF21,22 suggested the proton acceptor molecule was oriented
nearly perpendicular to the donor; a similar result was obtained for H2S...HC118. The nearly
perpendicular arrangement of complexes containing H2S, as compared to H2O, has been
confirmed recently at higher levels of theory. MP2/6-311+ +G(d,p) optimizations of
H2Y..HF found angles of 140° and 112°, for Y=O and S, respectively23.
One means of rationalizing the trends in the angle is via electrostatic arguments24. As
mentioned earlier, the planar geometry B, with = 180°, is favored by dipole-dipole inter-
actions between HX and YH2. This preference is illustrated in the left part of Fig. 2.4. On
the other hand, the two units are certainly close enough together that the quadrupole mo-
ment of YH2 can play a role as well. The bonding pattern of YH2 leads one to expect a neg-
ative element in the direction perpendicular to the molecule. The attraction between the neg-
ative charges of this quadrupole tensor element and the positive end of the HX dipole will
tend toward a 90° value for , illustrated by the right part of Fig. 2.4. The end result can be
considered a compromise between these two trends toward large and small angles. The fact
that the MP2 values of are smaller may be attributed in part to the correlation-induced re-
duction of the dipole moment of water, which would diminish the pull toward large angle.
Of course, this is an oversimplification and a thorough analysis of the reasons for the di-
rections would have to take into account forces other than electrostatic. Nonetheless, in-
sights gained from Coulombic concepts are extremely valuable, and can be superior to pre-
dictions based on detailed analysis of the wave function23.
One can also think more quantitatively about the energetic difference between pyramidal
and planar geometries. The first column of Table 2.9 illustrates that the fully planar C2v
structure of H 2 O ... HF is higher in energy than the optimized pyramidal geometry by only
0.1 kcal/mol at the SCF level, increasing to 0.5 kcal/mol with MP216,17 (an experimental
estimate is 0.4 kcal/mol25). The latter correlated barrier was later confirmed with a larger
basis set26. Similar values are found for H2O...HC1. Much higher energy differences arise
for complexes where H2O is replaced by H2S. These increases are consistent with the pref-
erence toward much smaller values of p.

Figure 2.4 Interactions between multipole moments. Dipole moments are indicated by arrows,
and an element of the quadrupole tensor by the double lobe.
Geometries and Energetics 65

Table 2.9 Energy required to bend each complex from its equilibrium pyramidal structure into
a planar C2v arrangement. Values in kcal/mol16,17.

H 2 O ... HF H20...HC1 H2S...HF H S...HC1


SCF 0.13 0.15 2.96 1.59
MP2 0.49 0.33 3.65 2.52

In cases where the pyramidal equilibrium geometry differs little in energy from the pla-
nar configuration, what are the consequences for experimental observation? This question
may be addressed by considering the potential function for bending. In the case where the
planar structure is most stable, there is little question but that experimental observations
would confirm this. The situation is less clear when the planar geometry is less stable than
a pyramidal structure but only marginally so. For example, Legon et al. concluded from
their gas-phase rotational spectroscopic measurements that the equilibrium geometry of
H2O...HX, X=Br,Cl was either planar or, if pyramidal, that the inversion barrier was very
low27,28.
Figure 2.5 illustrates a number of different cases that are possible. The potential energy
surface is illustrated as a double-well potential with respect to the flipping of the YH2 mol-
ecule about the planar position. In the case where the barrier to this flipping is high enough
that the lowest vibrational level occurs well below the top of the barrier, the square of the
wave function resembles that depicted by the lowest function in the figure. There are two

Figure 2.5 Vibrational wave functions corresponding to


each of several energy levels in a double-well potential,
with respect to "flipping angle" p.
66 Hydrogen Bonding

maxima in the probability density, each associated with a pyramidal structure. On the other
extreme is the case where the barrier is so low that the ground vibrational level is well above
its top. The highest of the three probability density functions in Fig. 2.5 illustrates that an
experimental measurement would indicate a planar structure despite the appearance of a
shallow maximum in the potential energy. The situation is most ambiguous when the vi-
brational level is below the barrier top, but only slightly so. The probability density func-
tion retains two maxima but these are close to the center and poorly defined. The function
may perhaps be better described as a single flat maximum extending on either side of the
planar structure, = 180°. The atomic motions would correspond to large-amplitude devi-
ations from a planar structure. It thus appears that the experimental elucidation of the pre-
cise nature of the potential energy function for this sort of wagging motion may not be a
trivial task. The simple fact that the vibrational level occurs below the maximum in the po-
tential function does not insure that one can detect a double-well potential.
The preceding discussion has been simplified by assuming that the wagging of the YH2
molecule is separable from other wags and stretches; a more thorough analysis would not
make this assumption. The small energy barriers computed for the H2O...HX systems cor-
respond to the cases where the lowest vibrational level is near the top of the barrier16. The
higher barriers that occur when H2O is replaced by H2S allows unambiguous determination
of a pyramidal structure.
The last two rows of Table 2.8 indicate that there is only a small amount of nonlinearity
in the equilibrium geometries of the H2Y...HX H-bonds. The deviations from fully linear
arrangements are typically less than 5°. In most cases, the bridging hydrogen lies "above"
the Y...X axis, in the sense of configuration A above, as indicated by the generally positive
values of a. This trend is consistent with the attempt by the proton to better align itself with
the dipole moment of the H2Y, when its hydrogens are bent down.
Szczesniak et al.16 have pointed out an interesting relationship between the stretch of the
hydrogen away from the X atom and the energetics of the interaction. They have shown that
Ar is very nearly linear, over a range of HX stretches, with respect to the contribution made
by electron correlation to the H-bond. The authors assumed the latter is dominated by dis-
persion, and so concluded that the stretch of the H—X bond causes an increase in the mol-
ecule's polarizability. They hence infer that a molecule whose polarizability is sensitive to
the X—H bond length can enhance its ability to form a H-bond by permitting a greater
stretch of the bond upon complexation.

2.2.3 Alternate Complexes and Geometries


The greater acidity of HX than of H2Y leads to the normal supposition that the former mol-
ecule will act as the donor in any interaction between the two molecules. Szczesniak and
Scheiner17 tested this presumption in the complex between HF and H2S. A structure in
which the normal bonding pattern is reversed, namely H2S donates a proton to HF, was in-
deed found to be a minimum on the potential energy surface. However, it was found to lie
some 3 kcal/mol higher in energy than FH...SH2. Novoa26 later explored the same question
for the HF,HOH pair, using a high quality basis set. They concluded that the potential en-
ergy surface probably does not contain a minimum corresponding to the "reverse" complex
wherein HF acts as the proton acceptor, although the question was not answered definitively
as they did note a "plateau" in that region of the surface.
Substitution of the hydrogen atoms by alkyl groups appears to exert a minimal impact
on the angular aspects of the calculated geometries. As an example, Amos et al.29 optimized
Geometries and Energetics 67

the geometry of the dimethylsubstituted Me2O...HCl and found difficulty in determining


the equilibrium value of since the energy profile for bending away from 180° was ex-
tremely flat. Calculations by Bouteiller et al.30 found only very minor differences in the
equilibrium geometries of H,O...HF and Me2O...HF. Experimentally, the geometries of
H2O...HC1 and MeHO...HCl are quite similar as well31.
Hannachi et al.32 calculated the relative stability of base..HX versus base...XH for com-
plexes in which water is the base. They refer to the former geometry as "H-bonded" and the
latter as "van der Waals," also known as "anti-H-bonding." In the case of the complex with
HC1, they only find one minimum on the potential energy surface, corresponding to the H-
bonded H2O-HC1. However, both types of complex were identified as minima for HBr and
HI. The first row of Table 2.10 illustrates the closer approach of the two subunits in the anti-
H-bonding arrangments. The energetics indicate that the H-bonding structure is greatly pre-
ferred for HBr; this preference is much weaker for HI, such that both structures might be
observed experimentally. The authors also monitored the shifts in electron density in the
monomers which accompanied the formation of the two types of complexes. For either H-
bonding or anti-H-bonding, the lone pairs of the oxygen atom suffer a loss of density, al-
beit stronger in the former case. Unlike the case of a H-bond, density is shifted from I to-
ward H in the HI subunit of H,O..IH. This shift acts to induce a dipole moment in IH which
aligns favorably with the moment of H2O. In fact, two separate geometries have been ob-
served for the complex between HI and H2O by matrix isolation IR spectroscopy, one H-
bonded and the other not33, although the details of the structure were not established.
The basis set superposition errors of the H2Y..HX complexes are comparable to those
observed for H3Z..HX, listed earlier in Table 2.3 for a variety of basis sets5. One interest-
ing difference is that whereas the MP2 contribution to the binding energy of H 3 N .. HF is at-
tractive, albeit by less than 1 kcal, the correlation contribution in H2O..HF is close to zero,
with some basis sets yielding a small repulsive contribution after primary BSSE is ac-
counted for.

2.2.4 Energy Components


Backskay et al.34 have partitioned their total interaction energies into components, using a
scheme similar to Kitaura-Morokuma, but with some modifications. Their results are dis-
played in Table 2.11 where it may be seen that the electrostatic component is the dominant
attractive term in all cases. ES is particularly large for the two bases with a first-row atom,
H3N and H2O. It is this pair of complexes which are most strongly bound, as indicated by
the last column of Table 2.11. Of comparable magnitude to ES, but of opposite sign, is the
exchange repulsion. As the only repulsive element, EX keeps the two subunits of each com-
plex from collapsing together. The polarization and charge transfer energies are particularly

Table 2.10 Comparison of H2O..HX and H2O..XH32. E refers to electronic contribution


only.
H2O..HBr H2O..BrH H2O..HI H2O..IH

R(O .. X), A 3.496 3.228 3.830 3.192


- ESCI , kcal/mol 4.10 0.22 2.47 1.71
- EMP2, kcal/mol 5.04 0.59 3.35 1.99
68 Hydrogen Bonding

Table 2.1 I Components of interaction energy of complexes involving HC1, calculated at


experimental geometries. Values in kcal/mol34.

ES EX POL CT Totala

H 3 N ... HC1 -17.36 17.67 -2.96 -3.88 -6.65


H2O...HC1 -9.21 6.94 -1.44 .-1.62 -5.35
H3P...HC1 -4.33 5.16 -0.69 -1.47 -1.37
H2S...HC1 -4.49 4.50 -1.35 -1.08 -2.45
a
Total — Eelec includes also "unassigned" contribution, so is not equal to sum of other terms.

large for H3N...HC1, and vary between —0.7 and —1.6 kcal/mol for the other complexes.
The authors also examined how the various terms behave as the two subunits in H 3 N ... HC1
are pulled apart. The electrostatic attraction dominates the interaction at long distances.
Slightly closer approach brings an exponential rise of the exchange repulsion. It is not un-
til intermolecular separations of about 3.5 A or less that the polarization or charge transfer
energies become significant.
Much accumulated data like that above have provided evidence that electrostatics is a
primary force which orients the bridging proton of a H-bond along the intermolecular axis.
While an oversimplification, this force may be thought of as composed of the alignment of
the dipole moments of the donor and acceptor molecules. Let us now consider a situation
where this force is steadily decreased. Table 2.12 reports data which illustrate that as the
halide atom of the HX molecule advances to lower rows of the periodic table, the molec-
ular dipole moment decreases32. Hence, as one changes this molecule from HC1, to HBr,
to HI, one can expect that the electrostatic drive toward a base...HX orientation will be sim-
ilarly weakened. At the same time, the molecules with the larger halide also are most po-
larizable, particularly along their molecular axis, as indicated by the values of zz in Table
2.12. This greater longitudinal polarizability leads the interaction between the two mol-
ecules to contain progressively greater amounts of induction and dispersion energy. Both
of the latter forces become rapidly more attractive as the molecules approach one another.
In fact, by approaching in an "anti-H-bonding" orientation, that is, heavy atom first as in
base"XH, the base can more closely approach the more polarizable halogen end of the HX
molecule. And as the X atom changes from Cl to Br to I, there is more to gain by ap-
proaching this way, and less electrostatic energy to lose since the HX dipole moment be-
comes so small.

Table 2.12 Experimental values of dipole moment,


, and dipole polarizabilities, a, of hydrogen halide
HX molecules32.
HC1 HBr HI

.D 1.094 0.819 0.447


a , aua 21.1 28.5 44.4
xx, au 19.6 22.3 32.9
a
The molecular axis is defined as the Z-axis.
Geometries and Energetics 69

2.3 HYH...ZH3

The importance of hydrogen bonds between amino and hydroxy groups has been amplified
in recent years by the finding that such interactions can guide the formation of well-ordered
supramolecular structures35,36. Because the ZH3 molecules are stronger bases than YH2,
one expects the former to act as proton acceptor in complexes with the latter. This has in-
deed been found to be the case in the complex between water and ammonia, the most stud-
ied of systems of this type. One of the two hydrogens of the YH2 is used to bridge the two
molecules in a classic H-bond that is nearly linear, incorporating the single lone pair of the
ZH3 molecule as illustrated in Fig. 2.6.
The geometrical aspects of the HOH...NH3 complex are reported in Table 2.13 at vari-
ous levels of theory37. (The +VPS basis set is related to 6-31+G**, except that orbital ex-
ponents have been reoptimized so as to reduce BSSE.) The intermolecular distance elon-
gates somewhat as the basis set is enlarged but diminishes to 2.94 A upon inclusion of
electron correlation. This distance is just slightly shorter than estimated by microwave/far-
IR data which lead to a value of 2.97-2.99 A38,39. The covalent bond to the bridging pro-
ton stretches by around 0.01 A upon formation of the H-bond, less at the SCF level, more
at MP2. This hydrogen lies within about 5° of the H-bond axis. The last two rows of Table
2.13 indicate that the NH3 molecule turns its lone pair up toward the connecting hydrogen
since the (ONHc) angles are larger by some 15° than (ONHt). Another study40 made the in-
teresting observation that the structure depicted is a true minimum in the MP2 potential en-
ergy surface, but is slightly less stable than that in which the NH3 is rotated 60° around the
H-bond axis when the surface is uncorrelated. That is, the "staggered" geometry is the min-
imum in the correlated surface but an "eclipsed" structure is preferred at the SCF level. The
energy differences in either case are exceedingly small, so the rotational barrier can be con-
sidered negligible. This finding is consistent with experimental estimates of a barrier of only
0.03 kcal/mol38. There is no evidence of a minimum for which the roles of proton donor
and acceptor are reversed, such as H2NH...OH2.
Calculated values for the binding energy of HOH...NH3 are listed for several basis sets
in Table 2.1437. SCF values of — E are in the 4.6-5.6 kcal/mol range. Correlation adds to
this amount, bringing the binding energy up near 6 kcal/mol. Del Bene41 performed a sim-
ilar set of calculations, but all values were somewhat higher due to BSSE which was left
uncorrected. Her results were nonetheless valuable in that they illustrated that MP2 values
were nearly identical to full MP4 interaction energies. The best value for Eelec seems to
be about —5.5 kcal/mol at this time. A lower-bound for — H of 2.9 kcal/mol comes from
molecular-beam electric-resonance optothermal spectroscopy42. Del Bene7 has applied a
more flexible basis set to this complex, with MP4 consideration of correlation. At the

Figure 2.6 Geometry of HYH...ZH3.


70 Hydrogen Bonding

Table 2.13 Calculated geometry of HOH...NFL complex37.


SCF MP2

6-31G** +VPS + VPs(2d)s +VPM

R(O..N), A 3.050 3.074 3.096 2.942


r(OH), A 0.008 0.008 0.007 0.013
a, degs 2.1 3.7 4.6 4.8
(ONHt), degs 100.5 100.8 101.1 101.8
(ONHc), degs 116.2 116.0 116.1 117.7

MP4/6-31 +G(2d,2p) level, a binding enthalpy at 298 K was calculated to be -4.7 kcal/mol,
which included vibrational corrections, and so forth.
The HOH...NH3 complex served as a recent test for symmetry-adapted perturbation the-
ory (SAPT). Basing their work on earlier formalism43, which was further elaborated, Lan-
glet et al.44 observed that a pure perturbation approach yielded an intermolecular separa-
tion that was somewhat too long, and underestimated the binding strength of the complex.
Better correlation with experimental quantities, as well as with other accurate computations,
is obtained by a "hybrid" approach, wherein the dispersion energy, computed by SAPT, is
added to the (counterpoise corrected) SCF portion of the interaction energy. This conclu-
sion was found to apply not only to HOH...NH3, but also to the homodimers of HF, H2O,
and NH3.
The complex between H2O and H3P is barely bound at all, with H298 only —0.8
kcal/mol41. In fact, this value might become positive were counterpoise corrections made
to the binding energy. Attempts at identifying another minimum on the surface in which
H2O and H3Z reverse their roles to proton acceptor and donor, respectively, failed for both
Z=N and P, which suggested there is no such local minimum.
A pairing of H2S with NH3 did yield a minimum, containing a linear H-bond with H2S
as donor7. This geometry conforms to molecular beam electric resonance data45 which
yields an intermolecular R(S .. N) of 3.639 A. This H-bond is somewhat shorter than a prior
ab initio computation of 3.79 A46. The HSH...NH3 complex is bound at the MP4/6-
31 +G(2d,2p) level, relative to the isolated monomers, by 3.6 kcal/mol. Again, this value
would likely be diminished by inclusion of zero-point energy and superposition error cor-
rections. Recent electric-resonance optothermal spectroscopic measurements47 place an up-
per bound of 2.8 kcal/mol on the binding energy of HSH...NH3, including zero-point vi-
brational corrections. This complex has a slightly smaller energy barrier to proton exchange
as compared to HOH...NH3, 1.5 versus 2.0 kcal/mol, which is taken as evidence of a less
directed H-bond in the former.

Table 2.14 Electronic contribution to binding energy


of HOH...NH3 complex. Data in kcal/mol37.
6-31G** +VPS +VPs(2d)s

- ESCF 5.61 5.10 4.59


_ E MP2 6.35 5.92 5.72
Geometries and Energetics 71

2.3.1 Substituents
Alkylating ZH3 might be expected to make this molecule a better proton acceptor, as the
larger substituents can better delocalize any charge accumulation. Calculated data are pre-
sented in Table 2.15 for the complex of methylamine with water48. Comparison with the
data in Tables 2.13 and 2.14 suggests that the methyl group on the N does in fact enhance
the H-bond energy by a small amount, along with a shortening of the intermolecular dis-
tance. Correlation enhances this binding, as in most other complexes of this type, but the
amount cannot be easily discerned because Zheng and Merz did not remove their BSSE,
which is apt to be rather appreciable with this basis set, in particular at the MP2 level.
Nonetheless, the last row indicates that AG is probably positive for the binding reaction,
due to the large negative AS. The sensitivity to basis set is evident from a comparison with
prior calculations using the 6-31G basis set49. The deletion of the polarization functions
on O and N, reduces the intermolecular separation by 0.1 A and increases the SCF inter-
action energy by 2.2 kcal/mol. A similar sort of analysis, but this time alkylating the hy-
droxyl group to form CH3OH...NH350, again confirmed very little influence of the methyl
group.
Another type of substitution places an aromatic group on the proton-donor oxygen atom.
SCF/6-31G** optimization of the complex between phenol and ammonia51 yields a
R(O..N) H-bond length of 2.891 A, somewhat shorter than the value of 3.050 A computed
at the same level of theory for HOH...NH337. The enhanced proton-donating capability pro-
vided by the aromatic group is verified by the energetics of binding. Eelec is — 8.5 kcal/mol
for this complex, as compared to —5.6 kcal/mol for HOH..NH3. After adding in MP2 cor-
relation, zero-point vibrational energies, and correcting for BSSE, the binding energy for
the phenol-ammonia complex is computed to be Do = 7.0 kcal/mol51. MP2-level correla-
tion is responsible for a contraction of the H-bond by 0.11 A52.

2.4 XH...XH

The structure of the complex between a pair of hydrogen halide molecules is depicted in
Fig. 2.7 where three lone electron pairs are placed on the proton-accepting molecule. In the
classical case of sp3 hybridization, one might expect an angle of some 109°. a measures
the nonlinearity of the H-bond as in the above cases. A nonzero value of a might be ex-
pected based on the direction of the dipole moment of the acceptor molecule.

Table 2.15 Optimized H-bond length and energetics


of complex between HOH and CH3NH2, calculated
with 6-31G* basis set48. Energetics not corrected
for BSSE.
SCF MP2

R(O..N), A 3.015 2.902


Eelec, kcal/mol -6.5 -9.1
H, kcal/mol -5.3 -7.8
S, cal mor--1' deg 1 -26.7
G, kcal/mol 2.7 0.1
72 Hydrogen Bonding

Figure 2.7 Dispositions of molecules and lone pairs in


HX dimer.

The principal features calculated for the geometry of the HF dimer are reported in Table
2.16 from which it may be seen that the angles are predicted reasonably well, even with
rather small basis sets and without correlation53-58. The equilibrium interfluorine distance
is approximately 2.76 A. The bridging hydrogen lies within at least 10° of the F..F axis,
probably more like 5°. The angle made by the proton acceptor molecule is somewhat more
sensitive to details of the calculation but appears to fall within the 112°-120° range. These
predictions conform to the experimental measurements reported in the last row of Table
2.16.
The binding energies reported in the last column of data in Table 2.16 indicate some sen-
sitivity to the type of basis set and method of computing correlation. A better feel for these
trends may be obtained from the data in Table 2.17, calculated by Del Bene59 for a range
of different basis sets, all within the M011er-Plesset scheme of correlation. Although there
is some degree of erratic behavior due to the failure to remove BSSE, there are some clear
patterns in evidence nonetheless. Addition of a single set of diffuse .sp-functions to F yields
a marked reduction in the interaction energy, probably due to the reduction in superposition
error. In contrast, addition of a second set of d- or p- functions has very little effect on E.
Enlarging from double-valence to triple valence in the core lowers the interaction energy,
again likely due to reduced BSSE. In all cases, correlation enhances the binding energy,
with MP2 being a satisfactory substitute for much more expensive full MP4. The best esti-
mate of Eelec achieved is —4.7 kcal/mol which would likely be reduced by incorporation
of a counterpoise correction.
The level of theory was raised once again in a recent set of calculations wherein a dif-
ferent type of basis set was used, in conjunction with coupled-cluster means of considering
electron correlation60. Table 2.18 lists the geometrical parameters optimized for the HF
dimer, with and without counterpoise corrections. These results were obtained with a very
large correlation-consistent (cc) set: [6s5p4d3f2g/5s4p3d2f]. The data indicate that MP4 and
coupled-cluster singles and doubles (with triples approximation) yield very similar results.
The interfluorine distance is some 2.73 A at either level. However, counterpoise correction
does lengthen the equilibrium value to something closer to 2.75 A. The angular aspects of
the equilibrium geometry are consistent from one type of correlation to the next. The pro-
ton acceptor is rotated some 110° from the H-bond axis and the nonlinearity within this bond
is just under 7°. The best theoretical estimate of the binding energy De is 4.5 kcal/mol, with
about 3.7 kcal/mol arising from the SCF level alone. This result is in excellent accord with
an experimental estimate of 4.6 kcal/mol in the gas phase, based on absolute infrared line
strengths61, and another estimate of 4.5 kcal/mol62. The authors conclude that the MP2
method offers a computationally efficient means to obtain the more accurate results which
require much more computationally demanding approaches. The above computed results
were confirmed to good accuracy by another correlated study that made use of basis sets
such as triple- plus double polarization functions and a set of higher angular momentum
functions 63 .
Table 2.16 Geometrical and energetic aspects of (HF)2 calculated at various levels.

R(F..F) (A) (degs) (degs) - Eelec (kcal/mol) Reference

SCF
4-31G 2.687 8.1 124.1 8.0 [53]
6-31+G* 2.788 7.9 117 4.7 [54|
6-311G* 2.81 9 115 4.7 |53]
+VPs(2d)s 2.82 7.2 116.5 3.7a [551
[11s7p2d/6slpl 2.83 6.0 123.2 3.8 [53J

Correlated
CC/TZP 2.768 6.4 120.1 4.6 [56J
MP2/6-31+G* 2.762 6.9 — 5.7 [571
MP2/6-311 + +G(2d2p) 2.759 5.5 112 5.0 [57]
expt 2.72 10±6 117±6 [58J
a
Corrected for BSSE
74 Hydrogen Bonding

Table 2.17 Calculated binding energies of HF dimer (— Eelec), in kcal/mol59.

Basis set SCF MP2 MP3 MP4

6-31G(d,p) 5.97 7.45 7.02 7.34


6-31+G(d,p) 3.98 4.69 4.62 4.71
6-31G(2d,p) 5.87 7.58 7.13 7.52
6-31 + G(2d,2p) 3.75 4.61 4.61 4.66
6-311G(d,p) 5.06 6.22 5.82 6.15
6-311+G(2d,2p) 3.71 4.66 4.64 4.71

The data in Table 2.19 indicate that the H-bond between a pair of HC1 molecules is some-
what weaker than in (HF)254,55,64-68. Best theoretical estimates of the binding energy are
less than 2 kcal/mol; the gas-phase estimate is 2.3 kcal/mol61. Part of the discrepancy is
likely due to the fact that dispersion is very important to this interaction. The latter phe-
nomenon requires particularly flexible basis sets for its saturation. The H-bond is likely less
linear than in (HF)2, with some estimates for a above 10°. Notable also is the smaller value
of p in (HC1)2 wherein the two HC1 molecules are nearly perpendicular to one another. Far
IR spectroscopic measurements confirm the near perpendicular nature of this complex in
the gas phase, with equal to about 100-110°69. An important contrast between the two
systems is the strong effect of correlation in reducing the interchlorine separation.
It is not immediately obvious which molecule would be the proton donor and which the
acceptor in a complex pairing HF with HC1. Calculations suggest the two possibilities are
nearly equal in energy55. This supposition was confirmed by later observation of both in
the gas phase70. The pertinent features of the complexes are reported in Table 2.20, from
which it may be observed that the stabilization energies of the two differ by less than 0.1
kcal/mol55. The binding energy is intermediate between the two homodimers (HF)2 and
(HC1)2. A recent state-to-state photodissociation study of this mixed complex71 yielded a
dissociation energy Do of 1.83 kcal/mol. Bearing in mind that the latter includes vibrational
energies, which the electronic contributions to the binding energy listed in Table 2.20 do
not, the computed values seem quite reasonable. The complex in which HF acts as proton
donor has a slightly shorter R(F..C1). Indeed, the MP2 H-bond lengths of 3.29 and 3.37 A
for HC1...HF and HC1...HF, respectively, are quite close to the experimental values of 3.28
and 3.37 A reported later70. It may be noted as well that the HC1 acceptor molecule is nearly
perpendicular to the H-bond axis, with = 93°, as in (HC1)2.

Table 2.18 Geometrical and energetic aspects of (HF)2 computed with aug-cc-pVQZ basis set.
Counterpoise-corrected values are indicated by "cc" notation60.
R(F .. F)(A) (degs) (degs) De (kcal/mol)

no cc cc no cc cc
SCF 2.821 2.824 6.8 119.7 3.71 3.66
MP2 2.737 2.753 6.4 111.6 4.63 4.38
MP4 2.735 2.749 6.6 110.4 4.68 4.44
CCSD 2.745 2.759 6.7 112.1 4.53 4.31
CCSD(T) 2.732 2.745 6.7 110.8 4.72 4.49
Geometries and Energetics 75

Table 2.19 Geometrical and energetic aspects of (HC1)2 calculated at various levels.

R(Cl..Cl) (A) (degs) (degs) - E (kcal/mol) Reference

SCF
4-31G 3.986 7 102 2.1 [64]
[6s4pld/2s1pld] 3.96 2.6 83.0 3.6a [65]
6-31+G* 4.156 14.6 — 1.1 [54]
6-31G** 4.111 10.3 97.3 1.0" [55]
+VPs(2d)s 4.210 11.3 90.3 0.5a [55]

Correlated
MP2/6-31G** 3.876 — — 1.4a [55]
MP2/+VPs(2d)s 3.838 — — 1.6a [55]
ACPF/[652/42] 3.912 6.6 91.1 1.7 [66]
ACPF/[6531/42] 3.887 6.1 91.4 1.7 [66]
MP2/[8s6p3d/6s3p]b 3.78 8.0 90.0 2.0 [67]
expt 3.80 [68]
a
Corrected for BSSE.
b
Bond functions added.

There are of course a host of different means of including electron correlation into the
computation of binding energies. It would be useful at this point to make a comparison of
some of these techniques. Table 2.21 lists the interaction energies computed for the HF and
HC1 homodimers, all with the same 6-31+G(d,p) basis set72. LCCM refers to a linearized
coupled cluster technique73,74 and ACPF to an approximate coupled-pair functional ap-
proach75. The configuration interaction technique, truncated after all single and double ex-
citations is designated CISD. As the latter approach is not size-consistent it is completely
unsuitable for study of molecular interactions unless some further steps are taken. Davl and
Dav2 indicate corrections proposed by Davidson76,77 which multiply the correlation energy
by a factor which includes the coefficient of the Hartree-Fock configuration in the normal-
ized CISD wave function. A further scaling, indicated by (s), was added to include the num-
ber of correlated electrons in the expression. The last type of correction considered is due
to Pople78 and is designed for identical 2-electron systems.
The first few rows of Table 2.21 show the enhancement of the SCF interaction energy
arising when M011er-Plesset correlation is added. MP2, 3, and 4 are little different from one

Table 2.20 Geometrical and energetic features calculated for


complex pairing HF with HC155. Eelec corrected for BSSE.

HC1-HF HC1...HF

R(F..C1)SCF, A 3.465 3.499


R(F..C1)MP2, A 3.294 3.365
(degs) 7.4 8.2
(degs) 93.2 119.7
- ESCF (kcal/mol) 1.77 1.93
- EMP2 (kcal/mol) 2.39 2.45
76 Hydrogen Bonding

Table 2.21 Comparative binding energies ( Eelec, in


kcal/mol) computed with different correlated
schemes, all with the 6-31 +G(d,p) basis set72.

(HF), (HC1),
SCF -4.3 -0.8
MP2 -5.0 -2.1
MP3 -4.9 -1.9
MP4 -5.0 -1.9
LCCM -4.9 -1.9
ACPF -4.9 -1.9
CISD +4.4 +7.6
Davl -2.4 + 1.0
Davl(s) -2.8 +0.5
Dav2 -3.3 -0.2
Dav2(s) -3.7 -0.6
Pople -4.5 -1.7

another, a common observation. However, it might be noted that the correlation-induced


enhancement is likely exaggerated as no corrections were made for basis set superposition.
The LCCM and ACPF methods yield results remarkably similar to MR CISD, on the other
hand, predicts that both complexes would be strongly unbound, with positive values of E.
This result is not a surprise, as the CISD method is not size-consistent. The various David-
son corrections seem to improve the energetics, particularly the Dav2 variant, which is
nearly as attractive as the methods in the preceding rows. Even better is the Pople correc-
tion in the last row of Table 2.21. Due to the computational efficiency of the M011er-Pies-
set technique, it would appear to still represent a very cost-effective workhorse for study of
H-bonding interactions.
A later study also focused on various means of computing the correlation contribution
to the interaction energy in the HF dimer79 and reached very similar conclusions. All of the
correlated methods (MP2, MP4, CCSD(T) and CISD) based on the Hartree-Fock reference
configuration gave essentially the same binding energy. The results deteriorate when mul-
tireference methods are used.
There have been calculations that extend the set of hydrogen halides investigated to var-
ious combinations of HBr, HI, and HO80. The studies were limited to the SCF level, and
made use of core pseudopotentials. The optimized geometries are reported in Table 2.22
along with the interaction energies, corrected for BSSE. The H-bond lengths exhibit the ex-
pected increases as one moves to bigger atoms as in Cl < Br < I. In the case of the proton
donor, the increment from Cl to Br is 0.1 A, but a larger stretch of nearly 0.3 A occurs upon
going from Br to I. The increments are larger, around 0.3 and 0.4 A, for the proton accep-
tor. As the proton donor changes from Cl to Br to I, the H-bond becomes progressively more
linear; the smallest a angles of 3° are associated with HI. The linearity of this bond is in-
fluenced, albeit to a lesser degree, by the character of the acceptor, with a becoming larger
for the heavier atoms. All of these complexes are very nearly perpendicular in the sense that
is close to 90°. Experimental confirmation for such a shape for the HI dimer comes from
recent high-n Rydberg time-of-flight measurements 81 . There is a definite trend for the in-
teractions to weaken as the proton acceptor atom is enlarged but little dependence upon the
Geometries and Energetics 7777

Table 2.22 Properties of binary complexes, calculated at the SCF level, using
core pseudopotentials80.

R(A) (degs) (degs) - Eeleca (kcal/mol)

HC1...HC1 4.11 13.4 90.9 1.0


HCl...HBr 4.22 7.8 90.4 1.1
HC1...HI 4.49 3.2 92.2 0.9
HBr...HCl 4.38 16.0 88.3 0.8
HBr...HBr 4.48 8.2 89.8 0.8
HBr...HI 4.76 3.2 92.1 0.7
HI...HC1 4.77 18.2 85.7 0.4
HI...HBr 4.87 8.0 89.8 0.4
HI-HI 5.14 3.3 91.7 0.4
a
Counterpoise corrected.

nature of the donor. HI is the poorest acceptor, with - Eelec equal to 0.4 kcal/mol in all
cases.
It is legitimate to question whether the interactions listed in Table 2.22 represent true H-
bonds or might be better described in terms of simple electrostatic or dispersive interac-
tions. Indeed, the stretches undergone by the HX bonds as a result of formation of each com-
plex are all 0.002 A or less. And the red shifts undergone by this bond are below 20 cm- 1 .
On the other hand, these calculations were limited to the SCF level and thereby ignore some
of the correlation effects that become particularly important for the heavier hydrogen
halides. Nor is the basis set very large, another source of error. For this reason, the interac-
tion energy computed for (HC1)2 by this work is only about half of that obtained at higher
levels (see earlier). One might conclude that the geometries of these complexes are consis-
tent with certain patterns observed in true H-bonds but the energetics and other features are
weaker than would normally be associated with such a bond.
A microwave structure for the complex between HF and HI82 indicates that the angle
becomes even more acute in HF...HF, as small as 70°. The I atom lies along the H—F axis,
with R(I..F) = 3.66 A. This "triangular" structure would argue against the presence of a H-
bond here.

2.5 HYH...YH2

The ubiquitous occurrence of water and its importance as a solvent medium have motivated
a great deal of research into the fundamental nature of the interaction between water mol-
ecules by theoretical as well as experimental means. Some of the more recent work has been
summarized in a review article83. Prior to experimental elucidation of the geometry of the
water dimer in the gas phase or to the ability of calculations to provide an unambiguous res-
olution to this question, a number of different candidate structures were considered. In ad-
dition to the "standard" linear arrangment wherein the bridging hydrogen lies near the O...O
axis in Fig. 2.8, cyclic and bifurcated structures were considered as illustrated in Fig. 2.9.
It is now widely accepted that the linear geometry is in fact the equilibrium structure, al-
though the energy cost in assuming other configurations remains under debate84-86. A rather
78 Hydrogen Bonding

Figure 2.8 Dispositions of molecules and lone pairs


in H2Y dimer.

extensive comparison of the details of the equilibrium geometry and binding energy ob-
tained by different basis sets and levels of theory has been contributed to the literature by
Frisch et al.57. The results are summarized in Table 2.23 which illustrates nicely that the
small basis sets like STO-3G and 3-21G strongly underestimate the intermolecular separa-
tion. There is a nearly consistent trend of more flexible basis sets yielding longer R(OO);
in most cases correlation reduces this distance. The H-bond energies in the last two columns
echo this trend in that larger basis sets yield a lower absolute value of Eelec, but that cor-
relation enhances the binding energy. While there is substantial scatter in the calculated
equilibrium (3 angles, most values are within or close to the experimental range of 113-133°.

2.5.1 Binding Energy of Water Dimer


More recent calculations have further improved on the theoretical method and yielded re-
fined values for the interaction energy of the water dimer. The Hartree-Fock limit of the
electronic contribution to E has been placed at —3.73 ± 0.05 by Szalewiczetal.87, a value
which was confirmed by others88. A slightly smaller estimate of —3.55 kcal/mol emerged
from studies of Feller89 whose basis sets included h functions on O and g on H. In contrast
to the SCF value which is relatively simple to obtain with moderate sized basis sets, the au-
thors were more pessimistic about correlation components, in particular the dispersion en-
ergy. Later work90 placed a fully correlated binding energy of — Eelec at 4.5-4.6 kcal/mol,
confirmed by others incorporating bond functions into the basis set88 or using other corre-
lation techniques91, van Duijneveldt-van de Rijdt optimized the geometry with BSSE cor-
rections included. Correlation seems to have only a minor influence upon angular features
of the equilibrium geometry88. Kim et al.92 found a binding energy of 4.66 kcal/mol at the

Table 2.23 Equilibrium geometries and binding energies (uncorrected for BSSE) calculated
for the linear water dimer at various levels of theory57.

R(O-O) (A) (degs) - E elec (kcal/mol)

Basis set SCF MP2 SCF MP2 SCF MP2

STO-3G 2.740 124.0 5.9


3-21G 2.797 2.802 124.6 107.9 10.9 12.6
6-31G* 2.971 2.913 117.5 102.7 5.6 7.4
6-31 +G* 2.964 2.901 130.3 128.9 5.4 7.1
6-311 ++G** 2.999 2.910 143.1 135.8 4.8 6.1
6-311 + +G(2d,2p) 3.035 2.911 130.8 123.2 4.1 5.4
6-31l + +G(3df,2pd) 3.026 — 133.2 — 4.0 —
expt 2.98 = 0.01 123±10
Geometries and Energetics 79

MP2 level, with a counterpoise correction and a (13s8p4d2f/8s4p2d) basis set. Following
appropriate additional terms, they computed a binding enthalpy H of —2.86 kcal/mol.
When combined with their S of -17.7 cal mol-1 K-1, a G of +3.72 kcal/mol was fi-
nally derived. Their optimized R(OO) of 2.958 A was in nice agreement with the experi-
mental value of 2.976 A. Feller's best correlated E is —5.1 kcal/mol89, somewhat larger
than other workers have found, but supported by recent calculations93.
In a recent effort94, bond functions, centered on regions between atoms rather than on
nuclei, have been added to the basis set. The results lead to an MP2 binding energy of —4.7
kcal/mol, following correction for basis set superposition error. The Hartree-Fock portion
of this interaction, —3.6 kcal/mol, is consistent with prior work. Full optimization at the
MP2 level, with counterpoise corrections included, yield R(OO) = 2.94 A. The nonlinear-
ity parameter, a, is 6° and the proton-accepting water molecule makes an angle of 123° with
the O..O axis. A followup of this work95 suggested that the bond functions were unimpor-
tant here, as any stabilization produced by their presence was largely cancelled by the large
BSSE that they introduce. Another calculation made use of a very large basis set, as many
as 574 functions96. Correlation was included by a method that approaches the MP2 method
in an approximate fashion. The authors concluded that their best estimate of — Eelec is
5.0 ± 0.1 kcal/mol, which leads then to a binding enthalpy at 375 K of 3.2 ± 0.1 kcal/mol.
Hence despite the inordinate attention paid to the water dimer and the application of
state-of-the-art methods, there remains some lingering ambiguity concerning the binding
energy, from both an experimental and theoretical perspective. The largest uncertainty prob-
ably lies in the dispersion part of the interaction energy which appears most resistant to sat-
uration by enlarged basis sets97. At the present time, it is probably safe to say that the elec-
tronic contribution to E is in the range between —4.5 and — 5.0 kcal/mol. About 1.0-1.5
of this amount arises from correlation.

2.5.2 Complexes Containing H2S


While experimental measurements of the water dimer in the gas phase had yielded an un-
ambiguous linear structure, the results for (H2S)2 were less clear98. Cyclic and bifurcated
geometries of the type shown in Fig. 2.9 were also proposed for this dimer.
Following earlier ab initio investigation of this question, van Hensbergen et al. applied
first-order exchange perturbation theory, coupled with an "effective-electron" model99.
This work was unable to clearly differentiate the more stable between the cyclic and linear
geometries, but found bifurcated clearly higher in energy. Later ab initio calculations54
found linear to be preferred to bifurcated, but only by a very small amount. The respective
values of Eelec were —0.9 and —0.7 kcal/mol at the SCF level with a 6-31 + G* basis set.
Taking the theory up to MP4SDQ raised the binding energies to — 1.4 and —1.1, still quite
close to one another. Indeed, addition of zero point energies left the two in a dead heat at
-0.6 kcal/mol. The intersulfur distance of the linear structure was optimized at the SCF
level to 4.524 A, with the bridging hydrogen within 1.4° of the S...S axis. Later calculations
carried the optimization to the MP2 level and found a strong bond contraction accompanied
inclusion of correlation; as seen in Table 2.24100, R(SS) diminished by 0.44 A. The proton
acceptor molecule is nearly perpendicular to the H-bond axis. The last column emphasizes
the weakness of the interaction. Other calculations101 involved a detailed comparison of the
three types of arrangement above using a 6-31G* basis set and found bifurcated to be more
stable than linear at MP2, but only by an amount less than 0.05 kcal/mol; cyclic was less
stable than linear by about 0.4 kcal/mol. A model potential which computes the total bind-
80 Hydrogen Bonding

Figure 2.9 Proposed structures of H2Y dimer.

ing energy as a sum of terras after making some simplifying assumptions identified the lin-
ear arrangement as more stable than bifurcated and cyclic by a significant margin102. One
must consider the relative stability of the bifurcated and linear structures of (H2S)2 an open
question at this point, particularly since the geometry optimizations of both have been lim-
ited to SCF.
A more recent examination of the linear H2S dimer, using a heavily polarized
[9s,6p,3d,f] basis set for S103, yielded an MP2 interaction energy Do of 1.7 kcal/mol, cor-
rected for BSSE. The authors projected an infinite-basis set limit of 1.9 kcal/mol for this
quantity. This calculation also confirmed the lack of any real binding at the SCF level.
Because of their similar proton affinities, it is unclear whether H2S or H2O would be the
proton acceptor in a complex combining the two. Calculations confirm the difficulty of an-
swering this question. Amos104 carried out SCF calculations for various combinations of
these two molecules including the mixed dimer, all in the linear arrangement. While he
found H2S to be the preferred proton donor, the difference in energy versus the case where
H2O is the donor was small enough that Amos considered the calculations not definitive.
The data in Table 2.25 do provide a meaningful comparison of molecular geometries, how-

Table 2.24 Equilibrium geometries and binding energies (uncorrected for BSSE) calculated
for the linear structure of (H2S)2 with 6-31G(2d) basis set100.
R(S-S) (A) (degs) (degs) Eelec (kcal/mol) ( E+ZPVE) (kcal/mol)

SCF 4.600 6.8 97.1 -0.8 0.1


MP2 4.161 3.2 89.6 -2.3 -1.1
Geometries and Energetics 81

Table 2.25 Equilibrium geometries calculated for complexes of H2S and H2O with 6-31G**
basis set at SCF level104.
HOH-OH2 HOH-SH2 HSH--OH2 HSH-SH2

R(YY) (A) 2.922 3.811 3.622 4.489


a (degs) 5 0.2 5 4
(degs) 117 100 131 104

ever. One can see the clear progression toward longer intermolecular distances as each O is
replaced by S. Note that in the mixed complex, R(SH) is shorter by perhaps 0.2 A when S
acts as proton donor. All bridging protons lie within about 5° of the Y-Y axis. It is also worth
noting that, regardless of the identity of the proton donor, the proton acceptor H2S molecule
is nearly perpendicular to the H-bond axis, with angles around 100°. A somewhat later
calculation of the mixed dimers was also unable to resolve the nature of the most stable
complex. After applying the necessary corrections to energetics computed at the MP4 level,
Del Bene41 found binding enthalpies of the HOH-SH2 and HSH--OFL, complexes of -1.3
and — 1.5 kcal/mol, respectively. It is unlikely that higher levels of theory will clearly dif-
ferentiate between the stabilities of these two conformers, since Del Bene's work went up
toMP4/6-31 + G(2d,2p).

2.5.3 Substituent Effects


A minor perturbation on a molecule such as water would be the replacement of one of the
hydrogens by an alkyl group. An early investigation of the interaction between methanol
and water, with the former acting as the donor, yielded an interaction energy of some 6
kcal/mol105, and suggested that dispersion effects would be quite small. A later, more com-
prehensive study including substitution by methyl and ethyl resulted in the binding ener-
gies reported in Table 2.26, computed at the SCF level with a 6-31G** basis set106. These
results show remarkable insensitivity to such substitutions. The most strongly bound of the
set, where ethanol replaces water as proton acceptor, is only 0.3 kcal/mol stronger than the
water dimer. The very similar energies of the water-methanol complexes, where the two

Table 2.26 SCF/6-31G** binding energies


(electronic contributions, in kcal/mol) without
BSSE correction 106.

Complex ~AE
HOH-OH2 5.54
HOH-OHCH3 5.42
HOH-OHC2H5 5.85
CH3OH OH2 5.52
CH3OH--OHCH3 5.42
C2H5OH--OH2 5.44
C2H5OH--OHC2H5 5.66
82 Hydrogen Bonding

molecules reverse their roles as donor or acceptor, has been confirmed by IR spectral ob-
servation: whereas water will act as donor in the gas phase or in Ar matrix107,108, it is the
methanol that is the donor in N2 matrix109.
Calculations of a similar nature110 have demonstrated that replacement of both hydro-
gens of water, yielding dimethyl ether, also has only a minor effect upon the nature of the
H-bond in the water dimer. With their polarized basis set, and with inclusion of corrections
for BSSE, dispersion, and intramolecular correlation effects, these authors found the first
methyl substitution raises the binding energy by 0.5 kcal/mol and the second by 0.6. The
authors cautioned that an unpolarized basis set would fail to pick up these small effects,
which they attribute to Coulomb and dispersion components of the interaction.
Methyl groups have been added to the sulfur analog as well. When water is combined
with S(CH3)2, the water molecule donates a proton to the S111 in a complex with C2v sym-
metry. The bridging hydrogen is computed to lie some 2.727 A from sulfur at the SCF/6-
31G* level and stretches away from the oxygen by 0.003 A as a result of forming the H-
bond. At the MP2/6-311 + +G** level, the binding enthalpy of this complex is -3.5
kcal/mol. This value is significantly larger in magnitude than a prior computation of
HOH SH241 so the methyl groups would appear to make the sulfur a better proton acceptor.
Other effects of substituents on the character of the H-bonding have been studied in one
investigation of complexes involving water with silanol112. Either molecule can act as pro-
ton donor, but the greater acidity of silanol makes the latter the preferred proton donor. With
a doubly polarized triple- basis set, the complex with silanol as proton donor is bound by
4.8 kcal/mol at the SCF level, 6.2 at MP2, but without correction of BSSE. Due to the greater
acidity of silanol, this complex is more strongly bound than the water dimer. The compa-
rable values of — E for the arrangement where water acts as proton donor are 3.1 and 4.6
kcal/mol, respectively, quite similar to that of the water dimer. The authors computed en-
tropic contributions to the binding, enabling them to arrive at the thermodynamic quanti-
ties listed in Table 2.27 from which it may be seen that replacement of the proton-accept-
ing water by silanol has little effect upon the energy or enthalpy of binding but that a
significant boost is obtained if the proton-donating molecule is changed to silanol. In all
cases, the Gibbs free energy of complexation is positive.
Ugliengo et al.113 have optimized the geometries of various pairs of CH3OH and
SiH3OH using basis sets of polarized double- quality. The geometries are all of the linear
variety with all (O—H-O) angles within 10° of 180°. The results are listed in Table 2.28
and indicate that the substitution with a SiH3 group makes for a stronger proton donor, since
the most strongly bound complexes involve this function for silanol. Comparison of the en-
ergetics with the 6-31G* data for the water dimer in Table 2.23 is clouded by the failure to
remove BSSE from the latter. Taking instead the BSSE-uncorrected data of Ugliengo et al.
with a 6-31G** basis set, the binding energies for the water and methanol dimers are vir-

Table 2.27 Calculated thermodynamics of binding of complexes involving water and silanol.
Data in kcal/mol112.
HOH-OH2 HOH-OHSiH3 SiH3OH-OH2

EclccMP2 -4.70 -4.56 -6.25


H° -2.96 -2.94 -4.64
G° 2.97 4.97 3.13
Geometries and Energetics 83

Table 2.28 Energetics of complexes involving methanol and silanol. Data in kcal/mol113 are
corrected for BSSE.

SiH3OH--OHCH3 SiH3OH-OHSiH3 CH3OH-OHCH3 CH3OH OHSiH3

- EelecSCF 5.69 4.59 4.21 3.61


- EelecMP2 6.60 5.59 5.09 4.30
- H (0K) 5.19 4.30 3.90 3.20

tually indistinguishable, suggesting methyl substitution has little effect upon the strength of
the H-bond. A similar conclusion has been drawn in comparisons of the binding of aceto-
nitrile or formamide with either HOH or HOCH3114, albeit with a small basis set.
Substitution of one hydrogen of HOH with an aromatic group leads to a phenol mole-
cule. When paired with methanol, phenol acts as the proton donor molecule in a structure
very much akin to the water dimer itself115. At the SCF/6-31G* level, the interoxygen dis-
tance is 2.89 A. The electronic contribution to the binding energy is computed to be 6.0
kcal/mol, after removal of BSSE, and 7.1 kcal/mol at the MP2 level with the same basis set.
Correction of the correlated result by ZPVE yields a Do of 5.8 kcal/mol, leading to the con-
clusion that phenol is a more potent proton donor than is water.
The structure of the phenol-water complex in the gas phase has been elucidated by mi-
crowave spectroscopy 116,117 and the phenol is indeed the proton donor in this complex. The
interoxygen H-bond length was measured to be in the 2.88-2.93 A range. Prior computa-
tions with basis sets as large as 6-311 + + G(d,p) were consistent with this structure118,119;
R(O O) was computed to be 2.94 A at the SCF level118, and one can expect a significant
reduction upon reoptimizing the structure at a correlated level. Indeed, another computa-
tion finds a distance of 2.83 A in the MP2/6-31G** optimized structure120. These calcula-
tions suggested that this structure is indeed the global minimum on the surface of the phe-
nol-water complex120. The BSSE-corrected interaction energy, Eelec, is calculated to be
— 6.1 kcal/mol at the MP2 level118. Addition of vibrational terms yields a Do of 4.3 kcal/mol.
Eclcc is computed with a more flexible aug-cc-pVTZ basis set to be —6.6 kcal/mol at the
MP2 level, somewhat deeper and indicating some lingering basis set sensitivity. A sec-
ondary minimum represents a reversal in the donor-acceptor roles in that water donates a
proton to the phenol120. (There was some evidence of a third minimum, wherein the water
oxygen atom approaches one of the phenyl C—H bonds.)
Another type of substitution, and one which is likely to have a stronger effect, is the re-
placement of one of the H atoms of water by a more electronegative atom like Cl. This sub-
stitution enhances the proton-donating ability of the water, so that HOC1 is the donor when
combined with a water molecule. Unlike the water dimer itself, for which the anti conformer
is the only stable minimum, both syn and anti arrangements represent minima on the sur-
face of C1OH OH2121 as illustrated in Fig. 2.10. There are no minima corresponding to a
reversal in which HOC1 acts as proton acceptor. However, the reader should be cautioned
that HOH--OHC1 may appear to be a minimum, even at fairly high levels of theory. It re-
quired MP2/6-311 + +G(d,p) to demonstrate it not to be a true minimum.
At all correlated levels of theory, the syn geometry is found to be slightly more stable
than anti. At the MP4//6-311 + + G(3df,3pd)//MP2/6-311 + + G(d,p) level, the electronic
contribution to the binding energy computed for syn is 8.2 kcal/mol, compared to 7.9 for
anti. After a 2.3 kcal/mol ZPVE correction, these binding energies are reduced to 5.9 and
84 Hydrogen Bonding

Figure 2.10 Syn and anti conformers of H2O + HOC1.

5.6 kcal/mol, respectively. The interoxygen distances are 2.78 A in these two geometries,
considerably shorter than in (H2O)2, and the H-bond is within 4° of linearity. The stronger
binding here, as compared to the water dimer, is commensurate with the greater acidity of
the hydrogen on HOC1.

2.6 (ZH3)2

Perhaps more than any other complex, the ammonia dimer has provided the most intrigu-
ing puzzle in piecing together its equilibrium geometry. It was presumed early on that this
dimer would form a H-bond much like the other small hydrides such as HF and H2O, even
if perhaps somewhat weaker. Indeed, there were indications from experimental work that
the equilibrium structure was in fact linear122. For that reason, most of the early ab initio
calculations focused on the linear type of structure.
A representative group of data is collected in Table 2.29 which indicates the sensitivity
of the calculated energetics and geometry to basis set123. The first row illustrates the inap-
propriateness of a minimal basis set, especially STO-3G, for this system. Most of the in-
teraction energy is composed of superposition error, leaving only 0.8 kcal/mol of "true"
binding energy. The intermolecular distance is grossly underestimated. The split-valence 4-
31G represents an improvement in that R(NN) elongates to a more realistic value. Never-
theless, this basis is also subject to a large BSSE, about as much as its real binding energy.
Addition of polarization functions lowers the BSSE to about 1/2 kcal/mol, and yields bind-
ing energies of 2.6 kcal/mol. While [541/31] does include polarization functions, the results
are surprisingly poor, with a large BSSE, and small corrected interaction energy. Probably
the best results emerge from the basis set in the last row of the table that contains two sets
of d-functions, permitting superior treatment of polarization energy as well as electrostat-

Table 2.29 Calculated properties of linear geometry of ammonia dimer at the SCF level.
Data123; all energies in kcal/mol.
R(NN) (A) - Eelec . BSSE -( E + BSSE) (NH3) (D)

STO-3G 3.08 3.8 3.05 0.8 1.66


4-31G 3.31 4.1 1.97 2.1 2.28
6-31G* 3.44 2.9 0.53 2.4 1.93
6-31G** 3.44 3.1 0.55 2.6 1.87
[541/31] 3.44 2.4 1.04 1.4 1.84
6-3IG(2d,lp) 3.54 2.4 0.68 1.7 1.50
Geometries and Energetics 85

ics. It is worth mentioning that the dipole moment calculated for the ammonia monomer
with this basis set, listed in the final column of Table 2.29, is quite close to the experimen-
tal measurement of 1.47 D124. Using a number of approximations123, it was estimated that
the conversion from Eelec in Table 2.29 to E at 298 K requires the addition of 1.6
kcal/mol. Hence, at the SCF level, the ammonia dimer would probably not be expected to
be bound at all, and if so by only a very small amount.
The view of the ammonia dimer changed radically in the mid 1980s with the report by
the Klemperer group that their microwave measurements argued against a linear arrange-
ment125-128. They contended that their data supported a geometry akin to a cyclic structure.
The "classic" linear and cyclic geometries are depicted in Fig. 2.11, along with the struc-
ture proposed from the microwave data.
This interpretation of the microwave spectrum stimulated a flurry of activity to unearth
the correct equilibrium geometry. It was suggested, for example, that photoelectron spec-
troscopy was not inconsistent with a cyclic structure 129 , whereas infrared photodissociation
and matrix infrared measurements suggested the two molecules are not equivalent130,131
and supported the microwave equilibrium geometry132. Another set of measurements led
to the notion that a tunneling motion, similar to that in the HF dimer, which interchanges
the roles of proton donor and acceptor, was responsible for the two IR bands observed in
the gas phase133. State selection in a hexapole electric field indicated that the dimer has a
small dipole moment and that it is not a symmetric top structure134.
On the computational front, Latajka and Scheiner135 considered an extensive region of
the entire potential energy surface as a function of the two angles which describe the ori-
entations of the two molecules, as well as the internitrogen distance. The only true mini-
mum located on this surface corresponded to a cyclic structure in which the two H-bond-
ing protons are displaced 42° from the N-N axis. A very shallow trough leads from this

Figure 2.1 I Three candidate equilibrium


geometries of the ammonia dimer.
86 Hydrogen Bonding

geometry to a linear structure, predicted to lie only 0.2 kcal/mol higher in energy at the
MP2/6-31G(2d,lp) level. This energy difference is unaffected by the inclusion of counter-
poise corrections for BSSE. The conversion from cyclic to linear stretches R(NN) from 3.15
to 3.34 A. As a rationale for the microwave spectral data that seem to point toward a geom-
etry that is not quite cyclic, the authors suggested motions on the surface which might tend
to drive the vibrationally-averaged structure part of the way along the path from cyclic to-
wards linear.
Nearly simultaneously, Frisch et al. addressed the same problem54 but drew a different
conclusion. They found the linear geometry to be the only minimum on their potential en-
ergy surface. The cyclic structure represents a transition state for conversion between the
two linear arrangements, but only 0.2 kcal/mol higher in energy. While this work went up
to fourth-order M011er-Plesset, and included zero-point vibrational energy, it did not attempt
to remove BSSE at any level. Frisch et al.57 reoptimized the geometries of the cyclic and
linear geometries at the MP2 level and confirmed their contention that linear is most sta-
ble. Other than a 0.15 A contraction of R(NN), the MP2-optimized linear structure differs
very little from the SCF geometry. After making counterpoise corrections, other calcula-
tions136 confirmed the very nearly equal energy of the linear and cyclic arrangements, but
their calculations were confined to the SCF level.
Sagarik et al.137 incorporated correlation into the potential via a coupled-pair functional
(CPF) approach which is size-consistent; their basis set was contracted to [642/31]. The
dimerization energies calculated by the ab initio calculations were fit to an analytical site-
site potential function for purposes of molecular dynamics. While the linear geometry was
most stable at the SCF level, CPF calculations yielded a cyclic unsymmetrical structure as
the minimum, wherein the angles of the two monomers differ by 14° with respect to the
N N axis. Although their surface was not adjusted for BSSE, counterpoise correction was
made to the global minimum, after which their binding energy was calculated to be 2.8
kcal/mol. The results confirmed the very shallow nature of the surface along the pathway
between linear and cyclic.
Another set of calculations in 1986 attacked the surface of the dimer by first computing
the electrical properties of the monomer at a high level138. By incorporating these proper-
ties such as multipole moments, polarizabilities, and hyperpolarizabilties into standard for-
mulae of molecular interactions, the authors were able to extract an "electrical interaction"
which includes not only electrostatic forces, but also polarization and dispersion energies.
However, since the exchange which prevents collapse of the complex does not appear in
their formalism, it was not possible to determine optimal intermolecular distances, and
hence to identify the global minimum, so the authors focused their attention upon the an-
gular features of the potential energy surface. Their results emphasized the flatness of the
potential energy surface, in agreement with prior ab initio calculations.
An attempt was made to distinguish the cyclic from the linear H-bond geometry of the
ammonia dimer by modifying the basis set so as to drastically minimize the superposition
error5. The results at the MP2 level favored the cyclic structure by a small amount, some
0.2 kcal/mol, with a doubly polarized basis set, containing one set of diffuse functions. It
was noted that a basis of the same general quality, but unmodified to reduce BSSE, could
yield the opposite conclusion as to which geometry was the more stable. Computations with
a larger 6-311G(2d,2p) basis set in 198759 agreed that the cyclic was preferred to the linear
structure by 0.2 kcal/mol at the MP4 level, but the two became indistinguishable when a set
of diffuse functions were added to the nitrogens. Later gradient search of the ammonia
dimer PES with the 6-31G* basis set7 led to a cyclic geometry; there was no minimum iden-
Geometries and Energetics 87

tified on the surface for the linear H-bond. Application of an MP4/6-31+G(2d,2p) elec-
tronic energy to this geometry, in combination with SCF/6-31G* vibrational frequencies,
led to a binding enthalpy, H298, of —1.6 kcal/mol, roughly a third of the earlier experi-
mental estimate of this quantity139.
The level of theory was increased in 1991 by Hassett et al.140, with basis sets as large as
6-311 +G(3d',2p) and correlation by MP2 and QCISD. They learned that the characteriza-
tion of a given geometry as a particular type of stationary point was subject to small changes
in the type of calculation. For example, while the cyclic structure is indeed a stationary
point, it is a minimum if the basis set does not contain diffuse functions; it is otherwise a
transition state for NH3 rocking motions. In agreement with most other calculations, the mi-
crowave geometry could not be located as a minimum on any surface. The energy differ-
ence between cyclic and linear is very small indeed; at their highest level of theory,
QCISD(T), the De values are 3.06 and 3.15 kcal/mol, respectively. Correction for BSSE
makes these energies identical. Inclusion of zero-point vibrational energies yields a bind-
ing energy of 1.44 kcal/mol. The authors criticized the earlier Sagarik calculations137 as
failing to identity true minima on the surface and imposing certain assumptions on the
geometry of the dimer.
Very large basis sets, including several polarization functions and bond functions, in con-
cert with MP2-4 for correlation, have been applied by Tao and Klemperer141, along with
counterpoise corrections. They find very little energetic separation between the cyclic and
linear structures, with the preference dependent upon presence or absence of bond func-
tions. Their findings echo the earlier work of Latajka and Scheiner135 in that, at their high-
est levels, the cyclic geometry is most stable, but a truly shallow energy path leads to the
linear structure. Table 2.30 reiterates the nearly equal stability of the two geometries. Us-
ing their very flexible basis set, the authors found linear to be preferred by some 0.1 kcal/mol
at the SCF level, but this difference is eliminated at any of the correlated levels, leaving an
essential dead heat. The data further confirm the excellent correspondence between MP2
and MP4 dimerization energies. The "microwave" geometry proposed by Klemperer et al.
was found to be rather unstable.
Cybulski142 considered the ammonia dimer using even larger basis sets, as many as 200
functions, also with bond functions included. He noted a delicate balance between the spe-
cific functions used, or the centers of the bond functions, and the relative stability of the lin-
ear and cyclic geometries and warns against using bond functions without first carefully

Table 2.30 Variation of interaction energy (— Eelec) of ammonia


dimer upon level of correlation. Data, in kcal/mol, were calculated
with [753/41] basis set, augmented by bond functions. Both
geometries fully optimized at MP2 level, including BSSE
corrections141.
Level Linear Cyclic

SCF 1.67 1.56


MP2 2.86 2.88
MP3 2.78 2.78
MP4SDQ 2.66 2.67
MP4 2.85 2.87
88 Hydrogen Bonding

balancing the functions centered on nuclei. He concludes the linear structure is more stable
by only 0.03 kcal/mol. Echoing earlier findings, the basis set superposition error was found
capable of introducing distortions into the PES which might appear to favor one confor-
mation over another, in this case linear versus nonlinear143. A CASSCF treatment leads to
a linear structure as the minimum prior to BSSE correction, but nonlinear is clearly pre-
ferred following such correction.
The apparent paradox between theory and experiment has been largely resolved by care-
ful measurements of the Saykally group, reported in 1992144,145. These workers deduced
from analysis of over 800 new far-IR absorption lines that the appropriate molecular sym-
metry group for the ammonia dimer is G144. Hence, interpretation of the spectra must al-
low for three different types of tunneling motion, including the unexpected umbrella in-
version tunneling. In addition, surprisingly large interchange tunneling splittings were
observed. These findings questioned the assumption made earlier by the Klemperer group
that the umbrella tunneling was completely quenched and donor-acceptor interchange tun-
neling nearly so. In conjunction with six-dimensional vibration-rotation-tunneling dynam-
ics calculations, Saykally et al. deemed it unlikely that the microwave geometry advanced
by Klemperer was an equilibrium structure at all.
More recent work has examined this question by formulating an empirical potential for
the ammonia dimer, based on experimental and theoretical data146,147, and includes the ef-
fects of off-diagonal Coriolis interactions and octupole moments for the electrostatic inter-
actions. Improved results for VRT states, far-IR frequencies, and properties of the protiated
and deuterated dimers were obtained. The authors add further support to the contention that
the ammonia dimer is highly nonrigid, and the essential nature of including vibrational av-
eraging effects148. Their potential contains an energy barrier to interchange of only 7 c m - 1 .
Olthof et al. go on to demonstrate that the data which led the Klemperer group to incorrectly
conclude the dimer was rigid would be better explained as the competing effects of: (1) a
stronger localization of (ND3)2 near one of its minima, and (2) its smaller ortho-para dif-
ference. The authors conclude that ab initio calculations are probably not yet up to the task
of the 7 cm-1 accuracy necessary for truly accurate representation of such a weakly bound
dimer. Based on the potential which best fits the experimental data, the authors147 hypoth-
esize that there are two equivalent minima in the PES that correspond to nearly linear H-
bonds. The top of the barrier for their conversion is a cyclic type of geometry. VRT-aver-
aging of the ground state leads to a geometry that is nearly cyclic. In their view, this is not
a H-bond, even though some of the features of a H-bond are present.
This work is counterpointed by a study of the analogous dimethylamine dimer in the gas
phase149 in which the authors deduce a geometry that appears to contain a distorted linear
H-bond. That is, both the H atom of the proton donor molecule and the lone pair of the ac-
ceptor are bent to one side of the N N internuclear axis. But this distortion is substantial,
and the designation as linear type is not entirely clear; indeed, the authors refer to their struc-
ture as "cyclic" although only one of the hydrogens can be conceivably involved in a H-
bond.
Other gas phase work has indicated the possibility for ammonia acting as proton donor
in a H-bond as well. Held and Pratt150 examined the complex of 2-pyridone with ammonia
by rotationally resolved spectra and found what they consider two H-bonds holding the two
molecules together. The shorter, and probably stronger of the two, has the N—H group of
the 2-pyridone acting as proton donor to the N of ammonia, with an H N distance of 1.99
A. A longer bond connects one of the protons of ammonia to the carbonyl oxygen of 2-pyri-
donc, with R(H--O) = 2.91 A. The designation of the latter as a H-bond is questionable due
Geometries and Energetics 89

to its length, but the possibility cannot be dismissed out of hand, especially as the authors
found evidence of its existence in the fluorescence excitation spectrum; the torsional bar-
rier to rotation of the ammonia is quite a bit larger than steric barriers normally encountered
in complexes of this type with no H-bond.
It is the unanimous conclusion of all high-level theoretical work that the potential en-
ergy surface of the ammonia dimer is quite flat, particularly the region connecting the cyclic
and linear structures. While there remains some disagreement as to which of the latter two
is more stable, the energy difference between them appears to be vanishingly small. It would
hence be more realistic to speak not so much of the single equilibrium geometry of the am-
monia dimer, as of a vibrationally averaged structure, with high-amplitude vibrational mo-
tions. The binding energy is rather low, probably less than 2 kcal/mol after vibrational en-
ergies are included. It is questionable whether this dimer should be classified as bound by
a hydrogen bond in the usual sense.
The presence of a H-bond in the PH3 dimer is even less likely. Frisch et al.54 found the
equilibrium structure to be of cyclic type with no other minima on the surface. The two P
atoms are 4.346 A from one another and each of the bridging hydrogens is located 88° from
the P P axis. The electronic contribution to the binding energy is less than 1 kcal/mol. Af-
ter reductions of this quantity by zero-point vibrational energies and removal of BSSE, it is
questionable whether this dimer would be bound at all.
A calculation of the mixed complex of NH3 with PH3 locates two minima on the sur-
face7. Both contain what appears to be a linear H-bond; NH3 is the proton donor in one min-
imum and PH3 in the other. At the MP4/6-31 +G(2d,2p) level, both complexes are bound
by 1.0 kcal/mol. However, since no correction has yet been made for superposition error,
nor is there any computation of zero-point vibrational energy, it is likely that these minima
will effectively disappear when these corrections are made.

2.7 Carbonyl Group

The doubly bonded oxygen atom of the carbonyl group presents an interesting contrast to
the hydroxyl of water and related molecules. Lewell et al.151 have optimized the geometry
of the complex between formaldehyde and water using a variety of fairly small basis sets
and find that the latter molecule acts as the proton donor and the carbonyl as the acceptor.
The bridging proton does not lie along the C=O axis but is off to one side, along a "lone
pair direction" as indicated in Fig. 2.12.
The trends in Table 2.31 illustrate that the carbonyl group obeys trends much like the
simpler hydrides 151 . As the basis set becomes more flexible, the interaction energy is low-
ered and the intermolecular distance lengthened. Much of this trend is due to the reduction

Figure 2.12 Pairing of formaldehyde with wa-


ter, containing nearly linear H-bond.
90 Hydrogen Bonding

Table 2.31 Optimized geometries (A and degs) and energetics (kcal/mol) of H-bond between
water and formaldehyde151. Results at SCF level, not corrected for BSSE. See Fig. 2.12 for
atom numbering scheme.

Basis set r(O2H1) r(CO2) r(O1H1) (O1H1O2) - E

STO-3G 1.88 0.002 -0.001 177.9 3.38


3-21G 1.97 0.006 0.002 141.0 9.14
6-31G 2.04 0.005 0.004 139.4 6.69
6-31G** 2.12 0.005 0.004 146.3 5.25

of the artificially attractive BSSE, which was not corrected by the authors. STO-3G yields
anomalous results here, as in other systems. The covalent bond between O1 and H1 under-
goes a small stretch, as a result of the formation of the H-bond. A new feature here is the
stretch of the carbonyl bond between C and O2. This stretch can be rationalized within the
context of this bond losing double-bond character as the bridging hydrogen moves toward
it. The authors pointed out also that, whereas the STO-3G and 6-31G basis sets yield a fully
planar geometry for the complex, 3-21G and 6-31G** predict that the peripheral hydrogen
of water will swing out of the plane by some 60°. The penultimate column indicates that there
is a fair degree of nonlinearity predicted within the H-bond for all basis sets except STO-3G.
Calculations on this system were extended to include correlation in 199090, although in-
ternal geometries of each subunit were held rigid. With a polarized double- basis set, the
interoxygen distance is calculated to be 3.00 A at both the SCF and CEPA-1 levels, only
slightly shorter at 2.96 A for MP2. The (COO) angle is 100 ± 2° at any level; there is a
20° nonlinearity in the H-bond. The binding energies are surprisingly insensitive to the in-
clusion of correlation. At all three levels, — Eelec is calculated to be in the range 4.2-4.4
kcal/mol, with counterpoise corrections included. This interaction energy is quite similar to
the binding computed by the same authors for the water dimer.
Kumpf and Damewood152 examined the entire surface of the water-formaldehyde com-
plex so as to consider all possible candidate geometries for the lowest energy. They found
structure II in Fig. 2.13 to be slightly more stable than structure I earlier. Structure II dif-
fers from I in that the H-bond is far from linear, that is, the (O1H1O2) angle is not close to
180°. This nonlinearity is compensated to some extent by the proximity of the water oxy-
gen toward one of the CH2 hydrogens. Another way of thinking of the reason for its stabil-
ity is that the dipole moments of the two subunits are nearly antiparallel.
The details of the optimized geometries of I and II are compared in Table 2.32, where it
may be seen that the interoxygen distance is shorter in II, but the distance from the bridg-

Figure 2.13 Alternate pairing of formaldehyde


with water, with strongly bent H-bond.
Geometries and Energetics 91

Table 2.32 Optimized geometries (A and degs) of geometries


1 and II for the complex between water and formaldehyde; data
calculated at the SCF level with 6-31G** basis set152.

I II

R(O..O) 3.042 2.945


r(O2..H1) 2.096 2.107
(O1H1O2) 176.7 146.7
r(CO2) 0.004 0.005
r(O1H1) 0.003 0.004

ing hydrogen, H1, to the acceptor oxygen is longer152. The H-bond is nonlinear by some
35° in II. Table 2.33 illustrates the very similar energies of the two structures, and how the
relative stability can in fact shift from one level of theory to the next. For example, struc-
ture II is more stable with the 6-31G** basis set, but the situation becomes murkier for 6-
311+G**. It is difficult to draw any certain conclusions, especially since the authors did
not attempt to remove BSSE, but structure II does appear to be at least slightly more stable.
The same workers considered a number of other types of geometries. However, it was un-
clear as to how many represent true minima since a number of geometries were optimized
under symmetry constraints and the numbers of imaginary frequencies were not reported.
Dimitrova and Peyerimhoff153 focused their work on geometry I and helped provide a
more accurate assessment of its interaction energy. Their highest level of theory stopped at
MP2 but incorporated a 6-311 + +G(2d,2p) basis set. The SCF part of the interaction en-
ergy is —4.79 kcal/mol, reduced to —4.04 when corrected for BSSE. The superposition-
corrected contribution from MP2 correlation is —0.97 kcal/mol, adding up to a value of
Eelec = — 5.0 kcal/mol, quite similar to the best estimates for the water dimer. When zero-
point vibrational corrections are added, this quantity lowers in magnitude to —3.3 kcal/mol.
Ramelot et al.154 probed the nature of the minimum with the highest level of correlation
to date, fully optimizing the geometries of stationary points. They verify the nature of struc-
ture II as a true minimum. At their highest level of theory, CCSD with a doubly-polarized

Table 2.33 Interaction energies (- Eelec, in


kcal/mol) of geometries I and II for the complex
between water and formaldehyde; data152 not
corrected for BSSE.
I II
SCF/6-31G** 4.6 5.2
MP2/6-31G** 5.6 6.7
SCF/6-311 + G** 4.0 3.9
MP2/6-311 + G** 4.7 5.0
MP3/6-311 + G** 4.6 4.9
MP4SDQ/6-311 + G** 4.5 4.8
SCF/6-31G** + ZPVE 2.9 3.0
MP2/6-31G** + ZPVE 3.9 4.6
92 Hydrogen Bonding

triple- basis set, the bridging hydrogen is 2.007 A from the carbonyl oxygen and the
(O 1 H 1 O 2 ) angle is 150°. The two O atoms are separated by 2.881 A. The authors disputed
the earlier claim by Kumpf and Damewood of a second weaker H-bond between the water
oxygen and a CH2 hydrogen, noting that their electron density plots indicated little of the
perturbation characteristic of a H-bond. While Ramelot et al. find structure II most stable,
they confirm the earlier conclusions by Kumpf and Damewood that a structure like I, with
a more linear H-bond, is only marginally less stable. Indeed Ramelot et al. find only a 0.05
kcal/mol difference in energy. The authors recommend that diffuse functions should be
added to this system whenever possible.
The complex pairing H2CO with HF was examined recently at a high level of theory155.
Optimization of the geometry at the MP2/6-311 + + G(2df,2pd) level yielded an intermo-
lecular R(O-F) distance of 2.627 A, close to an experimental value of 2.66 A156. The H-
bond is within 13° of linearity, with 0 (FH-O) = 167°. This H-bond is clearly shorter than
that in H2CO-H2O. It is stronger as well, with a binding energy — Eelec of 7.57 kcal/mol
(including counterpoise correction). Following the standard corrections, most notably the
zero-point vibrations, the full — E is computed to be 4.99 kcal/mol. Raising the level of
theory to MP4 lessens the latter quantity by 0.1 kcal/mol. The authors noted that inclusion
of diffuse (+) functions provided an important component to their interaction energy,
adding about 1 kcal/mol. Inclusion of these functions serves another important purpose.
They reduce the BSSE by a factor of three. Consequently, computations with such diffuse
functions are recommended to avoid certain spurious effects which might not be fully cor-
rected by the counterpoise technique.
Correlation was incorporated into the geometry optimization of the complex of
formaldehyde with HC1157 in 1988 using basis sets as large as doubly polarized triple- .
Table 2.34 lists the relevant properties of this complex at the SCF and two different corre-
lated levels, MP2 and a coupled-pair functional (CPF) approach. The best value obtained
for the electronic contribution to the binding energy from this study would appear to be
around 5.0 kcal/mol, which is reduced to about 3 kcal/mol after zero-point vibrations are
considered. Counterpoise corrections would likely further reduce this interaction, were they
to be introduced. A later work explicitly accounted for basis set superposition and found
these effects to indeed be important 158 . The binding energy, including ZPVE corrections,
was computed to be 2.65 kcal/mol at the MP4/6-31 l + +G(2df,2pd) level; a CCSD treat-
ment of correlation yields a similar value.

Table 2.34 Optimized geometries (A and degs) and energetics (kcal/mol) of H-bond between
HC1 and formaldehyde157. Energetics not corrected for BSSE.
SCF MP2 CPF

DZP TZ2P DZP TZ2P DZP TZ2P

R(Cl--O) 3.356 3.375 3.166 3.124 3.246 3.208


r(C02) 0.003 0.003 0.004 0.005 0.004 0.004
r(ClH) 0.008 0.008 0.018 0.021 0.011 0.014
(C1H,02) 172.7 169.1 168.0 164.4 167.8 163.2
(CO 2 H,) 132. 1 124.2 113.7 107.3 115.7 109.0
- Eelec 4.3 3.6 6.2 5.7 5.3 4.8
-( E + ZPVE) 2.9 2.2 4.3 3.8 3.6 3.1
Geometries and Energetics 93

A recent set of IR spectra in the gas phase159 suggest that the interaction energy is not
very sensitive to substitution on the carbon atom. The measured interaction energies of HC1
with acetone, 2-butanone, methyl formate, and methyl acetate are all within 0.5 kcal/mol
of one another.
The energetics in the last two rows of Table 2.34 illustrate that correlation acts to sig-
nificantly enhance the binding, even more for MP2 than for CPF157. This general trend is
evident in the optimized geometrical parameters as well. For example, the SCF equilibrium
R(Cl-O) distances are about 3.36 A, but are contracted by correlation, more so by MP2 than
by CPF. The CPF/TZ2P distance is in fact very close to the experimental measurement of
3.21 A160. Correction of the potential for BSSE increases the R(Cl--O) H-bond length by
some 0.05 A158. The C=O bond stretches by some 0.004 A upon complexation, accom-
panied by a longer stretch of the H—Cl bond. These bond stretches are highly sensitive to
correlation and the method of its inclusion. The CPF/TZ2P r(HCl) stretch is 0.014 A; it is
even longer, 0.020 A, at the MP2/6-311 + +G(2df,2pd) level158.
The bridging proton is predicted to lie within 15° of the H-bond axis and the HC1 mol-
ecule approaches along the general direction of a carbonyl lone pair, with (COH) angles
of 120 ± 13°. The authors emphasize that an angle of just this magnitude would be expected
based upon the electrostatics of the two subunits. In the analogous complex between
formaldehyde and HF, the rotational spectrum indicates a similar structure156, with a (COH)
angle of 115°, although high-level computations yield a slightly smaller angle in the
102-110° range155. Perhaps more important, the energy of the H2CO-HC1 complex is fairly
insensitive to this angle. The authors also draw a parallel between the optimized value of
this angle and the dipole moment computed for H2CO157. It stands to reason that the di-
pole-dipole interaction between H2CO and HC1 will drive the complex toward a linear
arrangement with 0 (COH) equal to 180° so deviations from this large angle will hinge upon
a smaller moment.
Similar arguments pertain to the much smaller angle of 91° computed for the sulfur-ana-
log, H2C=S-HF, at the MP2/6-311 + +G(d,p) level23. This nearly perpendicular arrange-
ment is in fact consistent with a survey of H-bonds in crystals where the 0 (C=S H) angle
distribution peaks at about 110°, as compared to a maximum probability of almost 130° for
6 (C=O H) angles23. The authors note that their results of a smaller angle for the sulfur-
containing systems are best explained by a set of Coulombic interactions in the complex.
While there does appear to be some degree of preference for proton acceptors to ap-
proach the carbonyl oxygen atom along a lone pair direction, there is not. much energy cost
to an approach of the proton donor along the C==O axis. In fact, the entire region in be-
tween the two lone pairs can be considered "fair game" for formation of a H-bond. When
the H-bond formed by H2CO and the proton donor is weakened by the replacement of HF
by HCN, the (COH) angle enlarges nearly 30° to 138°,161 and the energy barrier to in-
creasing this angle to 180° is observed to be low. This sort of result is confirmed by surveys
of crystal structures13'162-166.

2.7.1 Substituent Effects


A recent work has systematically examined the effects of replacing one of the H atoms of
H2CO by F or Cl, then pairing this proton acceptor with each of HF or HC1167. Geometries
were fully optimized at the MP2 level, using a flexible 6-311G(2d,2p) basis set. The es-
sential features of these complexes are listed in Table 2.35 from which it may be seen that
the H-bonds are somewhat shorter for the complexes containing HF, as compared to HC1.
94 Hydrogen Bonding

Table 2.35 Geometries (A) and energetics (kcal/mol) of complexes pairing HXCO with HX
(X = F or Cl). Data computed at MP2/6-311 G(2d,2p) level l67 .
R(O-H) r(HX) r(C=O) Eelec Eelec + ZPVE

HFCO-HF 1.867 0.007 0.006 -4.57 -2.52


HC1CO-HF 1.871 0.007 0.010 -4.12 -1.84
HFCO-HC1 2.087 0.010 0.005 -3.45 -1.85
HC1CO-HC1 2.088 0.009 0.009 -3.22 -1.60

More specifically, the equilibrium distance between the carbonyl oxygen and the bridging
proton is smaller by 0.2 A, as evident in the first column of data. The next column indicates
that the HF bond elongates by less than does the HC1 bond in forming these complexes. In-
corporation of a Cl atom makes the C=O bond more flexible: formation of the H-bond
causes this bond to stretch by twice as much in HC1CO, as compared to HFCO. The ener-
getics in the last two columns reveal a slightly stronger interaction for HF as compared to
HC1. There is also a smaller tendency for HFCO to form stronger H-bonds than HC1CO. It
is important to note a lingering basis set sensitivity. Comparable computations with a
slightly smaller 6-311G(d,p) basis set, containing fewer polarization functions167, indicate
different trends, favoring HClCO HCl as the most strongly bound, and HFCO HCl the
weakest.
Comparison of the results in the last column of Table 2.35 with unsubstituted
H2CO HC1, at a similar level of theory158, indicate that unsubstituted H2CO is a better pro-
ton acceptor; the counterpoise-corrected Eelec, with ZPVE included, amounts to —3.12
kcal/mol for H2CO HC1. The H-bond in H2CO HC1 is also shorter. R(O H) is equal to
1.85 A in H2CO HC1158, as compared to 2.09 A in HXCO HC1. The reduction in proton-
accepting ability, that appears to be associated with replacement of a hydrogen by a halo-
gen, can be understood simply on the basis of the greater electronegativity of the halogens
which drain electron density away from the carbonyl oxygen atom.

2.8 Carboxylic Acid

The carboxylic acid group is of particular interest as it contains both the hydroxyl —OH
and carbonyl C=O functionalities on the same molecule. Its acidic nature is exhibited, for
example, in crystal structures where its H-bonds tend to be shorter and straighter than those
of other proton donors168.
The geometries of complexes pairing HCOOH with HF and HC1 have been optimized
at the SCF level169 and the results presented in Table 2.36, based on the geometrical para-
meters described in Fig. 2.14. Note that when in position I, the HX molecule acts as proton
donor to the carbonyl oxygen, and as donor to the hydroxyl in site II.
The H-bond lengths are of course longer for HC1 than for HF, due principally to the larger
size of the Cl atom. The 4-31G basis set yields a shorter bond than 6-31G**. Of greatest
import, site I yields a uniformly shorter H-bond than site II, indicating a stronger H-bond
for the former. This expectation is confirmed by the energetics in the lower part of Table
2.36, where the HX molecule at site I is bound by an energy greater than that of site II by a
factor of approximately two. We conclude that the doubly bonded oxygen acts as a much
Table 2.36 Optimized geometries (A and degs) and energetics (kcal/mol) of H-bond configurations I and II (see
Fig. 2.14) between HF or HC1 and formic acid169. E refers to electronic contribution, not corrected for BSSE.

HF HC1

I II I II

4-31G 6-31G** 4-31G 6-31G** 4-31G 6-31G** 4-31G 6-31G**

R(O X) 2.586 2.636 2.618 2.740 3.180 3.308 3.262 3.435


r(XH) 0.019 0.015 0.009 0.004 0.011 0.009 0.008 0.004
r(C=O 1 ) 0.014 0.013 -0.005 -0.003 0.008 0.006 -0.004 -0.003
r(C-O2) -0.020 -0.016 0.018 0.012 -0.011 -0.009 0.011 0.008
6(X-OC) 97.5 98.3 112.8 110.3 105.9 114.3 126.0 129.1
(HX-O) 24.7 24.6 22.8 27.5 24.7 17.2 2.0 1.7
_ ESCF 14.52 11.75 8.60 5.03 6.98 5.08 4.06 2.53
_AE MP2 14.72 6.70 6.64 3.86
96 Hydrogen Bonding

Figure 2.14 Possible configurations of HCOOH


+ HX.

more effective proton acceptor than does the hydroxyl oxygen, and will be the preferred site
in most cases. The preference of site I over II is confirmed by observation of the
HCOOH---HF complex in solid Ar170. The interaction energies in the last row of Table 2.36
would probably be reduced by 1 or 2 kcal/mol were BSSE corrected. One may hence con-
sider the electronic contribution of the binding energy of HF to formic acid to be perhaps
12-13 kcal/mol, stronger than the case where formaldehyde is the acceptor molecule. Like-
wise, the interaction of formic acid with HC1 would be some 5 kcal/mol, only slightly
stronger than the comparable data for formaldehyde in Table 2.34.
Other indicators of the stronger interaction with the carbonyl oxygen emerge from the
perturbations of the internal geometries. Considerably larger stretches of the HF or HC1
bond are seen for site I, as much as 0.015 A for HF with the larger basis. As in the case of
formaldehyde, H-bonding to the carbonyl oxygen stretches the C=O bond. Interestingly,
in the case of formic acid, a concomitant contraction of the C—O bond to the hydroxyl oxy-
gen is observed as well. This contraction is larger in magnitude than the C=O stretch. Anal-
ogously, formation of a H-bond to the hydroxyl oxygen stretches the C—O bond and shrinks
C=O.
There is an interesting difference in angle of approach amongst the various types of H-
bonds. First, the 9 (X OC) angles are smaller for site I than for II. (These angles are insen-
sitive to basis set choice.) The specific halogen atom makes a difference in that the angles
are appreciably larger for HC1 than for HF. Another intriguing distinction is the greater de-
viation from linearity for HF, as indicated by the larger values of the (HX O) angle.
Zheng and Merz48 have paired HCOOH with water which serves as the proton donor.
The geometry optimized with a 6-31G* basis set is like geometry I for HCOOH plus HX,
and is illustrated in Fig. 2.15 as the syn geometry. Table 2.37 indicates that correlation
strengthens the H-bond and shortens the interoxygen distance by some 0.07 A. Compari-
son with Table 2.36 indicates the interaction of HCOOH with water is slightly weaker than
with HF, but certainly stronger than with HC1. The last row indicates that HOH and HCOOH
are bound with respect to one another, that is, G<0, only after correlation has been added.
Calculations at the same level 171 emphasized the "cyclic" nature of the syn geometry,
in the sense that O acts as a proton acceptor from HCOOH while the water proton is do-
nated to the carbonyl oxygen. The inability of such a geometry to form when the carboxylic
acid is in its anti configuration was stressed. The data represented an improvement over the
earlier data48 in that Nagy et al. included a counterpoise correction. This correction is quite
large since it reduces the SCF interaction energy of the syn structure from — 10.7 to —8.8
Geometries and Energetics 97

Figure 2.15 Syn and anti arrangements of HCOOH + HOH.

kcal/mol, and the MP2 value lowers from —14.3 to —10.1 kcal/mol. These results are es-
sentially unchanged when the basis set is enlarged to 6-311 + + G**. The interaction energy
in the syn structure is greater than that in the anti by some 1.3 kcal/mol at either the SCF or
MP2 level.
Instead of water or HF, one may pair HCOOH with a nitrogen base. The data on the right
side of Table 2.37 indicate that the greater basicity of methylamine versus water leads to a
stronger H-bond at either SCF or MP2 level of theory. Note, however, that the internuclear
distances are comparable.
The Cl atom in CH3C1 can unsurprisingly act as a proton acceptor when paired with
formic acid. In conjunction with this particular interaction, one of the H atoms of methyl
chloride can "curl around" and approach an oxygen atom of HCOOH172. Several geome-
tries of the dimer which contain this second potential H-bond are illustrated in Fig. 2.16.
The notation lists the proton donor group of the formic acid (which can be either the OH or
CH), followed by the proton acceptor oxygen (—O or —O).
The computed binding energies of these various configurations are listed in Table 2.38
at various levels of theory. All methods agree that the most stable is the OH/=O geometry.

Table 2.37 Optimized interoxygen distance and thermodynamics of H-bond between formic
acid and water or CH3NH2, calculated with 6-31O* basis set48. Energetics not corrected
for BSSE.

Water CH3NH2

SCF MP2 SCF MP2

R(O-Ba) (A) 2.796 2.726 2.834 2.718


Eelec (kcal/mol) -10.8 -14.5 -11.5 -16.3
H (kcal/mol) -8.3 -12.5 -9.6 -14.4
S ( c a l m o l - 1 (leg - 1 ) -30.1 -29.9
C (kcal/mol) 0.7 -3.5 -0.7 -5.5
a
B = O or N.
98 Hydrogen Bonding

Figure 2.16 Geometries of stationary points on surface of complex pairing


HCOOH with CH3C1.

This result is not surprising as the OH group is certainly more acidic than is CH of formic
acid and the carbonyl oxygen should serve as a better proton acceptor than the hydroxyl
oxygen. The electronic contribution to the binding energy is some 6.0 kcal/mol, reduced to
5.0 when zero-point vibrational energies are added. There is some ambiguity as to the next
most stable configuration, but it appears to be OH/—O, again due to the acidity of the car-
boxyl hydrogen. This structure is less stable than OH/=O in part because of the strain in-
curred by the OH---C1 grouping so as to enable the H atom of CH3C1 to approach the hy-
Geometries and Energetics 99

Table 2.38 Interaction energies, Eelec, (in kcal/mol) computed for various configurations of
complex between HCOOH and CH3Cl172. Notations from Fig. 2.16.

Basis set OH/=O OH/-O CH/=O CH/-O


6-311+G** SCF -3.89 -2.01 -2.35 -1.39
6-31G** MP2 -6.09 -3.32 -3.60 -2.43
MP2 + ZPVE -5.14 -2.75 -2.92 -1.92
6-31 + G* MP2 -5.98 -3.82 -3.45 -2.63
MP2 + ZPVE -5.03 -3.25 -2.77 -2.12

droxyl oxygen. Comparable in stability is CH/=O where the CH of formic acid substitutes
for OH as the proton donor group. The same is true of the least stable geometry, labeled as
CH/—O. The authors did not correct their interaction energies for BSSE, but noted that the
corrections are all 0.40 kcal/mol or less at the SCF level, while no attempt was made to as-
sess the MP2 error.
One can take the results above to argue that the CH group can act as a proton donor. In
the OH/=O and OH—O configurations, the CH group of CH3C1 does approach an accep-
tor atom so as to resemble a H-bond. Granted, it is true that this interaction can be consid-
ered as simply auxiliary to the "main" or conventional H-bond between the OH of formic
acid and the Cl atom. However, both of the H-bonds present in each of the other two con-
figurations, CH/=O and CH/—O, involve a CH group as proton acceptor, with no "con-
ventional" H-bonds at all. (The possible existence of CH H-bonds will be discussed in
greater detail in Chapter 6.)

2.8.1 Carboxylic Acid Dimers


When two carboxylic acids are paired together, it is possible to form two H-bonds, within
an eight-membered ring, as illustrated for formic acid in Fig. 2.17. Much of the early work
on this complex, applying SCF-type calculations to small unpolarized basis sets, illustrated
the sensitivity of the intermolecular separation to the basis set173-175. The geometrical pa-
rameters optimized for this dimer are reported in Table 2.39 for a variety of higher levels
of ab initio theory176. With respect to the internal geometry of each monomer, there is not
much change ongoing beyond the polarized double- basis set. Electron correlation elon-
gates both the carbonyl C=O bond and the OH bond. When the dimer is formed, both of
these bonds become somewhat longer. The C=O bond stretches by 0.14-0.18 A, with the
stretches on the longer end of this spectrum obtained with correlation included. Similarly
for the OH bond, stretches of as much as 0.025 A are observed at the MP2 level. Consis-
tent with these greater intramolecular perturbations at the correlated level, the H-bonds are

Figure 2.17 Geometry of formic acid dimer.


Table 2.39 Optimized parameters (A and degs) of the geometry of the formic acid dimer 176 .

Basis set r(C=O) r(C=O)a r(OH) r(OH)a R(O O) 6(OH O)

DZP SCF 1.191 0.014 0.967 0.015 2.767 172.3


MP2 1.221 0.018 0.998 0.025 2.670 177.8
TZ2P SCF 1.190 0.015 0.961 0.014 2.781 174.6
MP2 1.219 0.018 0.998 0.021 2.672 180.0
expt 1.217 1.03 2.70
a
Relative to monomer.
Geometries and Energetics 101

also somewhat shorter, as measured by R(O O), and more linear with values of (OH O)
closer to 180°.
The energetics of the complex are reported in Table 2.40176-182. The results indicate that
the electronic contribution to the binding energy is of the order of 12-13 kcal/mol with the
DZP and TZ2P basis sets. The effect of correlation is inconsistent from one basis to the next.
Note also that the results in Table 2.40 have been corrected for basis set superposition er-
ror. Uncorrected data are inflated by as much as 3-4 kcal/mol, nearly 8 kcal/mol at the MP2
level. After addition of zero-point vibrational energies and other factors, the binding en-
thalpies at 300 K are computed to lie in the range of 10.4-11.7 kcal/mol. These results have
not yet reached convergence. When a larger basis set is applied to the problem, the calcu-
lated binding enthalpy reaches 13.1 kcal/mol. The latter result is in reasonable coincidence
with experimental estimates that are in the 14-15 kcal/mol range.
From their explorations with various geometries, the authors conclude that a MP2 cal-
culation of the energy, using a geometry optimized at the SCF level, will typically lead to
errors of less than 1 kcal/mol for this dimer. They recommend a set of DZP quality for this
identification of an appropriate geometry. Two sets of polarization functions should provide
accuracy of 0.3 kcal/mol in the SCF binding energies (provided counterpoise corrections
are made). Correlation is more resistant to convergence; diffuse and semidiffuse d and f
functions can be recommended for quantitative accuracy.
From computations with basis sets of the polarized 6-31G type, at SCF and MP2 levels,
the change from formic acid dimer to acetic acid dimer has only a very minor influence
upon the energetics of binding183. This conclusion is indeed verified by resonant laser pho-
toacoustic spectroscopic measurements179,181,182 which find further that trifluoroacetic and
propionic acid dimers have a very similar binding enthalpy to formic dimer. One may ex-
trapolate that the formic acid dimer H-bond energy is probably applicable for most car-
boxylic acids with alkyl chains replacing the CH group.

2.9 Nitrile

Another bonding arrangement of the nitrogen atom involves the nitrile group, as in HC=N.
As the single lone pair of the nitrogen atom is collinear with the rest of the molecule, as
well as with the molecular dipole moment, there is little question as to the angular prefer-
ences for H-bonding. It is interesting to compare and contrast the H-bonds formed by this

Table 2.40 Energetics of binding (kcal/mol) in the formic acid


dimer. Data corrected for BSSE176.

Basis set - Eelec, - H(300° K)

DZP SCF 12.9 11.2


MP2 12.0 10.4
TZ2P SCF 12.1 10.5
MP2 13.3 11.7
VQZ2PP MP2 14.7 13.1
expt a,b 14.1-15.2

"See references 177-181.


b
Result for acetic acid dimer is 15.0 kcal/mol l82
102 Hydrogen Bonding

group with amines wherein the nitrogen has three separate bonding partners. Moreover, the
hydrogen in HCN is acidic enough that the molecule may act as an effective proton donor.
Several H-bonded complexes involving HCN were optimized at the SCF level with a
doubly polarized triple- basis set by Somasundram et al.184. When paired with NH3, HCN
acts as proton donor, with R(C-N) = 3.292 A. HF is the proton donor in FH-NCH, with
R(F-N) = 2.878 A. The latter bond length is only slightly longer than the value of 2.848 A
obtained in an earlier calculation185 with a polarized triple- basis set, with electron corre-
lation treated by an approximate coupled cluster approach186. The computed binding en-
ergy of FH---NCH agrees quite nicely with a high resolution FTIR measurement of 6.9
kcal/mol for Dc187. While FH---NCH is the more stable, it is also possible to trap the reverse
complex, NCH---FH, in Ar matrix188. A dimer of HCN is fully linear and the C and N atoms
are separated by 3.307 A. Binding energies for NCH---NH 3 and FH---NCH were reported to
be 7 kcal/mol, in comparison to the weaker binding of 5 kcal/mol for the HCN dimer.
A later calculation189 optimized the geometry of NCH---NH 3 at the correlated MP2 level
and with a much enlarged [6-31 +G(3df,2p)] basis set and confirmed its C3v structure. The
R(C-N) distance is 3.139 A in the complex, which is bound by 6.06 kcal/mol relative to the
isolated monomers, following counterpoise correction. This binding energy drops to 4.37
kcal/mol when zero-point vibrational effects are included. Curiously enough, a second min-
imum was identified on the PES pairing HCN with NH3. A geometry in which one of the
H atoms acts as a bridge to the N atom of HCN is a local minimum on the MP2 correlated
surface. This complex is much more weakly bound than the conventional structure where
NH3 is the proton acceptor. The binding energy of the structure in Fig. 2.18 is only 1.1
kcal/mol at the SCF/6-31G* level, with counterpoise corrections. It is likely not to be bound
at all following incorporation of vibrational effects. The latter geometry has not been de-
tected in experimental studies of this complex190,191 which have confirmed the presence of
a C3v geometry for NCH-NH3.
Replacement of NH3 by PH3, maintains the linear character of the H-bond in NCH--PH3,
although R(C P) is lengthened to 3.913 A in a microwave spectrum192,193. The presence
of a true H-bond is also questionable in NCH---SH 2 where a very long R(C S) of 3.809 A
is observed194. Using measured intermolecular stretching force constants as a criterion, the
interaction in this complex is the weakest of any complex of the type AH---YH2 (A=NC,
Cl, F;Y=O,S) 194 .
SCF/6-31G** calculations by Boyd195 showed that substitution of a methyl group for
the hydrogen of HCN increases the binding energy with HC1 from 4.3 to 5.3 kcal/mol, sug-
gesting the larger molecule is a better proton acceptor. Del Bene et al.196 also paired HC1
with acetonitrile. The geometry optimized at the MP2/6-31 + G(2df,2pd) level contains a
fully linear H-bond, with a R(Cl--N) distance of 3.316 A. The latter distance compares with
3.301 A measured by microwave spectra197. Applying MP4 to this geometry yields a bind-
ing energy, De, of 5.9 kcal/mol, lowered to 5.2 after correction for BSSE. This result is in
excellent agreement with a FTIR photometric measurement of 5.2 kcal/mol198. As ex-
pected, the H-bond is stronger in FH---NCCH3, with a binding energy measured to be 6.9

Figure 2.18 Alternate geometry for HCN + NH3.


Geometries and Energetics 103

kcal/mol199. In fact, computations suggest certain other substitutions might further strengthen
this interaction, for example, replacing NCCH3 by NCLi200.
The dimer of HCN was studied in more detail, with much improved basis sets, ranging
up near Hartree-Fock quality. Specifically, the largest set employed by Kofranek et al.201
was doubly polarized on C and N, and had a single set of p-functions on H. The complex is
fully linear, as illustrated in Fig. 2.19.
Beginning with the SCF data, the salient geometric and energetic aspects of this dimer
are listed in the upper part of Table 2.41201,202. Enlargement of the basis set leads to a pro-
gressive lengthening of the intermolecular distance and weakening of the binding energy.
The H-bond computed with the biggest basis set is too long by some 0.1 A, not surprising
since the attractive effects of correlation have not been considered. One can see also the usual
stretching of the bridging hydrogen away from the C atom, and small changes in the inter-
nal C N bonds. These trends are duplicated when correlation is included, and the inter-
molecular distance comes closer to the experimental value, as do the energetic quantities.

2.10 Imine

Intermediate between the N atom of amines, which is involved in single bonds, and the
triply bonded N in nitriles, lie the imines with their double bonds. Recent calculations203
have combined methyleneimine, and its methylated derivatives as proton acceptor, with wa-
ter and with methanol as donor. The general arrangement of water with methyleneimine is
illustrated in Fig. 2.20 where it is emphasized that the hydrogen of the water not partici-
pating in the H-bond prefers to lie outside of the molecular plane of H2C=NH. The same
is true of the methyl group when water is replaced by methanol.
The sensitivity of the energetic aspects of the H-bonds to the level of theory is reported
in Table 2.42. The inclusion of polarization functions reduces the interaction energy some-
what, probably due largely to reduction of BSSE. With these d-functions included, there is
little difference between water or methanol as proton donor. Correlation enhances the bind-
ing strength, as noted in other systems. Comparison with other complexes is difficult as the
authors did not remove the BSSE from their data. On the other hand, the authors have com-
piled a list of thermodynamic quantities for a number of related systems, all at the SCF/6-
31G level, so there is some basis for comparison here.
Comparison of the first two rows of Table 2.43 indicates that the doubly bonded N in the
imine is a better proton acceptor than is the doubly bonded carbonyl oxygen in formalde-
hyde. This trend is consistent with the single bonded analogues wherein amines are more
basic than hydroxyls. The succeeding rows address methyl substitution at various sites. It
appears that methyl substitution at either the proton-donating water or proton-accepting ni-
trogen slightly diminishes the H-bond energy. The two effects reinforce one another when
both are alkylated. In contrast, the H-bond is strengthened when one of the hydrogens of
the CH2 group of the imine is replaced by methyl. Unlike the carbonyl bond, which stretches
by around 0.005 A upon forming a H-bond, the double bond between C and N in the imine
stretches by less than 0.002 A.

Figure 2.19 Linear HCN dimer.


Table 2.41 Optimized geometries (A) and energetics (kcal/mol) of the HCN dimer. Data201,202 contain energetics not corrected for BSSE. See
Fig. 2.19 for atomic labeling scheme.

Basis set R(C N) r(CdHd) r(CdNd) r(CaAa) - Eelec -AH

[53/3] SCF 3.293 0.006 0.001 -0.002 5.60 4.72


[64/4] 3.345 0.007 0.001 -0.001 4.71 —
[641/41] 3.392 0.005 0.002 -0.001 4.39 3.59
[752/41] 3.405 0.005 0.001 -0.001 4.19 3.39
[752/421 3.401 0.005 0.001 -0.002 4.21 3.42

[641/41] CPF 3.364 0.005 0.000 -0.001 4.37 3.63


[752/41] 3.342 0.005 0.000 -0.001 4.35 3.61
[752/421 3.331 0.005 0.000 0.000 4.54 3.80
expt 3.29 4.4 2.7-3.8
Geometries and Energetics 105

Figure 2.20 H-bond arrangements in methyl-


eneimine-water complex.

2.1 I Amide

The prevalence of amide groups as structural elements in proteins has motivated a good deal
of interest in its H-bonding ability204. An early calculation, using unpolarized minimal and
split-valence basis sets205 noted the near equivalence of the strengths of the amide-amide
and amide-water H-bonds. Other computations illustrated the sensitivity of the formamide
dimer binding energy to the particular basis set. A range of 10 to 23 kcal/mol for this quan-
tity was obtained for basis sets varying from minimal (STO-6G) to 6-31G*206,207. These
values were undoubtedly inflated by the failure to correct them for superposition error.
There are various ways in which two amide molecules can come together to form a H-
bonded complex. Ostergard et al.208 assumed the general configuration shown in Fig. 2.21
and carried out geometry optimizations at various levels.
The data in Table 2.44 illustrate the stretching of the intermolecular separation that is
typically seen as small basis sets are enlarged. The 5 angle which characterizes the place-
ment of the proton donor relative to the carbonyl oxygen is rather sensitive to the choice of
basis set or to inclusion of correlation. The orientation of the donor molecule, denoted by
|3, is clearly in the range between 116° and 122°. None of the interaction energies are at a
high enough level to be considered definitive, particularly since no BSSE corrections have
been made. SCF computations with the 6-31G** basis set209 indicate that the interaction
energy is rather insensitive to the 8 angle and there is little dependence on the NH---O an-
gle, provided it remains within about 30° of linearity.
Again taking formamide as a model amide, other calculations examined its modes of
binding to water210. Four separate minima were located on the potential energy surface,
generated via full geometry optimizations with DZ and DZP basis sets. The most stable
structure found is illustrated as I in Fig. 2.22. It contains two H-bonds in a cyclic arrange-
ment, permitting H2O and HCONH2 to act as both proton donor and acceptor simultane-
ously. The carbonyl oxygen is proton acceptor in Structure II, whereas the NH proton is

Table 2.42 Binding energies (— Eelec in kcal/mol) of complexes


of memyleneimine with water and methanol. BSSE was not
removed in data203.
SCF MP2

6-31G 6-31G* 6-31G 6-31G*

H 2 C=NH-HOH 8.2 6.1 9.2 8.2


H 2 C=NH--HOCH 3 7.8 6.2
106 Hydrogen Bonding

Table 2.43 Thermodynamic quantities calculated203 for various H-bonded complexes. All
entries in kcal/mol, except AS in cal mol-1 deg - 1 .
Eelec AH° AS AG°

H 2 C=O HOH -6.7 -4.9 -24.2 2.3


H 2 C=NH-HOH -8.2 -6.2 -25.9 1.5
H2C=NH-HOCH3 -7.8 -6.0 -27.2 2.0
H2C=NCH3--HOH -7.9 -6.0 -25.8 1.7
H2C=NCH3 HOCH3 -7.6 -5.8 -27.1 2.3
CH3CH=NCH3--HOH -8.7 -6.7 -27.3 1.4
CH3CH=NCH3--HOCH3 -8.3 -6.5 -28.6 2.0
CH3CH=NH-HOH -9.0 -6.6 -27.2 1.0
CH3CH=NH-HOCH3 -8.7 -6.9 -28.7 1.7

donor in III and IV. The results of the calculations, and the existence of geometry I, were
later confirmed by microwave spectral data211. The experimental structure contains slightly
shorter H-bonds than calculated, but is otherwise quite similar. The complex pairing for-
mamide with HF is again similar to I in that the primary H-bond has HF donating a proton
to the carbonyl oxygen, with a secondary interaction between F and the NH group212.
The electronic contributions to the binding energy, — Eelec, for the various geometries
are listed in Table 2.45, computed at both the SCF and correlated (SCEP) levels210. The re-
sults are apt to be inflated since BSSE was not removed. It is nonetheless clear that the pres-
ence of two H-bonds in Structure I makes it the most stable by 2-3 kcal/mol. The carbonyl
appears to be a slightly better proton acceptor than the NH2 group is a donor, since geom-
etry II is more stable than III and IV. The similarity of the latter two indicates there is little
difference between the syn and anti hydrogen atoms.
There is some question concerning the ability of peptide groups to form H-bonds in aque-
ous versus apolar solvents213,214. A recent effort215 compares the strengths of H-bonds be-
tween peptide groups (modeled by N-methylacetamide, NMA) and water. The former dif-
fers from formamide principally in eliminating some of the possibilities of H-bonding to
one of the NH2 hydrogens, so a cyclic structure like I in Fig. 2.22 is excluded. Cognizant
of the similarities in strength between these sorts of bonds, the authors employed a polar-
ized double- basis set for full geometry optimizations and extraction of vibrational ener-
gies; energies were then computed with a much larger basis set at the correlated MP2 level,
and BSSE corrections evaluated.
Several configurations were considered in which a water molecule is paired with NMA.
The first pair of structures illustrated in Fig. 2.23 has the two methyl groups of NMA trans

Figure 2.21 Formamide dimer, illustrating


the definition of two intermolecular angles.
Geometries and Energetics 107

Table 2.44 Optimized geometries (A and degs) of formamide dimer and


interaction energy (kcal/mol)208; parameters defined in Fig. 2.21.

