Untitled

Download as pdf or txt
Download as pdf or txt
You are on page 1of 280

10718_9789813230446_tp.

indd 1 12/12/17 5:05 PM


Peking University–World Scientific Advanced Physics Series
ISSN: 2382-5960

Series Editors: Enge Wang (Peking University, China)


Jian-Bai Xia (Chinese Academy of Sciences, China)

Vol. 3 Computer Simulations of Molecules and Condensed Matter:


From Electronic Structures to Molecular Dynamics
by Xin-Zheng Li and En-Ge Wang

Vol. 2 Geometric Methods in Elastic Theory of Membranes in


Liquid Crystal Phases
Second Edition
by Zhanchun Tu, Zhongcan Ou-Yang, Jixing Liu and Yuzhang Xie

Vol. 1 Dark Energy


by Miao Li, Xiao-Dong Li, Shuang Wang and Yi Wang

Rachel - 10718 - Computer Simulations of Molecules.indd 1 29-12-17 9:52:31 AM


10718_9789813230446_tp.indd 2 12/12/17 5:05 PM
Published by
World Scientific Publishing Co. Pte. Ltd.
5 Toh Tuck Link, Singapore 596224
USA office: 27 Warren Street, Suite 401-402, Hackensack, NJ 07601
UK office: 57 Shelton Street, Covent Garden, London WC2H 9HE

Library of Congress Cataloging-in-Publication Data


Names: Li, Xin-Zheng, 1978– author. | Wang, En-Ge, 1957– author.
Title: Computer simulations of molecules and condensed matter : from electronic structures to
molecular dynamics / Xin-Zheng Li (Peking University, China),
En-Ge Wang (Peking University, China).
Other titles: Peking University-World Scientific advance physics series ; v. 3.
Description: Singapore ; Hackensack, NJ : World Scientific, [2017] |
Series: Peking University World Scientific advance physics series, ISSN 2382-5960 ; vol. 3 |
Includes bibliographical references.
Identifiers: LCCN 2017032699| ISBN 9789813230446 (hardcover ; alk. paper) |
ISBN 9813230444 (hardcover ; alk. paper)
Subjects: LCSH: Condensed matter--Computer simulation. |
Molecular dynamics--Computer simulation.
Classification: LCC QC173.457.C64 L58 2017 | DDC 530.4/10113--dc23
LC record available at https://lccn.loc.gov/2017032699

British Library Cataloguing-in-Publication Data


A catalogue record for this book is available from the British Library.

Copyright © 2018 by World Scientific Publishing Co. Pte. Ltd.


All rights reserved. This book, or parts thereof, may not be reproduced in any form or by any means,
electronic or mechanical, including photocopying, recording or any information storage and retrieval
system now known or to be invented, without written permission from the publisher.

For photocopying of material in this volume, please pay a copying fee through the Copyright Clearance
Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. In this case permission to photocopy
is not required from the publisher.

Printed in Singapore

Rachel - 10718 - Computer Simulations of Molecules.indd 2 29-12-17 9:52:31 AM


December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-fm page v

Preface

State-of-the-art computer simulation of molecules and condensed matters,


according to the Born–Oppenheimer (BO) approximation, requires an accu-
rate ab initio treatment of both the electronic structures and the nuclei’s
motion on (and sometimes even beyond) the corresponding potential energy
surfaces. As a student majoring in computational condensed matter physics,
one of the authors (Xin-Zheng Li, XZL) had taken many years to under-
stand this simple statement. During this time, his work in collaboration
with Prof. Jianbai Xia and Prof. Enge Wang as well as other scholars in
Europe (Prof. Angelos Michaelides, Prof. Dr. Matthias Scheffer, Dr. Ricardo
Gómez-Abal, and Prof. Dr. Claudia Draxl, etc.) has luckily covered some
topics in both regimes. Based on this limited, yet to a certain extent,
unique experience, the authors want to share their understanding of this
rapidly growing field with the readers, especially those Chinese graduate
students majoring in computational condensed matter physics or chemistry.
Special focus will be put on the basic principles underlying the present
electronic structure calculations and the molecular dynamics simulations.
A wide range of electronic structure theories will be introduced, includ-
ing the traditional quantum chemistry method, density-functional theory,
many-body perturbation theory, etc. Besides these electronic structures,
motions of the nuclei will also be described using molecular dynamics,
with extensions to enhanced sampling and free energy calculation tech-
niques including umbrella sampling, metadynamics, integrated tempering
sampling, etc., and the thermodynamic integration methods. As a further
extension beyond the standard BO molecular dynamics, some simulation
techniques for descriptions of the quantum nuclear effects will also be dis-

v
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-fm page vi

vi Computer Simulations of Molecules and Condensed Matter

cussed, based on Feynman’s path integral representation of the quantum


mechanics. With such a choice of theories on both the electronic structures
and molecular dynamics perspectives, hopefully, we can help those gradu-
ate students to find the proper method for tackling the physical/chemical
problems they are interested in.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-fm page vii

Contents

Preface v

1. Introduction to Computer Simulations


of Molecules and Condensed Matter 1
1.1 Born–Oppenheimer Approximation and the
Born–Oppenheimer Potential Energy Surface . . . . . . 2
1.2 Categorization of the Tasks in Computer Simulations
of Molecules and Condensed Matters . . . . . . . . . . . 6
1.2.1 Electronic Structure Calculations . . . . . . . . . 6
1.2.2 Geometry Optimization, Stationary Points on
PES, Local Minimum, and Transition State . . . 7
1.2.3 Metastable State and Transition State
Searching . . . . . . . . . . . . . . . . . . . . . . 8
1.2.4 Molecular Dynamics for the Thermal
Effects . . . . . . . . . . . . . . . . . . . . . . . . 11
1.2.5 Extensions of MD: Enhanced Sampling
and Free Energy Calculations . . . . . . . . . . . 12
1.2.6 Path Integral Simulations for the Quantum
Nuclear Effects . . . . . . . . . . . . . . . . . . . 13
1.3 Layout of the Book . . . . . . . . . . . . . . . . . . . . . 15

2. Quantum Chemistry Methods and


Density-Functional Theory 17
2.1 Wave Function-Based Method . . . . . . . . . . . . . . . 17
2.1.1 The Hartree and Hartree–Fock
Approximations . . . . . . . . . . . . . . . . . . 18
2.1.2 Beyond the Hartree–Fock Approximation . . . . 21

vii
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-fm page viii

viii Computer Simulations of Molecules and Condensed Matter

2.2 Density-Functional Theory . . . . . . . . . . . . . . . . . 22


2.2.1 Thomas–Fermi Theory . . . . . . . . . . . . . . 22
2.2.2 Density–Functional Theory . . . . . . . . . . . . 24
2.2.3 Exchange–Correlation Energy . . . . . . . . . . . 