R(O-N) 5 - Eelec

SCF
STO-3G 2.778 122.4 121.8 5.7
4-31G 2.949 163.7 116.1 8.4
6-31G** 3.074 143.1 118.7 6.3
6-31 + +G 3.089 158.2 118.9
MP2
4-31G 2.988 126.3 117.7 8.8
6-31G** 3.016 125. 121.7 —

to one another, while they are cis in the second two geometries. In WTd, the trans NMA
molecule acts as proton donor to the water molecule, while it is the acceptor in WTa.
NMA is the acceptor in WCa, but the proximity of the carbonyl oxygen and NH group of
NMA in its cis geometry permits a pair of H-bonds to be formed with a single water mol-
ecule. NMA is hence simultaneously proton donor and acceptor in WCad. The water-NMA
H-bond is compared with the bond between a pair of amide units in the last two structures
in Fig. 2.23. A single H-bond is formed when two trans geometries are paired in TdTa
whereas a pair of cis molecules can form two H-bonds such that each molecule is simulta-
neously donor and acceptor, as in CdaCda.

Figure 2.22 Structures of four minima identified on H2O +


HCONH 2 PES.
Table 2.45 Binding energetics (— Eelec in kcal/mol)
of complexes of formamide with water. Data
computed with DZP basis set and uncorrected
for BSSE210.

SCF SCEP

I -7.9 -9.5
II -5.8 -6.7
III -5.2 -6.0
IV -5.2 -6.2

Figure 2.23 Configurations pairing NMA with water or another molecule of


NMA.
Geometries and Energetics 109

Table 2.46 lists the energetics of binding for the various complexes. Comparison of the
first two rows indicates that the amide interacts more strongly with a water molecule when
the amide is the proton acceptor, that is, water is not a good proton acceptor. The similarity
of the second and third rows suggests that proton donation to the carbonyl oxygen is unaf-
fected by the cis or trans character of the amide. (Other calculations had indicated a simi-
larly small sensitivity of the H-bonding interaction to the rotations of the two methyl
groups216.) However, the cis arrangement provides another option to the complex in which
there are two H-bonds present. The stronger binding of the WCad configuration confirms
that despite any angular distortions, these two H-bonds are cumulatively stronger than a sin-
gle H-bond. When a pair of amides interact, the single H-bond in TdTa is similar in strength
to that in WTa or WCa. In other words, the N—H of the amide is comparable to HOH as a
proton donor. Again, the appearance of a second H-bond, as in CdaCda, strengthens the in-
teraction, this time by quite a bit. The binding energy in the latter configuration is roughly
twice that of TdTa where there is only one such bond. This near additivity is probably due
in part to the low degree of distortion of each H-bond in the NMA dimer, compared to the
bent H-bonds when one of the molecules is the smaller water.
The aforementioned trends are noted at either SCF or MP2 levels, and apply not only to
the electronic segment of the binding energy, but also to AH where vibrational contributions
have been considered. The authors conclude that a single amide-amide H-bond is similar in
strength to that between an amide and a water molecule. If one is interested in a binary com-
plex, there is a preference for the syn geometry of NMA, as it then becomes possible to form
more than one H-bond.
Of course, this preference would be removed as additional molecules are added to the
system, as discussed by Guo and Karplus217. This reasoning is amplified by inversion trans-
fer 13C NMR spectroscopic data of model systems204 which indicate that the amide-water
interaction is stronger than that between a pair of amides. (The work points out the anom-
alous character of formamide which may not make it a good model for the peptide units in
proteins.) Since the data suggest that amides form stronger H-bonds with water than with
other amides, the prevalence of peptide-peptide H-bonding in folded proteins is attributed
to the cooperativity that occurs as these peptide units form multiple H-bonds. (See Chapter
5 for a detailed discussion of the cooperativity phenomenon.) As a last parenthetical note,
there is a recent set of calculations218 that lead to the intriguing suggestion that interaction

Table 2.46 Binding energetics of complexes of amide with water


or another amide (see Fig. 2.23 for definitions). All data computed
with aug-cc-pVDZ basis set and corrected for BSSE, in
kcal/mol215.

- Eelcc - H298

SCF MP2 SCF MP2

WTd 3.5 4.8 2.0 3.3


WTa 5.7 7.0 4.0 5.3
wc. 5.6 7.0 3.9 5.2
wcad 6.8 8.9 4.9 7.0
T
dTa
5.4 7.2 3.8 5.6
cdacda 10.5 14.0 8.8 12.3
110 Hydrogen Bonding

of an amide with multiple water molecules will have a large effect on its internal geometry
in the excited * electronic state, and further, that this effect will be threefold larger and
in an opposite sense to geometry changes that occur in the ground state.
As reported above for the pair of N-methylacetamide molecules, two amides can form
a cyclic complex containing two H-bonds as the favored geometry. In the case of NMA, the
stability of this cyclic structure is very nearly equal to twice the binding energy of each in-
dividual H-bond. This question was examined for formamide as well176 and the results re-
ported in Fig. 2.24 and Table 2.47. Fig. 2.24a illustrates the two equivalent H-bonds within
the dimer in its optimized geometry. This type of geometry is consistent with the crystalline

Figure 2.24 Geometries of the formamide dimer, com-


puted at the SCF/DZP level 176 . Bond lengths refer to
distances between nonhydrogen atoms.
Geometries and Energetics I !I

Table 2.47 Binding energies (— Eelec in kcal/mol) of various


geometries of formamide dimer (see Fig. 2.24). Data calculated
with DZP basis set, corrected for BSSE176.
a b e d

SCF 10.6 5.6 7.1 3.8


MP2 11.4 5.6 7.1 4.0

structure of formamide219-221, and the computed H-bond length is only slightly longer than
the solid-phase value of 2.93 A.
Another study has examined the effect of level of theory upon the H-bond length in this
structure222. Correlation typically contracts the H-bond, but the amount of this reduction is
sensitive to basis set quality. For example, the SCF and MP2 H-bonds are nearly identical in
length for a double- basis set, whereas simply adding d-functions results in a correlation-
induced contraction of 0.08 A. The magnitude of this contraction generally grows as the ba-
sis set is further improved, reaching as high as 0.18 A for TZ(2df,p). Once polarization func-
tions have been added, the H-bond length is reasonably stable with respect to changes in
basis set222. At the SCF level of optimization, R(N--O) remains in the range between 2.99
and 3.02 A as the basis set is enhanced from DZ(d) to TZ(2df,2pd); the MP2 range is
2.84-2.92 A.
The NH--O bond stretches by 0.06 A when the two molecules are disposed as in a so that
only a single H-bond can form between them176. The binding energy of the cyclic dimer is
not quite twice that of the single H-bond in b, indicating little apparent cooperativity in the
energetics in this case. This result is consistent with the data described above for the di-
methylated amides215. The MP2 binding energy of the cyclic formamide dimer (a) of 11.4
kcal/mol, compares with a comparable calculation with the 6-31G** basis set223, which led
to a value of 12.4 kcal/mol. Following correction of this result by ZPVE, the enthalpy of
binding at 0° K comes to 9.6 kcal/mol. Another computation evaluated the binding enthalpy
at 298 K of the formamide dimer to be 12.1 kcal/mol224 at a fairly high level of theory:
MP2/6-311 + +G(2d,p), using a geometry optimized at the MP2 level. This value, however,
likely overestimates the true value as the authors failed to correct for BSSE. Based on an-
other set of computations222, enlargement of the basis set beyond the polarized double-
type is unlikely to have much influence on the computed binding energy.
Unlike NMA, formamide contains a CH group which has the potential to act as a pro-
ton donor in a H-bond. The ability of the CH group to substitute for N—H in this manner is
explored via structure (c)176. The R(C O) distance is rather long, and the (NH--O) distance
is 0.01 A longer than in structure (a). The weaker binding in (c) as compared to (a) is con-
firmed by the energetic data in Table 2.47 where the (CH--O) interaction adds only about
1.5 kcal/mol to the binding energy of a single H-bond (b). The longest H-bonds of all oc-
cur when both H-bonds are of the (CH--O) variety in (d). Note that, even summed together,
these two H-bonds are cumulatively weaker than a single (NH--O) interaction, as in (b).
It is worth mentioning that the relative populations of linear and cyclic structures can be
shifted by temperature. A recent work which couples experimental measurements with ab
initio calculations of quadupole coupling constants225 determined that rings of six for-
mamide molecules dominate in the liquid state at low temperatures, but are replaced by lin-
ear tetramers as T approaches 400° K.
112 Hydrogen Bonding

2.11.1 Interaction with Carboxylic Acid and Ester


The NH2 group of formamide contains two hydrogens which could be donated in a H-bond,
and the oxygen atom could serve as a proton acceptor. When paired with formic acid, which
also contains a separate proton donor and acceptor group, it is not surprising to see two H-
bonds formed, within an eight-membered ring, as illustrated in Fig. 2.25a176. Since oxygen
is more electronegative than nitrogen, and formic acid more acidic than formamide, the fact
that the R(OH--O) H-bond is shorter than R(NH--O) by some 0.25 A is also not unexpected.
By considering the open (noncylic) structures depicted in Fig. 2.25b and 2.25c, it was pos-
sible to assess the relative strengths of these two H-bonds. Comparison with Fig. 2.25a in-
dicates that the cyclization of the complex, that is, the formation of two simultaneous H-
bonds, results in a contraction of each. R(O-O) is shorter by 0.05 A in the cyclic complex
and R(N-O) shorter by 0.10 A.

Figure 2.25 Geometries optimized for com-


plex of HCOOH + HCONH2, at SCF/DZP
level 176 . Bond lengths refer to distances be-
tween nonhydrogen atoms.
Geometries and Energetics II3

The energetics, too, illustrate the cooperative nature of the two H-bonds. As reported in
Table 2.48, the formamide and formic acid molecules are bound together in the cyclic com-
plex by some 12.6 kcal/mol at the MP2/DZP level. The (OH-O) H-bond of structure (b)
contributes 7.0 kcal/mol while 3.8 kcal/mol more arises from the (NH--O) interaction. To-
gether, these two separate H-bonds add up to less than the full interaction in the cyclic struc-
ture. (Structures (b) and (c) do not represent true minima on the potential energy surface.)
It is interesting to note that correlation plays little apparent role in the computed binding
energies of this particular complex.
In addition to the eight-membered cyclic structure of Fig. 2.25a, one might envision a
7-membered ring such as pictured in Fig. 2.25d in which the NH2 proton-donating group is
replaced by the CH group. This structure is computed to be 3.0 kcal/mol less stable than
the global minimum (a). On the other hand, the (CH--O) interaction does appear to add
something to the (OH--O) H-bond. Complex (d) is bound 2.6 kcal/mol more strongly than
(b); moreover, formation of the cycle shortens R(OH--O) by 0.04 A. Structure (e) substi-
tutes NH2 for =O as the proton-accepting group in the other H-bond, also representing a
7-membered ring. This complex was found to be unbound, relative to the pair of isolated
monomers, and does not represent a minimum on the PES.
The studies of formic acid and formamide suggest that the CH group can act as a proton
donor. While there is some question as to whether it constitutes a true H-bond, there are
clearly certain stabilizing interactions present.
Formamide was paired with a small ester, methyl acetate, at the SCF and MP2 levels
in order to compare the strength of this interaction with that between a pair of amides226.
Three minima were identified on the surface, all containing a H-bond wherein the NH2
group of formamide acts as donor. The optimal acceptor is the carbonyl oxygen atom, as
compared to the alkoxy oxygen of the ester. As reported in Table 2.49, the former H-bond
is shorter than the latter by 0.14 A and is more linear. The energetics indicate an approxi-
mate 2 kcal/mol preference for the carbonyl oxygen atom as acceptor. Note also that cor-
relation appears to contribute perhaps 3 kcal/mol to the total interaction energy of the
complex.

2.12 Nucleic Acid Base Pairs

The nature of the interaction between nucleic acid bases has been a continual source of fas-
cination ever since the nature of the genetic code was unraveled. However, the numbers of
atoms and electrons in these systems erected a high barrier to the application of accurate
quantum mechanical methods for many years. For example, the guanine-cytosine (GC) pair
contains 29 atoms and 136 electrons.

Table 2.48 Binding energies ( — Eelec in kcal/mol) of complexes


of formamide with formic acid. Data calculated with DZP basis
set, corrected for BSSE176.
A B C D E

SCF 12.4 6.9 4.0 9.8 unbound


MP2 12.6 7.0 3.8 9.6 unbound
114 Hydrogen Bonding

Table 2.49 Hydrogen bonding parameters calculated


with 6-31G* basis set for complex pairing
formamide with carbonyl and alkoxy oxygen atoms
of methylacetate226. Geometry optimized at SCF
level. Counterpoise corrections for BSSE added to
all energies, reported in kcal/mol.

Carbonyl Alkoxy

R(N-O)(A) 3.073 3.215


(NH-O) (degs) 177.7 173.1
- Eelec (SCF) 5.9 2.9
- H298 (SCF) 3.9 1.4
- Eelec(MP2) 8.1 5.6
- H298 (MP2) 6.1 4.0

From semiempirical227-230 to ab initio231-236 the levels of treatment of this interaction


over the years parallel the development of computational quantum chemistry. Some of the
recent computations were able to optimize the molecular geometries using gradient proce-
dures237-240 while others have fit the results of ab initio calculations on smaller systems to
empirical functions so as to then study the full nucleic acid base pairs241-242. There has also
been work that relates the effect of adding or removing an electron to the base pair upon the
nature of the interaction243,244.
We will focus here on the "state of the art." That is, it is now possible in practice as well
as in theory to perform reliable ab initio calculations of a nucleic acid base pair. By this, it
is meant that polarized basis sets can be used, and the effects of electron correlation can be
included explicitly. A case in point is a recent series of computations that compared the sta-
bility of 30 different arrangements of DNA base pairs245. A 6-31G** basis set was used to
obtain gradient-optimized geometries, including normal harmonic vibrational modes and
energies. This vibrational analysis also confirmed the presence of a minimum or stationary
point of higher order. The interaction energies were then computed using the MP2 method
to include correlation, at the SCF-optimized geometries. Counterpoise corrections were
added at each level. Addition of the ZPVE led to an estimate of the binding enthalpy of each
pair.
To provide some estimate of the scale of these computations at this point in time, each
SCF geometry optimization, including vibrational analysis, required some 85-430 hours of
CPU time on a Cray Y-MP supercomputer. Computation of the MP2 energy at a single
geometry is much quicker, needing only 1-5 hours. As a push toward the outer limit of cur-
rent technology, optimization of the geometry of the cytosine dimer at the MP2 level re-
quired on the order of several weeks of CPU time. The authors estimated that were they to
attempt to compute electron correlation effects by the more time-consuming CCSD(T) ap-
proach on the cytosine dimer, enforcing C2h symmetry and using a 6-31G* basis set lack-
ing polarization functions on hydrogen, such a calculation would require days of CPU time,
25 GB of scratch disk space, and one full GB of memory.
Table 2.50 lists the binding energies of 13 of the 30 pairs studied in this work245,246. For
any given pair of bases, G and C for example, only the most strongly bound is included in
the table. The first two columns report the electronic contributions to the binding at the SCF
Geometries and Energetics 115

Table 2.50 Calculated energetics (kcal/mol) of binding of selected nucleic acid


base pairs. Data245 computed with 6-31G* basis set.

- Eelec -AH

SCF MP2 calc exptb

GCWC 23.4 23.8 21.9 21


GG 22.5 22.2 21.5
CC 15.9 17.5 15.5 16
GAa 11.5 14.1 13.3
GTa 12.6 13.9 13.3
AC 11.0 13.5 11.7
TAH 10.2 12.7 11.4 13
TARH 10.2 12.6 11.3
TAWC 9.6 11.8 10.5
TARWC 9.5 11.7 10.4
AA 8.1 11.0 9.3
TCa 8.3 10.7 9.5
TT 8.4 10.0 9.1 9
a
Not a true minimum; one negative eigenvalue in Hessian.
Train reference 246.

and MP2 levels, respectively, followed by the full enthalpy (using the MP2 binding energy)
in the last columns. There are various ways that the bases can be paired together, main-
taining good H-bonding contacts. In addition to the Watson-Crick arrangement, abbreviated
as WC by the authors and illustrated in Fig. 2.26 for G—C and A—T, there is also reverse
Watson-Crick (RWC), Hoogsteen (H), and reverse Hoogsteen (RH). These conformations
are depicted for the adenine-thymine pair in Fig. 2.27.
The first row of Table 2.50 indicates that the Watson-Crick pairing of guanine and cy-
tosine is predicted to be the most strongly bound of any pair, with a AH of — 21.9 kcal/mol.
There are three H-bonds in this complex: One NH---O H-bond length is 3.02 A and the other
is 2.92 A; R(NH-N) = 3.02 A. Another factor contributing to the strength of the interac-
tion is the very near linearity of the placement of the bridging proton: all three H-bond an-
gles are within 4° of 180°. Although it only contains two H-bonds, the GG pair is only
slightly less strongly bound than GC. This is probably due in part to the large dipole mo-
ment (7.1 D) computed for G and the ability of the two dipoles to align in an antiparallel
fashion in the GG complex. Also, the two H-bonds are short and linear; both are of NH---O
type with a bond length of 2.87 A and angle of 178°. It might be noted that both the GCWC
and GG pairs have only minor correlation contributions to their total binding energies. Like
GG, the CC pair also contains only two H-bonds. But the binding enthalpy is significantly
smaller, down to 15.5 kcal/mol. This weaker binding may be due to the lesser proton-
donating ability of the NH group in the two NH--N H-bonds, each of which are 3.05 A in
length.
The GA and GT combinations are somewhat less strongly bound. It was noted by the
authors that these structures are not true minima. The single negative eigenvalue probably
refers to an out-of-plane motion of a side group (the authors enforced full planarity), and
so is probably not a serious concern to this discussion. These two dimers have in common
the presence of a pair of H-bonds; nitrogen serves as the proton donor in all of these.
116 Hydrogen Bonding

Figure 2.26 Watson-Crick pairing of guanine-cytosine and


adenine-thymine. Only hydrogens involved in H-bonds
are shown explicitly.

Of particular interest is the comparison of the four types of TA dimers that were exam-
ined. The calculations245 indicate a clear preference for Hoogsteen pairing over Watson-
Crick; in both cases, it does not matter whether these are standard or reverse form. One can-
not attribute the discrepancy to the type of H-bonds present. In all four instances, there is
one NH—N and one NH---O bond. The adenine homodimer and thymine homodimer are
both slightly less strongly bound.
In summary, the purine-pyrimidine pairs follow the trend in binding energy:

If one compares purine/purine or pyrimidine/pyrimidine pairs:

Restricting the latter to homodimers only:

The authors stress that the ability of G and C to form the strongest interactions with other
bases may be traced to their large dipolc moments. They summarize that neither the total
number of H-bonds, nor their length or linearity, can be taken as sole arbiters of the strength
of the overall interaction, but that all factors must be considered together.
Geometries and Energetics 117

Figure 2.27 Reverse Watson-Crick (RWC), Hoogsteen (H),


and reverse Hoogsteen (RH) pairings of adenine with
thymine.

As a final test of the influence of correlation, the CC geometry was reoptimized at the
MP2 level, using a 6-31G** basis set245. The resulting changes, with respect to the Hartree-
Fock structure, accounted for a further stabilization of only 1.1 kcal/mol, or a 6% increase,
despite the contraction of one of the H-bonds from 3.05 to 2.92 A, a full 0.13 A. It was noted
that an estimate of the dispersion energy by the empirical London formulation does not re-
produce the correlation contribution to the interaction as computed by MP2.
Comparison of the computed binding enthalpies in Table 2.50 with the experimental
quantitites in the last column 246 confirm the ability of ab initio methods to treat such ex-
tended systems with a high degree of accuracy. The experimental enthalpies were obtained
118 Hydrogen Bonding

by mass spectrometry and provide no information about the actual geometries. The corre-
spondence with the calculated structure is made under the presumption that the structure
observed experimentally is in fact the most stable computed geometry. Not only do the cal-
culations correctly predict the relative order of the GC, CC, TA, and TT pairs, but the bind-
ing enthalpies are within about 1 kcal/mol.
A similar set of calculations was limited to the Watson-Crick structures of GC and AT,
and included also a Hoogsteen conformation of AT236. Although the computations were car-
ried out at the MP2 level, counterpoise corrections were limited to SCF only. Their best
computed enthalpies of interaction are all somewhat greater than those in Table 2.50. The
two studies agree, however, in the relative ordering. The Watson-Crick GC pair is most
strongly bound, and the Hoogsteen AT pair is bound by about 1 kcal/mol more than the
Watson-Crick geometry.
While one might expect that H-bonded conformations of nucleic acid base pairs would
be most stable, there is also the interesting possibility of stacked structures in which the
planes of the two bases are parallel to one another. Correlated calculations at the MP2/6-
31G* level suggest these stacked geometries are less stable than the H-bonded structures247.
Taking the cytosine dimer as an example, the interaction energy of the stacked structure is
roughly half that of the H-bonded conformer. While a large fraction of the binding energy in
the more stable structure originates in Coulombic attraction, it is dispersion which is chiefly
responsible for the binding in the stacked structure. Since dispersion is relatively insensitive
to the mutual orientations of the two molecules within their respective parallel planes, the
anisotropy of the electrostatic term guides the two molecules into their optimal positions.

2.13 H-Bonds versus D-Bonds

The question sometimes arises as to whether a hydrogen bond is stronger or weaker than a
deuterium bond. This question might better be posed in pairing a HF molecule with DF:
will FH--FD be more or less stable than FD---FH? Since H and D are chemically indistin-
guishable, the electronic part of the energies of these two dimers will be identical. The dif-
ferences are associated with the masses, and thence to the vibrational energies. The points
can best be illustrated by analyses of the vibrational frequencies of FH---FD and FD---FH.
The data calculated at both the SCF and MP2 levels248 are reported in Table 2.51, along
with the total zero-point vibrational energy (in c m - 1 ) in the last row.
The first two rows list the intramolecular HF and DF stretching frequencies. It is obvi-
ous that the HF stretch goes down if this molecule is the proton donor. On the other hand,
a red shift of comparable magnitude occurs in the DF stretch when DF acts as donor. Con-
sequently, these two effects pretty much cancel one another, as may be seen in the third row
of Table 2.51 where the total intramolecular zero-point vibrational energies of FH---FD and
FD---FH are very nearly identical, within 10 cm - 1 .
The source of the difference in energy between these two isotopic isomers may be traced
to the intermolecular frequencies in the lower part of Table 2.51. The intermolecular stretch-
ing frequencies of FH---FD and FD---FH are close to one another, as are the first of the two
in-plane bends. The biggest discrepancies occur in the next two intermolecular bending
modes. Both of these represent primarily the wagging motion of the bridging hydrogen. In
the case of the D-bond, that is, FD---FH, it is a heavy D nucleus which is moving, so the fre-
quency is reduced relative to the case of FH---FD where it is protium that acts as the bridge.
The reductions in each of these bending frequencies are of the order of 100 c m - 1 . As a re-
Geometries and Energetics 119

Table 2.51 Vibrational frequencies and zero-point vibrational energies (cm - 1 ) of the two iso-
topic dimers248.

SCF MP2

FH-FD FD-FH FH-FD FD-FH


HF stretch 4371 4418 3993 4073
DF stretch 3202 3169 2952 2895
intramolecular ZPVE 3787 3794 3473 3484
iritermolecular stretch 126 138 149 163
in-plane bend 169 183 194 210
out-of-plane bend 439 319 500 362
in-plane bend 495 407 568 466
inter-molecular ZPVE 615 524 706 601
total ZPVE 4401 4316 4178 4084

sult, the total vibrational energy of the D-bonded complex is lower by about 100 cm-1 com-
pared to the H-bond.
The total difference in zero-point vibrational energies of the FH---FD and FD---FH com-
plexes, displayed in the last row of Table 2.51, is 85 cm-1 at the SCF level, and 94 c m - 1
at MP2. It is interesting that this difference is relatively insensitive to correlation. The size
of the basis set is not crucial either, as earlier calculations with a smaller basis set249 had
obtained a value of 109 c m - 1 . While the potential energy surface of the dimer can perhaps
be calculated reasonably well, the biggest source of error in this analysis is the assumption
of harmonic frequencies. Analysis of high resolution near-IR data for the Cl analogue250 is
consistent with a "stronger" D-bond here as well: C1D---C1H is more stable than C1H---C1D
by 16±4 c m - 1 .
Despite the errors, the calculations permit a simple interpretation of the greater stability
of D-bonds versus H-bonds. The highest intermolecular frequencies are the bends that in-
volve the rocking of the proton donor molecule. The higher mass of D lowers these fre-
quencies when DF is the proton donor, and thereby reduces the total zero-point vibrational
energy. There are increases in the other two intermolecular frequencies associated with plac-
ing the lighter H nucleus on the acceptor molecule, but these changes are very much smaller
than those in the aforementioned bends.
Changing the subunit from HX to H2Y adds a number of new vibrational modes to the
dimer. Nonetheless, the D-bonded form of the water dimer is more stable than the H-bonded
complex251. The energy difference of 60 cm-1 is attributed chiefly to the out-of-plane mo-
tion that shears the bond which is of higher frequency for a proton than a deuteron, consis-
tent with the observations for the HF dimer.
The preference for D- versus H-bonding persists in solid matrices as well. When HDO
is paired with NH3, HOH, formaldehyde, or formamide, it is the D atom of the former mol-
ecule which acts as the bridge252-254. The same holds true for complexes betwen HDO and
olefins, where there is some question as to whether there is a true H-bond present255. In the
latter case, the preference for D acting as a bridge rather than H is quite small, inasmuch as
both isotopomers are observed. The difference in energy is estimated as less than 0.1
kcal/mol. An even smaller preference is noted between NCH---NCD and NCD---NCH, both
of which have been observed in the gas phase, leading to the suggestion that their energy
difference is only several c m - 1 at most256.
120 Hydrogen Bonding

Whereas the D-bond is stronger in some sense than the H-bond, the question of its length
is more complicated. In contrast to their greater length in the solid state, gas-phase results
indicate D-bonds are shorter than their protium analogs257. This discrepancy diminishes as
the H-bond becomes stronger. The authors explained the gas-phase shortening on the basis
of the high-frequency bending and stretching modes.
The preference for D-bonding versus placement of a protium in the bridging position ap-
pears to reverse in ionic H-bonds. Calculations at the SCF and MP2 levels of various iso-
topomers of the H 2 O-H + --OH 2 system258 indicate that any D atoms would tend to migrate
away from the bridging position, and toward one of the four peripheral positions. The equi-
librium constant for this preference lies in the neighborhood of 2-3, based on a strict har-
monic treatment of the ab initio force field. Using a temperature of 298° K, an equilibrium
constant of 2 translates into a free energy preference for D to act as bridging atom of 0.4
kcal/mol. These results confirm an earlier experimental study based on fractionation fac-
tors in the gas phase259.
Similarly, in the case of an analagous anion, like (CH 3 O-H--OCH 3 ) - , ab initio calcula-
tions with a 4-31G basis set260 indicate that there is a preference for a protium over D in
the bridging position. It is found from ion cyclotron resonance spectroscopic measurements
that the reaction

lies to the left, with an equilibrium constant of 0.3260. A similar preference for H versus D-
bonding has been observed in the gas phase for (CI-H--C1)- as well261.

2.13.1 Water Molecules


Recent ab initio calculations have attempted to probe the fundamental source of the rever-
sal of H/D preference in ionic as compared to neutral systems, using water as a test base262.
A harmonic analysis of the potential energy surface of the water dimer, computed with a 6-
31G** basis set, indicates that the preference for D in the bridging site can be explained in
a manner similar to that described earlier for HF--HF. The frequency of the bending motion
of the bridging atom is sensitive to its mass: this effect leads to a lower vibrational energy
of some 0.2 kcal/mol when the heavier D undergoes this motion. The computations indi-
cated that electron correlation has little effect upon this conclusion, even its quantitative as-
pects. While the treatment was purely harmonic in nature, other calculations263 have indi-
cated that anharmonicity effects yield very little distinction between one isotopomer and
the next.
The calculations262 considered higher levels of deuteration than single substitution and
found that this central conclusion remains unchanged. Several manifestations of this pref-
erence for a D-bond are:
1. When DOD is paired with HOH, it is the former molecule that acts as proton donor.
2. HOD--OH2 is lower in vibrational energy than is DOH OH.
3. DOD--OD2 is more strongly bound than is HOH--OH2.
The energetic preference for placement of a deuterium at a bridging position in the wa-
ter dimer carries over nearly unchanged in the trimer262. This complex forms a triangular
structure, almost equilateral. Each H-bond is distorted from linearity by the constraints of
the structure, with 6 (OOH) angles close to 20°. Nonetheless, it is again the vibrational
modes which distort this bridging atom from the H-bond axis that can be traced as the source
Geometries and Energetics 121

of the lower vibrational energy for isotopomers where D acts as bridge as opposed to H.
The close similarity between dimer and trimer in this respect leads one to suppose that the
principle would remain valid for larger aggregates, including the liquid state as well. In fact,
recent calculations verified the insensitivity of fractionation factors for the bridging hydro-
gens in oligomers of formamide264.
Ions of both positive and negative charge were analyzed as well, using (H2OH+--OH2)
and (HOH--OH)- as prototypes262. As in the case of the neutral dimer, deuterosubstitution
of the bridging atom lowers the total intermolecular zero-point vibrational energy (ZPVE)
of (H2OH+--OH2) by 0.2 kcal/mol more than if the replacement occurs at a nonbridging
(terminal) site. But the situation is quite different with respect to the intramolecular modes.
Whereas the intramolecular ZPVE of the neutral dimer is quite insensitive to the site of sub-
stitution, in the cation there is a 0.7 kcal/mol greater lowering of this quantity when D oc-
curs at a terminal, as opposed to a bridging, site. The latter effect outweighs the intermo-
lecular preference, so that in the case of (H 2 OH + --OH 2 ), a deuterium is favored to occupy
a terminal position by some 0.5 kcal/mol. This same trend extends to multiple deuterosub-
stitution.
The calculations indicated a remarkably small influence of correlation upon the trends
discussed above. This lack of sensitivity is particularly notable since the equilibrium geom-
etry of (H2OH+-OH2) is fundamentally different at the SCF and MP2 levels. The proton
transfer potential contains a pair of minima in the former case, whereas there is only a sin-
gle, centrosymmetric minimum present after correlation is included. The computations
were extended to the larger proton-bound dimer of methanol, and the results were found to
be similar, indicating that water is typical of general systems bound together through hy-
droxyl groups. The results of deuterosubstitution in the anion are of smaller magnitude, but
do appear to confirm the trend that H-bonds are slightly stronger than D-bonds.
Inclusion of aspects of the H-bond energy other than the electronic contribution, as well
as entropic effects, permit inferences to be drawn about the magnitude of AG for formation
of a H-bond at room temperature. It was found262 that the computed AS of the D-bridged
water dimer is more negative than for the H-bridge by perhaps 0.2-0.3 cal mol-1 deg - 1 .
As a result, the energetic preference for the D-bond in the neutral complex will slowly di-
minish as the temperature rises, eliminating some (but not all, by any means) of the 0.2
kcal/mol preference. Entropy has a similar "push" toward the H-bond in the ions as well,
in this case reinforcing the energetic preference for the H-bond.
While not directly relevant to H-bonds per se, a recent work compared the H/D frac-
tionation factors for a number of small molecules, as computed at various levels of the-
ory265. The results indicated that the relative free energy of protiated versus deuterated
species is rather insensitive to choice of basis set. Best results are achieved if polarization
functions are added to all atoms, and correlation is recommended, but even SCF computa-
tions with a basis set as small as 3-21G* can provide quite reasonable results.

2.14 Summary

The strength of the hydrogen bond rises with the acidity of the proton donor and the basic-
ity of the acceptor. Of the simple hydrides, then, complexes of the H3Z--HX type are the
strongest, with an electronic contribution to the binding energy of about 11 kcal/mol. Alky-
lation of the amine can improve its H-bonding ability, leading to H-bond energies on the
order of 20 kcal/mol. Indeed, when the acidity and basicity of the donor and acceptor, re-
122 Hydrogen Bonding

spectively, reach large enough proportions, the character of the complex can change to an
ion pair when the bridging proton is transferred across to the base. The strongest H-bonds
are formed between first-row atoms F, O, and N. The decrements in going from the second
to lower rows are smaller ones.
Associated with stronger H-bonds are typically shorter distances between the nonhy-
drogen atoms. One of the shorter contacts is R(N---F) of about 2.7 A in H3N---HF; a simi-
lar bond length is characteristic of H2O—HF as well, although the latter is not quite as
strongly bound. The formation of the H-bond stretches the A—H covalent bond, again by
a larger amount for stronger interactions. This stretch amounts to some 0.03 A in H3N HF.
Electron correlation is a rather important phenomenon in H-bonding. Its neglect can lead
to underestimation of the strength of the bond and its associated phenomena. Correlation
becomes progressively and proportionately more important for atoms of the second and
third rows of the periodic table. It is also important to correct for basis set superposition er-
ror, particularly at the correlated level, as this phenomenon can lead to spurious inflation of
the H-bond energy by several kcal/mol. Zero-point vibrational energies will typically re-
duce the dissociation energy as the complex contains more vibrational modes than the pair
of isolated monomers. This reduction is about 3 kcal/mol for a strongly bound complex like
H3N HF.
When a H2Y molecule acts as the proton acceptor, there is a delicate balance between
one orientation, in which the proton donor approaches along the direction of a Y lone elec-
tron pair, and another in which the donor is attracted toward the negative end of the H2Y
molecular dipole moment. Consequently, the energy profile for transition from a pyramidal
to planar complex is typically a rather flat one. Atoms of lower rows of the periodic table,
that is, S and Se, show a marked proclivity for pyramidalization while water tends more to-
ward planarity. In the case where the planar structure is only slightly higher in energy than
the pyramidal one, vibrational motions may hide the energy barrier to inversion around the
O atom, and experimental observation may suggest a planar equilibrium geometry. In the
case of a heavy atom such as I, an "anti-H-bonded" geometry such as H2O IH can become
competitive in energy to the H-bonded, albeit weakly so, H2O-HI geometry. While the lat-
ter is stabilized by the normal electrostatic interaction between the two subunits, dispersion
is maximized in the former.
When paired with a stronger H3Z base, H2Y acts as proton donor. While this interaction
has all the characteristics of a H-bond, it is considerably weaker than H3Z HX or even
H2Y--HX, suggesting H2Y is not a very effective proton donor. The electronic portion of
the binding energy of H3N--HOH is about 6 kcal/mol, only about half that in H3N---HF, and
the H-bond is longer by some 0.2 A.
When two HX molecules are placed together, they do not line up in collinear HX---HX
fashion. Rather, the proton acceptor rotates so as to present one of its three lone electron
pairs toward the bridging hydrogen. This reorientation makes for an angle in the 110°-120°
range. Because of the poor proton-accepting quality of HF, the dimerization energy of (HF)2
is only about 4.5 kcal/mol (the electronic contribution), comparable to, but slightly weaker
than, H3N--HOH. The thorough testing of various levels of theory suggest that MP2 is an
excellent means to evaluate the contribution of electron correlation, with results quite sim-
ilar to those achieved with MP4 or coupled cluster approaches. The interfluorine distance
is about 2.75 A. The HC1 dimer is more weakly bound, with a — Eelec of less than 2
kcal/mol. The acceptor is nearly perpendicular to the donor in (HC1)2. In the mixed com-
plex pairing HF with HC1, either molecule can act as donor or acceptor; there is little dif-
ference in energy between HP--HC1 and HC1---HF. The binding energy of either is slightly
Geometries and Energetics 123

greater than in the HC1 dimer. HX dimers involving Br and I are even more weakly bound.
In fact, it is questionable whether any of the HX dimers, other than (HF)2, really contain a
true H-bond.
Despite a great deal of speculation over the years as to the possibility of "bifurcated" or
"cyclic" H-bond geometries in the water dimer, it appears that the classic linear structure is
the equilibrium structure, and perhaps the only true minimum on the potential energy sur-
face. Calculations with large basis sets and thorough account of correlation have enabled a
particularly accurate estimate of the energetics of this H-bond. The SCF limit of the elec-
tronic part of the binding energy is 3.5-3.7 kcal/mol. Correlation adds another kcal/mol or
so, leading to a theoretical value of 4.5-5.0 kcal/mol. Thus, the water dimer is bound at
about the same strength as (HF)2, although the H-bond in (H2O)2 is about 0.2 A longer. Af-
ter correction for vibrational terms, AH comes to —2.9 kcal/mol for the water dimer. Since
the entropy of binding is negative, the Gibbs free energy becomes positive at 298° K, about
+ 3.7 kcal/mol. The equilibrium geometry of (H2S)2 remains unclear at this time, with lin-
ear and bifurcated likely candidates. In either case, this dimer is weakly bound, with a bind-
ing energy of 1 kcal/mol or so after zero-point vibrational corrections. When mixed to-
gether, it is not clear whether H2O or H2S would be the proton donor and which the acceptor.
Alkyl substitution of the hydrogens of H2O has little influence upon the H-bond energet-
ics. Replacement by an electronegative atom like Cl, on the other hand, significantly en-
hances the binding energy (by several kcal/mol) as HOC1 is a considerably stronger proton
donor than is HOH.
The potential energy surface of the ammonia dimer is so flat that the concept of a single
equilibrium geometry loses some of its meaning. The linear type of H-bond easily inter-
converts with a cyclic geometry with very little change in energy. As a result, the presence
of a true H-bond in this dimer is questionable. The system offered an opportunity to exam-
ine how small modifications in theoretical method can shift the balance between one geom-
etry and another, and provided a caution against drawing conclusions of relative stability
based upon small differences in energy. The electronic part of the interaction energy ap-
pears to be some 4 kcal/mol, slightly weaker than the water dimer.
Enlarging the types of systems studied beyond the simple hydrides adds interesting pos-
sibilities. Proton donors appear to prefer to approach the carbonyl oxygen atom along one
of its lone pairs (6(C=O-H)~120°) rather than directly along the C=O axis, although this
preference is not a pronounced one. In addition to the stretch of the A—H bond within the
donor molecule, noted for the hydrides, the C=O bond of the acceptor elongates as well,
and by a comparable amount. The carbonyl oxygen of formaldehyde is comparable in
strength as a proton acceptor to the O of water. By sacrificing a linear O—H O arrange-
ment, it is possible for a proton donor molecule like H2O to align itself with H2CO such
that a second stabilizing interaction is formed: the oxygen of the water can approach the
C—H of H2CO. The energetic cost of bending the H-bond is nearly equally compensated
by the second interaction. The two O atoms are a little closer together in the H 2 O—H 2 CO
complex than in the water dimer.
The carboxylic group is considerably more acidic than water. Nonetheless, the two O
atoms of COOH act as proton acceptors when paired with a hydrogen halide, due in part to
the poor proton accepting ability of the halogen atom. Not unexpectedly the C=O oxygen
is a considerably better proton acceptor than the —OH oxygen. Indeed this C=O within
the context of the carboxyl group is a superior acceptor to the C=O in formaldehyde; form-
ing a H-bond with HF in excess of 10 kcal/mol. On the other hand, this interaction energy
is not solely the result of a single H-bond; some arises as a result of a secondary H-bond as
124 Hydrogen Bonding

the F accepts a proton from the —OH group of carboxyl. The primary H-bond is rather short
as well, around 2.6 A. Other measures of the stronger H-bond are enhanced lengthening of
the HF and C=O bonds. When HF is replaced by the better proton acceptor, HOH, the "sec-
ondary" H-bond, wherein HOH accepts a proton from the —OH of COOH, is strengthened
at the expense of a somewhat weaker interaction with the C=O group where HOH acts as
donor. As a result, the complex takes on a bit more of a "cyclic" character with two nearly
equivalent H-bonds. The H-bonds with HCOOH donating a proton to either an O or N
acceptor appear to be some 2.7 A in length, with an interaction energy in excess of 10 kcal/
mol in either case. A pair of carboxylic acids will form a dimer, containing two strong
O—H O=C H-bonds. The total interaction energy approaches 15 kcal/mol, with each H-
bond about 2.7 A in length.
Although the N atom of NH3 or amines is not a good proton donor, it becomes much
more electronegative when involved in a triple bond as in a nitrile. The N of HCN will ac-
cept a proton from the strong donor HF. The ensuing H-bond is of normal length, less than
2.9 A, with a strength of some 7 kcal/mol. Replacing the hydrogen of HCN by an alkyl
group enhances the ability of the N to accept a proton. But more important than the proton-
accepting ability of the nitrile N is the acidity of the C H group. There is evidence of the
formation of the reverse complex, that is, NCH---FH, as a secondary minimum on the PES.
NCH clearly donates a proton when combined with NH3 Although this H-bond is fairly
long, perhaps 3.14 A, it amounts to some 6 kcal/mol. And a standard H-bond is present as
well in the dimer of HCN, albeit slightly weaker and longer than in NCH—NH 3 . (HCN)2
shows evidence of stretching of the C—H bond upon complexation, with a H-bond energy
of about 4 kcal/mol. When involved in a double bond, the N atom of an imine is not as good
a proton acceptor as in amines, but better than the doubly bonded O of the carbonyl group.
The amide group contains a carbonyl C=O and amine group, directly connected to each
other so that they perturb one another's properties. One can draw a resonance structure in
which a double bond connects the C and N atoms, leaving a negative charge on O and pos-
itive charge on N. This interaction enhances the proton accepting ability of the former and
makes the N—H group a strong donor. When an amide like formamide interacts with HOH
or some such molecule, a cyclic arrangement occurs which allows the N—H to donate to
the O while the carbonyl C=O accepts a water proton. The overall interaction energy is in
the range of 9 kcal/mol. This cyclic structure is 2 kcal/mol more stable than one in which
only a single H-bond is formed to the carbonyl oxygen; a single H-bond to the N-H group
of the amide is weaker still by the same amount. Even stronger is the interaction between a
pair of amides. If there is a free N—H cis to the C=O group, the two molecules can form
a pair of NH O=C H-bonds, neither of which is badly bent. The total interaction energy
here amounts to some 14 kcal/mol, twice the energy of a single H-bond between a pair of
amides. The N—H group in an amide appears to be comparable in strength as a proton donor
to the O—H of water. An analogous pair of H-bonds can be formed, within the context of
a cyclic geometry, when one of the amide molecules is replaced by a carboxylic acid.
Ab initio calculations are now capable of being applied directly to systems as large as
pairs of nucleic acid bases. One can obtain reliable data, with binding enthalpies within
about 1 kcal/mol of experiment, using polarized basis sets and MP2 correlation, applying
counterpoise correction, and fully optimizing the geometries of the entire pair using gradi-
ent procedures. The computations are able to provide information about the native binding
abilities of these bases with one another, in isolation from the other components and geo-
metrical constraints of the DNAbiopolymer. The results illustrate the strong binding of the
guanine-cytosine pair, followed closely by the guanine homodimer. It also appears on the
Geometries and Energetics 125

basis of the calculations that the Hoogsteen pairing of adenine and thymine is probably
somewhat more stable than the Watson-Crick combination, although this difference
amounts to only about 10% of the full binding energy. In any case, the binding enthalpy of
the AT pair is only about half as large as the GC pair. Comparison with a series of other pairs
makes it clear that this difference cannot be attributed solely to the presence of three H-
bonds in GC and only two in AT. While the interaction energy of the stacked bases is cer-
tainly weaker than the coplanar H-bonded geometries, the stacking energy can represent a
significant contribution to the overall stability of the DNA molecule.
Of course isotopic substitution has no effect on the potential energy surface of H-bonded
complexes or any system. However, atoms of different mass will leave the entire system
with differing normal modes of vibration and their associated frequencies. In simple sys-
tems like FH--FH, the zero-point vibrational energy is lowered more if the bridging hy-
drogen is replaced by deuterium, as compared to the terminal atom. That is, FD--FH has
less vibrational energy than does FH---FD. This difference can be traced to one of the in-
termolecular modes, the wagging motion of the bridging atom. Since this normal mode con-
sists chiefly of the motion of the hydrogen, its effective mass is lowered significantly by the
replacement of H by D. The energy difference amounts to some 100 cm - 1 , or about 0.3
kcal/mol. Similar results have been obtained for the water dimer, albeit a slightly smaller
energetic preference for the D-bond versus the H-bond. The trend reverses for strong ionic
complexes such as (H2OH OH2)+ or (HO H OH)- where it is the H that is preferred
over D to adopt the bridging position.

References
1. Marco, J., Orza, J. M, Notario, R., and Abboud, J.-L. M., Hydrogen bonding of neutral species
in the gas phase: The missing link, J. Am. Chem. Soc. 116, 8841-8842 (1994).
2. Hinchliffe, A., Ab initio study of the hydrogen-bonded complexes H3N HBr, H3P—HBr,
H 3 As-HF, H 3 As-HCl, and H 3 As-HBr, J. Mol. Struct. (Theochem) 121, 201-205 (1985).
3. Latajka, Z. and Schemer, S.,Ab initio study of FH—PH3 and CIH—PH3 including the effects of
electron correlation, J. Chem. Phys. 81, 2713-2716 (1984).
4. Latajka, Z. and Schemer, S., Ab initio comparison of H bonds and Li bonds. Complexes of LiF,
LiCl, HF, and HCl with NH3, J. Chem. Phys. 81, 4014-4017 (1984).
5. Latajka, Z. and Scheiner, S., Basis sets for molecular interactions. 2. Application to H3N—HF,
H3N-HOH, H20-HF, (NH3)2, and H 3 CH-OH 2 , J. Comput. Chem. 5, 674-682 (1987).
6. Sadlej, J. and Miaskiewicz, K., Ab initio calculations of the vibrational spectra of the ammonia
and fluoramide complexes with HF, J. Mol. Struct. (Theochem) 236, 427-441 (1991).
7. Del Bene, J. E., An ab initio molecular orbital study of the structures and energies of neutral and
charged bimolecular complexes of NH3 with the hydrides AHn (A = N, O, F, P, S, and Cl), J.
Comput. Chem. 10, 603-615 (1989).
8. Latajka, Z. and Scheiner, S., Three-dimensional spatial characteristics of primary and secondary
basis set superposition error, Chem. Phys. Lett. 140, 338-343 (1987).
9. Latajka, Z. and Scheiner, S., Primary and secondary basis set superposition error at the SCF
and MP2 levels: H 3 N—Li + a n d H 2 O - L i + , J. Chem. Phys. 87, 1194-1204 (1987).
10. Szczesniak, M. M. and Scheiner, S., Accurate evaluation of SCF and MP2 components of inter-
action energies. Complexes of HF, OH2, and NH3 with Li+, Coll. Czech. Chem. Commun. 53,
2214-2229 (1988).
11. Latajka, Z., Scheiner, S., and Ratajc/.ak, H., The proton position in amine-HX (X = Br,I) com-
plexes, Chem. Phys. 166, 85-96 (1992).
12. Bacskay, G. B. and Craw, J. S., Quantum chemical study of the trimelhylamine-hydrogen chlo-
ride complex, Chem. Phys. Lett. 221, 167-174 (1994).
126 Hydrogen Bonding

13. Taylor, R. and Kennard, O., Hydrogen-bond geometry in organic crystals, Ace. Chem. Res. 17,
320-326 (1984).
14. Hinchliffe, A., Ab initio study of the hydrogen-bonded complexes H2X HY (X = O,S,Se:
Y=F,Cl,Br), J. Mol. Struct. (Theochem) 106, 361-366 (1984).
15. Hannachi, Y., Silvi, B., and Bouteiller, Y., Structure and vibrational properties of water hydro-
gen halide complexes, J. Chem. Phys. 94, 2915-2922 (1991).
16. Szczesniak, M. M., Scheiner, S., and Bouteiller, Y., Theoretical study of H2O—HF and
H2O—HCl: Comparison with experiment, J. Chem. Phys. 81, 5024-5030 (1984).
17. Szczesniak, M. M. and Scheiner, S., Contribution of dispersion to the properties of H 2 S - H F
andH2S-HCl, J. Chem. Phys. 83, 1778-1783 (1985).
18. Goodwin, E. J. and Legon, A. C., Microwave rotational spectrum of a weakly bound complex
formed by hydrogen sulfide and hydrogen chloride, J. Chem. Soc., Faraday Trans. 2 80, 51-65
(1984).
19. Legon, A. C., Millen, D. J., and North, H. M., Experimental determination of the dissociation
energies Do and De of H2O HF, Chem. Phys. Lett. 135, 303-306 (1987).
20. de Oliveira, G. and Dykstra, C. E., Weakly bonded clusters of'H2S, J. Mol. Struct. (Theochem)
362,275-282(1996).
21. Viswanathan, R. and Dyke, T. R., The structure of H2S HF and the stereochemistry of the hy-
drogen bond, J. Chem. Phys. 77, 1166-1174 (1982).
22. Willoughby, L. C., Fillery-Travis, A. J., and Legon, A. C., An investigation of the rotational spec-
trum ofH2S HF by pulsed-nozzle, Fourier-transform microwave spectroscopy: Determination
of the hyperfine coupling constants Xaa(33S), XDaa, and D aa H(D)F , J. Chem. Phys. 81, 20-26
(1984).
23. Platts, J. A., Howard, S. T., and Bracke, B. R. F., Directionality of hydrogen bonds to sulfur and
oxygen, J. Am. Chem. Soc. 118, 2726-2733 (1996).
24. Singh, U. C. and Kollman, P. A., Ab initio calculations on the structure and nature of the hy-
drogen bonded complex H2S HF, J. Chem. Phys. 80, 353-355 (1984).
25. Legon, A. C. and Millen, D. J., Determination of properties of hydrogen-bonded dimers by ro-
tational spectroscopy and a classification of dimer geometries, Faraday Discuss. Chem. Soc. 73,
71-87 (1982).
26. Novoa, J. J., Planas, M., Whangbo, M.-H., and Williams, J. M., Accurate computation of the nor-
mal and reverse complexes between water and hydrogen fluoride, Chem. Phys. 186, 175-183
(1994).
27. Legon, A. C. and Willoughby, L. C., Identification and molecular geometry of a weakly bound
dimer (H2O,HCl) in the gas phase by rotational spectroscopy, Chem. Phys. Lett. 95, 449—452
(1983).
28. Legon, A. C. and Suckley, A. P., Br nuclear quadrupole and H,Br nuclear-spin-nuclear-spin
coupling in the rotational spectrum of H2O-HBr, Chem. Phys. Lett. 150, 153-158 (1988).
29. Amos, R. D., Gaw, J. F., Handy, N. C., Simandiras, E. D., and Somasundram, K., Hydrogen-
bonded complexes involving HF and HCl: The effects of electron correlation and anharmonic-
ity, Theor. Chim. Acta 71, 41-57 (1987).
30. Bouteiller, Y, Allavena, M., and Leclercq, J. M., Cubic force constants of the FH—OH2 and
FH O(CH 3 ) 2 hydrogen-bonded complexes. An analysis of computed ab initio SCF values,
Chem. Phys. Lett. 69, 521-524 (1980).
31. Cope, P., Legon, A. C., and Millen, D. J., The microwave spectrum and geometry of the methanol-
hydrogen chloride dimer, Chem. Phys. Lett. 112, 59-64 (1984).
32. Hannachi, Y, Silvi, B., Perchard, .1. P., and Bouteiller, Y, Ab initio study of the infrared photo-
conversion in the water-hydrogen iodide system, Chem. Phys. 154, 23-32 (1991).
33. Hannachi, Y, Schriver, L., Schriver, A., and Perchard, J. P., Infrared photodissociation of hy-
drogen-bonded complexes trapped in inert matrices. The water-hydrogen iodide system, Chem.
Phys. 135,285-299(1989).
34. Bacskay, G. B., Kerdraon, D. 1., and Hush, N. S., Quantum chemical study of the HCl molecule
Geometries and Energetics 127

and its binary complexes with CO, C2H2, C2H4, PH3, H2S, HCN, H2O, and NH3: Hydrogen
bonding and its effect on the 35Cl nuclear quadrupole coupling constant, Chem. Phys. 144,
53-69(1990).
35. Hanessian, S., Gomtsyan, A., Simard, M., and Roelens, S., Molecular recognition and self-
assembly by "weak" hydrogen bonding: Unprecedented supramolecular helicate structures
from diamine/diol motifs, J. Am. Chem. Soc. 116, 4495-4496 (1994).
36. Ermer, O. and Eling, A., Molecular recognition among alcohols and amines: Super-tetrahedral
crystal architectures of linear diphenol-diamine complexes and aminophenols, J. Chem. Soc.
Perkin Trans. II925-944 (1994).
37. Latajka, Z. and Schemer, S., Structure, energetics, and vibrational spectrum of H3N HOH, J.
Phys. Chem. 94, 217-221 (1990).
38. Stockman, P. A., Bumgarner, R. E., Suzuki, S., and Blake, G. A., Microwave and tunable far-
infrared laser spectroscopy of the ammonia-water dimer, J. Chem. Phys. 96,2496—2510(1992).
39. Herbine, P. and Dyke, T. R., Rotational spectra and structure of the ammonia-water complex, J.
Chem. Phys. 83, 3768-3774 (1985).
40. Yeo, G. A. and Ford, T. A., Ab initio molecular orbital calculations of the infrared spectra of hy-
drogen bonded complexes of water, ammonia, and hydroxylamine. Part 6. The infrared spectrum
of the water-ammonia complex, Can. J. Chem. 69, 632-637 (1991).
41. Del Bene, J. E., Ab initio molecular orbital study of the structures and energies of neutral and
charged bimolecular complexes of H2O with the hydrides AHn (A = N, O,F,P,S,and Cl), J. Phys.
Chem. 92, 2874-2880 (1988).
42. Eraser, G. T. and Suenram, R. D., Perturbations in the infrared spectrum of the NH3 umbrella
mode of H O H — N H 3 , J. Chem. Phys. 93, 7287-7297 (1992).
43. Hess, O., Caffarel, M., Huiszoon, C., and Claverie, P., Second-order exchange effects in inter-
molecular interactions. The water dimer, J. Chem. Phys. 92, 6049-6060 (1990).
44. Langlet, J., Caillet, J., and Cafferel, M.,Aperturbational study of some hydrogen-bonded dimers,
J. Chem. Phys. 103, 8043-8057 (1995).
45. Herbine, P., Hu, T. A., Johnson, G., and Dyke, T. R., The structure ofNH3 H2S and free inter-
nal rotation effects, J. Chem. Phys. 93, 5485-5495 (1990).
46. Akagi, K., Ab initio study in the hydrogen bonding complex between hydrogen sulfide and am-
monia by means of configuration analysis, J. Mol. Struct. 133, 67—82 (1985).
47. Hilpert, G., Eraser, G. T., Suenram, R. D., and Karyakin, E. N., Proton interchange tunneling
and internal rotation in H S H - N H 3 , J. Chem. Phys. 102, 4321-4328 (1995).
48. Zheng, Y.-J. and Merz, K. M. J., Study of hydrogen bonding interactions relevant to biomolecu-
lar structure and function, J. Comput. Chem. 13, 1151-1169 (1992).
49. Wee, S. S., Kim, S., Jhon, M. S., and Scheraga, H. A., Analytical intermolecular potential func-
tions from ab initio self-consistent field-calculations for hydration of methylamine and methyl-
ammonium ion, J. Phys. Chem. 94, 1656-1660 (1990).
50. Fraser, G. T., Suenram, R. D., Lovas, F J., and Stevens, W. J., Microwave spectrum of the
CH3OH NH3 complex, Chem. Phys. 125, 31-43 (1988).
51. Schiefke, A., Deusen, C., Jacoby, C., Gerhards, M., Schmitt, M., Kleinermanns, K., and Hering,
P., Structure and vibrations of the phenol-ammonia cluster, J. Chem. Phys. 102, 9197-9204
(1995).
52. Vener, M. V., How reliable is the Lippincott-Schroeder potential for the OH"N hydrogen bonded
fragment in the gas phase?, Chem. Phys. Lett. 244, 89-92 (1995).
53. Lischka, H., A note on the ab initio calculation of intermolecular potentials: The HF dimer,
Chem. Phys. Lett. 66, 108-110 (1979).
54. Frisch, M. J., Pople, J. A., and Del Bene, J. E., Molecular orbital study of the dimers (AHn)2
formed from NH3, OH2, FH, PH3, SH2, and CIH, J. Phys. Chem. 89, 3664-3669 (1985).
55. Latajka, Z. and Scheiner, S., Structure, energetics and vibrational spectra of H-bonded systems.
Dimers and trimers of HF and HCl, Chem. Phys. 122, 413-430 (1988).
56. Michael, D. W., Dykstra, C. E., and Lisy, J. M., Changes in the electronic structure and vibra-
128 Hydrogen Bonding

tional potential of hydrogen fluoride upon dimerization: A well-correlated (HF)2 potential en-
ergy surface, J. Chem. Phys. 81, 5998-6006 (1984).
57. Frisch, M. J., Del Bene, J. E., Binkley, J. S., and Schaefer, H. F., Extensive theoretical studies
of the hydrogen-bonded complexes (H2O)2, (H2O)2H+, (HF)2, (HF)2H+, F 2 H - , and (NH3)2,
J. Chem. Phys. 84, 2279-2289 (1986).
58. Howard, B. J., Dyke, T. R., and Klemperer, W., The molecular beam spectrum and the structure
of the hydrogen fluoride dimer, i. Chem. Phys. 81, 5417-5425 (1984).
59. Del Bene, J. E., Basis set and correlation effects on computed hydrogen bond energies of the
dimers(AHn)2:AHn = NHy OH2, and FH, J. Chem. Phys. 86, 2110-2113 (1987).
60. Peterson, K. A. and Dunning, T. H. J., Benchmark calculations with correlated molecular wave
functions. VII. Binding energy and structure of the HF dimer, J. Chem. Phys. 102, 2032—2041
(1995).
61. Pine, A. S. and Howard, B. J., Hydrogen bond energies of the HF andHCl dimersfrom absolute
infrared intensities, J. Chem. Phys. 84, 590-596 (1986).
62. Quack, M. and Suhm, M. A., On hydrogen-bonded complexes: The case of (HF)2, Theor. Chim.
Acta 93, 61-65 (1996).
63. Collins, C. L., Morihashi, K., Yamaguchi, Y., and Schaefer, H. F., Vibrational frequencies of the
HF dimer from the coupled cluster method including all single and double excitations plus per-
turbative connected triple excitations, J. Chem. Phys. 103, 6051-6056 (1995).
64. Hobza, P., Carsky, P., and Zahradnik, R., The dimer (HCl)2: Stationary points on the SCF 4-31G
energy hypersurface and the role of entropy in the dimer formation, Coll. Czech. Chem. Com-
mun. 44, 3458-3463 (1979).
65. Allavena, M., Silvi, B., and Cipriani, J., The in-equivalence of the HCl molecules in (HCl)2: An
SCF ab initio calculation, J. Chem. Phys. 76, 4573-4577 (1982).
66. Karpfen, A., Bunker, P. R., and Jensen, P., An ab initio study of the hydrogen chloride dimer: the
potential energy surface and the characterization of the stationary points, Chem. Phys. 149,
299-309(1991).
67. Tao, F.-M. and Klemperer, W., Ab initio potential energy surface for the HCl dimer, J. Chem.
Phys. 103,950-956(1995).
68. Ohashi, N. and Pine, A. S., High resolution spectrum of the HCl dimer, J. Chem. Phys. 81,73-84
(1984).
69. Blake, G. A., Busarow, K. L., Cohen, R. C., Laughlin, K. B., Lee, Y. T., and Saykally, R. J., Tun-
able far-infrared spectroscopy of hydrogen bonds: The Ka = 0(u) l(g) rotation-tunneling
spectrum of the HCl dimer, J. Chem. Phys. 89, 6577-6587(1988).
70. Fraser, G. T. and Pine, A. S., Microwave and infrared electric-resonance optothermal spec-
troscopy of HF—HCl and HCl-HF, J. Chem. Phys. 91, 637-645 (1989).
71. Oudejans, L. and Miller, R. E., State-to- state photodissociation of oriented HF—HCl complexes:
Isotopic and isomeric effects, J. Phys. Chem. 99, 13670-13679 (1995).
72. Del Bene, J. E. and Shavitt, I., Comparison of methods for determining the correlation contri-
bution to hydrogen bond energies, Int. J. Quantum Chem., Quantum Chem. Symp. 23, 445—452
(1989).
73. Cizek, J., On the use of the cluster expansion and the technique of diagrams in calculations of
correlation effects in atoms and molecules, Adv. Chem. Phys. 14, 35—89 (1969).
74. Laidig, W. D. and Bartlett, R. J., A multi-reference coupled-cluster method for molecular appli-
cations, Chem. Phys. Lett. 104, 424-430 (1984).
75. Gdanitz, R. J. and Ahlrichs, R., The averaged coupled-pair functional (ACPF): A size-extensive
modification of MR CI(SD), Chem. Phys. Lett. 143, 413-420 (1988).
76. Langhoff, S. R. and Davidson, E. R., Configuration interaction calculations on the nitrogen mol-
ecule, Int. J. Quantum Chem. 8, 61-72 (1974).
77. Siegbahn. P. E. M., Multiple substitution effects in configuration interaction calculations, Chem.
Phys. Lett. 55, 386-394 (1978).
78. Pople, J. A.. Seeger, R., and Krishnan, R., Variational configuration interaction methods and
Geometries and Energetics 129

comparison with perturbation theory, Int. J. Quantum Chem., Quantum Chem. Symp. 11,
149-163(1977).
79. Racine, S. C. and Davidson, E. R., Electron correlation contribution to the hydrogen bond in
(HF)2, J. Phys. Chem. 97, 6367-6372 (1993),.
80. Hannachi, Y. and Silvi, B., Structure and bonding of hydrogen halide complexes: An ab initio
calculation of the 1:1 species, J. Mol. Struct. (Theochem) 200, 483-496 (1989).
81. Zhang,!., Dulligan, M., Segall, J., Wen, Y., andWittig, C., Ultraviolet photochemistry and pho-
tophysics of weakly-bound (HI)2 clusters via high-n Rydberg time-of-flight spectroscopy, J. Phys.
Chem. 99, 13680-13690 (1995).
82. Bumgarner, R. E. and Kukolich, S. G., Microwave spectra and structure of HI—HF complexes,
J. Chem. Phys. 86, 1083-1089 (1987).
83. Scheiner, S., Ab initio studies of hydrogen bonds: The water dimer paradigm, Annu. Rev. Phys.
Chem. 45, 23-56(1994).
84. Muguet, F. R, Robinson, G. W., and Bassez-Muguet, M.-P, The intermolecular vibrations of the
bifurcated water dimer: An ab initio study, Int. J. Quantum Chem. 39, 449-454 (1991).
85. Astrand, P.-O., Wallqvist, A., and Karlstrom, G., On the basis set superposition error in the eval-
uation of water dimer interactions, J. Phys. Chem. 95, 6395-6396 (1991).
86. Dannenberg, J. J., and Mezei, M., Reply to the Comment on the application of basis set super-
position error to ab initio calculation of water dimer, J. Phys. Chem. 95, 6396-6398 (1991).
87. Szalewicz, K., Cole, S. J., Kolos, W., and Bartlett, R. J., A theoretical study of the water dimer
interaction, J. Chem. Phys. 89, 3662-3673 (1988).
88. van Duijneveldt-van de Rijdt, J. G. C. M. and van Duijneveldt, F. B., Convergence to the basis-
set limit in ab initio calculations at the correlated level on the water dimer, J. Chem. Phys. 97,
5019-5030(1992).
89. Feller, D., Application of systematic sequences of wave functions to the water dimer, J. Chem.
Phys. 96, 6104-6114(1992).
90. Vos, R. J., Hendriks, R., and van Duijneveldt, F. B., SCF, MP2, and CEPA-1 calculations on the
OH Ohydrogen bonded complexes(H2O)2 and (H2O-H2CO), J.Comput.Chem. 11,1-18(1990).
91. Chakravorty, S. J. and Davidson, E. R., The water dimer: Correlation energy calculations,
J. Phys. Chem. 97, 6373-6383 (1993).
92. Kim, K. S., Mhin, B. J., Choi, U.-S., and Lee, K.,Ab initio studies of the water dimer using large
basis sets: The structure and thermodynamic energies, J. Chem. Phys. 97, 6649-6662. (1992).
93. Saeb0, S., Tong, W., and Pulay, P., Efficient elimination of basis set superposition errors by the
local correlation method: Accurate ab initio studies of the water dimer, J. Chem. Phys. 98,
2170-2175(1993).
94. Wang, Y.-B., Tao, F.-M., and Pan, Y.-K., Accurate calculation of the binding energy of the wa-
ter dimer, J. Mol. Struct. (Theochem) 309, 235-239 (1994).
95. de Oliveira, G. and Dykstra, C. E., Bond functions in the description of the water dimer, J. Mol.
Struct. (Theochem) 337, 1-7 (1995).
96. Feyereisen, M. W., Feller, D., and Dixon, D. A., Hydrogen bond energy of the water dimer,
J. Phys. Chem. 100, 2993-2997 (1996).
97. Williams, H. L., Mas, E. M., Szalewicz, K., and Jeziorski, B., On the effectiveness of monomer-,
dimer-, and bond-centered basis functions in calculations of intermolecular interaction ener-
gies, J. Chem. Phys. 103, 7374-7391 (1995).
98. Dyke, T. R., Mack, K. M., and Muenter, J. S., The structure of water dimer from molecular beam
electric resonance spectroscopy, J. Chem. Phys. 66, 498-510 (1977).
99. van Hensbergen, B., Block, R., and Jansen, L., Effect of direct and indirect exchange interac-
tions on geometries and relative stabilities of H2O and H2S dimers in bifurcated, cyclic, and lin-
ear configurations, J. Chem. Phys. 76, 3161-3168 (1982).
100. Woodbridge, E. L., Tso, T.-L., McGrath, M. P., Hehre, W. J., and Lee, E. K. C., Infrared spectra
of matrix-isolated monomeric and dimeric hydrogen sulfide in solid O2, J. Chem. Phys. 85,
6991-6994(1986).
I 30 Hydrogen Bonding

101. Fernandez, P. F., Ortiz, J. V., and Walters, E. A., Calculations on species relevant to the photo-
ionization of the van der Waals molecule (H2S)2, J. Chem. Phys. 84, 1653-1658 (1986).
102. Spackman, M. A., A simple quantitative model of hydrogen bonding. Application to more com-
plex systems, J. Phys. Chem. 91, 3179-3186 (1987).
103. de Oliveira, G., and Dykstra, C. E., Large basis set study of the stability of(H2S)2: The impor-
tance of 3d functions in weak interaction of second row molecules, Chem. Phys. Lett. 243,
158-164(1995).
104. Amos, R. D., Structures, harmonic frequencies and infrared intensities of the dimers of H2O and
H2S, Chem. Phys. 104, 145-151 (1986).
105. Bolis, G., Clementi, E., Wertz, D. H., Scheraga, H. A., and Tosi, C., Interaction of methane and
methanol with water, J. Am. Chem. Soc. 105, 355-360 (1983).
106. Nishi, N., Koga, K., Ohshima, C., Yamamoto, K., Nagashima, U., and Nagami, K., Molecular
association in ethanol-water mixtures studied by mass spectrometric analysis of clusters gener-
ated through adiabatic expansion of liquid jets, J. Am. Chem. Soc. 110, 5246—5255 (1988).
107. Bakkas, N., Bouteiller, Y., Loutellier, A., Perchard, J. P., and Racine, S., The water-methanol
complexes. Matrix induced structural conversion of the 1-1 species, Chem. Phys. Lett. 232,
90-98 (1995).
108. Huisken, F. and Stemmler, M., On the structure of the methanol-water dimer, Chem. Phys. Lett.
180,332-338(1991).
109. Bakkas, N., Bouteiller, Y, Loutellier, A., Perchard, J. P., and Racine, S., The water-methanol
complexes. I. A matrix isolation study and an ab initio calculation on the 1—1 species, J. Chem.
Phys. 99, 3335-3342 (1993).
110. Kroon-Batenburg, L. M. J. and van Duijneveldt, F. B., Effect of methyl substitution at the ac-
ceptor in O—H O hydrogen bonds. Application of a moment-optimized DZP basis at the (self-
consistent field + dispersion + corrected point-charge) level, J. Phys. Chem. 90, 5431-5435
(1986).
111. Markham, G. D. and Bock, C. W., Interaction of the sulfides CH 3 SCH 3 and CH3SCH2CO2 - with
water molecules. An ab initio molecular orbital study, J. Phys. Chem. 99, 10118-10129 (1995).
112. Ugliengo, P., Saunders, V., and Garrone, E., Silanol as a model for the free hydroxyl of amor-
phous silica: Ab initio calculations of the interaction with water, J. Phys. Chem. 94, 2260-2267
(1990).
113. Ugliengo, P., Bleiber, A., Garrone, E., Sauer, J., and Ferrari, A. M., Relative propensity of
methanol and silanol towards hydrogen bond formation, Chem. Phys. Lett. 191,537-547 (1992).
114. Sathyan, N., Santhanam, V., and Sobhanadri, J., Ab initio calculations on some binary systems
involving hydrogen bonds, J. Mol. Struct. (Theochem) 333, 179-189 (1995).
115. Schmitt, M., Muller, H., Henrichs, U., Gerhards, M., Perl, W., Deusen, C., and Kleinermanns,
K., Structure and vibrations of phenol-CH 3 OH (CD3OD) in the electronic ground and excited
state, revealed by spectral hole burning and dispersed fluorescence spectroscopy, J. Chem. Phys.
103,584-594(1995).
116. Gerhards, M., Schmitt, M., Kleinermanns, K., and Stahl, W., The structure of phenol(H2O) ob-
tained by microwave spectroscopy, J. Chem. Phys. 104, 967-971 (1996).
117. Breden, G., Meerts, W. L., Schmitt, M., and Kleinermanns, K., High resolution UVspectroscopy
of phenol and the hydrogen bonded phenol-water cluster, J. Chem. Phys. 104, 972-982 (1996).
118. Schutz, M., Burgi, T., Leutwyler, S., and Fischer, T., Intermolecular bonding and vibrations of
phenol-H2O(D2O), J. Chem. Phys. 98, 3763-3776 (1993).
119. Schutz, M., Burgi, T., and Leutwyler, S., Structures and vibrations of phenol-H2O and d-phe-
nol-D2O based on ab initio calculations, J. Mol. Struct. (Theochem) 276, 117-132 (1992).
120. Feller, D. and Feyereisen, M. W., Ab initio study of hydrogen bonding in the phenol-water sys-
tem. J. Comput. Chem. 9, 1027-1035 (1993).
121. Dibble, T. S. and Francisco, J. S., Ab initio study of the structure, binding energy, and vibrations
of the HOCl—H2O complex, J. Phys. Chem. 99, 1919-1922 (1995).
122. Odutola, J. A., Dyke, T. R., Howard, B. ]., and Mucnter, J. S., Molecular beam electric deflec-
tion study of ammonia polymers, J. Chem. Phys. 70, 4884-4886 (1979).
Geometries and Energetics I3I

123. Latajka, Z. and Schemer, S., Effects of basis set and electron correlation on the calculated prop-
erties of the ammonia dimer, J. Chem. Phys. 81, 407-409 (1984).
124. Cohen, E. A. and Poynter, R. L., The microwave spectrum of I4NH3 in the v = 000° 11 state,
J. Mol. Spectrosc. 53, 131-139 (1974).
125. Nelson, D. D. J., Fraser, G. T., and Klemperer, W., Ammonia dimer: A surprising structure,
J. Chem. Phys. 83, 6201-6208 (1985).
126. Nelson, D. D. J., Klemperer, W., Fraser, G. T. Lovas,F. J., and Suenram, R. D., Ammonia dimer:
Further structural studies, J. Chem. Phys. 87, 6364-6372 (1987).
127. Fraser, G. T., Nelson, D. D. J., Charo, A., and Klemperer, W., Microwave and infrared charac-
terization of several weakly bound NH3 complexes, J. Chem. Phys. 82, 2535-2546 (1985).
128. Nelson, D. D. J., Fraser, G. T., and Klemperer, W., Does ammonia hydrogen bond?, Science 238,
1670-1674(1987).
129. Carnovale, F, Peel, J. B., and Rothwell, R. G., Ammonia dimer studied by photoelectron spec-
troscopy, J. Chem. Phys. 85, 6261-6266 (1986).
130. Huisken, F and Pertsch, T., Infrared photodissociation of size-selected small ammonia clusters,
Chem. Phys. 126, 213-228 (1988).
131. Perchard, J.-P, Bonn, R. B., and Andrews, L., Matrix infrared and Raman spectra of the in-
equivalent submolecules in the ammonia dimer, J. Phys. Chem. 95, 2707-2712 (1991).
132. S zer, S. and Andrews, L., FTIR spectra of ammonia clusters in noble gas matrices, J. Chem.
Phys. 87, 5131-5140(1987).
133. Snels, M., Fantoni, R., Sanders, R., and Meerts, W. L., IR dissociation of ammonia clusters,
Chem. Phys. 115, 79-91 (1987).
134. Ohashi, K., Kasai, T., andKuwata, K., Characterization of a pulsed supersonic beam of ammo-
nia monomer and clusters using the hexapole electric field, J. Phys. Chem. 92, 5954-5958
(1988).
135. Latajka, Z. and Scheiner, S., The potential energy surface of(NH3)2, J. Chem. Phys 84, 341-347
(1986).
136. Sadlej, J. and Lapinski, L., Ammonia dimer, linear or cyclic?, J. Mol. Struct. (Theochem) 139,
233-240(1986).
137. Sagarik, K. P., Ahlrichs, R., and Brode, S., Intermolecular potentials for ammonia based on the
test particle model and the coupled pair functional method, Mol. Phys. 57, 1247—1264 (1986).
138. Liu, S.-Y., Dykstra, C. E., Kolenbrander, K., and Lisy, J. M., Electrical properties of ammonia
and the structure of the ammonia dimer, J. Chem. Phys. 85, 2077-2083 (1986).
139. Cook, K. D. and Taylor, J. W., A supersonic molecular beam mass spectrometric study of hy-
drogen bonding in ammonia, Int. J. Mass Spectr. Ion Phys. 30, 345-357 (1979).
140. Hassett, D. M., Marsden, C. J., and Smith, B. J., The ammonia dimer potential energy surface:
Resolution of the apparent discrepancy between theory and experiment?, Chem. Phys. Lett. 183,
449-456 (1991).
141. Tao, F.-M. and Klemperer, W., Ab initio search for the equilibrium structure of the ammonia
dimer, J. Chem. Phys. 99, 5976-5982 (1993).
142. Cybulski, S. M., Extended basis set calculations of the interaction energy and properties of the
ammonia dimer, Chem. Phys. Lett. 228, 451-457 (1994).
143. Muguet, F. F. and Robinson, G. W., Towards a new correction method for the basis set super-
position error: Application to the ammonia dimer, J. Chem. Phys. 102, 3648-3654 (1995).
144. van Bladel, J. W. I., van der Avoird, A., Wormer, P. E. S., and Saykally, R. J., Computational ex-
ploration of the six-dimensional vibration-rotation-tunneling dynamics of (NH3)2, J. Chem.
Phys. 97, 4750-4763 (1992).
145. Loeser, J. G., Schmuttenmaer, C. A., Cohen, R. C., Elrod, M. J., Steyert, D. W., Saykally, R. J.,
Bumgarner, R. E., and Blake, G. A., Multidimensional hydrogen tunneling dynamics in the
ground vibrational state of the ammonia dimer, J. Chem. Phys. 97, 4727-1749 (1992).
146. Olthof, E. H. T., van der Avoird, A., and Wormer, P. E. S., Structure, internal mobility, and spec-
trum of the ammonia dimer: Calculation of the vibration-rotation tunneling states, J. Chem.
Phys. 101, 8430-8442 (1994).
I 32 Hydrogen Bonding

147. Olthof, E. H. T., van der Avoird, A., and Wormer, P. E. S., Is (NH 3 ) 2 hydrogen bonded?, J. Mol.
Struct. (Theochem) 307, 201-215 (1994).
148. Cotti, G., Linnartz, H., Meerts, W. L., van der Avoird, A., and Olthof, E. H. T., Stark effect and
dipole moments of the ammonia dimer In different vibration-rotation-tunneling states, J. Chem.
Phys. 104,3898-3906(1996).
149. Tubergen, M. J. and Kuczkowski, R. L., Microwave spectrum and structure of the dimethylamine
dimer: Evidence for a cyclic structure, J. Chem. Phys. 100, 3377-3383 (1994).
150. Held, A. and Pratt, D. W., Ammonia as a hydrogen bond donor and acceptor in the gas phase.
Structures of 2-pyridone-NH3 and 2-pyridone-(NH 3)2 in their S() and S 1 electronic states, J. Am.
Chem. Soc. 115, 9718-9723 (1993).
151. Lewell, X. Q., Hillier, I. H., Field, M. J., Morris, J. J., and Taylor, P. J., Theoretical studies ofvi-
brationa/frequency shifts upon hydrogen bonding, J. Chem. Soc., Faraday Trans. 2 84, 893-898
(1988).
152. Kumpf, R. A. and Damewood, J. R., Jr., Interaction of formaldehyde with water, .1. Phys. Chem.
93,4478-4486(1989).
153. Dimitrova, Y. and Peyerimhoff, S. D., Theoretical study of hydrogen-bonded formaldehyde-
water complexes, J. Phys. Chem. 97, 12731-12736 (1993).
154. Ramelot, T. A., Hu, C.-H., Fowler, J. E., DeLeeuw, B. J., and Schaefer, H. F, Carbonyl-water
hydrogen bonding: The H2CO—H2O prototype, J. Chem. Phys. 100, 4347-4354 (1994).
155. Nowek, A. and Leszczynski, J., Ab initio investigation on stability and properties ofXYCO ...HZ
complexes. II: Post Hartree-Fock studies on H2CO -HF, Struct. Chem. 6, 255-259 (1995).
156. Baiocchi, F. A. and Klemperer, W., The rotational and hyperfine spectrum and structure of
H2CO-HF, J. Chem. Phys. 78, 3509-3520 (1983).
157. Rice, J. E., Lee, T. J., and Handy, N. C., The analytic gradient for the coupled pair functional
method: Formula and application for HCl, H2CO and the dimer H 2 CO . . . HCl, J. Chem. Phys.
88,7011-7023(1988).
158. Nowek, A. and Leszczynski, J., Ab initio study on the stability and properties o f X Y C O . . . H Z
complexes. III. A comparative study of basis set and electron correlation effects for H2CO —HCl,
J. Chem. Phys. 104, 1441-1451 (1996).
159. Dudis, D. S., Everhart, J. B., Branch, T. M., and Hunnicutt, S. S., Hydrogen bond energies of hy-
drogen chloride-carbonyl complexes, J. Phys. Chem. 100, 2083-2088 (1996).
160. Fraser, G. T., Gillies, C. W., Zozom, J., Lovas, F. J., and Suenram, R. D., Rotational spectrum
and structure ofH2CO—HCl, J. Mol. Spectrosc. 126, 200-209 (1987).
161. Goodwin, E. J. and Legon, A. C., Effect of weakening the hydrogen bond on the angular geom-
etry of'H 2 CO . . . HX: Evidence from the rotational spectrum ofH 2 CO . . . HCN, J. Chem. Phys. 87,
2426-2432 (1987).
162. Gavezzotti, A. and Filippini, G., Geometry of the intermolecularX—H . . . Y(X,Y = N,O) hydro-
gen bond and the calibration of empirical hydrogen-bond potentials, J. Phys. Chem. 98,
4831-4837(1994).
163. Baker, E. N. and Hubbard, R. E., Hydrogen bonding in globular proteins, Prog. Biophys. Molec.
Biol. 44, 97-179(1984).
164. Taylor, R., Kennard, O., and Versichel, W., Geometry of the N—H . . . O=C hydrogen bond. I.
Lone-pair directionality, J. Am. Chem. Soc. 105, 5761-5766 (1983).
165. Murray-Rust, P. and Glusker, J. P., Directional hydrogen bonding to sp2- andsp-hybridizedoxy-
gen atoms and its relevance to ligand-macromolecule interactions, J. Am. Chem. Soc. 106,
1018-1025(1984).
166. Ceccarelli, C., Jeffrey, G. A., and Taylor, R., A survey of O—H . . . O hydrogen bond geometries
determined by neutron diffraction, J. Mol. Struct. 70, 255-271 (1981).
167. Nowek, A. and Leszczynski, J., An ah initio study on HXC=O . . . HY molecular complexes
( X , Y = F.Cl), Int. J. Quantum Chem. 57, 757-766 (1996).
168. Ramanadham, M., Jakkal, V. S., and Chidambaram, R., Carboxyl group hydrogen bonding in X-
ray protein structures analysed using neutron studies on arnino acids, FEBS Lett. 323, 203-206
(1993).
Geometries and Energetics I 33

169. Latajka, Z., Ratajczak, H., andZeegers-Huyskens, T., Ab initio studies of hydrogen-bonded com-
plexes between formic acid or methylacetate and HF or HCl, J. Mol. Struct. (Theochem) 164,
201-209 (1988).
170. Patten, K. O., Jr. and Andrews, L., Fourier-transform infrared spectra of HF complexes with
acetic acid and methyl acetate in solid argon, J, Phys. Chem. 90, 1073-1076 (1986).
171. Nagy, P. I., Smith, D. A., Alagona, G., and Ghio, C., Ab initio studies of free and monohydrated
carboxylic acids in the gas phase, J. Phys. Chem. 98,486-493 (1994).
172. Reynolds, C. H., Methyl chloride/formic acid van der Waals complex: A model for carbon as a
hydrogen bond donor, J. Am. Chem. Soc. 112, 7903-7908 (1990).
173. Hayashi, S., Umemura, J., Kato, S., and Morokuma, K., Ab initio molecular orbital study of the
formic acid dimer, J. Phys. Chem. 88, 1330-1334 (1984).
174. Dory, M., Delhalle, J., Fripiat, J. G., and Andre, J.-M., Equilibrium geometry and electrical po-
larizability of formic acid, formamide and their cyclic hydrogen-bonded pairs, Int. J. Quantum
Chem., QBS 14, 85-103 (1987).
175. Chang, Y.-T., Yamaguchi, Y, Miller, W. H., and Schaefer, H. F, An analysis of the infrared and
Raman spectra of formic acid dimer (HCOOH)2, J. Am. Chem. Soc. 109, 7245-7253 (1987).
176. Neuheuser, T., Hess, B. A., Reutel, C., and Weber, E., Ab initio calculations of supramolecular
recognition modes. Cyclic versus noncyclic hydrogen bonding in the formic acid/formamide sys-
tem, J. Phys. Chem. 98, 6459-6467 (1994).
177. Clague, A. D. H., and Bernstein, H. J., The heat of dimerization of some carboxylic acids in the
vapour phase determined by a spectmscopic method, Spectrochim. Acta A25, 593-596 (1969).
178. Mathews, D. M. and Sheets, R. W., Effect of surface adsorption on the determination by infrared
spectroscopy of hydrogen bond energies in arboxylic acid dimers, J. Chem. Soc. A 2203-2206
(1969).
179. Winkler, A., Mehl, J. B., and Hess, P., Chemical relaxation ofH bonds in formic acid vapor stud-
ied by resonant photoacoustic spectroscopy, J. Chem. Phys. 100, 2717-2727 (1994).
180. Curtiss, L. A. and Blander, M., Thermodynamic properties of gas-phase hydrogen-bonded com-
plexes, Chem. Rev. 88, 827-841 (1988).
181. Winkler, A. and Hess, P., Study of the energetics and dynamics of hydrogen bond formation in
aliphatic carboxylic acid vapors by resonant photoacoustic spectroscopy, L Am. Chem. Soc.
116,9233-9240(1994).
182. Sauren, H., Winkler, A., and Hess, P., Kinetics and energetics of hydrogen bond dissociation in
isolated acetic acid-dl and -d4 and trifluoroacetic acid dimers, Che . Phys. Lett. 239, 313-319
(1995).
183. Borisenko, K. B., Bock, C. W., and Hargittai, L, Geometrical consequences of intermolecular
hydrogen bond formation in the formic acid and acetic acid dimers from ab initio MO calcula-
tions, J. Mol. Struct. (Theochem) 332, 161-169 (1995).
184. Somasundram, K., Amos, R. D., and Handy, N. C., Ab initio calculation for properties of hy-
drogen bonded complexes H 3 N ... HCN, HCN ... HCN, HCN ... HF, H 2 O . . . HF, Theor. Chim. Acta
69,491-503(1986).
185. Benzel, M. A. and Dykstra, C. E., The nature of hydrogen bonding in the NN—HF, OC _ _ HF,
and HCN—HF complexes, J. Chem. Phys. 78, 4052-4062 (1983).
186. Bachrach, S. M., Chiles, R. A., and Dykstra, C. EL, Application of an approximate double sub-
stitution coupled cluster (ACCD) method to the potential curves of CO and NeHe: Higher order
correlation effects in chemically and weakly bonded molecules, J. Chem. Phys. 75, 2270-2275
(1981).
187. Wofford, B. A., Eliades, M. E., Lieb, S. G., and Bevan, J. W., Determination of dissociation en-
ergies and thermal functions of hydrogen-bond formation using high resolution FTIR spec-
troscopy, J. Chem. Phys. 87, 5674-5680 (1987).
188. Andrews, L. and Hunt, R. D., Matrix trapping of two structural arrangements of weak com-
plexes, J. Phys. Chem. 92, 81-85 (1988).
189. Chattopadhyay, S. and Plummer, P. L. M., Ab initio studies of the mixed heterodimers of ammo-
nia and hydrogen cyanide, Chem. Phys. 782, 39-51 (1994).
I 34 Hydrogen Bonding

190. Bohn, R. B. and Andrews, L., Matrix infrared spectra of the NH 3 _ _ HCN and NH3—(HCN)2
complexes in solid argon, J. Phys. Chem. 93, 3974__3979 (1989).
191. Fraser, G. T., Leopold, K. R., Nelson, D. D. J., Tung, A., and Klemperer, W., The rotational spec-
trum and structure ofNH 3 —HCN, J. Chem. Phys. 80, 3073-3077 (1984).
192. Legon, A. C. and Willoughby, L. C. The rotational spectrum and structure of the phosphine-
hydrogen cyanide complex, Chem. Phys. 85,443-150 (1984).
193. Legon, A. C. and Willoughby, L. C., First-order Stark effect and electric dipole moment of
H 3 P ... HC 15 N by pulsed-nozzle Fourier-transform microwave spectroscopy, Chem. Phys. Lett.
111,566-570(1984).
194. Goodwin, E. J. and Legon, A. C., Geometry of a weakly bound complex (H2S,HCN) determined
from an investigation of its rotational spectrum, J. Chem. Soc., Faraday Trans. 2 80, 1669-1682
(1984).
195. Boyd, R. J., Hydrogen bonding between nitriles and hydrogen halides and the topologicalprop-
erties of molecular charge distributions, Chem. Phys. Lett. 129, 62-65 (1986).
196. Del Bene, J. E., Mettee, H. D., and Shavitt, L, Structure, binding energy, and vibrational fre-
quencies ofCH 3 CN . . . HCl, J. Phys. Chem. 95, 5387-5388 (1991).
197. Legon, A. C., Millen, D. J., and North, H. M., Rotational spectrum of the hydro gen-bonded dimer
CH 3 CN ... HCl, J. Phys. Chem. 91, 5210-5213 (1987).
198. Ballard, L. and Henderson, G., Hydrogen bond energy ofCH 3 CN—HCl by FTIR photometry,
J. Phys. Chem. 95, 660-663 (1991).
199. Legon, A. C., Millen, D. J., and North, H. M., Dissociation energies of the hydrogen-bonded
dimers RCN ... HF (R = CH3, HCC) determined by rotational spectroscopy, J. Chem. Phys. 86,
2530-2535 (1987).
200. Boyd, R. J. and Choi, S. C., A bond-length-bond-order relationship for intermolecular interac-
tions based on the topological properties of molecular charge distributions, Chem. Phys. Lett.
120,80-85(1985).
201. Kofranek, M., Lischka, H., and Karpfen, A., Ab initio studies on hydrogen-bonded clusters. L
Linear and cyclic oligomers of hydrogen cyanide, Chem. Phys. 113, 53—64 (1987).
202. Kofranek, M., Lischka, H., and Karpfen, A., Ab initio studies on structure, vibrational spectra
and infrared intensities ofHCN, (HCN)2, and (HCN)3, Mol. Phys. 61, 1519-1539 (1987).
203. Migchels, P., Zeegers-Huyskens, T., and Peeters, D., Fourier transform infrared and theoretical
studies of alkylimines complexes with hydroxylic proton donors, J. Phys. Chem. 95, 7599-7604
(1991).
204. Eberhardt, E. S. and Raines, R. T., Amide-amide and amide-water hydrogen bonds: Implications
for protein folding and stability, J. Am. Chem. Soc. 116, 2149-2150 (1994).
205. Johansson, A., Kollman, P., Rothenberg, S., and McKelvey, J., Hydrogen bonding ability of the
amide group, J. Am. Chem. Soc. 96, 3794-3800 (1974).
206. Zielinski, T. J. and Poirier, R. A., Examination offormamide, formamidic acid, amidine dimers,
and the double proton transfer transition states involving these dimers, J. Comput. Chem. 5,
466-470(1984).
207. Sapse, A. M., Fugler, L. M., and Cowburn, D., An ab initio study of intermolecular hydrogen
bonding between smallpeptide fragments, Int. J. Quantum Chem. 29, 1241-1251 (1986).
208. stergard, N., Christiansen, P. L., and Nielsen, O. F., Ab initio investigation of vibrations in free
and hydrogen bondedformamide, J. Mol. Struct. (Theochem) 235, 423-446 (1991).
209. Novoa, J. J. and Whangbo, M.-H., The nature of intramolecular hydrogen-bonded and non-
hydrogen-bonded conformations of simple di- and triamides, J. Am. Chem. Soc. 113,9017-9026
(1991).
210. Jasien, P. G. and Stevens, W. J., Ab initio study of the hydrogen bonding interactions of for-
mamide with water and methanol, J. Chem. Phys. 84, 3271-3277 (1986).
211. Lovas, F. J., Suenram, R. D., Fraser, G. T, Gillies, C. W., and Zozom, J., The microwave spec-
trum of formamide-water and formamide-methanol complexes, J. Chem. Phys. 88, 722-729
(1988).
Geometries and Energetics 135

212. Bohn, R. B. and Andrews, L., Infrared spectra ofamide-HF complexes in solid argon, J. Phys.
Chem. 93, 5684-5692 (1989).
213. Sneddon, S. F, Tobias, D. J., and Brooks, C. L. L, Thermodynamics of amide hydrogen bondfor-
mation in polar and apolar solvents, J. Mol. Biol. 209, 817-820 (1989).
214. Tobias, D. J., Sneddon, S. F., and Brooks, C. L. L, Reverse turns in blocked dipeptides are in-
trinsically unstable in water, J. Mol. Biol. 216, 783-796 (1990).
215. Dixon, D. A., Dobbs, K. D., and Valentini, J. J., Amide-water and amide-amide hydrogen bond
strengths, J. Phys. Chem. 98, 13435-13439 (1994).
216. Guo, H. and Karplus, M., Ab initio studies of hydrogen bonding in N-methylacetamide: Struc-
ture, cooperativity, and internal rotation barriers, J. Phys. Chem. 96, 7273-7287 (1992).
217. Guo, H. and Karplus, M., Solvent influence on the stability of the peptide hydrogen bond: A
supramolecular cooperative effect, J. Phys. Chem. 98, 7104-7105 (1994).
218. Markham, L. M. and Hudson, B. S., Ab initio analysis of the effects of aqueous solvation on the
resonance Raman intensities of N-methylacetamide, J. Phys. Chem. 100, 2731-2737 (1996).
219. Shimizu, H., Nagata, K., and Sasaki, S., High-pressure Raman study of the hydrogen-bonded
crystalline formamide, J. Chem. Phys. 89, 2743-2747 (1988).
220. Ladell, J. and Post, B., The crystal structure of formamide, Acta Cryst. 7, 559-564 (1954).
221. Stevens, E. D., Low-temperature experimental electron density distribution of formamide, Acta
Cryst. B34,544-551(1978).
222. Suhai, S., Structure and bonding in the formamide crystal: A complete fourth-order many-body
perturbation theoretical study, J. Chem. Phys. 103, 7030-7039 (1995).
223. Florian, J. and Johnson, B. G., Structure, energetics, and force fields of the cyclic formamide
dimer: MP2, Hartree-Fock, and density functional study, J. Phys. Chem. 99, 5899-5908 (1995).
224. Shimoni, L., Glusker, J. P., and Bock, C. W., Energies and geometries of isographic hydrogen-
bonded networks. L The R22(8) graph set, J. Phys. Chem. 100, 2957-2967 (1996).
225. Ludwig, R., Weinhold, P., and Farrar, T. C., Temperature dependence of hydrogen bonding in
neat, liquid formamide, J. Chem. Phys. 103, 3636-3642 (1995).
226. Aleman, C., Navas, J. J., and Munoz-Guerra, S., Study of the amide . . . ster hydrogen bond in
small molecules and its influence on the conformation of polypeptides and related polymers,
J. Phys. Chem. 99, 17653-17661 (1995).
227. Devoe, H. and Tinoco, L, The stability of helical polynucleotide: Base contributions, J. Mol.
Biol. 4, 500-517(1962).
228. Pullman, B., Claverie, P., and Caillet, J., Van der Waals-London interactions and the configuration
of hydrogen-bondedpurine andpyrimidinepairs, Proc. Nat. Acad. Sci., USA 55, 904-912 (1966).
229. Kudritskaya, Z. G. and Danilov, V. I., Quantum mechanical study of bases interactions in vari-
ous associates in atomic dipole approximation, J. Theor. Biol. 59, 303—318 (1976).
230. Scheiner, S. and Kern, C. W., Molecular orbital investigations of multiply hydrogen bonded sys-
tems. Formic acid dimer and DNA base pairs, J. Am. Chem. Soc. 101,4081-4085 (1979).
231. Sagarik, K. P. and Rode, B. M., The influence of the Mg2+ ion on the hydrogen bonds of the ade-
nine-thymine base pair, Inorg. Chim. Acta 78, 177-180 (1983).
232. Ohta, Y., Tanaka, H., Baba, Y., Kagemoto, A., and Nishimoto, K., Solvent effect on the hydro-
gen-bonding interaction between adenine and uracil, J. Phys. Chem. 90, 4438-4442 (1986).
233. Hobza, P. and Sandorfy, C., Nonempirical calculations on alt the 29 possible DNA base pairs,
J. Am. Chem. Soc. 109, 1302-1307 (1987).
234. Anwander, E. H. S., Probst, M. M., and Rode, B. M., The influence ofLi+, Na+, Mg+, Ca2+,
andZn2+ ions on the hydrogen bonds of the Watson-Crick base pairs, Biopolymers 29, 757—769
(1990).
235. Hrouda, V., Florian, J., and Hobza, P., Structure, energetics and harmonic vibrational spectra
of the adenine-thymine and adenine*-thymine* base pairs: Gradient nonempirical and semi-
empirical study, J. Phys. Chem. 97, 1542-1557 (1993).
236. Gould, I. R. and Kollman, P. A., Theoretical investigation of the hydrogen bond strengths in gua-
nine-cytosine and adenine-thymine base pairs, J. Am. Chem. Soc. 116, 2493-2499 (1994).
I 36 Hydrogen Bonding

237. Aida, M., Characteristics of the Watson-Crick type hydrogen-bonded DNA base pairs: An ab
initio molecular orbital study, J. Comput. Chem. 9, 362-368 (1988).
238. Dive, G., Dehareng, D., and Ghuysen, J. M., Energy analysis on small to medium sized H-bonded
complexes, Theor. Chim. Acta 85, 409-421 (1993).
239. Jiang, S.-P., Raghunathan, G., Ting, K.-L., Xuan, J. C., and Jernigan, R. L., Geometries, charges,
dipole moments and interaction energies of normal, tautomeric and novel bases, J. Biomol.
Struct. Dyn. 12, 367-382 (1994).
240. Florian, J. and Leszczynski, J., Theoretical investigation of the molecular structure of the k
DNA base pair, J. Biomol. Struct. Dyn. 12, 1055-1062 (1995).
241. Pranata, J., Wierschke, S. G., and Jorgensen, W. L., OPLS potential functions for nucleotide
bases. Relative association constants of hydro gen-bonded base pairs in chloroform, J. Am.
Chem. Soc. 113, 2810-2819(1991).
242. Trollope, K. L, Gould, I. R., and Hillier, L H., Modelling of electrostatic interactions between
nucleotide bases using distributed multipoles, Chem. Phys. Lett. 209, 113-116 (1993).
243. Colson, A.-O., Besler, B., Close, D. M., and Sevilla, M. D., Ab initio molecular orbital calcula-
tions of DNA bases and their radical ions in various protonation states: Evidence for proton
transfer in GC base pair radical anions, J. Phys. Chem. 96, 661-668 (1992).
244. Aida, M., Kaneko, M., and Dupuis, M., An ab initio MO study on the thymine dimer and its rad-
ical cation, Int. J. Quantum Chem. 57, 949-957 (1996).
245. Sponer, J., Leszczynski, J., and Hobza, P., Structures and energies of hydrogen-bonded DNA
base pairs. A nonempirical study with inclusion of electron correlation, J. Phys. Chem. 100,
1965-1974(1996).
246. Yanson, I. K., Teplitsky, A. B., and Sukhodub, L. F., Experimental studies of molecular interac-
tion between nitrogen bases of nucleic acids, Biopolymers 18, 1149-1170(1979).
247. Hobza, P., Sponer, J., and Polasek, M., H-bonded and stacked DNA base pairs: cytosine dimer.
An ab initio second-order Moller-Plesset study, J. Am. Chem. Soc. 117, 792-798 (1995).
248. McDowell, S. A. C. and Buckingham, A. D., Isotope effects on the stability of the carbon monox-
ide-acetylene van der Waals molecule and the hydrogen fluoride dimer, Chem. Phys. Lett. 182,
551-555 (1991).
249. Curtiss, L. A. and Pople, J. A., Ab initio calculation of the force field of the hydrogen fluoride
dimer, J. Mol. Spectrosc. 61, 1-10 (1976).
250. Schuder, M. D. and Nesbitt, D. J., High resolution near infrared spectroscopy of'HCl—DCl and
DCl—HCl: Relative binding energies, isomer conversion rates, and mode specific vibrational
predissociation, J. Chem. Phys. 100, 7250-7267 (1994).
251. Engdahl, A. and Nelander, B., On the relative stabilities of H-and D-bonded water dimers,
J. Chem. Phys. 86, 1819-1823 (1987).
252. Nelander, B. and Nord, L., Complex between water and ammonia, J. Phys. Chem. 86, 4375-4379
(1982).
253. Engdahl, A. and Nelander, B., The intramolecular vibrations of the ammonia water complex. A
matrix isolation study, J. Chem. Phys. 91, 6604-6612 (1989).
254. Nelander, B., Infrared spectrum of the water formaldehyde complex in solid argon and solid ni-
trogen, J. Chem. Phys. 72, 77-84 (1980).
255. Engdahl, A. and Nelander, B., Water-olefin complexes. A matrix isolation study, J. Phys. Chem.
90,4982-1987 (1986).
256. Fillery-Travis, A. J., Legon, A. C., Willoughby, L. C., and Buckingham, A. D., Rotational spec-
troscopy of 15 N-hydrogen cyanide dimer: Detection, relative stability and D-nuclear quadrupole
coupling of deuterated species, Chem. Phys. Lett. 102, 126-131 (1983).
257. Legon, A. C. and Millen, D. J., Systematic effect ofD substitution on hydrogen-bond lengths in
gas-phase dimers B . . . HX and a model for its interpretation, Chem. Phys. Lett. 147, 484-489
(1988).
258. Edison, A. E., Markley, J. L., and Weinhold, F., Ab initio calculations of pmtium/deuteriumfrac-
tionation factors in O 2 H 5 + clusters, J. Phys. Chem. 99, 8013-8016 (1995).
Geometries and Energetics 137

259. Graul, S. T., Brickhouse, M. D., and Squires, R. R., Deuterium isotope fractionation within pro-
tonated water clusters in the gas phase, J. Am. Chem. Soc. 112, 631-639 (1990).
260. Weil, D. A. and Dixon, D. A., Gas-phase isotope fractionation factor for proton-bound dimers
of methoxide anions, J. Am. Chem. Soc. 107, 6859-6865 (1985).
261. Larson, J. W. and McMahon, T. B., Isotope effects in proton-transfer reactions. An ion cyclotron
resonance determination of the equilibrium deuterium isotope effect in the bichloride ion,
J. Phys. Chem. 91, 554-557 (1987).
262. Scheiner, S. and Cuma, M., On the relative stability of hydrogen and deuterium bonds, J. Am.
Chem. Soc. 118, 1511-1521 (1996).
263. Astrand, P.-O., Karlstrom, G., Engdahl, A., and Nelander, B., Novel model for calculating the
intermotecular part of the infrared spectrum for molecular complexes, J. Chem. Phys. 102,
3534-3554 (1995).
264. Edison, A. E., Weinhold, E, and Markley, J. L., Theoretical studies of protium/deuterium frac-
tionation factors and cooperative hydrogen bonding in clusters, J. Am. Chem. Soc. 117,
9619-9624(1995).
265. Harris, N. J., A systematic theoretical study of harmonic vibrational frequencies and deuterium
isotope fractionation factors for small molecules, J. Phys. Chem. 99, 14689-14699 (1995).
3

Vibrational Spectra

n contrast to the deep minimum in the potential energy surface that corresponds to the
I equilibrium geometry of a standard molecule, the equilibrium structure of a typical H-
bonded complex resides in a much shallower minimum. The PES of the complex is over-
all much flatter, yet commonly contains other minima, albeit not quite as stable, which are
energetically accessible from the global minimum. This chapter focuses on the nature of
the potential energy surface in the vicinity of the equilibrium structure, deferring until later
a discussion of more extended regions of the surface. A principal question addressed here
is the steepness of the surface as one moves away from the global minimum in various
directions, and the source of this slope. This point is directly examined by theoretical
calculations which can determine the energy of any given geometrical distortion of the
structure.
The shape of the minimum in the surface is experimentally probed by vibrational spec-
troscopy. It is here that the computations can make direct connection with experimental in-
formation. Formation of the H-bond from a pair of isolated molecules converts three trans-
lational and three rotational degrees of freedom of the formerly free pair of molecules into
six new vibrations within the complex. The frequencies of these modes are indicative of the
functional dependence of the energy upon the corresponding geometrical distortions. But
rather than consisting of a simple motion, for example, H-bond stretch, the normal modes
are composed of a mixture of symmetry-related atomic motions, complicating their analy-
sis in terms of the simpler motions. In addition to these new intermolecular modes, the in-
tramolecular vibrations within each of the subunits are perturbed by the formation of the H-
bond. The nature of each perturbation opens a window into the effects of the H-bond upon
the molecules involved. The intensities of the various vibrations carry valuable information
about the electron density within the complex and the perturbations induced by the forma-
tion of the H-bond.

138
Vibrational Spectra I 39

3.1 Method of Calculation

Most ab initio analyses of vibrational spectra invoke a "double-harmonic" assumption


wherein the potential energy surface in the vicinity of the minimum is fit to a function that
involves only quadratic dependence of the energy with respect to the nuclear motions. The
intensities of the normal vibrational modes are extracted from the derivatives of the dipole
moment, taken as linear with respect to nuclear coordinates. Within this approximation, the
intensities of the fundamentals are proportional to the square of the dipole moment deriva-
tives with respect to normal coordinates 1-3.
The harmonic approximation ignores third and higher-order terms in the dependence of
the energy upon nuclear coordinates. In the case of most molecules, the approximation of
the shape of the well as a parabola is a reasonable one, at least for small displacements from
equilibrium. It is a straightforward matter for computer programs that evaluate the energy
derivatives to go on and elucidate the normal vibrational modes and their associated fre-
quencies, using a standard GF matrix formulation4. Analysis of the modes allows one to de-
termine the real nature of a given vibration, for example, stretching, bend, or some combi-
nation. The frequency provides a measure of the "stiffness" of the potential with respect to
the particular motion involved.
Experience has shown that SCF-level calculations typically overestimate the magnitude
of the vibrational frequencies by some 5-10%, even with very large flexible basis sets. This
exaggeration is due in part to the inability of the SCF wave function to dissociate properly.
A compilation of a great deal of data5 has suggested that SCF/6-31G* frequencies should
be multiplied by a scaling factor of 0.893 for best reproduction of experimental values,
while the factor for MP2/6-31G* is 0.943. Such a prescription is estimated to reproduce ex-
perimental frequencies to 50 cm-1 as a root mean square error. In the case of H-bonds, the
intermolecular part of the potential is subject to greater degrees of anharmonicity, so har-
monic frequencies should be treated with caution, even if the energetics are computed with
a high level of theory.
The symbol v is normally used to express the frequency of any given vibrational mode,
in units of c m - 1 . Since most calculations are restricted to the harmonic approximation, the
use of this symbol in the computational literature likewise refers generally to harmonic fre-
quencies. In those cases where anharmonicity is added to the computations, the notation can
become confusing in that v usually refers to the anharmonic value, with reserved for the
harmonic approximation to this frequency. The reader should therefore exercise some cau-
tion in scanning the original literature. In this text, we will adopt the convention that v will
represent the harmonic frequency; in those cases where anharmonicity is included, the dis-
tinctions and notation will be clearly delineated.
The integrated infrared band intensity of the ith mode is measured experimentally as

where C and L refer to the concentration and the optical path length, respectively, and Io
and I the intensities of the incident and transmitted light. This same quantity may be ex-
pressed within the confines of the double harmonic approximation as
140 Hydrogen Bonding

where Qi is the normal coordinate and the dipole moment vector of the molecular system.
The sum extends over any g degeneracies of the mode. NA is Avogadro's constant, c the
speed of light, and o the vacuum permittivity. The ratio (vi/ i) allows for incorporation of
anharmonic effects if desired, but is generally taken as unity. If the summed expressions are
reported in the usual units, the intensity can be expressed in km mol-1 with the leading co-
efficient as below.

Somasundram et al.6 have derived an analogous expression for Raman intensities, based
upon derivatives of the isotropic, a, and anisotropic, , polarizabilities, with respect to the
normal coordinate Q:

As in the preceding equations, the symbols A and S will be used to refer to infrared and Ra-
man intensities, respectively, throughout this text.
Nuclear quadrupole coupling constants (NQCCs) constitute another window into the
fundamental attributes of the hydrogen bonding phenomenon. Bacskay et al.7 decompose
the NQCC, , at a nucleus A of a H-bonded complex in terms of this same property in the
unperturbed molecule, o, and its shift due to complexation, .

represents the instantaneous angle between the molecule HA and the inertial axis. The
brackets indicate librational motion averaging. For small angles, can be approximated as

The NQCC can be related to the electric field gradient through the nuclear quadrupole mo-
ment at atom A, eQ(A), and the vibrationally averaged component of the electric field gra-
dient vector

where the inertial axis is defined as the z-axis. One then can write an expression for the lat-
ter vibrationally averaged quantity as

in terms of the electric field gradient at A in free HA and its change due to formation of the
complex.
There are two vibrational modes that are of particular interest in H-bonding studies. The
first is the stretching vibration within the proton donor molecule, involving chiefly the
bridging hydrogen. This mode is commonly referred to as v . Over the years, it has been
learned that the shift which this frequency undergoes when the H-bond is formed, relative
to the uncomplexed A—H molecule, correlates very well with the strength of the H-bond.
This correlation is typically referred to as the Badger-Bauer rule 8-11 .
Vibrational Spectra 141

The other mode of particular interest is the stretching between the two subunits. This stretch,
denoted v , can be contaminated by minor amounts of internal distortions but is of much
lower frequency. Other intermolecular vibrations involve combinations of stretches, wags,
and bends; they are referred to by various nomenclature systems.

3.2 Accuracy Considerations

Before directly considering the perturbations of each subunit that occur upon formation of
a H-bond, it would be worthwhile to first establish the sorts of accuracy that one may ex-
pect in calculating vibrational spectra of individual molecules. A number of popular basis
sets, ranging from minimal to extended, were examined for this purpose12. Frequencies
computed for the HF stretch and for the three internal vibrations of H2O are listed in Table
3.1, along with experimental estimates in the final row. It is immediatly clear that SCF-level
calculations overestimate the frequencies whether they are stretches or bends. This is not
surprising as it is known that correlation is required to obtain better agreement with exper-
iment. The largest sets seem to be converging to a limit of accuracy. However, this limit
does not require very large basis sets, as comparable results can be achieved even with dou-
ble- polarized sets. Polarization functions seern to be required for good accuracy, even if
only a single set of such functions. With respect to the smaller sets, the minimal basis set
provides particularly large exaggerations of the frequencies. 3-21G seems to do surprisingly
well for a set of this size, at least for this pair of molecules. The MP2 and CISD correlated
methods reduce the frequencies a good deal, lowering the overestimation to only a few per-
cent. MP2 seems superior to CISD.

Table 3.1 Computed frequencies (cm - 1 ) of vibrational bands in HF and H2O. Data are at SCF
level unless otherwise indicated12.

H2O

Basis set HF v1 v2 v3

STO-3G 4475 4140 2170 4391


3-21G 4061 3812 1799 3946
DZ 4228 4028 1711 4204
TZ 4224 3990 1723 4156
6-31G(d) 4358 4070 1827 4189
MP2 3941 3748 1682 3894
CISD 4000 3800 1712 3920
DZP 4490 4152 1750 4267
6-31+G(d) 4314 4071 1797 4190
TZP 4454 4114 1748 4226
6-31 + +G(d,p) 4492 4143 1726 4245
TZ2P 4457 4116 1751 4221
6-31l + +G(3d,3p) 4486 4129 1757 4229
expt 3962 3657 1595 3756
142 Hydrogen Bonding

Table 3.2 Computed intensities (D 2. amu -1 • A - 2 ) of vibrational bands in HF and H2O. Data
are at SCF level unless otherwise indicated12.

H2O

Basis set HF v
1
V
2
V
3

STO-3G 0.90 1.05 0.17 0.71


3-2 1G 0.78 0.001 1.89 0.22
DZ 2.28 0.08 3.21 1.54
TZ 1.95 0.02 3.11 1.26
6-31G(d) 3.35 0.43 2.54 1.38
MP2 3.19 0.28 2.52 1.68
CISD 2.84 0.21 2.44 1.27
DZP 4.11 0.52 2.40 1.81
6-31+G(d) 4.28 0.55 2.82 2.06
TZP 3.68 0.42 2.32 1.57
6-31 + +G(d,p) 4.54 0.60 2.03 2.09
TZ2P 3.90 0.35 2.28 2.14
6-311 + +G(3d,3p) 3.74 0.35 2.22 2.10
expt 0.06 1.2-1.6 1.0-1.4

Table 3.2 contains analogous information concerning the intensities of the various vi-
brational modes. Comparison with the data in the last row illustrates the difficulty in com-
puting experimental intensities to a high degree of accuracy. As in the case of the frequen-
cies, it appears that the DZP basis set seems to perform about as well as any of the much
more extended sets in a number of instances. Deletion of the polarization functions leads to
erratic fluctuations in the intensities, particularly the V1 symmetric stretching frequency of
water. Minimal and 3-21G are especially bad and should be avoided. Correlation does not
influence the intensities in an obvious predictable manner.
A different study 13 provides some further information concerning the effects of electron
correlation upon calculated intensities. In addition to SCF and MPn type methods, these au-
thors considered also the coupled-cluster model which is an infinite-order generalization of
MP. Coupled-cluster, limited to single and double excitations (CCSD), was considered, as
well as CCSD + T which includes the effects of connected triple excitations. The double-
harmonic approximation was made, consistent with most calculations of H-bonded systems.
The [5s3pld/3slp] basis set is of the polarized double- variety, within reach of most
workers.
Results computed for the HF stretch and for the three internal vibrations of H2O are listed
in Table 3.3, along with experimental estimates of the intensities in the final row. Compar-
ison with the first row indicates that SCF intensities seem to be consistently too high, some-
times by a factor of less than two, but by an order of magnitude in the case of the symmet-
ric stretch of water. Inclusion of correlation immediately lowers all intensities. MP2 appears
satisfactory for all modes, with the exception of v1 for H2O. Higher levels of correlation be-
yond MP2 introduce smaller reductions, in some cases even lower intensities than experi-
mentally observed.
Comparison of various modes in additional molecules led to the conclusion that the cal-
culations of stretches of terminal A—H bonds are particularly demanding. With regard to
basis set requirements, there seems to be little benefit in enlarging the set beyond the DZP
Vibrational Spectra 143

Table 3.3 Computed intensities (km/mol) of vibrational bands in HF and H2O.


All data calculated with [5s3pld/3slp] basis set13.

H2O

Method HF V1 v2 v3

SCF 168.8 19.2 84.6 57.7


MP2 116.7 8.9 64.9 45.7
MP4 94.2 4.8 61.7 30.4
CCSD 99.4 5.8 64.1 31.9
CCSD+T 93.3 4.6 62.0 29.2
expt 100-102 2.2 54-64 40-48

level. On the other hand, it is essential that every atom, hydrogen included, contain at least
one set of polarization functions. The intensities are quite insensitive to the values chosen
for the exponents of these polarization functions.
Similar sorts of conclusions apply to the frequencies. A systematic study14 found that a
DZP basis set yields vibrational frequencies within about 9% of experimental (harmonic)
values. The discrepancy diminishes to 4% when correlation is included via CISD and to 2%
with a coupled cluster treatment. Another set of calculations15 confirmed the cost-effec-
tiveness of the MP2 treatment of vibrational frequencies, indicating better agreement with
experiment than MP3 on some occasions. Certain types of modes can be more sensitive to
the level of theoretical treatment than others. For example, out-of-plane bending motions
for -bonded systems can require triple- plus two sets of polarization functions, as well as
a set of f-functions in the basis set16.
In general, the results just reported, as well as other work,17-19, indicate that infrared in-
tensities are much more sensitive to the level of theory than are the vibrational frequencies.
In certain cases, correlation-induced changes can be small, but this is not a general rule.
Without polarization functions, the vibrational spectra are apt to contain errors so large as
to render the data potentially useless.

3.3 (HX)2

As a first example, let us consider the simplest possible H-bond between a pair of diatomics,
as in (HF)2. Prior to formation of the complex, each molecule has a single vibrational mode,
with a particular frequency and intensity. Table 3.4 collects data calculated over the years
at various levels, for the harmonic frequencies in the monomer and the dimer20-27. The up-
per part of the table illustrates the expected overestimate of the vibrational frequency by
SCF calculations. The correlated frequencies for the monomer in the first column of data
better match the experimental quantity in the last row of the table, although there still re-
mains some scatter from one method to the next. It does appear that a correlated treatment
such as MP2, in conjunction with a large flexible basis set, can successfully reproduce the
harmonic aspects of the experimental spectrum.
Of greater interest than the frequency of the monomer is the change this property un-
dergoes as a result of formation of a H-bond. The next two columns of Table 3.4 indicate
that the frequencies of both the proton donor and acceptor groups shift toward the red. One
144 Hydrogen Bonding

Table 3.4 Vibrational frequencies (cm - 1 ) of HF in the monomer and changes undergone in
the dimer.

v, dimer

Method v, monomer acceptor donor Reference

SCF
DZ 4103 -33 -82 [20]
6-31G** 4495 -41 -89 [211
6-31+G* 4314 -35 -72 [22|
DZP 4440 -43 -91 [201
+VPs(2d)s 4488 -40 -94 [23]
correlated
CI/DZ 3732 -10 -40 [20]
CI/DZP 4150 -47 -105 [20]
SCEP/TZP 3994 -18 -74 [24]
MP2/6-31+G* 3941 -36 -89 [25]
MP2/6-311 + +G(2d,2p) 4170 -43 -116 [25]
CCSD(T)/TZ2P(f,d) 4157 -38 -107 [26]
expt 4139a -31 -94 [27]
a
Harmonic frequency

may take this frequency drop as an indicator that each internal H—F covalent bond is weak-
ened somewhat by the complexation. There is a good deal of scatter amongst the calculated
data. In fact, in some respects, the SCF frequency shifts seem superior to the correlated data,
at least in the aggregate. For example, the correlated donor shifts vary in magnitude be-
tween 40 and 116 c m - 1 , whereas the SCF values are in the narrower range of 72-94 c m - 1 .
Because of their sensitivity to small redistributions of electron density, the computation
of the intensities of vibrational modes has proven to be more demanding than the frequen-
cies. Table 3.5 reports calculations of the intensities at the SCF and correlated levels. The
intensity of the internal vibration of the proton-acceptor molecule is changed only little by
the perturbation, but that of the donor undergoes a large increase by a factor of three or four.
The latter intensification is characteristic of H-bonds and will be seen repeatedly.
When two diatomics such as HF are combined, there are new intermolecular vibrational
modes generated. These modes can sometimes be easily recognized as a particular sort of

Table 3.5 Intensities of vibrational modes (km mol - 1 ) of HF in the monomer and dimer.
Dimer

Method Monomer acceptor donor Reference

SCF
6-31G** 132 182 376 [21]
+VPs(2d)s 168 167 459 [23]
CCSD(T)
TZ2P(f,d) 103 120 427 [26]
Vibrational Spectra 145

stretch or bend, but in many cases are instead a complex mixture of various types of dis-
tortion. In the case of (HF)2, there are three vibrations occurring within the plane of the
dimer and one out-of-plane distortion. In an early work, Curtiss and Pople28 carried out a
detailed analysis of the composition of these normal modes. The results were computed with
only a 4-31G basis set and at the SCF level but are illustrative of the motions nonetheless.
The intermolecular mode of frequency 226 c m - 1 in Table 3.6 is composed largely of the
stretch between the two F atoms, and can be designated as v . But the contamination by
bending of the two HF molecules is not negligible. There is also a good deal of this H-bond
stretching motion in the lowest-frequency mode, which represents primarily the bending by
the proton acceptor molecule. The mode at 588 cm-1 can be identified with bending of the
proton donor molecule, but not without appreciable bending of the acceptor as well. The
fourth mode is of different symmetry, and is a pure torsional motion, that is, an out-of-plane
(oop) bend. The reader should thus be alerted that designations of certain vibrational mo-
tions can be misleading since they are generally a combination of various pure motions.
How do these notions about intermolecular modes vary with larger basis sets and with
correlation effects? The frequencies listed in Table 3.7 indicate that the fundamental natures
of these modes change little. The intermolecular modes have frequencies in the range be-
tween about 120 and 600 cm - 1 , and as such are clearly separated from the much higher-
frequency intramolecular vibrations. The highest-energy distortion mode is an in-plane
bend. Analysis of this mode shows it to be composed largely of a wagging of the proton
donor molecule, distorting the linearity of the H-bond. The torsional or out-of-plane mo-
tion is perhaps 100 cm-1 smaller, followed by the other two in-plane distortions. That mode
which seems to most closely resemble an intermolecular stretch appears at around
200cm - 1 .
The intensities reported in Table 3.8 appear to be quite sensitive to particular features of
the basis set or to inclusion of electron correlation. For example, the intensity of the low-
est-frequency mode changes by a factor of 25 upon going from 6-31G** to the larger
+VP s (2d) s . Some of this difference is due to the change in the nature of the motion itself
since the two basis sets are associated with different potential energy surfaces. It might be
stressed finally that the particular nomenclature, that is, assignment, of certain bands can
be nettlesome. For example, both the V and in-plane bending modes of (HF)2 contain ele-
ments of F—F stretch and bends.

Table 3.6 Composition of intermolecular vibrational modes in (HF)228.

v, cm-1

171 -2.6( R) + 0.3(r d d) - 1.0 (ra a)


226 -3.2 ( R) - 0.3 (rd d) + 0.6 (r a)
588 -0.1 ( R) - 1.2(r d d) - 0.4 (ra a)
519 0.06(R )
146 Hydrogen Bonding

Table 3.7 Intermolecular vibrational frequencies ( c m - 1 ) of HF dimer.


ip Bend oop Bend v ip Bend Reference

SCF
DZ 519 475 189 165 [20]
6-31G** 601 455 230 127 [21]
6-31 + G* 551 463 221 154 [22]
DZP 529 442 193 143 [20]
+VP s (2d) s 522 433 206 142 [23]
correlated
CI/DZ 523 459 196 166 [20]
CI/DZP 607 486 218 156 [20]
SCEP/TZP 420 — 167 127 [24]
MP2/6-31+G* 607 497 243 174 [25]
MP2/6-311 + +G(2d,2p) 582 516 231 163 [25]
CCSD(T)/TZ2P(f,d) 567 458 157 210 [26]

Bunker et al.29,30 have attempted to calculate anharmonicities and tunneling splittings


for the HF dimer. Starting with a large number of single point energies on the potential en-
ergy surface, the authors fit an empirical analytic expression with 39 adjustable parameters.
Their calculated anharmonic frequencies for V1 and V2 are 3926 and 3874 cm - 1 , respec-
tively, which compare rather well with the experimental values of 3931 and 386831.
Some of the vibrational frequencies of the analogous HC1 dimer have been computed by
Karpfen et al.32 using a correlated scheme and rather large basis sets. The data in Table 3.9
indicate trends similar to (HF)2, albeit of lesser magnitude. That is, the stretches of both the
acceptor and donor are red-shifted, the latter by a greater amount. The intensity of the donor
stretching mode is considerably greater than in the case of the acceptor molecule. The cor-
responding data for the intermolecular modes are reported in Table 3.10. Comparison with
data in Tables 3.7 and 3.8 for the HF dimer reveal uniformly smaller frequencies for all
modes. The H-bond stretching motion is the lowest in frequency for (HC1)2, unlike (HF)2
where one of the bends is of lower frequency. v is quite small, less than 70 c m - 1 . In ad-
dition, this band has a very weak intensity of only 0.1 km/mol, several orders of magnitude
weaker than in (HF)2. Indeed, all of the intermolecular bands in (HC1)2 are quite weak, with
integrated intensities of less than 50 km/mol.
As a point of interest, Karpfen et al.32 have also computed the vibrational frequencies
of the cyclic geometry, which represents not a minimum on the PES, but rather a transition
state, and hence the imaginary frequency listed in Table 3.11 for the bending motion that
would transform this cyclic geometry into one of the minima. It might be noted that the
asymmetric and symmetric internal HF stretching modes in the cyclic dimer are of very sim-

Table 3.8 Intensities of intermolecular modes (km m o l - 1 ) of (HF)2.


ip Bend oop Bend v ip Bend Reference

SCF/6-31G** 167 279 111 154 [21]


SCF/+VPs(2d)s 219 230 160 6 [23]
CCSD(T)/TZ2P(f,d) 160 188 25 141 [26]
Vibrational Spectra 147

Table 3.9 Vibrational frequencies (cm - 1 ) and intensities (km/mol) of HC1 in the monomer
and dimer32.

v, dimer A, dimer

Method v, monomer acceptor donor acceptor donor

ACPF/[652/42] 3005 -11 -32 39 165


ACPF/[6531/42] 2993 -6 -30 44 188

ilar frequency. It is also of interest that the various intermolecular modes are of lower fre-
quency than is the case for the minimum energy configurations in Table 3.10.
In the absence of spectral information in the gas phase, it is common to compare calcu-
lated features of the vibrational spectrum to data measured in rare gas matrix, the premise
being that the latter medium perturbs the H-bonded system as little as possible. The influ-
ence of the medium was considered33 via a self-consistent reaction-field formalism wherein
inductive interactions between the polar system and the polarizable medium are incorpo-
rated into a model Hamiltonian34. The calculations made use of the 6-31G** basis set at the
SCF level.
The immersion of (HF)2 into a simulated Ar or Xe matrix led to a more linear configu-
ration, lowering a in Fig. 3.1 from 13.6° to 8°. The proton acceptor molecule also rotated
toward a less perpendicular arrangement, with increasing from 103° to 110°. There were
only very small changes observed in the internal bond lengths, although a small stretch of
the interfluorine distance of the order of 0.01 A occurred. The net result is a 10% increase
in the dipole moment of the (HF)2 complex. Ar matrix has little effect upon the gas-phase
properties of (HC1)2, whereas immersion in Xe yields results much like those for (HF)2.
The changes calculated to occur in the stretching frequencies of the proton donor and
acceptor HX molecules are reported in Table 3.12, along with experimental values in paren-
theses. It seems clear that this theoretical approach is a disappointment as it strongly un-
derestimates the effects of placing the complex within the matrix. As the authors point out,
this effort represents merely a first step toward an improved treatment of matrix effects.
While their formalism includes purely inductive effects, it neglects repulsive and disper-
sion forces which are likely important as well. Another direction toward improvement might
be permitting some geometrical relaxation of the matrix and complex, which would allow
coupling of vibrational and rotational motions.

Table 3.10 Intermolecular vibrational frequencies (cm - 1 ) and


intensities (km/mol) of (HC1)232.
ACPF/[652/42] ACPF/[6531/42]

Mode V A v A

bend 290 15 297 13


torsion 195 25 204 19
bend 98 43 129 31
stretch v 63 0.1 66 0.1
148 Hydrogen Bonding

Table 3.11 Intermolecular vibrational frequencies (cm - 1 ) and intensities


(km/mol) of the cyclic geometry of (HC1)2, which represents a transition state
on the surface32.
ACPF/[652/42] ACPF/[6531/42]

Mode v A V A

Intramolecular
Asymmetric stretch 2989 76 2982 87
Symmetric stretch 2986 0 2979 0
Intermolecular
Bend 250 0 252 0
Torsion 157 49 160 46
Stretch v 60 0 60 0
Bend 75i Hi

3.4 H3Z-HX

The linearity of the H-bond in a complex like H 3 N ... HF simplifies the analysis of the vi-
brational frequencies in some ways. The vibrational data for this complex are reported in
Table 3.13 from data computed at the SCF level with a 6-31G** basis set21. One may first
note that the red shift of the proton donor HF molecule is more than 400 c m - 1 , about five
times greater than in (HF)2. This larger shift is concordant with the stronger H-bond inter-
action in H3N...HF. The intensification of this band is also more profound in the latter sys-
tem. The ratio of intensity in the complex versus that in the HF monomer is 7.1 in H3N...HF,
as compared to 2.8 in (HF)2, with the same theoretical framework.
With regard to the proton acceptor molecule, the vibrational frequencies of NH3 are
largely unaffected by the complexation. The exception is the symmetric bending motion
which is blue-shifted by nearly 100 c m - 1 . This change is in contrast to the HF proton donor,
whose stretching frequency is red-shifted. In (HF)2, the intensity of the vibration of the pro-
ton acceptor molecule was not changed much in comparison to the monomer. When NH3
acts as proton acceptor, the result is quite different. Both stretching modes undergo strong
intensification, by factors of 16 and 35. The bending modes keep their intensities largely
intact upon complexation.
One might expect the principal result of the replacement of NH3 by PH3 to be an over-
all weakening of the trends due to the latter's weaker basicity. This is indeed found to be
the case for a number of the spectroscopic properties. Comparison of the first rows of Ta-
bles 3.14 and 3.13 shows a marked drop in the red shift of v , as well as its intensification
which drops from 7 to 4. Curious, though, are the changes within the base molecule.
Whereas the stretching frequencies of NH3 are essentially unaffected by complexation, both
symmetric and asymmetric stretching modes of PH3 shift to higher frequency by some
26-30 c m - 1 . Instead of a strong blue shift, as in H3N...HF, the symmetric bend of PH3 shifts
slightly to the red, while its asymmetric bend moves to higher frequency by 94 cm - 1 . Also
obeying different trends are the intensities of these modes. In contrast to the very strong in-
tensification of the stretching vibrations in H 3 N ... HF, the internal stretches of PH3 are made
less intense by the complexation.
Figure 3.1 Definition of and angles in (HX)2
2

Table 3.12 Changes in stretching vibrational frequencies (cm ')


induced by placement of complex in Ar or Xe matrix. Calculated
with 6-31G** basis set33. Experimental values in parentheses.

Acceptor Donor

Ar Xe Ar Xe

(HF)2 16(35) 15 11 (42) 9


(HC1)2 1(24) 4(40) 2 (39) 2(40)

Table 3.13 Frequencies (in c m - 1 ) and intensities (km m o l - 1 ) of HF and NH3,


and the changes resulting from formation of the H 3 N ... HF complex. Data, cal-
culated at SCF level with 6-31G** basis set21.

v Amon
Mode Vmon Adim/Amon

HF
4495 2432 132 7.1
NH3
v str (a 1 ) 3706 0 0.2 35
vstr(e) 3844 -2 0.8 16
V bend (a 1 ) 1141 93 217 1.0
Vbcnd(e) 1812 -7 21 1.2

Table 3.14 Frequencies (in c m - 1 ) and intensities (km mol - 1 ) of HF and PH3,
and the changes resulting from formation of the H 3 P ... HF complex. Data,
calculated at SCF level with 6-31G** basis set21.

Mode V mon v A
A
mon Adim/Amon
A

HF
4495 -134 132 4.0
PH3
Vstr(a1) 2596 26 106 0.4
vstr(c) 2600 30 128 0.7
Vbend(a1) 1112 -13 36 1.4
vbend(e) 1250 94 21 1.0
150 Hydrogen Bonding

3.4.1 Analysis of Intensities


Because the intensity of a given vibrational mode is connected with the changing molecu-
lar dipole moment associated with that particular motion of the atoms, analysis of these in-
tensities offers valuable insights into charge redistributions within the system. One can par-
tition the dipole changes into contributions from various atoms using an "atomic polar
tensor" (APT) formalism35-36 which is defined for an atom a as

where u^ is the ith (x, y, or z) component of the dipole moment vector of the entire system,
and qj represents the jth coordinate of atom a. It is generally most useful to define a local
coordinate system for each atom such that the z-axis corresponds to a bond. Also useful are
characteristics of the tensor which are invariant to choice of coordinate system37. An ef-
fective charge x is related to all nine elements of the APT, defined as the square root of 1/3
the sum of squares of the elements. The diagonal elements are stressed in the mean dipole
derivative, p, equal to 1/3 the trace of the matrix. Anisotropy is defined in terms of these
two quantities, as is the deformability of the electron cloud.
This type of analysis provides useful clues as to the behavior of the intensities in Tables
3.13 and 3. 14 as follows. Of particular relevance is the Pzz element for each hydrogen atom
which describes how much a stretch of the H atom along its bond axis will affect the mol-
ecular dipole moment in that same direction. The sharp increase in intensity of the HF
stretch upon H-bonding can thus be associated with a bigger increase in that accompanies
motion of the proton in the dimer, as compared to the monomer.
More intriguing than the HF stretch is the qualitatively different behavior of the NH3
and PH3 internal vibrational modes. Indeed, the intensification of the NH3 stretching modes
are greater in a relative sense than the HF stretch when the H 3 N ... HF complex is formed.
Again, the behavior observed for the intensities is mirrored by the Pzz element of the APT.
Further insight is provided by separating Pzz into two components. One can evaluate the
change in molecular dipole moment that would result from the simple motion of the hy-
drogen nucleus, along with its static partial charge. Subtraction of this quantity from PZZ is
a measure of how much the dipole moment is affected by redistribution of electron density,
that is, how much the partial charge of the hydrogen is changing as it is moved.
This type of analysis revealed that the hydrogens of NH3 in the monomer suffer a loss
of positive charge as they move away from the N center, that is, they pick up additional elec-
tron density. Hence, the stretch away from the N engenders two competing effects. Motion
of a H atom with a partial positive charge acts to increase the dipole moment, but this is
counteracted by the shift of excess electron density in the same direction. This cancellation
is responsible for the small intensities of the N — H stretches in the monomer noted in
Table 3. 13.
The situation changes when the ammonia is bound to HF. Formation of the H-bond trans-
fers some electron density from NH3 to HF, causing an increase in the positive charge of
the three hydrogens of NH3. The larger positive charge causes a greater change in the mol-
ecular dipole moment when the proton moves. Another effect of the presence of the HF is
to "lock up" the electron density of NH3, that is, it makes it more difficult for the density
to follow the ammonia hydrogens as they move away from N. This less flexible density
lowers the counteracting influence noted in the free ammonia, adding to the value of z/ z.
What accounts for the very different behavior in PH3? First, the P— H stretches are al-
ready rather intense, even in the isolated monomer. This difference can be traced to the
Vibrational Spectra 151

charges of the H atoms, which are slightly negative compared to their positive charge in
NH3. The different charges of hydrogen in the two molecules can be associated with the
lesser electronegativity of the P atom versus N. The negatively charged hydrogens lead to
like changes for the Pzz APT elements since motion of the hydrogen away from the P atom
will diminish the molecular dipole. As in NH3, the charge flux again acts to augment the
electron density around the shifting hydrogen. Instead of making the hydrogen less posi-
tive, as in NH3, this atom now becomes more negatively charged as it moves away from P.
So rather than counterbalancing one another, which made the intensities of the NH3
stretches so weak, the charge flux and partial charge of the H atoms in PH3 reinforce each
other, leading to the strong P—H stretches in the monomer.
Binding to the HF molecule again draws electron density away from the proton accep-
tor molecule. But as opposed to NH3 where the hydrogens were made more positive, this
removal decreases the negative charges of the hydrogens of PH3. The stretching motion of
the latter hydrogens hence yields a smaller decrement of the molecular dipole than in the
isolated monomer, and the result is a weaker intensity.
Extending the same partitioning to the HF stretches helps us understand the intensifica-
tion of vs. The charge on the H atom is positive in the monomer, and becomes slightly more
positive in the complex. But more important is the charge flux term. Upon formation of the
H-bond, the electron density can less effectively follow the proton as it moves away from
the F atom. The stretch therefore increases the hydrogen's positive charge and causes a large
increment in the molecular dipole. This finding is consistent with other calculations wherein
motion of the bridging hydrogen induces a flow of electron density in the opposite direc-
lion38,39.
The data calculated for the intermolecular modes of H 3 Z . . . F are reported in Table 3.15.
The 876 cm-1 frequency of the bending of the proton donor molecule in H 3 N ... HF is clearly
higher than any of those calculated for (HF)2, another indication of the stronger H-bond.
This motion corresponds primarily to imparting nonlinearity into the H-bond by pulling the
proton off of the axis. The wagging motion of the acceptor requires less energy, partially
explaining its lower frequency. The H-bond stretching frequency, v , is similar in magni-
tude to the acceptor bend. v is a relatively pure mode in the H3Z...HF complexes, unlike
other cases where the intermolecular stretch is combined with substantial contributions
from other stretches and bends.
The patterns of intensities can be understood on the basis of how much a given vibra-
tional motion changes the molecular dipole moment. The donor bend, for example, acquires
its high intensity because it turns the HF molecule with its large molecular moment in such
a way as to give the C3v complex a nonzero moment perpendicular to the H-bond axis. In
contrast, the stretch of the two subunits away from each other in the v mode does little to

Table 3.15 Vibrational spectra calculated for intermolecular modes21 at SCF/6-31G** level.
Frequency ( c m - 1 ) Intensity (km mol - 1 )

Mode H 3 N ... HF H 3 P ... HF H 3 N ... HF H3P...HF

H-bond stretch (v ) 240 113 3 I


donor bend 876 467 208 134
acceptor bend 237 112 10 6
I 52 Hydrogen Bonding

change either the magnitude or direction of the complex's moment. Analysis carried out by
Kurnig et al.21 indicated that of the total intensity of this mode in H3N...HF, 15% arises from
small mixing in of an internal HF stretch and rocking and internal bending of NH3. The re-
maining 85% arises from the interactions between the two molecules, for example, the
changing amount of charge transfer as the two subunits are pulled apart.
Indeed, low intensities are expected in any H-bonded complex for v which is consti-
tuted primarily of the intermolecular stretch. What lends intensity to this band in a complex
like (HF)2 is the contribution of bending motions that change the dipole moment vector di-
rection, and internal HF stretches that alter the subunit's moment. For example, Swanton et
al. had determined in their analysis of the water dimer40 that 70% of the v intensity can be
attributed to internal motions within the two water subunits.

3.4.2 Anharmonicity
Bouteiller et al. have devised a means of treating the anharmonicities of H-bonded com-
plexes41 by expanding the potential energy function as a fourth-order polynomial, fit to the
ab initio calculations. These investigators use two degrees of freedom, corresponding to the
X—H distance, r, and the intermolecular separation between X and Z, R.

The variational method is next used to solve the Schrodinger equation of vibrational mo-
tion, taking eigenfunctions of the harmonic oscillator as expansion functions. A group of
selected expansion coefficients are listed in Table 3.16 to illustrate the degree of coupling
and anharmonicity. With regard first to the second-order coefficients in the first three rows,
the much larger values of a20 reflect the greater sensitivity of the energy to small changes
in r(H—X) than to the intermolecular distance. But it is significant that a11 is comparable
in magnitude to a02, reflecting the coupling between r and R. The large value of a30 is an-
other source of anharmonicity as is the fourth-order term represented by a40.
The frequencies that arise from this treatment are listed for the three complexes in Table
3.17. Note first that the anharmonic frequencies of vs (XH in Table 3.17) are quite a bit
smaller than the harmonic values in the upper part of the table, differing by about 400 cm - 1 .

Table 3.16 Expansion coefficients calculated for H-bonded complexes at SCF


level (CI values in parentheses)41. All values in units of mdyn/A p+q+1 .
apq C1H..NH
3 C1H ..NH2CH3 BrH .. NR 3

a20 2.382(1.938) 2.239 1.816


a11 0.197(0.213) 0.190 0.144
a02
0.106(0.132) 0.109 0.078
a30 -5.572 (-4.503) -5.309 -3.765
a21 0.379(0.510) 0.805 0.566
a12 -0.459 (0.004) -0.193 -0.165
a03 - 0.08 (-0.240) -0.160 -0.106
a40 4.879 (3.835) 4.917 3.098
Vibrational Spectra 153

Table 3.17 Vibrational transitions calculated41 at SCF level, except those in


parentheses derived at CI level. All values in units of c m - 1 .

CIH NH3 CIH NH2CH3 BrH NH3

Harmonic approximation
XH 2888 (2606) 2800 2516
X..N 172(192) 146 135
Anharmonic data
XH 2501 (2249) 2365 2084
X..N 162 (201) 144 144

Nonetheless, there is a strong parallel between these vs frequencies and the a20 coefficients
in Table 3.16, with either the harmonic or anharmonic treatments. There is little such paral-
lel in the v stretches (X .. N) and the a02 coefficients, providing a warning against using sim-
ple force constants to estimate the frequency of this intermolecular mode. It is interesting,
however, that the harmonic and anharmonic values of the X..N stretch are quite similar.
With regard to various levels of theory, it is significant that inclusion of correlation (val-
ues in parentheses for C1H..NH3) lowers the vs frequency but has the opposite effect of in-
creasing v . It is also worth noting that the SCF and CI frequencies can differ by as much
as 300 c m - 1 , harmonic or anharmonic. Indeed, anharmonicity and electron correlation ef-
fects are additive here: the anharmonic CI vs frequency is smaller than the SCF harmonic
value by a full 639 c m - 1 .
A later work by Bouteiller et al.42 considered the set of complexes of NH3 with FH, FD,
and FLi. Comparison of the results of the first two illustrates isotope effects while substi-
tution with lithium introduces a different sort of bonding. The Vibrational transition ener-
gies calculated by the above anharmonic approach, extracted from energy surfaces at the
SCF and MP2 levels, are listed in Table 3.18. The correlated F—H and F—D stretching fre-
quencies are considerably smaller than the SCF values, whereas a small increase results
when correlation is included for FLi. Similar discrepant behavior in terms of correlation ef-
fects has been noted in geometries and energetics. There does not seem to be much of an
isotope effect on the F..N stretch, as results for H and D are nearly identical. Correlation ex-
erts a still significant effect here, raising the frequencies by perhaps 10%, but again the ef-
fect is opposite for FLi. The values in parentheses refer to the frequencies computed when
all coefficients apq are ignored when p + q > 2, again at the correlated level. One can see
that neglect of anharmonicity yields F—L frequencies that are too high, for all F—L in-
cluding FLi. The effects are much smaller on the F..N stretches.

Table 3.18 Vibrational transitions calculated42 in units of c m - 1 . Harmonic


values in parentheses.
.. .. ..
FH NH3 FD NH3 FLi NH3

SCF MP2 SCF MP2 SCF MP2

F-L 3773 3331(3514) 2752 2427(2527) 932 947(1010)


F..N 240 263(265) 236 255(262) 292 281(272)
154 Hydrogen Bonding

The Bouteiller approach to anharmonicity also permits extraction of energies of excita-


tion to vibrational levels beyond the first excitation. The various progressions are reported
in Table 3.19, arising from the correlated potentials. Proceeding down each column, the
spacing between successive overtones decreases as the quantum number rises, resulting
from the mechanical anharmonicity.

3.4.3 Other Properties


As mentioned earlier, there are additional properties of H-bonded systems accessible to cal-
culations of this sort. For example, Bacskay et al.7 have computed the vibrationally aver-
aged component of the electric field gradient (EFG) tensor along the symmetry axis of
C1H...NH3 and C1H...PH3. Table 3.20 reports the change in this component, at the Cl nu-
cleus, caused by formation of the H-bond, as Vzz(Cl), as well as the change in the mole-
cular dipole moment, also along the symmetry axis. Also listed in the last column is the root
mean square angle between HC1 and the symmetry axis, after vibrational averaging. The
calculations used a basis set of moderate size, with polarization functions for all atoms.
The first column of Table 3.20 indicates threefold more change in the EFG at the Cl nu-
cleus for the stronger complex with NH3 than with PH3. As a percentage of the EFG in the
uncomplexed HC1 monomer, the next column indicates a 17% change for C1H...NH3, in
reasonable agreement with an experimental measurement of 23%. The 5.5% change calcu-
lated for C1H...PH3 is also lower than the experimental estimate of 7.4%. The dipole mo-
ment of the complex changes much more for the stronger H-bond, too. The last column il-
lustrates the "floppier" character of the complex for a weaker H-bond. The root mean square
angle made by the HC1 molecule with the symmetry axis, resulting from vibrational aver-
aging is some 13° for C1H...NH3 but 19° for C1H...PH3. The authors noted that the basis set
superposition error was negligible for their EFG and other electronic properties.
Bacskay et al.7 also concerned themselves with the possible effects of electron correla-
tion on the aforementioned electronic properties. They consequently compared their SCF
data with results obtained using an approximate coupled pair functional (ACPF) approach
to correlation. As the authors had noted that geometries optimized at the SCF level typically

Table 3.19 Overtones and combination bands calculated with anharmonicity at


the correlated level. Data in cm-1. 42

vn v'n' FH NH3 FD..NH3 FLi..NH3

00 01 263 255 281


00 02 509 494 557
00 03 739 717 829
00 04 958 931 1096
00 05 1172 1141
00 10 3331 2427 947
00 11 3644 2735 1272
00 12 3921 2995 1551
00 13 4180 3239 1822
00 14 4425 3470
00 15 4658 3691

Note. v and n refer, respectively, to the F-L and F-N stretching modes.
Vibrational Spectra 155

Table 3.20 Changes of electric field gradient and dipole moment caused by
H-bonding (atomic units) and rms HC1 librational angle (degs) calculated with
[642/531/31] basis set for [C1,P/N/H]7.
2 1/2
- Vzz(Cl) - Vz//Vzz° - z < >
..
C1H NH3 0.618 0,17 0.412 12.7
C1H..PH3 0.196 0.06 0.250 18.8

had the two subunits too far apart, they evaluated their properties at the experimental
geometries in Table 3.21. Comparison of Tables 3.20 and 3.21 indicates a significant in-
crease in the magnitude of Vzz results from bringing the two molecules closer together, as
well as forcing each subunit into its experimental monomer geometry. For example, this
property has increased from 0.62 to 0.92 atomic units when the experimental geometry of
C1H...NH3 is adopted. Indeed, the sensitivity of this quantity to the intermolecular separa-
tion is embodied by the Vzz/ R term in Table 3.21. The 0.83 value for C1H...NH3 indicates
that a stretch of only 0.1 A would alter the EFG at the Cl nucleus by as much as 0.15 atomic
units. Comparison of the first two columns of data in Table 3.21 reveals that correlation re-
duces to some extent the sensitivity of the EFG to the H-bonding interaction by 7-17%. The
values in parentheses are not much different than their preceding values, indicating very lit-
tle basis set superposition influence upon this property. The dipole moment change result-
ing from the H-bond formation has surprisingly little sensitivity to inclusion of electron cor-
relation.

3.4.4 Relationship between H-Bond Strength and Spectra


An example of the Badger-Bauer relationship between the strength of the H-bond and the
red shift of the X—H stretching frequency is provided by recent correlated computations
of complexes pairing HC1 with a series of 4-substituted pyridines43. As illustrated by the
solid line in Fig. 3.2, the change in this frequency is very nearly linearly related to the cal-
culated strength of the H-bond. The dashed line refers to the intensity of this mode in the
dimer, as compared to the HC1 monomer. This property, too, is strongly correlated with the
strength of the H-bond. Note the magnitude of this enhancement, making the intensity in
the complex some two orders of magnitude higher than in the monomer.

Table 3.21 Changes of electric field gradient and dipole moment caused by H-bonding
(atomic units) calculated at experimental geometries with a [642/531/31] basis set
for [C1,P/N/H]7.
- Vzz (Cl) - z
Vzz/ R
SCF ACPF(+CC) a SCF SCF ACPF
..
C1H NH3 0.924 0.862(0.848) 0.83 0.486 0.499
C1H- PH 0.340 0.285 (0.282) 0.30 0.304 0.279
a
With full counterpoise correction.
156 Hydrogen Bonding

Figure 3.2 Illustration of the Badger-Bauer relationships for HCl...base


where the base is a sesries of 4-substituted pyridines. Data43 calculated
at MP2/6-31 +G(d,p) level. The dashed line represents the magnification
of the intensity in the dimer, relative to the moment. E refers to the
electronic contribution to the binding energy.

3.5 H2Y...HX

The complex combining water with HF furnishes a good example where the proton accep-
tor has two lone pairs. Somasundram et al.6 and Amos et al.44 have examined this complex
with two different basis sets, one singly and the other doubly polarized. The results for both
basis sets are reported in Table 3.22 and illustrate the expected strong red shift of the pro-
ton donor stretching frequency, albeit by a lesser amount than in the more strongly bound
H3N...HF. The shift calculated at the SCF level of around 260 cm-1 underestimates the ex-
perimentally observed value of 353 c m - 1 , but the correlated shifts are much closer to ex-
periment. The frequency changes in the proton acceptor are fairly consistent from one ba-
sis to the next: Both stretches (V1 and V3) are red shifted by some 7-10 c m - 1 whereas the
bending motion is only slightly affected. This behavior contrasts with NH3 in its complex
with HF where the stretching frequencies are altered less than the bends. The trends are
more or less intact after correlation, but the red shift of the highest frequency mode is larger.
Somasundram et al.6 computed the intensities not only for IR but also for Raman bands
and the results are listed in Table 3.23. The intensification of the IR proton donor stretch-
ing band is by a factor of nearly 5 for H 2 O ... HF, a little smaller than the magnification of 7
in H3N...HF. Whereas the two stretching motions in H3N were strongly intensified by com-
plexation with HF, these increases are much smaller in H2O. In both cases, the intensities
Vibrational Spectra 157

Table 3.22 Frequencies (in cm-1 )of HF and OH2, and the changes resulting from formation
of the H2O ...HF complex6,44.

SCF MP2
DZP TZ2P DZP TZ2P

Mode mon
v Vmon Av Vmon V Vmon vmon V

HF
4511 -259 4471 -264 4221 -340 4154 -363
H2O
V1 4166 -7 4128 -9 3913 2 3858 -10
V2
1752 5 1760 0 1671 -3 1657 -3
V3 4289 -10 4228 -10 4059 -18 3980 -20

of the bending motions are insensitive to the H-bond. The Raman bands undergo some mi-
nor modifications upon complexation but none larger than 1.6.
The intermolecular modes are listed in Table 3.24 along with their IR and Raman inten-
sities, all calculated with the doubly-polarized TZ2P basis set. The SCF H-bond stretching
frequency of 220 cm-1 is surprisingly similar to the experimental measurement of 198
cm - 1 , considering the harmonic and other approximations involved in its calculation. This
quantity is comparable to that in H3N HF, but the IR band is considerably more intense in
HLjO—HF. This greater intensity is linked to the coupling into the mode of bending motions
which permit a greater change of the dipole moment. Indeed, the weakest band in this part
of the IR spectrum, corresponding to a bending motion, probably owes its low intensity to
a certain degree of H-bond stretching motion. Of particularly low intensity are the Raman
bands, all less than 2 A4/amu. This finding is not surprising as there are only small changes
within the covalent bonds of the monomers that accompany these intermolecular motions.
The authors expressed their belief that their IR and Raman intensities of the intermolecular
vibrations are correct within an order of magnitude. It should be noted that correlation acts
to increase the frequencies of all the intermolecular modes by amounts varying from 20 to
100cm -1 .

Table 3.23 IR and Raman intensities of HF and OH2, and the changes
resulting from formation of the H2O...HF complex. Data calculated with TZ2P
basis set6.
IR (km m o l - 1 ) Raman (A4/amu)

Mode Amon_ A A dim /A mon Smon_ Sdim/Smon

HF
147 4.6 26 1.6
H2O
v1 14 5.5 70 1.0
v2 92 1.0 4 0.7
v3 70 1.6 29 1.0
158 Hydrogen Bonding

Table 3.24 Vibrational spectra calculated for intermolecular modes of H2O...HF6,44 with
TZ2P basis set.

Mode v S C F (cm - 1 ) vMPZ(cm-1) A IR (km mol - 1 ) SRamen (A4/amu)

H-bond stretch (v ) 220 270 87 0.4


bend 182 232 155 0.8
bend 234 252 3 2
shear 644 742 226 1
shear 786 862 194 0.2

Latajka and Scheiner45 carried out a vibrational analysis of H2O...HC1 using several dif-
ferent basis sets. The results with their best basis, at the SCF level, are presented in Table
3.25 where the red shift of the vs band of HC1 equals 105 c m - 1 . Its intensity is magnified
by a factor of 6.4 upon forming the complex. The frequencies of the proton-accepting wa-
ter are little affected and intensity changes are only moderate, none increasing by more than
a factor of two. These changes are all smaller here than in the more tightly bound H2O...HF.
The vibrational data for the intermolecular modes are reported in Table 3.26. The H-bond
stretching frequency, v , is only 118, comparable to the same quantity in the H 3 P ... HF com-
plex. The other frequencies are also in the same range as H3P...HF. The v is of notably low
intensity, as in the H 3 Z ... HF complexes, suggesting little mixing with the bending modes
that would add intensity via changing the dipole moment. The other modes are of higher in-
tensity, in the 30-80 km mol - 1 range.
Hannachi et al.46 have carried out their calculations of the spectrum of the full series
XH ... OH 2 , X=F,Cl,Br,I. The energies were computed with a core pseudopotential ap-
proach, specifically the PS-31G** basis set for the halogens, and standard 6-31G** for wa-
ter. The first clear trend in Table 3.27 is a progressive decrease in both the vs and v fre-
quencies as the halogen atom changes from F to Cl to Br to I. These trends are true for either
the harmonic or anharmonic frequencies. While the harmonic and anharmonic values of vs
are clearly different, the magnitude of this difference diminishes as one proceeds from F to
I. The red shift of this band is enhanced by anharmonicity in all cases. Again, inclusion of
anharmonicity effects do little to change the X..O stretching frequencies.

Table 3.25 Frequencies (in cm - 1 ) and intensities (km mol - 1 ) of HC1 and OH2,
and the changes resulting from formation of the H2O...HC1 complex. Data
calculated with + VPs(2d)s basis set45.

Mode V mon V A mon Adim/Amon

HC1
3141 -105 56 6.4
H2O
v1 4139 -4 19 1.9
v2 1759 1 103 0.9
v3 4244 -3 91 1.3
Vibrational Spectra 159

Table 3.26 Vibrational spectra calculated for intermolecular


modes of H2O...HC145.

Mode Frequency ( c m - 1 ) Intensity (km mol - 1 )

H-bond stretch (V ) 118 3


donor bend 459 77
donor bend 351 38
acceptor bend 143 33
acceptor bend 94 28

3.5.1 Alkyl Substituents


The effects of methyl substitution upon the Vibrational spectra may be determined from
comparison of the aforementioned results for H2O...HC1 in Table 3.26 with the data com-
puted by Amos et al.44 for (CH3)2O...HC1 in Table 3.28. Bearing in mind the results were
obtained with slightly different basis sets, it is nevertheless apparent that the H-bond stretch-
ing frequency is changed very little by the substitution, nor is the intensity of this band al-
tered by much. There seems to be an increase in the frequencies for bending the proton
donor, whereas the frequencies for bending the acceptor molecule are very small. This drop
is due in some measure to the large increase in the effective mass for this motion when the
two hydrogens of H2O are replaced by methyl groups.
The red shift of the HC1 stretch, listed in the last row of Table 3.28, is considerably larger
than in H2O...HC1. Nonetheless, this shift of 170 c m - 1 is only about half of the experi-
mental quantity of 316 c m - 1 47. The same trends are observed in solid matrix. When the
proton acceptor in the H2O...HC1 complex is changed to dimethyl (or diethyl) ether, the red
shift of the HC1 stretch increases by several hundred cm-1 48. Similarly increased red shifts
when the base is alkylated are noted for HF and HBr as proton donors. The sulfur analogs,
namely, H2S, Me2S, and Et2S, obey similar patterns when paired with HF, HC1, and HBr48.

3.5.2 Other Properties


As described above for complexes of HC1 with NH3 and PH3, Bacskay et al.7 have com-
puted the vibrationally averaged component of the electric field gradient (EFG) tensor along

Table 3.27 Calculated vibrational transitions46. All values in c m - 1 .


FH..OH2 C1H..OH2 BrH..OH2 IH..OH2

Harmonic approximation
XH 4243 3035 2656 2390
- vs 321 150 129 59
..
xo 235 155 120 98
Anharmonic data
XH 4019 2847 2501 2307
- vs 370 257 209 76
X..O 226 146 121 97
160 Hydrogen Bonding

Table 3.28 Vibrational spectra calculated for intermolecular


modes of (CH3)2O...HC144 with DZP basis set at SCF level.
mode frequency, cm-1 intensity, km mol-1

H-bond stretch (v ) 107 3


donor bend 507 37
donor bend 399 51
acceptor bend 10 4
acceptor bend 35 0
HO shift ( vs) -170

the inertial axis of C1H...OH2 and C1H...SH2. Table 3.29 reports the change induced in this
tensor by formation of the H-bond, Vzz(Cl), and the accompanying change in the molec-
ular dipole moment. < 2>1/2 refers to the vibrationally averaged, root mean square angle
between HC1 and the inertial axis.
As noted earlier, the stronger H-bond produces more of a change in the EFG and the di-
pole moment at the Cl nucleus. The trends in Table 3.29 are consistent with those noted for
the complexes of HC1 with XH3. The 12% change in the EFG for C1H...OH2 matches very
closely the experimental estimate and the 8.2% change in C1H...SH2 matches the 8.6% ex-
perimental result. The averaged librational angles for the two complexes are quite similar
to values calculated for the C1H...NH3 and C1H...PH3 analogues.
The effects of electron correlation on these electronic properties may be noted from
Table 3.30. As in the case of the C1H...ZH3 complexes, comparison of Tables 3.29 and 3.30
indicates an increased magnitude of Vzz results from bringing the two molecules closer
together, coupled with forcing each subunit into its experimental monomer geometry.
Vzz/ R in Table 3.30 shows a sensitivity of the EFG to intermolecular separation, although
not quite so much as for C1H...NH3. Comparison of the first two columns of data in Table
3.30 reveals that correlation reduces the sensitivity of the EFG to the H-bonding interaction
by about 8%. As for the other complexes studied by the authors, there is little basis set su-
perposition affecting this property. The dipole moment change resulting from the H-bond
formation is basically independent of electron correlation.

3.6 H2Y...HYH

Just as for other properties of H-bonded systems, the water dimer has been the subject of
perhaps the greatest scrutiny to its vibrational spectrum. Curtiss and Pople's seminal work49

Table 3.29 Changes of electric field gradient and dipole moment caused by
H-bonding (atomic units) and rms HC1 librational angle (degs) calculated with
[642/531/31] basis set for [C1,S/O/H]7.
- Vzz(C1) -Vzz/Vzz° -z z < 2 > 1/2
ClH..OH2 0.431 0.12 0.281 13.8
ClH .. SH 2 0.290 0.08 0.235 17.8
Vibrational Spectra 161

Table 3.30 Changes of electric field gradient and dipole moment caused by H-bonding
(atomic units) calculated at experimental geometries with a [642/531/31] basis set for
[C1,S/O/H]7.

- V zz (Cl) - z
vzz / R
a
SCF ACPF (+CC) SCF SCF ACPF

ClH..OH2 0.546 0.499 (0.488) 0.50 0.319 0.324


ClH..SH2 0.394 0.361 (0.359) 0.30 0.293 0.295
a
With full counterpoise correction.

consisted of a FG matrix analysis to obtain the normal modes, using SCF force constants.
It was learned that simple description of the intermolecular modes is complicated by a high
degree of mixing between the various internal coordinates. Nonetheless, the authors were
able to identify a mode which is composed largely of the hydrogen-bond stretching motion.
Less obvious, but still recognizable, were librational motions associated with nonlinearity
in the H-bond. One is primarily an in-plane wagging of the proton donor and the other an
out-of-plane bend. A rotation of the proton acceptor molecule about its internal symmetry
axis, that is, out of the H-bond plane, is of very low frequency, only 80 cm-1 or so.
Some very interesting calculations50 have addressed the question of how the geometry
of the H-bond directly affects the vibrational features of the complex, using the water dimer
in Fig. 3.3 as a model H-bonded system. The stretching force constant, k, of the bond be-
tween Od and Hd was evaluated as a function of the intermolecular geometrical parameters
R, , and . k is smallest at the equilibrium geometry, reflecting the weakening effect of the
H-bond. k rises much more slowly with increasing (3, as the proton acceptor molecule
swings away from its optimal angular orientation, than when the donor is rotated via an in-
crease in a. This stretching force constant rises toward its monomer value as the H-bond is
stretched. Indeed, the authors remark upon the similarities between the behavior of this par-
ticular stretching force constant and the interaction energy, E, itself.
The authors go on to conclude that the red shift of the vs band in this H-bonded complex
can be directly attributed to the lengthening of the O d - H d bond. By partitioning the inter-
action energy into various components, they show how the stretch of this bond makes it
both more polar and polarizable, which in turn, increases the induction and charge transfer
components of the interaction energy. Although the authors did not include correlation in
their treatment, the same could be said for dispersion energy which is directly related to po-
larizabilities of the individual monomers. It is for this reason that a nearly linear relation-
ship is observed between vs and r. Zilles and Person36 have reached a similar conclusion
that the polarity and polarizability of the O—H bond increases upon formation of the H-

Figure 3.3 Geometry of water dimer, defining


three geometrical parameters.
162 Hydrogen Bonding

bond, based upon their atomic polar tensor analysis of the wave function. Indeed, the latter
authors attribute the bulk of vibrational intensity changes seen in all normal modes upon
dimerization to the electron density shifts in this bond.

3.6.1 Polarizability
Swanton et al.51 investigated the effect of H-bond formation upon the electronic structure
of the water molecule, in particular its polarizability. These properties are related to exper-
imentally accessible quantities via Raman bands. Using the harmonic approximation, the
differential cross section perpendicular to the incident light can be described as

where g is the degeneracy of the mode and C a physical constant. The quantity in brackets
is referred to as the scattering activity. mean' is the derivative of the average polarizability

and ( ')2 the square of the polarizability-derivative anisotropy, where

and derivatives, indicated by prime, are taken with respect to the normal coordinate. One
can also define a degree of polarization, p, as

when the incident light is directed along the x-axis, polarized in the z-direction, and scat-
tered in the y-direction.
Coupled perturbed Hartree-Fock calculations at the SCF level were used to assess the
polarizability tensor elements, each of which is defined as

where represents the applied field.


The authors used a [5s4pld/4slp] basis set in their calculations. In order to focus on the
effects of the molecular interaction, they introduced the concept of a "noninteracting dimer"
wherein the dimer wave function is a simple product (non-antisymmetrized) of the unper-
turbed monomer functions. The effects of the interaction are thus in evidence by compari-
son of the two columns in Table 3.31 from which it may be seen that the average polariz-
ability is little affected, increasing from 16.48 to only 16.60. The anisotropy of the
polarizability, however, as measured by 2, undergoes a dramatic increase. Whereas the po-
larizability tensor is nearly spherical in the monomer, with all ii values between 16.3 and
16.6, xx is increased up to 18 when the two molecules interact with one another. This in-
crease is thus focused along the H-bond direction.
The changes in the polarizability quantities that are associated with each of the normal
intramolecular vibrational modes are presented in Table 3.32. Any changes that occur on
going from the monomer to the noninteracting dimer are due to the redefinition of the nor-
mal coordinate motion within the context of the dimer, rather than any changes in electronic
structure. For example, the symmetric stretching motion in the monomer couples together
the two O—H bonds. But this coupling is weakened within the dimer where the first donor
stretch correlates with the O—H bond of the bridging proton. This changing motion pro-
Vibrational Spectra 163

Table 3.31 Polarizability aspects of the water dimer and a dimer


with identical geometry in which the two molecules are prevented
from interacting with one another. Proton-donating water
molecule lies in xy plane; x-axis is approximately parallel to
O..O line. Data51 in units of ao2e2/Eh.
Noninteracting Interacting

mean 16.48 16.60


2 1.48 8.20
xx 16.63 18.09
yy
16.26 15.67
zz
16.54 16.04
-0.68 -0.02
xy

duces some strong effects. For example, the mean polarizability derivatives in the sym-
metric stretches of the two molecules split from 4.75 in the monomer to 5.84 and 2.60 in
the donor and acceptor molecules, respectively. The actual interaction causes a small in-
crease in the former and a decrease in the latter, resulting in a further splitting. The anti-
symmetric stretching motions in the monomer do not have any effect upon the mean po-
larizability because of the strict coupling between the two O—H bonds. But again, placed
within the context of the dimer, the two O—H stretches are uncoupled. The asymmetric
stretching mode of the donor correlates with the stretch of the donor O—H bond (the H not
involved in the H-bond) and causes a marked change in polarizability, as indicated by the
2.08 entry in Table 3.32. The polarizability is fairly insensitive to bending motions, either
within the monomer or the dimer.
The various changes described above translate into the analogous intensification and
weakening of the scattering activities in the rightmost section of Table 3.32. Significant
changes occur in the donor stretching modes and the acceptor symmetric stretch. The bulk
of these changes can be attributed to the changes in the normal modes that accompany
dimerization, with smaller effects resulting from the actual interaction between the two
monomers. The greatest intensification, by a factor of about two, is noted in the O—H
stretch of the donor. This increase is dwarfed by the much larger changes noted in the in-
frared spectrum when H-bonding occurs.
The authors also studied the polarization patterns associated with the intermolecular vi-
brational modes. Average polarizability derivatives were calculated to be quite small, yield-
ing small scattering activities, all below 10 (xlO -34 C 4 N -2 kg -1 ).

3.6.2 Comparison between (H2O)2 and (H2S)2


Vibrational frequencies and intensities were compared between the monomer and water
dimer by Amos in 198652 using a polarized basis set of the 6-31G** type. Also calculated
and reported for purposes of comparison is the analogous dimer of H2S. The Vibrational fre-
quencies and intensities of the monomers are listed in Tables 3.33 and 3.34, respectively,
along with the changes that occur upon dimerization53. One might make a preliminary note
that the frequencies are overestimates compared to experiment (shown in parentheses), as
are the intensities.
Table 3.32 Calculated data relevant to polarizabilities in water dimer for intramolecular vibrational modes, and Raman
scattering activities. 51a

mean' ( ')2 P S

mono nonint ilnt mono nonint int mono nonint int mono nonint int

a' donor stretch (sym) 4.75 5.84 6.28 26.9 18.0 49.8 0.07 0.03 0.08 704 970 1240
a' acceptor stretch 4.75 2.60 2.33 26.9 38.8 34.2 0.07 0.25 0.27 704 337 282
a' donor stretch (antisym) 0.0 -2.08 2.04 51.6 48.5 54.9 0.75 0.37 0.40 211 312 334
a" acceptor stretch 0.0 0.0 0.0 51.6 51.6 50.6 0.75 0.75 0.75 211 211 207
a' donor bend -0.24 -0.18 -0.09 1.78 2.05 1.88 0.55 0.64 0.71 8.8 9.2 7.9
a' acceptor bend -0.24 -0.27 -0.22 1.78 1.50 2.76 0.55 0.48 0.63 8.8 8.1 12.5
a
Units: mean' in aoe2Eh- 1 -1/2
; ( ') 2 in a o 2 e 4 E h - 2 -1
; p and S in l 0 - 3 4 C 4 N - 2 kg-
Vibrational Spectra 165

Table 3.33 Calculated frequencies and changes induced by H-bonding in intramolecular


modes, calculated with 6-31G** basis set, in c m - 1 . 52 Experimental values in parentheses.

(H2O)2 (H2S)2
v
V mon vd va vmon vd va

sym stretch 4149(3657) -45 -5 2874 (2614)a —5 1


bend 1772(1597) 27 -1 1321 (1183) 4 -2
asym stretch 4259 (3756) -25 -8 2887 (2619) -3 1
a
Taken from reference 53.

Focusing first on the water dimer, both of the stretching frequencies are lowered when
the complex is formed, but these changes are much more pronounced in the donor mole-
cule. The intensities are also enhanced, particularly for the first mode listed, corresponding
to v , which is amplified by an order of magnitude in the donor. The bending frequency un-
dergoes a significant blue shift in the donor and a slight weakening of its intensity, but the
acceptor bending mode is hardly affected. Very similar patterns emerge in the H2S dimer,
although the effects are smaller in magnitude. The red shift of vs is only 5 cm-1 and it is
intensified threefold. The changes in the acceptor frequencies are insignificant. Qualitative
differences between the two dimers are revealed in the intensities. Rather than the en-
hancement observed for both stretches in (H2O)2, the asymmetric stretches of donor and
acceptor of (H2S)2 are both reduced in intensity, as is the symmetric stretch of the acceptor.
The data in Table 3.35 refer to the intermolecular vibrational modes calculated for the
two systems by Amos52. All the frequencies are uniformly higher for the water dimer, at-
tributable to the stronger H-bond as compared to (H2S)2. For both complexes, the highest
frequency corresponds to the out-of-plane bending motion. The a' stretch, most closely cor-
responding to v , is considerably lower, particularly for (H2S)2. But other than this fre-
quency, the others seem to fall in approximately the same order for the two complexes. The
most intense intermolecular band would appear to be the a' bend. Second most intense for
(H2O)2 is the a" shear or the a" bend for (H2S)2. It is not clear exactly how distinct these
two modes are since they are of the same symmetry and can consequently mix extensively.
Another fairly intense mode is the a' shear. The marked difference between the two con-
geners with respect to the a' stretch is particularly interesting. It is unclear why this differ-
ence should arise, but may be due to some particular mixing of the a' modes.

Table: 3.34 Calculated intensities, in km/mol, and changes, Aa, induced by H-bonding
in intramolecular modes52. Experimental values in parentheses.
(H20)2 (H2S)2

Amon Ad Aa Amon
mon Ad Aa

sym stretch 17(2.2) 10.41 1.71 1.8 3.17 0.39


bend 97 (54) 0.91 1.06 6.7 0.94 1.09
asym stretch 58 (45) 1.79 1.52 5.0 0.54 0.56
a
A is calculated as the ratio between the dimer vs. the monomer: A d i m /A m o n .
166 Hydrogen Bonding

Table 3.35 Calculated frequencies and intensities of the intermolecular modes


in the dimers of H2O and H2S52.

Frequency (cm - 1 ) Intensity (km/mol)

(H2O)2 (H2S)2 (H2O)2 (H2S)2

a" shear 605 217 185 20


a' shear 375 130 78 21
a' stretch 175 43 127 0.7
a" torsion 145 67 64 9.5
a' bend 142 62 237 48
a" bend 121 40 142 34

3.6.3 Effects of Electron Correlation and Matrices


Recent work has addressed the issue of how much correlation influences the internal
frequencies of water when involved in its dimer54. First of all, comparison of the data in
Tables 3.33 and 3.36 indicate that the SCF frequencies and shifts of the 6-31G** and
6-311+G(2d,2p) basis sets are very similar, suggesting the extra functions in the latter
larger set have only minor effects on these quantities. Focusing now on Table 3.36 shows
that correlation lowers the two stretching frequencies. While the SCF and correlated fre-
quency shifts of the acceptor molecule are virtually identical, including correlation adds a
significant increment to the red shift of the donor. The magnitude of the shift of the sym-
metric stretch is doubled by correlation, while the asymmetric stretch rises from —21 to
—28 cm - 1 . It is further notable that extending the correlation treatment beyond MP2 ap-
pears to have no significant impact on the results. The authors were particularly concerned
with the discrepancy between gas-phase and matrix data, summarized in the last two rows
of Table 3.36. After noting that the calculated frequency shifts match the matrix results
much better than they do the gas phase, they went on to argue for a reinterpretation of the
latter measurements.
The abilities of different types of correlation treatments to properly handle the vibra-
tional frequencies were the subject of a careful inquiry55. Frequency shifts of the intramo-
lecular modes of the water dimer, calculated at various levels with the harmonic approxima-
tion, are exhibited in Table 3.37. All computed frequencies are higher than the experimental

Table 3.36 Calculated frequencies and changes (in c m - 1 ) induced by H-bonding in


intramolecular modes of (H2O)2, calculated with 6-311+ G(2d,2p) basis set, in c m - 1 . 54
sym stretch asym stretch

V
mon Vd Va V mon Vd Va

SCF 4146 -45 -7 4247 -21 -11


MP2 3865 -91 0
3986 -28 -13
MP3 3878 -87 -8 3986 -26 -12
MP4 3838 -89 -9 3947 -27 -12
expt (gas phase) 3657 -125 -57 3756 -26 -34
expt (Kr matrix) 3628 -59 -3 3724 -23 -6
Table 3.37 Calculated frequencies and changes (in c m - 1 ) induced by H-bonding in intramolecular modes of (H2O)2,
calculated with TZ2P basis set, in cm - 1 . 55

sym stretch asym stretch bend

vmon vd va vmon vd va vmon vd va

SCF 3894 -51 -5 4238 -22 -10 1764 2 23


MP2 3861 -95 -12 3983 -31 -17 1663 2 29
CISD 3943 -17 +46 4042 +31 +44 1698 14 41
CCSD(T) 3845 -72 -8 3951 -26 -13 1679 1 28
expt (Ar matrix) 3638 -64 -4 3733 -24 -7 1596 3 21
168 Hydrogen Bonding

values in the last row of the table, and correlated frequencies are smaller than SCF. The
MP2 and CCSD(T) results are similar to one another and closer to experiment than
CISD. The latter method yields especially poor shifts in the frequencies, when compared
to the monomer, even to the point of predicting the incorrect sign of the shift. Overall,
the SCF shifts are in surprisingly good coincidence with experiment for all three intra-
molecular modes. Correlated shifts are significantly larger in magnitude, with MP2 shifts
larger than CCSD(T). The CISD method should be avoided for calculations of this sort. The
authors conclude with an important point: "the post-HF frequency shifts are not necessar-
ily better than the HF ones unless the calculational level is high and the basis set used is
large."
Calculations performed by Woodbridge et al.53 allow a clear picture of the effects of
electron correlation upon the harmonic frequencies of the weakly bound H2S dimer. First
of all, as may be noted from Table 3.38, correlation has a substantial lowering effect upon
the frequencies of the monomer, typically by about 100 c m - 1 . Whereas the SCF calcula-
tions predict very minor changes in the frequencies upon H-bond formation, these changes
are much more pronounced at the MP2 level. A 35 cm-1 red shift is calculated for the sym-
metric stretch in the donor, corresponding to the vs mode. These stronger changes are con-
sistent with the strengthening effect of correlation upon the H-bond. One may conclude that
correlation is very important when considering the shape of the potential energy surface in
complexes like (H2S)2 which contain second-row atoms.
The reader may have noted that experimental spectra of H-bonded species are commonly
measured in either the gas phase or in inert gas matrices. Of course, there may be some dif-
ferences as the molecules of the matrix can interact in various ways with the H-bonded com-
plex. A recent set of measurements56 provides some estimates as to the perturbations caused
by the matrix. Table 3.39 reports in the first row the frequencies of the OH stretches of the
free and bridging hydrogens of the proton donor molecule of the water dimer in the gas
phase. The next row indicates that a Ne matrix has only a very small effect, perhaps 10
c m - 1 . The Ar and Kr matrices produce larger perturbations, reducing the frequencies by
about 30 cm - 1 . A smaller cluster of Ar atoms, averaging perhaps 50 such atoms yields a re-
sult very much like a full Ar matrix. With the single exception of the very small increase
for the free OH stretch in the Ne matrix, all matrices and the Ar cluster lower the frequen-
cies of both of the modes studied.

3.6.4 Substituent Effects


With regard to changes induced by replacement of hydrogens by other groups, it has been
mentioned above that when water is paired with methanol, it is not entirely clear which of

Table 3.38 Calculated frequencies and changes (in c m - 1 ) induced by H-bonding in


intramolecular modes of (H2S)2, calculated with 6-3lG(2d) basis set, in c m - 1 53.
SCF MP2
v
V mon vd va mon vd vd

sym stretch 2842 -1 +1 2694 -35 -5


bend 1332 +4 -1 1233 12 -3
asym stretch 2857 +2 - 1 2720 -10 -5
Vibrational Spectra 169

Table 3.39 Frequencies (cm - 1 ) measured for the proton donor


molecule of the water dimer in various media56.

free OH bridging OH

gas phase 3730 3601


Ne matrix 3734 3590
Ar matrix 3709 3574
Kr matrix 3701 3569
ArN cluster 3714 3576

the two will act as proton donor and which as acceptor. Table 3.40 provides vibrational spec-
tral data57 to illustrate the very different frequencies of the two competing complexes. In
the situation where water acts as proton acceptor, all three of the water intramolecular fre-
quencies are red-shifted by small amounts. This contrasts with the values in the last column
of the table where the shifts would be larger, and that of the bend would be to the blue, were
the water to serve as proton donor. The patterns in the methanol molecule are even more
discrepant. When water is the proton acceptor, the OH stretching frequency in the methanol
molecule is strongly diminished, and a small increase is noted for the C—O stretch. The
latter would be red-shifted if water were donor, and there would be little change in the OH
stretch.
Whereas experimental assessments of the frequency of the OH stretch in the donor mol-
ecule of the water dimer cover a range between 3500 and 3600 cm-1 in the gas phase58-60,
the assignment is clearer in the methanol dimer, at 3574 cm - 1 . 6 1 A recent work has opti-
mized the geometries of the dimers of water, methanol, and silanol at the MP2 level62. The
vibrational frequencies include correlation by this approach, and are then corrected for
BSSE and anharmonicity. The basis sets applied were DZP, as well as a triple- set, and is
polarized under the rubric VTZ(2df,2p).
The data in the first row of Table 3.41 refer to the red shifts of the OH stretching normal
mode for each of the three dimers. (The HOD dimer was used instead of (HOH)2 so as to

Table 3.40 Calculated frequencies and changes (in c m - 1 ) induced by


H-bonding in complex between water and methanol. The role of water as
either proton acceptor or donor is indicated by PA and PD, respectively. Data,
computed with 6-31G** basis set57.

FA PD

SCF MP2 obsd SCF

H2O
V1 -2 -8 -16 -50
V2 -3 -5 -4 + 28
-7 -23 -14 -28
CH3OH
vCO 12 33 14 -16
vOH -72 -111 -128 -0
170 Hydrogen Bonding

Table 3.41 Calculated frequency shifts (in c m - 1 ) of the OH stretch in the dimers of water,
methanol, and silanol. Data computed with VTZ(2df,2p) basis set62.

HDO..HOD (CH3OH)2 (SiH3OH)2

SCF MP2 SCF MP2 SCF MP2

normal -74 -133 -71 -111 _


decoupled -76 -136 -72 -156 -112 -208
CPa -73 -123 -71 -142 -111 -196
anharmonic -89 -164 -86 -191 -136 -259
CPa -87 -150 -85 -177 -135 -245
a
With counterpoise corrections.

decouple the OH and OD stretches that exist as symmetric and asymmetric stretches in
HOH.) The next rows consider the OH stretch in isolation from the other motions within
each monomer, using a one-dimensional treatment. Similarity of the first and second rows
illustrates this to be a valid approximation. The third row is similar to the second, except
that counterpoise corrections have been used to modify the potential. Such corrections ap-
pear to have a small, albeit significant, lowering influence upon the magnitude of the SCF
red shift. The last two rows add anharmonicity effects into the OH stretching frequency. The
anharmonicity magnifies the red shift quite appreciably, particularly at the correlated MP2
level.
The authors conclude with their contention that a correlated treatment of the OH stretch,
in isolation from other motions, and corrected by a one-dimensional anharmonic approach,
can produce frequency shifts within 10 c m - 1 of experiment. Comparison of the three sys-
tems indicates the OH stretching frequency suffers somewhat of a larger red shift in the
methanol dimer than in the water dimer; however this difference might not be observed at
the SCF level. The shifts in the silanol dimer are quite a bit larger in magnitude.
Recent calculations of the phenol-methanol pair provide results that compare remark-
ably well with experiment63. The experimental frequencies listed in Table 3.42 were ob-
tained by spectral hole burning and dispersed fluorescence spectroscopy which permitted
assignment of the various bands. The agreement with the SCF/6-31G* frequencies is all the
more impressive due to the absence of electron correlation and anisotropy effects.

Table 3.42 Comparison of calculated and experimental


intermolecular frequencies (in c m - 1 ) in the phenol-methanol
complex. Data calculated at the SCF/6-31G* level63.
Calc Expt

H-bond stretch, 158 162


rocking, p2 17 22
torsion, 30 35
wagging, 2 55 55
rocking, 1 70 65
wagging, 1 90 91
Vibrational Spectra 171

When one of the H atoms of a water molecule of (H2O)2 is replaced by Cl, the interac-
tion is strengthened64. In this complex, HOC1 acts as proton donor, due to its enhanced acid-
ity. The red shift of the OH stretch is calculated to be 185-195 cm - 1 , in nice agreement
with an experimental measurement of 229 cm-1 ,65 and considerably larger than in (H2O)2.

3.6.5 NMR spectra


There have been some calculations carried out to consider the NMR spectra of the water
dimer and related systems. Using a GIAO approach, and a small basis set66, it was demon-
strated that the most shielded direction for the bridging proton generally coincides with the
H-bond axis, in agreement with experiment. The calculated isotropic shifts for the water
dimer and its ionic counterparts fall in the same region as experimental data. The anisotropy
typically increases as the molecules are brought together. Other calculations67 have shown
that the deshielding of the bridging hydrogen can be attributed to three factors. The proton
loses overall density as the H-bond is formed. The proton acceptor molecule deshields the
perpendicular components and shields the parallel ones. The proton donor atom has a more
varied effect but also deshields the perpendicular components. The linear correlation noted
by these authors between isotropic and perpendicular shifts was later confirmed by multi-
pulse NMR data68. A means of incorporating a counterpoise correction for BSSE was later
developed69 and applied to the water dimer, using a larger 6-311G** basis set. These BSSE
corrections appear to be negligible for the bridging hydrogen, but effects upon the other hy-
drogens are significant.

3.7 Expected Accuracies

For anyone considering carrying out ab initio calculations of vibrational spectra, or simply
interested in a thumbnail critique of a given paper in the literature, it would be particularly
useful to have at hand some information as to what level of accuracy might be expected
with a given basis set. We therefore take a brief interlude to explore this question for the
dimers of HF and H2O, since both of these cases have witnessed the heaviest barrage of all
levels of theory over the years.
A comparison of a broad range of basis sets, varying from minimal to quite extended,
was carried out several years ago for the HF and water dimers70. The limitation of this study
is that it did not go beyond the SCF level, nor did it include anharmonicity, so comparisons
with experiment are tenuous. Nevertheless, the data do illustrate the trends and provide use-
ful information as to the types of errors likely to be incurred for a H-bonded system with
any given basis set type.

3.7.1 HF Dimer
Data computed for the intramolecular vibrational modes of HF and its dimer are reported
in Table 3.43. Taking the values in the last row, computed with a very extended basis set,
as a benchmark, it is immediately apparent that frequencies computed with small unpolar-
ized basis sets are several hundred c m - 1 too small. 3-21G is probably the worst offender
in this regard. Once polarization functions have been added, even a single set, the frequen-
cies are more in line with those of the better basis sets. The same patterns are observed in
the intensities which are significantly underestimated with the small unpolarized basis sets.
172 Hydrogen Bonding

Table 3.43 Calculated frequencies (in cm - 1 ) and intensities (km m o l - 1 ) of HF


and (HF)2. A and D refer to proton acceptor and donor, respectively70.

v A

(HF)2 (HF)2

Basis set HF A D HF A D

MINI-1 4183 4171 4131 8.6 23 279


3-21G 4064 4041 3958 33 79 371
4-31G 4124 4102 4047 77 110 395
DZ 4103 4070 4021
DZP 4440 4397 4349
6-31G(d,p) 4494 4453 4406 133 182 374
6-31 + G(d) 4314 4279 4242
+VPS 4485 4449 4401 171 170 450
+VPs(2d) 4488 4448 4394 168 167 459
Extended 4452 4418 4366

The limits of accuracy were probed in a study which focused on the gradient and force con-
stants in diatomics like HF71. As more basis functions are added to one of 6-31G** type,
changes of the order of 1 % occur in the force constant, at both correlated and SCF levels.
The superposition contribution to the gradient is fairly large, and can account for variation
of as much as 0.004 A in the bond length.
If one concedes that SCF vibrational spectra, with no correction for anharmonicity, are
unlikely to reproduce experiment, the next question would concern whether such calcula-
tions are capable of reproducing changes that occur in each molecule upon formation of the
H-bond. Table 3.44 lists the shifts in the HF stretching frequency that result upon dimer-
ization, along with the intensification, expressed as a ratio between that of the dimer ver-
sus that of the monomer. The results are in many ways a confirmation of the absolute val-

Table 3.44 Frequency shifts and intensification ratios


(dimer/monomer) resulting from dimerization of HF70.

v(cm-1) Ad/Ain
Basis set A D A D

MINI-1 -12 -52 2.7 32


3-21G -23 -106 2.4 11
4-3 1G -22 -77 1.4 5.1
DZ -33 -82
DZP -43 -91
6-31G(d,p) -41 -88 1.4 2.8
6-31 + G(d) -35 -72
+ VPS -36 -84 1.0 2.6
s
+VP (2d) -40 -94 1.0 2.7
Extended -34 -86
Vibrational Spectra I 73

ues in Table 3.43. That is to say, the results with the unpolarized basis sets are undepend-
able. The minimal basis set greatly underestimates the frequency shifts of both molecules.
3-21G and 4-31G are both double-valence type sets; nonetheless, they yield very different
shifts in the donor molecule, and the acceptor shifts are both too small. The DZ set yields
good results, but it is difficult to say if this is fortuitous. The polarized sets, on the other
hand, all yield frequency shifts in decent agreement with the extended set. Turning to the
magnifications in the intensities resulting from dimerization on the right side of Table 3.44,
MINI-1 and 3-21G both fare especially poorly. Unlike the larger sets that yield little change
in the intensity of the acceptor, and an enhancement of 2-3 in the donor, these two sets are
off by orders of magnitude. 4-31G is somewhat better, as is 6-31G(d,p). It would appear
that the intensity enhancements are somewhat more demanding in terms of basis set qual-
ity than are the frequency shifts.

3.7.2 Water Dimer


The water dimer adds a number of new dimensions to the problem since each water mole-
cule contains three vibrational frequencies instead of one. The two stretching modes are la-
beled v1 and v3; v2 refers to the symmetric bending motion. The frequencies computed for
the water monomer are reported in the first three columns of Table 3.45, followed by the
corresponding frequencies in the dimer. As in the case of (HF)2, the unpolarized basis sets
strongly underestimate the stretching frequencies in the monomer. On the other hand, the
bending frequency is computed reasonably well with all of the sets, albeit the small unpo-
larized sets do yield a bit of an overestimate. Rather similar patterns are evident in the dimer
as well. The unpolarized sets underestimate v1 and v3 and yield a small overestimate, by
less than 100 c m - 1 , of the frequency for v2.
The shifts in each intramolecular vibrational frequency that occur upon dimerization of
water are described in Table 3.46. Whereas even very small basis sets seem capable of pre-
dicting qualitatively correct shifts in the HF dirner, (H2O)2 apparently represents a more
stringent test. For example, the three larger basis sets predict that both stretching frequen-
cies of the proton-acceptor molecule will be lowered, and are in good agreement as to the
amounts. The smaller sets in the first three rows of Table 3.46, on the other hand, yield er-
ratic results. All three make the erroneous prediction of a blue shift in v1; there is no con-
sistency at all for v3. Correct prediction of the behavior of the bending frequency is appar-
ently more demanding. Only the two polarized basis sets which also contain diffuse +
functions are in agreement. Even 6-31G(d,p) yields an incorrect sign. It might be stressed
at this point that reproduction of the large basis set results are rendered particularly diffi-
cult in the case of the acceptor, due to the small magnitudes of the shifts involved, all less
than 10 c m - 1 .
As in the case of (HF)2, the shifts are larger in the donor molecule. All basis sets, even
the smallest, agree that both stretches suffer a red shift and that the frequency of the bend
is increased. There is discrepancy concerning the magnitudes of these shifts. The three po-
larized sets concur that the red shifts are some 40-50 cm-1 for v1 and 20-30 cm-1 for v3.
The shifts predicted by the three smaller basis sets are erratic and generally undependable.
Turning now to the intensities of the various modes, the data in Table 3.47 indicate that
a polarized basis set like 6-31 G(d,p) offers a reasonable alternative to a much more extended
set such as 6-31 + +G(2d,2p), even if not quantitatively very accurate. The split-valence 4-
31G is considerably poorer, and the data with 3-21G are much worse, even though formally
of split-valence type as well. The intensity magnification ratios induced upon dimerization
Table 3.45 Calculated frequencies (in km m o l - 1 ) of H2O and (H2O)27

(H2O)2

H2O A D

Basis set v1 v2 V3 v1 V2 V3 v1 V2 V3

MINI-1 3897 1816 4127 3900 1806 4115 3843 1852 4071
3-2 1G 3811 1799 3945 3817 1785 3945 3712 1845 3891
4-3 1G 3960 1767 4098 3979 1771 4121 3907 1813 4085
6-31G(d,p) 4149 1770 4267 4144 1767 4258 4102 1798 4240
6-31+G(d) 4071 1797 4190 4068 1806 4181 4028 1826 4165
6-3l + +G(2d,2p) 4128 1746 4235 4123 1752 4226 4076 1767 4213
HF-limit 4130 1747 4231
Vibrational Spectra 175

Table 3.46 Calculated frequency shifts (in c m - 1 ) occurring upon dimerization


of water70.

A D

Basis set V1 V2 V3 V1 V2 V3

MINI-1 3 -10 -12 -54 36 -56


3-21G 6 -14 0 -99 46 -54
4-3 1G 19 4 23 -53 46 -13
6-31G(d,p) -5 -3 -9 -47 28 -27
6-31 +G(d) -3 9 -9 -43 29 -25
6-31 + +G(2d,2p) -5 6 -9 -52 21 -22

are listed in Table 3.48. If one is interested in these properties, the 6-31G(d,p) basis would
appear satisfactory in most cases. 4-31G does fairly well, the primary exception being its
five-fold exaggeration of the enhancement of the first stretching mode in the donor mole-
cule. 3-21G should be avoided.
In summary, calculation of vibrational frequencies can be meaningful, even if restricted
to the SCF level and with no account of anharmonicity. The frequencies are less demand-
ing of basis set quality than are the intensities. In some cases, one can compute reasonable
estimates of dimerization-induced frequency shifts with basis sets of 4-31G type, although
polarization functions are strongly recommended for uniform quality of results. Intensity
calculations without polarization functions can be expected to yield only the crudest of es-
timates. Reasonable results can be achieved with only one set of such functions on each
atom.

3.8 HYH...NH3

Somewhat more complicated than the complexes discussed earlier is the combination of
water with ammonia. The IR spectral characteristics of this complex were calculated using
a variety of basis sets and the results are presented in Table 3.4972. Also presented in this
table are comparable data computed at a correlated (MP2) level73, for purposes of compar-
ison. The red shift of the proton donor water molecule vs band is 103 c m - 1 , quite a bit lower

Table 3.47 Calculated intensities (in km mol - 1 ) of H2O and (H2O)270.

(H2O)2

H2O A D

Basis set v1 v2 v3 v1 v2 v3 v1 v2 v3

3-21G 0.05 80 9 5 98 45 282 98 52


4-31G 4 125 54 8 172 131 283 138 110
6-31G(d,p) 16 105 58 26 112 89 184 99 104
6-31++G(2d,2p) 14 91 81 22 110 103 191 65 127
176 Hydrogen Bonding

Table 3.48 Magnifications of intensity of vibrational bands occurring upon


dimerization of water70.
A D

Basis set V1 V2 V3 V1 V2 V3

3-21G 100 1.2 5 5640 1.2 6


4-3 1C 1.9 1.4 2.4 71 1.1 2.0
6-31G(d,p) 1.6 1.1 1.5 11 0.9 1.8
6-31 + +G(2d,2p) 1.6 1.2 1.3 14 0.7 1.6

than in H3N...HF, and even smaller than in H 3 P ... HF or H2O...HF. This shift is similar to
that in H2O...HC1 where HC1 acts as the donor. Correlation increases the red shift to 167
c m - 1 , close to the 197-209 c m - 1 observed in Ne, Ar, and Kr matrices74-75. The intensifi-
cation of this band is surprisingly large at 14, and grows to 88 when correlation is added.
The patterns of frequency shifts and intensity changes for water in H 3 N ... HOH are very dif-
ferent than when water acts as the proton acceptor as in H2O...HF. This set of differences
may act as a simple marker when there is some question as to the nature of a given com-
plex in a spectrum. As a proton donor, the bending frequency of the water molecule is blue
shifted and the other stretch reduced by a small amount. Note that correlation increases the
magnitude of these shifts.
When complexed with HF, the stretching bends in NH3 were unshifted but intensified
by upwards of thirty-fold. The situation is different when NH3 is bound to H2O, as the in-
tensifications are much smaller, only an approximate doubling at the SCF level. However,
correlation drastically enhances these bands, intensifying them by an order of magnitude.
(Note that the lower symmetry of the H3N...HOH complex breaks the degeneracy of the
pairs of vibrational levels in NH3.) The effects of complexation on the bending modes of
NH3 are similar when either HF or H2O acts as proton donor. The lower-frequency bend is

Table 3.49 Frequencies ( c m - 1 ) and intensities (km mol - 1 ) of H2O and NH3, and the changes
resulting from formation of the H 3 N ... HOH complex. SCF data72 calculated with + VPS basis
set; MP2 values73 computed with 6-31G**.

V
V A n Adim/Amon
mon mo

Mode SCF SCF MP2 SCF SCF MP2


H2O
v1 4129 -103 -167 25 14 88
V2 1727 37 60 94 1.1 0.4
V3 4242 -16 -45 85 1.2 1.5
NH3
Vstr 3709 1 3 0.4 2.5 17
Vstr 3847 4 2 6 1.9 17
-15 0 2.0 46
Vbend 1096 83 36 242 0.9 0.9
Vbend 1794 0 -4 29 0.5 1.3
-2 -13 0.8 3.7
Vibrational Spectra 177

blue shifted and the other nearly unaffected; small changes in intensity are predicted, with
or without correlation.
The H-bond stretching frequency is calculated in Table 3.50 to be 115 c m - 1 , compara-
ble to that of H 3 P ... HF or H2O...HC1. The intensity of this mode is low, only 3.3 km mol - 1 ,
again similar to H2O...HC1 or indeed any of the H3Z...HF complexes, suggesting that this
mode represents a fairly pure H-bond stretching motion in all of these. The numerical value
does not reproduce very well the experimental measurement of 202 cm-1 in matrices74.
The other intermolecular modes are generally a little higher frequency and intensity than
those in H 2 O ... HCl. The two donor bending frequencies of 669 and 443 c m - 1 match the
experimental values of 662 and 430 surprisingly well. The same is true of one of the ac-
ceptor bends but the other is significantly in error.
Replacement of one of the hydrogens of H2O by an aromatic ring was indicated in the
preceding chapter to strengthen the H-bond with NH3. The effects upon the calculated vi-
brational spectrum are consistent with this observation76 in that the H-bond stretching fre-
quency of C6H5OH...NH3 is calculated to be 165 c m - 1 , somewhat larger than the 115 c m - 1
computed for H3N...HOH72. As indicated in Table 3.51, the computed frequencies match
the available experimental data for this complex exceedingly well, despite the lack of cor-
relation or anharmonicity effects.

3.9 (NH3)2

In 1987, Sadlej and Lapinski77 carried out a force-field analysis of the ammonia dimer with
the 4-31G basis set. Because of the difficulty in definitively locating the true minimum on
the potential energy surface, and the dubious ability of this small basis to correctly model
the interaction, the results are provided here for instructional purposes and should be taken
in that spirit only. Note first from the results in Tables 3.52 and 3.53 that the linear arrange-
ment is a transition state on the SCF/4-31G surface, indicated by the single imaginary fre-
quency. The symmetry of the linear structure allows one to assign each intramolecular mode
as primarily occurring within the donor or acceptor while this is not possible in the cyclic
geometry with a pair of equivalent molecules. It is stressed that assignment of certain bands
as "bends" is tenuous and does not differentiate between scissoring, rocking, or wagging

Table 3.SO Vibrational spectra calculated for intermolecular


modes of the H3N...HOH complex with +VP S basis set. Data at
SCF level72.

v(cm-1)
A(km mol - 1 )
a
Mode calc expt calc

H-bond stretch (v ) 115 202 3.3


donor bend 669 662 215
donor bend 443 430 124
acceptor bend 181 180 59
acceptor bend 219 411 37
torsion 50 20 0.5
a
See reference 74.
178 Hydrogen Bonding

Table 3.51 Vibrational frequencies (cm - 1 ) for intermolecular


modes of the H3N...HOC6Hg complex. Calculated data at
SCF/6-31G** level with +VPS basis set76.
V

Mode calc expt

H-bond stretch (v ) 165 162


wagging 305
rocking 242
wagging 64 62
rocking 31
torsion 37

motions. Red shifts of 50-60 cm-1 occur in the stretching modes of the proton donor mol-
ecule in the linear structure, while the cyclic arrangement has smaller shifts. The work in-
dicates red shifts in all stretching modes and shifts to higher frequency for internal bends,
whether cyclic or linear. The H-bond stretching motion in either configuration is in the
neighborhood of 100-150 cm - 1 , but again this mode is not pure by any means. It is inter-
esting finally to note that the total zero-point vibrational energies of the two geometries are
quite similar, 47.4 kcal/mol for the linear arrangement and 47.6 for the cyclic.

Table 3.52 Frequencies (cm - 1 ) and intensities (D2 A- 2 a m u - 1 )


calculated for the ammonia dimer in its linear configuration77.

Mode vmon v A

proton donor
a", stretch 3929 -28 0.09
a' , stretch 3898 -59 1.78
a', stretch 3738 -22 0.01
a', stretch 3710 -50 1.52
a", bend 1866 45 0.83
a', bend 1830 11 0.77
a', bend 822 200 10.07
proton acceptor
a', stretch 3926 -31 0.36
a", stretch 3922 -35 0.38
a', bend 1836 15 1.15
a", bend 1829 8 1.17
a', bend 817 195 15.71
intermolecular
Mode V A
a', v 112 0.76
a' 134 1.16
a', bend 381 1.71
a", bend 314 1.20
a' 112 0.76
a", bend imaginary
Vibrational Spectra I 79

Table 3.53 Frequencies (cm - 1 ) and intensities (D2 A-2 amu - 1 )


calculated for the ammonia dimer in its cyclic configuration77.

Mode vm o n v A

intramolecular
bu stretch 3917 -40 1.81
au, stretch 3916 -41 0.42
bg, stretch 3916 -41 0
ag, stretch 3916 -41 0
bu, stretch 3728 -38 0.27
ag, stretch 3724 -36 0
a u ,bend 1856 35 1.90
bg, bend 1836 15 0
ag, bend 1829 8 0
b u ,bend 1821 0 1.77
a g ,bend 869 247 0
bu, bend/torsion 771 149 27.0
intermolecular
Mode v A
137 0
ag, bend 477 0
au, bend 255 2.06
bg,bend 159 0
au, bend 96 0.61
bu, bend 93 5.55

Yeo and Ford78 worked out the atomic polar tensors of the cyclic and linear ammonia
dimers to help interpret the spectral intensities. They found that dimerization produced only
very minor changes, with the exception of the proton donor N and H atoms, particularly the
latter atom, for the linear structure. The cyclic geometry sees the largest changes in the two
bridging hydrogens. It is the charge flux that appears to be most important in the intensity
changes seen upon dimerization.
The vibrational spectra of the linear and cyclic geometries of the ammonia dimer were
computed at the correlated level, using a polarized basis set in 199279. The striking con-
trasts between the spectra of these two geometries are clear from the data in Table 3.54. For
example, a number of the modes in the cyclic geometry would have zero intensity in the
harmonic approximation, whereas some of the others would be considerably more intense
than in the linear configuration. Vibrational frequency shifts, too, exhibit discrepant be-
havior in the two geometries. Comparison of the data in Table 3.54 with those in Tables 3.52
and 3.53 points out some of the dramatic changes that occur when correlation effects are
included, along with a polarized basis set. As an example, the SCF frequencies of all stretch-
ing modes in the linear geometry are red-shifted by 20-60 c m - 1 , whereas MP2 calculations
indicate very little shift at all for a number of these stretches. The red shifts of these same
modes in the linear structure are also reduced considerably upon adding correlation.

3.10 Carbonyl Oxygen

The carbonyl oxygen provides a contrast to the hydroxyl atom in a number of ways. It is in-
teresting to examine how the C = O bond reacts when this group accepts a proton from a donor.
180 Hydrogen Bonding

Table 3.54 Frequency shifts and intensity ratios of the vibrational modes of
NH3 that accompany dimerization into the linear and cyclic ammonia complex.
Data computed at the MP2/6-31G* level79.

v(cm-1) A dimer /A monomer

Parent mode linear cyclic linear cyclic

a1 stretch 1 -11 0.3


-35 -7 335 75
e stretch 0 -12 9 —
-2
-45 153 3
0 -3 11 —
-11 -12 1.2 87
a1 bend 58 39 0.9 —
13 15 0.9 2.2
cbend 2 8 0.6 —
2 23 1.0 1.9
41 11 0.6 —
-7 -18 1.5 1.1

3.10.1 Relationship between E and v


Latajka and Scheiner80 considered the nature of the relationship between the stretching fre-
quency of this covalent C=O bond and the strength of the H-bond, or indeed other sorts of
interaction. For this purpose, a water molecule was brought up toward the carbonyl oxygen
of H2CO and allowed to approach to within a set of specific distances. For interaction en-
ergies in the 2.5-4.5 kcal/mol range, there appeared to be a linear relationship between E
and v, where the latter quantity is the red shift of the C=O vibration. The weakening of
the C=O bond, implied by the lower frequency, is consistent with a picture of the H-bond
wherein the group takes on a certain amount of C — O — H character. This notion is con-
firmed by the lengthening of the C=O bond associated with its acceptance of a proton, and
discussed in the preceding chapter. What was most interesting about this particular study
was the close linear correspondence between v and the H-bond energy.
The small values clouded numerical precision of the relationship, so stronger interac-
tions involving ions were considered as well. The ions that were allowed to interact with
H2CO were H 3 O + , Na + , and Mg +2 . These interactions covered a wide range of E be-
tween — 8 and —75 kcal/mol. A very nearly linear relation was noted between E and v,
which suggested that each 1 kcal/mol strengthening of the interaction would shift the C=O
stretch to the red by approximately 2 cm - 1 . These linear relationships have foundation in
experimental work, as reported by Thijs and Zeegers-Huyskens81, who also confirmed that
the slope of 2 cm-1 kcal-1 mol is consistent with observation in solution. (In fact, the red
shift induced by formation of a H-bond appears to occur even when the C=O group is part
of a monolayer assembly82.) There is an angular dependence to the relationship between
the stretching frequency and the interaction energy. When the proton donor approaches
along the C=O direction, the amount of red shift is less than when it is directed along a
lone pair, that is, 8 (C=O...X) ~ 120°, given the same interaction energy.
More recent calculations have considered the same problem from the perspective of the
proton donor83. When water was paired with a set of N and O-type proton acceptor mole-
Vibrational Spectra 181

cules, a linear relationship was noted between the shift of the O—H stretching frequency
and the calculated strength of the H-bond. The slopes were somewhat different for the two
types of acceptors: a 23 cm-1 shift corresponds to 1 kcal/mol change in H-bond strength
for O acceptors, while the same energy enhancement in N-bases is accompanied by a 27
c m - 1 shift. (The foregoing analysis was aided by use of HOD, rather than HOH, so as to
more clearly distinguish the stretches of the two hydrogen atoms of water.) In either case,
the shifts associated with the proton donor are much larger than those in the acceptor mol-
ecule.

3.10.2 Formaldehyde + Water


High-level calculations84 have predicted the frequency shifts for the various internal modes
when water binds with formaldehyde. The equilibrium geometry is of type II (see section
2.7) in which water donates a proton to the carbonyl oxygen, and the water oxygen atom
also approaches within about 2.7 A of a CH2 H atom, as a secondary and weaker interac-
tion. As in earlier cases, correlation has a marked effect upon the vibrational frequencies in
Table 3.55, in most cases reducing these quantities by a substantial amount. On the other
hand, the SCF and correlated frequency shifts arising from complexation are not very dis-
similar. The trend of red shifts of one of the O—H stretches of water and of its bending
mode, matches the same trend in the proton donor of the water dimer. Indeed, the quantita-
tive aspects are rather similar to those in Table 3.33 particularly given the different basis
sets. The C=O stretching vibration is barely affected at all, surprising in light of the change
in this bond's length upon forming the H-bond. Most altered is firstly the CH2 rocking mo-
tion, probably due to its proximity to the water molecule, which hinders this motion. The
two C—H stretching vibrations undergo a blue and red shift, respectively.
The intermolecular frequencies are reported in Table 3.56 at various levels of theory,
along with the calculated SCF intensities in the last column. With only one exception, the
correlated frequencies are somewhat higher than the SCF values. Because of the bent na-
ture of the H-bond in H2CO--HOH, the H-bond stretching frequency is not particularly pure.

Table 3.55 Frequencies (cm- 1 ) of H2O and H2CO, and the changes resulting
from formation of the complex 84.

SCF/TZ2P(f,d) CCSD/TZ2P
expta
Mode V
mon v V
mon v v

HOH
v1, stretch 4261 -43 4023 -61 -58
V 2 , stretch 4073 18 3802 15 21
v3, bend 1780 -17 1722 -20 -24
H2CO
V1, C—H stretch 3173 20 3055 25 16
v2, C—H stretch 3125 -15 3004 -14 -6
v3, C=O stretch 1980 0 1790 -3 1
V4, HCH scissor 1645 6 1550 7 5
V
5,
CH2 rock 1343 32 1259 42 16
CH2 wag 1341 8 1214 7 2
V

a
In solid Ar.
182 Hydrogen Bonding

Table 3.56 Vibrational frequencies and intensities calculated for intermolecular modes of the
H2O-H2CO complex84.

v(cm-')
A (km m o l - ' )
a
Mode SCF/TZ2P(f,d) CISD/TZ2P CCSD/TZ2P expt SCF/TZ2P(f,d)

H-bond stretch (v ) 151 179 180 31


bend 329 359 354 250,261 103
bend 65 95 101 50
oopbend 457 501 493 435,439 131
H2CO rotation 148 171 173 0
torsion 39 29 20 68 155
a
ln solid Ar.

Its frequency of about 180 cm-1 is quite similar to that of the water dimer. The intensity is
only 30 km/mol, considerably weaker than the same band in (H2O)2. The two in-plane bends
listed in Table 3.56 are of quite different frequency. The first, of higher frequency, corre-
sponds to a distortion in which the XO-H angle becomes straighter and the OHO angle
more acute, while the lower refers to an overall straightening of both aspects of the H-bond.
The other intermolecular bends are all of a" symmetry. The out-of-plane distortion is the
highest frequency mode of all and the torsion the lowest. The latter frequency is quite a bit
smaller than that of the torsional mode of water dimer. Agreement with available experi-
mental observation in solid Ar matrix is moderate.

3.10.3 Formaldehyde + HX
The combination of H2CO with HC1 was considered, along with two different approaches
to electron correlation, by Rice et al.85. The harmonic frequencies they predicted for the in-
tramolecular modes at various levels are reported in Table 3.57, along with the changes in-
duced by formation of the H-bonded complex. The numbering scheme was taken directly

Table 3.57 Frequencies (cm - 1 ) of HC1 and H2CO, and the changes resulting from formation
of the complex85.
SCF/DZP SCF/TZ2P MP2/DZP CPF/DZP

Mode Vmon V Vmon V Vmon V Vmon V

HC1
V v3, stretch 3144 -103 3130 -109 3061 -251 3022 -151
H2CO
v1,CH2 a stretch 3226 30 3168 28 3128 46 3089 43
V2, CH2 s stretch 3149 22 3096 19 3040 32 3012 30
v4, CO stretch 2009 -10 1992 -11 1774 -11 1796 -11
vv CH2 scissor 1659 -1 1655 0 1572 -4 1566 -3
v6, CH2rock 1370 0 1373 2 1284 1 1281 2
v10, CH2wag 1338 6 1339 7 1217 4 1197 7
Vibrational Spectra 183

from their paper, and refers simply to the numerical order of frequencies, grouped by sym-
metry. The red shift of the HC1 stretch is calculated to be between 100 and 110 cm-1 at the
SCF level, very similar to that predicted when H2CO is replaced by H2O. This shift is sub-
stantially increased when correlation is included, especially by MP2, wherein the shift is
more than doubled. In particular, the calculated shift of 251 cm-1 is in superb agreement
with a measurement of 242 cm-1 in Ar matrix86. An examination of the sensitivity of the
red shift of the HC1 stretch in this complex to basis set87 indicates it remains in the 200-300
crn-1 range for basis sets varying from 6-31G(d,p) to 6-311 + +G(2df,2pd), all at the MP2
level. The stronger H2CO--HF complex shows a red shift of 422 cm-1 at the MP2/6-
31 l + +G(2df,2dp) level88.
Both CH2 stretching modes undergo frequency increases, in contrast to H2CO--H2O
where one of these modes is red-shifted. The CO stretch is shifted to the red by 10 c m - 1 , a
result which is insensitive to basis set or correlation. Again, this result is in excellent agree-
ment with experiment where a shift of 12 cm-1 has been measured86. The calculated red
shift in H2CO--HF is 13 cm - 1 . 8 8 The pattern in these H2CO--HX complexes is also distinct
from the H2CO--H2O complex where little shift of this stretching frequency is calculated.
The rocking motion of the CH2 group is not shifted by complexation in H2CO-HC1, unlike
the complex with water where a significant frequency increase is calculated and in fact ob-
served experimentally.
A systematic comparison of the shifts in the HX stretching frequency for a series of
H2CO-HX complexes in Ar matrix86 shows a clear correlation with the strength of the H-
bond as the shift varies in the order HF > HC1 > HBr > HI. There is no real pattern ob-
served for the C=O stretching frequency which is in the range between 10 and 17 cm-1
for all four complexes.
The intermolecular frequencies for H2CO-HC1 in Table 3.58 again show certain simi-
larities with H2CO--H2O. The H-bond stretching frequency in the 120-127 cm-1 range for
the former is quite close to the 118 calculated for the latter, all at the SCF level. Note the
small but significant increases that result when correlation is added. Indeed, correlation in-
creases the frequencies of all of the intermolecular modes of H2CO--HC1, and by substan-
tial amounts. The greater strength of the H2CO--HC1 interaction is reflected in its intermol-
ecular frequencies. At the MP2/6-311 + +G(2df,2dp) level 88, the H-bond stretching
frequency, v , is computed to be 248 c m - 1 ; the two higher-frequency shearing motions oc-
cur in the 754-791 cm-1 range.
The intensities calculated for the intramolecular vibrational modes of the H2CO--HC1
complex are reported for two different basis sets in Table 3.59. The intensification of the vs
mode, v3, is estimated in the range between seven- and eightfold, similar to that predicted
for H2O..HC1. Within the H2CO molecule, the two CH2 stretches both lose some intensity;
increases are observed in the CO stretch and the CH2 scissor. Analogous data for the inter-

Table 3.58 Frequencies (cm - 1 ) of the intermolecular vibrational modes of H2CO..HC185.


Mode SCF/DZP SCF/TZ2P MP2/DZP CPF/DZP

v8, in-plane stretch (v ) 127 120 170 155


v7, in-plane shear 394 391 554 478
v9, in-plane bend 38 36 54 45
v11, out-of-plane shear 399 384 521 460
v 12 , out-of-plane shear 112 122 186 158
184 Hydrogen Bonding

Table 3.59 Infrared intensities (km m o l - 1 ) of the intramolecular vibrational


modes of H2CO..HC1, calculated at the SCF level, and the enhancement ratio as
compared to the monomer85.

DZP TZ2P
A
Mode A dim dim/Amon A
dim A
dim /A mon

HC1
v3, stretch 412 8.4 388 7.4
H2CO
v1, CH2 a stretch 85 0.7 70 0.7
v2, CH2 s stretch 57 0.8 41 0.7
V4, CO stretch 211 1.3 186 1.2
V5, CH2 scissor 21 1.5 26 1.5
V6, CH2 rock 17 0.9 19 0.9
v10, CH 2 wag 1.7 0.9 2.3 1.0

molecular modes are listed in Table 3.60 where it may be seen that the H-bond stretch is
computed to have an intensity of some 16-22 km/mol. The v mode in H2O..HF is com-
puted to be even stronger, with an MP2/6-311 + + G(2df,2dp) intensity of 31 km/mol88. The
result for H2CO..HC1 is smaller than in H2CO..HOH, but nearly an order of magnitude
larger than in H2O..HC1. The other intermolecular bands of H2CO..HC1 have comparable
intensity to va with the exception of the out-of-plane shear, v 12 , which has a very low in-
tensity.

3 . 1 1 Imine

Migchels et al.89 have evaluated the effect of interaction with a proton-donating water or
methanol molecule upon the internal vibrational frequencies of a number of imines. These
perturbations are reported in Table 3.61 at the SCF/6-31G level and indicate first that the
red shift of the hydroxyl group of methanol is quite a bit larger than that of water. This dis-
crepancy may be due to the symmetric nature of the water molecule which thoroughly mixes
the two O—H stretches into a symmetric and asymmetric pair. The frequency chosen by
the authors as a reference point for water is 4145 c m - 1 , quite distinct from the 4032 cm-1
of the purer O—H stretch in CH3OH. Nonetheless, the large shifts in the methanol corn-

Table 3.60 Infrared intensities (km m o l - 1 ) of the intermolecular


vibrational modes of H2CO..HC1, calculated at the SCF level85.
Mode DZP TZ2P

v8, in-plane stretch (v ) 16 22


v7, in-plane shear 53 43
v9, in-plane bend 12 12
v n . out-of-planc shear 53 4)
v12, out-of-plane shear 0.1 0.0
Vibrational Spectra 185

Table 3.61 Frequencies (in c m - 1 ) of imine molecules, and the


changes resulting from formation of complexes with HOH or
CH3OH as proton donor. Data calculated at SCF/6-31G level89.

Mode vmon Av(HOH) Av(CH3OH)

CH 2 =NH
V
OH
-45 -173
V
CN
1866 -5 -5
V
NH
3686 21 20
cis-CH 3CH=NH
V
OH
-49 -202
V
CN
1884 -4 -3
V
NH
3659 25 25
trans-CH3CH=NH
V
OH
-51 -210
V
CN 1896 -6 -5
-y
NH
3699 14 14
CH2=NCH3
V
OH
-46 -177
V
CN
1901 -2 -1
CH3CH=NCH3
V
OH
-52 -213
V
CN
1932 -5 -3

plexes are consistent with experimental data reported by the same authors who found this
quantity to be —315 cm-1 in a complex very much like CH3CH=NCH3---CH3OH. There
are very small red shifts of the C=N stretch on complexation, and there is very little dif-
ference whether the proton donor is water or methanol. There are larger blue shifts of the
NH band, again insensitive to the nature of the donor. The authors discussed how the for-
mation of the H-bond alters the nature of the C=N stretching mode. In a monomer such as
CH 2 =NH, this mode includes a contribution from a scissoring motion of the CH2 group.
When water forms a H-bond, this mode becomes mixed with a certain degree of N—H
stretching as well.

3.12 Nitrile

The nitrile group as a proton acceptor provides an interesting counterpoint to the nitrogen
atom of amines which is involved in all single bonds, or the double bonds in imines. In each
case, the proton donor lies directly along the direction of the N lone pair. As one example,
frequencies and intensities for the complex of HCN with HF were computed in 1973 by
Curtiss and Pople90 and later with more flexible basis sets by Somasundram et al.6 The in-
tramolecular frequencies obtained are presented in Table 3.62, along with experimental in-
formation. The calculated frequencies are in error by several hundred cm -1 when compared
to experimental harmonic frequencies. The enlargement of the basis set markedly increases
the HF stretching frequency, but less of an effect is noted in HCN. Probably of greater im-
portance than the frequencies themselves are the shifts arising from complexation. The red
shift of the HF stretch is only 127 cm-1 with 4-31G and as large as 189 cm-1 with the big-
186 Hydrogen Bonding

Table 3.62 Frequencies (in cm-1) of HF and HCN, and the changes resulting from formation
of the HCN..HF complex. Data calculated at SCF level6,90.
4-31G DZP TZ2P expta

Mode V V Vmon v Vmon v Vmon v


mon

HF
V
HF 4117 -127 4511 -189 4471 -171 4138 -251
HCN
V
HC 3695 -13 3638 0 3600 -7 3442
V
CN 2384 12 2406 31 2408 19 2129 24
V
bend 911 26 861 17 869 12 727
a
Harmonic frequencies.

ger sets. This result is still quite a bit smaller than the best experimental estimate. It also
represents less of a shift than when HF is paired with the much more basic NH3. The CN
stretching frequency of HCN is shifted toward the blue, as is the bending mode. In contrast,
the HC stretch suffers a lowered frequency. The quantitative aspects of these changes are
rather sensitive to basis set.
The intermolecular frequencies in Table 3.63 indicate that the larger sets predict pro-
gressively smaller frequencies for the H-bond stretch, but the trends are more erratic for the
two bending modes. The frequencies seem to be calculated rather well, even with the small-
est 4-31G basis set, surprising in light of the low order of theory and the neglect of anhar-
monic effects.

3.12.1 Correlation and Anharmonicity


Botschwina91 later used a doubly polarized basis set to study this complex, along with a
CEPA-1 treatment of electron correlation. The ab initio energetics were fit to an analytic
four-dimensional function in order to elucidate anharmonic effects. The results at various
levels of theory are presented in Table 3.64 along with experimentally measured quantities.
Comparison of the SCF and CEPA-1 data suggests that while correlation yields major
changes in the frequencies themselves, the shifts that occur upon complexation are sur-
prisingly insensitive to correlation. The same is true of introduction of anharmonicity with
one major exception. Whereas the frequency shifts of the stretches of the HCN proton ac-
ceptor molecule are little affected by introduction of anharmonicity, the red shift of HF is
increased by 46% from 168 to 245 c m - 1 . This latter result is in near perfect agreement with

Table 3.63 Vibrational frequencies (cm - 1 ) calculated for intermolecular modes


of the HCN .. HF complex at SCF level6,90.
Mode 4-31G DZP TZ2P expt

H-bond stretch (v ) 193 176 159 155±1O


Donor bend (shear) 561 645 581 55±3
Acceptor bend 86 108 84 7()±24
Vibrational Spectra ! 87

Table 3.64 Harmonic and anharmonic frequencies (in c m - 1 ) of HF and HCN, and the changes
resulting from formation of the HCN..HF complex. Data91 calculated with doubly polarized
basis set.
SCF CEPA-1 expt

Mode
vmon v vmon v vmon v

HF
V
HF harm 4458 -164 4160 -168
anharm 4295 -242 3982 -245 3716 -245
HCN
V
HC harm 3604 -2 3439 -3
anharm 3505 0 3325 1 3310 1

V
CN harm 2407 21 2155 28
anharm 2385 19 2128 27 2121 24

the experimental quantity in the last column of Table 3.64. Indeed, the shifts of all the
stretches are calculated to lie very close to experiment for this complex. Note however that
the same cannot be said of the stretching frequencies themselves which remain in substan-
tial error when compared to experiment.
Amos et al.44 considered the same complex using comparable basis sets, and evaluated
the anharmonic constants using standard second-order perturbation formulas, based upon
third and fourth derivatives of the SCF energy. This treatment evaluates each vibrational
frequency, Vi, in terms of a purely harmonic potential i, and anharmonic constants xij re-
lating the various modes i and j (assuming all modes are nondegenerate).

The results furnish an interesting comparison with Botschwina's work, both because of a
different means of including correlation (MP2 vs. CEPA-1) and due to the alternate treat-
ment of anharmonicity. The Amos group also tested the feasibility of adding anharmonic
corrections to the intermolecular frequencies, as indicated in Table 3.65. It is first note-
worthy that the MP2 correlation treatment significantly enhances the red shift of the HF
stretch, as opposed to CEPA-1 which had only marginal effects upon any of the vibrational
frequencies. The anharmonic correction lowers this red shift but only by a little, bringing
the final shift into reasonable agreement with experiment. The MP2 shifts of the internal
frequencies of the HCN molecule are considerably different than the SCF values, with an-
harmonicity having an enlarging effect in one case, the CN stretch, but reducing the shift
in the other two modes.
In the case of the intermolecular frequencies, all the SCF values are too high, in com-
parison to experiment. These frequencies are further raised at the MP2 level. But when an-
harrnonicity is included, they are lowered. As a result, the V and shearing frequencies are
quite close to experiment although the bending frequency of the proton acceptor remains
too high. The authors considered the question as to whether a second-order perturbation
treatment is appropriate for the case of H-bonds. They concluded that terms higher than
quartic should typically be considered if possible.
188 Hydrogen Bonding

Table 3.65 Harmonic and anharmonic frequencies (in c m - 1 ) of HF and HCN, and the changes
resulting from formation of the HCN .. HF complex, along with intermolecular modes. Data44
calculated with doubly polarized basis set.
SCF MP2
Anharmonic
Mode V
mon V V
mon V v expt

Intramolecular
HF
V
HF 4322 -189 3955 -238 -226
HCN
V
HC 3638 0 3511 5 2
V
CN 2437 31 2050 45 53
bend 878 17 727 8 0
Intermolecular frequencies
acceptor bend 108 122 101 72 ±4
H-bond stretch (v ) 176 195 169 163
donor bend (shear) 645 697 539 550

Both electrical and mechanical anharmonicity can be considered in the calculations. This
was done92 using higher energy and dipole moment derivatives. Despite the use of rela-
tively small basis sets, the anharmonicity constants were in surprisingly good coincidence
with experiment.
The infrared and Raman intensities computed for the HCN . . HF complex are listed in
Tables 3.66 and 3.67 for the intramolecular and intermolecular modes, respectively. Like
the red shift, the intensification of the HF stretch is smaller for this dimer than for H 3 N .. HF,
indicative of the weaker binding. The Raman band is strengthened by a factor of 2.5. The
three vibrations of the HCN monomer all have reasonable IR intensity, varying between 10
and 80 km m o l - 1 . While the CN stretch is intensified by about 2.6, the other two modes are
relatively unaffected by complexation with HF. Raman intensities are all increased slightly.
The intensity of the intermolecular H-bond stretching band is fairly small, only 3 km m o l - 1 ,
in HCN .. HF, which matches quite closely to the same quantity in H 3 N .. HF. This similar-

Table 3.66 IR and Raman intensities of HF and HCN, and the changes
resulting from formation of the HCN . . HF complex. Data calculated with TZ2P
basis set6.
IR (km mol-1 ) Raman (A4/amu)

Mode A A.. /A S S,. /S


HF
147 4.8 26 2.5
HCN
V
ITC 76 1.1 14.1 1.1
V
CN
9.7 2.6 48 1.1
V
bend 66 0.9 1.8 1.1
Vibrational Spectra 189

Table 3.67 Vibrational spectral intensities calculated for


intermolecular modes of HCN .. HF 6 with TZ2P basis set.
Mode AIR (km mol - 1 ) SRaman (A4/amu)

H-bond stretch (V ) 3 0
donor bend (shear) 346 4
acceptor bend 30 6

ity indicates that the intensity is less sensitive to the strength of the H-bond as a reflection
of the fact that the mode in both cases is a simple stretching motion that involves little re-
orientation of the subunits, and hence only small changes in dipole moment of the complex.
In common with H 3 N .. HF, the bending motion of the proton donor is rather intense, fol-
lowed by a weaker acceptor bend. It is notable that the latter two intensities are significantly
stronger in HCN .. HF than in H 3 N .. HF. The Raman intensities are weak for all three inter-
molecular modes, V in particular.
Bouteiller and Behrouz93 re-examined the question of anharmonicity in the HCN .. HF
complex, with a particular eye toward the effects of basis set superposition. As in ap-
proaches of this sort, a two-dimensional grid of points was constructed and the energies fit
to a polynomial of r(HF) and R(N .. F), up to fourth order. They found first that the SCF
BSSE for this complex is 0.58 kcal/mol, as compared to an uncorrected interaction energy
of 6.69 kcal/mol. The authors next reoptimized the r and R parameters of the optimized
geometry and found a change of less than 0.01 A in the former while R elongates by 0.02
A, from 2.872 to 2.892 A. At the correlated MP2 level, the interaction energy is reduced
25% by BSSE and the optimized R(N .. F) stretched by 0.06 A. In the next step, the two-di-
mensional PES V(r,R) was refit to the calculated grid of points, but with counterpoise cor-
rection of the BSSE for each energy.
The results are presented in Table 3.68 which is divided into frequencies derived using
the harmonic approximation and anharmonic data taking account of the fourth-order poly-
nomial description of the PES. It appears that the SCF harmonic frequencies are barely al-
tered at all by BSSE. In fact, the only harmonic frequency to be affected is the vs stretch,
v(FH), which is increased by 12 cm-1 at the MP2 level when the BSSE is included. When
the treatment is expanded to include anharmonicity, there is again virtually no effect on ei-
ther frequency from BSSE. However, MP2 calculations do show significant changes: ac-
count of superposition error raises the vs frequency by 32 cm-1 and lowers v(F..N) by a

Table 3.68 Vibrational transitions (in c m - 1 ) calculated for HCN ... HF 93 with
and without correction for BSSE.
Harmonic Anharmonic
..
v(FH) v(F N) v(FH) v(F .. N)

SCF 4202 167 4010 168


SCF+BSSE 4202 164 4009 151
MP2 3785 201 3604 201
MP2+BSSF 3797 201 3636 171
190 Hydrogen Bonding

similar amount. The latter changes would be consistent with a weakening of the interaction
between the two molecules, which is in fact a direct result of correcting the BSSE.
Shortly thereafter, Bouteiller94 continued on this same line of reasoning and considered
combination bands and intensities. The data in Table 3.69 represent transitions of some
combination of the v(FH) and v(F .. N) bands, as indicated. At both the SCF or MP2 levels,
inclusion of BSSE corrections reduce the transition energies for excitation of the v(F..N)
mode from the ground state: ]00> -> |0n>, as may be seen from the first four rows of Table
3.69. The values listed in parentheses refer to the spacing between the successive transi-
tions. The counterpoise corrections reduce this spacing. In all cases, this spacing diminishes
with n, indicative of the mechanical anharmonicity of the H-bond. It is worth noting, how-
ever, that there is less such reduction when superposition errors are corrected.
The next sets of transitions all include an excitation of the v(FH) mode by one quantum,
that is, |0n> -> ln'>. The main transition, |00> -> |10>, is lowered by BSSE correction
by a very small amount at the SCF level, but increases by 31 cm-1 for MP2. For any given
progression, |0n> —> ln'> with fixed n, there is a reduced spacing between lines as n'
rises, due again to mechanical anharmonicity. The SCF spacings are relatively unaffected
by BSSE whereas correction of this error yields consistently reduced spacing at the MP2
level. Whether or not BSSE is corrected, the MP2 spacings are consistently larger than their
SCF correlates.
The authors also investigated the intensities of their various combination excitations.
The results led to the conclusion that the greatest intensities arise from the simple |00> —>
|10> transition. Also rather intense are the |0n> -> |ln> transitions wherein the v(F..N)
mode remains at the same level.
Del Bene et al.95 computed the vibrational spectra of the methylsubstituted complex of
CH3CN..HC1. Their intermolecular frequencies for the C3v equilibrium structure were in

Table 3.69 Vibrational transitions (in c m - 1 ) calculated for HCN...HF94 with and without
correction for BSSE.a Values in parentheses indicate increase relative to transition in
preceding row.

SCF SCF + BSSE MP2 MP2 + BSSE


100> 101> 164 150 200 171
100> 102> 315(151) 287(137) 388(188) 326 (155)
100> 103> 455 (140) 415(128) 562(174) 470(144)
100> 104> 584 (129) 539(124) 725(163) 611 (141)

100> 110> 4028 4019 3611 3642


100> lll> 4217(189) 4209 (190) 3838 (227) 3839 (197)

101 110> 3864 3869 3410 3471


101> lll> 4053 (189) 4059 (190) 3638 (228) 3669 (198)
101 112> 4216(163) 4211(152) 3841 (203) 3843(174)

102> 110> 3713 3732 3223 3316


102> 111> 3901 (188) 3922(190) 3451 (228) 3513(197)
102> 112> 4064(163) 4074(152) 3654 (203) 3688(175)
I02> 113> 4213 (149) 4214(140) 3844(190) 3848 (160)
a
ij> refers to v(FH)and v(F . . N) excitations, respectively.
Vibrational Spectra 191

good agreement with experimental values96. For example the calculated V is 117 c m - 1 , as
compared to 97±3 from the FTIR spectrum. This value is somewhat smaller than estimates
of the unmethylated complex with HF in HCN..HF. The mode corresponding to bending of
the proton donor has calculated and experimental frequencies of 417 and 350± 100 cm - 1 ,
respectively. The out-of-phase bending of the monomers is of much lower frequency, 35
cm-1 in the calculations as compared to 40±20 cm-1 in the experiment.

3.12.2 HCN as Proton Donor


Unlike most C—H bonds which are poor donors, the triple bond to the carbon atom makes
HCN a rather effective proton donor molecule. The frequency shifts encountered in the two
subunits when HCN is combined with NH3 as proton donor are listed in Table 3.70. There
is a substantial red shift of the HC stretch, corresponding to the vs band. This shift is re-
produced surprisingly well with the TZ2P basis set, even though the calculations are at the
SCF level. The CN stretching frequency is also lowered while the bend undergoes a sub-
stantial shift to the blue. These changes are quite a bit different than when HCN acts as pro-
ton donor, as with HF in Table 3.62. In such a case the HC stretching frequency is changed
only very little and the CN stretch shifts a small amount to the blue. The bending mode is
increased in either case but by a much smaller amount when HCN acts as proton acceptor.
It is interesting as well to compare the behavior of NH3 when it interacts with HCN as com-
pared to a strong proton donor like HF. In the former case, both stretching frequencies shift
to the red a small amount, whereas little change occurs with HF. In either case, the a1 bend-
ing frequency increases by nearly 100 cm - 1 , while the other bending mode is essentially
unaffected.
The intensifications of the IR and Raman bands in Table 3.71 exhibit the expected in-
crease of the IR band of the HC stretch, comparable in magnitude to the change of the vs
band in HCN...HF. On the other hand, this same band is weakened in the Raman spectrum,
opposite to the HF stretching band in HCN...HF. It is particularly intriguing to note the even
stronger intensifications of both the IR and Raman bands for the CN stretching mode, again
very much more exaggerated than when HCN acts as proton acceptor. Even the bending
mode shows significant enhancement in NCH ... NH 3 . The effects of complexation on the IR

Table 3.70 Frequency shifts (in crn -1) of NH3 and HCN arising
from formation of the NCH ... NH 3 complex. Data calculated at
SCF level6.

Mode DZP TZ2P expta

HCN
V
HC -189 -162 -162
V
CN -29 -26 -11
V
bend 153 131
NH3
V
str( 1) a -9 -7
vstr(e) -18 -12
v b e n d (a 1 ) 93 68
vbend(e) 0 1
a
Harmonic frequencies.
192 Hydrogen Bonding

Table 3.71 IR and Raman intensities of NH3 and HCN, and the
changes resulting from formation of the NCH ... NH 3 complex.
Data calculated with TZ2P basis set6.

IR Raman
Mode A dim /A mon dim/smon
s

HCN
V
HC 4.7 0.7

V
CN 7.3 8.3
V
bcnd 1.2 1.8

NH3
v str (a 1 ) 100a 0.9
v str (e) 3.4 1.0
v bcnd (a 1 ) 1 0.2
vbend(e) 1.1 0.9
a
Intensity in monomer too small for ratio to be meaningful.

intensities of the NH3 subunit are comparable to those in FH ... NH 3 , although there are some
quantitative differences.
The intermolecular frequencies in Table 3.72 are similar in pattern to other H-bonded
complexes6'97. The H-bond stretching frequency, in excellent agreement with experiment,
is smaller than in HCN ... HF or FH...NH3, consistent with a weaker bond. The highest in-
termolecular frequency is again the pivoting of the proton donor molecule. A comparison
of SCF and MP2 data indicates that correlation does not have a profound effect upon these
intermolecular frequencies. Its principal effect is an increase in the H-bond stretch fre-
quency. The intensities in Table 3.73 are in keeping with comparable systems. The H-bond
stretch is of low IR intensity, with the donor bend much stronger.

3.12.3 HCNDimer
It is of interest also to examine the dimer of HCN where one molecule acts as donor and
the other as acceptor. The results in Table 3.74 fit and confirm the patterns when HCN is
involved in complexes with other molecules. When acceptor, the HC frequency is changed

Table 3.72 Vibrational frequencies (cm - 1 ) calculated for intermolecular modes of


NCH ... NH 3 .
SCFa MP2b

Mode DZP TZ2P 6-3 1G* 6-31+G(3df,2p) expt

H-bond stretch (v ) 160 146 190 158 141 ± 3


donor bend (shear) 323 302 308 283
acceptor bend 120 113 123 132
a
Data from Reference 6.
b
Data from Reference 97.
Vibrational Spectra 193

Table 3.73 Infrared and Raman intensities calculated for


intermolecular modes of NCH ... NH 3 6 with TZ2P basis set.

Mode AIR(km mol-1) SRaman (A4/amu)


H-bond stretch (v ) 2 0
Donor bend (shear) 158 0
Acceptor bend 4 10

very little, while the other two modes are blue shifted by some 10-30 cm - 1 . (Note how-
ever, that the experimental result appears to have a small increase for the CH frequency.)
The IR intensity of the CN mode is increased, but the two other modes are changed only
little. When HCN acts as donor, the HC stretch undergoes a strong red shift. A smaller red
shift occurs in the CN stretch and a surprisingly large blue shift is noted in the bending
mode, comparable in magnitude to the vs shift. Intensity increases are observed in both
stretching modes, and a smaller enhancement in the bend.
The calculated intermolecular H-bond stretching frequency is surprisingly close to the
experimental value in Table 3.75. This frequency of some 120 cm-1 is smaller than in the
HCN..HF complex or that in NCH...NH3, confirming that HCN is neither as strong a pro-
ton donor as HF nor as good an acceptor as NH3. The IR intensity of this band is again quite
weak, as in all the other complexes where symmetry restrains this mode from including
much reorientation of the two subunits. The donor bend remains the mode of highest fre-
quency and the acceptor bend the lowest. The intensity of the former vibration is by far the
strongest of the intermolecular modes.
Somewhat later calculations by Kofranek et al.98 were able to incorporate correlation
into the spectral data of the HCN dimer. The SCF data in Table 3.7699,100 confirm the pat-
terns obtained earlier by Somasundram et al.6 with a different basis set. The correlated re-
sults in the next two columns of the table indicate no major changes in patterns. Curiously
enough, the inclusion of correlation reduces all of the red shifts while enlarging all the blue

Table 3.74 Frequency shifts and IR intensity enhancements of the two


monomers in HCNH ... CN complex. Data calculated at SCF level using DZP
basis set6.
Frequency shifts ( c m - 1 ) Intensity factor (Adim/Amon)

Mode calc expta calc expt

HCN (donor)
V
HC -69 -65 4.1 1
V
CN
-11 —9 4.8 2
V,bend, 76 1.3
HCN (acceptor)
V
HC -2 11 0.8 1.0
V
CN
16 8 4.0 3
V
bend 13 1.1
a
Harmonic frequencies.
194 Hydrogen Bonding

Table 3.75 Vibrational frequencies and IR intensities calculated for


intermolecular modes of the NCH ... NCH complex at SCF level6.

v (cm - 1 )
A(kmmol-1)
Mode calc expt calc

H-bond stretch (v ) 122 119 2


donor bend (shear) 159 122
acceptor bend 60 40a 10
a
See references 104 and 105.

shifts to higher frequency. The comparison with the experimental shifts in the last column
of Table 3.76 is particularly encouraging.
From Table 3.77, it appears that correlation exerts only a minor effect upon the CH
stretching intensities. The absolute values of the monomers are diminished somewhat, but
the dimerization-induced magnification of both the donor and acceptor are increased rela-
tive to the SCF results. The internal HCN bends are virtually unaffected. These trends are
consistent with the SCF results of Somasundram et al.6 achieved with a different basis set,
and recorded in Table 3.74. But quite dramatic effects are observed in the CN stretches. In
the first place, correlation reduces the intensity of the HCN monomer band by an order of
magnitude. Regarding the intensification caused by dimerization, SCF calculations from
Tables 3.74 and 3.77 predict a factor of perhaps 3-5. But this magnification enlarges when
correlation is included: the dimer/monomer ratio is calculated to be 10 for the proton ac-
ceptor and 64 for the donor. Recent experimental measurements support the correlated data
in that enhancements of approximately 30 are seen.
Analogous information about the intermolecular frequencies and intensities are exhib-
ited in Table 3.78. The SCF data are consistent with those in Table 3.75 for a different ba-
sis set. It appears that correlation has a surprisingly small effect on any of the frequencies.
The intensity of the lower-frequency bending mode is also changed very little by correla-

Table 3.76 Harmonic frequencies (in c m - 1 ) of HCN, and the changes resulting from
formation of the (HCN)2 complex. Data98 calculated with polarized basis set [641/41].

SCF CPF expta

Mode V mon v V mon v V mon v

HCN (donor)
V
CH 3618 -64 3493 -56 3441a -65
V
CN 2418 -11 2170 -3 2129a -6
V
bend 868 60 695 76 721b 77
HCN (acceptor)
V
CH 3618 -7 3493 -1 344 la 11
V
CN 2418 9 2170 13 2129a 8
V
bend 868 13 695 15 721b 13
a
Harmonic freqilencies 99 .
b
See Reference 100.
Vibrational Spectra 195

Table 3.77 Calculated intensities (km mol - 1 ) of HCN, and the changes resulting from
formation of the (HCN)2 complex. Data98 calculated with polarized [641/41] basis set.

SCF CPF
expta
Mode mon
A dim/Amon Amon Adtm/Amon dim/Amon

HCN (donor)
V
CH 75.2 3.7 62.3 4.7
V
CN 10.3 3.9 0.12 64.3 30±10
V
bend 83.0 1.0 86.7 1.03
HCN (acceptor)
V
CH 75.2 0.93 62.3 1.03
V
CN 10.3 3.1 0.12 10.2 30±10
V
bend 83.0 0.88 86.7 0.84
a
See reference 99.

tion. On the other hand, correlation approximately doubles the intensity of the H-bond
stretch, while diminishing that of the other bend by about one third.

3.13 Amide

Following earlier calculations with smaller unpolarized basis sets101, the vibrational spec-
trum of the complex pairing together two formamide molecules was later computed by
stergard et al.102. The data obtained with their best basis set, 6-31 + +G**, are reported
for the intramolecular vibrations at the SCF level in Table 3.79. We use the authors' origi-
nal nomenclature for the individual modes wherein refers to a torsion, to a wag, v to a
stretch, 8 to scissoring, r to rocking, and to an out-of-plane bend. The v(NH2) stretch prob-
ably most closely corresponds to the vs band of the proton donor. Table 3.79 indicates a red
shift of about 30 c m - 1 , as contrasted to only 3 or 4 cm-1 in the acceptor formamide mol-
ecule. This shift is rather small for a H-bonded complex, smaller, for example, than the shift
of the C—H proton in (HCN)2. Also of particular interest is the frequency of the C=O
stretch in the acceptor molecule, which diminishes by 27 c m - 1 . This red shift contrasts with
the same C=O stretch in the complex of H2CO with H2O, where little change is seen. In
addition to these bands, there are a number of other in-plane motions whose frequencies
change by 30 cm-1 or less. The largest shift occurs for the out-of-plane wag of the NH2

Table 3.78 Vibrational frequencies and IR intensities calculated for


intermolecular modes of the NCH ... NCH complex98.
v(cm-1) A(km mol - 1 )
Mode SCF CPF SCF CPF

H-bond stretch (v ) 106 116 1.3 2.8


bend 139 125 128 84
bend 48 38 13 16
196 Hydrogen Bonding

Table 3.79 Frequencies (in c m - 1 ) of formamide, and the changes


resulting from formation of the dimer.a Data calculated at SCF
level with 6-31 + +G** basis set102.
Mode vvmon donor V
acccptor

in-plane(a')
(NCO) 618 14 4
r(NH2) 1154 30 11
v(CN) 1368 21 18
(CH) 1547 —2 1
(NH2) 1773 19 3
v(CO) 1966 -11 -27
v(CH) 3193 24 -4
v(NH2) 3835 -30
v(NH2) 3982 -29 -4
out-of plane(a")
(NH2) 252 253 88
(NH2) 664 78 15
(CH) 1175 6 5
a
refers to a torsion, to to a wag, v to a stretch, to scissoring, r to rocking, and 7 to
out-of-plane bend.

group. The donor frequency doubles upon formation of the H-bond, and an increase of 88
cm-1 occurs in the acceptor. This change can be accounted for based upon the low energy
cost of wagging in the isolated monomer. But this same motion would disrupt the H-bond
in the dimer, so the frequency rises accordingly.
Intermolecular frequencies, calculated with various basis sets, are listed in Table 3.80.
The authors point out that the normal modes are far from pure and that their nomenclature
is very approximate. For example, the H-bond stretch contains a strong element of donor
libration. Their designation of "strain" refers to simultaneous in-phase rotations of the two
molecules, while they use the term "bend" to indicate out-of-phase rotations. The imagi-
nary frequencies of the torsional modes are evidence that the true equilibrium geometry of
the formamide dimer is nonplanar. The v mode appears at about 120 c m - 1 , a little smaller
than that for H2CO...HOH where a water molecule donates a proton to the carbonyl oxy-

Table 3.80 Frequencies (in c m - 1 ) of intermolecular vibrational


modes of the formamide dimer. Data calculated at SCF level 102.

Mode 4-31G 6-31G** 6-31 + +G**

in-plane(a')
stretch, v a 144 122 114
bend 21 23 21
strain 43 59 48
out-of-plane(a")
torsion 3; 28; 15;
bend 30 28 21
strain 117 72 81
Vibrational Spectra ! 97

gen of formaldehyde, and fairly close to the V frequency in H2COH...C1. There is a mod-
erate level of sensitivity of the frequencies to the basis set. Outside of overestimating v and
the out-of-plane strain mode, 4-31G does surprisingly well.
The force constants in the cyclic formamide dimer were explicitly evaluated with the
6-31G** basis set103 at the SCF and MP2 levels. The intramolecular diagonal force con-
stants are listed in Table 3.81 for the CN, CO, and NH stretches, the latter being the H atom
that forms the H-bond bridge in the dimer. If one takes the magnitude of this force constant
as a measure of the bond strength, the CO bond is considerably stronger than CN (formally
a single bond), which is in turn stronger than NH. The entries for k in Table 3.81 repre-
sent the changes that occur in each force constant as the dimer is formed out of its con-
stituent monomers. The NH and CO bonds are both weakened while the CN bond becomes
stronger. These changes are consistent with the shifts of the corresponding vibrational fre-
quencies. The changes in the force constants are in the neighborhood of 7-10% at the SCF
level. Correlation reduces all of the force constants, the usual expectation for bonds of any
sort. Of greater interest are the effects of correlation upon the changes in the force constants
caused by dimerization. While the percentage change in the CO force constant is little af-
fected by correlation, the increase in the CN constant rises to over 10%. Even more dra-
matic is the N—H reduction which amounts to 18% at the MP2 level. This decrement is
nearly double that observed at the SCF level. This effect is consistent with the large corre-
lation-induced enhancement in the red shift of vs associated with H-bond formation.
It might be expected that the theoretical method that predicts the strongest interaction
between the two monomers should also yield the largest force constant for pulling the two
molecules apart. This supposition can be tested in Table 3.82 which indicates it is not fully
reliable. The minimal basis set predicts a very large force constant of 0.27 mdyn/A and a
binding energy of 16.0 kcal/mol. The next larger set has a stronger interation energy but a
much smaller force constant. As the method continues to improve, however, the pattern
does obey the expected connection between reduced binding energy and smaller force con-
stant.

3.14 Summary

SCF frequencies typically overestimate the various vibrational frequencies in H-bonded


complexes as well as in their constituent subunits. Once correlation is added, with MP2 usu-
ally a satisfactory approach, the computations can match fairly well the experimental spec-
tra, particularly if the basis set is large and flexible enough. Calculation of accurate vibra-

Table 3.81 Computed intramolecular force constants (mdyn/A) of the cyclic


formamide dimer103.
SCF MP2

Stretching mode k ka k ka

CN 8.98 0.76 8.45 0.89


CO 14.68 -1.10 12.26 -0.82
NH 7.71 -0.75 6.67 -1.18
a
Difference between force constant in dimer versus isolated monomer.
198 Hydrogen Bonding

Table 3.82 Comparison of force constant for intermolecular stretch of the cyclic formamide
dimer with the computed binding energy, without BSSE correction. Data103 are at SCF level
unless otherwise indicated.

MINI-1 MID1-1 4-31G 6-31G** MP2/6-31G**

k(v ), mdyne/A 0.27 0.17 0.15 0.12 0.15


— Eelec, kcal/mol 16.0 20.3 17.2 13.4 17.4

tional intensities is somewhat more demanding, and the results less stable with respect to
changes in basis set.
In conjunction with the stretch of the A—H bond that occurs in the donor molecule upon
formation of a H-bond, the stretching frequency of this bond is lowered by a good deal, as
much as several hundred c m - 1 . This red shift, coupled with a strong intensification of the
band, is characteristic of H-bond formation. The H-bonded complex contains a number of
intermolecular vibrational modes that do not exist within the separated monomers. Because
the H-bond is so much weaker than covalent bonds, the frequencies of these modes are gen-
erally less than 1000 c m - 1 , sometimes below 100 c m - 1 . The precise nature of these modes
differs from one system to the next but one can normally recognize several common ones,
such as a H-bond stretch, v , and bending motions of the donor. The frequencies are usu-
ally correlated with the strength of the interaction between the two subunits.
A normal mode analysis of (HF)2 reveals the mixing together of the various intermole-
cular stretches and bends, and underscores a certain amount of arbitrariness in their identi-
fication. The internal stretch in the proton donor molecule is calculated to shift to the red
by 100 c m - 1 , in good agreement with experimental measurement. The acceptor stretching
frequency is also diminished but by only about one third as much. The intensity of the donor
stretch is calculated to increase by a factor of 3 upon dimerization, while that of the accep-
tor is unchanged. The intermolecular stretch is in the range of 200 c m - 1 ; proton donor wags
are 500-600 cm~'. Vibrational spectral data for the analogous (HC1)2 are similar with the
proviso that the complex is more weakly bound. The red shifts of the HC1 stretches are con-
sequently smaller in magnitude; V is only around 60 c m - 1 , and the intermolecular donor
wags amount to less than 300 c m - 1 .
A complex like H 3 N ... HF is much more strongly bound so one might expect larger ef-
fects from formation of the H-bond. Indeed, the red shift in the proton donor HF molecule
amounts to more than 400 c m - 1 , about five times greater than in (HF)2, at the same level
of theory. The intensification of this band is greater than sevenfold, again several times
larger than in (HF)2. The proton acceptor NH3 has more internal modes than does HF. Most
of its frequencies are little affected by the complexation, with the exception of a 100 cm-1
increase in its symmetric bend. The intensities of the stretching modes of the NH3 mole-
cule are very strongly increased. Replacement of NH3 by PH3 leads to a weakening of most
of the patterns in H3N...HF, but there are certain anomalous changes as well. Analysis of
the intensities in terms of atomic polar tensors yields insights into the electronic redistrib-
utions that accompany the formation of each H-bond. For example, the intensification of
the v band is attributed to the fact that the bridging H is positive in the monomer and be-
comes more so in the complex. Coupled to this is the lesser ability of the charge cloud in
the complex to follow the motion of this proton; indeed, the density seems to move in the
Vibrational Spectra 199

opposite direction. Consequently, motion of the bridging proton yields a greater dipole en-
hancement than would occur in an isolated HF molecule.
The intermolecular modes also provide clear evidence of the greater strength of
H 3 N ... HF as compared to (HF)2. The frequency for wagging of the donor molecule is nearly
900 cm-1 and the H-bond stretch v is 240 cm - 1 . The C3v symmetry of this complex im-
parts to the latter mode a nearly pure stretching character. As the two molecules are pulled
apart, there is little off-axis motion occurring, so the dipole moment of the complex changes
very little. Consequently, the intensity of the v mode is quite weak. This sort of reasoning
can be generalized to rationalize the observed intensities on the basis of the balance between
stretches and bends in the mode in question.
Anharmonicity corrections can be quite large in complexes of this type, up to several
hundred c m - 1 . In FH...NH3 and C1H...NH3, for example, the anharmonic vs frequency is
lower than the harmonic value by about 400 c m - 1 . It is interesting also that whereas cor-
relation lowers the vs frequency, it has the opposite effect of increasing v . Nor can one ex-
pect cancellation to occur between anharmonic and correlation effects; in some cases, the
two can be additive. Mechanical anharmonicity leads to a progressively smaller spacing be-
tween successive overtones of the v mode as the quantum number increases.
Intermediate between the extremes of HP...HF and H3N...HF is the pairing of HF with
a water molecule. The red shift of the HF stretch in H2O...HF is somewhat smaller than in
H3N...HF, a little less than 400 c m - 1 ; the intensification of this band is about 5 in the for-
mer complex, as compared to 7 in the latter. Not many changes are observed in the fre-
quencies of the proton acceptor water. The intermolecular frequencies are similar to
H3N...HF; the H-bond stretch is around 200-300 c m - 1 , and the wags of the donor mole-
cule approach 1000 c m - 1 . Changing one atom from first to second-row type seems to have
similar effects, regardless of the nature of the atom chosen. The overall effect is a weaken-
ing of the H-bond, observed here in the spectral data, confirming the energetic information
in the previous chapter. For example, the stretching frequency of the H2O...HC1 complex is
comparable to that of H3P...HF. This weakening continues to progress as F is changed to
Cl, then to Br and I. The red shift of the HX band drops by severalfold, as does the inter-
molecular H-bond stretch frequency. Alkyl substitution on the water changes the spectrum
by only a little. There is a curious increase in the red shift of the HC1 stretch in
(CH3)2O...HC1, however, as compared to H2O...HC1. Stronger H-bonds also have enhanced
effects upon the electric field gradient; changes of the order of 10% are calculated for com-
plexes like H2O...HC1.
Despite significant mixing with other types of intermolecular motion in the water dimer,
one can recognize a H-bond stretch as one normal mode, as well as bends of the donor and
acceptor molecule. The force constant for the OH stretch in the proton donor molecule
obeys trends quite similar to the H-bond energy itself. That is, k is smallest for the equilib-
rium geometry and rises only slowly as the proton acceptor molecule is disoriented, but
more quickly for rotations of the donor. By one analysis, the drop in frequency in the pro-
ton donor stretching frequency is a direct consequence of the bond's stretch in the H-bond,
with r and vs being nearly linearly related. This longer bond is more polar and polariz-
able, enabling it to form a stronger H-bond. Indeed explicit computations of polarizability
support the notion that the formation of the H-bond induces a noticeable change in the po-
larizability in the H-bond direction.
When HOH is the proton donor molecule, the vs band is not as pure a single X H bond
stretch as for HX, since the normal modes of the HOH molecule contain a symmetric and
200 Hydrogen Bonding

antisymmetric stretch, each of which involve both H atoms. It is the symmetric stretch
which corresponds most closely to a vs band in the dimer in that its frequency shifts more
to the red and its intensity is increased by a greater amount. The latter intensification is com-
puted to be an order of magnitude, but it must be understood that some fraction of this in-
crease is due to the change in the character of the vibrational motion itself, as opposed to
perturbations of the electronic structure. The bending mode in the donor molecule is shifted
toward the blue and suffers a small diminution in intensity. The v frequency of the water
dimer is computed to be some 175 c m - 1 , similar to that in (HF)2, just as their H-bonds are
close in energy. The mode corresponding roughly to a proton donor wag is about 600 c m - 1 ,
also similar to the HF dimer.
In concurrence with the weak bonding between HOH and NH3, the red shift of the vs
band is rather small, only about 170 c m - 1 , at the correlated level, but the magnification of
its intensity is surprisingly large, nearly 90-fold. The V frequency reflects the weakness of
this H-bond, barely over 100 c m - 1 , although a value of double this amount is obtained from
experimental measurement in a matrix. The interaction between a pair of NH3 molecules is
even weaker. The vibrational spectrum of the linear arrangement of the H-bond (not a true
minimum on the surface) does not lend itself readily to clear identification of intermolecu-
lar bands as bends, stretches, and so on. The frequency corresponding most closely to a H-
bond stretching motion is barely over 100 c m - 1 . The largest red shift in the proton donor
is less than 50 cm-1 in one of its bending modes.
A comprehensive comparison of various basis sets for the homodimers of HF and H2O
offers hope that calculation of vibrational frequencies can be meaningful, even when re-
stricted to the SCF level and with no account of anharmonicity. The frequencies are less de-
manding of basis set quality than are the intensities. Minimal basis sets are to be avoided in
most cases, as are small split-valence sets such as 3-21G. In some cases, one can compute
reasonable estimates of dimerization-induced frequency shifts with basis sets of 4-31G
type; however, results with unpolarized basis sets can be deceptive. Polarization functions
are strongly recommended for uniform quality of results, particularly if one is interested
primarily in spectral changes induced by H-bond formation. Intensity calculations without
polarization functions can be expected to yield only the crudest of estimates. Reasonable
results can be achieved with only one set of such functions on each atom. In some cases, it
may be useful to include diffuse "+" functions as well.
When engaged in a H-bond as proton acceptor, the carbonyl C=O bond is weakened
somewhat, based upon a lowered stretching frequency. This red shift appears to be linearly
correlated with the H-bond energy. The relationship between frequency shift and energy is
linear also for the proton donor molecule, namely the O—H stretch of water. When water
is paired with H2CO, the vibrational spectrum of the donor (water) is very much like that
in the water dimer, as is the H-bond stretching frequency, v . The intensities are different,
probably due in large part to the different nature of the modes themselves in the two com-
plexes. Along this line, changes induced in the spectrum of other proton-donor molecules
like HC1 are similar for H2CO as compared to H2O. The C=N double bond in the imine
group is also shifted to the red when this group accepts a proton.
In contrast to some of the aforementioned cases, the stretching frequency of the C N
triple bond in the nitrile group shifts to the blue when this group accepts a proton in a H-
bond. This shift amounts to some 20-30 cm-1 when HCN is paired with HF, and the band
is intensified by a factor between two and three. Also shifted toward the blue is the bend-
ing frequency of the HCN molecule. The red shift in the HF stretch within the donor mol-
ecule is considerably smaller than when HF donates a proton to the more basic amines, as
Vibrational Spectra 201

is the magnification of the intensity of this mode. The H-bond stretching frequency, v , is
around 150 c m - 1 . Its purity as a simple stretching motion, which has little effect upon the
overall dipole moment of the complex, leads to a weak intensity, comparable to that in
H3N...HF. The effects of both correlation and anharmonicity are not yet entirely clear; dif-
ferent means of evaluating these factors lead to different conclusions. Nonetheless, the
methods concur that the C N stretch is blue-shifted on H-bond formation.
The triple bond in the nitrile group makes the C of HCN electronegative enough to act
as an effective proton donor. When complexed with a strong acceptor like NH3, the HC
stretching frequency is reduced by some 160 c m - 1 .Further evidence of a genuine H-bond
is the fivefold enhancement of its intensity. As opposed to the blue shift of the C N stretch
observed when the nitrile is a proton acceptor, this frequency is diminished when HCN do-
nates a proton, and the intensity magnified by a factor of seven. The intermolecular stretch-
ing frequency is about 140 c m - 1 , with a very low intensity, again due to the lack of any
bending character in this mode.
As the molecules are enlarged, it becomes progressively more difficult to identify any
single normal mode as the A—H stretch, vs. So it is that in the dimer of formamide this as-
signment is made to a stretching mode of the NH2 group of the donor molecule. The red
shift in this band incurred by H-bond formation is quite small, only about 30 c m - 1 .The
C=O stretch in the acceptor also drops by a similarly small amount. Similar impurity af-
fects the intermolecular modes; the vibration corresponding most closely to v is about 120
c m - 1 . Analysis of the force constants in the cyclic dimer is consistent with normal Lewis
structures: the C=O bond is stronger than C—N, and N—H is weaker still. Formation of
the H-bond strengthens C—N but weakens the other two bonds, again congruent with the
conventional picture of these bonds.

References
1. Swanton, D. J., Bacskay, G. B., and Hush, N.S., The infrared absorption intensities of the water
molecule: A quantum chemical study, J. Chem. Phys. 84, 5715-5727 (1986).
2. Botschwina, P., Rosmus, P., and Reinsch, E. A., Spectroscopic properties of the hydroxonium ion
calculated from SCEP CEPA wavefunctions, Chem. Phys. Lett. 102, 299-306 (1983).
3. Hess, B. A. J., Schaad, L. J., Carsky, P., and Zahradnik, R., Ab initio calculations ofvibrational
spectra and their use in the identification of unusual molecules, Chem. Rev. 86, 709-730 (1986).
4. Wilson, E. B. Jr., Decius, J. C., and Cross, P. C., Molecular Vibrations; Dover, New York, (1955).
5. Pople, J. A., Scott, A. P., Wong, M. W., and Radom, L., Scaling factors for obtaining fundamental
vibrational frequencies and zero-point energies from HF/6-31G* and MP2/6-31G* harmonic
frequencies, Isr. J. Chem. 33, 345-350 (1993).
6. Somasundram, K., Amos, R. D., and Handy, N. C., Ab initio calculation for properties of hy-
drogen bonded complexes H 3 N . . . HCN, HCN-HCN, HCN . . . HF, H 2 O . . . HF, Theor. Chim. Acta
69,491-503(1986).
7. Bacskay, G. B., Kerdraon, D. I., and Hush, N. S., Quantum chemical study of the HCl molecule
and its binary complexes with CO, C2H2, C2H4, PH3 ,H 2 S, HCN, H2O, and NH3: Hydrogen
bonding and its effect on the 35Cl nuclear quadrupole coupling constant, Chem. Phys. 144,
53-69 (1990).
8. Badger, R. M., and Bauer, S. H., Spectroscopic studies of the hydrogen bond. II. The shifts of the
O—H vibrational frequency in the formation of the hydrogen bond, J. Chem. Phys. 5, 839-851
(1939).
9. Rao, C. N. R., Dwivedi, P. C., Ratajczak, H., and Orville-Thomas, W. J, Relation between O—H
stretching frequency and hydrogen bond energy: Re-examination of the Badger-Bauer rule, .1.
Chem. Soc., Faraday Trans. 71, 955-966 (1975).
202 Hydrogen Bonding

10. Millen, D. J., and Mines, G. W., Hydrogen bonding in the gas phase, J. Chem. Soc., Faraday
Trans. 273, 369-377(1977).
11. Andrews, L., Fourier transform infrared spectra of HF complexes in solid argon, J. Phys. Chem.
88, 2940-2949 (1984).
12. Yamaguchi, Y., Frisch, M., Gaw, J., Schaefer, H. F., and Binkley, J. S., Analytic evaluation and
basis set dependence of intensities of infrared spectra, J. Chem. Phys. 84, 2262-2278 (1986).
13. Stanton, J. F., Lipscomb, W. N., Magers, D. H., and Bartlett, R. J., Correlated studies of infrared
intensities, J. Chem. Phys. 90, 3241-3249 (1989).
14. Besler, B. H., Scuseria, G. E., Scheiner, A. C., and Schaefer, H. F., A systematic theoretical study
of harmonic vibrational frequencies: The single and double excitation coupled cluster (CCSD)
method, J. Chem. Phys. 89, 360-366 (1988).
15. Alberts, I. L., and Handy, N. C., M ller-Plesset third order calculations with large basis sets, J.
Chem. Phys. 89, 2107-2115 (1988).
16. Simandiras, E. D., Rice, J. E., Lee, T. J., Amos, R. D., and Handy, N. C., On the necessity off
basis functions for bending frequencies, J. Chem. Phys. 88, 3187-3195 (1988).
17. Miller, M. D., Jensen, F., Chapman, O. L., and Houk, K. N., Influence of basis sets and electron
correlation on theoretically predicted infrared intensities, J. Phys. Chem. 93,4478-4502 (1989).
18. Handy, N. C., Gaw, J. F., and Simandiras, E. D., Accurate ab initio prediction of molecular
geometries and spectroscopic constants, using SCF and MP2 energy derivatives, J. Chem. Soc.,
Faraday Trans. 2 83, 1577-1593 (1987).
19. Michalska, D., Hess, B. A., Jr., and Schaad, L. J., The effect of correlation energy (MP2) on com-
puted vibrational frequencies, Int. J. Quantum Chem. 29, 1127-1137 (1986).
20. Gaw, J.F., Yamaguchi, Y., Vincent, M. A., and Schaefer, H. F., Vibrational frequency shifts in
hydrogen-bonded systems: The hydrogen fluoride dimer and trimer, J. Am. Chem. Soc. 106,
3133-3138(1984).
21. Kurnig, I. J., Szczesniak, M. M., and Scheiner, S., Vibrational frequencies and intensities of H-
bondedsystems. 1:1 and 1:2 complexes of NH3 and PH3 with HF, J. Chem. Phys. 87,2214-2224
(1987).
22. Frisch, M. J., Pople, J. A., and Del Bene, J. E., Molecular orbital study of the dimers (AHn)2
formed from NH3, OH2, FH, PH3 ,SH2, and CIH, J. Phys. Chem. 89, 3664-3669 (1985).
23. Latajka, Z., and Scheiner, S., Structure, energetics and vibrational spectra of H-bonded systems.
Dimers and trimers of HF and HCl, Chem. Phys. 122,413-430(1988).
24. Michael, D. W., Dykstra, C. E., and Lisy, J. M., Changes in the electronic structure and vibra-
tional potential of hydrogen fluoride upon dimerization: A well-correlated (HF)2 potential en-
ergy surface, J. Chem. Phys. 81, 5998-6006 (1984).
25. Frisch, M. 5., Del Bene, J. E., Binkley, J. S., and Schaefer, H. F., Extensive theoretical studies of
the hydrogen-bonded complexes (H2O)2, (H2O)2H+, (HF)2, (HF)2H+, F 2 H - , and (NH 3 ) 2 J.
Chem. Phys. 84, 2279-2289 (1986).
26. Collins, C. L., Morihashi, K., Yamaguchi, Y., and Schaefer, H. F., Vibrational frequencies of the
HF dimer from the coupled cluster method including all single and double excitations plus per-
turbative connected triple excitations, J. Chem. Phys. 103, 6051-6056 (1995).
27. Dinur, U., Bond contraction and spectral blue-shift in hydrogen-bonded dimers. An atom-based
molecular mechanics analysis, Chem. Phys. Lett. 192, 399-406 (1992).
28. Curtiss, L. A., and Pople, J. A., Ab initio calculation of the force field of the hydrogen fluoride
dimer, J. Mol. Spectrosc. 61, 1-10 (1976).
29. Bunker, P. R., Jensen, P., Karpfen, A., Kofranek, M., and Lischka, H., An ab initio calculation
of the stretching energies for the HF dimer, J. Chem. Phys. 92, 7432-7440 (1990).
30. Jensen, P., Bunker, P. R., Karpfen, A., Kofranek, M., and Lischka, H., An ab initio calculation
of the intermolecular stretching spectra for the HF dimer and its D-substituted species, J. Chem.
Phys. 93, 6266-6280 (1990).
31. Pine, A. S., Lafferty, W. J., and Howard, B. J., Vibrational predissociation, tunneling, and rota-
tional saturation in the HF and DF dimers, J. Chem. Phys. 81, 2939-2950 (1984).
Vibrational Spectra 203

32. Karpfen, A., Bunker, P. R., and Jensen, P., An ab initio study of the hydrogen chloride dimer: the
potential energy surface and the characterization of the stationary points, Chem. Phys. 149,
299-309 (1991).
33. Hannachi, Y., and Angyan, J. G., The role of induction forces in infra-red matrix shifts: Quan-
tum chemical calculations with reaction field model Hamiltonian, J. Mol. Struct. (Theochem)
232,97-110(1991).
34. Tapia, O., and Goscinski, O., Self-consistent reaction field theory of solvent effects, Mol. Phys.
29, 1653-1661 (1975).
35. Person, W. B., and Zerbi, G., ed., Vibrational intensities in infrared and Raman spectroscopy;
Elsevier, Amsterdam (1982).
36. Zilles, B. A., and Person, W. B., Interpretation of infrared intensity changes on molecular com-
plex formation. I. Water dimer, J. Chem. Phys. 79, 65-77 (1983).
37. Gussoni, M., Castiglioni, C., and Zerbi, G., Charge distribution for infrared intensities: Charges
on hydrogen atoms and hydrogen bond, J. Chem. Phys. 80, 1377-1381 (1984).
38. Schemer, S., Proton transfers in hydrogen bonded systems, 6. Electronic redistributions in
(N2H7)+ and(O2H5)+, J. Chem. Phys. 75, 5791-5801 (1981).
39. Janoschek, R., Weidemann, E. G., Pfeiffer, H., and Zundel, G., Extremely high polarizability of
hydrogen bonds, J. Am. Chem. Soc. 94, 2387-2396 (1972).
40. Swanton, D. J., Bacskay, G. B., and Hush, N. S., An ab initio SCF calculation of the dipole-mo-
ment derivatives and infrared-absorption intensities of the water-dimer molecule, Chem. Phys.
82,303-315(1983).
41. Bouteiller, Y., Mijoule, C., Karpfen, A., Lischka, H., and Schuster, P., Theoretical Vibrational in-
vestigation of hydrogen-bonded complexes: Application to CIH'-NH^, CIH"NH2CH3, and
BrH-NH3, J. Phys. Chem. 91, 4464-4466 (1987).
42. Bouteiller, Y., Latajka, Z., Ratajczak, H., and Scheiner, S., Theoretical vibrational study of
FX-NH3 (X=H,D,Li) complexes, J. Chem. Phys. 94, 2956-2960 (1991).
43. Del Bene, J. E., Person, W. B., and Szczepaniak, K., Ab initio theoretical and matrix isolation
experimental studies of hydrogen bonding: Vibrational consequences of proton position in 1:1
complexes ofHCl and 4-X-pyridines, Chem. Phys. Lett. 247, 89-94 (1995).
44. Amos, R. D., Gaw, J. E, Handy, N. C., Simandiras, E. D., and Somasundram, K., Hydrogen-
bonded complexes involving HF and HCl: the effects of electron correlation and anharmonic-
ity, Theor. Chim. Acta 71, 41-57 (1987).
45. Latajka, Z., and Scheiner, S., Structure, energetics and vibrational spectrum of H2O—HCl, J.
Chem. Phys. 87, 5928-5936 (1987).
46. Hannachi, Y., Silvi, B., and Bouteiller, Y., Structure and vibrational properties of water hydro-
gen halide complexes, J. Chem. Phys. 94, 2915-2922 (1991).
47. Millen, D. J., and Schrems, O., Comparative infrared study of hydrogen-bonded heterodimers
formed by HCl, DCl, HF and DF with (CH3)2O, CH3OH, and (CH3)3COH in the gas phase. As-
signment of vibrational band structure in (CH3)2O-HCl, Chem. Phys. Lett. 101,320-325 (1983).
48. Barnes, A. J., and Wright, M. P., Molecular complexes of hydrogen halides with ethers and sul-
phides studied by matrix isolation vibrational spectroscopy, J. Mol. Struct. (Theochem) 135,
21-30 (1986).
49. Curtiss, L. A., and Pople, J. A., Ab initio calculation of the vibrational force field of the water
dimer, J. Mol. Spectrosc. 55, 1-14 (1975).
50. van Duijneveldt-van de Rijdt, J. G. C. M., van Duijneveldt, F. B., Kanters, J. A., and Williams,
D. R., Calculations on vibrational properties of'H-bonded OH groups, as a function ofH-bond
geometry, J. Mol. Struct. (Theochem) 109, 351-366 (1984).
51. Swanton, D. J., Bacskay, G. B., and Hush, N. S., An ab initio SCF calculation of the polariz-
ability tensor, polarizability derivatives and Raman scattering activities of the water-dimer mol-
ecule, Chem. Phys. 83, 69-75 (1984).
52. Amos, R. D., Structures, harmonic frequencies and infrared intensities of the dimers ofHJO and
H2S, Chem. Phys. 104, 145-151 (1986).
204 Hydrogen Bonding

53. Woodbridge, E. L., Tso, T.-L., McGrath, M. P., Hehre, W. J., and Lee, E. K. C, Infrared spectra
of matrix-isolated monomeric and dimeric hydrogen sulfide in solid O2, J. Chem. Phys. 85,
6991-6994(1986).
54. Ventura, O. N., Irving, K., and Latajka, Z., On the dlmerization shift of the OH-stretching fun-
damentals of the water dimer, Chem. Phys. Lett. 217, 436 (1994).
55. Kim, J., Lee, J. Y., Lee, S., Mhin, B. J., and Kim, K. S., Harmonic vibrational frequencies of the
water monomer and dimer: Comparison of various levels ofab initio theory, J. Chem. Phys. 102,
310-317(1995).
56. Huisken, F., Kaloudis, M., and Kulcke, A., Infrared spectroscopy of small size-selected water
clusters, J. Chem. Phys. 104, 17-25 (1996).
57. Bakkas, N., Bouteiller, Y., Loutellier, A., Perehard, J. P., and Racine, S., The water-methanol
complexes. L A matrix isolation study and an ab initio calculation on the 1—1 species, J. Chem.
Phys. 99, 3335-3342 (1993).
58. Wuelfert, S., Herren, D., and Leutwyler, S., Supersonic jet CARS spectra of small water clus-
ters, J. Chem. Phys. 86, 3751-3753 (1987).
59. Page, R. H., Frey, J. G., Chen, Y.-R., and Lee, Y. T., Infrared predissociation spectra of water
dimer in a supersonic molecular beam, Chem. Phys. Lett. 106, 373-376 (1984).
60. Nelander, B., The intramolecular fundamentals of the water dimer, J. Chem. Phys. 88,
5254-5256 (1988).
61. Huisken, F., Kulcke, A., Laush, C., and Lisy, .1. M., Dissociation of small methanol clusters af-
ter excitation of the O-~H stretch vibration at 2.7 u, J. Chem. Phys. 95, 3924-3929 (1991).
62. Bleiber, A., and Sauer, J., The vibrational frequency of the donor OH group in the H-bonded
dimers of water, methanol and silanol. Ab initio calculations including anhannonicities, Chem.
Phys. Lett. 238, 243-252 (1995).
63. Schmitt, M., M ller, H., Henrichs, U., Gerhards, M., Perl, W., Deusen, C., and Kleinermanns,
K., Structure and vibrations of phenol'CH ^OH (CD3OD) in the electronic ground and excited
state, revealed by spectral hole burning and dispersed fluorescence spectroscopy, J. Chem. Phys.
103,584-594(1995).
64. Dibble, T. S., and Francisco, J. S., Ab initio study of the structure, binding energy, and vibrations
of the HOC1-H20 complex, J. Phys. Chem. 99, 1919-1922 (1995).
65. Johnson, K., Engdahl, A., Ouis, P., and Nelander, B., A matrix isolation study of the water com-
plexes ofCl2, CIOCI, OCIO, and HOC! and their photochemistry, J. Phys. Chem. 96,5778-5783
(1992).
66. Rohlfing, C. M., Allen, L. C., and Ditchfield, R., Proton chemical shift tensors in neutral and
ionic hydrogen bonds, Chem. Phys. Lett. 86, 380-383 (1982).
67. Rohlfing, C. M., Allen, L. C., and Ditchfield, R., Proton chemical shift tensors in hydrogen-
bonded dimers ofRCOOH and ROM, J. Chem. Phys. 79, 4958-4966 (1983).
68. Kaliaperumal, R., Sears, R. E. J., Ni, Q. W., and Furst, J. E., Proton chemical shifts in some hy-
drogen bonded solids and a correlation with bond lengths, J. Chem. Phys. 91,7387-7391 (1989).
69. Chesnut, D. B., and Phung, C. G., Functional counterpoise corrections for the NMR chemical
shift in a model dimeric water system, Chem. Phys. 147, 91-97 (1990).
70. Latajka, Z., Ratajczak, H., and Person, W. B., On the reliability ofSCF ab initio calculations of
vibrational frequencies and intensities of hydrogen-bonded systems, J. Mol. Struct. (Theochem)
194,89-105(1989).
71. Sellers, H., and Almlof, J., On the accuracy in ab initio force constant calculations with respect
to basis set, J. Phys. Chem. 93, 5136-5139 (1989).
72. Latajka, Z., and Scheiner, S., Structure, energetics, and vibrational spectrum of'H 3 N . . HOH, J.
Phys. Chem. 94, 217-221 (1990).
73. Yeo, G. A., and Ford, T. A.. Ab initio molecular orbital calculations of the infrared spectra of
hydrogen bonded complexes of water, ammonia, and hydroxylamine. Part 6. The infrared spec-
trum of the water-ammonia complex, Can. J. Chem. 69, 632-637 (1991).
74. Engdahl, A., and Nelander, B., The intramolecular vibrations of the ammonia water complex. A
matrix isolation study, .1. Chem. Phys. 91, 6604-6612 (1989).
Vibrational Spectra 205

75. Nelander, B., and Nord, L., Complex bet\veen water and ammonia, J. Phys. Chem. 86,
4375-4379 (1982).
76. Schieflce, A., Deusen, C., Jacoby, C., Gerhards, M., Schmitt, M., Kleinermanns, K., and Hering, P.,
Structure and vibrations of the phenol-ammonia cluster, J. Chem. Phys. 102, 9197-9204 (1995).
77. Sadlej, J., and Lapinski, L., Ab initio calculations of the vibrational force field and IR intensi-
ties of the ammonia dimers, J. Mol. Struct. (Theochem) 150, 223-233 (1987).
78. Yeo, G. A., and Ford, T. A., Ab initio molecular orbital calculations of the infrared spectra of
hydrogen bonded complexes of water, ammonia, and hydroxylamine, J. Mol. Struct. (Theochem)
168,247-264(1988).
79. Yeo, G. A., and Ford, T. A., The combined use ofab initio molecular orbital theory and matrix
isolation infrared spectroscopy in the study of molecular interactions, Struct. Chem. 3, 75-93
(1992).
80. Latajka, Z., and Scheiner, S., Correlation between interaction energy and shift of the carbonyl
stretching frequency, Chem. Phys. Lett. 174, 179-184 (1990).
81. Thijs, R., and Zeegers-Huyskens, T., Infrared and Raman studies of hydrogen bonded complexes
involving acetone, acetophenone and benzophenone I. Thermodynamic constants and frequency
shifts of the vQHandvc=o stretching vibrations, Spectrochim. Acta A 40, 307-313 (1984).
82. Nuzzo, R. G., Dubois, L. H., and Allara, D. L., Fundamental studies of microscopic -wetting on
organic surfaces. L Formation and structural characterization of a self-consistent series of poly-
functional organic monolayers, J. Am. Chem. Soc. 112, 558-569 (1990).
83. Gould, I. R., and Hillier, I. H., The relation bet\veen hydrogen-bond strengths and vibrational
frequency shifts: A theoretical study of complexes of oxygen and nitrogen proton acceptors and
water, J. Mol. Struct. (Theochem) 314, 1-8 (1994).
84. Ramelot, T. A., Hu, C.-H., Fowler, J. E., DeLeeuw, B. J., and Schaefer, H. R, Carbonyl-water
hydrogen bonding: The H2CO~H2O prototype, J. Chem. Phys. 100, 4347-4354 (1994).
85. Rice, J. E., Lee, T. J., and Handy, N. C., The analytic gradient for the coupled pair functional
method: Formula and application for HCl, H2CO and the dirtier H2CO'"HCl, J. Chem. Phys.
88,7011-7023(1988).
86. Bach, S. B. H., and Ault, B. S., Infrared matrix isolation study of the hydrogen-bonded com-
plexes between formaldehyde and the hydrogen halides and cyanide, J. Phys. Chem. 88,
3600-3604 (1984).
87. Nowek, A., and Leszczynski, J., Ab initio study on the stability and properties ofXYCO—HZ
complexes. III. A comparative study of basis set and electron correlation effects for H2CO- • • HCl,
J. Chem. Phys. 104, 1441-1451 (1996).
88. Nowek, A., and Leszczynski, J., Ab initio investigation on stability and properties ofXYCO'"HZ
complexes. H: Post Hartree-Fock studies on H2CQ-HF, Struct. Chem. 6, 255-259 (1995).
89. Migchels, P., Zeegers-Huyskens, T., and Peeters, D., Fourier transform infrared and theoretical
studies of alkylimines complexes with hydroxylic proton donors, J. Phys. Chem. 95, 7599-7604
(1991).
90. Curtiss, L. A., and Pople, J. A., Molecular orbital calculation of some vibrational properties of
the complex between HCNandHF, J. Mol. Spectrosc. 48, 413-426 (1973).
91. Botschwina, P. In Structure and Dynamics of Weakly Bound Molecular Complexes; Weber, A.,
ed. D. Reidel (1987) pp 181-190.
92. De Almeida, W. B., Craw, J. S., and Hinchliffe, A., Ab initio vibrational spectra of'HCN-HF arid
HF"HCN hydrogen-bonded dimers: Mechanical and electrical anharmonicities, J. Mol. Struct.
(Theochem) 200, 19-31 (1989).
93. Bouteiller, Y, and Behrouz, H., Basis set superposition error effects on electronic and VFX, VF..N
stretching modes of hydrogen-bonded systems FX . . . NCX (X=H,D), J. Chem. Phys. 96,
6033-6038 (1992).
94. Bouteiller, Y., Basis set superposition error effects on vFx, VFX ,.N stretching modes of hydrogen-
bonded systems FX-NCH (X=H,D), Chem. Phys. Lett. 198,491-497(1992).
95. Del Bene, J. E.. Mettee, H. D., and Shavitt, I., Structure, binding energy, and vibrational fre-
quencies ofCH 3 CN . . . HCl, L Phys. Chem. 95, 5387-5388 (1991).
206 Hydrogen Bonding

96. Ballard, L., and Henderson, G., Hydrogen bond energy of CH3CN—HCl by FTIR photometry,
J. Phys. Chem. 95, 660-663 (1991).
97. Chattopadhyay, S., and Plummer, P. L. M., Ab initio studies of the mixed heterodimers of am-
monia and hydrogen cyanide, Chem. Phys. 782, 39—51 (1994).
98. Kofranek, M., Lischka, H., and Karpfen, A., Ab initio studies on structure, vibrational spectra
and infrared intensities ofHCN, (HCN)2, and (HCN)r Mol. Phys. 61, 1519-1539 (1987).
99. Hopkins, G. A., Maroncelli, M., Nibler, J. W., and Dyke, T. R., Coherent Raman spectroscopy
of HCNcomplexes, Chem. Phys. Lett. 114, 97-102 (1985).
100. Pacansky, H., The infrared spectrum of a molecular aggregate. The HCN dimer isolated in an
argon matrix, J. Phys. Chem. 81, 2240-2243 (1977).
101. Wojcik, M. J., Hirakawa, A. Y., Tsuboi, M., Kato, S., and Morokuma, K., Ab initio MO calcula-
tion of force constants and dipole derivatives for the formamide dimer. An estimation of hydro-
gen-bond force constants, Chem. Phys. Lett. 100, 523-528 (1983).
102. stergard, N., Christiansen, P. L., and Nielsen, O. F., Ab initio investigation of vibrations in free
and hydrogen bonded formamide, J. Mol. Struct. (Theochem) 235, 423-446 (1991).
103. Florian, J., and Johnson, B. G., Structure, energetics, and force fields of the cyclic formamide
dimer: MP2, Hartree-Fock, and density functional study, J. Phys. Chem. 99, 5899-5908 (1995).
104. Jucks, K. W., and Miller, R. E., The intermolecular bending vibrations of the hydrogen cyanide
dimer, Chem. Phys. Lett. 147, 137-141 (1988).
105. Jucks, K. W., and Miller, R. E., Infrared spectroscopy of the hydrogen cyanide dimer, J. Chem.
Phys. 88, 6059-6067 (1988).
4

Extended Regions of Potential


Energy Surface

t is a common perception that when a molecule with an available proton is placed in the
I vicinity of another molecule that has a lone electron pair, the two will approach one an-
other and quickly adopt their most stable, equilibrium H-bonded geometry. While this may
be true in certain instances, the potential energy surface of a typical H-bonded complex is
surprisingly complex. It is not unusual to find more than one minimum; the secondary min-
ima can be only slightly higher in energy than the global minimum. The paths connecting
the various minima can be rather complex as well, as the surface is littered with stationary
points of second, third, and higher order, all bunched within a few kcal/mol of one another.
Whereas the previous chapters have focused their attention on the equilibrium geome-
tries of H-bonded complexes, and their immediate surroundings, this chapter examines
broader reaches of the surface. Of particular interest are the paths along the surface that con-
vert one minimum to another. The ammonia dimer furnishes an example of an extremely
weakly bound complex, wherein the presence of a true H-bond is questionable. Its surface
is extremely flat, furnishing a particularly stringent test of quantum chemistry to identify
the true global minimum. Complexes pairing water with a hydrogen halide molecule offer
the opportunity to compare H-bonding with other sorts of interaction. For example, the
H2O...HX geometry contains a H-bond, but H 2 O ... XH does not. Even in the absence of a
hydrogen bonding sort of interaction, the latter can be a true minimum on the surface of cer-
tain of these complexes.
The simplicity of the HX dimer permits an especially thorough search of its PES. One
can consider configurations that provide interesting contrasts to the traditional H-bond. The
geometry of the equilibrium structure contains a nearly linear H-bond, with the remaining
hydrogen nearly perpendicular to this axis. The pathway over the surface for interconver-
sion of the roles of proton donor and acceptor represents an especially interesting problem,
with spectroscopic manifestations. The dimer of water is intriguing in that it is not obvious
from first principles that the linear H-bonded structure must be the only minimum on the
surface, or indeed the global minimum. Other candidates include cyclic, bifurcated, trifur-

207
208 Hydrogen Bonding

cated, and stacked geometries. By dissecting the total interaction energy of the water dimer
into its constituents, it is possible to glean some insights into the underlying reasons for the
shape of its PES.
The larger groups add to the complexity of the potential energy surface and to the num-
ber of potential minima and stationary points. The pairing of H2O with H2CO is a case in
point. The strong H-bond resulting from the pairing of an amine with HX brings up the pos-
sibility that a proton could transfer from the acid to the base so as to yield an ion pair, as an
additional minimum on the PES.

4.1 Ammonia Dimer

The earlier discussion of the most stable geometry for the ammonia dimer focused on the
linear and cyclic arrangements. In one of the first considerations of extensive regions of
the potential energy surface, Latajka and Scheiner1 varied the two angles which describe
the orientations of the two molecules, as well as the internitrogen distance. The cyclic struc-
ture was found to be the only true minimum on the surface, but a very shallow trough leads
from this geometry to a linear H-bond. The conversion from cyclic to linear stretches R(NN)
from 3.15 to 3.34 A. The energy rises relatively rapidly if one climbs the walls on either
side of this valley. There is hence a definite direction to the high-amplitude vibrational oscil-
lations of the ammonia dimer. On the other hand, the complex can be easily pulled apart at
any point along the valley; a stretching force constant of less than 0.12 mdyn/A was com-
puted. Torsional motions of one of the NH3 molecules about its C3 axis are essentially free
rotations in the vicinity of the linear arrangement but are stiffer for the cyclic geometry,
where such a torsion would break any H-bonding interactions present. The results also pro-
vided a warning against limited searches, where certain parameters are held fixed. For ex-
ample, the authors found that if the search is limited to only slices of the PES, each of which
has a fixed internitrogen distance, the linear geometry can appear to be a "minimum." The
two symmetrically related linear structures are then connected by a cyclic "transition state."
Tao and Klemperer2 also evaluated the path connecting the two equivalent linear Cs geome-
tries of the ammonia dimer, and passing through the cyclic minimum at its midpoint. Con-
firming the earlier calculations', they found this path to be remarkably flat, with the energy
varying by only several cm - 1 .
A number of studies investigated large domains of the surface by avoiding direct ab ini-
tio computations of the energy of each point. Liu et al.3 made use of an empirical potential
function based upon the electrical properties of the ammonia monomer. Unfortunately their
potential did not include the exchange which is necessary to prevent the collapse of the
dimer. Hence, their surface was constructed as a slice, with constant R(NN), through the com-
plete potential. Two degrees of freedom were considered, the angles that describe the devi-
ation of the C3 symmetry axes of the two NH3 molecules from the N..N axis. Under these
constraints, the authors identified 18 symmetry-related, equivalent minima on the surface,
due to the 120° periodicity of the molecular rotations. Conversion from one minimum to
the next tracks over an energy barrier of some 0.7 kcal/mol, emphasizing the flatness of
the surface. Sagarik et al.4 examined the surface of the ammonia dimer via an empirical
potential, fit to the results of correlated ab initio calculations. The analytical site-site po-
tential contained separate terms modeling exchange repulsion, electrostatic interactions,
and dispersion. The authors found a very shallow potential for pivoting one of the two NH3
molecules from the cyclic dimer geometry.
Extended Regions of Potential Energy Surface 209

While Hassett et al.5 did not consider the entire surface, they did characterize various
candidate structures as to whether or not they were true minima. Their calculations were
valuable since they involved full optimization of all degrees of freedom. They found that
the linear type of dimer, wherein the two nonbridging hydrogens of the proton-donating
molecule eclipse the pertinent hydrogens of the acceptor, appears to be a true minimum at
most levels of theory, varying from SCF/6-31 +G(d) to MP2/6-311 +G(2d',p). Rotation of
one molecule to form the staggered linear complex results in a single imaginary frequency
at most levels tested, as does the cyclic structure. A "trifurcated" complex wherein all three
hydrogens of one molecule are oriented in the general direction of the N of the other, is of
high energy and is a second-order stationary point. The same is true of a variant of the cyclic
complex where each molecule contributes two hydrogens, rather than one, to the region be-
tween the nitrogens. The results of this paper also furnished a good estimate of the zero-
point vibrational contributions to the binding energies of each structure.

4.2 H2O...HX

Hannachi et al.6 sampled the energy surface of a trio of complexes of the type H2O...HX,
where X=C1, Br, and I. The internal geometries of the two subunits were held fixed and
the surface generated in terms of the intermolecular distance R and the angle 6 which mea-
sures orientation of the HX molecule. Calculations were limited to the SCF level, using an
effective core, pseudo-potential approach, with the PS-31G** basis set. The generated maps
covered a range of about 1.5 A in R and 240° segments of 9. For H2O...HC1, the surface
contains a single minimum. Attempts to change the angle away from 0° can destabilize the
system by roughly 8 kcal/mol for a perpendicular arrangement, before the energy starts to
go down again. A second minimum appears in the surface at = 180° when X=Br.
H2O...HBr is more stable than H2O...BrH by 4 kcal/mol. Indeed, this secondary minimum
might disappear were zero-point vibrations added to the surface. The system must surmount
an energy barrier of about 6 kcal/mol to reach this second minimum; the pathway involves
a small amount of O...Br stretching along the way. The two minima are much better defined
in H2O...HI and are more competitive in stability, differing by only some 0.6 kcal/mol. The
energy barrier for conversion appears to be 4 kcal/mol.

4.3 (HX)2

In an early effort to consider tunneling of hydrogens in the HF dimer, Curtiss and Pople7
carried out an abbreviated scan of the SCF/4-31G surface. They varied the angle made by
one of the HF molecules with the P...F axis in 20° increments, optimizing the other para-
meters at each point. The conversion from one linear type of H-bond to its equivalent, in
which the proton donor and acceptor switch roles, was found to pass through a transition
state, characterized by a C2h cyclic geometry. Both HF molecules make angles of 55° with
the F...F axis in this structure, which is 1.1 kcal/mol higher in energy than the minima.
Michael et al. 8 considered a more extensive region of the surface in an effort to analyze
the vibrational motions of the two HF molecules. The energies of over 200 configurations
were computed with a polarized triple basis set; correlation was added to 73 of these
points. They noted that correlation was important to obtain the proper shape of the poten-
tial. The correlation energy changes smoothly with variation in the H—F bond length, but
the dependence is slightly different for the dimer than for the free monomer.
210 Hydrogen Bonding

Similar intentions motivated Redmon and Binkley9 who expanded the surface coverage
to over 1300 points, all at the MP4/6-311G** level. The results were then fit to an analytic
function. The authors computed the Restricted Hartree-Fock multipole moments of the HF
molecule as a function of internuclear distance, important information in fitting the elec-
trostatic part of the intermolecular potential. Fig. 4.1 illustrates that the dipole moment in-
creases linearly with bond length, rising from 1.6 D when r(HF) = 0.7 A to 2.68 D when
stretched to 1.27 A. The longitudinal and perpendicular components of the quadrupole mo-
ment also vary with r(HF), and nearly linearly. The stretching of the bond leads to a reduc-
tion in the magnitude of 6zz, but 0xx becomes larger.
The computed energies show that there is a sharp maximum in the potential for a head-
to-head configuration: F—H . . . H—F, where the two hydrogens are pointed at one another,
about 40 kcal/mol higher in energy than the minimum. The tail-to-tail H—F . . . F—H geom-
etry is also a maximum, but much less sharp, and only about 7 kcal/mol higher than the min-
imum. Torsional rotation of one molecule around the F...F axis from the planar geometry is
mildly destabilizing; the maximum in the rotational profile is the 180° rotation wherein the
two hydrogens are located on the same side of the F...F axis.

4.3.1 Anisotropies of Energy Components


Szczesniak and Chalasinski10 focused their attention on the correlation segment of the in-
termolecular interaction, and in particular on the angular dependence of the dispersion en-
ergy. The latter component is typically considered to be quite isotropic so the authors de-

Figure 4.1 Variation of Restricted Hartree-Fock dipole ( . in D)


and quadrupole moments (0 in D.A) of HF with bond length,
calculated by Redmon and Binkley9. The molecular axis is
taken as the z-axis.
Extended Regions of Potential Energy Surface 21 I

cided to test this assumption on the HF dimer. They learned that there is considerable an-
gular dependence of the dispersion. This component is smallest in the head-to-head con-
figuration where the two hydrogens are pointed at one another, but climbs by a factor of 14
when the two fluorines approach one another. Even more anisotropic than the dispersion
energy is es(12), which reflects the change in the electrostatic energy as a result of includ-
ing correlation. This term switches sign: it is repulsive for the head-to-tail linear configu-
ration and becomes attractive as the proton acceptor molecule rotates around toward the
F—H . . . H—F geometry.
Figure 4.2 illustrates the anisotropy of the energetics of this system as the proton ac-
ceptor molecule is rotated, keeping the distance between centers of mass fixed at 2.8 A. The
solid curve indicates a shallow minimum in the SCF potential for (3 in the vicinity of 120°.
As (3 diminishes, the two hydrogens are brought closer to one another and the SCF energy
quickly becomes repulsive, due to the electrostatic repulsion between the two molecular
dipoles as well as steric repulsions between the two hydrogen atoms. Adding in the disper-
sion energy stabilizes the system, as illustrated by the long-dashed curve in Fig. 4.2. On the
full scale of Fig. 4.2, the SCF and SCF+disp curves are nearly parallel, indicating that the
anisotropy of the SCF potential far exceeds that of the dispersion. The short-dashed curve
represents the anisotropy of the full potential computed at the MP2 level. This curve is
nearly coincident with the SCF potential in the region of large [5, around the minimum, sug-
gesting that the attractive dispersion is approximately canceled by other facets of the sec-
ond order correlation here, chiefly es(12). The SCF+MP2 curve falls below SCF for smaller

Figure 4.2 Anisotropy of SCF and correlated components of in-


teraction energy in (HF)2'°. refers to the angle between right-
hand molecule and F..F axis. Centers of mass of two molecules
are held fixed at 2.8 A.
212 Hydrogen Bonding

angles, indicative that the dispersion outweighs the latter electrostatic correction for the rel-
evant configurations. The anisotropy of the correlated SCF4+MP2 curve is apparently best
mimicked by the sum of three terms, adding the dispersion and es(12) to the SCF energy,
as recommended by the authors.

4.3.2 Interconversion Pathways


Bunker et al.n computed 1061 points on the correlated surface of the HF dimer using a po-
larized basis set, including all six degrees of freedom of the complex. They then fit their re-
sults to an analytic function containing 42 adjustable parameters. Improving upon the ear-
lier work7, the authors were able next to determine the lowest energy paths for conversion
from the optimal geometry to other structures. The motion on the left of Fig. 4.3 decreases
a while is increased, taking the dimer through a linear transition state (C ) on a pathway
that moves each of the two hydrogens to the other side of the F..F axis from which they
started. An alternate route rotates the two hydrogens in opposite directions so as to reverse
the roles of the two molecules as donor and acceptor, respectively. This route passes through
a cyclic C2h structure as the transition state.
Fig. 4.4a illustrates the energetics of motion along either pathway. Starting from the min-
imum energy configuration at ~120° in the center of the figure, the energy climbs as one
moves toward larger and the linear geometry (to the left), or toward smaller and the
cyclic structure to the right. The two transition states are nearly equal in energy at 345 and
332 cm-1 above the minimum, respectively (1 kcal/mol). Fig. 4.4b indicates the manner in
which the other angle a changes along the two low-energy paths. Again, starting from
~ 120° in the center, a quickly diminishes from 8° to 0° as the linear geometry is approached
or climbs quickly toward 56° in the cyclic structure. The behavior of the interfluorine dis-

Figure 4.3 Interconversion pathways in (HF)2. The path on the right


switches the roles of proton donor and acceptor molecules.
Extended Regions of Potential Energy Surface 21 3

Figure 4.4 Variation of (a) energy, (b) angle,


and (c) interfluorine distance (1 au = 0.529 A)
along the lowest energy path of (HF)2, all as a
function of angle (3. The equilibrium structure
occurs at approximately = 120° in the center
of each figure11.

tance is illustrated in Fig. 4.4c from which it may be seen that R is in excess of 2.86 A for
the linear geometry but less than 2.72 A for the cyclic. The transition is rather sharp: R rises
steeply as increases beyond 110°, reaching its full value when this angle is only 140°.
These authors also explored the potential for rotating one molecule or the other around the
F..F axis. The torsional potential shows a single minimum at the equilibrium geometry and
a barrier of about 0.4 kcal/mol.

4.3.3 HClDimer
Karpfen et al.12 turned their attention from (HF)2 to the analogous dimer of HC1 in 1991,
using an averaged coupled pair functional (ACPF) approach13 to include correlation; the
two basis sets were of [6,5,2/4,2] and [6,5,3,1/4,2] quality. In addition to the minimum il-
lustrated in Fig. 4.5a, several other stationary points were identified on the PES.
The cyclic geometry in Fig. 4.5b is not a minimum but rather a first-order stationary
point, representing the transition state for interconversion of the proton donors and accep-
tors, just as for (HF)2. The authors find this point lies only 0.2 kcal/mol higher in energy
214 Hydrogen Bonding

Figure 4.5 Energies (kcal/mol) computed for various stationary points


on the PES of the HC1 dimer12. The values displayed for E refer to the
electronic contribution Eclec.

than the minimum. Less stable, but also a first-order stationary point is the nearly linear
structure in Fig. 4.5c, also of C2h symmetry. This minimum is not present in the analogous
PES for (HF)2. The two linear structures in Figs. 4.5d and 4.5e also represent transition
states on the surface, barely bound with respect to a pair of HC1 molecules, with binding
energies of around 0.25 kcal/mol.
The authors generated contour surfaces as functions of a and (3, at a series of fixed val-
ues of R(C1..C1). They noted firstly that this surface is considerably flatter than that of the
(HF)2 congener. Assuming no change in this distance, the minimum-energy path for trans-
tunneling from the minimum energy structure 4.5a through either transition state 4.5b or
4.5c is a linear one, that is, the changes in the two angles are both proportional to the
progress along the path.
The problem was reexamined in 1995 using an MP treatment of correlation; the
[8s6p3d/6s3p] basis set was augmented by bond functions 14. The results largely confirm
those of Karpfen et al.12. The MP2 barrier to interchange occurs at a geometry akin to that
in Fig. 4.5b, with a = 45°. This configuration lies 0.17 kcal/mol higher in energy than the
minimum, as in the earlier work. The collinear arrangement in Fig. 4.5d was found to be
bound by some 0.4 kcal/mol, and 4.5e is similarly weakly bound, with Eelec = —0.3
kcal/mol. The authors provide evidence that the MP2 anisotropy of the energetics of the
complex matches quite well the MP4 results.
Another surface has been generated by a fitting of microwave and far and near-infrared
spectroscopic quantities 15 . This "experimental" surface confirms most of the substantive
Extended Regions of Potential Energy Surface 21 5

predictions of the higher quality calculations. The global minimum has the two centers of
mass separated by 3.746 A, slightly closer together than derived from some of the corre-
lated ab initio calculations, and with a H-bond energy stronger by almost 0.5 kcal/mol. The
barrier to donor-acceptor interchange is 0.14 kcal/mol on this surface, also in excellent
agreement with ab initio predictions. The pathway along this experimental surface involves
a contraction in the intermolecular separation at the conversion barrier by 0.1 A, consistent
with an earlier calculation16.

4.4 Water Dimer

As arguably the most important of all H-bonded complexes, the water dimer has stimulated
a great deal of theoretical study. The "evolution" of its potential energy surface over the
years, as theoretical methods have become progressively more sophisticated and reliable,
offers a fascinating case study. Of particular interest are the attempts to determine which
geometry is the global minimum on the surface, and whether other candidate geometries
represent local minima, transition states for conversion of one structure to another, or if they
are stationary points of some higher order.

4.4.1 Characterization of Possible Minima and Stationary Points


4.4.1.1 Linear, Bifurcated, and Cyclic
Despite early assumptions that the linear arrangement is the most stable conformation of
the H-bond in the water dimer, a number of early works considered the possibility of other
geometries as well. Most notable among the candidates were the cyclic and bifurcated struc-
tures. Beginning in the early 1970s, Diercksen17 was probably the first to discuss the issue
from the perspective of a moderately large basis set, containing polarization functions. This
approach yielded a dipole moment for the water monomer within about 20% of the exper-
imental value. His searches of the potential energy surface suffered from several limitations.
In the first place, the work predated gradient algorithms so the surface was obtained as a
series of single points, without the possibility of unambiguous verification of any structure
as a true minimum or even as a stationary point of any order. Other sources of error were
the failure to remove BSSE or to include electron correlation. With these caveats estab-
lished, Diercksen found the linear structure to be the probable minimum, with R(OO) ~3.00
A. The two molecules could rotate about the H-bond axis with very low energy barriers.
The calculations indicated the bifurcated structure was not a minimum, but would imme-
diately convert to the linear structure with no barrier.
Matsuoka et al.18 added electron correlation to hthe comparison of linear, bifurcated, and
cyclic structures. Like their predecessors, these authors did not attempt to correct the su-
perposition error, which is now understood to typically be considerably larger for correlated
contributions than for the SCF interaction. This study confirmed the earlier finding that the
linear geometry was most stable, bound by 5.6 kcal/mol. The binding energies of the (dis-
torted) cyclic and bifurcated structures were, respectively, 4.9 and 4.2 kcal/mol, signifi-
cantly weaker than the linear dimer. On the other hand, the latter two structures could not
be verified to be true minima on the surface. To be more specific, while certainly generat-
ing minima in certain slices through the full potential energy surface as a function of R(OO),
there was insufficient probing of various angular parameters to be able to properly classify
these structures.
216 Hydrogen Bonding

Some of the prior problems were addressed in 1985 by Baum and Finney19 who at-
tempted to correct the BSSE of their correlated calculations. The CI procedure they em-
ployed was not size-consistent so they added a correction for this error as well. Their basis
set was of [5 s4p 1 d/3 s 1 p] quality, and the CI of the type that includes all single and double
excitations (except those involving the 1s orbitals of O). While unable to employ gradient
procedures, the scan of the surface permitted the workers to guess as to whether or not they
had located a minimum on the surface, since angles were varied as well as intermolecular
distance. They found that the trans type of linear H-bond is the most stable conformer. A
cis linear arrangement is not a minimum, but can readily convert to trans. The bifurcated
structure is also likely to rearrange to trans linear, but with a very small barrier. The authors
believed that even uncorrected SCF computations with their basis set should be adequate
to characterize the basic features of the PES.
At approximately the same time, Frisch et al.20 introduced correlation through M011er-
Plesset theory. Although no account was taken of BSSE, their methods were able to intro-
duce gradient optimization techniques so they could readily distinguish minima from other
stationary points. The results confirmed that the linear structure is a true minimum. The bi-
furcated arrangement is a saddle point and a planar cyclic geometry is a stationary point of
second order. At the MP4SDQ/6-31+G**//SCF/6-31+G* level, the linear minimum is
more stable than the bifurcated and cyclic structures by 1.6 and 1.5 kcal/mol, respectively.
However, these differences are reduced to 0.9 and 0.7 when zero-point vibrational energies
are considered.
Singh and Kollman21 dissected the interaction energies of the three candidate geome-
tries using the Morokuma decomposition scheme, within the framework of a doubly polar-
ized basis set. The ESX term in Table 4.1 is a summation of electrostatic, exchange repul-
sion, and "mixing." The dispersion energy, DISP, was evaluated by Singh and Kollman
using a three-term expansion in 6-8-10 powers of 1/R22, and added to the SCF interaction
energy to obtain the total interaction energies reported in the last column of Table 4.1. At
this level of theory, the combined electrostatic and exchange term favors the cyclic over the
linear geometry by a small amount, as does the dispersion energy. However, these differ-
ences are more than compensated by larger preferences of the charge transfer and polar-
ization energies for the linear arrangement. The bifurcated geometry is less stable than the
other two in all categories considered.

4.4.1.2 Stacked
Hobza et al.23 considered a different sort of geometry altogether. Their "stacked" structure
places the planes of the two water molecules parallel to one another. The O atoms are di-
rectly above one another, and similarly for the two hydrogens in the parallel arrangement
in Fig. 4.6.a.

Table 4.1 Morokuma energy components (kcal/mol) of three arrangements of


the water dimer21.
ESXa CT POL DISP Total

linear -2.68 -1.28 -0.90 -0.32 -5.19


cyclic -2.91 -0.87 -0.47 -0.42 -4.67
bifurcated -1.96 -0.37 -0.27 -0.28 -2.87
a
Summation of" electrostatic, exchange repulsion and "mixing."
Extended Regions of Potential Energy Surface 217

Figure 4.6 Parallel and antiparallel types of stacking of pairs of water molecules.

This complex was studied by the investigators as a function of the distance between the
two planes, identical to R(OO), with a variety of basis sets, ranging from minimal STO-3G
to a doubly polarized, double- set. The interaction was found to be repulsive at the SCF
level with all sets considered, which the authors attributed to the parallel alignment of the
molecular dipole moments. There was not much sensitivity of this potential to the particu-
lars of the basis set, except that the STO-3G results were not as repulsive as the others. In
contrast to this behavior, the correlated part of the potential was rather sensitive to basis set
choice. The authors found that the attractiveness of the correlation contribution was directly
related to the size of the set. But in no case was the negative correlation term large enough
to compensate for the strong repulsive nature of the SCF interaction, so the parallel dimer
above is clearly not a minimum on the surface. The authors also investigated the antiparal-
lel geometry illustrated in Fig. 4.6b. Since the dipoles are now more favorably aligned, this
interaction can be expected to be attractive. Indeed, such an attractive interaction, albeit a
weak one, was observed at the SCF level.

4.4.1.3 Trifurcated
Spurred by semiempirical AM1 calculations that suggested the lowest energy structure of
the water dimer might not be of the linear variety at all, Dannenberg24 applied well-corre-
lated ab initio methods to test the validity of the AM1 findings. He made use of the stan-
dard 6-311G** basis set and went up to MP4SDQ but did not attempt to correct for BSSE
or zero-point vibrational energies. It is somewhat difficult to fully interpret the results as
some of the structures were only partially optimized and cannot be characterized as min-
ima, or indeed as stationary points of any order. Dannenberg was chiefly concerned with
the relative stability of the linear type of structure, as compared to what he terms a "trifur-
cated" H-bond which is basically a variant of the cyclic configuration except that one of the
water molecules contributes two hydrogens to the region between the oxygen atoms, as il-
lustrated in Fig. 4.7.
Beginning from the AM1 structure as a starting point, Dannenberg carried out a geom-
etry optimization at the MP2/6-31G* level to arrive at a structure like that in Fig. 4.7, with
O .. HO angle equal to 129° (although it was not entirely clear from the paper that this was
indeed a truly optimized structure). The binding energy of this dimer was calculated to be
5.5 kcal/mol at the MP4SDQ/6-311G** level of theory. A similar type of optimization led
218 Hydrogen Bonding

Figure 4.7 Trifurcated type of arrangement of the


water dimer.

to the standard linear H-bond, bound by 5.8 kcal/mol, only 0.3 more than the trifurcated
structure. The author pointed out that the experimental determination of the equilibrium
geometry and binding energy had been carried out at 350 400° K. At a temperature this
high, it is probable that entropy would play an important role. He argued that if the trifur-
cated structure were only 10 eu lower in entropy than the linear structure, this would be suf-
ficient to render the latter one dominant at high T. The lower entropy of the trifurcated struc-
ture could easily be rationalized on the basis of "freezing" the positions of three of the
hydrogen atoms in H-bonds of some type, as compared to the linear configuration which
only has a single bridging proton.

4.4.1.4 Definitive Characterization


Smith et al.25 added some particularly useful information to the debate with their high level
study of various regions of the potential energy surface. This work employed MP4/6-
311 + G(2df,2p) single-point calculations at geometries optimized at the MP2/6-
311 + G(d,p) level. Counterpoise corrections were computed, but not added directly to the
energetics, as there was no way to include such corrections directly into geometry minimiza-
tions or characterization of stationary points on the potential energy surface. These work-
ers identified a number of stationary points and characterized the transition states separat-
ing them. Their global minimum corresponds to the linear geometry, and is bound with
respect to two isolated water molecules by 3.48 kcal/mol at 373° K, after correction for
BSSE and vibrational and thermal energies, which the authors compared with the thermal
conductivity measurement of 3.59 ± 0.5 kcal/mol26.
The process which interchanges the two hydrogens on the proton acceptor molecule con-
sists of a rotation of that molecule, as indicated in Fig. 4.8a. The transition state for this in-
terchange has no symmetry elements and lies 0.6 kcal/mol higher in energy than the global
minimum on the potential energy surface. While the latter structure is nonplanar as indi-
cated, its planar correlate is only 0.1 kcal/mol higher in energy. The interchange of the donor
or acceptor roles of the two molecules requires slightly more energy, passing over a barrier
of 0.9 kcal/mol, as illustrated in Fig. 4.8b. The transition state geometry is cyclic in nature,
with O..OH angles of 112°. The two hydrogens not participating in the H-bond lie alter-
nately above and below the plane; placing them both in this plane raises the energy by 0.4
kcal/mol. It takes more energy to interchange the two protons in the donor molecule since
one of them participates directly in the H-bond. A bifurcated geometry, in which two pro-
tons participate simultaneously in a pair of bent H-bonds, serves as the transition state to
this switch, with a barrier of 1.9 kcal/mol (Fig. 4.8c).
Smith et al. also reinvestigated the question of the trifurcated structure and found it to
be a second-order saddle point, nearly 2 kcal/mol higher in energy than the minimum.
This work reinforced the notion that the potential energy surface is rather flat, not sur-
Extended Regions of Potential Energy Surface 219

Figure 4.8 Paths for interconversion of one linear dimer to a symmetrically


related one. Energies of transition states for each path are given in
kcal/mol.

prising in view of the weak nature of the H-bond itself. The surface is punctuated by a large
number of small undulations, as the molecules reorient themselves. The fluctuations be-
come steeper for motions that more nearly break the H-bond. While the single, nearly lin-
ear, H-bond is most stable and corresponds to the unique minimum on the surface, other
geometries which contain multiple bent H-bonding interactions can be rather close in en-
ergy, although they correspond to first or higher-order saddle points. The flatness of the po-
tential makes its characterization sensitive to the level of theory. Semiempirical methods
seem prone to mistake higher-energy saddle points for minima. Even ab initio results must
be viewed with some caution. For example, Smith et al. noted a number of occasions where
the order of a given saddle point was altered upon including correlation, even with a flexi-
ble basis set.
At about the same time, Vos et al.27 published another investigation of various con-
formers of the water dimer using a polarized double- basis set. MP2 correlation was com-
pared directly with the CEPA-1 approach, and the binding energies were found nearly iden-
tical, to within about 0.1 kcal/mol or better. The cyclic geometry was found less stable than
the linear minimum by about 1 kcal/mol. The authors attributed most of this difference to
the deformation energy at the SCF level (non-Heitler-London effects). The bifurcated struc-
ture is further destabilized with respect to this SCF deformation energy and also contains
less dispersion energy as a result of the greater R(OO) separation. On the other hand, the
bifurcated arrangement does have stronger coulombic attraction due to the better orienta-
tion of the two molecular dipole moments. As a result of these competing effects, the bi-
furcated geometry is about 0.75 kcal/mol less stable than the cyclic. Since their results in-
dicate that distortion energies from equilibrium are dominated by Heitler-London effects,
the authors were optimistic that SCF calculations can provide reasonable equilibrium
geometries for H-bonded complexes.
Further examination by Marsden et al.28 of the nature of the bifurcated structure applied
a basis set containing three sets of diffuse functions, along with a doubly polarized quadru-
ple-Jj valence set, amounting to 146 contracted Gaussians. These workers concluded that
220 Hydrogen Bonding

the bifurcated structure is certainly a transition state on the SCF surface with an imaginary
frequency in the neighborhood of 200i c m - 1 . The authors expect this value to increase with
correlation, strengthening their contention of a saddle point. The energy barrier between the
true minimum and this transition state is estimated as 1.3 kcal/mol, increasing to perhaps
1.7 kcal/mol after inclusion of correlation.

4.4.2 Components of the Interaction Energy


The shape of the potential energy surface in the general vicinity of the minimum of the wa-
ter dimer has been addressed by examining the individual contributors to the total interac-
tion energy. Singh and Kollman21 examined how the various components of a Morokuma
decomposition vary as the proton acceptor molecule of the linear dimer "wags" as a func-
tion of in Fig. 4.9. As described in Section 4.4.1. the authors combined electrostatic and
exchange into a single term, ESX, along with the "mixing" energy. There is little change in
ESX as varies between 180° and 90°, with a shallow minimum at approximately 120°. The
behavior of the total interaction energy is very nearly parallel, although lower in energy as
a result of other stabilizing factors. The charge transfer, polarization, and dispersion were
all computed to be attractive and of smaller magnitude, in the order indicated. All three be-
come monotonically more attractive as decreases from 180° to 0°. Similar calculations
were carried out for the cyclic structure. Again, the ESX term closely parallels the full in-
teraction energy, as a function of angular orientation. In this case, the minimum is located
at 50°. The charge transfer, polarization, and dispersion are all maximized in magnitude
for smaller angles.

4.4.2.1 Electrostatic Contribution


Cybulski and Scheiner29 also considered the factors that contribute to the distortion energy
that must be overcome when the H-bond in the water dimer is bent. Both the proton donor
and acceptor molecules were subjected to 40° distortions from their optimal alignment, and
the energetics monitored. The authors were able to draw a strong parallel between these dis-
tortion energies and the change in the electrostatic component of the interaction energy.
These two quantities are nearly identical when R(OO) = 3.25 A but less coincident for
R(OO) = 2.75 A, shorter than the equilibrium H-bond. In contrast, the other contributors
to the interaction, namely, exchange repulsion, polarization, charge transfer, and a catch-all
"MIX" term, are much smaller in magnitude and do not change very much with misorien-
tation. Indeed, the distortions considered brought the amounts of these contributions down
close to zero.
Because of its importance in the water dimer, as well as in a number of other H-bonded
systems, the electrostatic interaction was partitioned into a multipole series in powers of
1/R, consisting of terms corresponding to interactions between dipole, quadrupole, and so

Figure 4.9 Geometry of the water dimer equilib-


rium structure.
Extended Regions of Potential Energy Surface 221

on moments29. Truncation of this series at the R-5 term led to excellent reproduction of the
full electrostatic interaction, particularly for the longer intermolecular separation, even
though the last term was not necessarily small in magnitude. Near cancellation between the
R-4 and R-5 terms was observed, leaving R-3 as a good indicator of the full electrostatic
energy for most modes of distortion examined. It is thus possible for the simple dipole-di-
pole interaction, constituting the R-3 term, to predict quite well the behavior of the full en-
ergetics with respect to angular distortions of this H-bond.

4.4.2.2 Contributions from Electron Correlation


Szczesniak et al.30 considered the factors leading to the degree of linearity of the H-bond
in the water dimer and the pyramidalization of the proton acceptor oxygen. The dependence
of the Hartree-Fock interaction energy was calculated as a function of both a and (3 (see ear-
lier), as were the dispersion energy, disp (20) ,and second-order M011er-Plesset correlation
energy, EMP(2). It is primarily the SCF components that lead to the equilibrium geometry
of the dimer. But particularly interesting was a comparison of the behavior of disp (20) and
EMP(2) which includes the former. The dispersion energy favors a linear H-bond, and
amounts to nearly 2 kcal/mol in this geometry. A 60° distortion of a reduces the dispersion
attraction by about half. The more complete second-order correlation energy behaves in an
opposite fashion, with EMP(2) favoring a nonlinear H-bond. The two quantities behave
more similarly when the proton acceptor molecule is rotated. Both disp(20) and EMP(2)
tend to push (3 away from 180°, that is, they favor a pyramidal arrangement.
The authors were able to rationalize the behavior of the dispersion energy based on a
model where this attractive component is built up from the interactions between occupied
molecular orbitals on the two subunits. The two lowest MOs of the water molecule, la1 and
2a1, are not very polarizable and so contribute relatively little to the dispersion. The third
MO is largely of O—H bonding character, while the fourth and fifth are the oxygen lone
pairs. Fig. 4.10 illustrates schematically the angular dependence of several of the more im-
portant interorbital dispersion interactions. For example, the interaction between the O—H
bond (orbital 3) of the donor and the a lone pair of the acceptor (orbital 4) is strongest due
to the maximal overlap when a = 0°, as pictured in the figure. Rotation of the donor turns
the O—H bond away so this term rapidly becomes less negative. Similar behavior, although
less dramatic, is observed in the 3-5 interaction involving the lone pair of the acceptor.
The bilobal character of the a lone pair leads to a substantial 4-4 interaction at a = 0°. A
maximum occurs when the two orbitals are perpendicular to one another, as indicated when
a is approximately 50°, after which the 4-4 interaction again becomes more negative. This
sort of analysis lends hope that the anisotropy of dispersion can be predicted from first prin-
ciples, based on knowledge of the general MO structure and polarizabilities, in the same
way that electrostatics can be decomposed into interactions between multipole moments of
the monomers.
An alternate approach, and one which is more commonly taken, is to partition the dis-
persion into interactions between atoms on the two subunits. Probably the simplest exam-
ple assumes an inverse sixth power dependence upon interatomic separation.
222 Hydrogen Bonding

Figure 4.10 Angular dependence of various interorbital pairwise


contributors to dispersion energy in the water diraer30. (3 was held
fixed at 180°.

The CXY are parameters fit to each pair of atoms X and Y on molecules A and B, and rXY
is the distance separating these atoms. Szczesniak et al.30 fit these parameters by a least-
squares method so as to reproduce their calculated dispersion as closely as possible.
The reason for the discrepancy between the angular dependence of disp (20) and EMP(2)
can be traced by examining the distance-dependence of these two terms. At long distance
the former remains negative while the latter becomes repulsive. This repulsive character re-
sults from another term that appears in the second-order correlation. Since the multipole
moments of the water monomers are lower in the correlated wave function than SCF, the
attractive electrostatic interaction becomes weaker when correlation is included. Hence, the
correlation correction to the electrostatic interaction, es(12), shows up as a repulsive term.
Since electrostatics dies off more slowly than dispersion, it is this repulsive correlation cor-
rection that remains at long distance. The authors left their results as a warning against at-
tempts to simulate a true correlated intermolecular potential by supplementing the SCF in-
teraction by dispersion alone. Another point is that there is a definite anisotropy to
dispersion, as there is to other correlation components, that should not be ignored in con-
struction of empirical potentials to model the interaction.

4.4.2.3 Natural Bond Orbitals


Most recently, Glendening and Streitwicser31 have decomposed the interaction energy of
the water dimer using natural bond orbitals. Their natural energy decomposition analysis
(NEDA) combines the normal electrostatic and exchange energies into a single ES term,
Extended Regions of Potential Energy Surface 223

and the DEF term refers to the energy required to deform the wave functions of the isolated
subsystems to those they will assume within the complex. In the case of a fully linear H-
bond, with a = 0°, the total interaction energy has a minimum (or at least within several de-
grees of 0°), as do the ES and charge transfer terms; DEF is at its most positive here. Dis-
torting the H-bond by rotating the proton donor molecule causes both ES and CT to rise
rather sharply. These two terms each become less negative by about 5 kcal/mol when a =
60°. But the total interaction energy increases by only 2 or 3 kcal/mol because this bending
motion reduces the (electronic) deformation energy DEF by about 8 kcal/mol.

4.5 Carbonyl Group

The complexity of the PBS for the pair of H2CO with H2O offers a rich field of candidates
for minima and stationary points. Kumpf and Damewood32 explored an extensive region
of the potential energy surface of the H2CO...HOH complex using a polarized basis set of
the 6-31G** variety. Correlation was added, as well as zero-point vibrations, once station-
ary points were identified. Thirteen different configurations were considered.
Some of these had formaldehyde as proton acceptor and some tested its capability as
donor. Linear H-bonds were compared with bifurcated arrangements and with a sort of
cyclic geometry in which both molecules can act simultaneously as donor and acceptor. The
authors also checked on the ability of the bond of H2CO to accept a proton. Some of the
more important of the configurations examined by Kumpf and Damewood are illustrated
in Fig. 4.11 where the same nomenclature has been used as in their original article for pur-
poses of consistency. The interaction energies, — Eelec, computed for each structure at the
MP2/6-311 +G**//SCF/6-31G** level are indicated along with the letter designation, fol-
lowed after the comma by the same quantity after correction by zero-point vibrational en-
ergies (but all with the 6-31G** basis set). The reader is cautioned at the outset that the au-
thors of this paper were not entirely clear as to when they had identified true minima or
stationary points on their surface and when the structures were the result of optimization
under some sort of constraint.
As pointed out in Chapter 2, the energetically preferred conformation is the "ringlike"
geometry (b). The energetic cost of the nonlinearity of the H-bond is presumably compen-
sated by the contact between the water oxygen and one of the hydrogens of H2CO. Also of
note is the nearly antiparallel alignment of the molecular dipole moments of the two sub-
units in (b). The more conventional single linear H-bond in (a) is slightly less stable. Struc-
ture (c) is very much like (a) except that the proton donor approaches the formaldehyde
along its C=O axis. The lesser stability of (c) is an indication that a lone-pair approach, as
in (a), is preferred by about 1 kcal/mol.
The bifurcated geometry in (d) is favored by the head-to-tail alignment of the two mol-
ecular dipoles. Nevertheless, the lack of a strong H-bond makes this structure less stable
than any of the aforementioned geometries. One can disrupt the latter partial H-bonds by
rotating one of the two molecules by 90° about the O..O axis, rotating the hydrogens out of
the plane of the carbonyl oxygen lone pairs. Such a rotation leaves intact the favorable align-
ment of the molecular dipoles. The highest level MP2/6-311+G** calculations predict an
energetic cost of 0.5 kcal/mol arising from this rotation.
Structure (f) permits both hydrogens of the water to interact with the carbonyl oxygen,
as well as allowing the water oxygen to approach one of the CH2 hydrogens. This geome-
try contains an antiparallel arrangement of the molecular dipoles. It was found to be only
224 Hydrogen Bonding

Figure 4.1 1 Interaction energies (in kcal/mol) of various complexes of H2CO with
H2O, computed at MP2/6-311 +G**//SCF/6-31G** level. The first number refers
to electronic part of E; the second includes ZPVE correction32.

slightly less stable than the other bifurcated structure (d). A 90° rotation of the H2CO mol-
ecule around its C=O axis removes the interaction between O of water and the H of CH2.
The rotation simultaneously displaces one of the C=O lone pairs from the water hydro-
gens. Nevertheless, the ensuing destabilization relative to (f) is only about 0.4 kcal/mol.
The arrangement in (h) permits the testing of the ability of H2CO to donate a proton to
the water. The authors found (h) to be a true minimum on the PES, but not surprisingly, less
stable than the H-bonds where HOH acts as the donor. An analogous reversal between the
roles as donor and acceptor, but probing a bifurcated type of H-bond, results in structure
(j). Like (d), the molecular dipoles are arranged head-to-tail, but the two bridging hydro-
gens are both associated with the H2CO molecule. As indicated in Fig. 4.11, this structure
is only about 0.8 kcal/mol less stable than the more conventional bifurcated geometry (d).
The authors tested the possibility that the water might donate a proton to the bond of
H2CO, but were unable to locate corresponding minima. They conclude that the PES is rel-
atively shallow, consistent with recent findings for the simpler water dimer.
Dimitrova and Peyerimhoff33 ignored structure (b), and instead focused their attention
on (a), (d), and (j') (where the prime indicates a 90° rotation to make a fully planar com-
plex). They optimized the geometries and added BSSE corrections, after which they ob-
tained binding energies, Eclcc, of 4.0, 2.4, and 2.0 kcal/mol, respectively, at the MP2/6-
311 + +G(2d,2p) level. Following ZPVE correction, these values become 2.2, 1.6, and 1.5
kcal/mol.
Extended Regions of Potential Energy Surface 225

Later work by Ramelot et al.34, incorporating full geometry optimizations at SCF and
correlated levels, verified the characterization of (a), (b), and (f) as stationary points on the
PES of H2CO...HOH. While (b) is a true minimum, (a) is a transition state. The rotation of
the water is associated with barrierless collapse to (b). Like (a), (f) is also a transition state
and not a true minimum. Rotation of the water again leads back to (b) without an energy
barrier. Its energy yields the conclusion that interchange of the two water hydrogens in (b)
passes through an energy barrier of 1.2 kcal/mol.

4.6 Amines

A surface of a different type was generated for the complexes of amines with hydrogen
halides. Brciz et al.35 considered the four possible pairs of HC1 and HBr with NH3 and
CH3NH2 with an eye toward determining if any would prefer a proton-transferred ion-pair
structure. Their surfaces were hence functions of the H—X distance, r, and the distance be-
tween the two heavy atoms, R(N .. X). For three of the systems, there is only one minimum
on the surface, corresponding to the neutral dimer. However, a sort of low-energy corridor
appears in each PES that would take the system toward the ion pair, even though the energy
rises monotonically along this pathway. It is only for the complex pairing HBr with
CH3NH2 that a second minimum appears, which corresponds to Br ...+HNH2CH3. Indeed,
the latter well is deeper than that for BrH...NH2CH3 by about 1.2 kcal/mol. The barrier
which must be surmounted to pass from one minimum to the next is 1.2 kcal/mol higher in
energy than the neutral complex.
Correlation was added to this picture several years later by Jasien and Stevens36 who
also employed gradient optimization techniques. These authors found the appearance of a
secondary, ion-pair minimum on the potential energy surface for C1H...NH3 and BrH ... NH 3
at the SCF level which disappears with correlation. In the case of IH...NH3, both the neutral
and ion-pair minima survive the inclusion of electron correlation and are within 1 kcal/mol
of each other. The authors located a transition state for conversion that lies 1.4 kcal/mol
higher in energy than the neutral pair. As an important point, the authors found that when
they added zero-point vibrations, the ion-pair disappears as a minimum from the surface.
Latajka et al.37 followed up this line of reasoning by considering complexes of HBr and
HI with ammonia and mono-, di-, and trimethylamine. They learned that the earlier finding
that correlation can change the character of the potential energy surface was indeed rather
general. While two minima were encountered in the SCF surfaces, most of the correlated
analogues contained only a single minimum. BrH ... NH 3 is present as the indicated neutral
pair while both neutral and ion pairs are present in the correlated surface of IH...NH3. Like
Brciz et al., the authors generated a potential energy surface for the latter complex in terms
of the two pertinent distances. The two minima are very close in energy, and an inter-
conversion pathway involves changes in R as well as the I—H distance. The ionic mini-
mum is not very highly developed and the system must surmount a barrier of only about
0.1 kcal/mol to climb out of it at the MP2 level. On the other hand, raising the level of cor-
relation to MP4 heightens the barrier to perhaps 0.6 kcal/mol, making the ion-pair more
likely to be observed.
As methyl groups are added to the amine, it becomes more basic and hence leads to a
stronger tendency for an ion pair. Only a single minimum is identified in the MP2 potential
energy surface of CH3NH2 with either HBr or HI. While the ion pair structure is clear for
HI, the proton is approximately midway between the nitrogen and halogen atoms in
Br .. H .. NH 2 CH 3 . Both dimethylation and trimethylation lead to only single, ion-pair min-
226 Hydrogen Bonding

ima in the complexes of the amines with HBr and HI. This sort of analysis has been ex-
tended by Bacskay and Craw38 who have illustrated that trimethylamine is a strong enough
base to extract a proton from HC1; the proton transfer potential of this complex contains a
single well, corresponding to the ion pair.

4.7 Summary

The potential energy surface of the ammonia dimer is unquestionably very flat. The calcu-
lations indicate that there is no clearly defined minimum. Rather, a very shallow trough ex-
ists on the surface connecting a configuration containing a C linear N—H--N arrangement
with an equivalent geometry in which the two NH3 molecules exchange places. This path
passes through a cyclic structure, which is comparable in energy to the two linear configu-
rations. Complexes pairing water with a hydrogen halide molecule were used as a testbed
to examine the energetic cost of large-scale nonlinearity in the H-bond. In the case of
H2O...HC1, rotation of the HC1 molecule by 90° completely eliminates the stability of the
complex, raising the energy by 8 kcal/mol from equilibrium. A secondary minimum occurs
for HBr, where H 2 O ... BrH is less stable than the H-bonded H2O...HBr by 4 kcal/mol. Tran-
sition from one minimum to the other must overcome a barrier of 6 kcal/mol. The two min-
ima are much closer in energy for HI: the energy of H2O...IH is less than 1 kcal/mol dif-
ferent from that of H2O...HI.
One mode of interconversion of proton donor and acceptor roles in HP...HP to FH...FH
passes through a transition state with C2h cyclic character. The energy barrier to this inter-
conversion is estimated as approximately 1 kcal/mol. A similar barrier is encountered in the
transition path which moves hydrogens from one side of the H-bond axis to the other. The
transition state consists of a fully linear HP...HF configuration of all four atoms. The two
transition states have interfluorine distances that differ by some 0.25 A.
The head-to-head F—H . . . H—F and tail-to-tail H—F . . . F—H linear configurations are
maxima on the surface. The former is particularly high in energy, 40 kcal/mol higher than
the minimum whereas the latter lies only about 7 kcal/mol above this point. Calculations
showed that the dispersion part of the interaction energy is more anisotropic than is usually
thought, particularly large in magnitude for the H—F . . . F—H configuration. In this case, the
attractive dispersion energy is approximately canceled by other correlation effects, most no-
tably the correlation-correction to the electrostatic interaction which is repulsive. When
added to the SCF interaction, the latter two terms can reproduce the fully correlated poten-
tial energy surface rather well.
As for (HF)2, the cyclic C2h structure of (HC1)2 is not a minimum, but rather a transition
state for interconversion between the two equilibrium geometries. In this case, however, the
cyclic geometry lies only about 0.2 kcal/mol higher in energy than the minima, rather than
the 1 kcal/mol in (HF)2. Also analogous to (HF)2, the fully linear C1H...C1H geometry rep-
resents a transition state on the surface of (HC1)2; here, this structure lies about 1.5 kcal/mol
higher in energy than the minimum, as compared to 1 kcal/mol in (HF)2. In general, the PES
of (HC1)2 is somewhat flatter than that for (HF)2.
Despite some notions in the literature that the PES of the water dimer contains minima
other than the equilibrium geometry, with a linear H-bond, no other geometries appear to
represent true minima. The bifurcated arrangement is a transition state for the interchange
of the two protons in the donor molecule, and a trifurcated structure is a stationary point of
second order. Although not true minima, some of these geometries arc not much higher in
energy than the linear H-bond, typically within 1 or 2 kcal/mol.
Extended Regions of Potential Energy Surface 227

The total interaction energy of the equilibrium water dimer was dissected into its com-
ponents to understand the nature of its anisotropy. The first-order component, prior to
dimerization-induced charge rearrangement, is fairly insensitive to wags of the proton ac-
ceptor, whereas the higher-order terms favor a large [3 angle, that is, planar versus pyrami-
dal. Further partitioning reveals that the electrostatic interaction is largely responsible for
the observed energetics of bending either the donor or acceptor away from the equilibrium
geometry. In certain cases, it is useful to represent the electrostatic contribution by its mul-
tipole series, and then focus on the first few terms, each of which has a simple physical in-
terpretation; for example, a dipole-dipole term. The anisotropy of correlated terms is smaller
but reveals an interesting trend. Whereas the dispersion energy favors a linear H-bond, that
is, a tends toward zero, a more complete treatment of correlation would tend toward non-
linearity. The behavior of the former can be rationalized on the basis of interactions between
individual MOs on each subunit. The trend toward nonlinearity on the part of the full cor-
relation component arises from the influence of the correlation upon the multipole moments
of the individual subunits. This effect can be embodied in es(12) which represents the ef-
fect of correlation upon the electrostatics of the interaction.
The potential energy surface of the H2CO,H2O pair contains a number of possible min-
ima. Most stable is the geometry wherein HOH acts as the donor to the C=O group of
H2CO in the primary H-bond, but a weaker interaction occurs between a C—H group of
H2CO and the water oxygen. This structure has been characterized as a true minimum on
the surface. Slightly higher in energy is a strictly linear H2CO...HOH H-bond, with no sec-
ondary interaction, but this geometry is a transition state. Forcing the proton donor HOH to
lie along the C=O axis, rather than a carbonyl lone pair, destabilizes the system by about
1 kcal/mol. Slightly higher in energy is a bifurcated H-bond wherein both HOH hydrogens
approach the carbonyl oxygen. A number of other geometries appear bound relative to the
isolated monomers, including a reverse C—H ... OH 2 interaction, where a C—H group acts
as proton donor.
Another sort of motion that a full PES must include is the transfer of the proton from the
donor molecule to the acceptor. Such a transfer would require a particularly strong acid,
coupled with a strong base, as it would yield an ion pair with a high degree of charge sep-
aration. The hydrogen halides are strong acids, just as amines are strong bases. It appears
that this proton transfer can take place for HBr or HI paired with alkylated amines. This
topic is explored in more detail in Chapter 6.

References
\. Latajka, Z., and Scheiner, S., The potential energy surface of (NH3)2, J. Chem. Phys 84, 341-347
(1986).
2. Tao, F.-M., and Klemperer, W., Ab initio search for the equilibrium structure of the ammonia
dimer, J. Chem. Phys. 99, 5976-5982 (1993).
3. Liu, S.-Y., Dykstra, C. E., Kolenbrander, K., and Lisy, J. M., Electrical properties of ammonia
and the structure of the ammonia dimer, J. Chem. Phys. 85, 2077-2083 (1986).
4. Sagarik, K. P., Ahlrichs, R., and Erode, S., Intermolecular potentials for ammonia based on
the test particle model and the coupled pair functional method, Mol. Phys. 57, 1247-1264
(1986).
5. Hassett, D. M., Marsden, C. J., and Smith, B. J., The ammonia dimer potential energy surface:
resolution of the apparent discrepancy between theory and experiment?, Chem. Phys. Lett. 183,
449-456(1991).
6. Hannachi, Y., Silvi, B., Perchard, J. P., and Bouteiller, Y., Ab initio study of the infrared photo-
conversion in the water-hydrogen iodide system, Chem. Phys. 154, 23-32 (1991).
228 Hydrogen Bonding

7. Curtiss, L. A., and Pople, J. A., Ab initio calculation of the force field of the hydrogen fluoride
dimer, J. Mol. Spectrosc. 61, 1-10 (1976).
8. Michael, D. W., Dykstra, C. E., and Lisy, J. M., Changes in the electronic structure and vibra-
tional potential of hydrogen fluoride upon dimerization: A well-correlated (HF)2 potential en-
ergy surface, J. Chem. Phys. 81, 5998-6006 (1984).
9. Redmon, M. J., and Binkley, J. S., Global potential energy hypersurface for dynamical studies
of energy transfer in HF—HF collisions, J. Chem. Phys. 87, 969-982 (1987).
10. Szczesniak, M. M., and Chalasinski, G., Anisotropy of correlation effects in hydrogen-bonded
systems: The HF dimer, Chem. Phys. Lett. 161, 532-538 (1989).
11. Bunker, P. R., Kofranek, M., Lischka, H., and Karpfen, A., An analytical six-dimensional po-
tential energy surface for (HF)2 from ab initio calculations, J. Chem. Phys. 89, 3002-3007
(1988).
12. Karpfen, A., Bunker, P. R., and Jensen, P., An ab initio study of the hydrogen chloride dimer: the
potential energy surface and the characterization of the stationary points, Chem. Phys. 149,
299-309(1991).
13. Gdanitz, R. J., and Ahlrichs, R., The averaged coupled-pair functional (ACPF): A size-extensive
modification of MR CI(SD), Chem. Phys. Lett. 143, 413-420 (1988).
14. Tao, F.-M., and Klemperer, W., Ab initio potential energy surface for the HCl dimer, J. Chem.
Phys. 103,950-956(1995).
15. Elrod, M. J., and Saykally, R. J., Determination of the intermolecular potential energy surface
for (HCl) 2 from vibration-rotation-tunneling spectra, J. Chem. Phys. 103, 933-949 (1995).
16. Latajka, Z., and Scheiner, S., Structure, energetics and vibrational spectra H-bonded systems.
Dimers and trimers of HF and HCl, Chem. Phys. 122, 413-430 (1988).
17. Diercksen, G. H. F., SCF-MO-LCGO studies on hydrogen bonding. The water dimer, Theor.
Chim. Acta 21, 335-367 (1971).
18. Matsuoka, O., Clementi, E., and Yoshimine, M., CI study of the water dimer potential surface,
J. Chem. Phys. 64, (1976).
19. Baum, J. O., and Finney, J. L., An SCF-CI study of the water dimer potential surface and the ef-
fects of including the correlation energy, the basis set superposition error, and the Davidson cor-
rection, Mol. Phys. 55, 1097-1108 (1985).
20. Frisch, M. J., Pople, J. A., and Del Bene, J. E., Molecular orbital study of the dimers (AHn)2
formed from NH3 ,OH2, FH, PH3 ,SH 2 , and CIH, J. Phys. Chem. 89, 3664-3669 (1985).
21. Singh, U. C., and Kollman, P. A., A water dimer potential based on ab initio calculations using
Morokuma component analyses, J. Chem. Phys. 83, 4033-4040 (1985).
22. Douketis, C., Scoles, G., Marchetti, S., Zen, M., and Thakker, A. J., Intermolecular forces via
hybrid Hartree-Fock-SCF plus damped dispersion (HFD) energy calculation. An improved
spherical model, J. Chem. Phys. 76, 3057-3063 (1982).
23. Hobza, P., Mehlhorn, A., Carsky, P., and Zahradnik, R., Stacking interactions: Ab initio SCF and
MP2 study on (H2O)2, <H2S)2, (HCN)2, (CH 2 O) 2 , and(C2H4)2, J. Mol. Struct. (Theochem) 138,
387-399 (1986).
24. Dannenberg, J. J., An AMI and ab initio molecular orbital study of water dimer, J. Phys. Chem.
92,6869-6871 (1988).
25. Smith, B. J., Swanton, D. J., Pople, J. A., Schaefer, H. F., and Radom, L., Transition structures
for the interchange of hydrogen atoms within the water dimer, J. Chem. Phys. 92, 1240—1247
(1990).
26. Curtiss, L. A., Frurip, D. J., and Blander, M., Studies of molecular association in H2O and D2O
vapors by measurement of thermal conductivity, J. Chem. Phys. 71, 2703—2711 (1979).
27. Vos, R. J., Hendriks, R., and van Duijneveldt, F. B., SCF, MP2, and CEPA-l calculations on the
OH.. O hydrogen bonded complexes (H2O)2 and (H2O- H2), J. Comput. Chem. 11,1-18(1990).
28. Marsden, C. J., Smith, B. J., Pople, J. A., Schacl'cr, H. F., and Radom, L., Characterization of
the bifurcated structure of the water dimer, J. Chem. Phys. 95, 1825-1828 (1991).
29. Cybulski, S. M., and Scheiner, S., Factors contributing to distortion energies of bent hydrogen
bonds. Implications for proton-transfer potentials, J. Phys. Chem. 93, 6565-6574 (1989).
Extended Regions of Potential Energy Surface 229

30. Szczesniak, M. M., Brenstein, R. J., Cybulski, S. M., and Scheiner, S., Potential energy surface
for dispersion interaction in (H2O)2 and (HF)2, J. Phys. Chem. 94, 1781-1788 (1990).
31. Glendening, E. D., and Streitwieser, A., Natural energy decomposition analysis: An energy par-
titioning procedure for molecular interactions with application to weak hydrogen bonding,
strong ionic, and moderate donor-acceptor interactions, J. Chem. Phys. 100, 2900-2909 (1994).
32. Kumpf, R. A., and Daraewood, J. R., Jr., Interaction of formaldehyde with water, J. Phys. Chem.
93,4478-4486(1989).
33. Dimitrova, Y, and Peyerimhoff, S. D., Theoretical study of hydrogen-bonded formaldehyde-wa-
ter complexes, J. Phys. Chem. 97, 12731-12736 (1993).
34. Ramelot, T. A., Hu, C.-H., Fowler, J. E., DeLeeuw, B. J., and Schaefer, H. P., Carbonyl-water
hydrogen bonding: The H2CO H2O prototype, J. Chem. Phys. 100, 4347-4354 (1994).
35. Brciz, A., Karpfen, A., Lischka, H., and Schuster, P., A candidate for an ion pair in the vapor
phase: Proton transfer in complexes R3N HX, Chem. Phys. 89, 337-343 (1984).
36. Jasien, P. G., and Stevens, W. J., Theoretical studies of potential gas-phase charge-transfer com-
plexes: NH3 + HX(X=Cl,Br,I), Chem. Phys. Lett. 130, 127-131 (1986).
37. Latajka, Z., Scheiner, S., and Ratajczak, H., The proton position in amine-HX (X=Br,I) com-
plexes, Chem. Phys. 166, 85-96 (1992).
38. Bacskay, G. B., and Craw, J. S., Quantum chemical study of the trimethylamine-hydrogen chlo-
ride complex, Chem. Phys. Lett. 221, 167-174 (1994).
5

Cooperative Phenomena

he formation of a H-bond causes certain changes in the internal geometries of each


T monomer. For example, the bridging hydrogen moves a little farther away from the do-
nating atom. In addition to the deformations of the nuclear positions, the H-bond formation
is accompanied by real redistributions in the electronic structure of each subunit. These po-
larizations are readily apparent from experimental spectra, as well as from molecular or-
bital calculations. Taking into account the above perturbations of the two molecules in-
volved in a H-bonded dimer, it is reasonable to presume that the ability of either of these
two molecules to form another H-bond is altered by their participation in the first bond1.
Consider, for example, the pair of molecules AH and BH, each of which has a proton to
donate in a H-bond, and each of which contains one or more lone electron pairs appropri-
ate to accept a proton. If they form a H-bond of the type AH-BH, the proton of BH is still
available to form a H-bond to another molecule, CH. But the CH molecule will encounter
two different situations depending on whether the BH molecule is involved in the afore-
mentioned dimer, or is a single isolated BH molecule. In fact, formation of the AH---BH
complex will remove electron density from the BH subunit, density which is transferred
across to AH. This loss of negative charge will make BH a more powerful proton donor so
that one can expect the BH---CH interaction in the AH--BH--CH trimer to be stronger than
in the simpler BH- CH dimer. Analogous reasoning would make the AH molecule in
AH---BH a better proton acceptor, in comparison to the isolated AH molecule.
These effects that make the "whole larger than the sum of its parts" wherein a chain of
H-bonds is more strongly bound together than any of the individual links would be in the
absence of the others, is an expression of the "cooperative" nature of H-bonds. It is this co-
operativity that leads to the common occurrence of long strings of H-bonds. It has been
noted from surveys of crystal studies, for example, that H-bonds that occur as parts of such
strings tend to be considerably shorter, and presumably stronger, than isolated H-bonds2.
Cooperativity is a powerful enough phenomenon that it is responsible for the formation in
liquid formamide of six-membered rings3. Rigorous calculations have confirmed this rea-

230
tCooperative Phenomena 231

soning, indicating that it is polarizability of this sort that is largely responsible for the non-
additivity effects described in greater detail below for H-bond chains4.
But it should be emphasized that multiple H-bonds are not always cooperative in a pos-
itive sense. Again referring to the AH ... BH dimer, the electron density removal from BH
makes this molecule not only a better proton donor, but also a poorer proton acceptor. So
BH would now be less inclined to accept a proton from another molecule like CH. For this
reason, one would expect the total interaction energy in a trimer like that in Fig. 5.1 to be
weaker than in the pair of dimers AH ... BH and CH...BH. This sort of general weakening is
sometimes referred to oxymoronically as "negative cooperativity."
It is usually energetically unfavorable for a molecule to act as a double proton acceptor
as BH would in Fig. 5.1. For similar reasons, cooperativity is typically negative also when
a molecule acts as double proton donor. Of course, even in the case of negative coopera-
tivity, formation of the second H-bond is usually energetically favorable when compared to
the complete absence of a second H-bond. That is, even though the CH ... BH interaction
above is weaker than it would be in the absence of the other proton donor, AH, this inter-
action energy is still negative, and so will form spontaneously. In other words, two H-bonds
are always better than one (or usually so).
For the sake of consistency of terminology, triads of molecules in which the central unit
acts simultaneously as both proton donor and acceptor will be termed "sequential" to dis-
tinguish such configurations from those in which the central molecule acts as double pro-
ton donor or double acceptor. A perhaps more quantitative expression of cooperativity is re-
ferred to in the literature as "nonadditivity." The latter term is commonly taken as the
difference between the total interaction energy of an aggregation of molecules on one hand
and the sum of all the pairwise interactions on the other.
If adding a third molecule to a H-bonded dimer in a sequential fashion strengthens all
interactions, it is natural to wonder if the same will occur for a fourth and fifth and so on.
Calculations can address the behavior of the energetics as the chain grows. It is also possi-
ble to consider the effects of multiple H-bonding upon the geometries of each element in
the chain and the charge distributions within each. Another point of interest concerns the
vibrational modes of the growing chain which reflect upon the bonding characteristics.
The first section in this chapter focuses on HCN as an element in a chain of H-bonded
units. The presence in this molecule of only one proton and one lone electron pair provides
a simple testing ground for ideas about cooperativity. The linearity of the complex also sim-
plifies analysis of electron density redistributions caused by multiple bonding. HCCH is
similar in some ways, but its absence of a lone electron pair causes significant distinctions.

Figure 5.1 Example of negative cooperativity where B


serves as proton acceptor to both AH and CH.
232 Hydrogen Bonding

Since the equilibrium dimer is T-shaped, there is little propensity of the oligomers to con-
sist of linear chains.
HX molecules form strong H-bonds as dimers and are good candidates to manifest co-
operativity. The bent geometry of the dimer suggests that a ring can be formed even by
oligomers as small as the trimer. These complexes also provide a fertile ground to examine
many-body effects upon the energetics, their basis set sensitivity, and the relative impor-
tance of SCF versus correlation contributions. Of greatest importance from a practical point
of view, perhaps, is the cooperativity in liquid water. For this reason, this phenomenon has
been the subject of the highest-level calculations, leading to the most well founded conclu-
sions and the deepest insights. Mixed systems, in which the molecules involved in the
oligomers are not identical to one another, are discussed last.

5.1 HCN Chains

We begin our discussion of past consideration of cooperativity with the homogeneous


(HCN)n series because the linearity of this molecule and the single lone pair of the HCN
molecule make a linear geometry likely for a chain of any length. The analysis is hence un-
clouded by more complex angular distortions that accompany most other chains. The pres-
ence of only a single proton and a single lone pair on each monomer also simplifies our un-
derstanding as one need not be overly concerned with any molecule acting as double proton
donor or acceptor.
Motivated by gas phase spectroscopic studies of (HCN)35, Kofranek et al.6 used a vari-
ety of fairly large basis sets to examine oligomers of HCN of up to five molecules. They
optimized the geometries of the complexes by gradient procedures.

5.1.1 Geometries
Considering first fully linear arrangements, Fig. 5.2 illustrates how the C—H and C=N
bonds change their length as the chain is built up, one subunit at a time. In the jump from
the monomer to dimer, Fig. 5.2a illustrates the expected elongation of the C—H bond in the
proton donor molecule, and the concomitant smaller increase in the C—H bond length of
the proton acceptor molecule. As the chain gets longer, the donor designation indicates the
leftmost molecule of NCH .. (NCH) n-2 .. NCH, and the furthest right molecule is considered
the acceptor. Fig. 5.2a indicates that these elongation trends continue with each additional
molecule added in the middle of the chain, although these bond lengths slowly approach
asymptotes as n continues to increase. For odd n, we also consider the central molecule
which has an equal number of molecules on its left as on its right. The data points in Fig.
5.2a show how the elongating effects of being a donor or acceptor reinforce one another in
the middle molecule which simultaneously plays both roles. Hence, the C—H bonds are
longest in the central molecule.
Analogous data are presented in Fig. 5.2b for the CN bonds in the HCN linear chains.
In this case, the act of donating a proton makes the CN bond longer while it becomes shorter
if the molecule acts as acceptor. Again, the bond lengths approach asymptotes for the ter-
minal molecules as the chain grows longer. As before, the central molecule acts both as pro-
ton donor and acceptor, but in this case the middle molecule shows little modification as
compared to the monomer, since these two roles produce opposite effects on the CN bond
length.
Figure 5.2 Optimized lengths of (a) C—H and (b) C N bonds in
linear arrangements of (HCN)n, n = 1-5. Donor refers to the left-
most molecule and acceptor to the rightmost in NCH .. (NCH) n _ 2 .. NCH.
The middle designation indicates the central molecule for n = 3 and
n = 5. Data are taken from SCF optimizations with a [53/3] basis
set6.
234 Hydrogen Bonding

5.1.2 Energetics
The binding energies of the various linear oligomers were also computed by Kofranek et
al.6 and the data are listed in Table 5.1. The first two columns report the energetics of as-
sembling each complex from n isolated monomers to represent the total binding energy and
enthalpy, respectively. For example, at the particular SCF/[53/3] level used, the electronic
contribution to the binding energy of the dimer is 5.6 kcal/mol. After inclusion of zero point
vibrational and other effects, the authors obtain a binding enthalpy of 4.7 kcal/mol. — Eelec
for assembly of the full trimer is 12.5 kcal/mol. Since this total is more than twice that of
— Eelec for the dimer, the two H-bonds in the trimer are stronger in sum than two isolated
H-bonds, as would occur in a pair of dimers. This difference is reported in the next two
columns of Table 5.1 as the cooperativity. Hence, for either Eelec or H°, the binding en-
ergy of the trimer exceeds that of two dimers by 1.3 kcal/mol.
The series progresses to n = 4 in the next row where total binding energies are nearly
20 kcal/mol. These totals exceed the sum of three dimers by 3 kcal/mol. The latter amount
is divided by two in the next two columns of Table 5.1 to facilitate comparison with the
trimer in the preceding row. That is, there are two "trimers" present within the tetramer so
the authors divide by 2 for a fair comparison. Note that these cooperativities of 1.5 kcal/mol
are larger than the values for the trimer. The cooperativity increases further in the pentamer.
One can imagine that these cooperativities continue to increase while approaching an as-
ymptote as n .
The convergence of the binding energy with chain length can be gleaned from the last
two columns of Table 5.1. — Qn/(n-l) refers to the average H-bond energy of a given
oligomer where the full interaction energy of the n monomers is divided by the number of
adjacent pairs in the oligomer. For example, while - Eelec is 5.6 kcal/mol for the dimer,
the average interaction energy of the four H-bonds present in (HCN)5 is approaching 7
kcal/mol.
This sort of analysis was later extended to the correlated MP2 level, using a 6-31+G**
basis set, with counterpoise corrections introduced7. Calculations up to the heptamer indi-
cated the anticipated approach toward an asymptote of the average binding energy. This as-
ymptote is somewhat smaller than in Table 5.1, tending toward about 5.8 kcal/mol at the
SCF level for — Eelec/(n-l). This asymptote represents an approximate enhancement of
40%, as compared to the dimer.

Table 5.1 Energetics of binding (kcal/mol) in linear oligomers (HCN)n, as calculated at the
SCF level6.
coopa — Qn/(n — 1)

n - Eelec -H - Eelec - H° - E.,.,,. - H°

2 5.60 4.72 5.60 4.72


3 12.52 10.75 1.32 1.31 6.26 5.38
4 19.89 17.21 1.55 1.53 6.63 5.74
5 27.47 23.87 1.69 1.66 6.87 5.97
a
Cooperativity is defined here as follows: if Qn is the property of interest for (HCN)n, coop is evaluated as [ Qn
(n- 1) Q 2 ] /(n- 2).
Cooperative Phenomena 235

5.1.3 DipoleMoments
It was stated earlier that formation of a first H-bond in a dimer will polarize both molecules.
One can hence expect a dipole moment in the dimer that is larger than the vector sum of the
individual moments of the isolated monomers. This supposition is confirmed by the SCF
moments of these oligomers which are reported in the second column of Table 5.2. The third
column lists the cooperativity, measured in a manner analogous to that for the energies. This
property is approximately 1 D and increases with n. The average dipole moment per mole-
cule in the chain is exhibited in the last column of Table 5.2 and echoes the cooperative na-
ture of the electronic rearrangements within the chain. Later experimental measurements
by Ruoff et al.8 indicated that the dipole moment of the linear trimer exceeds the vector sum
of the three monomers by 1.8 D, quite close to the estimation of 2.0 D in Table 5.2 (i.e.,
twice the value listed for n = 3 in column three).

5.1.4 Vibrational Spectra


Kofranek et al.6 carried out vibrational analyses of their various complexes. It must be stated
at the outset that the frequencies are not so clearly identified with any particular single bond
as are geometrical features. Nevertheless, the authors were able to identify various modes
as largely C—H or C N stretches, or intramolecular bends, even in the longer chains. With
respect to the C—H stretches, the longer oligomers exhibit a spread of frequencies. The
highest of these is fairly clearly identified with the proton acceptor molecule and the oth-
ers are within a narrow range of each other. We denote the lowest frequency as that of the
terminal donor molecule with the caveat that this designation is somewhat arbitrary. Anal-
ogous reasoning was applied to the CN stretches.
The calculated frequencies exhibit behavior very much like the bond lengths in some
ways. For example, the CH stretching frequencies in Fig. 5.3a suffer decreases as the chain
grows, with this red shift being more pronounced for the proton donor end of the chain.
These frequency drops are consistent with the bond stretches described in Fig. 5.2a. In fact,
the two plots are nearly perfect mirrors of one another. The CN stretching frequency of the
donor end of the chain, illustrated in Fig. 5.3b, diminishes with larger n, also consistent with
the analogous bond stretches of Fig. 5.2b; the behavior of the acceptor is again opposite in
sign. Note the approach to an asymptote for all frequencies in Fig. 5.3. Another important
aspect is the magnitude of changes in bond length and frequency. The CH bonds stretch by
nearly 0.01 A upon forming the serial H-bonds; the red shift of the associated frequency is

Table 5.2 Calculated dipole moments (D) in linear oligomers


(HCN)n, as calculated at the SCF level6.
n n coopa /n

1 3.20 3.20
2 7.29 0.89 /n
3.65
3 11.63 1.02 3.88
4 16.08 1.09 4.02
5 20.57 1.14 4.11
a
Cooperativity is defined here as [ n — n 1 ] / (n — 1).
Figure 5.3 Calculated stretching frequencies of (a) C—H and (b)
C N bonds in linear arrangements of (HCN)n, n = 1-5. Donor
refers to the leftmost molecule (assumed to be the lowest frequency)
and acceptor to the rightmost (highest v) in NCH .. (NCH) n _ 2 .. NCH.
Data derived at SCF/[53/3] level in6.
Cooperative Phenomena 237

some 150 c m - 1 . In contrast, the CN bonds change their length by less than 0.002 A, with
shifts of only 20 cm-1 or so.
The intermolecular frequencies were also computed and are presented in Fig. 5.4 as a
function of chain length. For the dimer, the stretching frequency, v , is 122 cm - 1 , brack-
eted by the two intermolecular bending modes. For each of these three modes, growth of
the chain leads to a drop in one frequency and an increase in the other. The proportional ef-
fects of chain length on these frequencies are enormous. For example, v nearly doubles
upon going from dimer to pentamer; the lower-frequency bend drops by 80%.
The same authors followed up their earlier work with an elaboration of the oligomers up
to trimer which computed the intensities, as well as frequencies, of the vibrational transi-
tions9. This work included electron correlation through the coupled pair functional (CPF)
technique, using polarized basis sets. Their results for the trimer are summarized in Table
5.3. Correlation appears to dampen the red shift of the three CH stretching frequencies while
making all of the CN frequencies more positive. The changes in the bending frequencies
are not quite as regular. But correlation does not appear to produce major changes in any of
these shifts. The band intensifications, listed in the last two columns of Table 5.3, also show
slight correlation-induced increases in the CH stretches, with virtually no changes occur-
ring at all in the bending intensities. However, correlation produces marked changes in the
CN stretches, raising the intensification by an order of magnitude or more in each case. This
large discrepancy between SCF and correlated calculations is consistent with that noted
above for the dimer.

Figure 5.4 Calculated intermolecular frequencies in linear


arrangements of (HCN)n n = 2-5. Highest and lowest frequency
of each type are illustrated. Data derived at SCF/[53/3] level in6.
238 Hydrogen Bonding

Table 5.3 Calculated frequency shifts and intensification ratios in linear trimers
of HCN, as compared to the monomer9.

vCcm-1) A tri/Amon

Mode SCF CPF SCF CPF

CH stretch -8 -3 0.9 1.0


-76 -68 0.01 0.02
-86 -76 8.6 11.1
CN stretch 10 15 5 59
-4 3 9.8 258
-14 -7 3.0 33
HCN bend 85 96 0.5 0.5
74 90 1.4 1.5
20 10 0.9 0.9

Just as one can consider the effects of dimerization upon the frequencies or intensities
of the monomer modes, it is useful to ponder how juxtaposing two "dinners" within a trimer
affects the vibrational spectra of the dimer. Kofranek et al.9 provided such information, re-
ported in Table 5.4, which suggests that the SCF data graphed in Fig. 5.4 are quite repre-
sentative of the correlated level as well. All of the trends in the SCF frequencies are mir-
rored in the CPF data. The similarity of the last two columns in the table indicates no major
perturbations of the intermolecular band intensifications caused by electron correlation.
Anex et al.10 later repeated some of this work using a 6-31G* basis set and including
electron correlation via MP2. Their data verified the earlier calculations of (HCN)n and
went further by describing the various normal modes that correspond to the ab initio force
field. The highest frequency mode was found to correspond to independent C—H stretch-
ing motion of the terminal proton acceptor molecule (with 8% C N stretching mixed in).
The C—H stretches of the other two molecules mix with each other in the next two modes.
Karpfen 11 was able to consider an infinitely extended linear chain of HCN molecules
using an ab initio crystal-orbital approach, with basis sets ranging from STO-3G to
[641/41], albeit without any consideration of electron correlation. Polynomials of second
and third order were fit to the results, computed for a grid of values of r(CH), r(CN), and

Table 5.4 Calculated frequency shifts and intensification ratios in the


intermolecular modes of the linear trimers of HCN, as compared to the dimer9.

v3-v2 (cm-1) Atri/Adim

Mode SCF CPF SCF CPF

H-bond stretch 36 38 0 0
-24 -27 1.5 0.7
H-bond bend 31 27 1.5 2.0
-17 -13 0 0
11 4 2.6 2.0
-23 -27 0.05 0.06
Cooperative Phenomena 239

Table 5.5 Quadrupole coupling constants (in MHz) in linear


dimers of DCN, computed at SCF/6 31+G* level12.

n D 14N

1 0.225 -4.758
2 a 0.223 -4.426
d 0.209 -4.660
3 a 0.223 -4.366
d-a 0.204 -4.345
d 0.205 -4.638
4 a 0.222 -4.347
d-a 0.200 -4.317
d 0.204 -4.629
5 a 0.222 -4.340
d-a 0.199 -4.252
d 0.204 -4.625
6 a 0.222 -4.336
d-a 0.197 -4.241
d 0.204 -4.623

r(H .. N). He found with his best basis set that the formation of such an infinite chain di-
minishes the CH stretching frequency, relative to the monomer, by 138 cm - 1 , while that of
the CN stretch is constant to within 1 cm - 1 . In comparison to a calculated binding energy
of 4.4 kcal/mol for the single H-bond in the dimer, this same basis set yielded an average
of 6.1 kcal/mol in the chain.

5.1.5 Quadrupole Coupling Constants


The effects of cooperativity can also extend to other properties such as quadrupole coupling
constants. These quantities were computed for the D and 14N nuclei in linear clusters of
(HCN)n, with n varying up to 6, at the SCF level with a 6-31+G* basis set12. The calcu-
lated values are listed in Table 5.5 where d refers to the proton donor molecule on one end
of the chain, a to the acceptor on the other end, and d-a to a combined donor-acceptor mol-
ecule in the middle of each chain.
The coupling constants of both the D and 14N atoms of the monomer are lowered in mag-
nitude when the H-bond is formed in the dimer. The value of D is not affected much in the
proton acceptor molecule; conversely that of 14N is lowered much more in the proton ac-
ceptor. As the chain becomes longer, the quadrupole coupling constants of the D atoms of
the molecules on the two ends of the chain change very little, indicating little cooperativ-
ity. On the other hand, the value for a molecule in the middle of the chain, serving as both
donor and acceptor, is slightly smaller than for either terminal molecule.
Cooperativity is more obvious in the 14N coupling constants. The values for the termi-
nal molecules continue to diminish as the chain grows beyond the dimer level. As in the
case of the D atom, the coupling constants are smaller for the molecules in the middle of
the chain. The authors noted a close linear correlation between the 14N quadrupole coupling
constants and the lengths of the pertinent H-bonds. This same property was found to corre-
late also with the VCH stretching frequencies.
240 Hydrogen Bonding

5.1.6 Cyclic Chains


The terminal HCN molecule on one end of a linear chain contains a nitrogen lone electron
pair that is not involved in a H-bond. Nor is there any involvement by the proton of the HCN
on the other end of the chain. If these two entities could be paired up, an additional H-bond
might be formed, presumably stabilizing the entire system. However, in order to bring these
two groups into proximity to one another, by forming a ring, the deformation from nonlin-
earity would destabilize the existing H-bonds. Hence, the relative stability of a cyclic
oligomer rests on two competing effects: formation of the last H-bond versus deformation
of all the others. In the case of a long chain (large n), the latter deformation can be rather
small since each H-bond must bend by only a small amount. But when there are only a small
number of molecules, the angular deformations are considerable. Hence, one can expect lin-
ear chains to be preferred for small n, but the cycles will become more favorable as n in-
creases past some threshold value. Of course, the foregoing analysis is based only upon en-
ergetics; the higher constraints in the cyclic geometry would disfavor this arrangement from
entropic considerations.
In the case of (HCN)n, Kofranek et al.6 found the threshold of conversion of linear to
cyclic, on energetic grounds, to occur at approximately n = 4. The linear trimer is favored
over the cyclic by 2.1 kcal/mol, but the latter tetramer is more stable than the linear by 0.1
kcal/mol, again at the SCF/[53/3] level.
Kurnig et al.13 reexamined the relative stabilities of the linear and cyclic trimers, moti-
vated by experimental indications that two conformers coexist8,10,14. They included corre-
lation via the ACPF method15 which they consider equivalent in this case to CPF, along
with a basis set of [321/31] quality. Their results in Table 5.6 show that the linear trimer is
more stable than the cyclic by 2 kcal/mol, with respect to binding energy at the SCF level,
but this difference is reduced to 0.7 with correlation included. Including zero-point vibra-
tions and other corrections yields the AH° data in the last row which nearly mirror the trends
in E. Given the errors which remain at this level of calculation, the calculated preference
for the linear trimer by 0.5 kcal/mol in AH° was not considered by the authors as definitive,
and the smallness of this quantity accounted for the presence of both geometries in the ex-
periments.

5.2 HCCH Aggregates

The C—H group of HCCH is of comparable strength as a proton donor to that in HCN, al-
beit somewhat less acidic, but the former molecule does not contain a lone electron pair.

Table 5.6 Comparative stabilities of linear and cyclic trimers of


HCN13. Data shown represent binding relative to three isolated
monomers (kcal/mol)
SCF ACPF

linear cyclic linear cyclic

- Eelec
- E
elec
12.5 10.4 10.1 9.4
- H° 10.7 9.0 8.4 7.9
Cooperative Phenomena 241

Hence, the interaction between a pair of acetylene molecules places the proton in some
proximity to the alternate source of electron density, the cloud between the two C atoms.
The weakness of the interaction, with — Eelec less than 2 kcal/mol, and the absence of ap-
preciable perturbations in the monomer geometries, makes questionable the characteriza-
tion of the interaction in (HCCH)2 as a H-bond. It is legitimate to wonder if perhaps the in-
teraction might be strengthened by the effects of cooperativity. That is, since the interaction
between any pair of HCN molecules in a chain is enhanced relative to that in the dimer, the
same phenomenon might also occur for oligomers of HCCH, such that the pairwise inter-
action might be unambiguously classified as a H-bond. With this in mind, we peruse the lit-
erature.

5.2.1 Trimers
Alberts et al.16 were the first to consider (HCCH)3 applying a DZP approach at the SCF
level. Keeping in mind that the absence of a lone pair on either end of the HCCH molecule
makes any linear arrangement unlikely, it is not surprising that this system tends toward a
cyclic arrangement. Optimization led to the C3h structure pictured in Fig. 5.5. One can en-
vision how this geometry permits one H atom of each molecule to interact with the cloud
of its neighbor. This configuration is stable by 2.6 kcal/mol, relative to three monomers, or
0.9 kcal/mol for each pairwise interaction. The latter quantity is nearly identical to — Eelec
calculated at the same level for the dimer (T-shaped), so the evidence of cooperativity here
is weak. On the other hand, this equality is indicative of a certain degree of cooperativity
since each of the pairwise interactions is distorted from a true T-shape. Moreover, this is
only the SCF result and earlier work with the dimer illustrated that a good deal of the in-
teraction takes place in the context of electron correlation effects.
With the inadequacy of SCF treatments for a system of this type in mind, the trimer was
reexamined shortly thereafter17, employing the MP2 treatment of correlation. A number of
different configurations were first optimized at the SCF level so as to identify stationary
points. In addition to the minimum identified previously and illustrated in Fig. 5.5, a num-
ber of other configurations were identified as minima, albeit of lesser stability. Fig. 5.6 il-
lustrates the other two minima wherein the protons of two different molecules attack the
same cloud (a) and the conjugate situation where the two hydrogens of a single central
molecule approach the clouds of separate molecules (b). In comparison to the cyclic
arrangement, with a binding energy of 2.3 kcal/mol, (a) and (b) have — Eelec values of 1.4
and 1.5 kcal/mol, respectively. Reoptimization of the cyclic structure in Fig. 5.5 incorpo-

Figure 5.5 C3h geometry of the HCCH trimer 16 .


242 Hydrogen Bonding

Figure 5.6 Alternate geometries of the HCCH trimer17. The


central molecule acts as a double proton acceptor in (a) and
as a double donor in (b).

rating correlation through MP2, enhanced the binding energy to 3.0 kcal/mol; enlarging the
basis set to TZ2P raised this quantity still more, to 3.8 kcal/mol, again at the MP2 level.
There is evidence of weak cooperativity in the geometries. The MP2/DZP distance be-
tween centers of mass is 4.310 A in the dimer, but only smaller by 0.02 A in the cyclic trimer.
The binding energy of this trimer is 3.0 kcal/mol at the MP2/DZP level, so that each of the
three pairwise interactions contributes 1.0. Since this value is equal to the binding energy
of the dimer at the same level of theory, we again conclude that the evidence of coopera-
tivity in the cyclic trimer is present but weak.
One may consider a comparison between dimer and cyclic trimer as unfair in that the
pairwise geometries are different, distorted T-shapes in the latter. A better comparison might
involve the other trimer minima in Fig. 5.6 which contain true T-shaped interactions. The
interaction energies of the trimers (a) and (b) are both less than twice the binding energy of
the dimer, indicating a negative cooperativity. This result is consistent with the idea that a
single molecule cannot act effectively as either double proton acceptor, as in (a), or
donor, (b).

5.2.2 Tetramers and Pentamers


Yu et al.18 took matters a step further and computed optimized geometries of acetylene
oligomers up to n = 5. Taking as a cue the fact that monomer geometries are virtually un-
changed within the dimer, they held internal geometries fixed during the course of their op-
timizations. Trimer and tetramer geometries were optimized at the SCF and MP2 levels with
a 3-21G basis set, and then MP2/6-311G** was applied to obtain the interaction energy al
Cooperative Phenomena 243

these geometries; pentamers were considered only at the SCF/3-21G level. The tetramer
was interesting in that experimental work19 had indicated a cyclic S4 complex very much
like the trimer in Fig. 5.5 except that each molecule is tilted by nearly 30° with respect to
the overall molecular plane. This configuration is illustrated in Fig. 5.7a while a fully pla-
nar C4h equivalent is shown in Fig. 5.7b. Calculations revealed no discernible difference in
energy between these two configurations at any level of theory. The MP2/3-21G interac-
tion energy of the entire complex is 9.3 kcal/mol, which works out to 2.3 kcal/mol for each
"H-bond," that is, pairwise interaction, as indicated in the second column of data in Table
5.720. This value is a little larger than 2.0 kcal/mol in the trimer or the T-shaped dimer. The
SCF values in the preceding column of the table indicate only small fluctuations in the H-
bond energy as the size of the complex increases from n = 2 to n = 5. The pentamers stud-
ied by Yu et al. were comparable to the tetramers in Fig. 5.7 in that both were cyclic, one
fully planar and the other staggered. The 8° distortions in the latter stabilized the system
slightly, by around 0.15 kcal/mol.
Bone et al.20 also considered the higher oligomers, n = 4, 5. They improved on the ear-
lier methodology by employing a DZP basis set which contains polarization functions, and
optimized some geometries at the MP2 level, including internal parameters. Their opti-
mized tetramer was planar, as pictured in Fig. 5.7b, with the monomer centers of mass defin-
ing the vertices of a square. However, closer inspection of the SCF frequencies revealed
this structure to be a saddle point. The small values of a number of imaginary frequencies
(less than 1i cm-1) prevented a determination of the order of this saddle point. The authors
hence searched for a S4 structure which is probably the true minimum, but the flatness of
the surface prevented an unambiguous identification of such a minimum. They concluded
the planar C4h geometry is in reality the global minimum, but the surface is extremely flat
in this region, such that large-amplitude excursions from planarity will be experienced by
the cluster.
In the case of the pentamer, Bone et al.20 considered two possible geometries: a five-
membered version of the planar ring in Fig. 5.7, as pictured in Fig. 5.8a, or a different
arrangement which contains two three-membered rings, clustered around a central mole-
cule, illustrated in Fig. 5.8b. Both were found at the SCF/DZP level to be stationary points
on the PES. The double ring (b) was bound relative to five monomers by 4.7 kcal/mol, as
compared to 4.5 kcal/mol for (a). The authors attributed the greater stability of (b) to the

Figure 5.7 Configurations of the HCCH tetramer18. The complexes


in (a) and (b) are nonplanar and planar, respectively.
244 Hydrogen Bonding

Table 5.7 Interaction energies per H-bond (— Eelec in kcal/mol)


in oligomers of HCCH18,20.
3-21G DZP

SCF MP2 SCF MP2

dimer 1.6 2.0 0.78 0.97


trimer 1.5 2.0 0.78 1.01
tetramer 1.8 2.3 0.91 1.15
pentamer 1.7 0.89 1.13

presence of six pairwise interactions, as compared to only five in (a). The six pairwise con-
tacts in (b) average 0.8 kcal/mol each, as compared to precisely the same amount in the T
dimer at this level of theory, indicating little cooperativity. Frequency calculations revealed
both (a) and (b) to represent transition states on the surface, with the true minima probably
nonplanar and puckered, analogous to the tetramers.
The last two columns of Table 5.7 illustrate the average binding energies of the various
cyclic structures of the HCCH oligomers, calculated with the DZP basis set, and with BSSE
corrections included. These values are considerably smaller than the 3-21G data of Yu et
al., due in large part to the poor performance of the latter set for many molecular interac-
tions, coupled with its large BSSE (left uncorrected by the authors). Nonetheless, both sets
of data evidence similar trends. There is little cooperativity and what there is seems to peak
at n = 4. This may be due to the perfect perpendicular arrangement of each pair in the
tetramer. The antisymmetric C—H stretching frequencies also provide little evidence of co-
operativity. The two such frequencies in the T-dimer are 3572 and 3576 cm-1. Forming the
cyclic trimer in Fig. 5.5 yields three frequencies of 3567 c m - 1 , which is reduced only
slightly to 3564 in the cyclic tetramer and 3565 in Fig. 5.8a.

Figure 5.8 Possible geometries of the HCCH pentamer20.


Cooperative Phenomena 245

5.3 Hydrogen Halides

HX molecules contain a single proton but three lone electron pairs on the halogen atom.
Since none of these lone pairs is collinear with the HX bond, and the HX dimer does not
contain all four atoms on a single axis, it is unlikely that a chain of such molecules will be
linear.

5.3.1 Open versus Cyclic Trimers


One likely geometry for a trimer would contain a pair of dimer structures as in Fig. 5.9a,
where each H-bond is linear, or nearly so, and the H . . X—H angle is in the 100-130° range.
As in some of the cases already described, a third H-bond can be formed if the proton on
the last molecule is brought near the X atom of the first, forming a cyclic structure as in Fig.
5.9b. Of course the price paid for this third H-bond is some angular distortion of all three.
Some early studies of the comparative stability of these two arrangements of the HF
trimer with small basis sets had yielded ambiguous answers. In 1983, Karpfen et al.21 ap-
plied more extended basis sets and optimized the geometries with gradient procedures. They
found the cyclic trimer more stable than the open one by 2.2 kcal/mol. Their results con-
firm the greater strain in each of the three H-bonds of the trimer: the interaction energy per
H-bond is 3.9 kcal/mol in the cyclic structure, as compared to the 4.7 kcal/mol in each of
the two bonds of the open form. Soon thereafter, another study of the same system focused
on the vibrational frequencies of the cyclic structure22. Each of the three H-bonds was cal-
culated to contribute 5.0 kcal/mol, slightly more than computed earlier by Karpfen et al.
Calculations in 198623 confirmed the greater stability of the cyclic geometry, as well as
its planar nature, and suggested that the open form could convert to this geometry with lit-
tle or no energy barrier. (Indeed, other calculations in the same year dispensed with the no-
tion that the open form of the HF trimer represents a real minimum on the potential energy
surface24.) It was also pointed out that all experimental measurements were indicative of
the existence of the cyclic structure25. It was found here that the H-bond energy per bond

Figure 5.9 Open and cyclic arrangements of the HX


trimer.
246 Hydrogen Bonding

in the trimer (4.3 kcal/mol) is in fact slightly greater than the interaction energy of the dimer
(4.1) despite the strain in the former. One may presume that the cooperativity effect around
the ring (each molecule is both donor and acceptor) is compensating for the unfavorable
strain. This same study considered the ramifications of electron correlation and found little
qualitative change. All H-bonds, whether in dimer or trimer are strengthened by a small
amount. The correlated per-bond H-bond energy in the trimer is 4.7 kcal/mol, compared to
4.6 in the dimer. Interestingly, the authors pointed out a sharp increase in the correlated H-
bond energy per bond to 6.4 kcal/mol in the tetramer. This increase is probably the result
of a drop in the strain energy.
Latajka and Scheiner26 considered the trimer of HF, along with (HC1)3, with extended
basis sets designed to minimize basis set superposition error. The geometry optimizations
led to C3h minima for both trimers, conforming to experiment27-30. Electron correlation in-
duced a contraction in the interhalogen distances, more noticeable in (HC1)3. With correla-
tion included and BSSE corrected by the counterpoise procedure, the binding energy in
(HF)3 is 4.2 kcal/mol per H-bond, considerably stronger than the 1.2 kcal/mol for each H-
bond in (HC1)3. Confirming earlier calculations, the binding energy in the HF trimer indi-
cates a compensation between the strengthening caused by cooperativity, and the weaken-
ing that accompanies the angular distortions necessary to form the triangular trimer. The
interchlorine distance computed for (HC1)3 is 3.78 A, in surprisingly good coincidence with
an experimental measurement of 3.69 A between centers of mass29, in light of the fairly
small basis set, 6-31G**, used in the calculation.

5.3.2 Three-Body Interaction Energies


The question of cooperativity was further probed by computation of the three-body inter-
action energy, E326. This quantity is defined as the difference between the total interac-
tion energy in the entire trimer, and three times the two-body interaction, E2 (one for each
pair of subunits), all computed as Eelcc. This property differs from the prior means of con-
sidering cooperativity chiefly in that the geometry of all species are frozen in that of the op-
timized trimer. Thus, the two-body interaction is not computed in the optimized structure
of the dimer, but that of the trimer, so the effects of angular strain are present here, too.
The resulting quantities are reported in Table 5.8 where the "cc" superscript indicates
counterpoise correction of basis set superposition error. The two-body terms in the first row
underscore the greater H-bond interactions in oligomers of HF, as compared to HC1. When
corrected for superposition error, MP2 adds little to the interactions between HF monomers,
but is responsible for a 25% enhancement in HC1. It is important to mention that erroneous

Table 5.8 Two and three-body interaction energies ( Eelec in kcal/mol) in the
trimers of HF and HC1, computed with +VPS basis set26.

(HF)3 (HC1)3

SCF MP2 SCF MP2

E2 -3.67 -4.04 -0.94 -1.58


CC
E2 -3.45 -3.35 -0.90 -1.13
E3 -1.47 -1.39 0.31 -0.36
ECC
3 -1.66 -1.73 -0.23 -0.22
Cooperative Phenomena 247

conclusions of an exaggerated MP2 component would be drawn if counterpoise corrections


were not included (compare with first row of data). This exaggeration is even more pro-
found with other basis sets, particularly the "standard" ones such as 6-31G*26. The three-
body terms are sizable: E3CC is roughly half the magnitude of E2CC for (HF)3 at the MP2
level, and E3CC/ E2CC = 20% for (HC1)3. Quite similar ratios pertain to the SCF calcula-
tions. It is worth stressing that correlation contributes very little to the three-body terms.
This conclusion is confirmed by more rigorous partitioning of interaction energies31,
which suggest that SCF computations are adequate to study most aspects of 3-body inter-
actions involving H-bonds4. The nonadditivity of the cyclic HX trimer was partitioned into
various contributions by intermolecular perturbation theory, combined with M011er-Plesset
treatment of correlation, in 198932. After demonstrating the importance of correcting basis
set superposition error of all terms, the authors stress that the three-body effect is dominated
by its SCF component, as demonstrated by the data in Table 5.9. The latter is, in turn, pri-
marily composed of the term attributed to electron density deformation, as the Heitler-
London exchange nonadditivity amounts to only 5% of the total. The SCF portion of the
three-body term is rather insensitive to improvements of the basis set beyond 6-31G* *. Cor-
relation adds very little to the nonadditivity, making up only 5% of the total. The three-body
terms arising from each level of correlation alternate in sign, making for cancellation. That
is, the MP2 and MP4 terms are both negative (attractive), while MP3 contributes a positive
quantity. There is surprisingly little basis set sensitivity of any of these terms.
The manner in which the two- and three-body terms change as the three molecules are
pulled apart is illustrated graphically in Fig. 5.10. In the vicinity of the equilibrium geom-
etry (R = 2.73 A), the SCF terms clearly dwarf MP2 contributions. It is interesting to note
the crossing of the two- and three-body MP2 terms. The three-body SCF term diminishes
toward zero more quickly than does the two-body term, which approaches its asymptote
gradually. Most of the nonadditivity can be traced to the deformation terms within the SCF
approximation. The two- and three-body MP2 contributions become comparable in magni-
tude for interfluorine distances of 4 A or so.

Table 5.9 Two and three-body interaction energies (kcal/mol)


in the trimers of HF and HC1. Data calculated with medium-
polarized [7s5p2d/5s3p2d/3s2p] basis set for [Cl/F/H]32.

(HF), (HC1)3
2-body terms
SCF -3.20 -0.70
Heitler-London -1.89 -0.41
Deformation -1.31 -1.24
MP2 -0.20 -0.66
MP3 -0.07 0.17
SCF + MP2 + MP3 -3.47 -1.20
3-body terms
SCF -1.59 -0.25
Heitler-London -0.11 -0.00
Deformation -1.48 -0.24
MP2 -0.07 0.00
MP3 0.05 0.02
SCF + MP2 + MP3 -1.62 -0.23
248 Hydrogen Bonding

Figure 5. \ 0 Variation of SCF and MP2 two-body (solid lines) and


three-body (broken lines) terms for cyclic (HF)332.

Variations in these quantities as the H atoms are pivoted around the fluorines are exhib-
ited in Fig. 5.11. The two-body SCF term is clearly the most anisotropic. Its minimum in
the vicinity of a = 30° dominates the determination of the equilibrium geometry of the
trimer. Both the SCF and MP2 three-body terms favor smaller angles, with the SCF com-
ponent both of larger magnitude and more sensitive to a. More recent work using a some-
what different formalism33 has largely supported the angular sensitivity reported here al-
though there are certain points of lingering discrepancy.

5.3.3 Larger Oligomers


The cyclic complexes of HP investigated were extended to the hexamer34 in 1990. Karpfen
applied a polarized basis set, and included correlation via the averaged coupled pair func-
tional method15. He restricted his investigations to C h planar symmetries; application of
correlation prohibited gradient optimizations. Later work by the same group35 applied a
much larger basis set, albeit without explicit correlation included. Nonetheless, similar
trends are observed. Fig. 5.12 illustrates the progressive stretching of the HF bond as the
ring builds up to the hexamer level, in concert with the continued contraction of the dis-
tance between F atoms. The correlated data, indicated by the solid curves in Fig. 5.12, show
more of a sensitivity to ring size than do the SCF values, reported as the broken curves. The
data in Table 5.1034,35 reveal the progressively smaller deviation from linearity of each H-
bond as the ring is enlarged. This better arrangement, when coupled to the cooperativity.
leads to the increase in H-bond energy (per H-bond) such that this quantity in the hexamer
is 66% larger than the interaction energy of the fully optimized dimer.
Cooperative Phenomena 249

Figure 5.11 Angular dependence of SCF and MP2 two-body


(solid lines) and three-body (broken lines) terms for (HF)3, with
R(FF) held at 2.73 A. a is defined as the deviation of the HF bond
from the F..F axis, as indicated32.

As in the case of linear, open oligomers, the stretching frequencies of the HF bonds un-
dergo progressive red shift as the number of monomers increases. These shifts are illus-
trated in Fig. 5.13 where they may be seen to surpass 400 cm-1 in the tetramer. The reader
should understand that in a cyclic structure (i.e., n > 2), each molecule acts as both donor
and acceptor, and the symmetry is such that vibrations cannot be distinguished as belong-
ing to one particular molecule. It is also worth pointing out that a third v(FH) stretch is pre-
sent in the tetramer, which at 3571 c m - 1 , amounts to a red shift of 611 c m - 1 from the
monomer.
Just as the aforementioned HF bond stretches are magnified when correlation is in-
cluded, so too are the vibrational frequencies more sensitive to ring enlargement when com-
puted at a correlated level. The MP2 red shifts of the HF stretching frequency in the Cnh HF
oligomers, resulting from each addition of another HF molecule, are approximately double
the SCF values36. Three normal modes were investigated in this work, using a rather large
6-311 + +G(3df,3pd) basis set. vl refers to the symmetric HF stretching mode. It differs
from those reported in Fig. 5.13 in that it represents the totally symmetric combination of
all HF stretching motions, the lowest of such HF stretching frequencies. v2 corresponds to
motion of all atoms toward the center of the ring, and a symmetric bend of all molecules is
characterized by v3.
The behavior of these three frequencies as the ring is enlarged is illustrated in Fig. 5.14.
v1 undergoes the characteristic sharp drop as n is increased, like the other HF stretches in
Fig. 5.13. The enhancement in this sensitivity at the correlated level is evident by a com-
250 Hydrogen Bonding

Figure 5.12 Calculated internal and intermolecular distances in cyclic


(HF)n. Solid curves refer to correlated values with polarized basis set34;
dashed lines to SCF data with 6-311 + +G(2d,p)35.

parison of the solid and broken curves. In contrast to this reduction, v3, the bending mode,
is increased as the ring is enlarged. This increase is enhanced when correlation is included.
The mode corresponding roughly to a "swelling" of the ring, v2, is rather small in magni-
tude and is not strongly affected by the number of molecules in the ring.
As the size of the cyclic oligomers increases, there is progressively greater overlap be-
tween the experimental bands originating from various values of n. This spectral conges-
tion makes it difficult to identify those of a given size oligomer with certainty. This com-
plication fueled speculation that the pentamer might not be found in the gas phase, or might

Table 5.10 Angular deformation of each H-bond and average interaction


energy in cyclic (HF)n (SCF data with 6-311 + +G(2d,p) basis set35, correlated
values34).

(FFH) (degs) — Eelec/n (kcal/mol)

n SCF ACPF SCF ACPF

2 55.8 a
6.0 3.2 5.0a
3 26.5 23.6 4.2 5.1
4 13.2 11.6 5.6 7.2
5 7.8 6.2
6 4.0 2.4 6.4 8.3
a
Optimized dimer, not cyclic structure.
Cooperative Phenomena 25 I

Figure 5.13 Computed harmonic vibrational frequencies in cyclic


(HF)n34. Solid and dashed lines refer to smallest and second small-
est red shifts, respectively, relative to monomer.

exist in only very low proportions, being replaced by tetramers and hexamers. However,
the pentamer has recently been identified by FTIR spectra of continuous and pulsed super-
38
sonic jet expansions of HF37,38 at low temperature. An analysis of the surface generated
from a large number of points calculated at the MP2/DZP level led to the conclusion that
sizable anharmonicities are present in the HF stretching fundamental frequency shifts of
such oligomers but that these changes can be canceled by other errors (BSSE, zero-point
bond-weakening, etc.) in certain cases38.
Although perhaps not global minima, the open, chain-like oligomers were reexamined
in 1994 using large basis sets by Karpfen and Yanovitskii35,39, who considered how the
length of the chain influences the properties of the subunits within. As the chain elongates,
it is found that the internal bond lengths and intermolecular separations of most of the "in-
ternal" molecules, that is, those far from the ends, are fairly uniform, with sharp dropoffs
noted in the last three or four molecules at each end. The r(HF) bonds in the internal mol-
ecules are longer than those of the end molecules by 0.01-0.02 A; the F..F separations are
smaller in the middle of the chain by about 0.1 A.
As the chain elongates with additional HF molecules, there are progressively more HF
stretching frequencies. These frequencies are all red-shifted relative to the monomer, and
these shifts increase as the chain becomes longer. There is always one which is changed by
only a small amount, less than 100 cm - 1 , even for a long chain. The largest shift was noted
for the longest chain, n = 19, which had one HF stretching frequency smaller by 800 c m - 1 ,
relative to the monomer. This compares to a shift of only 42 cm-1 for the highest frequency
when n = 19. Intensities were also computed for the various HF stretching modes. Some
252 Hydrogen Bonding

Figure 5.14 Frequencies of three fully symmetric normal modes of


Cnh (HF)n, computed with a 6-311 + +G(3df,3pd) basis set36. Solid
curves refer to MP2 values and dashed lines to SCF.

salient findings are as follows: The intensity of the lowest-frequency mode, which corre-
sponds to the simultaneous, in-phase motion of all hydrogens, increases the most as n is in-
cremented. The intensity of this mode, for n = 19, is 200 times larger than for the monomer.
Even dividing this quantity by 19 to obtain an enhancement per molecule, one finds an in-
crease of 10 or so. The behavior of the other modes can be modeled by a vibrating string.
That is, the intensities of modes with an odd number of nodes in the phase relation are zero.
For example, when n = 18, the intensity vanishes for modes containing 1, 3, and 9 nodes.

5.4 Water

The cooperativity in aggregates of water is particularly important for attempts to understand


the behavior of the liquid and aqueous solvation. It is in part responsible for the ability of
water to maintain H-bonds up to very high temperatures, probably even above 800 K in su-
percritical water40-42. This importance has motivated numerous attempts to incorporate
nonadditivity into water-water potentials43-51 and, more recently, related liquids such as al-
cohols52,53.
The question of optimal orientation in oligomers of water differs from the (HX)n case
since each water molecule contains two lone pairs and two protons. Based upon the general
principles of cooperativity, it is reasonable to presume that if each water molecule has two
neighbors, it will prefer to act as donor to one and acceptor to the other, as compared to
serving as double donor or double acceptor. A full complement of four neighbors would per-
mit the central water to serve as donor in two H-bonds and acceptor in two others.
Cooperative Phenomena 253

After some earlier studies had yielded uncertainties as to the magnitude of the coopera-
tivity, Clementi et al.54 considered the sensitivity of this quantity in the water trimer to the
type of basis set used. Rather than perform a geometry optimization, the authors assumed
certain particular configurations. Despite the earliness of this work, the authors corrected
some of their results for basis set superposition error. The nonadditivity in most configura-
tions examined was rather small, in most cases less than 0.5 kcal/mol. The authors con-
cluded that minimal basis sets were capable of providing nonadditivities comparable in ac-
curacy to more extended sets, provided the former are well balanced and BSSE is corrected.
As a harbinger of later work, these authors attributed a fraction of the nonadditivity to in-
duction energy. Calculation of the latter quantity, based upon atomic point charges and bond
polarizabilities of the individual molecules, was able to reproduce the ab initio data with
surprising accuracy.

5.4.1 Extended Open Chains


The charge redistributions that accompany formation of a H-bond, and the consequences
for a string of water molecules, were investigated with two rather small basis sets55. By ex-
trapolating the results up to the pentamer level, Scheiner and Nagle found the component
of the dipole moment, in the direction of the chain, is increased by 40-60% as compared to
the moment in the water monomer. Shortly thereafter, Karpfen and Schuster used a differ-
ent formalism wherein infinite chains could be considered explicitly to examine the same
question56. Their means of analysis yielded a reduced enhancement of the moment of only
16%; they attributed the discrepancy to edge effects.
A later work57 continued the investigation of extended chains of water molecules, in-
corporating the effects of electron correlation. As in the oligomers of HF, the length of the
H-bond contracts as the chain enlarges. The small nonlinearity present in the dimer van-
ishes as well. Crystal orbital techniques were employed to consider infinitely extended
chains. Some of the more interesting features of the infinite chain are listed in Table 5.11

Table 5.11 Computed features of infinite chain of water molecules57.

- Eeleca r(OHb)
Basis set (kcal/mol) R(O .. O) (A) r(OH)

SCF
DZ 10.74 2.672 0.973 0.950
DZ(d,p) 6.72 2.837 0.956 0.944
DZ(2d,p) 6.24 2.877 0.953 0.942
TZ(2df,p) 5.82 2.882 0.967 0.941
TZ(2df,2pd) 5.61 2.880 0.959 0.939
TZ(3d2f,3p2d) 5.48 2.881 0.958 0.938
MP2
DZ 11.07 2.619 1.008 0.976
DZ(d,p) 7.89 2.728 1.010 0.963
DZ(2d,p) 7.51 2.734 1.010 0.960
TZ(2df,p) 7.16 2.731 1.004 0.959
TZ(2df,2pd) 6.98 2.742 1.003 0.956
TZ(3d2f,3p2d) 6.84 2.729 1.001 0.954
a
Average for each H-bond, counterpoise corrected.
b
Bridging hydrogen.
254 Hydrogen Bonding

as a function of basis set. There appears to be a clear convergence toward a H-bond energy
of 5.5 kcal/mol, at the SCF level and after counterpoise correction. This value represents an
enhancement of 1.7 kcal/mol or 45%, relative to the dimer, which can be attributed to the
cooperativity in the chain. The interoxygen distance is rapidly approaching an asymptote
of 2.88 A. The O—H bond of the hydrogen involved in the H-bond is longer by 0.02 A than
the other O—H bond. The correlated asymptote for the H-bond energy is somewhat larger,
close to 7.0 kcal/mol. The MP2 interoxygen distance is shorter by 0.15 A; the correlated
R(O..O) of 2.73 A is rather close to the interoxygen separation of 2.74 A in ice. Raising the
level to MP4 would likely increase the calculated distance by 0.01 A, making for even bet-
ter agreement. This accord is surprisingly good in light of the difference between the one-
dimensional chain considered here and the three-dimensional lattice of ice which includes
not only sequential chains in which each molecule is both donor and acceptor, but the
monomers in ice are involved in more than two H-bonds each. The differential between the
two O—H bond lengths in the water molecule is 0.05 A after correlation.
Another work considered finite chains of varying length58. The water molecules were
connected in a uniform sequential fashion as illustrated in Fig. 5.15. All interoxygen dis-
tances were taken as 2.84 A, chosen to mimic the liquid. The internal geometry of each mol-
ecule was frozen in its experimental structure. The basis set employed was DZP, with MP2
evaluation of correlation. Since the geometries were not optimized, normal modes could
not be calculated. Instead, the workers focused on what they termed "uncoupled OH stretch-
ing frequencies," derived from pointwise computation of the energy for the system with dif-
ferent OH bond lengths, holding fixed the remainder of the geometry of the chain. The har-
monic frequency was then evaluated as the second derivative of this potential curve. A
variational solution of the Schrodinger equation for OH oscillator motion enabled a calcu-
lation of the anharmonic frequency.
Some of the more interesting computations of the multibody interaction energies in the
chain are reported in Table 5.12. With regard first to the two-body interactions in the first
five rows, these terms die off fairly quickly with distance. The 1-4 interaction is mildly re-
pulsive. Moving on to trimers in the next five rows, the entries in the two-body column of
Table 5.12 refer to the sum of all three pairwise interaction energies. The 3-body interac-
tion represents less than 10% of the total binding energy in the consecutive 1-2-3 trimer.
It is much smaller in those trimers containing a "gap," for example, 1-2-4 or 1-2-5,
amounting to about —0.04 kcal/mol in these cases. Turning now to the tetramers, the four-
body interaction energies are much smaller still, the largest of them being only —0.06 kcal/
mol for the consecutive tetramer with no gaps; higher-order terms such as 5-body terms can
be safely ignored.
The highest nonadditivity, defined by these authors as the difference between the total
interaction energy and that computed by summing pairwise interactions, is observed for the
full heptamer, where it amounts to 16% of the total binding energy. Fig. 5.16 illustrates

Figure 5.15 Arrangement of water molecules in sequential


chain58.
Table 5.12 Multibody interaction energies (kcal/mol) in the sequential open
oligomers of water. Calculated at MP2/DZP level58.

Molecules" 2-body 3-body 4-body 5-body

1-2 -4.62
1-3 -0.81
1-4 0.03
1-5 -0.12
1-6 -0.01
1-2-3 -10.16 -1.03
1-2-4 -5.41 -0.05
1-2-5 -4.72 -0.03
1-3-4 -5.50 -0.04
1-4-5 -4.81 -0.04
1-2-3-4 -15.66 -2.15 -0.06
1-2-3-5 -11.06 -1.11 -0.00
1-2-4-5 -10.22 -0.17 -0.01
1-2-3-4-5 -24.77 -21.28 -3.34 -0.00
a
SeeFig. 5.15 for numbering scheme.

Figure 5.16 Percentage contribution of nonadditivity to total


binding energies of linear and ring oligomers of water58.
256 Hydrogen Bonding

the percentage contribution of nonadditivity to the total binding energy of the linear chains
of water. The solid line denotes a chain like 1-2-3-4 where there are no gaps present, as
compared to the broken line which corresponds to a chain containing one gap, for example,
1-2-4 or 1-2-3-5. Note that the greatest nonadditivity is present in the consecutive chains
and that the gap significantly reduces the amount of nonadditivity present. The percentage
contribution of nonadditivity appears to level off for the longer chains, and one might pro-
ject it to asymptotically approach something just less than 20% for an infinitely long chain.
For ring oligomers, on the other hand, the degree of nonadditivity shows no sign of level-
ing off, for n as large as 5. So cooperativity may be said to be appreciably greater in the
rings than in the chains for n > 3.
The effects upon the computed OH stretching frequencies of lengthening the chain are
illustrated in Fig. 5.17. The harmonic frequencies are illustrated by the broken lines and
anharmonic v by solid. Molecule "1" refers to the first in the chain. It shows a red shift of
100-150 c m - 1 for the dimer, and a large further shift of about 70 c m - 1 when a third mol-
ecule is added. Chain elongation beyond this point results in only small further red shifts.
Molecules ensconsed in the middle of the chain show the largest red shifts of all. Just as for
the terminal molecule, these shifts continue to climb, albeit slowly, as the chain elongates.
It is important to stress the very parallel behavior of the harmonic and anharmonic data,
suggesting trends can be accurately deduced from study of the harmonic data. Effects are
even stronger in the pentamer ring. The red shift of the anharmonic frequency is computed
to be 377 cm - 1 , with a corresponding harmonic shift of 321 c m - 1 .

Figure 5.17 Calculated frequency shifts of OH stretches, involv-


ing the bridging hydrogen. Solid lines refer to anharmonic and
broken to harmonic frequencies. The first molecule in the chain is
designated "1," and the central molecule corresponds to that with
largest frequency near center of chain 5 8 .
Cooperative Phenomena 257

5.4.2 Branching Clusters


The observation that water molecules are about 0.2 A closer together in ice than in the gas-
phase dimer has motivated a number of theoretical inquiries. Yoon et al.59 considered how
the R(O .. O) shrinkage might be tied together with stretches in r(OH). Their polarized basis
set was satisfactory in yielding both a low BSSE and a good moment for the monomer. The
three-body terms computed here were considerably larger than those obtained earlier by
Clementi et al.54, albeit for different configurations: Yoon et al. limited their analysis to
geometries corresponding to the structure of ice. An optimization of the water cluster geom-
etry using the nonadditivity term approached within 0.06 A of the interoxygen distance in
ice; the shrinkage relative to the dimer was thus attributed mainly to the three-body term.
The cooperativity involved in a single water molecule, surrounded by four others in a
tetrahedral arrangement, was investigated at the SCF level60, with all interoxygen distances
set to 2.90 A. The central molecule in the pentamer acts therefore as simultaneously dou-
ble proton donor and double acceptor. There are various subset trimers of this full complex:
one with the central molecule as double donor, one as double acceptor, and several where
the central molecule is both donor and acceptor; that is to say, "sequential." As expected,
the three-body interaction energies of the double-donor and double-acceptor trimers are
positive (destabilizing), but negative values are obtained for the sequential types. The mag-
nitudes of these terms vary between 0.5 and 1.0 kcal/mol, as compared to two-body terms
between 3 and 4 kcal/mol; four-body terms are all much smaller, less than 0.03 kcal/mol.
Because of large-scale cancellation between the various three-body terms, which are both
positive and negative, the total of all these terms accounts for only about 5% of the sum of
two-body interaction energies (most of which are negative).
The author60 also considered the charge shifts occurring as a result of pairwise and
higher-order interactions. In comparison to changes in Mulliken charges of the central mol-
ecule caused by dimerization which are of the order of 0.005-0.030 e, the (3-body) nonad-
ditivity in the shifts amounts to less than 0.010, more commonly 0.002; four-body nonad-
ditivities are typically less than 0.001. The authors conclude that the additive contributions
to the electron redistributions are responsible for the bulk of the nonadditivity in the energy.

5.4.3 Cyclic Oligomers


The importance of electron correlation to the cooperativity in a cyclic tetramer was inves-
tigated61 in 1987. The choice of system was based on its experimental detection62,63. The
geometry was fully optimized via gradient algorithms, and superposition error was cor-
rected by the Boys-Bernardi counterpoise procedure. The formation of the cyclic tetramer
stretches the donor OH bonds by three times more than the stretch in the dimer, and R(O..O)
is shorter by 0.15 A. With regard to energetics, the per-bond H-bond energy of the tetramer
is about 27% larger than the dimer interaction energy. This enhancement is virtually un-
changed when correlation is included. The authors attribute the majority of this cooperative
effect to three-body interactions, with the remainder due to two-body terms such as mutual
polarization.

5.4.3.1 Vibrational Spectra


The vibrational modes of clusters of water molecules were emphasized in a study up to the
tetramer level64, motivated by IR and coherent anti-Stokes Raman spectroscopic data. Cal-
258 Hydrogen Bonding

culations were restricted to the SCF level, using primarily the 4-31G and 6-31G* basis sets.
Optimization of the cyclic trimer with the larger basis set yielded an isosceles triangle of
the three O atoms: one of the R(O..O) is slightly longer than the other two. Its cyclic char-
acter is consistent with experimental observation62. A normal mode analysis was carried
out within the harmonic approximation and yields what the authors refer to as "delocaliza-
tion of intramolecular modes" wherein all three molecules participate in each normal mode.
The v1 and V2 modes of the monomer can be recognized within the trimer, but each of these
monomer modes appears as a set of three vibrations, each with participation from all three
molecules. The exception is v3 which remains localized in the trimer. The bending modes
are closely related to v2 in H2O while the vibrational amplitudes of the bridging and free
O—H stretching coordinates are quite unequal in the stretching modes of the trimer. The
stretching modes which place more emphasis on the H-bonding hydrogens are red-shifted
versus the monomer by 82-130 c m - 1 , while the shift to lower frequency in the other set of
stretches is 37-42 c m - 1 . In contrast, the bending modes in the trimer are higher than that
in the monomer by 15-37 c m - 1 . The tetramer is more symmetric than the trimer, being very
close to S4. All of its intramolecular modes are highly delocalized, in contrast to the trimer
which had varying degrees of this phenomenon. This work was continued several years later
with a study of the effects of deuterium substitution on the vibrational modes of (D2O)365.
The size of the ring was extended to the pentamer, again limited to the basis sets men-
tioned above66. The ring was found to be very close to planar, except for the nonbridging
hydrogens. Two of these lie on one side of the ring, and three on the other, with a nonpla-
narity of roughly 45°. Each H-bond is very nearly linear, at least within 2-3°. The pentamer
is "floppier" than the trimer and tetramer, with a very low-frequency 21 c m - 1 librational
mode. The same study considered also (H2O)8, and found a quasicubical D2d structure.

5.4.3.2 Energy Components


The size of the cyclic cluster under investigation was increased to six by White and David-
son67. Because of their underlying interest in ice, the authors took the O..O distance as 2.75
A, rather than the larger separation in the gas phase; r(OH) was fixed at 0.99 A. The struc-
ture examined was not of the classic type where all six molecules act as both donor and ac-
ceptor: one molecule (#5) serves as double donor and another (#4) as double acceptor, as
illustrated in Fig. 5.18.
The two- and three-body interaction energies in the water hexamer were decomposed
via the Morokuma procedure, without counterpoise correction, and some of the results are
listed in Table 5.13. Beginning with the two-body terms, the results for all adjacent mole-
cules are identical to the data for the 1-2 pair in the first row of the table. This similarity is
explained by the fact that all adjacent pairs constitute a single H-bond; the concept of dou-

Figure 5.18 Proton donor and acceptor characteristics


in water hexamer67.
Cooperative Phenomena 259

Table 5.13 Morokuma components (kcal/mol) of two and three-body interactions computed
for the water hexamer illustrated in Fig. 5.1867.

Eelec ES EX POL CT

total in hexamer -23.4 -81.0 86.9 -10.9 -18.5


2-body terms
1-2 -2.8 -13.1 14.5 -1.4 -2.8
1-3 -1.2 -1.1 0.0 0.0 0.0
2-4 -0.6 -0.6 0.0 0.0 0.0
3-5 0.8 0.8 0.0 0.0 0.0
4-6 0.4 0.4 0.0 0.0 0.0
1-5 -0.6 -0.6 0.0 0.0 0.0
1-4 0.2 0.2 0.0 0.0 0.0
2-5 0.2 0.2 0.0 0.0 0.0
3-6 -0.4 -0.4 0.0 0.0 0.0
SUM ( E2) -19.2 -81.0 87.0 -8.4 -16.8
3 -body termsa
1-2-3 -1.4 0.0 0.0 -0.8 -0.6
2-3-4 -1.0 0.0 0.0 -0.6 -0.4
3-4-5 0.8 0.0 -0.1 0.5 0.4
4-5-6 1.2 0.0 0.2 0.4 0.5
1-3-5 0.0 0.0 0.0 0.0 0.0
2-4-6 0.0 0.0 0.0 0.0 0.0
1-2-4 -0.1 0.0 0.0 0.0 0.0
1-3-4 -0.1 0.0 0.0 -0.1 0.0
1-4-5 0.2 0.0 0.0 0.1 0.1
2-5-6 -0.2 0.0 0.0 -0.1 0.0
SUM ( 3E
) -3.8 0.0 -0.1 -2.2 -1.4
Z2 E+ E3 -23.0 -81.0 86.9 -10.6 -18.3
a
Not all terms are listed in the table.

ble donor or acceptor is only meaningful within the context of three or more molecules. The
interaction energy amounts to —2.8 kcal/mol. The attractive electrostatic energy is canceled
by the exchange repulsion; polarization and charge transfer energies are both attractive. For
all nonadjacent pairs, the EX, POL, and CT terms are quite small, leaving only electrosta-
tic energy in the pairwise interactions. Because of the long-range character of the ES term,
there are significant contributions even from molecules on opposite ends of the ring; for ex-
ample, 1-4 or 3-6. The signs of these nonadjacent pairwise electrostatic energies can be un-
derstood on the basis of the orientations of the particular molecules. For example, mole-
cules 3 and 5 have H atoms pointed at one another, leading to the repulsive 3-5 term. When
summed together, the two-body terms amount to —19.2 kcal/mol, less attractive than the
full interaction energy in the hexamer by 4 kcal/mol. With respect to the individual com-
ponents, the electrostatic term is by nature fully additive, so the sum of two-body terms is
equal to the full ES energy of the hexamer. The exchange is very nearly additive, with a dis-
crepancy of only 0.1 kcal/mol. The sum of two-body polarization and charge transfer com-
ponents are each about 2 kcal/mol less attractive than the full components in the hexamer.
The three-body terms are listed in the lower part of Table 5.13. The first several entries
represent triplets of consecutive molecules around the ring. It is notable that these inter-
actions can be of either sign. Repulsive terms are associated with triplets like 3-4-5 and
260 Hydrogen Bonding

4-5-6 that contain either a double-donor or double-acceptor; others are attractive. The ES
contributions are identically zero and the exchange components are quite small. The three-
body terms are composed of similar amounts of polarization and charge transfer compo-
nents. It is worth noting that some of the three-body terms, for instance, 1-2-3, are of larger
magnitude than certain pairwise interactions, particularly those between nonadjacent pairs.
Nonconsecutive triplets may contain no adjacent pairs, as in 2-4-6, or one adjacent pair, as
in 1-2-4. In the former case, the three-body energies are less than 0.1 kcal/mol; the latter
are all less than 0.2 kcal/mol.
The total of all three-body interactions is —3.8 kcal/mol, as compared to —19.2 kcal/mol
for the sum of all two-body interactions. When added together, the total of all pairwise and
three-body interactions comes within 0.4 kcal/mol of the total interaction energy of —23.4
kcal/mol in the hexamer. With respect to the individual components, there is very little non-
additivity in ES or EX. The total nonadditivity of some 4 kcal/mol is approximately equally
divided between POL and CT.

5.4.3.3 Anisotropy of Energy Components


The nature of the cooperativity in the water trimer was dissected by perturbation theory,
coupled with extended basis sets, in 199168. Rather than a full geometry optimization, the
authors assumed an equilateral triangle C3h arrangement, with all atoms in a common plane,
as illustrated in Fig. 5.19. A basis set, especially designed for molecular interactions was
employed, with a [5s3pld/3slp] contraction, and counterpoise corrections were applied at
all stages.
The equilibrium geometry of this cyclic C3h arrangement has interoxygen separations
of 3.0 A, with an orientation angle a of 75°. Fig. 5.20 emphasizes that the optimal angle for
this trimer is controlled by pairwise forces present in the SCF wave function. The
anisotropies of three-body effects, at either SCF or MP2 levels, or of the two-body MP2 in-
teractions, are dwarfed by the strong angular dependence of two-body SCF terms. The sum
of SCF pairwise energies is also much larger in magnitude than the latter terms which are
all less than 2 kcal/mol.
We focus on the three-body forces in Fig. 5.21 where the SCF portion is compared with
its components extracted before (Heitler-London, HL) and after (deformation, def) the wave
functions of the monomers are perturbed by one another. It is apparent that the latter SCF-
def forces are largely responsible for the anisotropy of the SCF three-body terms. The
Heitler-London component is weaker, and resembles a mirror of SCF-def, with a maximum
where SCF-def contains a minimum. The extremum at 20° corresponds to a configuration

Figure 5.19 Planar C 3h geometry of water trimer 68 . a


refers to the angle between O..O axis and HOH bisector.
Cooperative Phenomena 261

Figure 5.20 Anisotropy of two and three-body interaction ener-


gies in the water trimer, at SCF and MP2 levels68. See Fig. 5.19
for definition of a.

where one H atom of each water molecule is pointing toward the center of the equilateral
triangle, a so-called H-to-H geometry. It is thus not surprising to see a maximum in ex-
change repulsion-type forces at this angle. One can conclude that HL acts to ameliorate the
deformation effects at the SCF level.
It is instructive to compare the SCF three-body interaction, represented by the solid
curve in Fig. 5.21, with the induction energy, with which there is a tendency to approximate
it in the literature. In this context, it should be noted that the induction curve is far too at-
tractive, by a factor of more than 2 in the vicinity of 20°. Other characteristics of its shape
differ from the full SCF curve or deformation energy in Fig. 5.21 as well. The three-body
forces at the correlated MP2 level are very small in magnitude, and insensitive to angular
characteristics of the trimer. Most of these conclusions have been verified by later calcula-
tions69 and by symmetry-adapted perturbation theory calculations, although there were a
number of discepancies as well33. The issue is not entirely closed.

5.4.3.4 Comparison with Open Trimers


In addition to the closed cyclic trimer, the open trimers displayed in Fig. 5.22 in which the
central molecule acts as (a) simultaneous donor-acceptor, d-a, (b) double donor, d-d, and
(c) double acceptor, a-a were considered as well68. In the first case, one would expect pos-
itive cooperativity, which is confirmed by an attractive total three-body interaction energy
equal to about 10% of the two-body contribution. Most of this three-body term is due to the
SCF-deformation, as in the cyclic structure. The double-donor is bound by only about 1/10
262 Hydrogen Bonding

Figure 5.21 Anisotropy of various contributions to the 3-body


forces for the cyclic water trimer68. See Fig. 5.19 for definition
of . HL refers to Heitler-London term which prohibits modifi-
cation of the charge clouds of each molecule in the presence of
the others, and SCF-def to the result of such deformation, both at
the SCF level. Three-body induction is computed directly via
perturbation theory.

the strength of the donor-acceptor trimer. This weaker total interaction is due in large mea-
sure to the pairwise interaction between the two terminal molecules, which are oriented with
their O atoms pointing directly toward one another. The three-body term is attractive but
only very slightly. The double-acceptor trimer is more strongly bound than the double
donor, with about half the total binding energy of d-a. The three-body term here is a repul-
sive one, and is roughly 10% of the magnitude of the cumulative two-body terms. This re-
pulsive character is due chiefly to the deformation of the wave functions.
It is concluded that total nonadditive effects are dominated by SCF terms, which are di-
rectly attributed to electric polarization. However, since the polarization is constrained by the
Pauli exclusion principle, classical models which neglect exchange phenomena may incur cer-
tain errors, especially in regions of strong overlap between electron clouds of the monomers.
Correlation contributions to nonadditivity do appear small enough that they can be safely ig-
nored, with efforts better concentrated on an accurate portrayal of the SCF phenomena.

5.4.4 Identification of True Minima


The structures investigated for the water trimer up to this point have been idealized, with
no guarantee that any are true minima on the full potential energy surface (PES). There are
Cooperative Phenomena 263

Figure 5.22 Three open trimer geome-


tries68.

many more possibilities to be considered. Since each water molecule contains two hydro-
gens and two O lone pairs, it is possible in principle for it to participate in up to four inter-
actions: as a double proton donor and/or double acceptor, all simultaneously. A diverse col-
lection of trimer configurations, taking into account these various possibilities, was
examined in 199270, using extended polarized basis sets, up to 6-311 + + G(2df,2p). Geome-
tries were optimized and harmonic frequencies obtained. While counterpoise corrections
were introduced, the authors decided that since these were of the approximate magnitude
0.5 — 0.7 kcal/mol (1.2 — 1.4 at the MP4 level) for all configurations examined, they were
unimportant for purposes of comparative stability.
The most stable geometry found is a cyclic one like that in Fig. 5.19, except that the non-
bridging hydrogens do not lie in the plane. The cyclic character is consistent with observa-
tions in the gas phase62,71 and in Ar and Kr matrices72. (Indeed, the hydroxyl groups in the
related phenol molecules will also form a cyclic trimer in gas-phase supersonic jets73.) This
complex is bound, relative to three isolated water molecules, by 12.0 kcal/mol at the SCF/6-
311 + +G(2df,2p) level, which is increased by 3.7 kcal/mol for MP2. After including MP4
and zero-point vibrational corrections, the total binding energy becomes 10.8 kcal/mol. This
represents an increase of 1.1 kcal/mol, relative to three times the interaction energy of an
optimized dimer.
A sequential dimer of the d-a type in Fig. 5.22 is not a true minimum on the full PES of
the trimer. Optimizations beginning with such a structure bring the two terminal molecules
closer together until the trimer closes up to the cyclic arrangement. The d-d structure in Fig.
5.22 does in fact represent a minimum on the PES. Its total binding energy is less than that
of the cyclic trimer by 5-6 kcal/mol. In contrast, the double acceptor geometry a-a in Fig.
5.22 is not a minimum at all, but rather a second-order saddle point. On the other hand, Ihe
264 Hydrogen Bonding

double-acceptor does represent a minimum when the two terminal molecules approach one
another to form a cyclic trimer as indicated in Fig. 5.23. This geometry is predicted to be
very slightly (1 kcal/mol) more stable than the open double-donor chain.
The structure of the cyclic water trimer was optimized also using correlation-consistent
polarized basis sets, augmented with additional diffuse functions74. Whereas the three
R(O..O) distances are all different at the SCF level, MP2 correlation brings them within
0.002 A of being equal to one another. The three H-bonds are all nonlinear by about 30°,
that is, (OH..O)~150°.
Replacement of one of the water molecules by phenol causes no fundamental changes
in the geometry of the trimer75. Of course, the three R(O..O) distances are now all clearly
different, varying between 2.85 and 2.98 A at the SCF/6-31 + +G(d,p) level. The shortest,
and presumably strongest, H-bond is that where the phenol acts as the proton donor mole-
cule. Deviations from nonlinearity are also affected by the phenyl substitution, but to only
a minor degree. The phenyl ring is rotated about 120° from the plane of the three O atoms.
Restricting calculations to the SCF level, and with counterpoise correction, the 6-31G(d,p)
interaction energy (- Eelec) of the entire trimer is 16.3 kcal/mol, as compared to 15.5
kcal/mol for the water trimer at the same level of theory. The enhanced binding introduced
by the phenol molecule carries over when ZPE corrections are added, yielding a DO value
of 11.5, as compared to 9.7 kcal/mol for the water trimer.
The positions of the three nonbridging hydrogens of the water trimer were the focus of
more recent work. A potential energy surface was constructed as a function of the out-of-
plane bending angles of these three atoms76. The energetics arise from MP2 level compu-
tations, with counterpoise corrections, using a singly polarized basis set, that includes func-
tions centered on each H-bond. The structure with all of the nonbridging hydrogens on one
side of the plane appears to be a minimum in the surface, albeit higher in energy than the
global minimum in which one hydrogen is on the side of the plane opposite from the other
two76. An all-planar structure, with all nonbridging hydrogens in the plane of the ring, is
described as a third-order stationary point on the PES. After fitting the ab initio data to an
analytical functional form, it was found77 that the lowest torsional levels of the non-bridg-
ing hydrogens lie well above the torsional barrier of 0.28 kcal/mol. This observation leads
to the conclusion that the corresponding nuclear wave functions are fully delocalized
amongst the six symmetry-related minima.
The tetramer is also cyclic74, as illustrated in Fig. 5.24, and with a greater degree of sym-
metry than the trimer, belonging to the S4 point group. The ring is sequential in the sense
that each molecule acts as simultaneous donor and acceptor. This type of structure has been
confirmed by analysis of the far-IR vibration-rotation tunneling spectrum78. The correlated
R(O .. O) distances are all equal to 2.743 A, about 0.06 A shorter than in the trimer. (The ex-

Figure 5.23 Cyclic water trimer containing a double-


acceptor molecule70.
Cooperative Phenomena 265

Figure 5.24 Cyclic water tetramer with S4. symmetry.

perimental estimate78 is 2.79 A in the tetramer.) The larger ring permits less nonlinearity in
the four H-bonds as 6 (OH..O) rises to 168°. A later work79 found that inclusion of electron
correlation is mandatory if one wishes to study the torsional motions of the hydrogens in
the tetramer. Fortunately, the results indicated that correlation effects beyond second-order
are unlikely to make significant contributions to the tetramer, consonant with findings for
the dimer and trimer.
The global minimum, with two hydrogens above the molecular plane and two below, is
more "classical" than the trimer, in the sense that the torsional zero-point vibrational level
lies below the saddle points for interconversion79. The barriers to torsional motions of the
nonbridging hydrogens are about 1.2 kcal/mol. These results were largely confirmed by
later optimizations of the trimer and tetramer with a variety of polarized basis sets80 where
the 0.16 A reduction in H-bond length on going from trimer to tetramer was stressed. The
VRT measurements78,81 support the notions advanced here concerning tunneling from one
minimum to a symmetrically related one.
Just as in the case of the trimer75, replacement of one of the water molecules of the cyclic
tetramer by the hydroxyl-bearing phenol molecule yielded much the same structure82. Of
course the S4 symmetry classification no longer applies. The tendency of the phenyl ring to
make its hydroxyl more acidic is reflected in the four R(O..O) distances. The shortest (2.784
A at the SCF/6-31G** level) corresponds to that in which the phenol acts as proton donor,
and the longest (2.900 A) when phenol is the acceptor. The combination of phenol as a bet-
ter proton donor, but poorer acceptor than water, leads to little net effect on the binding en-
ergy of this complex. With correlation included along with BSSE corrections, the binding
energy De is computed to be quite similar to that of the water tetramer (0.4 kcal/mol
stronger), although this difference increases to 1.7 kcal/mol when zero-point vibrations are
considered.
It was not possible74 to optimize the larger oligomers at the correlated level, so Xanth-
eas and Dunning reported SCF optimizations instead. Both the pentamer and hexamer were
found to be cyclic, sequential rings, in the same spirit as (H2O)3 and (H2O)4. There is some
puckering of the ring of the five O atoms in the pentamer, with two of them lying within
about 20° of the plane of the other three. This sort of geometry in the pentamer is supported
by recent VRT tunneling spectral data83. (In an interesting aside, crystal diffraction data
suggest that water molecules may assemble in cyclic pentamers in hydrophobic clefts of
proteins84.) The hexamer belongs to the S6 point group, and has the general structure of
chair cyclohexane. There is virtually no angular distortion of the H-bonds which are within
2.5° of linearity.
266 Hydrogen Bonding

When the number of water molecules grows beyond six, the situation begins to change.
In terms of energy alone, a di-bipyramid shape of the heptamer is computed with a 6-
311G* * basis set to be some 4 kcal/mol more stable than a cyclic geometry85. Consequently,
it is the former structure that would be observed at 0 K. However, the entropy of the di-
bipyramid is quite a bit less than that of the seven-membered cycle. As a result, the cyclic
geometry is favored with respect to free energy by 2.6 kcal/mol at room temperature.
This situation is echoed in the octamer where the cyclic structure is superceded in sta-
bility by a geometry in which the oxygen atoms occupy the corners of a cube86. This sort
of arrangement permits each molecule to be involved in three, rather than two, separate H-
bonds. Calculations at the SCF/6-311G** level find the cube to be more stable than an eight-
membered ring by 13 kcal/mol. This result confirms an earlier finding with the smaller 4-
31G basis set66. On the other hand, the latter ring has a very high entropy so that its free
energy is within about 0.2 kcal/mol of the cube at 298 K86. This trend is confirmed when
correlation is added to the calculations by the MP2 method and the results are corrected for
superposition error87. In fact, this particular set of correlated computations indicates the
cyclic structure is unquestionably of lower free energy at room temperature. Other geome-
tries that lie within about 10 kcal/mol of the global minimum cube can be characterized as
five- (or six-) membered clusters, with three (or two) additional molecules, and a pair of
four-membered rings, connected side to face. Altogether this series of calculations identi-
fied 29 separate minima on the PES of the water octamer86.

5.4.4.1 Vibrational Spectra


Xantheas and Dunning74 computed the vibrational spectra of these water clusters in some
detail, providing not only the frequencies at both the SCF and MP2 levels, but also IR in-
tensities for each and Raman activities. Table 5.14 lists the shifts that occur in the in-
tramolecular frequencies, relative to the monomer, and all computed at the same correlated
MP2 level with the same basis set. Considering first the bending modes, the increase in this
frequency with each progressive enlargement of the oligomer is apparent. Smaller blue
shifts are observed in the stretch of the free O—H. More dramatic than these are the red
shifts in the stretching frequencies of the bridging O—H group, which diminish by several
hundred cm-1 as each new water molecule is added. The relationship between the latter
shifts and the optimized length of the corresponding O—H bonds is very close to linear,
whether at the correlated or SCF levels, conforming to the Badger rule88,89. The trends in
the computed data were later confirmed by measurements of the spectra in the gas phase90.
The harmonic intermolecular modes are reported in Table 5.15. As for the intramolec-
ular modes, the MP2 frequencies are uniformly higher than the SCF values. In the case of

Table 5.14 Shifts in harmonic intramolecular vibrational frequencies (cm - 1 ) of cyclic


oligomers of water, relative to the monomera, computed at MP2/aug-cc-pVDZ level74.
n Bends Free O—H stretch Bridging O—H stretch

2 1,28 57,35 -73, -160


3 9, 12, 37 26, 24, 20 -231, -240, -299
4 14,30,30,60 15, 15, 15, 14 -350, -388, -388, -481
a
Frequencies in monomer are 1623, 3807, 3936, respectively. O—H stretches are referenced to average of two O — H stretch-
ing frequencies in monomer.
Cooperative Phenomena 267

Table 5.15 Harmonic intermolecular vibrational frequencies of cyclic oligomers of water


(cm - 1 ) computed at MP2/aug-cc-pVDZ level74.
n

2 141 147 155 185 342 632


3 158 173 185 193 218 235 343 351 444 571 667 863
4 51 79 200 211 237 237 255 255 261 291 403 435 451
451 754 826 826 996

the dimer, the two higher frequencies correspond to bending motions of the monomers, fol-
lowed by the intermolecular stretch at 185 cm - 1 . The authors were not clear about the types
of motions that correspond to the frequencies listed for the trimer or tetramer.
A later set of computations69 was more explicit about identification of the various modes
in the trimer. Frequencies of "flipping" motions of the three nonbridging hydrogens were
in the 200-270 cm-1 range. Intermolecular stretches were somewhat lower, 170-220 c m - 1 .
The remaining intermolecular modes correspond to librations and vary from 350 to 830
c m - 1 , with the highest of in-plane type.

5.4.4.2 Many-Body Contributions


An update of this work91 emphasized the many-body contributions to the binding energies.
MP4 results are very close to MP2, as evident in Table 5.16. The three-body term contributes
some 12% of the total binding energy of the trimer at the SCF level, which increases to 17%
at MP4. The results for the tetramer indicate the four-body term is quite small, on the order
of —0.5 kcal/mol. One would hence introduce only a 2% error by ignoring this term com-
pletely. The combined three-body terms make up some 25% of the total binding energy of
the tetramer. Correlation seems to nearly double the three and four-body terms, but has a
much smaller enhancing effect upon the pairwise interaction. The total binding energy of
the tetramer is 23.8 kcal/mol, the first reported correlated value for this quantity.
Another manner of viewing the cooperativity is through the progressive increase in the
average binding energy of the oligomers. Table 5.17 reports the average H-bond energy as

Table 5.16 Total multibody energetics (kcal/mol) for cyclic water oligomers
with aug-cc-pVDZ basis set91.
2-body 3-body 4-body

dimer SCF -3.71


MP2 -4.42
MP4 -4.36
trimer SCF -9.75 -1.38
MP2 -11.80 -2.45
MP4 -11.57 -2.40
tetramer SCF -16.04 -3.52 -0.22
MP2 -18.55 -6.24 -0.54
MP4 -17.97 -6.16 -0.56
268 Hydrogen Bonding

Table 5.17 Average binding energy per molecule, — Eelec/n,


(kcal/mol) in water oligomers with aug-cc-pVDZ basis set91.

SCF MP2 MP4

dimer 1.86 2.21 2.18


trimer 3.67 4.62 4.55
tetramer 4.88 6.08 5.95
pentamer 5.19
hexamer 5.40

(- Eelec/n) at various levels. The data clearly indicate the cooperativity as the binding en-
ergy rises from 1.9 kcal/mol (at the SCF level) for the dimer up toward 5.4 as the number
of molecules reaches six. (However, the data for the dimer may be misleading as the com-
plex is not cyclic and so contains only a single H-bond.) This study concludes that two and
three-body interactions can provide most of the total interaction. Correlation is recom-
mended, but the MP2 level appears satisfactory.

5.4.4.3 Higher Level Searches for Minima


The theoretical treatment of the cyclic trimer was improved when BSSE corrections were
added to the geometry optimizations and to the harmonic vibrational frequencies92. The
work was not truly a full optimization in that each molecule was frozen in its experimental
monomer geometry. The three O atoms were presumed to occupy the vertices of an equi-
lateral triangle, with the bridging hydrogens lying in this O—O—O plane. In this manner,
the optimization was reduced to only three parameters: R(O..O), (HO..O), and the dihe-
dral angle of rotation of the nonbonding hydrogens out of the plane, . The optimal value
for the latter out-of-plane angle was found to be 48°. This value differs by only 2° from what
it would be in the absence of cooperativity. It deviates from a previously optimized value
of 30°70 which may be a result of the counterpoise corrections included here. A 0.07-0.09
A shortening of the interoxygen separation was attributed to cooperativity, but another fac-
tor is in operation. Specifically, "bent" H-bonds tend to have shorter intermolecular dis-
tances, so some of the contraction can be associated with the nonlinearity imposed on each
H-bond by formation of a cyclic trimer.
The energy of interaction in the trimer was found in this work to be —14.7 kcal/mol, 0.6
kcal/mol stronger than three times the binding energy of the dimer. It should again be em-
phasized that this small net enhancement is really indicative of a larger cooperativity since
each of the three H-bonds in the trimer is significantly distorted from linearity, in this case
by 20°. The nonadditivity in the trimer amounts to —2.0 kcal/mol, perhaps a better indica-
tor of cooperativity here. When zero-point vibrations were included in the analysis, the dis-
sociation energy of the trimer, Do, came to 10.2 kcal/mol, larger than three times that of the
dimer by 2.4 kcal/mol.
In comparison to the magnitude of the effects of nonadditivity upon the energetics in the
trimer, much larger changes are observed in the shifts of the O—H stretching frequencies,
vOH, which are 70% larger than the dimer value 92 . The major component in the nonaddi-
tivity was traced to the SCF deformation, consistent with prior calculations. Whereas this
effect is a relatively minor contributor to the total interaction energy, it dominates the fre-
Cooperative Phenomena 269

quency shift. Another important finding is the high sensitivity of frequency shift to the in-
termolecular distance (and out-of-plane angle as well). The authors reason that the differ-
ences in calculated frequency shifts from one theoretical study to another may be largely
due to differences in R(O..O) used in these works. The reader is thus cautioned to carefully
consider intermolecular distances when comparing theoretical calculations of such fre-
quency shifts.
One of the highest level searches of the potential energy surface of the water trimer to
date added diffuse functions to a DZP basis set, and included correlation via single and dou-
ble excitation coupled cluster (CCSD)93. At this level of theory, the structure is that pic-
tured in Fig. 5.25, with two non-bridging hydrogen atoms above the plane of the oxygens
and one below. The three R(O..O) distances are not quite equivalent, being equal to 2.825,
2.828, and 2.837 A. Each of the H-bonds is significantly distorted from linearity, with
(OH..O) in the range 147-150°. The non-bridging hydrogens lie out of the plane by some
40-50°. The geometry obtained by the CCSD treatment of correlation is very similar indeed
to the MP2 results earlier74. Forcing full planarity of the cyclic trimer, and a C3h geometry,
results in a stationary point of index 3, that is, it is not a minimum and there are three imag-
inary frequencies.
The harmonic vibrational frequencies computed with this basis set are listed in Table
5.18, along with the change resulting from adding correlation effects. There is a clear pat-
tern in that the nine higher frequencies, representing the intramolecular modes, above 1000
cm-1 are all diminished by correlation, and the twelve intermolecular frequencies are
raised. The lowering of the intramolecular stretching frequencies are of the order of 6-9%,
with slightly smaller increases in the three intramolecular bending modes, 3-5%. The in-
crements in the intermolecular frequencies caused by correlation are larger, generally
10-20%, but varying all the way up to 39% for the lowest frequency. The lowest of these
intermolecular frequencies is some 170 c m - 1 , about double that of a band that has recently
been observed at 87 cm-1 94.
Another recent set of calculations69 confirmed the structure of the minimum, using a
modification of the standard MP2 correlation procedure, and with counterpoise corrections
applied. Stationary points were located at the MP2/6-311 + +G(d,p) level, and vibrational
frequencies determined. While the global minimum has two of the hydrogens above and
one below the plane of the three O atoms, a secondary minimum occurs when all three hy-
drogens are on the same side of the plane. This so-called "crown" configuration is 0.8
kcal/mol less stable than the global minimum. A barrier of 0.2 kcal/mol must be crossed to
transit from the global to the secondary minimum. A configuration with all nonbridging hy-
drogens lying in the molecular plane lies 1.2 kcal/mol above the global minimum.

Figure 5.2.5 Equilibrium geometry of water trimer,


belonging to C1 point group93.
270 Hydrogen Bonding

Table 5.18 Harmonic vibrational frequencies (cm - 1 ) of the water


trimer, computed at SCF and correlated level, with DZP+diff
basis set, along with percentage change93.

SCF CCSD CCSD-SCF (%)

4234 3957 -6.5


4233 3953 -6.6
4229 3951 -6.5
4072 3769 -7.4
4068 3762 -7.5
4038 3692 -8.6
1776 1720 -3.2
1754 1671 -4.7
1750 1667 -4.7
741 941 27
614 664 8
481 536 11
392 443 13
314 356 13
305 331 9
207 264 28
184 208 13
166 195 17
151 186 23
144 177 23
122 170 39

The complex is bound by 16.3 kcal/mol, relative to three isolated water molecules. This
value is higher by 1.3 kcal/mol than three times the 5.0 kcal/mol binding energy of the dimer
at the same level. The three-body component is negligible at the MP2 level, and more than
2 kcal/mol at the SCF level. This very small correlation component is in agreement with a
previous work68 but at odds with a later calculation91. The authors believe the large MP2
three-body term identified by Xantheas is due to a change in geometry rather than a true co-
operative effect.
The authors went on to compute the energies of seventy points on the PES, as a function
of the torsional angles of the three non-bridging hydrogen atoms. These data were then fit
to an analytical function95 so that the dynamics of these hydrogen motions could be exam-
ined. It appears that these three atoms undergo large-amplitude flipping and torsional mo-
tions, even in the lowest vibrational states.

5.4.5 Substituent Effects


Some of the effects of replacement of a nonbridging hydrogen by a phenyl group have al-
ready been described earlier in Section 5.4.4 in the context of interactions between water
and phenol in various oligomers.
The effect of a progressively larger alkyl group replacing one of the hydrogens of water
was probed by considering a series of alcohols96. Study was restricted to cyclic trimers, at
the SCF level with a 6-3 1G basis set, and with no correction for superposition error. The
Cooperative Phenomena 271

Table 5.19 Binding energies (kcal/mol) of cyclic trimers of mixed water (W)
and alcohols, computed at SCF/6-31G level. Data computed at SCF/6-31G
level96.

ROH W3 W2 ROH W (ROH)2 (ROH)3


CH3OH 24.3 23.9 23.3 22.8
C2H5OH 24.3 23.6 22.9 22.1
C3H7OH(1) 24.3 23.6 22.8 22.0
C3H7OH(2) 24.3 23.5 22.7 22.0

cyclic character of these trimers, and indeed of higher oligomers as well, is consistent with
molecular beam electric deflection results71. The binding energetics in Table 5.19 indicate
the nature of the alcohol, whether it is methanol, ethanol, or 1- or 2-isopropanol, has little
effect on the strength of the interactions. The number of alcohol molecules is a minor con-
cern as well: As each water molecule of the trimer is replaced by an alcohol, there is a very
minor decrement in the binding energy, on the order of 0.5 kcal/mol.
Unlike the energetics, there is an elongating effect of alkyl substitution upon the in-
teroxygen distances. Since the complexes are not fully symmetrical, and there are three dif-
ferent values of R(O..O), it is the average values that are listed in Table 5.20. Reading across
a given row illustrates that each replacement of a water molecule by an alcohol stretches
the average H-bond by some 0.01 A. The size of the alkyl group on the alcohol has a small
effect as well, with the larger groups elongating the H-bonds by slightly more. There ap-
pears to be a linear correlation between energetics and geometry in the sense that the
strongest total binding energy is associated with the shortest mean interoxygen separation.
The similarity of energetic and geometric data for ethanol and 1-propanol suggests the 2-
carbon chain is sufficient to model a much longer one. Comparison of the cyclic ho-
motrimers with the corresponding dimers reveals that only in the case of water is the bind-
ing energy of the trimer larger than three times that in the dimer. The three alcohol trimers
all show a slightly negative cooperativity. Nevertheless, the important result is that in all
cases, the energetics of distortion of the three H-bonds are approximately compensated by
the cooperativity factor.
A far better 6-3 11 + + G(2d,2p) basis set was applied to the cyclic methanol trimer97, al-
though the calculations were limited to the SCF level. At this level of calculation, the
R(O .. O) distances are shorter by 0.004 A than in the water trimer, in contrast to the study
at the lower level where the contraction is five times greater. In both trimers, the stretches
of the OH bonds, relative to the monomer, are 0.007 A. The energetic features of coopera-
tivity in the methanol trimer are very similar to those for water.

Table 5.20 Average interoxygen distances (A) in cyclic trimers. Data computed
at SCF/6-31G level96.
W3 2.702
ROH W2 ROH W (ROH)2 (ROH)3
CH3OH 2.711 2.718 2.724
C2H5OH 2.717 2.728 2.738
C 3 H 7 OH(1) 2.716 2.728 2.738
C3H7OH(2) 2.720 2.734 2.755
272 Hydrogen Bonding

Table 5.21 Shifts in harmonic intramolecular vibrational


frequencies ( c m - 1 ) of dimer and trimer of memanol, relative to
the monomer, computed at SCF/6-311 + +G(2d,2p) level97.

n O—H bends O—H stretch C — O stretch

2 2,45 -5, -69 -14, 10


3 34, 42, 82 -98, -103, -128 -2,0, 10

Table 5.21 reports the shifts that arise in the intramolecular vibrational frequencies, rel-
ative to the methanol monomer. It is possible to draw inferences by comparison with the
data for water in Table 5.14, even though the latter are correlated and obtained with a dif-
ferent basis set. In either case, the bending frequencies are blue shifted in the dimer, more
so in the trimer. A red shift occurs in the O—H stretch which is magnified in the trimer. The
authors pointed out that these shifts are significantly smaller than experiment, and attribute
the discrepancy to a lack of correlation. This contention is in part confirmed by the much
larger red shifts in the corresponding modes of the water trimer, following correlation. Un-
like the blue shifts noted in the O—H stretches for the free (nonbridging) hydrogen of wa-
ter, there is much less change in the C—O stretching frequencies of methanol dimer or trimer.

5.5 Mixed Systems

There has been a good deal of work on the cooperativity of H-bonds involving mixed sys-
tems. As an example, a pair of HX (X=F,C1) molecules were added to NH3 in order to ex-
amine the effects of multiple H-bonds on the charge distributions within each monomer98.
Rather than optimize the geometry, the NH3 molecule was retained in its experimental
monomer structure. The symmetry of each complex was taken as C3v, presuming that both
HX molecules are collinear with the N lone electron pair. Calculations were limited to the
SCF level. Comparison of the first two columns of Table 5.22, representing the charges in
the isolated monomers, with those for the 1:1 XH ... NH 3 complexes, illustrates that forma-
tion of the initial H-bond causes a polarization of the two molecules. The N atom of NH3
increases its negative charge while the hydrogen becomes more positive. An even stronger

Table 5.22 Mulliken atomic charges in complexes of NH3 with HF and HC198.

XH+NH3 XH...NH 3 X XHo...XHi...NH3

X= F Cl F Cl F Cl

NH3 H 0.243 0.243 0.269 0.264 0.274 0.267


N -0.727 -0.727 -0.793 -0.776 -0.796 -0.781
HXi H 0.380 0.197 0.436 0.274 0.461 0.292
X -0.380 -0.197 -0.449 -0.291 -0.481 -0.298
HX() H 0.380 0.197 0.411 0.210
X -0.380 -0.197 -0.420 -0.211
Cooperative Phenomena 273

polarization occurs in HF and HCl The last two columns show the continuation of these
trends as the second HX molecule is added to the complex. The changes in NH3 upon adding
the second molecule are quite small, 5 me or less. The redistributions in the HXi molecule,
now the central of the three molecules, are substantially larger: The H atom loses about 25
me of electron density while F picks up an additional 30 me. The inner HX molecule is quite
a bit more polarized than is the outer one. In the case of HF, the inner H atom has 50 rne
less density than does Ho, with a comparably larger negative charge on Fi.

5.5.1 Geometries
These systems were examined in further detail when a 6-31G** basis set was applied to
H3Z...HF...HF, Z=N,P99. The geometries of all complexes were fully optimized at the SCF
level. Unlike the earlier assumption98 that the trimer would contain fully linear H-bonds
and belong to the C3v point group, the equilibrium structure is highly bent, as illustrated in
Fig. 5.26. This structure is sensible in light of the two dimers contained within. That is, one
can consider the complex from the standpoint of taking H3Z...HF as a starting point. This
dimer would contain a linear H-bond. The HF molecule would act as proton acceptor to an
additional HF molecule, which would orient itself toward one of the inner F (Fi) atom's lone
pairs. Were there no other interactions present, then, the (Z..Fi..Fo) angle would tend toward
the 100° or so, characteristic of HF..HF. However, when this angle decreases, it is possible
for the Fo atom to form a favorable interaction with a proton of H3Z, which can be aided by
nonlinearities developing in each of the two H-bonds already present. These nonlinearities
are denoted by the a angles in Fig. 5.26.
Table 5.23 reports the salient features of the geometries of the various complexes. It is
immediately clear that addition of a second HF molecule to H3Z...HF contracts the H-bond
in the latter 1:1 complex. When Z==N, this contraction amounts to 0.13 A; it is 0.17 A in
the case of P. The interfluorine distance in either trimer is shorter than in the simple HF
dimer. This is particularly true in the case of H3N, which is a better proton acceptor than is
H3P. The effects of cooperativity are also apparent in the internal r(HF) bond lengths. The
bond which winds up in the central molecule, denoted ri, stretches by 0.03 A when H3N is
added to HP..HF. The addition of the H3Z molecule also lengthens the outer HF bond, al-
beit by a smaller amount. With regard to the angular features, it might be noted that addi-
tion of the second HF molecule to H 3 Z .. HF causes a 12-21° nonlinearity. Coupled with this
effect is a 17° nonlinearity in the new H-bond to (HF)o, and a reduction in the 0 angle from
102° in HP..HF to only about 70° in the full trimer. All of these angular changes operate to
allow the (HF)o molecule to more closely approach H3Z.

Figure 5.26 Illustration of geometry of H 3 Z ... HF ... HF complex", and definition


of Ebend.
274 Hydrogen Bonding

Table 5.23 Geometric features of H3Z...HF...HF, Z=N,P. See Fig. 5.26 for definition of
parameters. Data in A and degs99.

R(Z..F) R(F..F) ri r
o i %
..
H 3 N HF 2.756 0.918 0.0
H3P..HF 3.479 0.906 0.0
HF..HF 2.725 0.904 0.905 14.3 101.8
H3N..HF..HF 2.623 2.605 0.935 0.912 12.0 16.1 72.1
H3P..HF..HF 3.307 2.662 0.913 0.908 21.0 17.4 66.8

5.5.2 Energetics
The above geometrical changes are all indicative of a positive reinforcement in the strengths
of the various H-bonds. Further evidence of this cooperativity arises from consideration of
the energetics. The binding energies of complexes are reported in Table 5.24. The data is
organized into two separate paths for building the 1:2 trimer. Route 1 first combines a H3Z
molecule with HF, then adds the second HP molecule. The same complex is attained in
Route 2 by first combining the two HF molecules, then adding H3Z. The total energy of the
complex is of course independent of the route chosen. Comparison of reaction 2a with 1b
in Table 5.24 shows that HF binds more strongly to the HF end of the preformed H 3 Z ... HF
complex than it does to the isolated HF molecule. This enhanced binding is equal to 5.70
kcal/mol when Z=N and about half that amount for Z=P, as indicated by the entry in Table
5.24 for Ecoop. This effect may be understood on the basis of the H3Z molecule transferring
some electron density to HF in the H 3 Z ... HF complex, making the HF a better proton ac-
ceptor. The stronger cooperativity associated with H3N relative to H3P is due to the better
proton-accepting capability of the former. Of course there are other factors, such as polar-
ization of HF, which contribute to the cooperativity. An analogous comparison between la

Table 5.24 Complexation energies ( Eelec) and measures of cooperativity in


H3Z...HF...HF. Data in kcal/mol99.
Z=N Z=P

route 1
la: H3Z + HF H3Z..HF -11.81 -4.13
1b: H 3 Z .. HF + HF H3Z..HF..HF -11.67 -8.78
route 2
2a: HF + HF HF..HF -5.97 -5.97
2b: H3Z + HF..HF H3Z..HF..HF -17.51 -6.94
totala -23.48 -12.91
Ecoopb -5.70 -2.81
Eabcc -4.09 -1.42
Ebend -1.73 -1.06

"la - l b ( = 2a + 2b)
b
lb - 2a (=-- 2b - la)
°Eabc = K(ahc) - |K(ab) + E(ac) I E(bc)l + [E(a) + E(b) + E(c)]
d
Ebend = E(Cs) - E(C3v), see Fig. 5.26.
Cooperative Phenomena 275

and 2b illustrates that the H3Z molecule binds more strongly to the preformed HF dimer
than to a single HF molecule. The reasons are quite similar, with the dimerization making
the proton more accessible to the approaching lone electron pair of H3Z.
The above definition of cooperativity ignores any interaction between the terminal HF
molecule and H3Z, folding any attraction into the cooperativity term, Ecoop. One can re-
move this interaction from the effect by computing this interaction energy directly. Specif-
ically, the three-body term is defined in the usual manner: the total binding energy of the
complex, minus the interaction energy of each pair of subunits:

Another important point of distinction with Ecoop is that the three-body term is evaluated
with all interactions computed in the precise geometry, internal as well as intermolecular,
of the trimer. In contrast, Ecoop permits relaxation of the geometry of each entity,
monomer, dimer, or trimer. The appropriate row of Table 5.24 reports these three-body
terms for the 1:2 complexes, which are each reduced relative to E coop . This reduction is
understandable since Eabc removes the interaction between the two terminal molecules.
This formulation suggests the cooperativity associated with H3N is nearly three times
larger than that of H3P. One can obtain a measure of this interaction between terminal mol-
ecules by forcing the geometry to adopt a fully linear, C3v, geometry, as illustrated in Fig.
5.26. Of course, the bending energies, reported as Ebend in Table 5.24, reflect also the
strain that full linearity imposes on the HP..HF interaction, which prefers a 9 angle of 102°,
much smaller than the 180° in the C3v structure. Note that the sum of Ebend and Eabc is
roughly equql to Ecoop.

5.5.3 Vibrational Spectra


The same systems were probed also for the effects of cooperativity upon the vibrational fre-
quencies and intensities100. The data in Table 5.25 focus on the stretching vibrations of the
HF molecule. The results are organized along the same two reaction schemes presented in
Table 5.24 so as to parallel the energetics. The first row illustrates the red shift induced in
the HF stretching frequency when it donates a proton to H3Z. This shift is much larger for
H3N due to its stronger basicity. The formation of the H-bond also substantially increases
the intensity of this band, by a factor of seven for H3N...HF, as indicated in the last two

Table 5.25 Frequency shifts and intensity enhancements in HF stretch of central subunit of
H 3 Z ... HF ... HF 100 .
a
v (cm--1) Ap /Ara

Z=N Z=P Z=N Z==P

route 1
la: H3Z + HF H3Z..HF -432 -134 7.1 4.0
Ib: H3Z..HF + HF H3Z..HF..HF -389 -143 1.0 1.1
route 2
2a: HF + HF HF..HF -41 -41 1.4 1.4
2b: H3Z + HF..HF H 3 Z .. HF .. HF -780 -236 5.2 3.1
a
Ratio of intensity of product to that of reactant.
276 Hydrogen Bonding

columns of Table 5.25. Addition of the second HF molecule has little effect upon the in-
tensity of the stretching band of the HF molecule already bound to H3Z. On the other hand,
this second H-bond introduces a further red shift, comparable in magnitude to that occur-
ring when HF forms its first H-bond with H3Z. Comparison of the 1b row with the data in
row 2a illustrates the cooperativity in this complex. That is, the binding of one HF to an-
other only red shifts the HF stretching frequency of the proton acceptor by 41 cm-1. But if
the latter HF molecule is already bound to H3Z, the formation of the bond with HF lowers
the frequency by several times that amount. Another measure of the cooperativity emerges
from comparison between la and 2b. As indicated above, bonding of H3Z to HF lowers the
frequency of the latter molecule's stretch by a substantial amount. But the last row of Table
5.25 illustrates that this shift is magnified by a factor of about two if the HF molecule has
already attached itself to another HF, prior to the approach of H3Z.
The intermolecular modes also provide some insights into the nature of cooperativity in
these complexes. The computed harmonic frequencies are reported in Table 5.26, using the
nomenclature developed by Bertie and Falk101 wherein v has its usual meaning of the H-
bond stretch. The bending motions of the proton donor molecule are denoted vb and vt,
whereas v 1 and v 2 refer to the bends of the acceptor. Because of the angular characteris-
tics of the H-bond, and the low effective mass for a proton motion, the proton donor bend-
ing motions have the highest frequencies of the intermolecular modes in the binary com-
plexes. The weaker nature of the H-bond in H3P...HF, as compared to H3N...HF, explains
the fact that the frequencies in the former complex are roughly half those in the latter. Of
particular relevance here is a comparison of the 1:1 complexes in the first two columns with
the 1:2 complexes in the last two columns. In each case, addition of the second HF mole-
cule acts to increase the frequency of the intermolecular mode within H3Z...HF. This in-
crement can be as small as 27 cm-1 in the V band of H3P...HF, or as much as 300 cm-1 in
a number of cases.
The intensities of the intermolecular modes are compiled in a similar organizational
scheme in Table 5.27. The intensities of the proton donor bending modes of the 1:1 com-
plexes are by far the strongest, with those corresponding to H 3 N ... HF consistently larger
than those of H3P...HF. Comparison of the first two columns with the last two again brings
out the cooperative effects in the trimer. Adding the second HF molecule intensifies all the
bands by a substantial amount; the only exception is the v 2 band of H3P...HF...HF.

Table 5.26 Frequencies ( c m - 1 ) of intermolecular vibrational modesa in 1:1 and


1:2 complexes of H3Z with HF100.
..
H3Z HF H3Z...HF...HF

Z=N Z=P Z=N Z=P

V 240 113 290 140


v 876 467 1162 800
v 877 464 1024 603
V 1 236 108 417 247
V 238 117 307 175
a ...
v refers to Z F stretch, vb and v to bending motions of proton donor: bonds of acceptor arc denoted
by v and v 101.
Cooperative Phenomena 277

Table 5.27 Intensities (km/mol) of intermolecular vibrational modes in 1:1 and


1:2 complexes of H3Z with HF100.

H3Z...HF H3Z...HF...•HF

Z=N Z=P Z=N Z=P

v 3 1 13 5
vb 208 134 264 247
v 208 134 245 275
V 10 6 49 10
V
10 6 13 0.1

5.5.3.1 Analysis of Intensities


The intensity of a given vibrational band is directly related to the change in the molecular
dipole moment that is associated with the motions of the atoms corresponding to that par-
ticular normal mode. Hence, the intensities offer a point of contact between the computed
electronic distributions and experimental observation. Just as atomic charges can be defined
by integrating the computed electron density in a given region surrounding a particular cen-
ter, another means of assigning a charge is via the dipole moment change associated with
the atom's motion 102-105. Following a scheme designed some years ago106, an atomic po-
lar tensor (APT) is defined for a given atom a as

where i represents one of the three components of the dipole moment and qj is the x,y,
or z coordinate of atom a. For example, Pxz refers to the change in the x-component of
the molecular dipole moment when atom a is moved a certain distance along the z-axis.
An effective charge for an atom a is defined as the sum of the squares of all nine elements
of Pij.
Beginning with the HF bond, when this bond direction is taken as the z-axis, the PZZH
element describes the change occurring in the component of the dipole moment in this di-
rection, when the H atom is stretched along the HF bond. In the isolated HF molecule, this
element is equal to 0.36100. It increases to 0.96 in H3N...HF, and then to 1.14 when the sec-
ond HF molecule is added, forming H3N...HP...HF. The intensities calculated for the HF
stretch are approximately proportional to these values of PzzH. One can conclude that as the
H-bond is strengthened through the cooperativity, the charge cloud around HF changes such
that the molecular dipole is more easily altered by displacement of the proton along the HF
axis. Perhaps a better way to view this process is along the following lines: when the bridg-
ing proton of a H-bond moves in one direction, electron density shifts in the opposite di-
rection107. Thus the formation of a H-bond causes the motion of this proton to drastically
increase its "net" positive charge, resulting in a very large change in the moment, and thus
to a stronger intensity.
The intensity changes occurring in the proton acceptor molecules H3N and H3P offer
some real insights into the effects of multiple H-bonding upon electron distributions100.
Whereas the intensity of the symmetric stretch in the H3N monomer is quite weak, it is in-
tensified by a factor of 35 when complexed with HF, and then by a further factor of 40 when
278 Hydrogen Bonding

the second HF is added to the complex. In concert with these magnifications in the inten-
sity of this mode are large increases in the PZZH element of the atomic polar tensor of the
H3N protons. It is possible to further partition a dipole moment derivative like z/ z into
separable contributions106,108.109. Of utility here is a fractional charge q and a "charge flux,"
CF. The latter term accounts for the loss or gain of electron density as the atom is displaced
along a given axis. Within this framework, the intensity patterns were rationalized along the
following lines.
Beginning with the H3N monomer, the NH stretching modes are of low intensity due to
a cancellation between two factors. Each H atom has a fractional positive charge, so its dis-
placement away from N would cause the molecular moment to change. But as the H atom
moves, electron density accumulates on it, lowering its positive charge, so the moment is
changed very little, resulting in the low intensity. When the H3N molecule is complexed
with HF, each of the three H atoms becomes more positively charged. Working in concert
with this effect is the tendency of the H-bond to "restrict" the electron cloud, making it less
able to accompany the H atoms as they stretch away from N. Together, these effects account
for the large dipole moment change associated with the N—H stretch, and the consequently
intense band. The further intensity enhancement resulting from the addition of the second
HF molecule is due to a continuation of the above two effects: greater positive charge on
N—H protons, and less deformable electron cloud.
A comparison of the above patterns of H3N with those observed when H3P acts as pro-
ton acceptor is enlightening as well100. Whereas the N—H stretching intensities increased
manyfold when the HF molecules were added, the analogous P—H stretches show rela-
tively small perturbations. In fact, addition of the first HF molecule reduces the intensity of
the symmetric stretch by a factor of 2.5. The first point of contrast between the two proton
acceptor molecules is that the H atoms of H3P are negatively charged. As for H3N, the elec-
tron cloud follows these H atoms, enhancing their negative charge. As a result, the sym-
metric stretch in isolated H 3 Phas substantial intensity, 500 times larger than in H3N. When
the H-bond is formed with HF, the withdrawal of density from the H3P molecule now acts
to diminish the negative charge of the hydrogens. At the same time, the H-bond causes the
electron cloud to become less deformable, reducing its ability to follow the H atoms as they
stretch away from P. Together, these two effects act to lower the intensity of the symmetric
stretch. In summary, the profoundly different behavior of the intensities of the H3N and H3P
stretches can be simply explained on the basis of the opposite charges of the H atoms in the
two molecules. The H-bond has similar consequences for both the charges on the atoms and
the deformability of the electron cloud.

5.5.4 Effects of Electron Correlation


This style of analysis was followed up by another group of coworkers who replaced the ZH3
molecule by water and also included the effects of electron correlation110. The 6-31G** set
was used, as was the larger + (VP)s(2d)s containing two sets of d-functions, plus a diffuse
sp-set111. As in the earlier cases, the absolute minimum is of cyclic type, with a weak H-
bond between the H2O and the second HF molecule, as illustrated in Fig. 5.27. Since H2O
is a poorer proton acceptor than is NH3, one would expect the H-bond to HF to be weaker,
and the cooperativity within the trimer to be less extensive. On the other hand, the better
proton-donating ability of H2O should strengthen the last H-bond that completes the cyclic
character of the trimer.
Cooperative Phenomena 279

Figure 5.27 Illustration of geometry of H2O...HF...HF com-


plex, including definition of geometrical parameters110.

The geometric features of the binary complexes are listed in Table 5.28, along with the
1:2 complex in the last two rows. The expected contraction of the O...F distance as the sec-
ond HF molecule is added is immediately apparent. This contraction amounts to 0.10 A at
the SCF level, less than that in the H 3 N ... HF ... HF complex where the shrinkage was 0.13
A. Correlation has little influence on the H-bond reduction; the MP2 contraction is 0.11 A.
From the perspective of adding the proton acceptor molecule to the HF dimer, R(F..F) is re-
duced by 0.12 A for both H2O and NH3, at the SCF level. The distinction between the lat-
ter two molecules is perhaps most clearly seen in the bond length of the inner HF molecule,
ri This bond stretches by 0.017 A when the outer HF molecule is added to H3N...HF, but
by only 0.009 A if H3N is replaced by H2O. Note, however, that when correlation is added,
this stretching doubles.
Table 5.29 reports the energetics of step-by-step construction of the H2O...HP...HF com-
plex, using the same scheme established earlier for H3N...HF...HF99. In addition to consid-
ering the effects of electron correlation, the investigation of H2O...HF...HF added counter-
poise corrections for the complexation energies, reported as the "cc" data in Table 5.29. The
comparison with H3N...HF...HF is most consistent if the first column of this table is placed
alongside the first column of Table 5.24, all results at the SCF level with a 6-31G** basis
set. The first rows verify the better proton accepting ability of NH3, as the binary complex
with HF is stronger by some 2.7 kcal/mol. The cooperativity fostered by NH3 and H2O are
apparently similar, based on the second rows. Addition of the second HF molecule to
H2O...HF accounts for 12.1 kcal/mol; replacement of H2O by NH3 makes this process ex-
ergonic by 11.7 kcal/mol. Another means of assessing this cooperativity is Ecoop , the en-
hancement of the binding energy of the second HF molecule after the proton acceptor has
attached to the first HF. This quantity is comparable for H2O and NH3 in Tables 5.24 and
5.29, as are the three-body terms Eabc.

Table 5.28 Geometric features of H2O...HF...HF. See Fig. 5.27 for definition of parameters.
Data calculated with +(VP)s(2d)s basis set, in A and degs110.

R(O..F) R(F..F) r
i r
o a 0 0
..
H2O HF SCF 2.723 0.910 1.7
MP2 2.661 0.939 1.5
HF..HF SCF 2.822 0.901 0.903 7.5 116.5
MP2 2.764 0.924 0.927 6.2 111.2
H 2 O .. HF .. HF SCF 2.623 2.696 0.919 0.908 12.7 20.2 70.2
MP2 2.550 2.622 0.956 0.937 11.2 18.1 68.6
280 Hydrogen Bonding

The remaining data in Table 5.29 provide an opportunity to examine the effects of BSSE,
basis set, and correlation upon the cooperativity of energetics. The 6-31G** basis set is sub-
ject to a fairly large superposition error, accounting for the quite large discrepancies be-
tween unconnected and corrected interaction energies in the first two columns of the table.
For example, 4.5 kcal/mol of the total of 12.1 SCF binding energy of the second HF mole-
cule to H2O...HF is due to this error. The superposition error is much reduced with the
+(VP)s(2d)s basis set. Once the counterpoise corrections are added, the results with the
two basis sets are comparable (within about 1 kcal/mol of one another), if not identical.
Consistent with expectations, correlation adds to the binding energy of each H-bond. The
correlation is apparently responsible for an increase in the cooperativity effect as well. The
counterpoise-corrected values of Ecoop are —2.9 and —4.1 kcal/mol at the SCF and MP2
levels, respectively, and a similarly larger value of Eabc is obtained with MP2.

5.5.5 Other Mixed Trimers


The H2O...HF...HF complex was reexamined more recently in conjunction with a compar-
ison with H 2 O ... H 2 O ... HF 112 . The focus of this work was an exploration of the potential en-
ergy surface to identify all the local minima. In addition to the cyclic structure of Fig. 5.27,
a bifurcated geometry, in which the water serves as double proton acceptor, was also lo-
cated as a minimum of the PES of H2O...HP...HF. In the complex containing a pair of wa-
ter molecules and one HF, the only minimum located was of cyclic type. Both of these struc-
tures are illustrated in Fig. 5.28.
The binding energy for the cyclic H2O...HF...HF complex in Fig. 5.27 and the analogous
H 2 O ... H 2 O ... HF in Fig. 5.28 are calculated to be quite similar to one another. As illustrated
in Table 5.30, the SCF-level calculations suggest the former is marginally more strongly
bound, whereas including correlation shifts the balance toward the latter. In any case, it is
clear that there is little difference between the two. Computation of the cooperativity within

Table 5.29 Complexation energies and measures of cooperativity in H2O...HP...HF. Data


in kcal/mol110.
6-31G** +(VP)s(2d)s

SCF SCF MP2

no cc cc no cc cc no cc cc

route 1
la: H20 + HF H2O..HF -9.06 -8.00 -7.36 -7.23 -8.86 -7.83
1b: H2O..HF + HF H2O..HF..HF -12.09 -7.54 -6.76 -6.61 -9.04 -8.07
route 2
2a: HF + HF HF..HF -5.97 -4.23 -3.83 -3.75 -4.60 -4.02
2b: H2O + HF..HF H2O..HF..HF -15.18 -11.31 -10.29 -10.09 -13.30 -11.88
totala -21.15 -15.54 -14.12 -13.84 -17.90 -15.90
Ecoopb -6.12 -3.31 -2.93 -2.86 -4.44 -4.05
E
abc
-3.16 -3.08 -2.18 -2.28 -3.94 -4.12
a
la + l b ( = 2a + 2b)
b
Mb - 2a (= 2b - la)
c
E abc = E(abc) [E(ab) + E(ac) + H(bc)] + [K(a) - li(b) i K(c)]
Cooperative Phenomena 281

Figure 5.28 Secondary minimum of H2O...HF...HF com-


plex, and single minimum for H2O...H2O...HF112.

the two complexes, analogous to Ecoop in Table 5.29, leads to MP2 values of —1.38 and
- 2.41 kcal/mol for H 2 O ... HF ... HF and H2O...H2O...HF, respectively.
Another set of calculations replaced the HF molecule of H2O...H2O...HF by HC1 and ob-
tained the optimized geometry of H2O...H2O...HC1 at the MP2 level with a series of three
different polarized basis sets80. The structure is analogous to the C1 geometry of
H2O...H2O...HF in Fig. 5.28. The interoxygen separation is in the range 2.75-2.81 A and
R(C1...O) is 2.99-3.06 A (depending upon particular basis set). The H-bonds are within
about 10° of linearity. A harmonic frequency analysis of the MP2 data revealed that the red
shift of the water molecule's OH stretch (relative to the monomer) increases from 98 cm-1
in the water dimer to 210 cm-1 when HC1 is added to form H2O...H2O...HC1. The former
refers to the OH ... O H-bond which is apparently strengthened by addition of HC1; the red
shift in OH...C1 is 48 c m - 1 , verifying the expected weaker nature of this bond These val-
ues are in good agreement with experimental measurements of 172 and 52 cm-1 for this
complex in Ar matrix 113 .
The energetics of formation of the H2O....H2O...HCl complex are reported in Table 5.31,
in a format comparable to that for H2O...HF...HF in Table 5.29. All data were obtained at
the MP2 level and are corrected for BSSE by the counterpoise procedure. The three basis
sets all contain polarization functions and yield comparable results. The binding of HC1 to
the pre-formed water dimer (step 1 b) is considerably more favorable than its binding to a
water monomer, as in step 2a. This enhancement is referred to as ECOOp in Table 5.31, and
amounts to between 2 and 3 kcal/mol. (The same quantity is identically equal to the en-
hancement of binding of water to H2O...HC1 as compared to HC1.) It is worth noting that
whereas the total binding energy in the appropriate row of Table 5.31 is rather insensitive

Table 5.30 Computed binding energies of cyclic H2O...HF...HF


and H2O...H2O...HF. Data corrected for superposition error,
in kcal/mol112.
H2O...HF...HF H2O...H20...HF

6-31G** -16.02 -15.95


6-31 + +G** -16.49 -15.46
6-31 + + G(2d,2p) -14.41 -13.97
MP2/6-31 + +G** -17.25 -17.58
282 Hydrogen Bonding

Table 5.31 Complexation energies ( Eelec) and measures of cooperativity in H2O...H2O...HCl.


Data computed at MP2 level and corrected for BSSE, in kcal/mol80.
DZP 6-31G(2d,p) Poll

route 1
la: H2O + H2O H2O..H2O -5.04 -4.58 -4.32
1b: H2O..H2O + HCl H2O..H2O..HC1 -7.32 -7.67 -7.86
route 2
2a: H2O + HC1 H2O..HC1 -5.09 -5.25 -4.92
2b: H2O + H2O..HC1 H2O..H2O..HC1 -7.26 -7.00 -7.26
total" -12.36 -12.25 -12.18
Ecoopb -2.23 -2.42 -2.94
E
coop,vib
-1.52 -2.12 -2.96
a
la + 1b (= 2a + 2h)
b
lb - 2 a ( = 2 b - la)
c
Same as Ecoop, with zero-point vibrational corrections.

to details of the basis set, there is a clear and steady increase in the cooperativity as the ba-
sis set becomes more flexible.
It is intriguing that this 2-3 kcal range of Ecoop is quite similar to the 2.4 kcal/mol com-
puted for H2O...H2O...HF, also at the MP2 level and with BSSE correction112, despite the
substitution of HF by HC1 in the present system. Perhaps Ecoop would have been larger in
the H2O...H2O...HF system had a basis set more flexible than 6-31 ++G** been used. At
any rate, this cooperativity is significantly smaller than the 4 kcal/mol reported in the last
column of Table 5.29 for H 2 O ... HP ... F where there are two HF molecules present, again
obtained at the MP2 level with BSSE corrections.
The last row of Table 5.31 refers again to the energy of cooperativity, except that all
binding energies have been corrected for zero-point vibrations. Ecoop,vib pretty much du-
plicates the results without such corrections, except that the sensitivity to basis set flexi-
bility is heightened. So much so that there is a doubling in this quantity between the "stan-
dard" DZP basis set and the Pol 1 type. One might conclude that even though it is possible
to obtain good estimates of the binding energies of H-bonded clusters, accurate assess-
ment of the degree of cooperativity in multiply H-bonded systems can be particularly de-
manding in terms of a flexible basis set. Such a rule is commensurate with the notion that
the cooperative effects originate largely from polarizations of electron clouds induced by
H-bond formation.

5.6 Summary

The C—H bonds of HCN become progressively longer as the number of molecules in
(HCN)n increases. This elongation is greater for the proton donor molecule on the end and
least for the acceptor. Longest of all are the C—H bonds of molecules in the center of the
chain. The C N bond of the donor is elongated while that of the acceptor is shortened as
the chain grows; those in the middle exhibit little change. The energetic consequences of
cooperativity grow as the chain is enlarged, in that the total interaction energy in an n-mer
is more than n — 1 times larger than the interaction in a simple dimer. The average H-bond
Cooperative Phenomena 283

energy within an infinitely long chain is projected to be about 25% larger than the interac-
tion energy in the dimer. Cooperativity also makes for a dipole moment of an assembly of
HCN molecules that is greater than the sum of moments of each monomer. The stretching
frequencies behave much like the bond lengths in that bond stretches correspond to red
shifts and contractions to frequency increments. The longest C—H stretches of the termi-
nal proton donor molecule correlate with progressively larger red shifts as the chain grows.
The frequency shifts and bond length changes in the CN bonds are considerably smaller
than for CH. Cooperativity exhibits particularly dramatic effects upon the intermolecular
frequencies. For example, there are two H-bond stretching frequencies, v , in the trimer.
One of these is much larger than the value of V in the dimer, while the other is much smaller.
These trends continue as the chain elongates.
Although each NCH ... NCH connection prefers full linearity, a bending of each such con-
nection in an oligomer would enable the chain to bend around on itself, eventually allow-
ing the terminal lone pair of the first molecule to pair up with the proton of the last. The ad-
vantage would be the formation of an additional H-bond, but this would come at the expense
of angular distortion of each of the other H-bonds in the oligomer. Calculations indicate that
for a chain containing four or more HCN molecules, the former outweighs the latter and a
cyclic complex will be favored over a linear arrangement.
There is some evidence for a small amount of Cooperativity in the HCCH trimer, albeit
of a lesser degree due to the absence of a true H-bond. The interaction energy of the cyclic
trimer is very nearly equal to three times that of a single dimer, even though each pairwise
geometrical relationship is distorted from the true T-shape favored by the dimer. The
tetramer and pentamer are also likely to be of cyclic type, but again little variance is ob-
served in the H-bond energy per pairwise interaction.
The HF trimer adopts a cyclic geometry wherein the geometric distortion of each H-bond
is compensated by the Cooperativity as each molecule serves as both proton donor and ac-
ceptor. The lesser degree of deformation in the cyclic tetramer leads to a significantly
greater apparent Cooperativity. Similar trends were observed in the trimer of HC1. Three-
body interactions, another measure of Cooperativity phenomena, are well modeled at the
SCF level. The primary contributor arises from the deformation of electron density, akin to
polarization effects. The two-body SCF interaction is primarily responsible for the radial
and angular anisotropy of the energetics of (HF)3. Continued enlargement of the ring leads
to a progressive elongation of the HF bonds, coupled with reduction in the interfluorine dis-
tances. Also in evidence is a progressively larger red shift of the HF stretches as the ring in-
cludes more members. If maintained in an open chain structure, the molecules in the mid-
dle of the chain exhibit the longer HF bonds and the shortest F...F contacts.
Calculations of an infinitely long open chain of water molecules indicate the interoxy-
gen separation is 2.73 A, and the bridging hydrogen is stretched away from the O atom by
0.05 A. The H-bond energy (per bond) is enhanced by nearly 50% relative to the isolated
dimer. Four-body terms are considerably smaller than three-body and can probably be ig-
nored, along with higher-order terms. The sum of two-body interactions in a long chain,
with idealized geometry, is projected to account for about 80% of the total interaction en-
ergy, the rest arising from terms of higher order. The cumulative sum of three-body inter-
actions can be much smaller in a cluster wherein some triads of molecules contain a dou-
ble proton donor or acceptor (destabilizing) and other triads are sequential (stabilizing). Red
shifts of the OH bond involving the bridging hydrogen in the chain become progressively
larger as the chain elongates; patterns for anharmonic frequencies bear a striking resem-
blance to harmonic data.
284 Hydrogen Bonding

Experimental data indicate that small oligomers of water prefer a cyclic geometry. As in
the dimer, it becomes difficult to assign a particular band as the H-bond stretching mode.
Moreover, most of the normal modes involve significant motions of atoms on more than
one molecule. The sum of pairwise interactions in a cyclic hexamer accounts for only about
80% of the total interaction energy. The electrostatic energy is rigorously additive and the
exchange repulsion is nearly so. The majority of the nonadditivity arises in the charge trans-
fer and polarization terms. Consistent with prior findings on simpler systems, the three-body
term tends to be attractive for sequential triads (where the central molecule is simultane-
ously donor and acceptor) and repulsive for a triad containing a double donor or double ac-
ceptor.
The cyclic trimer contains H-bonds that are all significantly distorted from linearity. The
orientation of each water molecule is controlled largely by two-body terms present at the
SCF level. The anisotropy of the three-body term is due primarily to forces that arise from
deformation of the electron clouds; Heitler-London forces are opposite in sign and smaller
in magnitude. Induction energy per se is not a good substitute for the full SCF three-body
term. While the shapes of the two are not too dissimilar, with respect to angular distortion,
the former is far too attractive. As in the closed, cyclic trimer, the three-body terms in the
open trimers are dominated by SCF deformation forces. The three-body term in the double
proton donor configuration is weakly attractive, but this force is counteracted by a strongly
repulsive pairwise two-body interaction between the two terminal molecules. Correlation
effects are minimal in examining three-body terms.
Geometry optimization leads to the conclusion that the cyclic trimer is indeed the global
minimum on the PES of the water trimer. Despite the distortion of each of the three H-bonds,
the total interaction energy is greater than three times that of a dimer. The 3-body term con-
tributes 17% to the total binding energy of the trimer and 25% for the tetramer; 4-body terms
in the tetramer are nearly negligible. The three R(OO) distances are not quite equivalent,
but all are about 2.83 A. Another minimum on the PES, albeit much less stable than the se-
quential cyclic one, is one in which the central molecule acts as double donor. However,
even this structure takes on cyclic character as the two terminal molecules approach one an-
other to form a third H-bond. Enlarging the system to the tetramer brings about a cyclic
structure, as is true as well for the pentamer and hexamer. The latter is large enough that
there is little nonlinearity that must be suffered by the six H-bonds. There are large red shifts
in the O—H stretching frequencies, amounting to several hundred cm-1 as each new wa-
ter molecule is added. This change is apparently due to SCF deformation effects. The av-
erage H-bond energy in the oligomer rises with the number of molecules, surpassing 5.4
kcal/mol at the SCF level, more than 6 kcal/mol when correlated. Alkylation of the oxygen
causes only small perturbations of these results.
A system combining NH3 with a pair of HX (X=F,CI) molecules illustrates the sorts of
charge rearrangements that accompany the cooperativity. After the redistributions that oc-
cur in the proton donor NH3 molecule upon formation of XH ... NH 3 , the further changes
upon elongation to XH ... XH ... NH 3 are rather minimal. However, the central HX molecule
undergoes significant redistribution when the second HX molecule is added. Optimization
shows this complex to be bent, resembling a cyclic structure, although the interaction be-
tween the terminal ZH3 (Z=N,P) and HF molecules is a weak, distant one which would
not fit the description of a true H-bond. Addition of the second HF molecule induces
changes in the geometry of FH ... ZH 3 that are characteristic of positive cooperativity: The
H-bond contracts by more than 0.1 A. From the other perspective, that is, the influence of
ZH3 upon the geometry of FH ... FH, there is a similar reduction in the F...F distance. Addi-
Cooperative Phenomena 285

tion of the second H-bond also yields further stretches in the internal F—H bonds. The co-
operativity emerges from comparison of binding energies as well. HF binds more strongly
to the H 3 Z ... HF complex than it does to HF. Similarly, H3Z forms a stronger interaction with
the HF dimer than to a single HF molecule. Using another measure of cooperativity, the
three-body term in the complex is negative, signaling a stabilizing force.
The vibrational spectra provide additional insights into the cooperativity in these trimers.
While addition of a second HF molecule to FH...ZH3 has little effect upon the intensity of
the F—H stretch, the frequency of this mode is red shifted by an amount comparable to the
shift occurring when the first H-bond was formed to ZH3. It is worth stressing that this red
shift of the HF stretch in FH ... ZH 3 , is many times larger than would occur were this FH
molecule not first bound to ZH3. Similarly, the red shift occurring when ZH3 adds to
FH ... FH is considerably larger than when it adds to the FH monomer. In other words, a pre-
existing H-bond is more amenable to a change in frequency than is the unbound monomer.
In terms of intermolecular modes, addition of the second HF molecule raises the frequency
of the intermolecular stretch within H3Z...HF. This increment can be as small as 30 cm-1
in the v band of H3P...HF, or as much as 300 cm-1 for some of the bending motions. Adding
the second HF molecule intensifies most bands by a substantial amount.
Analysis of the changes in the molecular dipole moment reveals that when the H-bond
is strengthened by cooperativity, the charge cloud around HF changes such that the molec-
ular dipole is more sensitive to displacement of the proton along the HF axis. Any motion
of the proton is accompanied by a displacement of electronic charge in the opposite direc-
tion. The intensity enhancements in the internal NH stretches of NH3 when engaged in a H-
bond are due to a combination of more positively charged hydrogens and a charge cloud
less able to follow the protons. The opposite effect in PH3 is due to the negative charges on
these hydrogens.
Similar sorts of trends are witnessed in complexes such as FH...FH...OH2, H2O...H2O...HF,
and H 2 O ... H 2 O ... HC1 that contain OH2 instead of NH3.

References
1. Frank, H. S., and Wen, W.-Y., Structural aspects of ion-solvent interaction in aqueous solutions:
A suggested picture of water structure, Faraday Discuss. Chem. Soc. 24, 133-140 (1957).
2. Ceccarelli, C., Jeffrey, G. A., and Taylor, R., A survey of O—H . . . O hydrogen bond geometries
determined by neutron diffraction, J. Mol. Struct. 70, 255-271 (1981).
3. Ludwig, R., Weinhold, F., and Farrar, T. C., Experimental and theoretical studies of hydrogen
bonding in neat, liquid formamide, i. Chem. Phys. 102, 5118-5125 (1995).
4. Chalasinski, G., and Szczesniak, M. M., Origins of structure and energetics of van der Waals
clusters from ab initio calculations, Chem. Rev. 94, 1723-1765 (1994).
5. Hopkins, G. A., Maroncelli, M., Nibler, J. W., and Dyke, T. R., Coherent Raman spectroscopy
of HCN complexes, Chem. Phys. Lett. 114, 97-102 (1985).
6. Kofranek, M., Lischka, H., and Karpfen, A., Ab initio studies on hydrogen-bonded clusters. I.
Linear and cyclic oligomers of hydrogen cyanide, Chem. Phys. 113, 53—64 (1987).
7. King, B. F., and Weinhold, F., Structure and spectroscopy of (HCN)n clusters: Cooperative and
electronic delocalization effects in C — H . . . N hydrogen bonding, J. Chem. Phys. 103, 333—347
(1995).
8. Ruoff, R. S., Emilsson, T., Klots, T. D., Chuang, C., and Gutowsky, H. S., Rotational spectrum
and structure of the linear HCN trimer, J. Chem. Phys. 89, 138-148 (1988).
9. Kofranek, M., Lischka, H., and Karpfen, A., Ab initio studies on structure, vibrational spectra
and infrared intensities of HCN, (HCN)2, and (HCN) 3 , Mol. Phys. 61, 1519-1539 (1987).
286 Hydrogen Bonding

10. Anex, D. S., Davidson, E. R., Douketis, C., and Ewing, G. E., Vibrational spectroscopy of hy-
drogen cyanide clusters, J. Phys. Chem. 92, (1988).
11. Karpfen, A., Ab initio studies on hydrogen-bonded chains. III. The linear, infinite chain of hy-
drogen cyanide molecules, Chem. Phys. 79, 211-218 (1983).
12. King, B. F., Farrar, T. C., and Weinhold, R, Quadrupole coupling constants in linear (HCN)n
clusters: Theoretical and experimental evidence for cooperative effects in C—H . . . N hydrogen
bonding, J. Chem. Phys. 103, 348-352 (1995).
13. Kurnig, I. J., Lischka, H., and Karpfen, A., Linear versus cyclic (HCN) 3 . An ab initio study on
structure, vibrational spectra, and infrared intensities, J. Chem. Phys. 92, 2469-2477 (1990).
14. Jucks, K. W., and Miller, R. E., Near infrared spectroscopic observation of the linear and cyclic
isomers of the hydrogen cyanide trimer, J. Chem. Phys. 88, 2196-2204 (1988).
15. Gdanitz, R. J., and Ahlrichs, R., The averaged coupled-pair functional (ACPF): A size-extensive
modification of MR CI(SD), Chem. Phys. Lett. 143, 413-420 (1988).
16. Alberts, I. L., Rowlands, T. W., and Handy, N. C., Stationary points on the potential energy sur-
faces of (C2H2)2, (C2H2)3, and (C 2 H 4 ) 2 ,J. Chem. Phys. 88, 3811-3816 (1988).
17. Bone, R. G. A., Murray, C. W., Amos, R. D., and Handy, N. C., Stationary points on the poten-
tial energy surface of (C 2 H 2 ) 3 , Chem. Phys. Lett. 161, 166-174 (1989).
18. Yu, J., Su, S., and Bloor, J. E., Ab initio calculations on the geometries and stabilities of acety-
lene complexes, J. Phys. Chem. 94, 5589-5592 (1990).
19. Bryant, G. W., Eggers, D. F., and Watts, R. O., High resolution infrared spectrum of acetylene
tetramers, Chem. Phys. Lett. 151, 309-314 (1988).
20. Bone, R. G. A., Amos, R. D., and Handy, N. C., Ab initio studies of acetylene tetramer and pen-
tamer, J. Chem. Soc., Faraday Trans. 86, 1931-1941 (1990).
21. Karpfen, A., Beyer, A., and Schuster, P., Ab initio studies on clusters of polar molecules. Stabil-
ity of cyclic versus open-chain trimers of hydrogen fluoride, Chem. Phys. Lett. 102, 289—291
(1983).
22. Gaw, J. R, Yamaguchi, Y, Vincent, M. A., and Schaefer, H. F., Vibrational frequency shifts in
hydrogen-bonded systems: The hydrogen fluoride dimer and trimer, J. Am. Chem. Soc. 106,
3133-3138(1984).
23. Liu, S.-Y, Michael, D. W., Dykstra, C. E., and Lisy, J. M., The stabilities of the hydrogen fluo-
ride trimer and tetramer, J. Chem. Phys. 84, 5032-5036 (1986).
24. Scuseria, G. E., and Schaefer, H. F., Vibrational frequencies and geometries for the open HF
trimer, Chem. Phys. 107, 33-38 (1986).
25. Kolenbrander, K. D., Dykstra, C. E., and Lisy, J. M., Torsional vibrational modes of (HF)3:
IR—IR double resonance spectroscopy and electrical interaction theory, J. Chem. Phys. 88,
5995-6012(1988).
26. Latajka, Z., and Scheiner, S., Structure, energetics and vibrational spectra of H-bonded systems.
Dimers and trimers of HF and HCl, Chem. Phys. 122, 413-430 (1988).
27. Michael, D. W., and Lisy, J. M., Vibrational predissociation spectroscopy of (HF)3, J. Chem.
Phys. 85, 2528-2537(1986).
28. Andrews, L., and Bohn, R. B., Infrared spectra of isotopic (HCl) 3 clusters in solid neon, J. Chem.
Phys. 90, 5205-5207 (1989).
29. Han, J., Wang, Z., Mclntosh, A. L., Lucchese, R. R., and Bevan, J. W., Investigation of the ground
vibrational state structure of H35Cl trimer based on the resolved K, J substructure of the v5 vi-
brational band, J. Chem. Phys. 100, 7101-7108 (1994).
30. Suhm, M. A., and Nesbitt, D. J., Potential surfaces and dynamics of weakly bound trimers: Per-
spectives from high resolution IR spectroscopy, Chem. Soc. Rev. 45—54 (1995).
31. Moszynski, R., Wormer, P. E. S., Jeziorski, B., and van der Avoird, A., Symmetry-adapted per-
turbation theory of nonadditive three-body interactions in van der Waals molecules. I. General
theory, J. Chem. Phys. 103, 8058-8074 (1995).
32. Chalasinski, G., Cybulski, S. M., Szczesniak, M. M., and Scheiner, S., Nonadditive effects in HF
and HCl trimers, J. Chem. Phys. 91, 7048-7056 (1989).
Cooperative Phenomena 287

33. Tachikawa, M., and Iguchi, K., Nonadditivity effects in the molecular interactions of H2O and
HF trimers by the symmetry-adapted perturbation theory, J. Phys. Chem. 101, 3062-3072
(1994).
34. Karpfen, A., Ab initio studies on hydrogen-bonded clusters: Structure and vibrational spectra of
cyclic (HF)n complexes, Int. J. Quantum Chem., Quantum Chem. Symp. 24, 129 (1990).
35. Karpfen, A., and Yanovitskii, O., Cooperatively in hydrogen bonded clusters: An improved ab
initio SCF study on the structure and energetics of neutral, protonated and deprotonated chains
of neutral, cyclic hydrogen fluoride oligomers, J. Mol. Struct. (Theochem) 314,211-227 (1994).
36. Liedl, K. R., Kroemer, R. T., and Rode, B. M., Hydrogen transitions between (HF)n Cnh struc-
tures (n = 2-5) via Dnh transition states as models for hydrogen tunneling in hydrogen fluoride
clusters, Chem. Phys. Lett. 246, 455-62 (1995).
37. Quack, M., Schmitt, U., and Suhm, M. A., Evidence for the (HF)5 complex in the HF stretching
FTIR absorption spectra of pulsed and continuous supersonic jet expansions of hydrogen fluo-
ride, Chem. Phys. Lett. 208, 446-452 (1993).
38. Luckhaus, D., Quack, M., Schmitt, U., and Suhm, M. A., On FTIR spectroscopy in asynchro-
nously pulsed supersonic free jet expansions and on the interpretation of stretching spectra of
HF clusters, Ber. Bunsenges. Phys. Chem. 99, 457-468 (1995).
39. Karpfen, A., and Yanovitskii, O., Structure and vibrational spectra of neutral, protonated and
deprotonated hydrogen bonded polymers: an ab initio SCF study on chain-like hydrogen fluo-
ride clusters, J. Mol. Struct. (Theochem) 307, 81-97 (1994).
40. Postorino, P., Ricci, M. A., and Soper, A. K., Water above its boiling point: Study of the tem-
perature and density dependence of the partial pair correlation functions. I. Neutron diffraction
experiment, J. Chem. Phys. 101, 4123-4132 (1994).
41. Gorbaty, Y. E., and Kalinichev, A. G., Hydrogen bonding in supercritical water. 1. Experimen-
tal results, J. Phys. Chem. 99, 5336-5340 (1995).
42. Mizan, T. I., Savage, P. E., and Ziff, R. M., Temperature dependence of hydrogen bonding in su-
percritical water, J. Phys. Chem. 100, 403-408 (1996).
43. Saint-Martin, H., Medina-Llanos, C., and Ortega-Blake, I., Nonadditivity in an analytical inter-
molecular potential: The water-water interaction, J. Chem. Phys. 93, 6448-6452 (1990).
44. Lybrand, T. P., and Kollman, P. A., Water-water and water-ion potential functions including
terms for many body effects, J. Chem. Phys. 83, 2923-2933 (1985).
45. Cieplak, P., Kollman, P., and Lybrand, T., A new water potential including polarization: Appli-
cation to gas-phase, liquid, and crystal properties of water, J. Chem. Phys. 92, 6755-6760
(1990).
46. Campbell, E. S., and Mezei, M., Use of a non-pair-additive intermolecular potential function to
fit quantum-mechanical data on water molecule interactions, J. Chem. Phys. 67, 2338—2344
(1977).
47. Watanabe, K., and Klein, M. L., Effective pair potentials and the properties of water, Chem.
Phys. 131, 157-167(1989).
48. Wojcik, M., and Clementi, E., Single molecule dynamics of three body water, J. Chem. Phys. 85,
3544-3549 (1986).
49. Berendsen, H. J. C., Grigera, J. R., and Straatsma, T. P., The missing term in effective pair po-
tentials, J. Phys. Chem. 91, 6269-6271 (1987).
50. Brodholt, J., Sampoli, M., and Vallauri, R., Parameterizing a polarizable intermolecular poten-
tial for water, Mol. Phys. 86, 149-158 (1995).
51. Gregory, J. K., and Clary, D. C., Three-body effects on molecular properties in the water trimer,
J. Chem. Phys. 103, 8924-8930 (1995).
52. Gao, J., Habibollazadeh, D., and Shao, L., A polarizable intermolecular potential function for
simulation of liquid alcohols, J. Phys. Chem. 99, 16460-16467 (1995).
53. Caldwell, J. W., and Kollman, P. A., Structure and properties of neat liquids using nonadditive
molecular dynamics: Water, methanol, and N-methylacetamide, J. Phys. Chem. 99, 6208-6219
(1995).
288 Hydrogen Bonding

54. Clementi, E., Kolos, W., Lie, G. C., and Ranghino, G., Nonadditivity of interaction in water
trimers, Int. J. Quantum Chem. 17, 377-398 (1980).
55. Scheiner, S., and Nagle, J. P., Ab initio molecular orbital estimates of charge partitioning be-
tween Bjerrum and ionic defects in ice, J. Phys. Chem. 87, 4267-4272 (1983).
56. Karpfen, A., and Schuster, P., Ab initio studies on hydrogen bonded chains. V. The structure of
infinite chains of methanol and water molecules, Can. J. Chem. 63, 809-815 (1985).
57. Suhai, S., Cooperative effects in hydrogen bonding: Fourth-order many-body perturbation the-
ory studies of water oligomers and of an infinite water chain as a model for ice, J. Chem. Phys.
101,9766-9782(1994).
58. Ojamae, L., and Hermansson, K., Ab initio study of cooperativity in water chains: Binding en-
ergies and anharmonic frequencies, J. Phys. Chem. 98, 4271-4282 (1994).
59. Yoon, B. J., Morokuma, K., and Davidson, E. R., Structure of ice Ih. Ab initio two- and three-
body water-water potentials and geometry optimization, J. Chem. Phys. 83, 1223—1231 (1985).
60. Hermansson, K., Many-body effects in tetrahedral water clusters, J. Chem. Phys. 89,2149-2159
(1988).
61. Koehler, J. E. H., Saenger, W., and Lesyng, B., Cooperative effects in extended hydrogen bonded
systems involving O—H groups. Ab initio studies of the cyclic S4 water tetramer, J. Comput.
Chem. 8, 1090-1098(1987).
62. Dyke, T. R., and Muenter, ]. S., Molecular beam electric deflection studies of water polymers,
J. Chem. Phys. 57, 5011-5012 (1972).
63. Vernon, M. E, Krajnovich, D. J., Kwok, H. S., Lisy, J. M., Shen, Y. R., and Lee, Y. T., Infrared
vibrational predissociation spectroscopy of water clusters by the crossed laser-molecular beam
technique, J. Chem. Phys. 77, 47-57 (1982).
64. Honegger, E., and Leutwyler, S., Intramolecular vibrations of small water clusters, J. Chem.
Phys. 88, 2582-2595 (1988).
65. Schiitz, M., Burgi, T., Leutwyler, S., and Burgi, H. B., Fluxionality and low-lying transition
structures of the water trimer; J. Chem. Phys. 99, 5228-5238 (1993).
66. Knochenmuss, R., and Leutwyler, S., Structures and vibrational spectra of water clusters in the
self-consistent-field approximation, J. Chem. Phys. 96, 5233-5244 (1992).
67. White, J. C., and Davidson, E. R., An analysis of the hydrogen bond in ice, J. Chem. Phys. 93,
8029-8035(1990).
68. Chalasinski, G., Szczesniak, M. M., Cieplak, P., and Scheiner, S., Ab initio study of intermolec-
ular potential of H2O trimer, J. Chem. Phys. 94, 2873-2883 (1991).
69. Klopper, W., Schiitz, M., Luthi, H. P., and Leutwyler, S., An ab initio derived torsionalpotential
energy surface for (H 2 O) 3 . H. Benchmark studies and interaction energies, J. Chem. Phys. 103,
1085-1098(1995).
70. Mo, O., Yanez, M., and Elguero, J., Cooperative (nonpairwise) effects in water trimers: An ab
initio molecular orbital study, J. Chem. Phys. 97, 6628-6638 (1992).
71. Kay, B. D., and Castleman, Jr., A. W., Molecular beam electric deflection study of the hydrogen-
bonded clusters (H 2 O) N , (CH 3 OH) N , and(C 2 H 5 OH) N , J. Phys. Chem. 89, 4867-4868 (1985).
72. Engdahl, A., and Nelander, B., On the structure of the water trimer. A matrix isolation study, J.
Chem. Phys. 86, 4831-1837 (1987).
73. Ebata, T., Watanabe, T., and Mikami, N., Evidence for the cyclic form of phenol trimer: Vibra-
tional spectroscopy of the OH stretching vibrations of jet-cooled phenol dimer and trimer, J.
Phys. Chem. 99, 5761-5764 (1995).
74. Xantheas, S. S., and Dunning, T. H. J., Ab initio studies of cyclic water clusters (H2O)n, n = 1-6.
1. Optimal structures and vibrational spectra, J. Chem. Phys. 99, 8774-8792 (1993).
75. Gerhards, M., and Kleinermanns, K., Structure and vibrations of phenol(H2O)2, J. Chem. Phys.
103,7392-7400(1995).
76. van Duijneveldt-van dc Rijdt, J. G. C. M., and van Duijneveldt, F. B., Ab initio potential energy
surface for the low-frequency out-of-plane bending motions of the water trimer, Chem. Phys.
Lett. 237, 560-567(1995).
Cooperative Phenomena 289

77. Sabo, D., Bacic, Z., Burgi, T., and Leutwyler, S., Three-dimensional model calculation of tor-
sional levels of (H2O)3 and (D2O)3, Chera. Phys. Lett. 244, 283-294 (1995).
78. Cruzan, J. D., Braly, L. B., Liu, K., Brown, M. G., Loeser, J. G., and Saykally, R. L, Quantify-
ing hydrogen bond cooperativity in water: VRT spectroscopy of the water tetramer, Science 271,
59-62 (1996).
79. Schiitz, M., Klopper, W., Luthi, H.-P., and Leutwyler, S., Low-lying stationaiy points and torsional
interconversiom of cyclic (H2O)4: An ab initio study, J. Chem. Phys. 103, 6114-6126 (1995).
80. Packer, M. J., and Clary, D. C., Interaction of HCl with water clusters: (H2O)nHCl, n = 1-3, J.
Phys. Chem. 99, 14323-14333 (1995).
81. Liu, K., Cruzan, J. D., and Saykally, R. J., Water clusters, Science 271, 929-933 (1996).
82. Burgi, T., Schiitz, M., and Leutwyler, S., Intermolecular vibrations of phenol-(H 2 O) 3 and
d 1 -phenol-(D 2 O) 3 in the So and S1 states, J. Chem. Phys. 103, 6350-6361 (1995).
83. Liu, K., Brown, M. G., Cruzan, J. D., and Saykally, R. J., Vibration-rotation tunneling spectra
of the water pentamer: Structure and dynamics, Science 271, 62-64 (1996).
84. Teeter, M. M., Water structure of a hydrophobic protein at atomic resolution: Pentagon rings of
water molecules in crystals of crambin, Proc. Nat. Acad. Sci., USA 81, 6014-6018 (1984).
85. Jensen, J. O., Krishnan, P. N., and Burke, L. A., Theoretical study of water clusters: heptamers,
Chem. Phys. Lett. 241, 253-260 (1995).
86. Jensen, J. O., Krishnan, P. N., and Burke, L. A.., Theoretical study of water clusters: octamer,
Chem. Phys. Lett. 246, 13-19 (1995).
87. Kim, J., Mhin, B. J., Lee, S. J., and Kim, K. S., Entropy-driven structures of the water octamer,
Chem. Phys. Lett. 219, 243-246 (1994).
88. Badger, R. M., The relation between the internuclear distance and force constants of molecules
and its application to polyatomic molecules, J. Chem. Phys. 3, 710-714 (1935).
89. Badger, R. M., A relation between internuclear distance and bond force constants, J. Chem.
Phys. 2, 128-131 (1934).
90. Huisken, R, Kaloudis, M., and Kulcke, A., Infrared spectroscopy of small size-selected water
clusters, J. Chem. Phys. 104, 17-25 (1996).
91. Xantheas, S. S., Ab initio studies of cyclic water clusters (H2O)n, n = 1-6. II. Analysis of many-
body interactions, J. Chem. Phys. 100, 7523-7534 (1994).
92. van Duijneveldt-van de Rijdt, J. G. C. M., and van Duijneveldt, F. B., Ab initio calculations on
the geometry and OH vibrational frequency shift of cyclic water trimer, Chem. Phys. 175,
271-281 (1993).
93. Fowler, J. E., and Schaefer, H. R, Detailed study of the water trimer potential energy surface, J.
Am. Chem. Soc. 117,446-452 (1995).
94. Liu, K., Loeser, J. G., Elrod, M. J., Host, B. C., Rzepiela, J. A., Pugliano, N., and Saykally, R.
J., Dynamics of structural rearrangements in the water trimer, J. Am. Chem. Soc. 116,
3507-3512(1994),
95. Burgi, T., Graf, S., Leutwyler, S., and Klopper, W., An ab initio derived torsional potential en-
ergy surface for (H2O)3. I. Analytical representation and stationaiy points, J. Chem. Phys. 103,
1077-1084(1995).
96. Peelers, D., and Leroy, G., Small clusters between water and alcohols, J. Mol. Struct.
(Theochem) 314, 39-47 (1994).
97. Mo, O., Yanez, M., and Elguero, J., Cooperative effects in the cyclic trimer of methanol. An ab
initio molecular orbital study, J. Mol. Struct. (Theochem) 314, 73-81 (1994).
98. Hinchliffe, A., Ab initio study of the hydrogen-bonded complexes NH 3 . . . HX, PH 3 . . . HX and
NH 3 . . . (HX) 2 , where X=F,Cl, J. Mol. Struct. (Theochem) 105, 335-341 (1983).
99. Kurnig, I. J., Szczesniak, M. M., and Scheiner, S., Ab initio study of structure and cooperativity
in H 3 N . . . HF . . . HF and H 3 P . . HF . . . HF, J. Phys. Chem. 90, 4253-4258 (1986).
100. Kurnig, I. J., Szczesniak, M. M.. and Scheiner, S., Vibrational frequencies and intensities of H-
bonded systems. 1:1 and 1:2 complexes of NH3 and PH3 with HF, J. Chem. Phys. 87, 2214-2224
(1987).
290 Hydrogen Bonding

101. Bertie, J. E., and Falk, M. V., The infrared spectrum of the hydrogen-bonded molecule dimethyl
ether... hydrogen chloride in the gas phase, Can. J. Chem. 51, 1713—1720 (1973).
102. Ramos, M. N., Gussoni, M., Castiglioni, C., and Zerbi, G., Ab initio counterpart of infrared
atomic charges. Comparison with charges obtained from electrostatic potentials, Chem. Phys.
Lett. 151,397-402(1988).
103. Gussoni, M., Ramos, M. N., Castiglioni, C., and Zerbi, G., Ab initio counterpart of infrared
atomic charges, Chem. Phys. Lett. 142, 515-518 (1987).
104. Gussoni, M., Jona, P., and Zerbi, G., Atomic charges and charge flows from infrared intensities:
C—H bonds, J. Chem. Phys. 78, 6802-6807 (1983).
105. Cioslowski, J., A new population analysis based on atomic polar tensors, J. Am. Chem. Soc. 111,
8333-8336(1989).
106. Person, W. B., and Zerbi, G., eds. Vibrational intensities in infrared and Raman spectroscopy;
Elsevier, Amsterdam (1982).
107. Scheiner, S., Proton transfers in hydrogen bonded systems. 6. Electronic redistributions in
(N2H7)+ and (O2H5)+, J. Chem. Phys. 75, 5791-5801 (1981).
108. Zilles, B. A., and Person, W. B., Interpretation of infrared intensity changes on molecular com-
plex formation. I. Water dimer, J. Chem. Phys. 79, 65-77 (1983).
109. Gussoni, M., Castiglioni, C., and Zerbi, G., Charge distribution for infrared intensities: Charges
on hydrogen atoms and hydrogen bond, J Chem. Phys. 80, 1377-1381 (1984).
110. Hannachi, Y., Silvi, B., and Bouteiller, Y, Ab initio study of the structure, cooperativity, and vi-
brationalproperties of the H2O:(HF)2 hydrogen bonded complex, J. Chem. Phys. 97,1911-1918
(1992).
111. Latajka, Z., and Scheiner, S., Basis sets for molecular interactions. 1. Construction and tests on
(HF)2 and (H2O)2, J. Comput. Chem. 5, 663-673 (1987).
112. Rovira, C., Constans, P., Whangbo, M.-H., and Novoa, J. J., Theoretical study of the structure
and Vibrational spectra of the (H 2 O) 2 . . . HF and H 2 O . . . (HF) 2 molecular complexes, Int. J. Quan-
tum Chem. 52, 177-189 (1994).
113. Amirand, C., and Maillard, D., Spectrum and structure of water-rich water-hydracid complexes
from matrix isolation spectroscopy: Evidence for proton transfer, J. Mol. Struct. 176, 181—201
(1988).
6

Weak Interactions, Ionic H-Bonds,


and Ion Pairs

he classic picture of a H-bond involves the approach of a pair of neutral molecules. One
T contains a hydrogen atom covalently attached to a highly electronegative atom like O
or F. The other molecule also contains an electronegative atom, associated with which is at
least one lone pair of nonbonding electrons. It is perhaps oversimplistic to expect that a spe-
cific interaction either does or does not represent a true H-bond,with no grey area between.
For example, one can imagine a scenario where a genuine H-bond, as in FH...FH, becomes
weaker and weaker as one or the other F atom is replaced by less electronegative halogen
atoms. The HI dimer is clearly not associated via a H-bond, whereas C1H...C1H arguably
contains a H-bond, albeit a weak one. Another example replaces the proton-accepting FH
molecule by a much less polar one, such as FF. Does the absence of a molecular dipole mo-
ment in the proton acceptor preclude the existence of a H-bond, even if the latter does con-
tain a lone electron pair? Still another case in point takes as a starting point the equilibrium
geometry of a classic H-bond such as FH...FH. As the two molecules are pulled apart, the
interaction clearly becomes weaker and weaker. At what point would one cease to catego-
rize this interaction as a H-bond?
It would probably be best to think of interactions as spanning a continuum. In the mid-
dle of this continuum are the classic H-bonding interactions. On one end are those that are
much weaker and clearly not of the H-bonded variety. On the other are those that are con-
siderably stronger, typically dominated by electrostatic factors. It then becomes somewhat
arbitrary as to where on this continuum one draws the line between a true H-bond and a dif-
ferent sort of interaction. Interaction energy can be used as one measure, and one can as-
sign a minimum and maximum strength to the definition of a H-bond. Or the magnitude of
the red shift of the A—H stretching frequency can be used as an indicator, with an arbitrary
threshold assigned. But regardless of how carefully one designs the criteria, it must be un-
derstood from the outset that one person's H-bond is another person's van der Waals com-
plex, or another's Coulombic interaction.

291
292 Hydrogen Bonding

Some of the foregoing chapters have included discussions of borderline H-bonds. The
ammonia dimer, for example, is very weakly bound and its potential energy surface is so
flat that there is no clearly defined equilibrium geometry. The definition as H-bonds of some
of the systems that contain second or third-row atoms is also questionable.
This chapter considers a number of other types of interactions that are somewhere near
the boundaries of a true H-bond. We discuss the details of these interactions and the mag-
nitudes of some of the indicators. One issue discussed is the proton-accepting ability of an
electronegative atom when involved in a bond to another electronegative atom, leaving the
bond of low or zero polarity. Hydrogen atoms bonded to carbon are typically of low acid-
ity, so their ability to participate in H-bonds is questionable as well.
On the opposite extreme are H-bonds wherein one of the two partners bears an electric
charge. These interactions are severalfold stronger than H-bonds involving a pair of neutral
molecules. They are dominated by electrostatics and the bridging proton tends to drift so
far from the proton donor atom that it becomes questionable as to which unit is the donor
and which the acceptor. The strong role of electrostatics can call into question their cate-
gorization as H-bonds. For example, the fundamental nature of the interaction in
(H 3 NH +... NH 3 ) is very similar indeed to that in (K +... NH 3 ) where there is clearly no H-
bond present1.
In addition to the so-called ionic H-bonds in which one of the two partners bears an elec-
tric charge, there is the further possibility that both of the subunits might be charged. An ion
pair, represented in general as A - . . . + H B , is also dubbed a "salt bridge" on occasion. The
opposite charges of the two ions have the potential of producing a very strong attractive
force between them; such forces have been implicated in stabilizing certain peptide con-
formations, for example 2-4 . There is some question as to whether a system of this type is
truly a H-bond or would be better characterized as a simple electrostatic interaction. What
makes them particularly fascinating is the fact that the entire nature of this system can be
changed to the more conventional H-bond between neutral molecules, AH ... B, by a simple
proton transfer from the cation to the anion. It becomes an interesting and relevant question
as to what conditions would cause a neutral pair to convert to an ion pair.

6.1 Weak Acceptors

When two HX molecules are paired together, it is clear that a H-bond can form between the
proton of one molecule and the X atom of the other. But the situation is less clear cut if the H
atom of the proton acceptor molecule is replaced by another halogen atom. In such a case, the
dihalogen will not be polarized much at all, lessening the electrostatic part of the interaction.

6.1.1 Dihalogens
Ab initio calculations5 considered the interaction of HF with C1F and with C12. Two min-
ima were found in each case. The first corresponds to the standard sort of arrangement
which can be denoted as H-bonded, at least from a geometric perspective. The other mini-
mum is clearly not H-bonding as it is the F atom of HF that approaches the halogen atom.
As indicated in Fig. 6.1, the latter geometry is designated as "F-bonding."
The authors used 4-31G to probe the general character of the potential energy surface.
They then used a larger basis set, representing H by [3slp], F by [5s3pld], and Cl by
[7s5pld], basically a TZP set. Correlation was included via a coupled pair functional (CPF)
Weak Interactions, Ionic H-Bonds, and Ion Pairs 293

Figure 6.1 Two types of geometries


for HF + C1X (X = F,C1).

formalism6, which considers single and double excitations and adds the effects of quadru-
ple and higher excitations in an approximate way. It was found that there is a delicate bal-
ance between the stabilities of the two geometries. Failure to include counterpoise correc-
tions, correlation, or zero-point vibrational effects can lead to the wrong conclusion as to
which configuration is preferred. Indeed, earlier calculations at the SCF level7,8 had com-
puted comparable stabilities of these two structures, and been unable to account for the ex-
perimental observation of only one of them9.
The sensitivity is illustrated in Table 6.1 which shows that at the SCF level the two
geometries are within 0.1 kcal/mol of each other, but that the H-bonded structure is favored
after correlation is included. Once counterpoise corrections are added, however, the F-
bonded structure is preferred by 0.1 kcal/mol at either level. This preference is amplified
by zero-point vibrational energies, resulting in a greater stability of the F-bond by 0.6
kcal/mol. Later computations using a larger basis set10 confirmed the preference for the F-
bonded structure at the correlated level, in this case by 0.3 kcal/mol. This same preference
is confirmed by experimental observation9 that only the F-bonded geometry occurs in the
gas phase. One caution arises from earlier computations that noted that despite a computed
electronic contribution to the binding energy of these complexes by as much as —3
kcal/mol, addition of vibrational and entropic factors leads to a much smaller free energy
of complexation at 100° K; G becomes positive at higher temperatures, reaching +3
kcal/mol at 298° K7.
Calculations5 lead to a similar conclusion for the HF + C12 pair: both of these structures
have been subsequently observed by their infrared spectra in solid Ar and Ne11. Due to the
questionable existence of the H-bonded geometry, as well as the very weak interaction en-
ergy of only 2 kcal/mol or less, one can conclude that the dihalogen molecule does not act
as a proton acceptor in a H-bond.
It is interesting to note that dihalogens can interact also with a base like NH3, in a com-
plex of the H 3 N ... XX' type, even though there is no proton to act as bridge12-15. In fact, the

Table 6.1 Interaction energies of HF + GIF complex, computed with TZP basis set. Data
in kcal/mol5.
SCF C PF

F-bonded H-bonded F-bonded H-bonded

E
elec
-2.44 -2.52 -2.97 -3.31
E
elec + CCa -2.01 -1.90 -2.08 -2.00
e
E lec + CCa + ZPVE -133 -0.72 -1.43 -0.83

Counterpoise correction.
294 Hydrogen Bonding

experimental results suggest that the binding energy of H3N...C1F may even be compara-
ble to that in the clearly H-bonded H 3 N ... HF complex16.

6.1.2 CO
If HF will not form a H-bond to a nonpolar dihalogen molecule, what sort of interaction
might form with a molecule that has slightly more polar bonds? Calculations to answer this
question were carried out 17-19 incorporating electron correlation. It was found that CO par-
ticipates in two different minima with HF, both of which have the general structure of a H-
bond. They are both linear with the proton pointing toward the O or C atom: OC ... HF and
CO...HF. As listed in Table 6.2, the former is favored over the latter by some 1.8 kcal/mol17.
In fact, the latter is barely bound at all after zero-point vibrational energies are included.
The greater stability of OC ... HF is consistent with the idea that the dipole moment of CO
is negative on the C side, so it is this atom that can better attract the partially positively
charged hydrogen of HF. Continuing this qualitative reasoning, UV photofragmentation
measurements of this complex lead to an enthalpic preference for OC ... HF by about 0.1
kcal/mol20.
The C...F distance is 2.99 A for the more stable structure, and R(O ... F) = 3.00 A for the
other, both on the long end of the range of H-bond distances. Arguing for the presence of a
H-bond is a red-shift in the HF stretch of 154 cm-1 in OC...HF, but this shift is only 20
cm-1 in the weaker complex. Also notable is the stretch in the HF bond brought about by
complexation. This stretch is 0.006 A in OC...HF, but practically zero for the other. It ap-
pears that OC...HF has many of the characteristics expected of a H-bond, but CO ... HF
clearly does not.
When CO is paired with water, one again sees the former molecule acting as a formal
proton acceptor. SCF calculations had suggested that the role of acceptor atom can be served
equally well by both the C and O atoms21, although inclusion of correlation led to a marked
preference for OC...HOH. The binding energy was computed to be weaker than that in
OC ... HF by 1 or 2 kcal/mol. The details of the complex in the gas phase22 confirm the pres-
ence of OC...HOH and would lead one to doubt that it is a true H-bond. The O and C atoms
are separated by some 3.37 A and the bridging hydrogen is located 11.5° from the H-bond
axis.
Improved calculations utilizing a polarized basis set designed to accurately mimic mole-
cular properties of the monomer23 were able to reproduce the experimental tilt angle of 11°
and reaffirmed the preference for the OC ... HOH geometry. This nonlinearity was traced to a
balance between electrostatic attraction and exchange repulsion. Later calculations demon-

Table 6.2 Characteristics of various complexes, computed at MP2/TZP level, and corrected by
counterpoise procedure17.

- Eelec - E + ZPVE R(A...F) r(HF) v(HF)


(kcal/mol) (kcal/mol) (A) (A) (cm -1 )

OC ... HF 3.59 1.83 2.99 0.006 -154


CO ... HF 1.14 0.03 3.00 0.001 -20
OCO ... HF 2.42 1.12 2.86 0.003 -47
NNO ... HF 2.99 1.59 2.87 0.004 -85
ONN ... HF 1.93 0.67 2.91 0.004 -87
Weak Interactions, Ionic H-Bonds, and Ion Pairs 295

strated that the nonlinearity does not require a particularly large basis set for quantitative re-
production; diffuse functions are generally sufficient24. The electronic contribution to the
binding energy was computed to be 1.9 kcal/mol at the MP4 level with a basis set including
f functions. This interaction is due largely to the electrostatic and dispersion energies; terms
corresponding to deformation of the electron density (e.g., induction) contribute little. Cou-
pled-cluster computation of the binding energy, CCSD(T), using a 6-31 + +. G(d,p) basis set
and correcting for BSSE, yielded an electronic binding energy of 1.4 kcal/mol24. The weak-
ness of this interaction casts further doubt upon its categorization as a H-bond.
Other minima present on the surface correspond to CO...HOH, bound by less than 1
kcal/mol, and a higher-energy T-shape wherein the O atom of water approaches the C—O
midpoint. However, a later work classified the T-structure as a saddle point on the PES25.
The possible categorization of OO...HOH as a H-bond is discounted by the r(OH) stretch
which is only 0.0007 A (although some other flexible basis sets predict stretches of as much
as 0.002 A24). The OH stretching frequencies of water were computed at the MP2 level to
be red-shifted by only 15 to 18 c m - 1 , further arguments against the presence of a H-bond25.
The intermolecular stretching frequency, V , was calculated to be 101 c m - 1 . Other com-
putations comparing a range of basis sets24 found the red shift of the asymmetric OH stretch
to be anywhere from 5 to 22 c m - 1 . The shifts for the higher-energy CO...HOH minimum
are to the blue.
In summary, while one might convincingly argue for a H-bond in the complex between
CO and HF, such an interaction in OC...HOH is more dubious.

6.1.3 C02
When CO is replaced by CO2, the proton acceptor no longer has a net dipole moment at all.
On the other hand, the C=O bond's polarity might be expected to favor a H-bond, even if
a weak one. Calculations reported in Table 6.2 revealed a linear OCO...HF structure as the
only minimum on the potential energy surface17, conforming to experimental observation
[26] (although the bending potential is extraordinarily flat27). Indeed, the insensitivity of
the energy to bending is underscored by later studies where MP4 calculations confirmed a
linear structure28 whereas others obtained a bent geometry, with (F..OC) = 146°29. The
interaction energy of this structure is intermediate between OC ... HF and CO...HF, but the
H-bond length is shorter than in either of the others. The red shift of the HF stretching mode
is rather small, only 47 c m - 1 , and the HF stretch is 0.003 A. This arrangement would prob-
ably be considered as a marginal sort of H-bond at best.
The structure of the complex appears to retain its linear character when HF is replaced
by HC1, on the basis of SCF computations, coupled with observation of IR and Raman spec-
tra in Ar matrices30 or IR absorption31 and microwave spectroscopy in the gas phase32.
When CO2 is paired with HBr, the geometry loses its H-bond character: the H atom ap-
proaches the C atom, with the HBr axis perpendicular to OCO31,33,34. (Another recent mi-
crowave investigation suggested that it is the Br atom that more closely approaches the C
of OCO35, with the HBr molecule undergoing large-amplitude oscillations about the C .. Br
axis). A similar T-shaped geometry occurs as well when HF is replaced by HCN: in this
case, it is the N atom which approaches the C from above36. Calculations37 find this struc-
ture less stable than the linear configuration but only by about 0.2 kcal/mol, easily within
the margin of error.
Whereas HF appears to be a strong enough proton donor that it forms what has some of
the characteristics of a H-bonded complex with OCO, the same is apparently not true of wa-
296 Hydrogen Bonding

ter. When paired with OCO, the water molecule approaches OCO O-atom first to form the
complex indicated in Fig. 6.238, not resembling a H-bond in any way. The structure can be
described in terms of a T shape, permitting the negative end of the H2O dipole to approach
the partially positively charged C atom. This finding confirmed experimental measure-
ments39,40 and earlier calculations41 concerning this structure, but the latter computations
had noted another minimum on the surface, also illustrated in Fig. 6.2, as containing what
appears to be a H-bond. At the MP2 level, with a 6-311+G* basis set, the H-bonding geom-
etry is bound by 2.0 kcal/mol (without counterpoise correction), as compared to 3.4
kcal/mol for the non-H-bonded structure. Due to the weak nature of the interaction, the
stretch of only 0.001 A in the bridging O—H bond, and the long distance between O atoms
(more than 3.25 A), it would not be appropriate to refer to this interaction as a true H-bond.
A slightly larger basis set (6-31+G(2d,2p)) obtained similar binding energies, De, of 3.0
and 2.2, respectively40, for the two geometries in Fig. 6.2, but another work could not lo-
cate the less stable of the two42.
Later computations of this same system43 were improved in the sense that counterpoise
corrections were added and a larger basis set was employed (D95 + +(3d,2p)). All parame-
ters were fully optimized and minima verified by frequency analysis. This work verified
that the second structure in Fig. 6.2, which has certain characteristics of a H-bond, is a min-
imum on the surface. The electronic contributions to the binding energies of the T-shape
and secondary minima are 2.2 and 1.3 kcal/mol, respectively, at the MP2 level. Despite the
very weak binding energy of the secondary minimum, the OH bond of the water is stretched
by 0.005 A as a result of the complexation. On the other hand, this indication of a possible
H-bond is belied by the observation that the stretching frequencies of the water molecule
are virtually unchanged in the complex.
Hence, in complexes involving both CO and OCO, HF appears to be a strong enough
proton donor to form a H-bond (albeit a debatable one) whereas there is no such interaction
present in complexes with water.

6.1.4 NNO
NNO is isoelectronic with OCO. There are several possible sites where a proton might be
accepted were a H-bond to form. Calculations of HF + NNO17 locate two minima reported
in Table 6.2. When HF adds to the end N site, the entire complex is linear as in OCO...HF.
On the other hand, interaction with the O atom leads to a bent structure as indicated in Fig.
6.3. Both are found to be minima on the PES, but the interaction with the oxygen atom is
significantly stronger, 3.0 kcal/mol prior to addition of zero-point vibrational energies. The
presence of both minima, and their shapes, conform to experimental observations44-46. The
red shift of the HF stretch in the more stable geometry is 85 c m - 1 , and 87 cm-1 in the less

Figure 6.2 Structures of complexes of H2O with CO2.


Weak Interactions, Ionic H-Bonds, and Ion Pairs 297

Figure 6.3 Two minima in PES of HF + NNO.

stable linear structure, lending credence to the presence of a H-bond here. This contention
is supported by the 0.004 A stretches of the HF bond length.
The study of NNO + HF was later extended to NNO + HC147 in which case two min-
ima were identified roughly corresponding to those in Fig. 6.3, in which the HC1 molecule
donates a proton to either the N or O end of the NNO molecule (although the C1H...ONN
complex is nearly linear). But a third minimum was also located, which does not have the
appearance of a H-bond. As illustrated in Fig. 6.4, the two molecules are nearly parallel, and
it is this structure which has been observed experimentally48,49. The lesser proton donating
ability of HC1 versus HF is thus responsible for the loss of any H-bonding in the complex
with NNO. Reproduction of the preference for this structure required surprisingly high lev-
els of theory, as most calculations erroneously predicted one of the H-bonded geometries to
be more stable. Hence, the complex between NNO and HC1 does not contain a H-bond but
lower-level calculations of this sort of system could easily mislead the unwary researcher.

6.1.5 SO
2

The SO2 molecule is like NNO in that it contains several potential sites, including O atoms,
that might accept a proton. HF forms a H-bonded complex with the oxygen atoms of the
SO2 molecule, with a nearly linear F—H . . . O arrangement50. This complex can either be of
cis or trans type, but there is little difference in energy between the two. At the MP2/TZP
level, and with appropriate counterpoise correction of the BSSE, the electronic contribu-
tions to the binding energy are —4.44 and —4.35 kcal/mol for the cis and trans geometries,
respectively. (In light of the later results described below for HCN, it is important to note
that the geometry optimizations assumed a plane of symmetry in this complex.)
HCN is comparable in strength as a proton donor to HF so it too can serve to test the
proton accepting ability of certain molecules. Calculations that pair HCN with SO251 were
unable to locate as a minimum on the potential energy surface a geometry that would cor-
respond to proton donation by HCN to any site on the SO2 molecule. The only stable geom-
etry identified has the line of the HCN molecule lying nearly perpendicular to the plane of

Figure 6.4 Preferred geometry of complex of NNO with HC147.


298 Hydrogen Bonding

SO2, with the N of the former approaching the S atom of the latter, consistent with spec-
troscopic indications52,53. One can conclude that the electron pairs of the S and O atoms
are not competitive to attract the HCN proton, when compared with the factors leading to
the geometry observed.
SO2 can be paired with another proton donor, H2O, which is a weaker base than the N
of HCN so one would not expect a complex of the type formed with HCN. Two different
structures appear to represent minima on the H2O/SO2 surface54. The more stable of the
two has the planes of the two molecules roughly parallel, with no H-bond present, and is
consistent with microwave data55. The secondary minimum has some of the geometric fea-
tures of a H-bond between a proton of H2O and one of the SO2 oxygens. However, this com-
plex has only half the binding energy of the primary minimum. And one might question the
existence of a true H-bond here as the OH bond is stretched by only 0.002 A and this stretch-
ing frequency undergoes a red shift of only 8 cm-1 upon forming the complex. Further-
more, the bridging proton is 2.144 A distant from the acceptor oxygen atom.
Nor is a H-bond present in the complex formed when SO2 is paired with the weaker pro-
ton donor H2S, neither on the basis of calculations54, nor from microwave measurements56.
Taking this into consideration, it is hence not surprising that the weak proton donor, HCCH,
also does not form a H-bonding interaction with SO257, nor does CH3OH58.

6.1.6 CC12
Carbenes, wherein a central carbon atom is bound to only two other atoms via single bonds,
present the possibility of a singlet and triplet electronic state that are close in energy. In the
case of CC12, it appears that the unusual electronic structure reverses the normally expected
electronegativities. More specifically, the carbon seems to act as a better proton acceptor
than do the chlorine atoms. When paired with water as a proton donor, the only minimum
on the SCF/DZP surface has a classical H-bonding arrangement, wherein the water proton
acts as a bridge to the carbon59. The H-bond is longer than is typical: R(O..C) is equal to
3.32 A. However, the strength of the interaction is in the normal H-bonding range. After in-
clusion of correlation via MP2, with appropriate BSSE correction, the electronic contribu-
tion to the binding energy is —3.9 kcal/mol. Addition of ZPVE leads to a best estimate of
the dissociation energy of 2.4 kcal/mol.

6.2 C—H as Proton Donor

The matter of whether the carbon atom can act as a proton donor in a H-bond has been dis-
cussed for some time60,63. A systematic analysis of crystallographic data for a large col-
lection of molecules had shown a statistically meaningful tendency for C—H hydrogens to
approach oxygen, as compared to C or H atoms64. In these cases, the C—H hydrogen lies
within about 30° of the plane containing the lone electron pairs of the O atom and ap-
proaches within the sum of van der Waals contact radii. The data also suggest that N and Cl
can act as proton acceptors. The authors were reluctant to classify these interactions as true
H-bonds, based as they are on purely geometric considerations. Examining H-bonds within
the context of a crystal also subjects the results to crystal packing forces which might be
misleading65.
Implementation of neutron diffraction data allowed much better refinement of the posi-
tion of the hydrogens, and thus of the putative H-bonds themselves. Of particular interest
Weak Interactions, Ionic H-Bonds, and Ion Pairs 299

due to their thorough study, and the prevalence of C—H donor groups and O acceptors, are
the carbohydrates66,67. A thorough study of 26 carbohydrate structures68 found the short-
est C—H . . . O contact to have r(H ... O) = 2.27 A, considerably longer than is usually con-
sidered a H-bond; most of the contacts identified are even longer. The C—H bond stretches
no more than 0.004 A in even the shortest contact, at the limit of experimental accuracy. A
later extension examined structures in which water serves as the proton acceptor69. Again,
there were no contacts found with r(H ... O) < 2.3 A. Instead of arguing for the existence of
true H-bonds, the authors point out that the proton acceptor molecules may resort to a C—H
donor rather than leaving a lone pair with no interaction at all. A survey of crystal structures
of organometallics70 provides additional indications that C—H protons might form a Hi-
bond to oxygen. The former can approach the latter from many directions, with a peak in
the neighborhood of 0 (CO..H) = 140°. On the other hand, any such H-bond is rather long
with C..O distances typically greater than 3.3 A, peaking at around 3.4-3.5 A.

6.2.1 Alkynes
It is well known that carbon can act as a proton donor within the context of a molecule like
HCN where the triply bonded C behaves like a more electronegative atom. The bridging
hydrogen and the proton-accepting N atom of the N=CH ... NH 3 complex, for example, are
separated by 2.33 A71. Like HCN, acetylene also contains a hydrogen atom which is acidic
due to its placement on a carbon which is involved in a triple bond. When paired with a po-
tential proton acceptor like water, HCCH will indeed use its proton to act as a bridge72,73,
as it will when alkynes are paired with other oxygen proton acceptors74 or with N-bases75.
In fact, there is even evidence that alkynes will donate a proton to bases containing second
or third-row acceptor atoms76. The H-bond length between acetylene and water is only
slightly longer than in the N CH...OH2 complex. On the other hand, the rotational spec-
tra of complexes pairing O-bases oxirane and formaldehyde with HCCH and HCN indicate
a much stronger contrast between the proton-donating abilities of the latter two molecules.
The proton of HCN approaches within 2.21 A of the oxygen of formaldehyde, 1.99 A for
oxirane. These distances are considerably longer when HCN is replaced by HCCH; 2.48
and 2.40 A, respectively77. The dubious nature of a H-bond in the acetylene complexes is
further underscored by a 30-40° CH ... O nonlinearity and nearly perpendicular direction of
approach toward the oxgyen atom. Not surprisingly, when paired with a strong proton donor
like HC1, HC CH acts as proton acceptor, via its density78.
Unlike the situation in HCN, HC CH does not contain an atom with an available lone
electron pair which can act as the proton acceptor in a H-bond. The next best pool of elec-
trons is the cloud between the C atoms in the triple bond. While this is certainly a rich
source of electron density, it is natural to wonder if a bonding set of electrons can serve the
same purpose as a nonbonding pair. For this reason, it is of fundamental interest to enquire
as to whether acetylene molecules can form H-bonds with one another.
One would expect the sort of bonding described above to lead to a T-shaped complex,
as indicated in Fig. 6.5a. It is also conceivable that this same attraction of the H atom to the
cloud could be represented by a sort of "cyclic" structure in Fig. 6.5b wherein there are
two such interactions possible. This arrangement, a kind of staggered parallel geometry, has
also been referred to as "S-shaped." There has been some controversy in the literature as to
which is actually observed in the gas phase, or whether both coexist79-87.
The theoretical literature offers a chronological picture that illustrates the pitfalls en-
countered when applying low level or incomplete theoretical treatments to weak inlerac-
300 Hydrogen Bonding

Figure 6.5 (a) T and (b) S-shaped complexes of HCCH dimer.

tions. For example, Sakai et al.80 applied an empirical intermolecular potential, based upon
dispersion and exchange, through a 6-12 function, and the electrostatic interaction between
molecular quadrupole moments. Their calculations found the staggered parallel geometry
of Fig. 6.5b to be the most stable, with an interaction energy nearly double that of the T-
shape. Ab initio calculations which followed soon thereafter88 contradicted this conclusion
when the 6-31G calculations predicted a more stable T-geometry. Alberts et al.89 later ap-
plied an enhanced (DZP) basis set, and included MP2 treatment of correlation. Another ma-
jor improvement was their ability to identify stationary points on the surface and charac-
terize them as true minima or saddle points. They found both the T and S shapes to be
stationary points on the SCF surface. While T does indeed represent a minimum, the S is a
transition state; its single imaginary frequency corresponds to displacement toward the T.
The latter is stable by 0.9 kcal/mol, with respect to dissociation to two monomers, includ-
ing zero-point vibrations and a counterpoise correction. Reoptimization of the T structure
at the correlated level leads to a contraction of the intermolecular separation by 0.4 A and
a binding energy of 1.6 kcal/mol.
In fact, the T structure would appear to be that which occurs in the gas phase 84-86 , and
the S conformers lie along the path for interconversion of one T to the next. MP2 calcula-
tions confirm the earlier findings of Alberts et al. that the T is more stable than S90. The
calculated distance from the bridging hydrogen to the C^C midpoint of 2.677 A is only
slightly shorter than the experimental estimate of 2.743 A85. After computing harmonic fre-
quencies of the dimer, Bone et al. estimate a barrier of 20 cm-1 for interconversion of var-
ious T conformers.
The results cast doubts as to whether a true H-bond exists between the acetylene mole-
cules. The T-shape is precisely what one would expect based solely upon electrostatic con-
siderations91,92. The symmetry of HCCH yields a zero dipole moment, so the moment of
lowest order is a quadrupole. A T-arrangement would best allow the approach of the two
quadrupole moments. The interaction, probably less than 2 kcal/mol, is less than normally
expected for a H-bond. Finally, the ease of rotation of one molecule around the other con-
trasts with the directionality of most H-bonds.
As it appears that the pool of electrons in the triple bond is not adequate to accept a pro-
ton from acetylene, perhaps a more conventional acceptor might be successful. Various N
and O acceptors were paired with an acetylene derivative in an argon matrix and monitored
Weak Interactions, Ionic H-Bonds, and Ion Pairs 301

by IR spectroscopy93. Table 6.3 lists the various bases in increasing order of basicity, as
measured by gas-phase proton affinity, in the first row. Starting with the weakest base,
CH3CN in the first column, all of the shifts are less than 100 cm - 1 , so none of these would
probably be classified as a H-bond. On the other end of the spectrum is the very basic
(CH3)3N which induces red shifts of at least 100 cm - 1 , approaching 300 c m - 1 in some
cases. It would hence be easy to argue that the alkynic H is being donated to this base and
that a legitimate H-bond has formed. There are gradations in between these two extremes.
Depending on what minimum threshold red shift one wishes to establish for the presence
of a H-bond, one can argue that there are or are not H-bonds present for the intermediate
situations. In any case, it seems clear that the alkynic C—H group is indeed capable of form-
ing a H-bond under certain circumstances.
Perhaps another litmus test of the ability of the alkynic C—H group to donate a proton
in a H-bond arises when a molecule of this type is paired with a hydrogen halide, HX. One
then has two distinct possibilities. The X atom, although a weak proton acceptor by nature,
can form a complex of the C—H . . . XH type. An alternative would have the XH acting as
the proton donor, with the electron-rich alkyne triple bond acting as the acceptor. Experi-
mental measurements94,95 indicate the latter is the more stable of the two alternatives. In-
deed, a similar sort of geometry is adopted when HF approaches the system of ethylene96,
even though the electron source in this double bond is less rich than in the triple bond of an
alkyne.
An experimental study of diacetylene and HF in solid argon97 suggested both sorts of
complexes (a and b in Fig. 6.6) were present and that they are of comparable stability. Cor-
related (MP2) calculations with a 6-31 + +G(d,p) basis set98 found the perpendicular com-
plex (a in Fig. 6.6), wherein FH approaches one of the two triple bonds of diacetylene, is
more stable than is complex b wherein C—H acts as proton donor. The electronic contri-
butions to the binding energies of complexes a and b are calculated to be —3.8 and —2.6
kcal/rnol, respectively. However, these values are surely inflated by the failure to correct
them for BSSE. One can conclude that the triple bond is a better proton acceptor than the
alkynic C—H is a donor, at least when paired with the rather strong acid HF. The prefer-
ence for this sort of geometry is confirmed by gas-phase measurements, and are valid also
when HF is replaced by HC199. The importance of using a satisfactory level of theory for
such complexes is reinforced by comparison with earlier SCF-level calculations100 which
predicted a structure like b to be most stable.

Table 6.3 Experimentally measured red shifts (cm - 1 ) of H—C band of


substituted alkynes (R—C=CH), when paired with various O and N bases in
Ar matrix93.
R CH3CN (CH3)2O (CH3)2CO NH, (CH3)3N

proton affinitya 188 192 197 204 225


-Cl 69 103 95 138
-CH3 49 69 64 94 156
-H 57 73 54 115 160
-CH2C1 63 111 103 125 200
-COCH3 82 131 121 146 241
-CF3 82 125 118 170 288
a
Kcal/mol.
302 Hydrogen Bonding

Figure 6.6 Two complexes pairing diacetylene with HF.

There was some discussion in an earlier chapter of the propensity of the C—H group of
molecules in which the C atom is involved in a double bond (e.g., HCOOH or HCONH2)
to act as a proton donor in a H-bond. While there was certainly evidence presented of a sta-
bilizing interaction between this CH group and the electronegative O atom of the partner
molecule, it was not entirely clear whether this interaction truly constitutes a H-bond.

6.2.2 Alkanes
We turn our attention now to C—H groups in which the C atom participates in single bonds
only. While the H atom of CH4 is clearly not capable of forming a H-bond under most cir-
cumstances, it can be made more acidic by replacing some of the H atoms by more elec-
tronegative substituents. An early theoretical study in the mid-1970s101 predicted that CHF3
would donate a proton to NH3, for example. The presence of such a H-bond in this com-
plex was later confirmed by experimental work in the gas phase102, where its strength was
found comparable to that in HCCH...NH3. Water was taken as a potential proton acceptor
and paired up with both CF3H and CC13H, and studied at the SCF level with a 4-31G basis
set103. These complexes were found to be bound by a H-bond nearly as strong as that in the
water dimer. Similarly, these two proton donors form H-bonds with formamide of strength
comparable to that of the formamide dimer104.
This line of inquiry was later extended to the series CHmCln, m+n = 4105. The calcu-
lations were performed with the 4-31G basis set and at the SCF level and, therefore, are
crude by modern standards. On the other hand, the data were improved by removing the
BSSE and adding in the effects of dispersion by an atom-atom empirical expression. The
data listed in Table 6.4 indicate that, as expected, there is only a very weak interaction be-
tween CH4 and OH2. However, the replacement of one of the methane H atoms by chlorine
immediately imparts sufficient acidity to the other H atoms that a reasonably strong H-bond
can be formed with water. The C...O distance drops precipitously, another indication of a
much strengthened interaction. (The nonlinearity of the H-bond is likely due to the elec-
trostatic interaction between the dipoles of CC1H3 and OH2.) As additional H atoms are re-
placed by Cl, the interaction undergoes additional strengthening and the H-bond becomes
progressively shorter. Of course, there is no H-bond when all four H atoms are replaced.
The qualitative aspects of the above study with the 4-31G basis have recently been con-
firmed at a higher level and including correlation106. The interaction of a series of fluoro-
substituted methanes with water was computed at the MP2/6-31G** level. The results are
Weak Interactions, Ionic H-Bonds, and Ion Pairs 303

Table 6.4 Computed properties of complexes pairing H2O with CHmCln. Data
computed with 4-31G basis set, with counterpoise correction, and dispersion
term added105.

- Eelec (kcal/mol) R(C ... O) (A) a (degs)a

CH4..OH2 0.7 3.79 0.0


CClH3..OH2 5.6 3.22 29.7
CC12H2..OH2 6.4 3.16 2.6
CC13H..OH2 8.1 3.07 0.0
CC14..OH2 1.1 4.88 0.0
a
Nonlinearity of H-bond, (O..CH).

exhibited in Table 6.5, from which it may be seen that the interaction energy is 1.4 kcal/mol
for CF3H. This value is considerably smaller than the 5.6 kcal/mol computed in the earlier
work for CC13H, a discrepancy due in part to the difference between F and Cl, but proba-
bly due more to the difference in theoretical approach. Comparison of the data in Tables 6.4
and 6.5 suggests that the earlier results for the chlorosubstituted methanes likely exagger-
ated the binding. But there are nonetheless clear similarities apparent. In either case, the re-
placement of each H by a halogen atom adds an increment to the binding energy and short-
ens the intermolecular distance. (It should be noted that the latter distances refer to H...O
contacts in Table 6.5.) These changes are approximately 1 kcal/mol and 0.1 A, respectively,
for the fluoromethanes. The interaction energy of monofluoromethane with water lies on
the lower end of the energy spectrum of what are usually considered H-bonds, but the di-
and trifluoromethanes are more strongly bound.
When the various chloromethanes are paired with HF rather than water, there is an ob-
vious tendency for the strong HF acid to act as the proton donor. For this reason, the min-
ima identified at the correlated MP2 level with a 6-31 +G(d,p) basis set contained F—H ... C1
H-bonds for the most part107. The mono-, di-, and trichloromethanes did form "cyclic" com-
plexes which contain elements of a strongly bent C—H . . . F H-bond along with a distorted
F—H . . . Cl bond. But it is difficult to separate the properties of the former from those of the
latter, so one cannot ascertain whether there is a true C—H . . . F H-bond present in these struc-
tures. An analogous study replaced HF by the weaker proton donor HC1180. The results were
similar in that cyclic geometries were obtained for the mono-, di-, and trichloromethanes.
Again, in no case was an unambiguous C—H ... Cl H—bond identified in any of the minima
present in the surface.

Table 6.5 Computed properties of complexes pairing H2O with


CHmFn. Data computed at MP2/6-31G** level, with counterpoise
correction106.

- Eelec (kcal/mol) R(H ... O) (A)

CFH3..OH2 1.41 2.51


CF2H2..OH2 2.13 2.39
CF3H..OH, 3.16 2.28
304 Hydrogen Bonding

Electron-withdrawing substituents other than halogens are also capable of making


methane a stronger proton donor. When paired with ammonia, nitromethane forms a com-
plex which contains what can be described as a C—H . . . N hydrogen bond, in addition to
auxiliary interactions between the O atoms and the hydrogens of ammonia109. The elec-
tronic contribution to the binding energy of this complex is computed at the MP2 level with
a large polarized basis set to be —4.4 kcal/mol, after correction for BSSE. After addition of
vibrational and other corrections, H° is —3.0 kcal/mol.
The absence of a true H-bond between unsubstituted CH4 and OH2 was confirmed by
higher level calculations with much larger basis sets, and with correlation included explic-
itly 110-112 . In the optimum geometry of this complex111, one of the methane hydrogens is
pointing directly toward the O atom of water as in Fig. 6.7a, consistent with the earlier 4-
31G results105. The binding energy is only 0.5 kcal/mol, remarkably similar to the cruder
earlier value, as is the C... distance of 3.75 A. It is interesting that when OH2 is replaced
by SH2, the type of complex reverses and it is now SH2 which is the nominal proton donor,
as illustrated in Fig. 6.7b. Nonetheless, there is no true H-bond present as the interaction
energy here is still only 0.5 kcal/mol.
Further verification of the absence of a H-bond came from a basis set near the Hartree-
Fock limit113. The geometry of CH4 + OH2 illustrated in Fig. 6.7a was found there to be
the only minimum on the PES but its binding energy was computed to be only 0.6 kcal/mol,
a value which is largely confirmed by later calculations as well 112,114 . Only 34% of this in-
teraction energy is present at the SCF level, another indication that it is not a H-bond. In
partial contrast, a later careful study of a wider swath of the PES112 suggested that while
Fig. 6.7a does indeed represent a local minimum on the surface, the global minimum of the
methane-water complex is in fact akin to that in Fig. 6.7b. That is, HOH is the nominal pro-
ton donor rather than CH4. This structure was computed to be more stable than Fig. 6.7a by
some 0.2 kcal/mol. The total binding energy of this particular complex, with C as proton
acceptor, is 0.8 kcal/mol.
Just as changing the O of water to its second-row analog, S, does not lead to a H-bond,
the same is true when the C of CH4 is replaced by Si. The optimized geometry in this situa-
tion is illustrated in Fig. 6.7c115. Van Mourik and van Duijneveldt argue114 that the relatively
short C...O distances sometimes encountered in organic crystals are a result of the "softness"
of the carbon atom, as opposed to any H-bonding character in the interaction itself.
Amplifying on evidence that C—H can act as a proton donor in certain circumstances,
a recent work116 reports spectroscopic evidence that this group can donate a proton to a sec-
ond-row atom. Unusually short distances were found in the C—H . . . Se interaction in the
crystal structure of diselenocin; the proton was located 2.92 A from the Se center. IR data
indicate a 53 cm-1 shift in the C—H stretch. Arguing against this interpretation is the large
deviation from linearity, closer to a right angle.
When combined with water, methane may act as a proton donor, even if only in a geo-
metric sense. Suppose that OH2 is replaced by a molecule which is a superior proton donor:

Figure 6.7 Optimala geometries of complexes of CH4 and SiH4 with OH2 and SH2.
Weak Interactions, Ionic H-Bonds, and Ion Pairs 305

will the geometry reverse such that methane acts as proton acceptor? Rotational spec-
troscopy furnishes an answer to this question. When combined with HC1 or HCN, one does
indeed see a geometry indicative of proton donation to the C atom of CH4117,118, as illus-
trated in Fig. 6.8a. However, it is questionable if this structure is indicative of a true H-bond
any more than the complex of CH4 with SH2 in Fig. 6.7b, particularly because the distance
between C and Cl is nearly 4 A. Calculations with a 6-31G** basis set at the MP2 level119
indicate the H of HC1 approaches the C of CH4 along a C3 symmetry axis. The red shift of
the HC1 bond is computed to be 16 c m - 1 .
When HF replaces HC1, the structure loses all semblance of a H-bond: a C—H of CH4
points toward the F atom119. The structure is as shown in Fig. 6.8b. This geometry has cer-
tain features of a cyclic structure with two weak H-bonds. However, calculations carried
out of CH4 + HF at the SCF/6-31G** level117 revealed the total binding energy to be only
1.0 kcal/mol, 2.1 after correlation is included via MP2. Moreover, these results are likely to
be reduced substantially were BSSE removed. The absence of a true H-bond in this com-
plex is confirmed by the long distance between C and F of 3.60 A. Spectral data of the com-
plex between HF and CH4 in an inert matrix120 appear to confirm the absence of an H-bond
geometry. That is to say, the structure likely resembles Fig. 6.8b. A structure similar to that
in Fig. 6.8 is obtained when ethane is combined with HCN121. Again, the structure has cer-
tain angular features of a H-bond but R(C .. C) is very long, estimated to be 3.76 A.
Infrared matrix isolation studies support the possibility that CH4 can act as a proton ac-
ceptor when complexed with HNO3122. The presence of a H-bond is questionable, how-
ever, since the red shift of the bridging OH stretch is only 43 c m - 1 .

6.2.2.1Anionic Acceptors
Whereas the hydrogen atoms of methane are certainly unable to form H-bonds in general,
the situation changes when the proton acceptor is an anion. One then has a greatly ampli-
fied electrostatic component, in addition to the ability of the ion to polarize the neutral mol-
ecule, increasing the polarization energy as well. Methane was paired up with the series of
halides and the geometries optimized at the MP2 level using a 6-31 + +G(d,p) basis set; en-
ergies were then obtained using larger sets approaching the Hartree-Fock limit123. The types
of structures examined are illustrated in Fig. 6.9. In the lexicon of H-bonds, the first might
be denoted linear, the second bifurcated, and a trifurcated geometry is represented in
Fig. 6.9c.
The interaction energies reported in Table 6.6 make it immediately evident that the lin-
ear structure is most stable for each anion. Indeed, geometries b and c are not true minima
on the PES. The binding energies of the linear structure diminish as the anion becomes
larger, varying from 5.8 kcal/mol for F~ down to less than 2 for the larger anions. Using

Figure 6.8 Experimental geometries of complexes of CH4 with HCI


and HF.
306 Hydrogen Bonding

Figure 6.2 Possible geometries of CH4 I

energy as a criterion, one might consider methane to be capable of forming a H-bond to F


and probably also to Cl- as well. Applying a basis set approaching the Hartree-Fock limit,
and removing counterpoise errors, the binding energy of the linear structure is computed to
be 5.6 kcal/mol for CH4 + F-, and 2.5 kcal/mol when F- is replaced by C l - , quite close
to the data reported in Table 6.6.
What about other signs of a H-bond? The R(C..F) distance in the first complex is 3.05
A, rather long for a H-bond but not entirely unreasonable. The C—H bond involving the
bridging hydrogen is stretched relative to the other bonds by 0.011 A, another sign of a true
H-bond. These same properties would argue against a H-bond for C l - , however. R(C..C1)
is 3.79 A, and the bridging C—H bond is stretched by only 0.001 A. The harmonic fre-
quency corresponding to the H-bond stretch v is computed to be 159 cm-1 for CH4 + F-,
but less than 70 cm-1 for the other complexes. The C—H stretching frequency of CH4 un-
dergoes a red shift of 105 c m - 1 in CH4 + F- but 20 c m - 1 or less for the other anions.
In summary, one could argue that methane forms a H-bond with the fluoride anion. The
geometries are similar for the other anions, but the interaction is considerably weaker as are
the other indicators of a H-bond.

6.2.3 Metal Atoms as Acceptors


Another sort of unconventional type of H-bond that has been proposed in the literature in-
volves electron-rich transition metals as proton acceptor124-128. These ideas are derived
from spectroscopic data but the most explicit information arises from diffraction studies of
crystals, where the evidence is largely geometrical. For example, an NH group approaches
unexpectedly close to a Pt atom in a square-planar geometry, with R(Pt .. H) = 2.262 A129.
The arrangement is nearly linear as the 9 (N—H ... Pt) angle is 167°. The authors point to
prior crystal structures in the literature where N—H and even C—H groups are positioned

Table 6.6 Computed binding energies (— Eelec in kcal/mol) of


complexes pairing CH4 with halide anions. Data computed at
MP2 level with 6-31 + +G(d,p) basis set, uncorrected for
BSSE123.

Linear Bifurcated Trifurcated

F- 5.8 2.7 1.8


C1- 2.7 1.3 1.2
Br- 2.0 0.7 0.5
I- 1.3 0.2 0.3
Weak Interactions, Ionic H-Bonds, and Ion Pairs 307

directly above the plane of a square complex with the metal in a d8 configuration. Mention
is made also of possible interactions with d10 metals such as Co and Ni130,131. For exam-
ple, theR(Co .. H) distance is 2.613 A, less than the sum of the van der Waals radii, and the
(N—H ... Co) arrangement is very close to linear132. Moreover, the N—H distance is mea-
sured to be surprisingly long, 1.054 A, an indication of possible H-bonding activity. How-
ever, as the authors point out, there is a fine line between a true H-bond and a simple elec-
trostatic interaction between the cation and anion.

6.2.4 Hydride as Proton Acceptor


A different sort of interaction involving hydrogens has been recently noted in crystal stud-
ies of certain systems. It has been observed by X-ray diffraction and NMR techniques that
two hydrogens can approach one another unexpectedly closely if one of them is covalently
bonded to an electronegative atom such as N, and the other to a metal atom like Ir. It is
supposed that the Ir-H hydride acts as a sort of "base" to accept the proton from N—H.
An example of this type arises in the iridium complex with trans PCy3 groups and a cis
1
-SC5H4NH ligand133 where the two hydrogens in question are separated by 1.75 A. NMR
data indicate this interhydrogen distance persists in solution as well. A comparably short
Ir-H .. HO interaction has been observed when the OH is part of an iminol group134; the in-
terhydrogen distance here may be as short as 1.58 A. The two hydrogens are separated by
some 1.77 A when an alcohol is paired with a W—H hydride in a tungsten derivative in an
intermolecular complex135.
Estimates of the strength of this type of interaction, based upon experimental quantities,
vary from 2.5 to as much as 6.9 kcal/mol in certain cases135-137. Geometry optimizations
carried out at the SCF level, using a moderate-sized basis set including polarization func-
tions136, found the Ir-H and N—H hydrogens could approach within 1.96 A in a model sys-
tem. The authors attributed the source of the attraction to the opposite charges calculated
on the two hydrogens, +0.22 on one and —0.26 on the other. Another contributing factor
is the ease of polarization of the Ir-H bond. Although a positive bond order was computed
between the two hydrogens in question, the small value of only 0.012 might lead one to
question whether the interaction contains much covalent character. An analysis of various
energetic quantities yielded a theoretical estimate of the interhydrogen interaction energy
of 5.65 kcal/mol.
The fundamental nature of this interaction as primarily electrostatic attraction between
the opposite charges of the two hydrogens was supported by later calculations138. When the
geometry of the complex of FH with HMn(CO)5 was optimized at the SCF/3-21 G(*) level,
the H atoms of the two subunits were separated by 1.68 A. This distance is likely too short
due to the large BSSE common to this sort of basis set. Likewise, the computed interaction
energy of 6.55 kcal/mol is probably exaggerated by the same phenomenon. Nonetheless,
the results do support the notion of a short interhydrogen contact when one H is covalently
bonded to an electronegative atom and the other to a metal.
Similar sorts of interactions have recently been proposed as well when the metal is re-
placed by a boron atom. Crystallographic data yields a number of NH ... HB interhydrogen
separations of less than 2.2 A139. The nitrogenic hydrogen seems to prefer to approach the
other hydrogen from a direction nearly perpendicular to the BH bond. The authors attrib-
uted this significantly nonlinear geometry to the unfavorable alignment of bond dipoles that
would result in the case of a linear N — H . . . H — B arrangement.
308 Hydrogen Bonding

6.3 Symmetric Ionic Hydrogen Bonds

On one end of the spectrum of H-bonds are those that are very weak, as in the case of nu-
merous C—H . . . X interactions. On the other end of this continuum are very strong interac-
tions. HX molecules are quite acidic. If paired with a proton acceptor that has a negative
charge, the interaction can become very strong indeed. In fact, the H-bond can become so
short that the proton loses its association with one molecule versus the other. That is, the
proton can be drawn into the midpoint of the X...X axis. Such a H-bond is referred to as
centrosymmetric.

6.3.1 Hydrogen Bihalides


Simple examples are the hydrogen bihalide anions FH..F- and C1H .. C1 - . Polarized basis
sets were used to examine this pair of systems, along with CI treatment of correlation140.
At the SCF level, their binding energies were computed to be 39 and 15 kcal/mol, re-
spectively. These values increase to 41 and 21 kcal/mol at a level which approximates a
full CI treatment. The computed results mimic experimental data rather well. The bind-
ing energy in FH..F- had been measured to be 42 kcal/mol 141,142 ; the range obtained for
C1H..C1- is 13-19 kcal/mol143. In both cases, the optimized position of the proton is mid-
way between the two halogen centers. The interfluorine distance is calculated to be 2.26
A, one of the shortest contacts known (2.278 A in the gas phase144); R(Cl .. Cl) is 3.13 A,
also very short.
The hydrogen bifluoride anion was considered at higher levels of theory, focusing upon
its infrared spectrum145. CISDT was applied, in conjunction with a TZ3P basis set, aug-
mented by very diffuse Rydberg(R)-type functions. Anharmonic effects were considered by
evaluating the full quartic force field of this triatomic. At all levels, SCF through CISDT,
and DZP through TZ3P+R+d, the H-bond was found to be centrosymmetric, and the lin-
ear (FHF)- anion belongs to the D h point group. Interfluorine distances lie in the narrow
range of 2.24-2.28 A. The harmonic frequencies of the symmetric (H-bond stretch) and an-
tisymmetric (proton motion) stretches are computed to be 668 and 1195 cm - 1 , respectively.
Because of the large dipole moment change associated with the proton motion, the inten-
sity of the latter mode is very high indeed, 3724 km/mol. Anharmonicity lowers the first
frequency by 36 cm - 1 , but the asymmetric stretching mode is much more sensitive, in-
creasing by 265 cm^'. With regard to sensitivity to level of quantum mechanical treatment,
the symmetric stretch does not change much as the basis set is altered, nor as correlation is
added, remaining in the 650-700 cm-1 range at all levels of theory. The antisymmetric
stretch, on the other hand, is highly sensitive, with frequencies varying between 627 and
1538 cm-1 depending upon level of theory. One may conclude that computation of the vi-
brational characteristics of the mode involving the proton transfer between the two F cen-
ters is very demanding, whereas the other modes of the complex may be treated with only
reasonable levels of theory. The sensitivity of the antisymmetric stretch to particular theo-
retical treatment is paralleled by a similar sensitivity to experimental conditions: v3 has been
measured to lie anywhere between 1284 and 1740 cm - 1 , depending upon the particular
host146; the frequency in the gas phase is 1331 cm - 1 . 1 4 4
Later work evaluated the two-dimensional potential energy surface using various corre-
lation treatments including many-body perturbation theory and coupled cluster tech-
niques 147 . Evaluation of the vibrational spectrum was explicitly anharmonic in nature, mak-
Weak Interactions, Ionic H-Bonds, and Ion Pairs 309

ing use of a highly flexible analytic function to fit the calculated point-wise surface. The re-
sults agreed with earlier work in that the symmetric stretch is relatively insensitive to the
variety of correlation treatment, in the range of 570-600 c m - 1 for all methods examined.
The asymmetric stretch, too, is relatively stable with respect to the order of MP or CCSD
treatment, lying in the range between 1380 and 1450 c m - 1 . It was mentioned that the use
of a standard quartic potential leads to poor reproduction of experimental frequencies. An-
other treatment of the vibrational modes in FH..F- computed the energetics of 710 points
at the CID level with a [3s2pld/2slp] basis set148. These results were then fit to an analyt-
ical function (a superposition of Morse potentials plus other terms), which provided clear
evidence of the nonharmonic nature of the potential surface, and illustrated the transition
from a centrosymmetric single minimum to a pair of equivalent minima as the interfluorine
distance was increased. This transition occurs at approximately R(F..F) = 2.4 A on this sur-
face. The authors' results agreed with the other study in that the surface is more complex
than that presuming a quartic force field.
Ahigh level correlated study of the FH..F- complex149 yielded a MP4/6-311 +G(2d,2p)
interaction energy — Eelec of 44.3 kcal/mol, although this value was not corrected for su-
perposition error; a computation with an even more extended basis set resulted in a H-bond
energy of 45.6 kcal/mol150. Comparable calculations, including counterpoise correc-
tions 151 , reduced the binding energy to 39 kcal/mol, in good agreement with an experi-
mental measurement of 39152.
The (Cl—H—Cl) - analog represents a particularly interesting system as crystal struc-
tures are inconclusive regarding the nature of the proton transfer potential. The C1..C1 dis-
tance is observed to vary between 3.14 and 3.22 A in various different crystalline environ-
ments153. When in its shorter range, a centrosymmetric H-bond is observed whereas the
proton is displaced as much as 0.24 A from the C1..C1 midpoint for longer distances. The
interchlorine separation is measured to be 3.15 A in the gas phase154. The data suggest a
double-minimum potential for this anion, but the lowest vibrational level is close enough
to the top of the barrier that a centrosymmetric H-bond is observed154,155. An early SCF
study treated the (Cl—H—Cl) - system with a DZP basis set156. R(Cl..Cl) was optimized
to be 3.24 A. Consistent with experimental data for longer interchlorine distances, the pro-
ton transfer potential contained two separate minima in each of which the proton is offset
from the bond midpoint by 0.23 A. The barrier separating the two wells is rather low, only
0.6 kcal/mol. The H-bond energy of the (Cl—H—Cl) - complex was computed to be 20
kcal/mol, in reasonable agreement with an experimental measurement of 24 kcal/mol157.
When correlation is added, the barrier appears to vanish and the H-bond takes on cen-
trosymmetric character151,158-161. An extensive basis set of [11s9p2dlf] character for Cl,
with H represented by [5s2pld], was applied to this system, and correlation applied via
MP2-4159. The results in Table 6.7 illustrate the change in character from C v to D h that
occurs upon application of correlation. That is, the proton moves to the center of the C1..C1
bond. Along with this change in character comes a shortening of the H-bond by some 0.2
A. As in the earlier case of (FHF) - , the symmetric stretching frequency is rather insensi-
tive to the level of theory while the asymmetric stretch shows a good deal of variation. This
result is consistent with the flatness of the potential for proton transfer, and the change in
character upon adding correlation. After adding in zero-point vibrational corrections, the
enthalpy of association in the (Cl—H—Cl) - complex was computed to be —23.5 kcal/mol
at the MP4 level, in nice agreement with experimental estimates of —23.1 and
— 23.7 152,162,163 . This interaction is quite a bit weaker than in the FH..F- analogue. The
data emphasize the sensitivity of the nature of the H-bond in this complex to level of the-
3 10 Hydrogen Bonding

Table 6.7 Calculated characteristics of optimized (Cl—H—Cl) - complex159.

Basis set Level Symmetry R(Cl .. Cl) (A) ra (A)

DZP SCF C V 3.343 0.32


[Ils9p2dlf/5s2pld] SCF C v 3.331 0.32
DZP MP2 D h 3.126 0.0
DZP MP4 D h 3.126 0.0
a
Displace of proton from C1..C1 midpoint.

ory, or to any environmental effects that might tend to shorten or lengthen the H-bond. It is
also important to note that due to the change from a noncentrosymmetric to centrosym-
metric H-bond that occurs upon inclusion of electron correlation, the SCF and correlated
vibrational structures are quite different160, so one would be ill-advised to add SCF zero-
point vibrational corrections to the correlated energetics.
Correlated computations have also been performed on the next in the series,
(Br .. H .. Br )-164 . Using a DZP basis set (more specifically [9s7p2d/3slp]) and a multirefer-
ence CI approach, this anion was also found to be centrosymmetric, with R(B . . Br) = 3.429
A. This distance is estimated to be about 0.27 A shorter than the sum of the van der Waals
radii. The dissociation energy to HBr and Br- is 15.4 kcal/mol, bracketed by experimental
estimates of the same quantity152,165. Note that the computed value was not corrected for
superposition error. The zero-point vibrational correction to the dissociation energy was cal-
culated to be negligible. Electron correlation is absolutely critical to proper treatment of this
system.

6.3.2 Comparison with Other Anionic H-bonds


A systematic comparison of various symmetric anionic complexes provides a solid basis
for comparison151. The data listed in Table 6.8 illustrate the rapid decline in binding energy
as the electronegativity of the atoms diminishes, or as one passes from first to second-row
atoms. Concomitant with this weakening of the interaction is the lengthening of the H-bond.
In a number of cases, there is a fine distinction as to whether the H-bond is centrosymmet-

Table 6.8 Binding energies, H-bond lengths, and proton displace-


ments of anions, computed at MP4/6-311+G(d,p) level151.
Complex — Eeleca (kcal/mol) R(A) rb (A)

(FH..F)- 38.8 2.30 0.0


(HOH..OH)- 23.2 2.44 0.oc
(H2NH..NH2)- 10.2 2.91 0.40
(ClH..C1)- 18.0 3.10 0.0C
(HSH..SH)- 7.1 3.28 0.0c
(H2PH..PH2)- unbound
a
BSSF and ZPE corrections added.
b
Displacement of proton from H-bond midpoint.
c
Barrier lies below first vibrational level.
Weak Interactions, Ionic H-Bonds, and Ion Pairs 3II

RIc or not. It is not uncommon to find a proton transfer potential that contains two wells, but
the barrier is quite low.

6.3.2.1 (HOH..OH)-
The (HOH..OH)- anion serves as a classic case of the latter situation. For example, calcu-
lations with basis sets ranging from 4-31G to polarized such as 6-311G* and larger166-170
showed how correlation and/or zero-point vibrations can change a double to a single-well
potential, and how the results might differ depending upon the particular means of treating
the correlation. But all surfaces were flat, regardless of whether there were two minima pre-
sent or one. The highest barriers obtained were less than 1 kcal/mol, so one can conclude
that this barrier would lie below the first vibrational level. Consequently, the complex would
behave for all intents and purposes as though it contained a centrosymmetric bond. Of
course, the amplitude for proton motion would be quite large. In cases such as these, it is
probably more appropriate to compute the binding energy of the centrosymmetric geome-
try, even though this configuration represents a first-order saddle point on the purely elec-
tronic potential energy surface.
The fine balance between a centrosymmetric and noncentrosymmetric H-bond is con-
firmed by crystal studies. Lithium hydrogen phthalate monohydrate, for example, contains
H-bonds of the OH..O type about 2.4 A in length 171 . While one of these has the bridging
hydrogen closer to one O atom than to the other, the other H-bond is centrosymmetric (or
very nearly so). Another centrosymmetric H-bond is observed in the crystal structure of
methyl ammonium hydrogen succinate monohydrate172, this one being 2.44 A long. (The
authors do not rule out the possibility of a double-well potential for proton transfer, albeit
one with a low barrier.) Perhaps the shortest such H-bond observed occurs in the crystalline
complex of l,8-bis(dimethylamino)naphthalene with 1,2-dichloromaleic acid173; the 2.38
A H-bond is apparently centrosymmetric. The authors provide a graphic of measured r(OH)
bond lengths in similar complexes that illustrates the transition from noncentrosymmetric
to centrosymmetric as R(O..O) contracts down to 2.4 A.
Minor external effects can also influence the shape of the proton transfer potential be-
tween carboxylate anions. Whereas there are apparently two minima present in complexes
of this type in aqueous solution, the H-bond takes on centrosymmetric character when im-
mersed in nonpolar solvent174. Calculations have suggested the change from single to dou-
ble-well character is due not to the surrounding dielectric but rather to interactions with dis-
crete molecules175. The binding energy computed for (HOH..OH)- at a relatively high
level, 23 kcal/mol151, is in reasonably good coincidence with experimental measurements
of the enthalpy of binding of 27 kcal/mol176,177. Computed enthalpies, which include rota-
tional and translational corrections160,178 were even closer at 28.6 kcal/mol, although nei-
ther study removed BSSE.
Recent NMR measurements of a series of complexes in the solid state179 indicate that
many of the same principles apply to H-bonds of the FH..F type. Centrosymmetric H-bonds
occur only for short interfluorine distances, about 2.3 A or less. The gas-phase bond length
of 2.2,78 A is therefore consistent with the ab initio finding of a centrosymmetric H-borid
in (F...H...F)-. The crystalline environment can alter the proton transfer potential to a non-
centrosymmetric type by just a small H-bond elongation179.
Methyl substitution apparently induces little change in this behavior. MP2 calcu-
lations with a 6-31+G* basis set180 yield a binding enthalpy of 26.3 kcal/mol for
312 Hydrogen Bonding

(CH 3 OH ... OCH 3 )- . The small barrier present in the proton transfer potential at the SCF
level (2.2 kcal/mol) is eliminated by the inclusion of zero-point vibrational energies and/or
by consideration of electron correlation. Experimental measurements indicate the interac-
tion energy is increased by 1-2 kcal/mol by methyl substitution. Tandem flowing afterglow-
selected ion flow tube181 and variable-temperature pulsed high-pressure mass spectromet-
ric measurements182 of (CH3OH...OCH3)- have led to a dissociation energy of 28-29
kcal/mol. This change in binding energy accompanying methyl substitution is rather small,
considering that the deprotonation energy of CH3OH is lower than that of HOH by 9.6
kcal/mol182. And like the simpler (HOH .. OH) - , the proton transfer potential in
(CH3OH...OCH3)- in the gas phase contains either a single central minimum or a small
barrier, as determined by ion-cyclotron resonance spectroscopic measurements183. Even a
more major change in the character of the oxygen bases yields a relatively small perturba-
tion in the H-bond energy. For example, replacement of the two methoxide ions in
(CH3OH...OCH3)- by a pair of acetate anions increases the enthalpy of binding by only 0.5
kcal/mol184.

6.3.2.2 Anions with Triple Bonds


The C=N - anion is of particular interest as it is capable of forming a H-bond from either
end. Table 6.9 reports the properties of the H-bonds formed between this anion and HCN
(or HNC), as well as the analogous (HC=CH .. C=CH) - which also contains a triple
bond185. The first column reflects the greater acidity of HNC as compared to HCN, as the
former forms stronger H-bonds when it acts as proton donor. The weaker acidity of HCCH
is apparent from the relatively low interaction energy in (HCCH .. CCH) - . The binding en-
thalpies at 300 K in the next column are similar in magnitude to the electronic data in the
preceding column. The stronger H-bonds are also associated with shorter intermolecular
separations. The noncentrosymmetries of the H-bonds are illustrated in the r column; all
are predicted to be noncentrosymmetric at the SCF level. The degree of asymmetry is in-
versely related to bond strength. The values of E+ listed in the next columns refer to the bar-
riers in the proton transfer potential. There is a clear correlation between strong H-bonds
on one hand and low barriers on the other. It is important to stress that inclusion of electron
con-elation lowers all barriers, to the point where the NH—N interaction collapses to a cen-
trosymmetric (CN .. H .. NC) - H-bond.

Table 6.9 Properties of anionic complexes of triply bonded species, computed with 6-31 + G*
basis set185.

a Et,d (kcal/mol)
- Eelec (kcal/mol) - Hb (kcal/mol) R (A) rc (A)
Complex MP2 MP2 SCF SCF SCF MP2

(CNH .. NC) - 27.5 26.3 2.75 0.34 3.4 0.0


(NCH..CN)- 19.5 18.5 3.20 0.51 9.5 5.3
(HCCH..CCH)- 10.8 9.7 3.35 0.60 13.1 7.6
a
Corrected for BSSE.
b
Evaluatedat 300 K.
e
Displacement of proton from H-bond midpoint.
d
Proton transfer barrier.
Weak Interactions, Ionic H-Bonds, and Ion Pairs 3 I3

6.3.2.3 Anions with Double Bonds


An equivalent study was carried out for the analogous systems containing double instead
of triple bonds186. Because the doubly-bonded C or N atoms are less acidic, the H-bond en-
ergies in Table 6.10 are smaller than those in Table 6.9. The H-bonds are also uniformly
longer and more noncentrosymmetric. Other systematic trends remain consistent. The N
atom is more acidic than C, and the C of HN=CH 2 is more acidic than H2C=CH2. The H-
bonds lengthen as the interaction weakens. The noncentrosymmetry of the proton position
grows as the barrier for proton transfer becomes higher. Correlation lowers these barriers
but not to the point where any of them vanish entirely. (However, the lowest barrier, that in
(H 2 CNH .. NCH 2 )- , does indeed vanish when zero-point vibrational energies are added.)

6.3.2.4 Carboxylate
When bound together by a proton, a pair of formate anions (HCOO - ) can in principle adopt
a number of different conformations since each O atom has associated with it a syn and anti
lone electron pair. Using a fairly large basis set containing diffuse functions on O, along
with polarization functions, the geometry obtained for this complex is illustrated in Fig.
6.10a187. The proton donor is designated as anti because the OH proton is opposite the car-
bonyl oxygen; the formate is referred to as syn since the proton which it is accepting ap-
proaches a syn lone pair of that O atom. This particular arrangement also permits the CH
proton of the HCOOH to interact favorably with the oxygen of the HCOO- which is not
involved in the H-bond with —OH. The principal R(O .. O) H-bond length is 2.592 A, while
the distance separating the C—H hydrogen from the O of the acceptor group is 2.51 A. Since
this is much longer than would normally be expected for a H-bond, this secondary interac-
tion would probably be better characterized as an electrostatic interaction.
The binding energy (- Eelec) of this structure is calculated to be 28.8 kcal/mol at the SCF
level, increasing to 33.0 when MP2 correlation is added. These results are smaller than the
gas-phase measurement of H° of 36.8 kcal/mol184. The disagreement is amplified by the
fact that BSSE and ZPE corrections have not been added to obtain a theoretical enthalpy.
Like several of the systems mentioned above, the noncentrosymmetric H-bond illustrated
in Fig. 6. l0a, changes to a more symmetric structure, and the proton transfer potential con-
verts from double to single-well character when correlation effects are accounted for. At the

Table 6.10 Properties of anionic complexes of doubly bonded species, computed with
6-31+G** basis set186.
Etd (kcal/mol)
- Eelec (kcal/mol) - Hb (kcal/mol) R (A) rc (A)
Complex MP2 MP2 SCF SCF SCF MP2

(H2CNH..NCH2)- 15.4 13.8 3.03 0.50 8.6 1.7


(HNCH2..CHNH)- 10.3 9.6 3.52 0.68 18.2 11.7
(H2CCH2..CHCH2)- 5.6 4.8 3.74 0.78 19.2 12.8
a
Corrected for BSSE.
b
Evaluated at 300 K.
^Displacement of proton from H-bond midpoint.
d
Proton transfer barrier.
314 Hydrogen Bonding

Figure 6.10 Geometrical configurations of


HCOOH ...-- OOCH 187 .

approximate bottom of the correlated transfer potential, the R(O..O) distance is 2.47 A. The
authors estimate that the secondary interaction in the anti-syn configuration in Fig. 6.10a
accounts for perhaps an additional stabilization of 2 kcal/mol.
For purposes of comparison, the anti-anti and syn-syn geometries were also studied by
the authors187. The configurations pictured in Fig. 6.10b and 6.10c were optimized. The
anti-anti geometry was found to be more stable than syn-syn by about 3 kcal/mol. At the
SCF level, the anti-anti is 4 kcal/mol higher in energy than anti-syn geometry, but this dif-
ference is reduced to only 1 kcal/mol at MP2.
A later calculation disputed these results in that anti-anti was computed to be the most
stable conformer188. This work, however, was limited to the SCF level, using a 6-
31 + +G** basis set. (If consulting the original paper, it is important to note that this set of
workers reversed the standard nomenclature for syn and anti.) Anti-syn was computed to
be somewhat less stable, followed by syn-anti and syn-syn. These results were compared
to a statistical survey of such interactions in proteins where a marked propensity was ob-
served for the syn-syn structure, the least stable in the gas phase. In contrast, the anti-anti
configuration, most stable in the gas phase, is observed rarely in proteins. The authors at-
tributed these discrepancies to crystal packing forces. There also seems to be a preference
in the crystals for centrosymmetric (or nearly so) H-bonds between carboxyl and carboxy-
late, although noncentrosymmetric H-bonds are also present in large numbers. Centrosym-
Weak Interactions, Ionic H-Bonds, and Ion Pairs 315

metric H-bonds in the crystal tend toward R(O--O) distances of about 2.45 A; the noncen-
trosymmetric bonds are a little longer, in the general range of 2.5-2.6 A. In either case, the
crystal seems to exert a mild lengthening effect upon the H-bonds.

6.3.2.5 Environmental Effects


The effects of the forces present in the crystal upon the H-bond connecting a pair of for-
mate units was examined explicitly by placing positive point charges in the vicinity of the
complex, and then employing a self-consistent procedure to evaluate point charges of the
remainder of the crystal189. Rather than carry out a full geometry optimization, the posi-
tions of all atoms were taken from a neutron diffraction study of potassium hydrogen di-
formate, with the exception of the bridging hydrogen atom. The study was hence restricted
to an R(O-O) distance of 2.437 A, with a geometry corresponding roughly to anti-anti in
Fig. 6.10. Prior to inclusion of the crystal forces, the SCF/DZP proton transfer profile con-
tains a pair of minima, separated by an energy barrier of 1 kcal/mol; the MP2 H-bond is
centrosymmetric. The noncentrosymmetric nature of the crystal forces induces a displace-
ment of the position of the proton in the MP2 potential, moving the minimum 0.053 A from
the O--O midpoint, in nice agreement with the experimental shift of 0.052 A. For this par-
ticular case, then, it is apparent that asymmetric forces present within a crystal can shift the
equilibrium position of the proton away from the midpoint of the H-bond, while maintain-
ing its single-well character. Another crystal structure, that of sodium hydrogen bis(for-
mate), seems to contain all noncentrosymmetric H-bonds between formate anions190. In
aqueous solution, on the other hand, a large observed primary isotope effect suggests the
proton moves within a double-minimum potential191.
The ability of the environment to influence the nature of the proton transfer potential has
been underscored by work192 that differentiates between a centrosymmetric single-well po-
tential and one containing two wells, by isotopic perturbation of the NMR spectrum. It was
found that when certain H-bonded systems that have a single well in the crystal are dis-
solved in aqueous media, the potential changes its character to a double-well. One expla-
nation is that solvation can better stabilize a localized charge as would occur when the pro-
ton is shifted toward one subunit or another, as compared to a centrosymmetric geometry
where the charge is delocalized over the entire complex. This contention is confirmed to the
extent that systems of this type revert to a single-well proton transfer potential when placed
in less polar solvents. The disorder of water can be invoked as a second and supplementary
explanation, in that it would be unlikely for the two ends of the H-bond to be solvated to
the same degree, thereby preferentially stabilizing one side versus the other.

6.3.2.6 HSO4
Perhaps the strongest of H-bonds might be expected when H2SO4 combines with its con-
jugate base, H S O 4 , since three separate and distinct H-bonds can occur. SCF optimiza-
tions of the geometry of this complex193 with a 6-31G* basis set led to a structure of this
complex as illustrated in Fig. 6.11. Applying the MP2/6-31+G* level of theory to the op-
timized geometry, the total binding enthalpy was computed to be 47 kcal/mol. While this
result probably suffers from some inflation due to BSSE, it is certainly quite a strong in-
teraction. The potential for proton motion from the neutral to the anion contains a pair of
equivalent minima, separated by a barrier of 2 kcal/mol at the SCF level. It is likely that this
low barrier would vanish entirely were correlation applied to the transfer potential.
3 16 Hydrogen Bonding

Figure 6.11 Geometry of complex between


H2SO4 and HSO4- l93 .

6.3.3 Cationic H-bonds


In addition to complexes pairing a neutral molecule like HF with its anion, F--, one can con-
struct a cationic complex where HF is combined with the protonated H2F+. Such complexes
are also expected to be quite strong. Following earlier work at the SCF level194, a high-level
computation of this particular system195 obtained a binding energy of 31.9 kcal/mol. This
result was obtained at the fourth-order M011er-Plesset level with a polarized basis set. As il-
lustrated in Table 6.11, the interfluorine distance is quite short, only 2.29 A, similar to the
H-bond length in the anion but not quite as strong. As in the anion, the H-bond is cen-
trosymmetric, with the proton located in the center of the F-F axis.

6.3.3.1 (H 2 OH .. OH 2 ) +
The cationic analogue of (HOH..OH)-- is (H 2 OH .. OH 2 ) + , in which two water molecules
are held together by a fifth proton. The nature of the geometry in this complex, like the an-
ion, has generated a good deal of theoretical study. Early studies at the SCF level, and with
small basis sets, had indicated the complex contains a short, centrosymmetric H-
bond196-200.
This question has been probed by much higher levels of calculation recently201, 202 and
it appears that the previous lower-level work was largely correct. More precisely, the bot-
tom of the proton transfer potential appears to be only very slightly asymmetric if at all; the
barrier, if it exists, is less than 0.5 kcal/mol. As in the (HOH..OH)-- anion, the character of
the potential alternates from single to double well upon small changes in level of theory, the
former being favored by electron correlation. The calculations, employing CCSD(T) treat-

Table 6.11 Properties of cations, computed at MP4SDQ/6-31 +G(d,p) level195.


Complex - Ea (kcal/mol) R(A) rb (A)

(HFH..FH)+ 31.9 2.29 0.0


(H2OH..OH2)+ 33.0 2.48 0.0
(H3NH-NH3)+ 23.6 2.85 0.37
(HC1H-C1H)+ 12.2 3.16 0.0
(H 2 SH--SH 2 ) h 10.5 3.75 0.52
(H3PH-PH.,) ' 7.3 4.18 0.70
a
Includes ZPVE correction.
b
Displacement of proton from H-bond midpoint.
Weak Interactions, Ionic H-Bonds, and Ion Pairs 3 17

ment of correlation, in conjunction with a TZ2Pf basis set, predict R(O..O) to be some 2.40
A; the H-bond is close to linear, but not quite, with a (O--H--O) angle of about 173°201. The
binding enthalpy is calculated to be 33.4 kcal/mol, agreeing rather nicely with experimen-
tal estimates in the 32-33 kcal/mol range203"205. An MP2 treatment of correlation178, with
a 6-311 + +G(d,p) basis set yielded a binding enthalpy of 35.8 kcal/mol; this overestimate
is likely due to the failure to remove BSSE. It is notable that this particular calculation in-
dicated a centrosymmetric H-bond, as did other comparable calculations202'206, further ev-
idence of the sensitivity of this question to the particular theoretical approach. A later study
of the anharmonicity of the OH vibration involving the bridging proton202 indicates that
this effect would reduce the computed binding enthalpy by some 0.5 kcal/mol.
Other calculations have tested whether the complex pairing H3O+ with H2O contains a
single H-bond or more than one207. Whereas the single bond does appear to be preferred in
the neutral water dimer, the question is different when one of the two species is charged. It
is conceivable that the ion-dipole interaction might favor the closer approach that is possi-
ble in the double (D) or triple (T) H-bonds illustrated in Fig. 6.12 as the intervening hy-
drogen gets "out of the way" of the two oxygens. (These interactions can also be described
as bifurcated and trifurcated, respectively.) At the MP4/6-31G* level, the single H-bond is
in fact preferred by 8.0 kcal/mol over the D configuration, which is in turn favored by 4.8
kcal/mol over T. Moreover, R(O-O) elongates from 2.456 A in S to 2.527 A in D and 2.633
A in T, consistent with the progressive weakening. In fact, the D geometry is not a mini-
mum on the potential energy surface, but represents instead a transition state on the path of
interchange of one bridging hydrogen in S to another.

6.3.3.2 (H3NH..NH3)+
The nitrogen-containing cation, (H3NH--NH3)+, unlike the others, does appear to have a
legitimate double-well potential. Early studies at the SCF level208'209 indicated that R(N-N)
should be about 2.81 A, and that the potential for proton transfer contains a pair of equiva-
lent wells, separated by a barrier of 3.8 kcal/mol. The minimum in the potential has the pro-
ton displaced by 0.28 A from the N..N midpoint. A preliminary study210 was inconclusive
as to whether electron correlation would raise or lower this barrier: A generalized valence
bond approach increased the barrier while a POL—CI treatment yielded a lowering, al-
though both approaches suggested the H-bond is lengthened by correlation. M011er-Plesset
treatments yielded a barrier of less than 1 kcal/mol, suggesting it might lie below the first
vibrational level211. H-bond strengths in the 22-31 kcal/mol range were obtained from these
calculations. MP4 calculations195 confirmed that the barrier in this complex is probably in
the neighborhood of 1 kcal/mol. After zero-point vibrational energies were added to the in-
teraction energy, a value of —23.6 kcal/mol was obtained, in nice agreement with an ex-
perimental enthalpy of —25 kcal/mol212. The effect of correlation was made more explicit

Figure 6.12 Candidate geometries pairing H 3 O + with H2O, S, D, and T re-


fer to single, double, and triple H-bonds, respectively.
318 Hydrogen Bonding

in a later study213 where a contraction of R(N .. N) by 0.085 A was observed; the MP2/6-
31G* value is 2.731 A. The noncentrosymmetry of the complex, as measured by the devi-
ation of the proton from the N..N midpoint, decreased from 0.35 to 0.25 A. Applying
MP4SDQ to this geometry, along with a larger 6-311 + +G(d,p) basis set, so as to obtain vi-
brational frequencies, the binding enthalpy of the (H 3 NH .. NH 3 ) + complex was calculated
to be —24.9 kcal/mol at 298 K (no BSSE correction was applied)213. Another study com-
puted a AH of —27.1 kcal/mol with the same basis set at the MP2 level178, with this value
becoming slightly less attractive, —25.5 kcal/mol, when the level was raised to MP4/6-
311 + +G(2d,2p). As in the case of H5O2+, H 7 N 2 + also prefers a single linear H-
bond207,214. The double (or bifurcated) H-bond, analogous to D in Fig. 6.12, is higher in
energy than the single bond by 7.4 kcal/mol.

6.3.3.3 Comparison with Second-Row Analogs


Calculations have addressed the second-row analogs of the aforementioned cations. An
early study of (H2SH..SH2)+ indicated a noncentrosymmetric H-bond. At the SCF level the
intersulfur distance was computed to be 3.48 A215, and the proton is displaced 0.28 A from
the S..S midpoint. A barrier of 0.6 kcal/mol was predicted to separate this minimum from
its equivalent following proton transfer. Without BSSE correction or correlation effects, the
binding energy was estimated to be 17 kcal/mol. Correlation was later introduced into this
complex, as well as the other second-row cations216 and resulted in an improved binding
energy. The enthalpy computed at the MP4SDQ/DZP level is 12.9 kcal/mol, in the range of
experiment of 12.8 — 15.4 kcal/mol203, 217. The noncentrosymmetry of the proton position
diminishes from 0.51 to 0.36 A upon adding correlation. The binding enthalpy of the
(HQH..C1H)+ complex was computed to be about 2 kcal/mol stronger, while that of
(H3PH..PH3)+ was found weaker by 4-5 kcal/mol. In each case, correlation tends to bring
the bridging hydrogen atom closer to the H-bond midpoint.
It might be noted parenthetically, however, that the single H-bond studied for
(H3PH..PH3)+ may not in fact be the most stable. Calculations have suggested that a tri-
furcated geometry, similar to the T structure in Fig. 6.12, may be preferred to the single H-
bond by as much as 2 kcal/mol207. This result suggests that the interaction should perhaps
be considered less of a H-bond and more of a simple ion-dipole interaction. The R(P..P)
distance is only 3.64 A in the trifurcated arrangement, as compared to 4.17 A in S where
the intervening hydrogen holds the two P atoms apart.

6.3.4 Comparisons between Cations and Anions


Table 6.11 presents the results of a systematic study of the proton-bound cationic dimers,
all calculated at a fairly high level. These data may be compared with Table 6.8 to obtain
some insights about the cations and their corresponding anions. There are some clear pat-
terns that emerge. Considering first the H-bond lengths, as the electronegativity of the atoms
diminishes, F > O > N, the bond elongates rather quickly. The two F atoms can come as
close as 2.3 A to one another, while R(O..O) is approximately 2.45 A, and the internitrogen
distance is just under 3 A. The F and O ions contain centrosymmetric H-bonds, whereas the
NH .. N bond is noncentrosymmetric (although the barrier separating NH .. N from N .. HN is
probably less than 1 kcal/mol). In the case of second-row atoms, the C1..H..C1 bond is cen-
trosymmetric. The anionic (HSH .. SH) complex probably is as well, while the cationic
(H 2 SH .. SH 2 ) + apparently contains a double-well potential. It is emphasized that the spe-
Weak Interactions, Ionic H-Bonds, and Ion Pairs 319

cific nature of the proton transfer potential, that is, single or double-well, height of energy
barrier, and so on, is very sensitive to particular level of theory so results can vary widely
from one study to the next, even with nominally similar approaches. At any rate, most of
these systems appear to have a potential which is quite broad at its bottom, and a harmonic
treatment is prone to errors.
With regard to the strengths of these ionic H-bonds, (FH..F)--- contains the strongest,
nearly 40 kcal/mol. Its cationic analog is considerably weaker, with a binding energy of 32
kcal/mol. The same is true for the second-row analogs where (C1H--C1)" is more strongly
bound than is (HC1H..QH)+. The situation reverses for the oxygen-containing systems
where it is the cationic (H2OH-OH2)+ that is more strongly bound than is (HOH .. OH) -- ,
by some 10 kcal/mol (as well as for the S analogs). The greater strength of cationic versus
anionic H-bonds containing O-bases has been observed in comprehensive experimental sur-
veys177, 184. This latter pattern holds for N as well where (H3NH..NH3)+ is much more
tightly bound than (H 2 NH .. NH 2 ) -- . The P-containing anion is probably not bound at all, as
noted in Table 6.8. It is emphasized that the H-bonds in the second-row ions are consider-
ably weaker than the first-row species. Nonetheless, even in such cases, the ionic character
of the complexes leads to interaction energies that tend to be larger than many of the
strongest neutral H-bonds.
The comparisons in bond strength between cation and anion are highlighted in Table
6.12 which illustrates that the aforementioned patterns are true at SCF as well as correlated
levels218. Correlation tends to add to the binding energy, particularly in the case of second-
row atoms. MP2 appears to provide results close to MP4 in most cases. MP4 results are il-
lustrated in Table 6.13 to nearly coincide with coupled cluster including triples219. (FH-F)~
is apparently the strongest H-bond, with an interaction energy of some 44 kcal/mol.
Many of the earlier calculations, up to 1987, have been compiled along with experi-
mental H-bond enthalpies220. Although somewhat dated, as most of the calculations are lim-
ited to SCF level with fairly small basis sets like 4-31G, many of the trends are illustrative
of what is seen at higher levels of theory.

6.3.5 Alkyl Substituents


Methyl substitution appears to have little influence upon the cationic complexes. For ex-
ample, R(O..O) in (CH3OH)2H+ is longer by only 0.01 A than in (HOH)2H+221. In both
cases, the H-bond is centrosymmetric. This insensitivity to methyl substitution has been
confirmed by experimental measurements222: The binding enthalpy of (CH3OH)2H+ is
within experimental error of that of (HOH)2H+. A small decrease is noted when both hy-

Table 6.12 Comparison of binding energies (kcal/mol) of cations


and anions. Data uncorrected for BSSE218.
Complex SCF MP2 MP4

(FH..F)- 42.5 45.3 44.6


(HFH..FH)+ 29.9 33.2 33.1
(C1H..C1)- 15.0 21.8 20.4
(HC1H .. CIH) 1 8.2 15.5 14.4
(HOH-OH) 23.9 28.3 27.7
(H2OH..OH2) ' 31.2 35.9 35.2
320 Hydrogen Bonding

Table 6.13 Comparison of correlated binding energies (kcal/mol)


of cations and anions. Data, uncorrected for BSSE, calculated
with aug'-cc-pVTZ basis set219.

Complex MP4 CCSD+T(CCSD)

(FH-F)- 43.3 44.2


(HFH..FH)+ 33.1 33.1
(HOH..OH)- 26.2 26.9
(H2OH..OH2)+ 33.8 33.8
(H2NH..NH2)- 13.9 14.1
(H 3 NH .. NH 3 )+ 26.4 26.3

drogens of HOH are replaced by methyl groups as in (CH3)2O)2H+. Similar results have
been noted with the NH +.. N interaction. As (H3N)2H+ is changed to proton-bound methy-
lamine, dimethylamine, and trimethylamine dimers, the binding enthalpy decreases slowly
from 26 kcal/mol for (H3N)2H+ to 22 kcal/mol for ((CH3)3N)2H+222. When various sym-
metric substitutions are made on the pyridinium-pyridine complex, there is little change ob-
served in the internitrogen separation (2.5 A), even though these substitutions account for
a span of over four pK units in the monomers223. Nor do these substitutions have much ef-
fect upon the proton transfer barrier calculated at the HF/6-31G level, which lies between
18 and 20 kcal/mol for all complexes.

6.3.6 Other Considerations


Strong ionic H-bonds can be formed also when the O atom is engaged in a double bond.
The proton-bound acetaldehyde dimer, for example, is computed to be bound by 28
kcal/mol at the SCF/4-31G* level224. R(O..O) is equal to 2.52 A, and the H-bond is non-
centrosymmetric at this level, but the two wells in the proton transfer potential are sepa-
rated by a barrier of only 1.7 kcal/mol. This barrier will likely vanish entirely when corre-
lation is added. Similar results were obtained for proton-bound formaldehyde dimer225,
with R(O .. O) = 2.51 A and a transfer barrier of 1.4 kcal/mol226. Replacing one hydrogen
of each formaldehyde by F leads to a slightly deeper pair of wells, separated by a barrier of
2.4 kcal/mol226. But again, these barriers are likely to disappear when correlation is in-
cluded. The combination of a pair of carboxylate anions and a proton also results in a po-
tential that has two minima at the SCF level, but a centrosymmetric H-bond when correla-
tion is included187. The cationic analog on the other hand, pairing two neutral formic acids
and a bridging proton, appears to retain a double-minimum potential even in its MP2 trans-
fer potential, for which R(O..O) is shorter by 0.05 A than in the SCF-optimized structure227.
The strengthening of H-bonds that occurs when one of the two partners is charged can
be sufficient to form H-bonds between C atoms. For example, a recent crystal structure228
revealed that a proton can bind together the C atoms of a pair of imidazole species. The H-
bond is apparently noncentrosymmetric with r(C—H) distances of 2.03 and 1.16 A; the
bond is within 8° of being linear.
There are certain manifestations of H-bonds which are either centrosymmetric or in
which the transfer potential contains a broad minimum, or a pair of minima separated by a
low barrier. In such cases, the proton position can be easily shifted by external influences
and its vibration extends over a wide amplitude. Consequently, the infrared spectrum con-
Weak Interactions, Ionic H-Bonds, and Ion Pairs 321

tains a very broad, intense band. The "proton polarizability" can be two orders of magni-
tude larger than would be the case in more traditional H-bonds223,229 233. One reason for
the large changes in moment associated with displacement of the proton along the H-bond
is that the electron density tends to shift in the direction opposite to proton motion, thus am-
plifying the entire effect234,235.

6.4 Asymmetric Ionic Systems

Now let us consider how the H-bond energy might be expected to vary when the two part-
ners are chemically different species. This sort of bond, where A and B of AH...B are dif-
ferent, is referred to as an "asymmetric" H-bond. The reader should be wary of the litera-
ture, however, where the use of this term can be ambiguous as it often refers to a
"noncentrosymmetric" H-bond as well. Taking the proton bound ammonia dimer as an ex-
ample, this H-bond would be considered symmetric in that the two partners are identical,
and the proton transfer potential is similarly symmetric. The proton transfer potential con-
tains a pair of minima, corresponding to (H 3 NH +... NH 3 ) and (H 3 N ...+ HNH 3 ), in each of
which the bridging hydrogen is closer to one N atom than to the other. This noncentrosym-
metric geometry is sometimes given the label "asymmetric" in the literature. To reword this
description, the proton transfer potential in any symmetric H-bond, containing a pair of
identical partners, must by definition also be symmetric (provided there are no geometric
restraints that impose an asymmetry into the system). The H-bond can be centrosymmetric
if the profile contains a single, central minimum, or noncentrosymmetric when there are
two equivalent minima present, separated by a local maximum when the proton is equidis-
tant between the two partners.

6.4.1 General Principles


As a starting point for our discussion, consider the proton-bound dimer between A and B0.
If these two species have the same proton affinity, the AH+ + BO pair would have the same
energy as A + H+BO, as depicted at the top of Fig. 6.13. The stabilization achieved when
the two interact to form a H-bond, A...H+BO, is indicated by the arrow labeled Eo in Fig.
6.13. (We assume for the moment that the H-bond would be noncentrosymmetric with the
proton closer to B0 than A.) Now we change BO to a more basic molecule, B 1 .Because the
latter has been defined to be more basic, H + B 1 is more stable than H + B 0 , so the energy of
the A + H + B 1 pair would be lower than that of the original A + H+BO. As illustrated in
Fig. 6.13, this greater stability is manifested also in the complex, with A ... H + B 1 more sta-
ble than A ... H + B 0 . The net result is that the more basic B1, forms a stronger H-bond with
AH + , as expected. That is, E1 is greater than Eo. The same trend continues as B becomes
progressively more basic as B2, and so forth. In summary, the H-bond formed between AH+
and B becomes progressively stronger as the basicity of B increases.
The reader may have noted that the formation of the H-bonds in Fig. 6.13 has involved
also the transfer (or at least partial transfer) of the proton from AH+ to the base within the
complex. That is, since B is a stronger base than A, A ... H + B is preferred over AH +... B. It
is thus of interest also to consider the H-bond energy of each complex from the perspective
of the "other" reactants: A + H + B. The arrows labeled E' in Fig. 6.13 refer to the reac-
tion A + H+B A...H + B where B serves as proton donor rather than A. It is evident that
the stabilization of the complex arising from the increasing basicity of B is not as large in
322 Hydrogen Bonding

Figure 6.13 Schematic diagram of the H-bond energies aris-


ing from formation of proton-bound dimers of A with Bi.

magnitude as the stabilization of H+B itself. As a result, the arrows shorten as B becomes
more basic; that is, less acidic: E'2 < E' l < E'O. In other words, the H-bond energy di-
minishes as the acidity of H+B drops. This trend is fully consistent with the pattern dis-
cussed above for AH+ + B where A donates the proton to the H-bond.
To encapsulate the qualitative trends illustrated in Fig. 6.13, the H-bond will be ener-
getically strengthened if either (i) the proton acceptor becomes more basic, or (ii) the acid-
ity of the donor increases. These rules will be in force whether or not the formation of the
H-bond incorporates the partial transfer of the proton from one group to the other within
the context of the complex.
Finally, it is worth stressing that one must be careful about the precise meaning of the
H-bond energy of a complex such as A ... H + B. It is clear from Fig. 6.13 that E'1 differs
from E1 ,for example. In other words, the H-bond energy is quite different, depending on
whether one takes as reactants AH+ + B1 or A + H + B 1 .(In fact, the discrepancy between
these two measures of the H-bond strength is equal to the difference in proton affinity be-
tween A and B1.)

6.4.1.1 Quantitative Relationships


At this point, it is useful to inquire into any quantitative relationship between the H-bond
energies in Fig. 6.13 and the proton affinities of the two partners. Numerous experimental
studies have indicated a strong correlation between these two quantities157,184,236 238. The
relationship usually tested is a linear one where the H-bond enthalpy varies as

where the slope m is less than unity, typically in the range 0.2 — 0.4239 244. The proton
affinity difference, PA, which obeys this relationship is fairly wide, extending over a range
of perhaps 60 kcal/mol239. H-bond types considered include ionic interactions between N,
0,C, and S bases184,239 242.
Weak Interactions, Ionic H-Bonds, and Ion Pairs 323

A number of concepts have been proposed over the years to explain the linearity of this
relationship, and deviations therefrom. Probably the most popular is due to R. A. Marcus
whose original ideas were developed for electron transfer and then extended to proton trans-
fer reactions245 249. These ideas apply primarily to the height of the barrier for these reac-
tions and how it is influenced by the exothermicity of the overall reaction. But one can cer-
tainly imagine how "flipping the energy profile upside down" could allow the same notions
to apply to the depth of energy minima.
In fact, the idea of applying Marcus theory to systematize the well depths of H-bonded
complexes has been described in the literature on a number of occasions250 253. The con-
cepts behind this approach are summarized in Fig. 6.14 which "fills out" the energy pro-
files for some of the reactions in Fig. 6.13. The symmetric system, that in which the proton
affinities of A and Bo are identical, is taken as a base point, and the depth of the energy well,
Eo, is denoted an "intrinsic" well depth. As Bo is changed to a stronger base B 1 ,the in-
crease in its proton affinity is described by PA and the new proton transfer profile is rep-
resented by the dashed curve in Fig. 6.14.
The Marcus formulation relates the new well depth ( E1) to the depth in the "unper-
turbed," (that is, symmetric) system, Eo, and the change in basicity of B, PA:

This equation is derived from a number of simplifying assumptions254. In summary, it is


presumed that each profile in Fig. 6.14 can be constructed by a pair of inverted parabolas.
The energy of the total system is taken to be that of the first parabola, but then switches

Figure 6.14 Profiles for the transfer of a proton from A to B0 and from A to
B1.
324 Hydrogen Bonding

abruptly to the second parabola at the point that they intersect. The increased basicity of the
base B is incorporated into the model by lowering the second parabola relative to the first,
which in turn lowers their point of intersection and increases E1 relative to EO. One of
the prime objections to this prescription is that it would lead to an incorrect shape for the
transfer profile, namely a sharp "cusp" rather than the rounded bottom. Nonetheless, the ap-
proach has seen some surprising successes.

6.4.2 Test of Quantitative Relationships


The accuracy of Eq (6.2) was rigorously tested using ab initio calculations in 1986 by tak-
ing as a starting point the proton-bound water dimer255 where A and Bo in Fig. 6.14 are both
modeled by H2O. Eo then becomes the computed energetics of Reaction (6.3):

The proton affinity of the base B was enhanced by replacing H2O by CH3OH, then by
C2H5OH, and by (CH3)2O. With the use of a 4-31G basis set, each of these substitutions
leads to respective increases in proton affinity of the O atom, relative to H2O, by 16.8,21.3,
and 26.4 kcal/mol256. The binding energies computed for each of these bases with H 3 O + ,
also at the SCF/4-31G level, are reported in Fig. 6.15 as the three data points. It is evident
that these binding energies are predicted quite well by the Marcus Eq (6.2), represented by
the solid curve in Fig. 6.15. Similar computations were carried out with the nitrogen analogs

Figure 6.15 Energetics of binding of H3O + to various alkylated O


bases. The data points were computed explicitly and the solid
curve is derived from Equation (6.2)255.
Weak Interactions, Ionic H-Bonds, and Ion Pairs 325

of Reaction (6.3), pairing NH4+ with various alkylated amines, and the results again con-
firmed the accuracy of Eq (6.2).
It might be noted that the solid curve in Fig. 6.15 is nearly a straight line, even though
the proton affinity difference extends over a 30 kcal/mol range. This approximate linearity
is consistent with the experimental gas-phase findings mentioned above. Making use of Eq
(6.2), the linearity can be traced to the fairly large value of the binding energy of the sym-
metric system, Eo, which reduces the impact of the third (quadratic) term in the equation.
As an example, even at the far right of Fig. 6.15, where PA is equal to 30 kcal/mol, the
third term is equal to only 1.3 kcal/mol, as compared to 15 kcal/mol for the second, linear
term. This approach would lead one to expect a much less linear relationship between H-
bond energy and the basicity difference of the two partners when the former quantity is
smaller in magnitude, that is, for weaker H-bonds.
It should be stated parenthetically that the Marcus formulation is not the only one that
could in principle reproduce the patterns of H-bond energy changes precipitated by proton
affinity changes. There are other extant theories257 261 that differ from the Marcus equa-
tion chiefly in the last term, a rather unimportant one in many cases as the relationship is
very nearly linear anyway.
Later work attempted to extend the applicability of this sort of analysis to systems which
are more asymmetric. For example, instead of altering the degree of alkyl substitution on
two O-bases in a H-bond, the atoms directly involved were different, for example, O and
N, or O and C262. The agreement was not as good, a fact which might be attributed to the
distinction between single and double-well character in the various systems, resulting in
geometry mismatches.
It might also be noted that computations at a higher level, and extending the interaction
energies from the electronic contribution to E to the full enthalpy AH, with allowances
made for zero-point vibrational energies, confirm the original conclusions. The addition of
a methyl group to one end of the nitrogen system, namely CH3NH4+...NH3, reduces AH by
3.5 kcal/mol at the MP2/6-311 + +G(d,p) level178, relative to NH 4 +... NH 3 , as compared to
a reduction in Eelec of 3.7 kcal/mol for SCF/4-31G256.
We finally address a fundamentally interesting question that may best be understood by
way of example. The transfer potential of the proton-bound ammonia dimer contains a pair
of minima, (H 3 NH +... NH 3 ) and (H3N...+HNH3), separated by an energy barrier255. As one
of the nitrogens, say the one on the left, is alkylated so as to make it more basic, the left-
hand minimum becomes more stable than the one on the right. As the process continues and
the left minimum is progressively lowered, the right minimum becomes merely a shoulder
in the potential and eventually disappears entirely, leaving only a single-well potential:
(R 3 NH +... NH 3 ). So beginning with a symmetric system that contains a pair of minima, the
potential is transformed into an asymmetric single-well potential when the proton affinity
of one base exceeds that of the other by a certain amount.
Chu and Ho have attempted to quantify this idea via a parameter, , which is directly re-
lated to the difference in protonation energies (PEs) of the two bases A and B:

where B is the less basic of the two. The authors found that when exceeds about 0.08, the
collapse from double- to single-well proton transfer potential occurs. In other words, one
might expect only a single minimum when the difference in proton affinity is greater than
326 Hydrogen Bonding

8% of the proton affinity of the subunits themselves. However, their data sample was rather
small, restricted to carbonyl oxygen atoms as bases. And it is clear that in some classes of
systems, one would not expect a double-well transfer potential regardless of the value of .
An example is provided by the (R 2 OH +... OR' 2 ) family of H-bonded hydroxyl groups,
where even the unsubstituted, symmetric (H2O..H+..OH2) contains a single centrosym-
metric minimum.

6.5 Syn-Anti Competition in Carboxylate

The carboxylate anion is of particular importance to enzymatic activity as it serves as the


functional portion of the Asp and Glu residues in proteins. As described above, there are a
number of geometrical dispositions in which the carboxylate can form H-bonds with other
groups. Of particular interest is the question as to whether this group will form a stronger
H-bond on its syn or anti side, with these terms defined in Fig. 6.16, with respect to the two
O lone pairs.
From the point of view of an incoming proton donor, with a partial positive charge, the
O atoms carry the bulk of the negative charge of the anion. (Indeed, one study places a neg-
ative charge in excess of — 1 on each of these atoms in formate263.) The donor might be ex-
pected to prefer the syn approach so that it could foster an interaction with both of these
atoms. Indeed, the difference in basicity of these two sides of the O atom has been consid-
ered to play an important role in general base catalysis264,265, so is potentially an important
issue.

6.5.1 Ab Initio Calculations


The water molecule has been the most commonly studied proton donor with regard to in-
teractions with the carboxylate anion. A number of modes of H-bonding that one might en-
vision are illustrated in Fig. 6.17. In addition to the aforementioned H-bonds between the
two O atoms in the "pure" syn and anti configurations, there is the possibility of the "lin-
ear" (lin) geometry, in which the water molecule lies directly along the C—O axis rather
than on either side of it. Another candidate is the "bifurcated" structure (bif) in which there
are two H-bonds present simultaneously. Although both of these bonds would be bent to
some degree, this geometry might be stabilized by the excellent alignment between the di-
pole moments of water and HCOO .
It is important to note the possibility of a smooth transition from one configuration to
the next, with only small motions of the water molecule necessary. In other words, the sys-

Figure 6.16 Definition of syn and anti directions in


the carboxylate group.
Weak Interactions, Ionic H-Bonds, and Ion Pairs 327

Figure 6.17 Idealized geometries for interaction


between a water molecule and the carboxylate an-
ion.

tern can smoothly transition from and to bif, passing through first lin and then syn along the
way. It is also worth stressing that each of these categorizations refers not so much to a sin-
gle structure as to a class of them. That is, the anti designation can be used for any geome-
try which places the water on that side of the pertinent O atom. One would not expect the
C—O . . . O angle to necessarily be precisely 180.00° in lin so the distinction between this
configuration and anti or syn is an arbitrary one.
An early study of this problem was limited to the SCF level with relatively small basis
sets266. The geometries were not fully optimized but were restricted instead to adjustment
of the intermolecular distance only. The bifurcated geometry was found most stable, with
an interaction energy in the 19-29 kcal/mol range. The lin geometry was considerably less
stable; syn and anti were not considered. A later work267 confirmed the stability of the bi-
furcated geometry, in comparison with syn and anti. Their data suggested that one might
reach an erroneous conclusion were one to use a basis set that is too small, like STO-3G.
At the SCF/6-31G** level, the bifurcated structure is favored over syn by 2.2 kcal/mol; the
preference is increased to 2.7 kcal/mol with MP2 corrections. The electronic contribution
to the binding energy of bif is 23.6 kcal/mol at the latter level. Geometries of syn and anti
were fully optimized soon thereafter268, within the framework of the 6-31G* basis set (syn
and lin were not considered here); bif was favored by 4.6 kcal/mol over anti. In 1988, Her-
mansson et al.269 computed the energies of nearly 600 points on the PES of this complex
at the SCF/DZP level so as to derive an analytical form of the surface. The internal geome-
tries of the HCOO and HOH species were frozen, so the surface was concerned with in-
termolecular variables only. BSSE was corrected by the counterpoise procedure at each
point. Their data confirmed the greater stability of the bif geometry.
The geometries of bif and anti were fully optimized and compared to syn, using a
4-31+G* basis set270. Bif was preferred with a binding energy of 20.5 kcal/mol at the MP2
level. The syn and anti geometries were virtually indistinguishable on energetic grounds,
both with binding energies of 17 kcal/mol. (The nonbridging water hydrogen was rotated
180° from that pictured in Fig. 6.17 to prevent the collapse of syn into the bif structure.) A
more recent MP2 optimization, employing a 6-31 + +G**(d,p) basis set178 confirmed the
bif structure as preferred on energetic grounds, favored by 2.8 kcal/mol over anti. This en-
ergy difference remains nearly the same when elevated to AH by including zero-point vi-
brational terms. The Gibbs free energy also favors bif over anti by 1.3 kcal/mol, after fac-
toring in entropic factors, although this difference diminishes to only 0.3 kcal/mol when the
temperature is raised to 500° K. When paired with ammonia as proton donor instead of wa-
ter, the syn and anti structures are equal in energy to within 1-2 kcal/mol 271 .
328 Hydrogen Bonding

The conclusion from these various ab initio calculations is that a water molecule would
prefer to approach the carboxylate anion so as to form the bifurcated geometry. If this struc-
ture is distorted by stretching one of the two H-bonds and pulling the water toward a lone
pair of the other oxygen, so as to form the syn geometry, this deformation would represent
an energetic cost of perhaps 3 kcal/mol. The syn structure per se, (not the bifurcated arrange-
ment) is very similar in energy to the anti geometry in Fig. 6.17, in which the water mole-
cule approaches the O atom from the other direction.

6.5.2 Experimental Findings


The theoretical conclusions are counterpointed by experimental data. There is some indi-
cation that a linear complex versus the bifurcated structure is favored in the gas phase184
although this preference is based on entropic rather than energetic factors. A survey of crys-
tal structures in which a carboxylate group interacts with various metals272 finds 63% of
the interactions are with the syn lone pair, as compared to 23% with anti (the remainder fall
into the bif category). However, the interactions between carboxylate and metal ions like
potassium and copper would be expected to be dominated by electrostatics in contrast to
the H-bonds when water acts as the partner. Moreover, it is difficult to isolate the intrinsic
preferences from the perturbing influence of crystal packing forces. Indeed, a survey of pro-
tein structures which focused on true H-bonds273 reached the conclusion of a weakened
preference for syn. While the ratio of syn to anti H-bonds of the Glu carboxylate group was
significantly greater than 1 (57:43), the same quantity was within statistical error of unity
for Asp. Together, the syn:anti ratio for all groups examined is 53:47. Another survey, this
time focusing on H-bonds to small-molecule carboxylate groups274, was able to account for
and remove crystal packing forces. The authors found no statistical preference for syn over
anti H-bonds (in the absence of steric interference) over a database spanning 876 interac-
tions. It would hence appear that crystal studies confirm the computational finding that there
is little energetic distinction between H-bonds to the syn and anti sides of the carboxylate
oxygen atom.

6.5.3 Carboxylic Group


If there is no strong distinction between the syn and anti sides of the carboxylate oxygen
atom with regard to formation of a H-bond, one might wonder what would happen when
the hydrogen is much closer, within range to form a covalent bond. Calculations have shown
that syn carboxylic acids (the Z-rotamers) are more stable than anti (E), by some 5-7
kcal/mol in the case of HCOOH270,275,276, consistent with microwave data that indicate an
energy difference of 4 kcal/mol277. It is in this context that the syn lone pair of HCOO
may be said to be more basic. This greater stability may be due to a favorable alignment of
the C=O bond dipole with that of the O—H bond in the acid, since the Z-rotamer has a
much smaller molecular dipole moment than E276,270. Similar arguments relating to align-
ments of bond dipoles have been invoked in explaining anomalous acidities of esters278,279.

6.5.4 Solvent Effects


In contrast to these calculations and microwave information, there have been numerous
studies of model systems that question the magnitude of the preferential stabilization of the
syn geometry280 282, which does not appear to exceed 1 pKa unit 283 285. It was conjectured
that perhaps the resolution of the discrepancy resides in solvent effects282 since the equi-
Weak Interactions, Ionic H-Bonds, and Ion Pairs 329

librium ratio of E to Z isomers of primary amides is known to be strongly solvent depen-


dent286,287. This premise was confirmed by later calculations that indicated that the differ-
ence in energy between the syn and anti geometries of acetic acid is reduced from 6 to 1
kcal/mol when the molecule is taken from the gas phase into aqueous solvent288,289. This
result can be taken as a simple consequence of the much larger dipole moment of the anti
geometry as compared to syn (about threefold), which is subject to proportionately larger
stabilization in solvent. Thus, while there is a clear energetic preference for the syn (Z) con-
former of carboxylic acids in the gas phase, this preference is reduced or eliminated alto-
gether when in aqueous solvent.

6.5.5 Resolution of the Question


Let us return now to the gas phase where the syn conformer of the carboxylic acid is fa-
vored. Is this observation in contradiction to the finding that there is no real preference for
syn or anti H-bonding of the carboxylate? The answer to this question is negative because
there is a difference between the two phenomena274. That is, just because a proton has a
preference for the syn side when it has formed a covalent bond with the O atom in question
does not necessarily indicate that the same side will be preferred when the proton is more
distant, as in HCOO ...HA. The distinction can perhaps best be explained for a simpler car-
bonyl group, as in formaldehyde. A proton would be added to the O atom so as to yield a
C — O — H angle in the neighborhood of 110° due to the hybridization of the O atom. How-
ever, as the proton is pulled away, there would be a tendency for the molecule to rotate so
that the C=O bond dipole can point toward the positive charge developing on the proton
acceptor, leading to a C—O . . . H angle closer to 180°290 292.
An example of how this principle would apply to the carboxyl group is illustrated in
Fig. 6.18270,271. While the syn geometry of a carboxylic acid is intrinsically more stable
than the anti, if the proton extracted by carboxylate comes from a neutral molecule, the re-
sulting —COOH acid will be paired with an anion. As indicated in Fig. 6.18, a destabiliz-
ing electrostatic interaction can occur with the partially negatively charged O atom when
the anion is in the syn position, but not if it is anti. This destabilization can dampen, or negate
entirely, the intrinsic energetic preference of syn versus anti carboxylic acids. This concept

Figure 6.18 Illustration of the potential destabilizing elec-


trostatic interaction between the carbonyl O atom of the
carboxylic acid and the anion in the syn position (S). No
such destabilization occurs if the anion is anti (A).
330 Hydrogen Bonding

was further elaborated and shown to be in operation in some form regardless of whether the
carboxylic acid interacts with an anion or a neutral molecule293.
In summary, there is little distinction between the syn and anti lone pairs of the car-
boxylate oxygen atom with regard to forming H-bonds with proton donors. When a proton
has approached closely enough to form a covalent bond, the syn position is favored, but
only in the gas phase. The syn and anti conformers of the carboxylic acid are close in en-
ergy in aqueous solution. Even in the gas phase, the preference for the syn configuration of
the isolated carboxylic acid can be eliminated when it forms a H-bond, due to more favor-
able electrostatic interactions between the partner and the anti geometry of the carboxyl.

6.6 Neutral Versus Ion Pairs

As the acidity of the proton donor AH increases, and there is a concomitant enhancement
in the basicity of the acceptor B, the H-bond will become stronger. But it also stands to rea-
son that for a strong enough pair of acid and base, the proton can simply transfer across
from A to B, converting the system into an ion pair:

An important force working against such a transfer is the energetic difficulty of separating
charge, so such a transfer will occur only for the strongest acid-base combination. On the
other side of the equation, once formed, the ion pair will be held together by a powerful
electrostatic ion-ion attraction, a strong consideration in its favor. In addition, moving the
complex from complete isolation in the gas phase into an environment where the medium
can interact with the system to help stabilize the charges will clearly favor the ion pair. This
bias will become stronger as the polarity of the surrounding molecules or solvent increases.

6.6.1 Amine-Hydrogen Halide


The question as to what it takes to generate an ion pair (sometimes referred to as a "salt
bridge"294) has generated a good deal of inquiry in the literature2. Much of this work
has focused on the hydrogen halides as the strong acids, and the amines as the requisite
strong base. Indeed, a very early ab initio study of a proton transfer reaction was directed
at the possibility that H 3 N ... HC1 might spontaneously undergo the reaction to form
H3NH+... C1295, although a somewhat later calculation disputed this prediction296. A later
and more thorough examination as to how the nature of the complex is affected by the level
of theory297 illustrated that certain "mid-range" basis sets are prone to error in this matter.
The minimal basis sets tested, with and without polarization functions, indicated it is only
the neutral pair that is present. This result is confirmed by more flexible sets, containing
multiple sets of polarization functions. In contrast, application of the 3-21G basis set sug-
gested it is the ion pair that is favored in the gas phase. Addition of diffuse functions leads
to the same conclusion, whereas inclusion of polarization functions on non-hydrogen atoms
points instead to the neutral pair. Another split-valence set, MIDI-1, shows the same pat-
tern; still another, 4-31G, indicates two separate minima in the potential energy surface, one
corresponding to the neutral pair, the other to the ion pair. This set of calculations297,298 un-
derscored the necessity to use an appropriately flexible basis set in examining the relative
stabilities of the neutral and ion pairs. Interestingly, inclusion of electron correlation via
MP2 does not change the central conclusion of a neutral pair for this complex298.
Weak Interactions, Ionic H-Bonds, and Ion Pairs 33 I

Since H3N...HC1 exists only as a neutral pair and HF is less acidic than HC1, it is not sur-
prising that H3N...HF, too, does not form an ion pair299. Similarly for H3P...HC1 and
H3P...HF, as H3P is less basic than H3N300. The latter set of calculations is particularly sig-
nificant as it included the effects of electron correlation. The list of complexes found to be
neutral-pair only at the correlated level was later extended to H3P...HBr, H3As...HCl, and
H3As...HBr301. Other systems examined at the SCF level, and failing to show evidence of
an ion pair, were H 3 N ... HBr and H3As...HF302.
If HC1 is not acidic enough and NH3 not basic enough to form an ion pair, it was thought
that perhaps HBr might be sufficiently acidic and/or CH3NH2 basic enough. The potential
energy surfaces of NH3 and CH3NH2 paired with HC1 and HBr were scanned with a DZP
type basis set at the SCF level303. The data confirmed the lack of an ion pair for H3N...HC1.
The surface for CH3NH2...HC1 is somewhat flatter, but again contains only a single mini-
mum corresponding to the neutral pair. This result contradicted an earlier SCF/4-31G find-
ing of a CH3NH2H+... C1 ion pair304, reemphasizing the sensitivity to basis set. The sur-
face for H3N...HBr303 was similar, but did contain a hint that there might perhaps be a
second minimum for the ion pair. However, if it did, this minimum would probably not be
deep enough to sustain a vibrational level, and would lie some 4 kcal/mol higher in energy
than the neutral pair. In the case of CH3NH2...HBr, on the other hand, the surface contained
two clear minima. The neutral and ion pairs are of comparable energy (the ion pair is fa-
vored by 1.2 kcal/mol) and both appear deep enough to contain at least one vibrational level.
In contrast to the neutral pair for HC1 + methylamine, combining the same acid with the
more basic trimethylamine does appear to produce an ion pair at the MP2 level305.
The acidity was turned up one notch when NH3 was paired with HI306. This study in-
cluded the effects of electron correlation. In agreement with prior results, the surfaces of
both H3N...HC1 and H 3 N ... HBr contain only a single neutral-pair minimum. H 3 N ... HI, on
the other hand, exists both as the neutral and ion pairs. Whereas the neutral is favored by 5
kcal/mol at the SCF level, the two types of complex are nearly equal in energy when cor-
relation is added. On the other hand, zero-point vibrational contributions to the energy are
quite distinct for the two minima. Whereas this effect destabilizes the neutral pair by 1.6
kcal/mol, the ion pair is raised in energy by 5.6 kcal/mol, leading to the near disappearance
of the minimum in the potential. It is concluded unlikely that H3NH... I would actually
be observed.
Because of the possible occurrence of two minima in CH3NH2...HBr, this complex was
the subject of further scrutiny, this time including electron correlation307. The calculations
confirmed the possibility of two minima at the SCF level; moreover, the earlier hint of a
possible second minimum for H 3 N ... HBr was verified here. The data suggested further that
the relative energies of the neutral and ion pairs are really quite sensitive to the details of
the basis set. But the most important finding of this set of calculations is that correlation
can cause the collapse of a double-well potential to one with only a single minimum. The
MP2 proton transfer potential in either case contains a single broad minimum. Changing
the base from H3N to CH3NH2 shifts the bottom of this well slightly away from the Br and
toward the N. This trend opened up the possibility that it is perhaps incorrect to look for ei-
ther the classical neutral or ion pairs in all cases; the transition may in fact be a gradual one.
In other words, as the base and acid are strengthened, there may be some point where the
structure would be better described as A ..+ H--B, with the proton somewhere near the mid-
dle of the bond.
The calculations have demonstrated that the correlated potential energy surface of this
sort of system typically contains a single minimum, and that the position of this minimum
332 Hydrogen Bonding

shifts progressively from acid to base as the strength of the two increases. A comprehensive
examination of the strongest HX acids was completed, with the full spectrum of methylated
amines from H3N to (CH3)3N308. HI forms what can best be characterized as an ion pair
with all of the amines. The only exception is with the weakest base, H3N, where the sur-
face contains not a well-defined minimum, but rather a long shallow valley connecting
H3NH+... I with H3N...HI. In other words, the proton will undergo very large amplitude
vibrations between these two configurations and the complex is some mixture of the two.
HBr is a weaker acid than HI. Nevertheless, it too forms ion pairs with the more basic
amines. In the case of its complex with methylamine, the proton is very nearly equally
shared between the N and Br atoms. The less basic H3N is not capable of extracting the pro-
ton from HBr, and so this complex is best described as a neutral pair.
Table 6.14 contains a current summary of the nature of the complexes of the various
amines, coupled with each of the HX acids, as determined by ab initio calculations. HF does
not form ion pairs at all, whereas HI forms ion pairs with all (with the exception of the shal-
low valley in the potential with NH3). In the case of HBr and HC1, one can see the transi-
tion from neutral to ion pair as the amine becomes more basic.
The aforementioned trend of a gradual shift of the proton from acid to base, taking the
complex in stages from the neutral toward the ion pair, can even be observed with a mini-
mal basis set at the SCF level, given enough care. Complexes pairing HF, HC1, and HBr
with H3N and its mono-, di-, and trimethyl derivatives were constructed and studied first in
the gas phase309. Despite the use of a minimal basis set, the computed basicities of the
amines were in surprisingly good agreement with experimental protonation energies. Each
substitution of a hydrogen atom of the amine by a methyl group adds an increment of 5-9
kcal/mol to the protonation energy; similar steps were calculated for the deprotonation en-
ergy of the series HBr, HC1, HF. Each increment in acidity of HX or basicity of the amine
pushes the bridging proton a little closer to the base in the complex.
The progressive shift of the position of the single minimum was reaffirmed in a set of
correlated calculations at the MP2/6-31+G(d,p) level310. A single acid, HC1, was paired
with a series of nitrogen bases, in this case 4-substituted pyridines. As the pyridine became
more basic, the Cl—H bond was progressively elongated, until eventually the proton's equi-
librium position was approximately equidistant between the Cl and the pyridine. Further
enhancement of the basicity led to a full transfer of the proton, forming an ion pair, as may
be seen in Fig. 6.19.

Table 6.14 Character of complexes pairing hydrogen halides with amines, as


determined by ab initio calculations.a
HF HC1 HBr HI

NH3 NP NP NP NP IP valley
MeNH2 NP NP 50% IP
Me2NH NP ? IP IP
Me3N NP IP IP IP
a
NP = neutral pair; IP = ion pair; 50% indicates equilibrium proton position about halfway between N
and halide atoms. ? indicates the situation is still questionable.
Weak Interactions, Ionic H-Bonds, and Ion Pairs 333

Figure 6.19 Calculated values of the R(C1—H) bond length (solid line)
when HC1 is paired with a series of progressively stronger bases, 4-substi-
tuted pyridines. Data310 calculated at the MP2/6-31 +G(d,p) level. Also
shown (dashed line) is the stretching frequency involving the central
proton.

6.6.1.1 Quantitative Measure of Degree of Proton Transfer


None of the complexes pairing an amine with a hydrogen halide were computed to be of
the pure ion pair variety with a minimal basis set309. However, there were some that were
close. More specifically, since the equilibrium position of the proton need not shift precip-
itously from one atom to the other, but rather can move gradually as the acidity and basic-
ity increase, a proton-transfer parameter was devised to indicate the degree of transfer of
the proton from the acid to the base. The quantity was defined as

where r(NH) is the stretch in the N—H bond, relative to the isolated protonated amine,
and r(XH) has a similar meaning for the hydrogen halide HX. A negative value of
indicates a smaller stretch for HX than for the amine, and is hence a sign of a neutral pair;
in the reverse situation of a greater stretch for HX, becomes positive and one can con-
sider the complex to approach an ion pair. A zero value for p denotes equal sharing of the
proton.
The values calculated for this proton-transfer parameter are shown in Fig. 6.20 as a func-
tion of the acidity of A and the basicity of B. More precisely, the horizontal scale is a "nor-
334 HydrogenBonding

Figure 6.20 Proton transfer parameter, , computed for various


methylated amines paired with hydrogen halides. The numerical label
n on each data point refers to the number of methyl substitutes on the
amine N(CH3)nH3_n. The horizontal scale is a normalized difference
in proton affinity between the amine and the halide309.

malized proton affinity difference" (NPAD)311 between the two species competing for the
proton, the amine (Am) and X .

Since the proton affinity of the halide is greater than that of the neutral amine, NPAD is of
negative sign. The less negative is NPAD, the closer in magnitude will be the intrinsic pulls
on the proton by the two species.
It might first be noted that is less than zero in all cases, indicating that the optimized
complex resembles a neutral pair more than it does an ion pair. There is a clear trend in all
cases for to become less negative as the number of methyl groups on the amine increases.
In other words, the more basic amine increases its pull on the proton, leading the complex
closer toward an ion pair. In the case of HF, this effect is negligible: the complex is clearly
a neutral pair for all amines and the number of methyl groups has little effect on p. On the
other extreme is the set of complexes with HBr. Whereas < —0.3 for H3N...HBr, this pa-
rameter rises quickly, reaching almost zero for (CH3)3N...HBr. In other words, the latter
complex can be characterized as about halfway between a neutral and ion pair. HC1 is in-
termediate in its behavior, with less sensitivity of the proton equilibrium position to the ba-
sicity of the amine.
The data in Fig. 6.20 illustrate an important point concerning neutral and ion pairs. At
first blush, one might expect that should approach zero as the proton affinities of the amine
Weak Interactions, Ionic H-Bonds, and Ion Pairs 335

and the halide equalize. However, that notion would ignore a very important fundamental
difference between the two types of complex. Whereas the neutral pair AH ... B is held to-
gether by the "standard" H-bond forces, typically in the range of 5-15 kcal/mol, A ...+ HB
has a much stronger attractive force consisting of the electrostatic interaction between the
cation and anion. As a consequence, B does not have to have so strong an intrinsic proton
affinity as does A, in order to equalize the energies of the neutral and ion pairs.
Referring to a projection of the HBr plot in Fig. 6.20, this curve would intersect the =
0 axis at approximately NPAD = —0.21. Making use of Eq (6.7), the proton affinity of the
base which would be able to equally share the proton with Br would be 235 kcal/mol, 125
kcal/mol smaller than the proton affinity of Br . The latter quantity, then, represents ap-
proximately the enhanced binding energy between the entities in an ion pair, as compared
to that in the neutral pair, at least for this particular system. This value agrees surprisingly
well with an estimate of 113 kcal/mol, representing the purely Coulombic interaction en-
ergy betwen a pair of point charges, separated by the experimentally determined R(N ... Br)
in the ion pair formed between HBr and trimethylamine312.
The reader is reminded that Fig. 6.20 represents uncorrelated calculations with a mini-
mal basis set. Although quantitatively unreliable, the data illustrate useful trends nonethe-
less.

6.6.1.2 Environmental Effects


As indicated, it is possible to preferentially stabilize the ion-pair side of the proton transfer
potential by allowing the complex to interact with a polarizable medium. By incrementally
raising the polarizability, one can manually fine-tune the relative stability of the ion pair rel-
ative to the neutral pair. This lowering of one end of the proton transfer potential relative to
the other is very much akin to an adjustment of the acidity and basicity of the partners. When
the systems described above were immersed in a dielectric continuum model of a polariz-
able medium, the equilibrium position of the proton did indeed shift toward the base309. It
was possible to raise to the point where it became positive, indicating the complex had
more ion pair character than neutral pair. Just as the sensitivity of p to the basicity of the
amine increases in the order HF < HC1 < HBr, so too does its sensitivity to the polariz-
ability of the medium. Whereas complexes with HF show little displacement of the proton's
equilibrium position as the polarizability of the medium increases, those in which HBr acts
as the acid experience a rapid increase in with more polarizable medium. This treatment
was later confirmed at the correlated MP2 level in that the H3N...HC1 neutral pair is su-
perceded in stability by the corresponding ion pair when immersed in a dielectric contin-
uum 298.

6.6.2 Carboxyl/Carboxylate Equilibrium


The carboxyl group occupies a prominent place in protein structure and function and is com-
monly taken to be ionized as a carboxylate, particularly when paired in a salt bridge with a
base4,294,313,314. Whereas this may be the case in a protein environment, it is not obvious
that a —COOH group will donate its proton to a base to form an ion pair in the gas phase.
For example, SCF calculations with a polarized 6-31G basis set indicate that neither methy-
lamine315 nor an arginine model 316 is able to extract a proton from acetic acid so as to form
an ion pair. On the other hand, there is some indication that correlation might stabilize the
ion pair for formic acid + methylamine 317.
336 Hydrogen Bonding

This question was explored in some detail by ab initio calculations pairing formic acid
with methyleneimine, in which the N atom participates in a double bond318. As illustrated
in Fig. 6.21, the two configurations examined place the N atom either syn or anti with re-
spect to the carboxyl group. The calculations were carried out with a 4-31G basis set at the
SCF level. (These results were later confirmed at the correlated MP2 level with a 6-31G*
basis set319.) The protonation energies of HCOO~ and NHCH2 at the SCF/4-31G level of
theory are 360 and 230 kcal/mol, respectively. When adjusted by vibrational and other
terms, the calculated values of AH(300 K) are 352.6 and 221.9 kcal/mol, only a little bit
larger than the experimental quantities of 345.2 and 214.3 kcal/mol, respectively. Most im-
portant for our purposes, the theoretical overestimation is equal to 7 kcal/mol in both cases;
consequently, the calculated protonation enthalpy difference of 130.7 kcal/mol is very close
indeed to the experimental value of 130.9 kcal/mol. And it is this proton affinity difference
which is the key element in considering the question of neutral versus ion pairs.
When the intermolecular R(O..N) distance is taken to be 3.0 A, the neutral pair is fa-
vored over the ion pair in both the syn and anti geometries. This preference amounts to some
15 kcal/mol for the former configuration and more than 40 kcal/mol for the latter. The par-
ticular instability of the anti ion pair can be understood on the basis of electrostatics. Fol-
lowing proton transfer, one of the negatively charged O atoms of HCOO will be turned
away from the positively charged +NH2CH2 in the anti geometry, a particularly unfavor-
able situation, but not in syn where both oxygens will be oriented toward the cation. Other
calculations320 have confirmed the preference for the neutral pair, even in the absence of a
restriction on R(O .. N). Relaxing the restriction of the H-bond distance, permitting the
geometry to be fully optimized, yields a neutral pair as the only minimum on the potential
energy surface.
As pointed out earlier, permitting the complex to interact with a polarizable medium
would preferentially stabilize the ion pair. Using a self-consistent reaction field formalism
to model such interactions, the calculations did indeed witness such preferential effects 318.
Increasing the dielectric constant of the medium from unity (gas phase) to 2 lowered the
ion pair energy of the syn structure to the point where it became competitive with the neu-
tral pair; further increases in made the ion pair more stable. This reversal is an important
one as other studies estimate the average dielectric constant in the protein interior as being
in the range between 2 and 4321. These results suggest that even small interactions of the
complex with the surroundings can produce strong effects on the neutral/ion pair balance.
This idea is confirmed by the influence of an inert matrix, described below.
Although differing in quantitative aspects, other ab initio calculations using a different
model of solvation, based upon the Born equation, verify that interaction with a medium

Figure 6.21 Syn and anti configurations of the complex pairing formic acid and
methyleneimine.
Weak Interactions, Ionic H-Bonds, and Ion Pairs 337

will preferentially stabilize the ion pair in the imine-carboxyl complex, to the point that the
ion pair may be preferred322. The same sort of preference for the ion pair has been noted in
the complex of formic acid with trimethylamine when a polarizable medium is simu-
lated 323.
Still another means of simulating some of the effects of solvation are to introduce an
electric field as a perturbing influence upon the H-bonded solute system324. However, even
a field as strong as 5 x 107 V/cm is not capable of inducing a second well in the proton-
transfer potential of the complex between methanesulfonic acid and dimethyl sulfoxide,
which prefers to remain a neutral pair. The preceding calculations were carried out at the
MP2 level using a 3-21G* basis set. Another model of solvent effects would be to place a
number of discrete solvent molecules around the system of interest. Calculations have sug-
gested that the addition of just two water molecules is sufficient to convert the equilibrium
tautomer of glycine to its zwitterionic form, containing the COO and NH3+ groups 325.
Other calculations indicate little difference between these two approaches, (continuum ver-
sus discrete molecules) at least with respect to the (H 3 N .. H +.. OH 2 ) system 326.

6.6.3 Experimental Confirmation


During the period when theorists were addressing the question as to the possible existence
of ion pairs, there was a good deal of experimental inquiry as well. The neutral pair nature
of H3N...HC1 was confirmed in the gas phase by its rotational spectrum327. The intermole-
cular distance derived from the spectral data was rioted to be in remarkable agreement with
an earlier ab initio computation with MP2 correlation 299. The following year saw the mea-
surement of the microwave spectrum of H 3 N ... HBr 328, and this complex too had the char-
acteristics of a neutral pair. Nor are any of the complexes pairing a hydrogen halide with
H3P of the ion-pair type.
The stronger base trimethylamine was paired with the acid HCN in the gas phase329. Ap-
parently, HCN is simply not a strong enough acid to form the ion pair, even with trimethyl-
amine. The same is true of HF, which is also unable to donate its proton to trimethyl-
amine330. HC1 is another story, however. When this acid is combined with trimethylamine,
there is evidence for an ion pair in the gas phase331, particularly from the nuclear quadru-
pole coupling constants. The data are consonant with the broad minimum in the proton
transfer potential, with a minimum near the Cl atom, suggested by correlated ab initio cal-
culations. It is intriguing that when HC1 is paired with (CH3)3P, a neutral pair is the result 332.
This point is notable as the proton affinities of (CH3)3P and (CH3)3N are nearly identical.
The difference in behavior may be due to the larger size of the P atom as compared to N.
As a result, the two ions cannot get as close together should they be formed, and the di-
minished Coulombic interaction would mitigate against formation of the ion pair.
Perhaps the most comprehensive confirmation of the principles arising from the ab ini-
tio calculations comes from the series of complexes pairing mono and trimethylamines with
HBr and HC1312. The nuclear quadrupole coupling constants and stretching force constants
support the idea that increasing acidity of HX or basicity of the amine leads to a smooth in-
crease in the extent of proton transfer from X to N. Both HBr and HC1 appear to form an
ion pair complex with trimethylamine, but not with the monomethyl derivative nor with
H3N. Formulating an ad hoc measure of the degree of proton transfer, based upon the cou-
pling constants, the authors concluded that the proton is 62% transferred in (CH3)3N...HC1,
and 75% in the Br analog333. In agreement with the ab initio calculations, the pair of the
strongest acid, HI, with the strongest base, trimethylamine, leads to a product which is
338 Hydrogen Bonding

clearly an ion pair334. H3N...HI, on the other hand, would appear to remain a neutral pair.
Work in solution also offers some verification of this sort of behavior. For example, by pair-
ing 1 -methylimidazole with a series of acids with varying pKa, NMR and IR data indicated a
gradual shift of the proton from the acid to the base335. Similar properties have been noted in
the complex involving acetic acid and pyridine, in a variety of concentration situations336, and
also when substituted pyridines are paired with trifluoroacetate in dichloromethane solvent 337.

6.6.3.1 Matrix Isolation Studies


The situation in the gas phase, then, appears to be entirely consistent with results obtained
with ab initio calculations, particularly those including electron correlation. Matrix isola-
tion studies are germane as well, although the forces between the H-bonded system and the
matrix may influence the results. While the geometries of the structures cannot be worked
out by rotational spectroscopy in a matrix, a good deal of information is available through
the vibrational absorptions. Of particular relevance to the question of the existence of ion
pairs is the stretching frequency of the proton donor, vs. Collecting together the spectra of
a number of complexes pairing a hydrogen halide with O and N bases 311, Pimentel and
coworkers constructed a "vibrational correlation diagram" to illustrate the transition from
neutral to ion pair in N2 matrix. When the red shift of the vs mode (actually the percentage
shift vs ,/vs ) is plotted against NPAD between the halide anion and the base (see earlier), a
deep minimum is noted in the vicinity of NPAD = —0.23. (Since vs/vs is a negative quan-
tity, this minimum corresponds to a sharp increase in the absolute magnitude of the shift.)
The behavior of the vibrational correlation diagram was described in terms of the
strengths of the bonds to the bridging hydrogen in question. Considering first the neutral
pair AH ... B, it is well understood that increasing the acidity of AH and/or the basicity of B
produces a stronger H-bond. The latter is accompanied by a stretch of the A—H bond and
a lowering of the stretching frequency of this bond. One would expect this pattern to con-
tinue as the bond grows stronger. The same principles apply to the opposite ion pair as a
reference point. Starting from A ...+HB, an increase in the basicity of A or in the acidity
of +HB should yield a displacement of the proton away from B and towards A, accompa-
nied by a red shift in the vs frequency, now associated with the H—B bond. Since vs drops
off from either the AH ... B or the A ...+HB end of the continuum, it is not surprising that
this frequency will attain a minimum when the bridging proton's equilibrium position lies
somewhere around the middle of the A...B bond.
It is particularly notable that when the frequency is plotted as a relative red shift, vs/vs,
and the acidity/basicity of A and B is taken as a normalized proton affinity difference, it is
possible for a single function to fit the spectral data for a range of different acids and bases.
A central conclusion of these results matches nicely with what was found from both ab ini-
tio calculations and from gas-phase measurements: the system shifts smoothly from a neu-
tral to an ion pair as the proton affinities of the partners change.
When the proton affinities of A and B are precisely equal, NPAD would be zero. This
is not where the minimum in the vibrational correlation diagram was found. This displace-
ment from zero is related to an earlier explanation that the ion pair A ...+ HB has a much
stronger attractive interaction between the partners than does the neutral pair AH ... B. So the
proton affinity of A must be that much larger than that of B in order to compensate for this
fact. Consequently, the minimum in the diagram was found to occur at NPAD = -0.23.
Note the similarity of this value to that obtained from crude minimal basis set estimates of
the complexes pairing HBr with amines above in Fig. 6.20.
Weak Interactions, Ionic H-Bonds, and Ion Pairs 339

Some of the quantitative aspects of a diagram of this type can be gleaned from inspection
of Fig. 6.19 in Section 6.6.1, which illustrates the calculated frequencies in the complex be-
tween HC1 and a series of 4-substituted pyridines310. For weak bases, on the left side of the
figure, the complex is described as a neutral pair, ClH—pyridine, and it is the C1H stretching
frequency that is plotted. This frequency is shifted to the red as the base becomes stronger.
For much stronger bases, the proton is fully shifted and the complex is best represented by
the Cl •••H+pyridine ion pair, and the frequency of interest refers to the H—N stretch. In the
middle region, where the proton is shared between the two species, the harmonic modes are
more complex. A single intense mode is replaced by a pattern of several strong ones, all of
which are of much lower frequency. (It is the highest of these that is plotted in Fig. 6.19.)
One can also construct a diagram of similar character by plotting the isotopic ratio of the
vs frequencies (VAH/VAD) against a measure of the shift in this frequency 338. Specifically,
the latter quantity is defined as (vs/vs°) where vs° refers to the frequency measured in the
uncomplexed monomer. A sharp minimum is observed in this plot when the relative fre-
quency shift is about 0.5.
It might also be pointed out that correlation diagrams of this type have relevance in solu-
tion work as well, even though the vibrational bands are much broader. The maximal shift in
the spectral band for complexes of pyridines with carboxylic acids occurs for a proton affin-
ity difference of 109 kcal/mol; pairing of pyridine N-oxides with these acids peaks at 102
kcal/mol339. When plotted as relative frequency change versus NPAD, the functions have
minima at values of the latter parameter somewhat less negative than the —0.23 noted above.
It is interesting to compare the extent of proton transfer determined for complexes in. in-
ert matrices with those observed in the gas phase and predicted by ab initio computations.
Consistent with the latter, H3N..HF is found to be a neutral pair in Ar matrix340, but the
complex is somewhat more polar than it is in the gas phase. In fact, the frequency shift of
the HF stretch rises quite noticeably as the polarizability of the matrix increases from Ar to
N2. One can hence expect the matrix results to typically indicate greater proton transfer
character than observed in the gas phase. Nonetheless, HF does not appear to form an ion
pair with even stronger bases such as trimethylamine in Ar341, nor with H3P342. In fact, HF
does not release its proton to form an ion pair even in the 1:2 H3N..HF..HF complex where
cooperativity would tend to enhance the acidity of the central molecule 343.
A higher degree of proton transfer in N2 as compared to Ar was observed also for the
similar H3N..HC1 complex, and in complexes with HI344. A systematic examination of a se-
ries of related complexes 345 indicated that in Ar matrix, the proton is very nearly equally
shared between HC1 and methylamine, unlike the gas phase where this complex is a neu-
tral pair. This same point, that is, equal sharing, occurs for the less basic H3N in N2. The
degree of transfer continues to progress to the point where the complex is an ion pair when
HC1 is paired with trimethylamine in N2. The latter complex is probably an ion pair even
in the gas phase. On the other hand, HC1 does not seem to form an ion pair with H3P in
Ar342. Turning to the more acidic HBr346, the point of equal sharing occurs when it is paired
with ammonia in Ar. The degree of proton transfer, that is ion-pair character, increases for
the alkylated amines or in N2 matrix.

6.6.4 Long Chains


The transfer of a proton from the proton donor to the acceptor in an amide-amide dimer
would, of course, result in formation of a high energy ion pair, as depicted in Fig. 6.22, and
is consequently unlikely.
340 Hydrogen Bonding

Figure 6.22 Opposite charges acquired by two amide


molecules as a result of a proton transfer from donor to
acceptor.

If one has a large number of amide groups lined up in a long H-bonded chain, the ter-
minal molecules acquire the charges indicated in Fig. 6.22, but the remaining molecules
throughout the chain remain electrically neutral. The primary change occurring on each of
these non-terminal chain molecules is a tautomerism as indicated in Fig. 6.23. Such large-
scale tautomerism has been proposed to be the principal component in the mechanism of
long-range proton transfer in proteins347 350. Inelastic neutron scattering experiments of N-
methylacetamide and polyglycine have lent support to the possibility of dynamic exchange
of the proton between N and O351,352.
There are, however, a number of observations which argue against the viability of such
a string of transfers. Firstly, the imidic acid resulting from the tautomerism is considerably
higher in energy than the amide. Theoretical and experimental data provide an estimate of
an 11-12 kcal/mol energy difference in formamide353. The tautomeric equilibrium constant
is only some 10 8.192 The proton transfer would be impossibly slow: it would take more
than a month to produce an imidic acid intermediate at pH 7. Moreover, neutron diffraction
measurements of the acetanilide crystal further refute these notions of massive tau-
tomerism354. There is no evidence found for any significant degree of proton transfer from
Weak Interactions, Ionic H-Bonds, and Ion Pairs 341

Figure 6.23 Tautomerism in each of the amide units in a chain resulting from a series
of proton transfers from one group to the next.

N to O along the amide chain in the temperature range between 15 and 295 K. The authors
estimate that any such tautomerization would account for no more than 1% of the total sta-
tistical population of the crystal.

6.7 Summary

6.7.1 Low Polarity of Acceptor

The typical proton acceptor molecule contains an electronegative atom which is involved
in a bond with another atom of much lesser electronegativity. The ensuing bond dipole is
an important factor in formation of a H-bond. A number of systems were investigated in
which the dipole of the pertinent bond was fairly low.
342 Hydrogen Bonding

Despite the lone pairs on an electronegative atom, a dihalogen molecule like C12 or FC1
does not appear to act as a strong proton acceptor. The very low polarity of the dihalogen
bond greatly reduces the electrostatic component of the interaction energy. As a result, the
configuration wherein the proton of HF approaches one of the available lone pairs becomes
less stable than an entirely different dimer geometry in which the hydrogen plays no direct
part in the interaction.
Similar considerations apply to C=O, which is a highly nonpolar molecule. The proton
of HF prefers to approach the C atom rather than the O, consistent with the direction of the
dipole moment of the CO molecule. The red shift of the HF stretch is severalfold larger for
OC---HF than for CO"HF; the stretch of this HF bond in the former complex, although
small, is consistent with a H-bond, while there is no measurable stretch in CO--HF. The en-
ergetics indicate a H-bond may indeed exist in OC'"HF, with an electronic contribution to
AE of some —3.6 kcal/mol, as compared to only —1.1 kcal/mol in CO-HE
CO2 is completely nonpolar as a molecule, but the quadrupole moment is substantial and
consistent with O atoms with a partial negative charge. There is a noticeable red shift of the
HF stretch in OCO--HF, and r(FH) stretches a small amount. The interaction energy is
small, — 2.4 kcal/mol, but greater than in CO-HE The structure of this complex is funda-
mentally altered when HF is replaced by HBr, forming a T-shape with the hydrogen ap-
proaching the C atom of OCO from above. Interaction of HF with the isoelectronic NNO
makes the oxygen a slightly better proton acceptor while the terminal nitrogen is inferior to
the O in OCO. It is curious that, despite the different strength of binding to the N or O ends
of NNO, the red shifts induced in the HF stretch are virtually identical. When HF is replaced
by HC1, the equilibrium geometry is of parallel structure, with little in the way of H-bond-
ing character. Although HF can form what appear to be weak H-bonds with a molecule like
OCO, the same is not true of HOH. The complex pairing HOH with OCO has the O atom
of the former molecule approaching the C atom of the latter, held together in part by a di-
pole-quadrupole interaction. OSO is too weak a proton acceptor to form a H-bond with a
donor such as HCN.

6.7.2 C-H Donors


Geometric data extracted from crystal studies point toward the tendency of C—H bonds to
orient themselves in structures consistent with the existence of H-bonds. Careful examina-
tion reveals most of these to contain intermolecular contacts considerably longer than usu-
ally considered for H-bonds; moreover, the C—H bond stretches by only very small
amounts as a result of this contact.
An exception arises when the C atom is involved in a triple bond, as in HC=N. The
much more electronegative character of this carbon makes for a potent proton donor. Less
acidic, but still possessing some proton donating ability, are the alkynes. However, the lack
of a lone electron pair in a molecule like HC=CH prevents the formation of a H-bond in
the acetylene dimer. The most stable geometry has the hydrogen of one atom approach the
rich electron cloud around the central girth of the other, forming a T-shape. One would prob-
ably not consider this a H-bond due both to its length and to its weakness, less than 2
kcal/mol. Its stability can be rationalized on purely electrostatic grounds (quadrupole-
quadrupole), without recourse to other factors.
Pairing an alkyne with a more conventional proton acceptor does appear to form true H-
bonds in certain cases. Using the red shift of the C—H stretch as a barometer of H-bond
strength, there is a continuum that depends upon the strengths of the acid and the base.
Weak Interactions, Ionic H-Bonds, and Ion Pairs 343

CH3CN is a weak enough base that when paired with R—CCH, the red shift is less than
100 cm-1 regardless of the nature of R. On the other end of the spectrum is the stronger
base (CH3)3N, for which the red shifts vary from 140 to nearly 300 cm-1 as the changing
character of R makes the alkyne a stronger acid. When the C atom is involved in a double
rather than a triple bond, its electronegativity drops, as does its ability to donate a proton.
It seems unlikely that the C—H group of HCOOH, for example, acts as proton donor in a
true H-bond.
The single bonds of the alkyl group make these C—H bonds particularly reluctant to do-
nate protons in a H-bond. Methane, for example, does not engage in H-bonding, neither as
proton donor nor as acceptor. Its acidity can be increased the requisite amount by substitu-
tion of halogen atoms. CF3H and CC13H, for instance, are both capable of forming a H-
bond with a suitable acceptor. In fact, it is probably only necessary that one halogen atom
be present, as in CH3C1. Another means to overcome methane's inability to engage in H-
bonding is to "cheat" and pair it with an anion. The CH4"-X~ complex adopts a structure
with a linear CH---X arrangement. The binding energies vary from 6 kcal/mol when X=F
to about 1 kcal/mol for CH4--I". The signs of H-bonding also include a stretch of the C—H
bond by 0.01 A for CH4"-F~, whereas the other halides induce very little elongation. Con-
sistent with a picture of H-bonding for the fluoride, and a simpler electrostatic interaction
for the other halogens, is the red shift of the C—H frequency of 105 cm-1 in the former
case compared to less than 20 cm-1 for the others.

6.7.3 Ionic H-Bonds


If one of the partners is charged, the electrostatic attraction will be amplified, as will other
factors such as polarization effects. The interaction between HF and a F anion, certainly
one of the strongest H-bonds, amounts to more than 40 kcal/mol. Ionic H-bonds can be-
come so strong and so short, and the bridging proton stretched so far from the donor atom,
that it approaches the midpoint between the two partners. The (F H F)" system is a case
in point, with an interfluorine distance of less than 2.3 A. Other ionic systems, for exam-
ple, (C1"H"C1)~ and (HO-H-OH)", also contain short, centrosymmetric H-bonds. In some
other related systems, the barrier for transfer between the two minima in the proton trans-
fer potential is so low that the wave function corresponding to the first vibrational level
would likely correspond to a centrally located proton. It is a common observation that elec-
tron correlation lowers the barrier to proton transfer in these ionic H-bonds, or even elimi-
nates it entirely, converting a double-well potential to one with a symmetric single well. The
potentials for proton displacement in these systems tend to be very flat, with high-ampli-
tude proton motion. Enlargement of these anionic systems by alkyl substitution has little ef-
fect upon the energetics of binding or proton transfer.
There are certain features of these systems which differ from conventional H-bonds be-
tween neutral molecules. For one thing, the binding energy can be several times larger than
even the strongest neutral H-bond. This interaction is dominated by electrostatics, with rel-
atively minor contributions from other attractive forces. And in some cases, a proton donor
cation (e.g., NH 4 + ) can be replaced by another cation, one without a proton at all (K + ), and
the energetics of complexation are barely altered1. The hydrogen nucleus drifts so far from
the original donor atom that the distinction between a donor and acceptor group becomes a
murky one. Nonetheless, it would probably be best to retain their categorization as H-bonds
based largely upon their equilibrium geometry which maintains a hydrogen nucleus as the
glue which cements the entire complex together.
344 Hydrogen Bonding

The triply bonded nitrile and alkyne groups also engage in ionic H-bonds. Unlike the
case when the atoms involved are F or O, the H-bonds in which the C or N atoms of these
species participate tend to be noncentrosymmetric, that is, the proton transfer potential con-
tains a pair of minima. The exception is (C=NH--N=C) which is strongly bound by some
27 kcal/mol and which appears to contain a centrosymmetric H-bond. The H-bonds are all
weaker when the triply bonded species are replaced by their double-bonded analogs. For
example, whereas R(C--C) is equal to 3.35 A for (HC=CH"C=CH)~, with a dissociation
energy of 11 kcal/mol, the bond stretches to 3.7 A for (H2C=CH2"CH=CH2)~ and the H-
bond weakens to 5 kcal/mol.
The pairing of a neutral HCOOH molecule with HCOO leads to an interaction of about
33 kcal/mol. The primary interaction, the OH--O H-bond, is centrosymmetric with an in-
teroxygen separation of 2.5 A. There is perhaps an additional, much weaker interaction be-
tween the C—H group of the neutral and the other O atom of the anion. It appears from
studies of numerous crystals that centrosymmetric H-bonds occur when R(O-O) is less than
2.5 A; the transfer potential contains two minima for longer separations. Crystal forces can
influence the single or double-well character of the transfer potential, specifically by
stretching the H-bond, and asymmetries of the surroundings can shift the proton's equilib-
rium position. Immersion of this system in aqueous solvent changes the nature of the pro-
ton transfer potential to double-well.
One of the strongest interactions occurs between H2SO4 and its conjugate base, HSO4~.
The basis of this interaction energy, computed to be nearly 50 kcal/mol, lies in the three
OH"O H-bonds which bind the two species together. It appears likely that the proton trans-
fer potential contains a single well.
Just as a neutral proton donor can strongly associate with an anionic acceptor, one might
expect that a neutral acceptor can accept a proton from a cation to form a strong bond.
(HFH--FH)+, for example, is indeed strongly bound, with an interaction energy of 32
kcal/mol. While not quite as strong as its anionic analog, (FH--F)~, the interfluorine dis-
tance is comparably short, less than 2.3 A. The cation contains a centrosymmetric H-bond,
just as does the anion. Whereas the anionic (HO-H--OH)" complex was considerably less
strongly bound than (FH--F)~, the same is not true in the cations where (H2OH"OH2)+
contains a H-bond of comparable strength as is present in (HFH—FH) + . In a related per-
spective, the anionic interfluorine H-bond is stronger than the cation (also true of chlorine),
while the reverse is true for oxygen where the cationic H-bond is stronger then the anion,
bound by about 33 kcal/mol. Any barrier in the proton transfer potential of (H2OH--OH2)+
is so small as to be effectively nonexistent. In contrast to the interfluorine and interoxygen
ionic H-bonds, the internitrogen distance in (H3NH"NH3)+ is considerably longer, 2.73 A.
Its proton transfer potential contains two distinct minima, although the barrier is probably
only on the order of 1 kcal/mol. The interaction has the characteristic strength of ionic H-
bonds, with a binding energy of 25 kcal/mol.
The H-bonds in second-row analogs of the foregoing systems are weaker, but still rep-
resent interactions that are stronger than neutral H-bonds. The binding enthalpies of
(HC1H-C1H)+, (H2SH-SH2)+, and (H3PH--PH3)+ are about 15, 13, and 8 kcal/mol, re-
spectively. The first system contains a centrosymmetric H-bond whereas the others are char-
acterized by a double-well proton transfer potential.
Regardless of whether the ionic H-bond is of the cationic or anionic type, the H-bond
elongates as the electronegativity of the atoms involved diminishes. The shortest is the in-
terfluorine distance of 2.3 A in (HFH---FH) + and (FH"-F)~; the binding energy in the lat-
ter is in excess of 40 kcal/mol. Ionic H-bonds involving F, Cl, and O are centrosymmetric;
Weak Interactions, Ionic H-Bonds, and Ion Pairs 345

S is on the borderline, while N and P bonds are characterized by double-well potentials.


Methyl substitution has little effect upon the nature of these ionic H-bonds.
One can rationalize a simple relationship between the strength of an ionic H-bond
(A---H + B) on one hand and the difference in proton affinity between the two partners A and
E! on the other. The Marcus formulation provides a convenient framework for predicting the
H-bond energy of an arbitrary system based on knowledge of the interaction energy in a
symmetric system (A=B), and the difference in proton affinity between the two partners.
This model has proven successful in a series of interoxygen and internitrogen H-bonds.
One issue that has generated a good deal of interest over the years, in part for its poten-
tial importance in enzymatic activity, has been the competition for a proton donor between
the two lone pairs (syn and anti) of the O atom of the carboxylate group. When this group
is paired with a water molecule as donor, the preferred geometry belongs to the C2v point
group, wherein there are two H-bonds present: each hydrogen of the water acts as a bridge
to one of the carboxylate O atoms. But there is little energetic difference observed between
the approach of the water to the syn and anti sides of a single oxygen. This conclusion, of
nearly equal H-bonding potential of the syn and anti lone pairs, is confirmed by crystal stud-
ies which attempt to subtract out crystal packing forces and steric interference.
In contrast to this equal propensity toward formation of a H-bond, there is a definite pref-
erence for a proton to locate on the syn side when coming much closer and forming a co-
valent bond. That is, the syn conformer of RCOOH is more stable than the anti structure by
about 5 kcal/mol in the gas phase. If this is the case, why then is there so little apparent pKa
difference between the syn and anti structures? The answer to this question resides in sol-
vation phenomena. The presence of the polarizable medium preferentially stabilizes the anti
conformer of the carboxylic acid due to its higher dipole moment.
But even if we restrict our attention to the gas phase, one might wonder about a contra-
diction in that the syn conformer of RCOOH is clearly favored over anti, but that there is
no such preference when RCOO" forms a H-bond with a proton donor. This puzzle can be
resolved if one recalls that there are different forces in operation for a covalent bond as op-
posed to a H-bond. That is, the syn structure of RCOOH is indeed more stable than anti,
due to internal forces within this molecule. But as the proton acceptor is brought into prox-
imity with the carboxyl group, its partial negative charge will be repelled by the partial neg-
ative charge of the other O atom, the one with which it is not interacting directly. This re-
pulsion would act to destabilize the syn H-bond, as compared to the anti which would not
have such a force in operation.

6.7.4 Neutral Versus Ion Pairs


Starting with a conventional H-bond between a pair of neutral molecules, a proton transfer
from donor to acceptor will alter the fundamental character of the interaction to a pair of
oppositely charged ions. Such a transfer is disfavored by the energetic cost of generating
the high degree of separation of charge in the ion pair, so will occur only for a particularly
strong acid paired with a strong base. Hydrogen halides comprise some of the strongest pro-
ton donors, and amines, particularly alkylated amines, represent a class of powerful accep-
tors. Consequently, binary systems containing these species have provided a testbed for no-
tions about neutral/ion pair transitions and proton transfers.
Calculations have underscored the dominating importance of electron correlation in un-
derstanding the behavior of these systems. Whereas potential energy surfaces computed at
the Hartree-Fock level typically indicate the presence of two distinct minima, one corre-
346 Hydrogen Bonding

sponding to the neutral and the other to the ion pair, inclusion of correlation causes the co-
alescence of these two minima into a single one, in which the proton is located at an inter-
mediate position along the H-bond axis. Correlated calculations lead to the general con-
clusion that many surfaces contain a single, broad minimum. Increasing the basicity of the
acceptor and/or the acidity of the donor shifts the position of the minimum. The transition
from neutral to ion pair is best described as a gradual one as the nature of the donor and ac-
ceptor are changed.
Starting with a highly acidic donor, HI appears to donate a proton to amines that have
one or more methyl substituent, forming what can best be characterized as an ion pair. When
paired with NH3, however, the surface contains not a well-defined minimum, but rather a
long shallow valley connecting H 3 NH + ---~I with H 3 N---HI. HBr, too, forms ion pairs with
the more basic amines. The proton is nearly equally shared in a single minimum located be-
tween Br~ and CH3NH2, whereas the H 3 N---HBr complex is clearly a neutral pair. Turning
next to the still less acidic HC1, neither NH3 nor CH3NH2 is capable of extracting its pro-
ton, while (CH3)3N forms the ion pair. (The situation with (CH3)2NH remains not fully re-
solved.) HF is too weakly acidic to form an ion pair with any of these amines.
It is possible to quantify these relationships between the degree of proton transfer from
acid to base and the relative proton affinities of the two partners. This treatment illustrates
the unlikelihood that HF will form an ion pair with any acceptor, no matter how basic. HC1
is more susceptible to the nature of the acceptor: its proton can be pulled away from the Cl
as the basicity of the acceptor is raised. HBr is more sensitive still, as the nature of its com-
plexes with the amines take on more and more ion pair character for the more basic amines.
The crossover point, at which the equilibrium position of the proton is about midway be-
tween the donor and acceptor, does not occur when the two have equal proton affinities: the
proton affinity of the acceptor can be significantly less than that of the donor. The reason
for this observation resides in the nature of the force of interaction. Whereas the neutral pair
is held together by the "standard" components of H-bonding which amount to no more than
15 kcal/mol, the electrostatic attraction in the ion pair is greatly magnified by the opposite
charges of the two subunits. There is thus a natural energetic bias toward the ion pair, less-
ening the proton affinity of the acceptor that is required for this configuration to form.
Another means of shifting the equilibrium position of the proton is through the sur-
roundings. A polarizable medium preferentially stabilizes the ion pair, causing the proton
to translate toward the base. The same sensitivity of HX to basicity of the amine (i.e., HBr
loses its proton more readily than HC1 or HF as the amine becomes more basic) is noted as
the polarizability of the medium is enhanced.
The carboxyl group is an oxygen acid that is comparable in acidity to the hydrogen
halides. Like most of the latter, RCOOH will not donate a proton to form an ion pair with
a nitrogen base in the gas phase. But raising the dielectric constant of the medium does per-
mit this transfer to occur and at only fairly small enhancement of the medium's polariz-
ability over that of a vacuum.
Experimental measurements confirm many of the conclusions reached by the calcula-
tions. Combination of a hydrogen halide with an amine produces an ion pair in the gas phase
for only the strongest acid/base pairs. In concert with spectral information obtained in solid
matrix, the transition from neutral to ion pair appears to be a gradual one as the strengths
of the acid and base reach threshold values. There are a number of systems examined which
do not fall neatly into either the pure neutral or ion pair category, but are better described
as having a broad minimum in their proton transfer potential; the bottom of this well occurs
Weak Interactions, Ionic H-Bonds, and Ion Pairs 347

near the middle of the H-bond. Interaction with a matrix preferentially stabilizes the ion ver-
sus the neutral pair, leading to a higher degree of proton transfer than in the gas phase.

References
1. Liebman, J. E, Romm, M. J., Meot-Ner, M., Cybulski, S. M., and Schemer, S., Isotropy in ionic
interactions. 2. How spherical is the ammonium ion? Comparison of the gas-phase clustering
energies and condensed-phase thermochemistry of'K+ andNH4+, J. Phys. Chem. 95,1112-1119
(1991).
2. Scholtz, J. M., Qian, H., Robbins, V. H., and Baldwin, R. L., The energetics of ion-pair and hy-
drogen-bonding interactions in a helical peptide, Biochem. 32, 9668-9676 (1993).
3. Cans, P. J., Lyu, P. C., Manning, M. C., Woody, R. W., and Kallenbach, N. R., The helix-coil
transition in heterogeneous peptides with specific side-chain interactions: Theory and compar-
ison with CD spectral data, Biopolymers 31, 1605-1614 (1991).
4. Anderson, D. E., Becktel, W. J., and Dahlquist, F. W., pH-induced denaturation of proteins: A
single salt bridge contributes 3-5 kcal/mol to the free energy of folding of T4 lysozyme, Biochem.
2.9, 2403-2408 (1990).
5. Rendell, A. P. L., Bacskay, G. B., and Hush, N. S., An ab initio quantum chemical study of the
hydrogen- and "anti"-hydrogen-bonded HF/CIF and HF/C12 dimers, J. Chem. Phys. 87,
535-544(1987).
6. Ahlrichs, R., Scharf, P., and Erhardt, C., The coupled pair functional (CPF). A size consistent
modification of the CI(SD) based on an energy functional, J. Chem. Phys. 82, 890-898 (1985).
7. Hobza, P., Szczesniak, M. M., and Latajka, Z., HF—CIF: Minima on the 4-31G and4-3lG* en-
ergy hypersurfaces and thermodynamics of formation, Chem. Phys. Lett. 82, 469-472 (1981).
8. Rendell, A. P. L., Bacskay, G. B., and Hush, N. S., The validity of electrostatic predictions of the
shapes of van der Waals dimers, Chem. Phys. Lett. 117, 400-408 (1985).
9. Novick, S. E., Janda, K. C., and Klemperer, W., HFCIF: Structure and bonding, J. Chem. Phys.
65,5115-5121 (1976).
10. R0eggen, L, and Ahmadi, G. R., A quantum chemical study of the F- and H-bonded isomers of
HF/CIF, J. Mol. Struct. (Theochem) 307, 9-22 (1994).
11. Hunt, R. D., and Andrews, L., Infrared spectra of diatomic halogen complexes with hydrogen
fluoride in solid argon andneon, J. Phys. Chem. 92, 3769-3774 (1988).
12. Legon, A. C., Lister, D. G., and Thorn, J. C., Non-reactive interaction of ammonia and molecu-
lar chlorine: Rotational spectrum of the charge-transfer complex NH3~C12, J. Chem. Soc. Fara-
day Trans. 90, 3205-3212 (1994).
13. Bloemink, H. I., and Legon, A. C., The complex R3N---Br2 characterized in the gas phase by ro-
tational spectroscopy, J. Chem. Phys. 103, 876-882 (1995).
14. Bloemink, H. I., Hinds, K., Holloway, J. H., and Legon, A. C., Characterization of a pre-reac-
tive intermediate in gas-phase mixtures of fluorine and ammonia: The rotational spectrum of the
H3N-F2 complex, Chem. Phys. Lett. 245, 598-604 (1995).
15. Bloemink, H. I., Legon, A. C., and Thorn, J. C., Charge-transfer complexes of ammonia with
halogens: Nature of the binding in H3N"BrCl from its rotational spectrum, J. Chem. Soc. Fara-
day Trans. 91, 781-787 (1995).
16. Bloemink, H. L, Evans, C. M., Holloway, J. H., and Legon, A. C., Is the gas-phase complex of
ammonia and chlorine monofluoride H3N CIF or [H3 NCl]'"F~'? Evidence from rotational
spectroscopy, Chem. Phys. Lett. 248, 260-268 (1995).
17. Alberts, I. L., Handy, N. C., and Simandiras, E. D., The structure and harmonic vibrational fre-
quencies of the weakly bound complexes formed by HF with CO, CO2 and N2O, Theor. Chim.
Acta 74, 415-428(1988).
18. Curtiss, L. A., Pochatko, D. J., Reed, A. E., and Weinhold, F., Investigation of the differences in
stability of the OC-HF and CO-HF complexes, J. Chem. Phys. 82, 2679-2687 (1985).
348 Hydrogen Bonding

19. Benzel, M. A., and Dykstra, C. E., Conformational energetics in hydrogen-bonded dimers. The
unobserved CO—HF complex, Chem. Phys. 80, 273-278 (1983).
20. Schatte, G., Willner, H., Hoge, D., Knozinger, E., and Schrems, O., Selective trapping of the com-
plexes [OC—HF] and [CO'"HF] by photodissociation of matrix-isolated formyl fluoride, J.
Phys. Chem. 93, 6025-6028 (1989).
21. Reed, A. E., Weinhold, R, Curtiss, L. A., and Pochatko, D. J., Natural bond orbital analysis of
molecular interactions: Theoretical studies of binary complexes of HF, H2O, NH3, N2, O2, F2,
CO and CO2 with HF, H2O, and NHr J. Chem. Phys. 84, 5687-5705 (1986).
22. Yaron, D., Peterson, K. I., Zolandz, D., Klemperer, W., Lovas, F. J., and Suenram, R. D., Water
hydrogen bonding: The structure of the water-carbon monoxide complex, J. Chem. Phys. 92,
7095-7109(1990).
23. Sadlej, J., and Buch, V., Ab initio study of the intermolecular potential of the water-carbon
monoxide complex, J. Chem. Phys. 100, 4272-4283 (1994).
24. Lundell, J., A MPPT2 investigation of the H2O~CO dimer. A test of geometries, energetics, and
vibrational spectra, J. Phys. Chem. 99, 14290-14300 (1995).
25. Sadlej, J., Rowland, B., Devlin, J. P., and Buch, V., Vibrational spectra of water complexes with
H2, N2, and CO, J. Chem. Phys. 102, 4804-4818 (1995).
26. Lovejoy, C. M., Schuder, M. D., and Nesbitt, D. J., Direct IR laser absorption spectroscopy of
jet-cooled CO2HF complexes: Analysis of the v, HF stretch and a surprisingly low frequency v3
intermolecular CO2 bend, J. Chem. Phys. 86, 5337-5349 (1987).
27. Nesbitt, D. J., and Lovejoy, C. M., Rigid bender analysis of van der Waals complexes: The in-
termolecular bending potential of a hydrogen bond, }. Chem. Phys. 96, 5712-5725 (1992).
28. Sadlej, J., and Roos, B. O., A quantum chemical study of the hydrogen bonding in the CO2"HF
andN2O-HF complexes, Theor. Chim. Acta 76, 173-185 (1989).
29. Del Bene, J. E., and Frisch, M. J., An ab initio study of the structures and stabilities of the com-
plexes of the bases N2O, CO2, and CO with the acids FH, H+, and Li+, Int. J. Quantum Chem.,
Quantum Chem. Symp. 23, 371-380 (1989).
30. Fourati, N., Silvi, B., and Perchard, J. P., The carbon dioxide-hydrogen chloride complexes. A
matrix isolation study and an ab initio calculation on the 1-1 species, J. Chem. Phys. 81,
4737-4744(1984).
31. Sharpe, S. W., Zeng, Y. P., Wittig, C., and Beaudet, R. A., Infrared absorption spectroscopy of
CO2—HX complexes using the CO2 asymmetric stretch chromophore: CO2HF(DF) and
CO2HCl(DCl) linear and CO2HBr bent equilibrium geometries, J. Chem. Phys. 92, 943-958
(1990).
32. Altman, R. S., Marshall, M. D., and Klemperer, W., The microwave spectrum and molecular
structure of C02~HCl, J. Chem. Phys. 77, 4344-4349 (1982).
33. Zeng, Y. P., Sharpe, S. W., Shin, S. K., Wittig, C., and Beaudet, R. A., Infrared spectroscopy of
CO2-D(H)Br: Molecular structure and its reliability, J. Chem. Phys. 97, 5392-5402 (1992).
34. Muenter, J. S., Potential functions for carbon dioxide-hydrogen halide and hydrogen halide
dimer van der Waals complexes, J. Chem. Phys. 103, 1263-1273 (1995).
35. Rice, J. K., Lovas, F. J., Fraser, G. T., and Suenram, R. D., Pulsed-nozzle Fourier-transform mi-
crowave investigation of the large-amplitude motions in HBr—CO2, J. Chem. Phys. 103,
3877-3884 (1995).
36. Leopold, K. R., Fraser, G. T., and Klemperer, W., Rotational spectrum and structure of the com-
plex HCNC02, J. Chem. Phys. 80, 1039-1046 (1984).
37. De Almeida, W. B., Ab initio investigation of the stationary points on the potential energy sur-
face for the CO2-HCN binary complex, Chem. Phys. Lett. 166,589-598 (1990).
38. Merz, K. M. J., Gas-phase and solution-phase potential energy surfaces for CO2 + nH2O (n =
1,2), J. Am. Chem. Soc. 112, 7973-7980 (1990).
39. Peterson, K. J., and Klemperer, W., Structure and internal rotation of H2O~CO2, HI)O—CO2,
andD2O—CO2 van der Waals complexes, J. Chem. Phys. 80, 2439-2445 (1984).
40. Block, P. A., Marshall, M. D., Pedersen, L. G., and Miller, R. E., Wide amplitude motion in
Weak Interactions, Ionic H-Bonds, and Ion Pairs 349

the water-carbon dioxide and water-acetylene complexes, J. Chem. Phys. 96, 7321-7332
(1992).
41. Damewood, J. R., Jr., Kumpf, R. A., and Miihlbauer, W. C. F, Interaction of CO2 and water in-
vestigated by a combination of ab initio and SOLDRI-MM2 techniques, J. Phys. Chem. 93,
7640-7644 (1989),
42. Makarewicz, J., Ha, T.-K., and Bauder, A., Potential energy surface and large amplitude mo-
tions of the water-carbon dioxide complex, J. Chern. Phys. 99, 3694-3699 (1993).
43. Sadlej, J., and Mazurek, P., Ab initio calculations on the water-carbon dioxide system, J. Mol.
Struct. (Theochem) 337, 129-138 (1995).
44. Joyner, C. H., Dixon, T. A., Baiocchi, F. A., and Klemperer, W., The structure of'N2O~HF, J.
Chem. Phys. 74, 6550-6553 (1981).
45. Lovejoy, C. M., and Nesbitt, D. J., The near-infrared spectrum of ONNHF—Direct evidence for
geometric isomerism in a hydrogen bonded complex, J. Chem. Phys. 87, 1450 (1987).
46. Lovejoy, C. M., and Nesbitt, D. J., The infrared spectra of nitrous oxide-HF isomers, J. Chem.
Phys. 90, 4671-4680(1989).
47. Adamowicz, L.,NNO—HCl complex. Ab initio calculations with the coupled cluster method and
first-order correlation orbitals, Chem. Phys. Lett. 176, 249-254 (1991).
48. Zeng, Y. P., Sharpe, S. W., Reifschneider, D., Wittig, C., and Beaudet, R. A., Infrared absorption
spectroscopy of gas-phase N2O—HX (X=F,Cl,Br) weakly bonded complexes utilizing the N2O
v3 chromophore, J. Chem. Phys. 93, 183-195 (1990).
49. Pauley, D. J., Roehrig, M. A., Adamowicz, L., Shea, J. C., Haubrich, S. T., and Kukolich, S. G.,
Microwave measurements and theoretical calculations on the structures of NNO—HCl com-
plexes, J. Chem. Phys. 94, 899-907 (1991).
50. Gerber, S., and Huber, H., The sulfur dioxide-hydrogen fluoride complex. Additional informa-
tion to the experiment from ab initio calculations, Coll. Czech. Chem. Commun. 53, 1989-1994
(1988).
51. Chattopadhyay, S., and Plummer, P. L. M., Ab initio studies on the dimer of sulfur dioxide and
hydrogen cyanide, J. Chem. Phys. 93, 4187-4191 (1990).
52. Goodwin, E. .1., and Legon, A. C., The rotational spectrum and molecular geometry of an anti-
hydrogen-bonded dimer of sulfur dioxide and hydrogen cyanide, J. Chem. Phys. 85, 6828—6836
(1986).
53. Dayton, D. C., and Miller, R. E., Infrared spectroscopy of the SO2~HF and HCN~SO2 binary
complexes, J. Phys. Chem. 94, 6641-6646 (1990).
54. Plummer, P. L. M., Quantum mechanical studies of weakly bound molecular clusters, J. Mol.
Struct. (Theochem) 307, 119-133 (1994).
55. Matsumura, K., Lovas, F. J., and Suenram, R. D., The microwave spectrum and structure of the
H2O-S02 complex, J. Chem. Phys. 91, 5887-5894 (1989).
56. Bumgarner, R. E., Pauley, D. J., and Kukolich, S. G., Microwave spectra and structure for
SO2-H2S, SO2-HDS, and SO2-D2S complexes, J. Chem. Phys. 87, 3749-3752 (1987).
57. Andrews, A. M., Hillig, K. W. L, Kuczkowski, R. L., Legon, A. C., and Howard, N. W., Mi-
crowave spectrum, structure, dipole moment, and deuterium nuclear quadrupole coupling con-
stants of the acetylene-sulfur dioxide van der Waals complex, J. Chem. Phys. 94, 6947-6955
(1991).
58. Sun, L., Tan, X.-Q., Oh, J. J., and Kuczkowski, R. L., The microwave spectrum and structure of
the methanol-SO2 complex, J. Chem. Phys. 103, 6440-6449 (1995).
59. Pliego, J. R. J., and De Almeida, W. B., Searching for the ylide structure. An ab initio study of
the H2O-CC12 complex, Chem. Phys. Lett. 249, 136-140 (1996).
60. Glasstone, S., The structure of some molecular complexes in the liquid phase, Trans. Faraday
Soc. 33, 200-214 (1937).
61. Dippy, J. F., The dissociation constants of monocarboxylic acids; Their measurement and their
significance in theoretical organic chemistry, Chem. Rev. 25, 151-211 (1939).
62. Sutor, D. .1., The C—H~O hydrogen bond in crystals, Nature 195, 68-69 (1962).
350 Hydrogen Bonding

63. Green R. D., Hydrogen Bonding by C—H Groups; Wiley Interscience, New York (1974).
64. Taylor, R., and Kennard, O., Crystallographic evidence for the existence of C—H~O, C—H—N,
and C—H-Cl hydrogen bonds, J. Am. Chem. Soc. 104, 5063-5070 (1982).
65. Taylor, R., and Kennard, O., Hydrogen-bond geometry in organic crystals, Acc. Chem. Res. 17,
320-326 (1984).
66. Ceccarelli, C., Jeffrey, G. A., and Taylor, R., A survey of O—H—O hydrogen bond geometries
determined by neutron diffraction, J. Mol. Struct. 70, 255-271 (1981).
67. Jeffrey, G. A., and Saenger W., Hydrogen Bonding in Biological Structures; Springer-Verlag,
Berlin (1991).
68. Steiner, T., and Saenger, W., Geometry of C—H"'O hydrogen bonds in carbohydrate crystal
structures. Analysis of neutron diffraction data, J. Am. Chem. Soc. 114, 10146-10154 (1992).
69. Steiner, T., and Saenger, W., Role of C—H—O hydrogen bonds in the coordination of water mol-
ecules. Analysis of neutron diffraction data, J. Am. Chem. Soc. 115, 4540-4547 (1993).
70. Braga, D., Grepioni, R, Biradha, K., Pedireddi, V. R., and Desiraju, G. R., Hydrogen bonding in
organometallic crystals. 2. C—H—O hydrogen bonds in bridged and terminal first-row metal
carbonyls, J. Am. Chem. Soc. 117, 3156-3166 (1995).
71. Fraser, G. T., Leopold, K. R., and Klemperer, W., The structure of NH3-acetylene, J. Chem. Phys.
80, 1423-1426 (1984).
72. Peterson, K. I., and Klemperer, W., Water-hydrocarbon interactions: Rotational spectroscopy
and structure of the water-acetylene complex, J. Chem. Phys. 81, 3842-3845 (1984).
73. Peterson, K. I., Fraser, G. T, Nelson, D. D. J., and Klemperer, W., In Comparison of'Ab lnitio Quan-
tum Chemistry with Experiment for Small Molecules; Bartlettt, R. J., ed.; Reidell, (1985) 217-244.
74. DeLaat, A. M., and Ault, B. S., Infrared matrix isolation study of hydrogen bonds involving C~H
bonds: Alkynes with oxygen bases, J. Am. Chem. Soc. 109, 4232-4236 (1987).
75. Jeng, M.-L. H., DeLaat, A. M., andAult, B. S., Infrared matrix isolation study of hydrogen bonds
involving C~H bonds: Alkynes with nitrogen bases, J. Phys. Chem. 93, 3997-4000 (1989).
76. Jeng, M.-L. H., and Ault, B. S., Infrared matrix isolation study of hydrogen bonds involving
C—H bonds: Alkynes with bases containing second- and third-row donor atoms, J. Phys. Chem.
94, 13232-11327(1990).
77. Legon, A. C., Nonlinear hydrogen bonds O'"H—C and the role of secondary interactions. The
rotational spectrum of the oxirane"'acetylene complex, Chem. Phys. Lett. 247, 24-31 (1995).
78. Lee, R., and Henderson, G., A rigid rod normal coordinate analysis, and van der Waals force
field of acetylene-HCl, J. Chem. Phys. 81, 5521-5526 (1984).
79. Pendley, R. D., and Ewing, G. R, The infrared absorption spectra of(HCCH)2 and (DCCD)2, J.
Chem. Phys. 78, 3531-3540 (1983).
80. Sakai, K., Koide, A., and Kihara, T., Intermolecular forces for hydrogen, nitrogen and acetylene,
Chem. Phys. Lett. 47, 416-420 (1977).
81. Fischer, G., Miller, R. E., Vohralik, P. R, and Watts, R. O., Molecular beam infrared spectra of
dimers formed from acetylene, methyl acetylene, and ethene as a function of source pressure and
concentration, J. Chem. Phys. 83, 1471-1477 (1985).
82. Bryant, G. W., Eggers, D. F., and Watts, R. O., High-resolution infrared spectroscopy of acety-
lene clusters, J. Chem. Soc., Faraday Trans. 2 84, 1443-1455 (1988).
83. Ohshima, Y., Matsumoto, Y, Takami, M., and Kuchitsu, K., The structure and tunneling motion
of acetylene dimer studied by free-jet infrared absorption spectroscopy in the 14um region,
Chem. Phys. Lett. 147, 1-6 (1988).
84. Prichard, D. G., Nandi, R. N., and Muenter, J. S., Microwave and infrared studies of acetylene
dimer in a T-shaped configuration, J. Chem. Phys. 89, 115-123 (1988).
85. Fraser, G. T., Suenram, R. D., Lovas, F. J., Pine, A. S., Hougen, J. T., Lafferty, W. J., and Muenter,
J. S., Infrared and microwaveinvestigations of interconversion tunneling in the acetylene dimer,
J. Chem. Phys. 89, 6028-6045 (1988).
86. Muenter, J. S., An intermolecular potential function model applied to acetylene dimer, carbon
dioxide dimer, and carbon dioxide acetylene, J. Chem. Phys. 94, 2781-2793 (1991).
Weak Interactions, Ionic H-Bonds, and Ion Pairs 35 I

87. Craw, J. S., De Almeida, W. B., and Hinchliffe, A., Theoretical vibrational spectra of acetylene
dimer: "T" and cyclic configurations, J. Mol. Struct. (Theochem) 201, 69-74 (1989).
88. Aoyama, T., Matsuoka, O., and Nakagawa, N., Chem. Phys. Lett. 67, 508 (1979).
89. Alberts, I. L., Rowlands, T. W., and Handy, N. C., Stationary points on the potential energy sur-
faces of(C2H2)2, (C2H2)3, and (C2H4)2, J. Chem. Phys. 88, 3811-3816 (1988).
90. Bone, R. G. A., and Handy, N. C., Ab initio studies of internal rotation barriers and vibrational
frequencies of (C2H2)2, (CO2)2, and C2H2-CO2, Theor. Chim. Acta 78, 133-163 (1990).
91. Buckingham, A. D., and Fowler, P. W., A model for the geometries of Van der Waals complexes,
Can. J. Chem. 63, 2018-2025 (1985).
92. Buckingham, A. D., Permanent and induced molecular moments and long-range intermolecu-
lar forces, Adv. Chem. Phys. 12, 107-142 (1967).
93. Jeng, M.-L. H., and Ault, B. S., Infrared matrix isolation study of hydrogen bonds involving
C-H bonds: substituent effects, J. Phys. Chem. 93, 5426-5431 (1989).
94. McDonald, S. A., Johnson, G. L., Keelan, B. W., and Andrews, L., Infrared spectra of hydrogen-
bonded complexes between hydrogen halides and acetylene, J. Am. Chem. Soc. 102,
2892-2896(1980).
95. Legon, A. C., Aldrich, P. D., and Flygare, W. H., The rotational spectrum and molecular struc-
ture of the acetylene-HCl dimer, J. Chem. Phys. 75, 625-630 (1981).
96. Andrews, L., Johnson, G. L., and Kelsall, B. J., Infrared spectra of the hydrogen-bonded pi com-
plex C2H4—HF in solid argon, J. Chem. Phys. 76, 5767-5773 (1982).
97. Patten, K. O. J., and Andrews, L., Infrared spectra of diacetylene-hydrogen fluoride complexes
in solid argon, J. Phys. Chem. 90, 3910-3916 (1986).
98. Chandra, A. K., Pal, S., Limaye, A. C., and Gadre, S. R., Structure, energetics and bonding of
diacetylene complexes with hydrogen fluoride. A theoretical investigation, Chem. Phys. Lett.
247,95-100(1995).
99. Kisiel, Z., Fowler, P. W., Legon, A. C., Devanne, D., and Dixneuf, P., An investigation of hy-
drogen bonding between HCl and vinylacetylene: A molecule with two different --acceptor sites,
J. Chem. Phys. 93, 6249-6255 (1990).
100. Medhi, C., Majumdar, D., and Bhattacharyya, S. P., Theoretical studies on the topographical fea-
tures and energetics of diacetylene-hydrogen fluoride complexes, J. Mol. Struct. (Theochem)
280, 191-197 (1993).
101. Kollman, P., McKelvey, J., Johansson, A., and Rothenberg, S., Theoretical studies of hydrogen-
bonded dimers. Complexes involving HF, H2O, NH3, HCl, H , PH3, HCN, HNC, HCP, CH2NH,
H2CS, H2CO, CH4, CF3H, C2H2, C4H4, CgH6, F~, andH3O+, J. Am. Chem. Soc. 97, 955-965
(1975).
102. Fraser, G. T., Lovas, F. J., Suenram, R. D., Nelson, D. D., Jr., and Klemperer, W., Rotational
spectrum and structure ofCF3H—NHy J. Chem. Phys. 84, 5983-5988 (1986).
103. Hobza, P., Mulder, F., and Sandorfy, C., Quantum chemical and statistical thermodynamic in-
vestigations of anesthetic activity. 1. The interaction between chloroform, fluoroform, cyclo-
propane, and an O—H-O hydrogen bond, J. Am. Chem. Soc. 103, 1360-1366 (1981).
104. Hobza, P., Mulder, R, and Sandorfy, C., Quantum chemical and statistical thermodynamic in-
vestigations of anesthetic activity. 2. The interaction between chloroform, fluoroform, and a
N-H-0=C hydrogen bond, J. Am. Chem. Soc. 104, 925-928 (1982).
105. Hobza, P., and Sandorfy, C., Quantum chemical and statistical thermodynamic investigations of
anesthetic activity. 3. The interaction between CH4, CH3Cl, CH2C12, CHC13, CC14, and an
O—H-O hydrogen bond, Can. J. Chem. 62, 606-609 (1984).
106. Alkorta, L, and Maluendes, S., Theoretical study of CH—O hydrogen bonds in H2O—CHf,
H2O—CH2F2, andH2O~CHF3, J. Phys. Chem. 99, 6457-6460 (1995).
107. Del Bene, J. E., and Mettee, H. D., An ab initio study of the complexes of HF with the
chloromethanes, J. Phys. Chem. 97, 9650-9656 (1993).
108. Del Bene, J. E., and Shavitt, I., An ab initio study of the complexes of HCl with the
chloromethanes, J. Mol. Struct. (Theochem) 314, 9-17 (1994).
352 Hydrogen Bonding

109. Turi, L., and Dannenberg, J. J., Molecular orbital studies of the nitromethane-ammonia complex.
An unusually strong C—H-N hydrogen bond, J. Phys. Chem. 99, 639-641 (1995).
110. Latajka, Z., and Scheiner, S., Basis sets for molecular interactions. 2. Application to H3N—HF,
H3N—HOH, H2O-HF, (NH,)2, and H3CH-~OH2, J. Comput. Chem. 5, 674-682 (1987).
111. Woon, D. E., Zeng, P., and Beck, D. R., Ab initio potentials and pressure second virial coeffi-
cients for CH4-OH2 and CH4~SH2, J. Chem. Phys. 93, 7808-7812 (1990).
112. Szczesniak, M. M., Chalasinski, G., Cybulski, S. M., and Cieplak, P., Ab initio study of the po-
tential energy surface of CH4-H2O, J. Chem. Phys. 98, 3078-3089 (1993).
113. Novoa, J. J., Tarron, B., Whangbo, M.-H., and Williams, J. M., Interaction energies associated
with short intennolecular contacts of C—H bonds. Ab initio computational study of the C—H"O
contact interaction in CH4'OH2, J. Chem. Phys. 95, 5179-5186 (1991).
114. van Mourik, T., and van Duijneveldt, F. B., Ab initio calculations on the C—H~"O hydrogen-
bonded systems CH4—H2O, CH3N]H2-H2O, and CH3NH3+—H2O, J. Mol. Struct. (Theochem)
341, 63-73 (1995).
115. Sennikov, P. G., Sharibdjanov, R. I., and Khoudoinazarov, K., The molecular orbital study of the
structures and energies of bimolecular complexes of CH4 and SiH4 with water, J. Mol. Struct.
270,87-97(1992).
116. Iwaoka, M., and Tomoda, S., First observation of a C—H— Se "hydrogen bond", J. Am. Chem.
Soc. 116,4463-4464(1994).
117. Legon, A. C., Roberts, B. P., and Wallwork, A. L., Rotational spectra and geometries of the gas-
phase dimers (CH4,HF) and(CH4,HCl), Chem. Phys. Lett. 173, 107-114 (1990).
118. Ohshima, Y., and Endo, Y., Rotational spectrum and internal rotation of a methane-HCl com-
plex, J. Chem. Phys. 93, 6256-6265 (1990).
119. Govender, M. G., and Ford, T. A., Ab initio molecular orbital calculations of the properties of
the van der Waals complexes MH4HX(M=C, Si, X=F,Cl). Part 2. The infrared spectra, L Mol.
Struct. (Theochem) 338, 141-153 (1995).
120. Davis, S. R., and Andrews, L., Infrared spectra and ab initio SCF calculations for alkane-
hydrogen fluoride complexes, J. Am. Chem. Soc. 109, 4768-4775 (1987).
121. Legon, A. C., Wallwork, A. L., and Warner, H. E., Do methyl groups form hydrogen bonds? An
answer from the rotational spectrum of ethane-hydrogen cyanide, Chem. Phys. Lett. 191,97—101
(1992).
122. Mielke, Z., Tokhadze, K. G., Hulkiewicz, M., Schriver-Mazzuoli, L., Schriver, A., and Roux, E,
Infrared matrix isolation studies of methane-nitric acid system, J. Phys. Chem. 99, 10498-10505
(1995).
123. Novoa, J. J., Whangbo, M.-H., and Williams, J. M., Interaction energies associated with short
intennolecular contacts of C—H bonds. 4. Ab initio computational study of the C—H"anion in-
teractions in CH4-X~ (X=F,CI, Br, I), Chem. Phys. Lett. 180, 241-248 (1991).
124. Hon, F. H., and Tidwell, T. T., Steric crowding in organic chemistry. III. Spectral properties, con-
formations, and reactivities of highly substituted ferrocenylcarbinols, J. Org. Chem. 37,
1782-1786(1972).
125. Kazarian, S. G., Hamley, P. A., and Poliakoff, M., Intennolecular hydrogen-bonding to transi-
tion metal centres; Infrared spectroscopic evidence for O—H—lr bonding between [( 5-
C5Me5)Ir(CO)2] and fluoroalcohols in solution at room temperature, J. Chem. Soc., Chem.
Commun. 994-997 (1992).
126. Kazarian, S. G., Hamley, P. A., and Poliakoff, M., Is intermolecular hydrogen-bonding to un-
charged metal centers of organometallic compounds widespread in solution? A spectroscopic
investigation in hydrocarbon, noble gas, and supercritical fluid solutions of the interaction be-
tween fluoro alcohols and ( 5-C5RJML2 (R=H,Me; M= Co.Rh.Ir; L= CO, C 2 H 4 ,N 2 ,PMe 3 ) and
its relevance to protonation, L Am. Chem. Soc. 115, 9069-9079 (1993).
127. Shubina, E. S., Krylov, A. N., Kreindlin, A. Z., Rybinskaya, M. I., and Epstein, L. M., Inter-
molecular hydrogen bonds with d-electrons of transition metal atoms. H-complexes with metal-
locenes of the iron subgroup, J. Mol. Struct. 301, 1—5 (1993).
Weak Interactions, Ionic H-Bonds, and Ion Pairs 353

128. Epstein, L. M., Krylov, A. N., and Shubina, E. S., Novel types of hydrogen bonds involving tran-
sition metal atoms and proton transfer (XH-M, [MH]+-B, [MH]+-A~), J. Mol. Struct. 322,
345-352 (1994).
129. Brammer, L., Charnock, J. M., Goggin, P. L., Goodfellow, R. J., Orpen, A. G., and Koetzle, T. R,
The role of transition metal atoms as hydrogen bond acceptors: A neutron diffraction study of
[NPr"4]2[PtCl4]cis-[PtCl2(NH2Me)Jat20K, J. Chem. Soc. Dalton Trans. 1991, 1
(1991).
130. Calderazzo, R, Fachinetti, G., Marchetti, R, and Zanazzi, P. R, Preparation and crystal
and molecular structure of two trialkylamine adducts of HCo(CO)4 showing a preferential
NR3H+-[(OC)3Co(CO)r interaction, J. Chem. Soc., Chem. Commun. 1981,181-183 (1981).
131. Cecconi, R, Ghilardi, C. A., Innocenti, P., Mealli, C., Midollini, S., and Orlandini, A., Reductive
intramolecular hydrogen transfer in a d8 metal hydride promoted by CO addition. An experi-
mental study and its theoretical implications, Inorg. Chem. 23, 922-929 (1984).
132. Brammer, L., McCann, M. C., Bullock, R. M., McMullan, R. K., and Sherwood, P.,
Et3NH+Co(CO)4^: Hydrogen-bonded adduct or simple ion pair? Single-crystal neutron dif-
fraction study at 15 K, Organometallics 11, 2339-2341 (1992).
133. Lough, A. J., Park, S., Ramachandran, R., and Morris, R. H., Switching on and off a new in-
tramolecular hydrogen-hydrogen interaction and the heterocyclic splitting of dihydrogen. Crys-
tal and molecular structure of [Ir{H(r/>-SC 5H4NH)l2(PCy3)2]BF4-2.7CH2Cl2, J. Am. Chem.
Soc. 116, 8356-8357 (1994).
134. Lee, J. C., Rheingold, A. L., Muller, B., Pregosin, P. S., and Crabtree, R. H., Complexation of an
amide to indium via an iminol tautomer and evidence for an Ir-H—H—O hydrogen bond, J.
Chem. Soc., Chem. Commun. 1021-1022 (1994).
135. Shubina, E. S., Belkova, N. V., Krylov, A. N., Vorontsov, E. V., Epstein, L. M., Gusev, D. G.,
Niedermann, M., and Berke, H., Spectroscopic evidenceforintermolecularM—H—H—OR hy-
drogen bonding: Interaction of WH(CO)2(NO)L2 hydrides with acidic alcohols, J. Am. Chem.
Soc. 118, 1105-1112(1996).
136. Peris, E., Lee, J. C. J., Rambo, J. R., Eisenstein, O., and Crabtree, R. H., Factors affecting the
strength of'X—H-H—M hydrogen bonds, J. Am. Chem. Soc. 117, 3485-3491 (1995).
137. Peris, E., Lee, J. C., Jr., and Crabtree, R. H., Intramolecular N—H-X—Ir (X=H,F) hydrogen
bonding in metal complexes, J. Chem. Soc., Chem. Commun. 2573-2573 (1994).
138. Liu, Q., and Hoffmann, R., Theoretical aspects of a novel mode of hydrogen-hydrogen bonding,
J. Am. Chem. Soc. 117, 10108-10112(1995).
139. Richardson, T. B., de Gala, S., Crabtree, R. H., and Siegbahn, P. E. M., Unconventional hydrogen
bonds: InternwlecularB-H-H—N interactions, J. Am. Chem. Soc. 117, 12875-12876 (1995).
140. Sannigrahi, A. B., and Peyerimhoff, S. D., Ab initio SCF and CI calculations of the hydrogen
bonding energy of hydrogen bifluoride and bichloride ions, J. Mol. Struct. (Theochem) 122,
127-131 (1985).
141. Jenkins, H. D. B., and Pratt, K. F., Ion model independent studies of hydrogen bonding and the
thermochemistry of bifluoride salts, J. Chem. Soc., Faraday Trans. 2 73, 812-821 (1977).
142. Heni, M., and Illenberger, E., The stability of the bifluoride ion (HF2~) in the gas phase, J. Chem.
Phys. 83, 6056-6057 (1985).
143. Thomson, C., Clark, D. T., Waddington, T. C., and Jenkins, H. D. B., Theoretical studies on the
hydrogen dichloride, HC12~, ion and radical HC12~, J. Chem. Soc., Faraday Trans. 2 71,
1942-1947(1975).
144. Kawaguchi,K., and Hirota, E., Diode laser spectroscopy of the v, and v2 bands of FHF~ in 1300
cm"' region, J. Chem. Phys. 87, 6838-6841 (1987).
145. Janssen, C. L., Allen, W. D., Schaefer, H. R, and Bowman, J. M., The infrared spectrum of the
hydrogen bifluoride onion: Unprecedented variation with level of theory, Chem. Phys. Lett. 131,
352-358 (1986).
146. Kawaguchi, K., and Hirota, E., Infrared diode laser study of the hydrogen bifluoride onion:
FHF- and PDF', J. Chem. Phys. 84, 2953-2960 (1986).
354 Hydrogen Bonding

147. Spirko, V., Diercksen, G. H. R, Sadlej, A. J., and Urban, M., Vibrational spectrum ofFHF^from
high-level calculations of potential energy surfaces, Chem. Phys. Lett. 161, 519-526 (1989).
148. Epa, V. C., Choi, J. H., Klobukowski, M., and Thorson, W. R., Vibrational dynamics of the bi-
fluoride ion. I. Construction of a model potential surface, J. Chem. Phys. 92, 466-472 (1990).
149. Del Bene, J. E., Basis set and correlation effects on computed negative ion hydrogen bond en-
ergies of the dimers AHnAHn—l :AHn=NH3, OH2, and FH, International Journal of Quantum
Chemistry, Quantum Biology Symposia 14, 27-35 (1987).
150. Frisch, M. J., Del Bene, J. E., Binkley, J. S., and Schaefer, H. R, Extensive theoretical studies of
the hydrogen-bonded complexes (H2O)2, (H2O)2H+, (HF)2, (HF)2H+, F2H~, and (NH3)2, J.
Chem. Phys. 84, 2279-2289 (1986).
151. Gronert, S., Theoretical studies of proton transfers. 1. The potential energy surfaces of the iden-
tity reactions of the first- and second-row non-metal hydrides with their conjugate bases, J. Am.
Chem. Soc. 115, 10258-10266 (1993).
152. Larson, J. W., and McMahon, T. B., Gas-phase bihalide and pseudohalide ions. An ion cyclotron
resonance determination of hydrogen bond energies in XHY~ species (X,Y = F,Cl,Br,CN), In-
org. Chem. 23, 2029-2033 (1984).
153. Schroeder, L. W., and Ibers, J. A., Geometry of the bichloride ion. Preparation and crystal struc-
ture of cesium chloride 1/3 hydronium bichloride, J. Am. Chem. Soc. 88, 2601-2602 (1966).
154. Kawaguchi, K., Gas-phase infrared spectroscopy of ClHCl', J. Chem. Phys. 88, 4186^189
(1988).
155. Larson, J. W., and McMahon, T. B., Isotope effects in proton-transfer reactions. An ion cyclotron
resonance determination of the equilibrium deuterium isotope effect in the bichloride ion, J.
Phys. Chem. 91, 554-557 (1987).
156. Almlof, J., Ab initio MO-SCF calculations of equilibrium geometry and Vibrational structure for
the bichloride ion, HC12~, J. Mol. Struct. (Theochem) 85, 179-184 (1981).
157. Yamdagni, R., and Kebarle, P., Hydrogen-bonding energies to negative ions from gas-phase mea-
surements of ionic equilibria, J. Am. Chem. Soc. 93, 7139-7143 (1971).
158. Botschwina, P., Sebald, P., and Burmeister, R., Calculated spectroscopic properties for ClHCl~,
J. Chem. Phys. 88, 5245-5246 (1988).
159. Ikuta, S., Saitoh, T., and Nomura, O., The ClHCl~ union: Its chemical bond, vibrations, and free
energy, J. Chem. Phys. 91, 3539-3548 (1989).
160. Del Bene, J. E., An ab initio study of the structures and enthalpies of the hydrogen-bonded com-
plexes of the acids H2O, H2S, HCN, and HCl with the anions OH~, SH~, CN~, and Cr, Struct.
Chem. 1, 19-27 (1988).
161. Saitoh, T., Mori, K., Sasagane, K., and Itoh, R., Ab initio SCF-SDCI prediction of type II spec-
tra and geometry of (ClHCl)~ hydrogen bond complex. I. One dimensional vibrational analysis,
Bull. Chem. Soc. Jpn. 56, 2877-2888 (1983).
162. Yamdagni, R., and Kebarle, P., The hydrogen bond properties of'ClHCr and Cr(HCl)n, Can.
J. Chem. 52, 2449-2453 (1974).
163. French, M. A., Ikuta, S., and Kebarle, P., Hydrogen bonding of O—Hand C—H hydrogen donors
to Cl~. Results from mass spectrometric measurements of the ion-molecule equilibria RH + Cl~
= RHCr, Can. J. Chem. 60, 1907-1918 (1982).
164. Sannigrahi, A. B., and Peyerimhoff, S. D., Ab initio CI study of the hydrogen bibromide ion, J.
Mol. Struct. (Theochem) 165, 55-63 (1988).
165. McDaniel, D. H., and Vallee, R. E., Strong hydrogen bonds. I. The halide-hydrogen halide sys-
tems, Inorg. Chem. 2, 996-1001 (1963).
166. St0gard, A., Strich, A., Almlof, J., and Roos, B., Correlation effects on hydrogen-bond poten-
tials. SCF-CI calculations fro the systems HF2~ andH3O2^, Chem. Phys. 8, 405^11 (1975).
167. Roos, B. O., Kraemer, W. P., and Diercksen, G. H. F., SCF-CI studies of the equilibrium struc-
ture and the proton transfer barrier H 3 ) 2 , Theor. Chim. Acta 42, 77-82 (1976).
168. Szczesniak, M. M., and Scheiner, S., M ller-Plesset treatment of electron correlation in
(HOHO]H)-, J. Chem. Phys. 77, 4586^1593 (1982).
Weak Interactions, Ionic H-Bonds, and Ion Pairs 355

169. Rohlfing, C. M., Allen, L. C., Cook, C. M., and Schlegel, H. B., The structure of (H3O 2 )~, J.
Chem. Phys. 78, 2498-2503 (1983).
170. Xantheas, S. S., Theoretical study of hydroxide ion-water clusters, J. Am. Chem. Soc. 117,
10373-10380 (1995).
171. Kiippers, H., Takusagawa, T., and Koetzle, T. F., Neutron diffraction study of lithium hydrogen
phthalate monohydrate: A material with two very short intramolecular O—H—O hydrogen
bonds, J. Chem. Phys. 82, 5636-5647 (1985).
172. Flensburg, C., Larsen, S., and Stewart, R. P., Experimental charge density study of methylam-
monium hydrogen succinate monohydrate. A salt with a very short O—H—O hydrogen bond, J.
Phys. Chem. 99, 10130-10141 (1995).
173. Wozniak, K., He, H., Klinowski, J., Jones, W., and Barr, T. L., ESCA, solid-state NMR, andX-
ray diffraction monitor the hydrogen bonding in a complex of l,8-bis(dimethylamino)naphtha-
lene with 1,2-dichloromaleic acid, J. Phys. Chem,. 99, 14667-14677 (1995).
174. Perrin, C. L., and Thoburn, J. D., Symmetries of hydrogen bonds in monoanions of dicarboxylic
acids, J. Am. Chem. Soc. 114, 8559-8565 (1992).
175. Tunon, I., Rinaldi, D., Ruiz-Lopez, M. F., and Rivail, J. L., Hydroxide ion in liquidwater: Struc-
ture, energetics, and proton transfer using a mixed discrete-continuum ab initio model, J. Phys.
Chem. 99, 3798-3805 (1995).
176. Paul, G. J. C., and Kebarle, P., Thermodynamics of the association reactions of OH~ + H2O =
HOHOH- and CH3Q- + CH3OH = CH3OHOCH- in the gas phase, J. Phys. Chem. 94,
5184-5189(1990).
177. Meot-Ner, M., Comparative stabilities of cationic and anionic hydrogen-bonded networks.
Mixed clusters of water-methanol, J. Am. Chem. Soc. 108, 6189-6197 (1986).
178. Pudzianowski, A. T., MP2/6-311 + + G(d,p) study of ten ionic hydrogen-bonded binary systems:
Structures, normal modes, thermodynamics, and counterpoise energies, J. Chem. Phys. 102,
8029-8039 (1995).
179. Panich, A. M., NMR study of the F—H'"F hydrogen bond. Relation between hydrogen atom po-
sition and F—H-F bond length, Chem. Phys. 196, 511-519 (1995).
180. Wolfe, S., Hoz, S., Kim, C.-K., and Yang, K., Barrier widths, barrier heights, and the origins of
anomalous kinetic H/D isotope effects, J. Am. Chem. Soc. 112, 4186-4191 (1990).
181. Barlow, S. E., Dang, T. T., and Bierbaum, V. M., Hydrogen-deuterium-exchange reactions of
methoxide-methanol clusters, J. Am. Chem. Soc. 112, 6832-6838 (1990).
182. Meot-Ner, M., and Sieck, L. W., Relative acidities of water and methanol and the stabilities of
the dimer unions, J. Phys. Chem. 90, 6687-6690 (1986).
183. Ellenberger, M. R., Farneth, W. E., and Dixon, D. A., Gas-phase isotope fractionation factor for
the proton-bound dimer of the ethoxide anion, J. Phys. Chem. 85, 4-7 (1981).
184. Meot-Ner, M., and Sieck, L. W., The ionic hydrogen bond and ion salvation. 5. OH—O~ bonds.
Gas phase solvation and clustering ofalkoxide and carboxylate unions, J. Am. Chem. Soc. 108,
7525-7529(1986).
185. Cybulski, S., and Scheiner, S., Hydrogen bonding and proton transfers involving triply bonded
atoms. HC=NandHC=CH, J. Am. Chem. Soc. 109,4199-4206 (1987).
186. Scheiner, S., and Wang, L., Effect of bond multiplicity upon hydrogen bonding and proton trans-
fers. Double bonded atoms, J. Am. Chem. Soc. 114, 3650-3655 (1992).
187. Basch, H., and Stevens, W. J., The strong hydrogen bond in the formic acid-formate anion sys-
tem, J. Am. Chem. Soc. 113, 95-101 (1991).
188. Nusser, T., Turi, L., Naray-Szabo, G., and Simon, K., Transferability of the —COOH--QOC—
dyad. Geometry from the gas phase to crystals and proteins, Theor. Chim. Acta 90, 41-50
(1995).
189. Gejji, S. P., Taurian, O. E., and Lunell, S., Theoretical study of the short asymmetric [O---H—O]
hydrogen bond in solid potassium hydrogen diformate, including electron correlation, J. Phys.
Chem. 94, 4449-4452 (1990).
190. Spinner, E., Vibrational spectrum of solid sodium hydrogen bis(formate). An unsymmetrical very
356 Hydrogen Bonding

short OH"'O bond in a complex containing spectroscopically distinct HCO2~ and HCO2H en-
tities, J. Am. Chem. Soc. 105, 756-761 (1983).
191. Fenn, M. D., and Spinner, E., Proton and deuteron magnetic resonances of the strongly hydro-
gen-bonded complexes (HCO2)H~ and (HCO2)D ' in aqueous solution. The primary isotope ef-
fect on chemical shift, and isotopic fractionation of labile hydrogen in H—D mixtures, J. Phys.
Chem. 88, 3993-3997 (1984).
192. Perrin, C. L., Symmetries of hydrogen bonds in solution, Science 266, 1665-1668 (1994).
193. Evleth, E. M., Theoretical characterization of the triply H-bonded complex formed from HSO4~
andH2SO4, J. Mol. Struct. (Theochem) 307, 179-185 (1994).
194. Diercksen, G. H. R, von Niessen, W., and Kraemer, W. P., SCF LCGO MO studies on thefluo-
roniwn ion FH2+ and its hydrogen bonding interaction with hydrogen fluoride FH, Theor. Chim.
Acta 31, 205-214 (1973).
195. Del Bene, J. E., Frisch, M. J., and Pople, J. A., Molecular orbital study of the complexes
(AHn)2H+ formed from NH3, OH2, FH, PH3, SH2, and CIH, J. Phys. Chem. 89, 3669-3674
(1985).
196. Kraemer, W. P., and Diercksen, G. H. F., SCF MO LCGO studies on hydrogen bonding. The sys-
tem (H2OHOH2)+, Chem. Phys. Lett. 5, 463-465 (1970).
197. Newton, M. D., and Ehrenson, S., Ab initio studies on the structures and energetics of inner- and
outer-shell hydrates of the proton and hydroxide ion, J. Am. Chem. Soc. 93,4971-4990 (1971).
198. Busch, J. H., and de la Vega, J. R., Symmetry and tunneling in proton transfer reactions. Proton
exchange between methyloxonium ion and methyl alcohol, methyl alcohol and methoxide ion,
hydronium ion and water, and water and hydroxyl ion, J. Am. Chem. Soc. 99,2397-2406 (1977).
199. Scheiner, S., Proton transfers in hydrogen bonded systems. Cationic oligomers of water, J. Am.
Chem. Soc. 103, 315-320 (1981).
200. Desmeules, P. J., and Allen, L. C., Strong, positive-ion hydrogen bonds: The binary complexes
formed from NH3, OH2, FH, PH3, SH2 and CIH, J. Chem. Phys. 72, 4731-4748 (1980).
201. Xie, Y., Remington, R. B., and Schaefer, H. P., The protonated water dimer: Extensive theoret-
ical studies of H5O2+, J. Chem. Phys. 101, 4878-4884 (1994).
202. Ojamae, L., Shavitt, L, and Singer, S. J., Potential energy surfaces and vibrational spectra of
H5O2+ and larger hydraledproton complexes, Int. J. Quantum Chem., Quantum Chem. Symp.
29,657-668(1995).
203. Meot-Ner, M., and Field, F. H., Stability, association, and dissociation in the cluster ions
HjS'-nH^, H30+-nH20, andH2S+-H2O, J. Am. Chem. Soc. 99, 998-1003 (1977).
204. Cunningham, A. J., Payzant, T. D., and Kebarle, P., A kinetic study of the proton hydrate
H+(H2O)n equilibria in the gas phase, J. Am. Chem. Soc. 94, 7627-7632 (1972).
205. Meot-Ner, M., and Speller, C. V., Filling of solvent shells about ions. 1. Thermochemical crite-
ria and the effects of isomeric clusters, J. Phys. Chem. 90, 6616-6624 (1986).
206. Edison, A. E., Markley, J. L., and Weinhold, F., Ab initio calculations of protium/deuterium frac-
tionation factors in O2H5+ clusters, J. Phys. Chem. 99, 8013-8016 (1995).
207. Evleth, E. M., Hamou-Tahra, Z. D., and Kassab, E., Theoretical analysis ofnonconventional hy-
drogen-bonded structures in ion-molecule complexes, J. Phys. Chem. 95, 1213—1220 (1991).
208. Merlet, P., Peyerimhoff, S. D., and Buenker, R. J., Ab initio study of the hydrogen bond in
[H 3 N—H-NH 3 ] + -, J. Am. Chem. Soc. 94, 8301-8308 (1972).
209. Scheiner, S., Proton transfers in hydrogen-bonded systems. 4. Cationic dimers of NH3 and OH2,
J. Phys. Chem. 86, 376-382 (1982).
210. Scheiner, S., and Harding, L. B., Proton transfers in hydrogen-bonded systems. 2. Electron cor-
relation effects in (N2H7r, J. Am. Chem. Soc. 103, 2169-2173 (1981).
211. Scheiner, S., Szczesniak, M. M., and Bigham, L. D., Ab initio study of proton transfers includ-
ing effects of electron correlation, Int. J. Quantum Chem. 23, 739-751 (1983).
212. Payzant, J. D., Cunningham, A. J., and Kebarle, P., Gas phase solvation of the ammonium ion by
NH3 and H2O and stabilities of mixed clusters H' (NH Ja(H2O)^, Can. J. Chem. 51, 3242-3249
(1973).
Weak Interactions, Ionic H-Bonds, and Ion Pairs 357

213. Ikuta, S., Electron correlation effect on the geometries and energetics: Proton-bound ammonia
dimer, (H3N—H—NH3)+, J. Chem. Phys. 87, 1900-1901 (1987).
214. Platts, J. A., and Laidig, K. E., A theoretical study of the proton-bound ammonia dimer, J. Phys.
Chem. 99, 6487-6492 (1995).
215. Bigham, L., Scheiner, S., Comparison of proton transfers between first and second row atoms:
(H2SHSH2)+ and (H2OHOH2)+, J. Chem. Phys. 82, 3316-3321 (1985).
216. Ikuta, S., Geometries and energetics in the proton-bound dimers: H+(PH3)2, H+(H2S)2, and
H+(HCl)2, J. Mol. Struct. (Theochem) 152, 89-100 (1987).
217. Hiraoka, K., and Kebarle, P., Gas phase ion equilibrium studies of the proton in hydrogen sul-
fide and hydrogen sulfide—water mixtures. Stabilities of the hydrogen bonded complexes:
H+(H2S)JH2O)r, Can. J. Chem. 55, 24-28 (1977).
218. Del Bene, J. E., and Shavitt, I, Comparison of methods for determining the correlation contri-
bution to hydrogen bond energies, Int. J. Quantum Chem., Quantum Chem. Symp. 23, 445-452
(1989).
219. Del Bene, J. E., and Shavitt, L, Basis-set effects on computed acid-base interaction energies us-
ing the Dunning correlation-consistent polarized split-valence basis sets, J. Mol. Struct. (Theo-
chem) 307, 27-34 (1994).
220. Deakyne,C.A.,In Molecular Structure and Energetics, vol. 4 VCH Publishers, (1987), 105-141.
221. Hillenbrand, E. A., and Scheiner, S., Effects of molecular charge and methyl substitution on pro-
ton transfers between oxygen atoms, J. Am. Chem. Soc. 106, 6266-6273 (1984).
222. Meot-Ner, M., Intermolecular forces in organic clusters, J. Am. Chem. Soc. 114, 3312-3322
(1992).
223. Rabold, A., Bauer, R., and Zundel, G., Structurally symmetric N+H~N <=* N-H+Nbonds. The
proton potential as a function of the pKa of the base. FTIR results and quantum chemical cal-
culations, J. Phys. Chem. 99, 1889-1895 (1995).
224. Chu, C.-H., and Ho, J.-J., Theoretical studies of proton transfer in (CHfHO-H-OCHCH^,
J. Phys. Chem. 99, 1151-1155 (1995).
225. Chu, C.-H., and Ho, J.-J., Ab initio study of ion transfer in (H2CO—H—OCH2)+and
(H2CO—Li-OCH2)+, Chem. Phys. Lett. 221, 523-530 (1994).
226. Chu, C.-H., and Ho, J.-J., Effect of formaldehyde substituents on potential energy profiles for
proton transfer in [ABCO~HOCXH]+, J. Phys. Chem. 99, 16590-16596 (1995).
227. Zhang, R., and Lifshits, C., Ab initio calculations of hydrogen-bonded carboxylic acid cluster
systems: Dimer evaporations, J. Phys. Chem. 100, 960-966 (1996).
228. Arduengo, A. J. L, Gamper, S. R, Tamm, M., Calabrese, J. C., Davidson, R, and Craig, H. A., A
bis(carbene)-proton complex: Structure of a C—H— C hydrogen bond, J. Am. Chem. Soc. 117,
572-573 (1995).
229. Brzezinski, B., and Zundel, G., Collective H+, Li+, and Na+ motions and cation polarizabili-
ties of the cation-bonded systems within 1,11,12,13,14-pentahydroxypentacene salts: A FTIR
study, J. Phys. Chem. 98, 2271-2274 (1994).
230. Zundel, G., Brzezinski, B., and Olejnik, J., On hydrogen and deuterium bonds as well as on Li+,
Na+, and Be+2 bonds: IR continua and cation polarizabilities, J. Mol. Struct. 300, 573-592
(1993).
231. Zundel, G., Proton polarizability and proton transfer processes in hydrogen bonds and cation
polarizabilities of other cation bonds—their importance to understand molecular processes in
electrochemistry and biology, Trends Phys. Chem. 3, 129-156 (1992).
232. Zundel, G., and Eckert, M., IR continua of hydrogen bonds and hydrogen-bonded systems, cal-
culated proton polarizabilities and line spectra, J. Mol. Struct. (Theochem) 200, 73-92 (1989).
233. Eckert, M., and Zundel, G., Proton polarizability, dipole moment, and proton transitions of an
AH—B ^>A~—H+B proton-transfer hydrogen bond as a function of an external electrical field:
An ab initio SCF treatment, J. Phys. Chem. 91, 5170-5177 (1987).
234. Scheiner, S., Proton transfers in hydrogen bonded systems. 6. Electronic redistributions in
(N2H7yl- and (OJAJ v, J. Chem. Phys. 75, 5791-5801 (1981).
358 Hydrogen Bonding

235. Schemer, S., Energetics and electronic rearrangements of proton transfer in (H}NHOH2)+, Int.
J. Quantum Chem. 23, 753-764 (1983).
236. Yamdagni, R., and Kebarle, P., Gas-phase basicities of amines. Hydrogen bonding in proton-
bound amine dimers and proton-induced cyclization of a,m-diamines, J. Am. Chem. Soc. 95,
3504-3510 (1973).
237. Caldwell, G., Rozeboom, M. D., Kiplinger, J. P., and Bartmess, J. E., Anion-alcohol hydrogen
bond strengths in the gas phase, J. Am. Chem. Soc. 106, 4660-4667 (1984).
238. Feng, W. Y., Ling, Y, andLifshitz, C., Reactivity of mixed and neat proton bound dimers of ace-
tonitrile and methyl acetate, J. Phys. Chem. 100, 35-39 (1996).
239. Meot-Ner, M., The ionic hydrogen bond and ion salvation. 1. NH+~O, NH+~N, and OH+-O
bonds. Correlations with proton affinity. Deviations due to structural effects, J. Am. Chem. Soc.
106, 1257-1264 (1984).
240. Meot-Ner, M., The ionic hydrogen bond and ion salvation. 2. Solvation of onium ions by one to
seven H2O molecules. Relations between monomolecular, specific, and bulk hydration, J. Am.
Chem. Soc. 106, 1265-1272 (1984).
241. Meot-Ner, M., The ionic hydrogen bond and ion solvation. 7. Interaction energies of carbanions
with solvent molecules, J. Am. Chem. Soc. 110, 3858-3862 (1988).
242. Meot-Ner, M., Models for strong interactions in proteins and enzymes. 2. Interactions of ions
with the peptide link and with imidazole, J. Am. Chem. Soc. 110,3075-3080(1988).
243. Meot-Ner, M., and Sieck, L. W., The ionic hydrogen bond. 4. SH+-O andNH+-S bonds. Cor-
relations with proton affinity. Mutual effects of weak and strong ligands in mixed clusters, J. Am.
Chem. Soc. 89, 5222-5225 (1985).
244. Speller, C. V., and Meot-Ner, M., The ionic hydrogen bond and ion salvation. 3. Bonds involv-
ing cyanides. Correlations with proton affinities, J. Phys. Chem. 89, 5217-5222 (1985).
245. Marcus, R. A., Chemical and electrochemical electron-transfer theory, Annu. Rev. Phys. Chem.
15,155-196(1964).
246. Marcus, R. A., Theoretical relation among rate constants, barriers, and Br0nsted slopes of chem-
ical reactions, J. Phys. Chem. 72, 891-899 (1968).
247. Kreevoy, M. M., and Oh, S.-W., Relation bet\veen rate and equilibrium constants for proton-
transfer reactions, J. Am. Chem. Soc. 95, 4805-4810 (1973).
248. Kresge, A. J., What makes proton transfer fast?, Acc. Chem. Res. 8, 354-360 (1975).
249. Albery, W. J., The application of the Marcus relation to reactions in solution, Annu. Rev. Phys.
Chem. 31,227-263(1980).
250. Dodd, J. A., and Brauman, J. L, Marcus theory applied to reactions with double-minimum po-
tential surfaces, J. Phys. Chem. 90, 3559-3562 (1986).
251. Murdoch, J. R., Rate-equilibria relationships and proton-transfer reactions, J. Am. Chem. Soc.
94,4410-4418(1972).
252. Murdoch, J. R., and Magnoli, D. E., Kinetic and thermodynamic control in group transfer reac-
tions, J. Am. Chem. Soc. 104, 3792-3800 (1982).
253. Magnoli, D. E., and Murdoch, J. R., Kinetic and thermodynamic contributions to energy barri-
ers and energy wells: Application to proton-bound dimers in gas-phase proton-transfer reac-
tions, J. Am. Chem. Soc. 103, 7465-7469 (1981).
254. McLennan, D. J., A simple geometric model for the Marcus theory of proton transfer, J. Chem.
Ed. 53, 348-351 (1976).
255. Scheiner, S., and Redfern, P., Quantum mechanical test of Marcus theory. Effects of alkylation
upon proton transfer, J. Phys. Chem. 90, 2969-2974 (1986).
256. Redfern, P., and Scheiner, S., Effects ofalkylation upon the proton affinities of nitrogen and oxy-
gen bases, J. Comput. Chem. 6, 168-172 (1985).
257. Johnston, H. S., and Parr, C., Activation energies from bond energies. I. Hydrogen transfer re-
actions, J. Am. Chem. Soc. 85, 2544-2551 (1963).
258. Miller, R. A., A theoretical relation for the position of the energy harrier between initial and fi-
nal states of chemical reactions, .1. Am. Chem. Soc. 100, 1984-1992 (1978).
Weak Interactions, Ionic H-Bonds, and Ion Pairs 359

259. Agmon, N., and Levine, R. D., Energy, entropy and the reaction coordinate: Thermodynamic-
like relations in chemical kinetics, Chem. Phys. Lett. 52, 197-201 (1977).
260. Bell, R. P., The Proton in Chemistry, Cornell University Press, Ithaca, NY (1959).
261. Basilevsky, M. V., Weinberg, N. N., and Zhulin, V. M., Estimating the positions of transition
states from experimental data, Int. J. Chem. Kinet. 11, 853-865 (1979).
262. Cao, H. Z., Allavena, M., Tapia, O., and Evleth, E. M., Theoretical analysis of proton transfers
in symmetric and asymmetric systems, J. Phys. Chem. 89, 1581-1592 (1985).
263. Siggel, M. R. R, Streitwieser, A., Jr., and Thomas, T. D., The role of resonance and inductive ef-
fects in the acidity of carboxylic acids, J. Am. Chem. Soc. 110, 8022-8028 (1988).
264. Candour, R. D., On the importance of orientation in general base catalysis by carboxylate,
Bioorg. Chem. 10, 169-176 (1981).
265. Gandour, R. D., Nabulsi, N. A. R., and Fronczek, F. R., Structural model of a short carboxyl-im-
idazole hydrogen bond with a nearly centrally located proton: Implications for the Asp-His dyad
in serine proteases, J. Am. Chem. Soc. 112, 7816-7817 (1990).
266. Lukovits, I., Karpfen, A., Lischka, H., and Schuster, P., Ab initio LCMO studies on the hydra-
tion of formate ion, Chem. Phys. Lett. 63, 151-154 (1979).
267. Alagona, G., Ohio, C, and Kollman, P., Bifurcated vs. linear hydrogen bonds: Dimethyl phos-
phate and formate union interactions with water, J. Am. Chem. Soc. 105, 5226-5230 (1983).
268. Gao, J., Garner, D. S., and Jorgensen, W. L., Ab initio studies of structures and binding energies
for anion-water complexes, J. Am. Chem. Soc. 108, 4784-4790 (1986).
269. Hermansson, K., Lie, G. C., andClementi, E.,An ab initio pair potential for the interaction be-
tween a water molecule and a formate ion, Theor. Chim. Acta 74, 1-10 (1988).
270. Cybulski, S. M., and Scheiner, S., Hydrogen bonding and proton transfers involving the car-
boxylate group, J. Am. Chem. Soc. 1ll, 23-31 (1989).
271. Li, Y, andHouk, K. N., Theoretical assessments of the basicity and nucleophilicity of carboxy-
late syn and anti lone pairs, J. Am. Chem. Soc. 111,4505-4507 (1989).
272. Carrell, C. J., Carrell, H. L., Erlebacher, J., and Glusker, J. P., Structural aspects of metal ion-
carboxylate interactions, J. Am. Chem. Soc. 110, 8651-8656 (1988).
273. Ippolito, J. A., Alexander, R. S., and Christianson, D. 'W., Hydrogen bond stereochemistry in pro-
tein structure and function, J. Mol. Biol. 215, 457-471 (1990).
274. GSrbitz, C. H., and Etter, M. C., Hydrogen bonds to carboxylate groups. Syn/anti distributions
andsteric effects, J. Am. Chem. Soc. 114, 627-631 (1992).
275. Hillenbrand, E. A., and Scheiner, S., Analysis of the principles governing proton-transfer reac-
tions. Carboxyl group, J. Am. Chem. Soc. 108, 7178-7186 (1986).
276. Wiberg, K. B., and Laidig, K. E., Barriers to rotation adjacent to double bonds. 3. The C — O
barrier in formic acid, methyl formate, acetic acid, and methyl acetate. The origin of ester and
amide "resonance", J. Am. Chem. Soc. 109, 5935-5943 (1987).
277. Hocking, W., The other rotamer of formic acid. cis-HCOOH, Z. Naturforsch. 31 A, 1113-1121
(1976).
278. Wiberg, K. B., and Laidig, K. E., Acidity of (Z)- amd (E)-methyl acetates: Relationship to Mel-
drum's acid, J. Am. Chem. Soc. 110, 1872-1874 (1988).
279. Wang, X., and Houk, K. N., Theoretical elucidation of the origin of the anomalously high acid-
ity ofMeldrum's acid, J. Am. Chem. Soc. 110, 1870-1872 (1988).
280. Tadayoni, B. M., Parris, K., and Rebek, J., Jr., Intramolecular catalysis of enolization: A probe
for stereoelectronic effects at carboxyl oxygen, J. Am. Chem. Soc. 1ll, 4503-4505 (1989).
281. Menger, F. M., and Ladika, M., Fast hydrolysis of an aliphatic amide at neutral pH and ambi-
ent temperature. A peptidase model, J. Am. Chem. Soc. 110, 6794-6796 (1988).
282. Allen, F. H., and Kirby, A. J., Stereoelectronic effects at carboxylate oxygen: Similar basicity of
the E and Z lone pairs in solution, J. Am. Chem. Soc. 113,8829-8831 (1991).
283. Zimmerman, S. C., and Cramer, K. D., Stereoelectronic effects at carboxylate: A syn oriented
model for the histidine-aspartate couple in enzymes, J. Am. Chem. Soc. 110, 5906—5908
(1988).
360 Hydrogen Bonding

284. Huff, J. B., Askew, B., Duff, R. J., and Rebek, J., Jr., Stereoelectronic effects and the active site
of the serine proteases, J. Am. Chem. Soc. 110, 5908-5909 (1988).
285. Cramer, K. D., and Zimmerman, S. C., Kinetic effect of a syn-oriented carboxylate on a proxi-
mate imidazole in catalysis: A model for the histidine-aspartate couple in enzymes, J. Am. Chem.
Soc. 112,3680-3682(1990).
286. Pen-in, C. L., Johnston, E. R., Lollo, C. P., and Kobrin, P. A., NMR studies of base-catalyzed
proton exchange in amides, J. Am. Chem. Soc. 103, 4691-4696 (1981).
287. Perrin, C. L., Lollo, C. P., and Hahn, C.-S., Imidate anions: Stereochemistry, equilibrium, nitro-
gen inversion, and comparison with proton exchange, J. Org. Chem. 50, 1405—1409 (1986).
288. Gao, J., and Pavelites, J. J., Aqueous basicity of the carboxylate lone pairs and the C—O bar-
rier in acetic acid: A combined quantum and statistical mechanical study, J. Am. Chem. Soc.
114, 1912-1914(1992).
289. Pranata, J., Relative basicities of carboxylate lone pairs in aqueous solution, J. Comput. Chem.
14,685-690(1993).
290. Hillenbrand, E. A., and Scheiner, S., Modification of pK values caused by change in H-bond
geometry, Proc. Nat. Acad. Sci., USA 82, 2741-2745 (1985).
291. Scheiner, S., and Hillenbrand, E. A., Comparison between proton transfers involving carbonyl
andhydroxyl oxygens, J. Phys. Chem. 89, 3053-3060 (1985).
292. Scheiner, S., Bent hydrogen bonds and proton transfer, Acc. Chem. Res. 27, 402-408 (1994).
293. Montzka, T. A., Swaminathan, S., and Firestone, R. A., Reversal of syn-anti preference for car-
boxylic acids along the reaction coordinate for proton transfer. Implications for intramolecular
catalysis, J. Phys. Chem. 98, 13171-13176 (1994).
294. Marqusee, S., and Baldwin, R. L., Helix stabilization by Glu~-"Lys+ salt bridges in short pep-
tides ofde novo design, Proc. Nat. Acad. Sci., USA 84, 8898-8902 (1987).
295. Clementi, E., Study of the electronic structure of molecules. II. Wavefunctions for the NH3 + HCl
-> NH4Cl reaction, J. Chem. Phys. 46, 3851-3880 (1967).
296. Raffenetti, R. C., and Phillips, D. H., Gaseous NH4Cl revisited: A computational investigation
of the potential energy surface and properties, J. Chem. Phys. 71, 4534-4540 (1979).
297. Latajka, Z., and Scheiner, S., Influence of basis set on the calculated properties of (H3N—HCl),
J. Chem. Phys. 82, 4131-4134 (1985).
298. Chipot, C., Rinaldi, D., and Rivail, J.-L., Intramolecular electron correlation in the self-consis-
tent reaction field model of solvation. A MP2/6-3JG** ab initio study of the NH^—HCl com-
plex, Chem. Phys. Lett. 191, 287-292 (1992).
299. Latajka, Z., and Scheiner, S., Ab initio comparison of H bonds and Li bonds. Complexes of LiF,
LiCl, HF, and HCl with NH3, J. Chem. Phys. 81, 4014-4017 (1984).
300. Latajka, Z., and Scheiner, S., Ab initio study of FH—PH3 and CIH~PH3 including the effects
of electron correlation, J. Chem. Phys. 81, 2713-2716 (1984).
301. Alabart, J. R., and Caballol, R., The hydrogen bonding in H^Y-HX (Y=P, As; X=Cl, Br) com-
plexes: An ab initio study, Chem. Phys. Lett. 141, 334-338 (1987).
302. Hinchliffe, A., Ab initio study of the hydrogen-bonded complexes H3N'"HBr, H3P—HBr,
H 3 As-HF, H^As-HCl, and H3 As-HBr, J. Mol. Struct. (Theochem) 121, 201-205 (1985).
303. Brciz, A., Karpfen, A., Lischka, H., and Schuster, P., A candidate for an ion pair in the vapor
phase: Proton transfer in complexes R 3 N-HX, Chem. Phys. 89, 337-343 (1984).
304. Szczesniak, M. M., Latajka, Z., Ratajczak, H., and Orville-Thomas, W. J., Properties of strong
hydrogen-bonded systems. The formation of hydrogen-bonded ion pair in amine—HCl systems,
Chem. Phys. Lett. 72, 115-118 (1980).
305. Bacskay, G. B., and Craw, J. S., Quantum chemical study of the trimethylamine-hydrogen chlo-
ride complex, Chem. Phys. Lett. 221, 167-174 (1994).
306. Jasien, P. G., and Stevens, W. J., Theoretical studies of potential gas-phase charge-transfer com-
plexes: NH_, + HX (X=Cl,BrJ), Chem. Phys. Lett. 130, 127-131 (1986).
307. Latajka, Z., Scheiner, S., and Ratajczak, H., The proton position in hydrogen halide-amine com-
plexes. BrH~NH3 and BrH-NH2CHf Chem. Phys. Lett. 135, 367-372 (1987).
Weak Interactions, Ionic H-Bonds, and Ion Pairs 361

308. Latajka, Z., Scheiner, S., and Ratajczak, H., The proton position in amine-HX (X=Br,I) com-
plexes, Chem. Phys. 166, 85-96 (1992).
309. Kurnig, I. J., and Scheiner, S., Ab initio investigation of the structure of hydrogen halide-amine
complexes in the gas phase and in a polarizable medium, Int. J. Quantum Chem., QBS 14,47-56
(1987).
310. Del Bene, J. E., Person, W. B., and Szczepaniak, K., Ab initio theoretical and matrix, isolation
experimental studies of hydrogen bonding: Vibrational consequences of proton position in 1:1
complexes of HCl and 4-X-pyridines, Chem. Phys. Lett. 247, 89-94 (1995).
311. Ault, B. S., Steinback, E., and Pimentel, G. C., Matrix isolation studies of hydrogen bonding.
The vibrational correlation diagram, J. Phys. Chem. 79, 615-620 (1975).
312. Legon, A. C., and Rego, C. A., The extent of proton transfer from X toN in the gas-phase dimers
(CH3)3__nHN-HX. Evidence from rotational spectroscopy, Chem. Phys. Lett. 162, 369-375
(1989).
313. Deng, H., Zheng, J., Burgner, J., and Callender, R., Molecular properties of pyruvate bound to
lactate dehydrogenase: A Raman spectroscopic study, Proc. Nat. Acad. Sci., USA 86,
4484-4488 (1989).
314. Honig, B. H., and Hubbell, W. L., Stability of "salt bridges" in membrane proteins, Proc. Nat.
Acad. Sci., USA 81, 5412-5416 (1984).
315. Sapse, A. M., and Russell, C. S., Theoretical studies of the binding of methylamine and guani-
dine to carboxyiate, J. Mol. Struct. (Theochem) 137, 43-53 (1986).
316. Melo, A., and Ramos, M. J., Proton transfer in arginine-carboxylate interactions, Chem. Phys.
Lett. 245, 498-502 (1995).
317. Deerfield, D. W., Nicholas, H. B. J., Hiskey, R. G., and Pedersen, L. G., Salt or ion bridges in
biological systems: A study employing quantum and molecular mechanics, Proteins Struct. Func.
Genetics 6, 168-192 (1989).
318. Scheiner, S., and Duan, X., Effect of intermolecular orientation upon proton transfer within a
polarizable medium, Biophys. J. 60, 874-883 (1991).
319. Scheiner, S., and Kar, T., The nonexistence of specially stabilized hydrogen bonds in enzymes, J.
Am. Chem. Soc. 117, 6970-6975 (1995).
320. Fugler, L., Russell, C. S., and Sapse, A. M., Ab initio calculations on imine-carboxyl complexes,
L Phys. Chem. 91, 37-41 (1987).
321. Gilson, M. K., and Honig, B. H., The dielectric constant of a folded protein, Biopolymers 25,
2097-2119(1986).
322. Jain, D. C., Sapse, A. M., and Cowburn, D., Solvent effect on some imine-carboxyl complexes,
J. Phys. Chem. 92, 6847-6849 (1988).
323. Remko, M., Liedl, K. R., and Rode, B. M., Cation binding effect on hydrogen bonding and the
energetics of proton transfer in the system (CH3)3NH+'"~OCOH, J. Mol. Struct. (Theochem)
335,7-15(1995).
324. Geppert, S., Rabold, A., Zundel, G., and Eckert, M., Theoretical treatment of the spectroscopi-
cal data of a strong hydrogen bond with a broad single-minimum proton potential, J. Phys.
Chem. 99, 12220-12224 (1995).
325. Jensen, J. H., and Gordon, M. S., On the number of water molecules necessary to stabilize the
glycine zwitterion, J. Am. Chem. Soc. 117, 8159-8170 (1995).
326. Tortonda, F. R., Pascual-Ahuir, J.-L., Silla, E., and Tunon, L, Solvent effects on the thermody-
namics and kinetics of the proton transfer between hydronium ion and ammonia. A theoretical
study using the continuum and the discrete models, J. Phys. Chem. 99, 12525-12528 (1995).
327. Goodwin, E. J., and Legon, A. C., The rotational spectrum of '5N-ammonium chloride vapour:
characterisation of the hydrogen bonded dimer H3N'"HCl, Chem. Phys. Lett. 131, 319-324
(1986).
328. Howard, N. W., and Legon, A. C., An investigation of the hydrogen-bonded dimer H3N'"HBr by
pulsed-nozzle Fourier-transform microwave spectroscopy of ammonium bromide vapor, J.
Chem. Phys. 86, 6722-6730 (1987).
362 Hydrogen Bonding

329. Rego, C. A., Batten, R. C., and Legon, A. C., The properties of the hydrogen-bonded dimer
(CH3)3N—HCN from an investigation of its rotational spectrum, J. Chem. Phys. 89, 696-702
(1988).
330. Legon, A. C., and Rego, C. A., H,1VF nuclear-spin-nuclear-spin coupling in the rotational spec-
trum of(CH3)3N-HF and the lengthening of the HF bond, Chem. Phys. Lett. 154, 468-472
(1989).
331. Legon, A. C., and Rego, C. A., An investigation of the trimethylammonium chloride molecule in
the vapor phase by pulsed-nozzle, Fourier-transform microwave spectroscopy, J. Chem. Phys.
90, 6867-6876 (1989).
332. Legon, A. C., and Rego, C. A., Rotational spectrum of (CH3)3Pl"HCl and a comparison of prop-
erties within the series of axially symmetric dimers R3Y"HCl, where Y = N or P and R = H or
CH3, J. Chem. Soc. Faraday Trans. 86, 1915-1921 (1990).
333. Legon, A. C., Wallwork, A. L., and Rego, C. A., The rotational spectrum and nature of the het-
erodimer in trimethylammonium bromide vapor, J. Chem. Phys. 92, 6397-6407 (1990).
334. Legon, A. C., and Rego, C. A., Rotational spectrum of the trimethylamine-hydrogen iodide
dimer: An ion pair (CH3)3NH+-T in the gas phase, J. Chem. Phys. 99, 1463-1468 (1993).
335. Tobin, J. B., Whitt, S. A., Cassidy, C. S., and Frey, P. A., Low-barrier hydrogen bonding in mol-
ecular complexes analogous to histidine and aspartate in the catalytic triad of serine proteases,
Biochem. 34, 6919-6924 (1995).
336. Golubev, N. S., Smirnov, S. N., Gindin, V. A., Denisov, G. S., Benedict, H., and Limbach,
H.-H., Formation of charge relay chains between acetic acid and pyridine observed by low-
temperature nuclear magnetic resonance, J. Am. Chem. Soc. 116, 12055-12056 (1994).
337. Dega-Szafran, Z., and Dulewicz, E., Infrared and 'H nuclear magnetic resonance studies of hy-
drogen bonds in some pyridine trifluoroacetates and their deuteriated analogues in dichloro-
methane, J. Chem. Soc. Perkin Trans. II 345-351 (1983).
338. Zeegers-Huyskens, T., Isotopic ratio vHf/vHD of hydrogen fluoride complexes in solid argon,
Chem. Phys. Lett. 157, 105-107 (1989).
339. Dega-Szafran, Z., and Szafran, M., Complexes ofcarboxylic acids with pyridines and pyridine
N-oxides, Heterocycles 37, 627-659 (1994).
340. Johnson, G. L., and Andrews, L., Matrix infrared spectrum of the H3N—HF hydrogen-bonded
complex, J. Am. Chem. Soc. 104, 3043-3047 (1982).
341. Andrews, L., Fourier transform infrared spectra of HF complexes in solid argon, J. Phys. Chem.
88, 2940-2949 (1984).
342. Arlinghaus, R. T., and Andrews, L., Infrared spectra of the PH3, AsH3, and SbH 3 —HX hydro-
gen bonded complexes in solidargon, J. Chem. Phys. 81, 4341-4351 (1984).
343. Andrews, L., Davis, S. R., and Johnson, G. L., FTIR spectra of methyl-substituted amine-
hydrogen fluoride complexes in solidargon and nitrogen, J. Phys. Chem. 90,4273—4282 (1986).
344. Schriver, L., Schriver, A., and Perchard, J. P., Spectroscopic evidence for proton transfer within
the bimolecular complex HI~~NH3 trapped in cryogenic matrices, J. Am. Chem. Soc. 105,
3843-3848 (1983).
345. Barnes, A. J., Kuzniarski, J. N. S., and Mielke, Z., Strongly hydrogen-bonded molecular com-
plexes studied by matrix-isolation vibrational spectroscopy. Part 2. Amine-hydrogen chloride
complexes, J. Chem. Soc., Faraday Trans. 2 80,465-476 (1984).
346. Barnes, A. J., and Wright, M. P., Strongly hydrogen-bonded molecular complexes studied by ma-
trix-isolation vibrational spectroscopy. Part 3. Ammonia-hydrogen bromide and amine-hydro-
gen bromide complexes, J. Chem. Soc., Faraday Trans. 2 82, 153-164 (1986).
347. Dunker, A. K., and Marvin, D. A., A model for membrane transport through a-helical protein
pores, J. Theor. Biol. 72, 9-16 (1978).
348. Kayalar, C., A model for proton translocation in biomembranes based on keto-enol shifts in hy-
drogen bonded peptide groups, J. Membrane Biol. 45, 37-42 (1979).
349. Krimm, S., and Dwivedi, A. M., Infrared spectrum of the purple membrane: Clue to a proton
conduction mechanism?, Science 216, 407-408 (1982).
Weak Interactions, Ionic H-Bonds, and Ion Pairs 363

350. Tuchsen, E., and Woodward, C., Mechanism of surface peptide proton exchange in bovine pan-
creatic trypsin inhibitor. Salt effects and O-protonation, J. Mol. Biol. 185, 421-430 (1985).
351. Kearley, G. J., Fillaux, F, Baron, M.-H., Bennington, S., and Tomkinson, J.,A new look at proton
transfer dynamics along the hydrogen bonds in amides and peptides, Science 264, 1285—1289
(1994).
352. Fillaux, R, Fontaine, J. P., Baron, M.-H., Kearley, G. J., and Tomkinson, J., Inelastic neutron-
scattering study of the proton dynamics of N-methylacetamide at 20 K, Chem. Phys. 176,
249-278 (1993).
353. Schlegel, H. B., Gund, P., and Fluder, E. M., Tautomerization offormamide, 2-pyridone, and 4-
pyridine: An ab initio study, J. Am. Chem. Soc. 104, 5347-5351 (1982).
354. Johnson, S. W., Eckert, J., Barthes, M., McMullan, R. K., and Muller, M., Crystal structure of
acetanilide at 15 and 295 K by neutron diffraction. Lack of evidence for proton transfer along
the N—H-O hydrogen bond, J. Phys. Chem. 99, 16253-16260 (1995).
This page intentionally left blank
Index of Complexes

This index is provided so that the reader might locate information about a particular com-
plex of interest. It is organized as follows: The first section contains the neutral binary com-
plexes, followed by charged dimers, and then by larger complexes in the last section. Within
each section, the complexes are ordered by the type of molecules contained. The order is as
follows: XH, YH2, ZH3, carbonyl, carboxyl, imine, amide, nitrile, alkyne, alkene, alkane,
others. (X refers to any halogen: F, Cl, Br, I; Y to O, S, etc.; and Z to N, P, etc.) Any alky-
lated or similar substitutions follow immediately after their parent group.
Thus, any binary complex containing HF is listed first. Within the set of neutral binary
complexes containing HF as one partner, the complexes are listed in the order HF, HC1,
OH2, O(CH3)2, SH2, SeH2, NH3, NH2CH3, NH(CH 3 ) 2 ,... OCH2, OCHF, SCH2, HCOOH,
OCHNH2, NCH,.. . H2CCH2, CH4, CC1H3, and so on.
Each complex is listed only once. So, for example, the complex pairing HF with OH2 is
listed as FH-OHL, and not as HOH-FH.

Neutral Binary Complexes bending potential 64-66


vibrational spectrum 156-8, 171-3
FH- O(CH3)2 67
FH 291 SH2
decomposition 38 binding energy 62-63
energetics 72-77 bending potential 64-66
geometry 72-74, 209-13 SeH2 62
isotope effects 118-9 NH 3
superposition error 26, 27 binding energy 54
vibrational spectrum 143-7 BSSE 56-59
ZPVE 118-9 vibrational spectrum 148-52
CIH 74 anharmonicity 153
OH2 proton transfer 331, 334-5, 339
binding energy 62-64 NH2CH3 332,334-5

365
366 Index of Complexes

FH... (continued) SeH2 62


NH(CH3)2 332, 334-5 NH3
N(CH3)3 332, 334-5, 337, 339 binding energy 54
PH3 anharmonicity 152
binding energy 54 EFG 154
vibrational spectrum 148-52 proton transfer 225-6, 330-1, 334-5, 337,
proton transfer 331, 339 339
AsH3 54, 331 NH2CH3152, 225-6, 331-2, 334-5, 337, 339
OCH2 92, 183 NH(CH3)2 332, 334-5
OCHF 94 N(CH3)3 332, 334-5, 337, 339
OCHC1 94 PH3
SCH2 93 binding energy 54
HCOOH 94-96 EFG 154
HCONH2 93, 106 proton transfer 331, 339
NCH 102, 185-90 P(CH3)3 337
NCCH3 102 AsH3 54, 331
HCCCCH 301 OCH2 92-93, 182-4
H2CCH2 301 OC(CH3)2 93
CH4 305 OCHF 94
CC1H3 303 OCHCI 94
CC12H2 303 HCOOH 94-96
CC13H 303 NCCH3 102, 190-1
CO 294 HCCH 299
CO2 295 HCCCCH 301
N2O 296 CH4 305
SO2 297 CC1H3 303
C12 292-3 CC12H2 303
C1F 292-3 CC13H 303
HMn(CO)5 307 CO2 295
N2O 297
FD-NH3 153 2-butanone 93
FLi-NH3 153 methyl formate 93
methyl acetate 93
C1H- pyridine 155, 332-3, 339
C1H
energetics 74-77 BrH--
geometry 74, 213-5 BrH 76-77
isotope effects 119 IH 76-77
vibrational spectrum 146-7 OH2
BrH 76-77 binding energy 62-63
IH 76-77 bending potential 64-66, 209
OH2 vibrational spectrum 158
binding energy 62-63 SH2 62
bending potential 64-66 SeH2 62
vibrational spectrum 158 NH3
EFG 160 binding energy 54
OH(CH3) 67, 159 anharmonicity 152
0(CH3)2 67 proton transfer 225-6, 331-2, 334-5, 337,
SH2 339
binding energy 62-63 NH2CH3 60-61, 225-6, 331-2, 334-5, 337
bending potential 64-66, 209 NH(CH3)2 60-61, 225-6, 331-2, 334-5
EFG 160 N(CH3)3 60-61, 225-6, 331-2, 334-5, 337
Index of Complexes 367

PH3 54, 331 CF3H 302, 303


AsH3 54, 331 CC1H3 302-3
CO2 295 CC12H2 302-3
CC13H 302-3
IH- CC14 302-3
IH 76-77 SiH4 304
OH2 158,209 CO 294-5
NH3 225-6, 331-2, 338 CO2 296
NH2CH3 60-61, 225-6, 331-2 SO2 298
NH(CH3)2 60-61, 225-6, 331-2 CC12 298
N(CH3)3 60-61, 225-6, 331-2, 337
CH3OH-
HOH- OHCH3 81-83, 169
OH2 19-22 OHC6H5 83, 170
anharmonicity 15 OHSiH3 82-83
decomposition 33, 35 NHCH2 103, 185
energetics 78-79 NHCHCH3 185
geometry 77-78, 215-23 NCH3CH2 185
isotope effects 119-21 NCH3CHCH3 185
polarizability 162-3 SO2 298
superposition error 25-26, 27
total energy 24 C2H5OH-OHC2H5 81
vibrational spectrum 160-8, 173-5 C6H5OH-NH3 71, 177
OHCH3 81, 169-70 SiH3OH---OHSiH3 82-83, 169
OHC2H5 81
O(CH3)2 82 HSH-
OHC6H5 83 SH2 79-80, 163-6
OHSiH3 82 NH3 69-70
OHC1 83-84, 171 HCN 102
SH2 80-81 CH4 304
NH3 SO2 298
geometry 69-70
energetics 69-70 H3N-
isotope effects 119 NH3 292
vibrational spectrum 175-7 geometry 84-88, 208-9
NH2CH3 71 vibrational spectrum 177-9
PH3 69-70 ZPVE 178
OCH2 PH3 89
geometry 89-92, 223-5 HCN 102
isotope effects 119 CH3NO2 304
vibrational spectrum 180-2 2-pyridone 88-89
HCOOH 96-97 CIF 294
NHCH2 103, 185
NHCHCH3 185 CH3NH2-HCOOH 97
NCH3CH2 185 (CH3)2NH-NH(CH3)2 88
NCH3CHCH3 185 H3P-PH3 89
HCONH2 105-6, 119 H3P-HCN 102
CH3CONHCH3 106-9
HCCH 299 HCOOH-
CH4 302-4 NH2CH3 335
CFH3 303 N(CH3)3 337
CF2H2 303 HCOOH 99-101
368 Index of Complexes

HCOOH- (continued) (CH3)3NH+-N(CH3)3 320


CH3COOH 101 H3PH+-PH3 316, 318
NHCH2 336 CH3CHOH+-OCHCH3 320
HCONH2 112-3 HCOOH2+-HCOOH 320
CH3C1 97-99 HCNH2OH2+-OCHNH2 320
FCNH2OH2+-OCFNH2 320
CH3COOCH3-HCONH2 113 pyridinium+—pyridine 320
HCONH2-HCONH2 105, 110-1, 195-7 Li+-NH3 59-60
CH3CONHCH3-CH3CONHCH3 106-9 Na+-OCH2 180
K+-NH3 292
NCH- Mg+2-OCH2 180
NH3 191-2
N(CH3)3 337
OCH2 93 Anionic Complexes
NCH 102-3, 119, 192-5 FH-F- 308-11,319
CH4 305 F--CH4 305-6
H3CCH3 305 ClH-Cl- 308-11, 319
CO2 295 C1--CH4 305-6
SO2 97-8 BrH-Br- 310
Br--CH4 305-6
HCCH-HCCH 241, 299-300 T-CH 4 305-6
HCCH--OSO 298 HOH-OH- 120-1, 310-12, 319
RCCH-O(CH3)2 301 HOH-OOCH- 326-8
RCCH-NH3 301 CH3OH-OCH,- 120, 312
RCCH-N(CH3)3 301 HSH-SH- 310-11
RCCH-OC(CH3)2 301 H2NH-NH2- 310-11, 319
RCCH-NCCH3 301 H2NH-OOCH- 327
F3CH-NH3 302 H2PH-PH2- 310-11
HN03-CH4 305 HCOOH-OOCH- 313-5, 320
DNA base pairs CH3COOH-OOCCH3- 312
dispersion energy 31 H2CNH-NCH2- 313
geometry 113-8 NCH-CN- 312
dipole moments 116 HCCH-CCH- 312
stacked structure 118 H2CCH2-CHCH2- 313
H2SO4-SO4H- 315
Cationic Complexes

HFH+-FH 316 Neutral Trimers and Oligomers


HC1H+-C1H 316, 318-9 (HF)3 245-8
H2OH+-OH2 120-1, 316-7, 319, 324-5 (HF)n 248-52
H2OH+-OHCH3 324-5 (HC1)3 245-8
H2OH+-O(CH3)2 324-5 FH-FH-OH2 278-82
H2OH+-OHC2H5 324-5 FH-OH2-OH2 280-2
H2OH+-NH3 337 FH-FH-NH3 272-80, 339
H2OH+-OCH2 180 FH-FH-PH3 273-78
CH3OH2+-OHCH3 121, 319 C1H-OH2-OH2 281-2
H2SH+-SH2 316, 318 C1H-CIH-NH3 272-3
H3NH+-NH3 33, 38, 292, 316-9, 325 (H2O)3 120-1, 260-70
H3NH+-NH2CH3 325 (H2O)n 253-8, 264-6
CH3NH3+-NH2CH3 320 cyclic clusters 257-62
(CH 3 ) 2 NH 2 + -NH(CH 3 ) 2 320 vibrational spectra 256, 266-7
Index of Complexes 369

ROH-OH2-OH2 270-2 (HCN)3 232-40


ROH-ROH-OH2 270-2 (HCN)n 232-40
(ROH)3 270-2 (HCONH2)n 340-1
C6H5OH-OH2-OH2 264 (HCCH)3 241-2
C6H5OH-(OH2)3 265 (HCCH)n 242-4
This page intentionally left blank
Subject Index

acetic acid 329 triple-zeta 5


anharmonicity 15, 22, 120, 139-40, 169 well-tempered 7, 38, 59
in ionic complexes 308, 317 basis set extension (see basis set superposition
in neutral binaries 146, 152-4, 186-90 error, secondary)
in oligomers 251, 254, 256 basis set superposition error
anti-H-bonding 67-68 in decomposition terms 37, 39, 260
asymmetric H-bond, definition 321 definition 23-28
atomic charge 18 in dimers with FH 55-56, 293
atomic polar tensor 179 in dimers with H2O 216, 224, 296
definition 150, 162, 277-8 in H2NH-NH3 84-88
average H-bond energy 234, 244-6, 267-8 in DNA base pairs 114
in ionic complexes 309, 327
Badger-Bauer rule (see relationship between in oligomers 234, 244, 246, 254, 257
H-bond strength and frequency shift) effect on EFG 160
barrier effect on NMR spectrum 171
interconversion 208-9, 212-5, 218-20, 225, effect on vibrational spectrum 189-90
300 secondary 24-25, 56-59
proton transfer 225-6, 311, 312, 315-7, 325 spatial aspects 56-59
rotational 69, 215, 300 bifurcated structure 53, 77-80, 215-20, 223-5,
torsional 89, 213, 265, 269 280, 305
basis set 4-7 in ionic complexes 317, 326-8
atomic natural orbital 7 bond functions 6, 57, 78, 87, 214
complete 27 Born equation 336
correlation-consistent 7, 72, 264 Born-Oppenheimer approximation 3, 21
dirner-centered 24 branching clusters 257
double-zeta 4
minimal 4, 25-26, 84, 217, 330, 332 carbenes 298
monomer-centered 24 carbohydrates 299
split-valence 5, 330 CASSCF 10, 88

371
372 Subject Index

centrosymmetric 308-11, 314-5 decomposition


charge flux 150-1, 278 anisotropy 260—1
charge transfer energy 31, 32, 35, 37 definitions 28-39
chemical Hamiltonian 26-27 of dispersion energy 221-2
CH group as proton donor 99, 113, 191-5, 224, Kitaura-Morokuma 32—34, 67
240-4, 298-307 natural bond orbitals 34, 222-3
CISD 8, 75, 141-3, 168, 308 nonadditivity 260
cis/trans competition 216 perturbation schemes 37-39
combination band 190 reduced variational space method 34
complexation energy (see interaction energy) of dimers with H2O 67-68, 216, 220-3
configuration interaction 7 of H2O clusters 258-60
contraction 4 deformation energy (see also induction energy)
convergence 234 219, 260, 268
cooperativity degree of polarization 162
definition 234, 275 density difference map 18
negative 231, 242, 271 dielectric constant 336
correlation (see electron correlation) dipole moment
Coriolis interaction 88 alignment 90
Coulombic energy (see electrostatic energy) average 235, 253
counterpoise (see also basis set superposition change caused by correlation 64, 222
error) 26 change caused by H-bond formation 58-59,
definition 24 155, 235
coupled cluster 8-9 change caused by vibration 157, 277
in (HX)2 72, 75 change caused by bond stretch 210, 277, 308
in H2CO complexes 91-92 directionality 12
in oligomers 269 dispersion energy 66
in ionic complexes 308, 316, 319 anisotropy 210-2, 221-2
in weak complexes 295 definition 31, 37-38
vibrational spectra 142-3, 168, 181 in (HC1)2 74
coupled pair 8-9 of dimers with H2O 70, 78, 221-2, 295
in HX complexes 75, 186, 213 in DNA base pairs 117
in H2O complexes 219 dissociation energy (see also interaction energy)
in NH3 complexes 86,154 17
in H2CO complexes 93 double proton acceptor 231, 242, 261, 263
in oligomers 237, 240, 248 double proton donor 231, 242, 261, 263
in weak complexes 292-3 dynamics 14
coupled perturbed Hartree-Fock 162
crystal orbital method 238, 253 effective core potential 7, 62, 76, 158
cyclic chains electric field 337
of HX 245-52 gradient 154, 159-60
of H2O 257-64, 268-72 electron correlation 4
of HCN 240 contribution to H-bond 54, 63, 72, 221-2
of HCCH 241-4 definition 7-10
cyclic structure effect on proton transfer 309-11, 313, 315,
of (HX)2 146, 213-4 320, 331-2
of (H2O)2 53, 77-80, 215-9 electrostatic energy
of (NH3)2 85-88, 177-9, 208-9 correlation correction to 31, 39, 211
of HOH complexes 105 , 223-5 as factor in geometry 68, 220-1
of (CHONH2)2 111 electrostatic potential 18
enthalpy 17
damping factor 26 entropy 17-18, 21
Davidson correction 75-76 in dimers containing H2O 71, 78, 218
Subject Index 373

in FH-CIF 293 ion pair (see proton transfer)


in (H2O)n 266 isotopic substitution 21, 118-21, 153, 258, 315,
in ionic complexes 327 339
of isotopic substitution 121
equilibrium constant 340 LCAO 4
of formation 52, 120 linear combination of atomic orbitals (see
ESCA 14 LCAO)
exchange energy 30, 32, 37 localized molecular orbitals 35
London forces (see dispersion)
F-bonding (see reverse complex)
FG matrix (see GF matrix) Marcus theory 323-6
force constant 197, 208, 337 matrix, effects of 147, 166-8, 338-9
formic acid 328 MCSCF 9-10
four-body interactions 257, 267-8 minimum
fractionation factor 121 global vs. secondary 11, 15, 102
functional counterpoise (see counterpoise) true (see true minimum)
mixing term 32
Gaussian functions 4, 5, 26 multipole expansion 220-1
GF matrix 139, 161
ghost orbitals (see also basis set superposition) nomenclature 52
24, 39, 56-59 of vibrational modes 140-1
GIAO 19, 171 nonadditivity
Gibbs free energy 18, 21 definition 231
in dimers containing H2O 71, 78, 96 perturbational approach 39
in FH-CIF 293 contribution of correlation 262
in (H2O)n 266 nuclear quadrupole coupling (see also
in ionic complexes 327 quadrupole coupling constant) 140
of isotopic substitution 121 nuclear relaxation energy 16
gradient algorithms 10, 26, 216, 225, 245
orbital exponent, choice of 7, 56
Hamiltonian 3, 9, 26, 27
SCRF model 147 partition function 16
harmonic approximation 22, 139 Pauli blockade 34
Hartree-Fock approximation 3-4 Pauli exchange principle 26, 33-34
H-borid energy (see interaction energy) perturbation theory
Heisenberg uncertainty principle 3 many-body 9, 308
Heitler-London energy 30, 35, 260 M011er-Plesset 9, 38, 247
Hessian matrix 10 Rayleigh-Schrodinger 37
Hoogsteen geometry 115-8 symmetry-adapted 27, 37, 70, 261
hybridization 62 polarizability
hydrogen bond energy (see interaction energy) effects of ghost functions 59
proton 321
IGLO 19 polarization energy
independent electron pair approximation 8 definition 31, 32, 37
induction energy 261, 295 relation to Pauli principle 33
definition 30-31, 37 polarization function 5-6
intensity, vibrational spectrum Pople correction 75-76
formulation 139-40 potential energy surface (see also barrier)
Raman (see Raman intensity) shape of 11, 308-9, 311
interaction energy 13, 15-17, 322 bending 65-66
intrinsic well depth 323 primitive function 4, 7
374 Subject Index

principal moment of inertia 19, 21 rotational energy 16


proton affinity 322 Rydberg functions 308
difference 322-6, 334, 338-9
proton transfer salt bridge (see proton transfer)
correlation effects 61 scaling factor 139
effect of environment 311, 314-5, 328, scattering activity 163
335-7 Schrodinger equation 3, 26, 152, 254
parameter 333-5, 337 SCRF 147
potential 14, 61, 225-6, 313, 330-5 self-consistent field 3, 336
in amines 225-6, 330-2, 334-5, 337-339 sequential H-bond chain 231, 339-41
in arsines and phosphines 331, 337, 339 shifts, electronic (see redistributions)
pseudopotential (see effective core potential) size-consistency 8, 9, 75-76, 216
pyramidal geometry 64-66, 221-2 Slater-type orbitals 4, 26
solvent effects 328-9, 336-7
QCISD 8, 87 spectrum
quadrupole coupling constant 19, 239, 337 electron resonance 14
NMR 14, 19, 171, 307, 311, 315
Raman intensity 188-9, 191 resonant photoacoustic 14
definition 140, 156-8 vibrational 13-14
redistributions, electronic 13, 18 spin-spin coupling constant 14
caused by H-bond formation 57, 67 stacked geometry 216-7
reference configuration 7-8 steric repulsion (see exchange energy)
relationship STO (see Slater-type orbitals)
between frequency shift and bond stretch strength of H-bond (see interaction energy)
161, 266 stretch of AH bond
between frequency shift and isotopic ratio in XH 55, 77, 96
339 in HCOOH 99
between H-bond strength and bond stretch in NCH oligomers 232
66 stretch of carbonyl bond 90, 93-94, 96, 99,
between H-bond strength and force constant 180
161, 197 stretch of CN bond 232
between H-bond strength and frequency shift substituent effects
140, 155, 180, 183 alkylation 60-61, 66-67, 81, 103, 159, 169,
between H-bond strength and intensity 155 185, 270-2
between H-bond strength and proton affinity effect on proton transfer 311, 319
difference 322-4 chloro 83-84, 93, 171
between H-bond strength and proton transfer fluoro 93
barrier 161, 197 phenyl 71, 83, 170, 177, 263-5
between H-bond length and quadrupole silyl 82, 169-70
coupling constant 239 supermolecule approach 32
between isotropic and perpendicular shifts syn/anti competition 326-30
171
between proton transfer potential and thermodynamic quantities 15-18
proton affinity difference 322-6, three-body interactions 246-8, 254-62,
333-6, 338 267-70, 275, 279
between stretching frequency and quadrupole anisotropy 260-1
coupling constant 239 trans/cis (see cis/trans competition)
reverse complex translational energy 16
in HF complexes 66, 102, 209-10, 212, 292, trifurcated geometry 209, 217-8, 305
294 in ionic complex 317, 318
in HOH complexes 66, 169-70, 224, 304 true minimum 10
Subject Index 375

in complexes with H2O 69, 91, 215-6, 224 correlation diagram 338
in (NH3)2 85, 177, 208-9 levels 65-66
in complexes with HCONH2 113, 196
in nucleic acid base pairs 115 Watson-Crick geometry 115-8
in water cluster 262-4 wave function 3, 34, 64
tunneling 88 vibrational 65-66, 264

van der Waals zero-point vibrational energy 16, 17


force 12 effect on proton transfer 312, 331
van der Waals radius 12, 310 means of calculation 20-22
variation principle 23 in neutral dimers 56, 224, 225, 293
vibrational in oligomers 234, 240, 282
averaging 89 in ionic complexes 309, 317, 331

S-ar putea să vă placă și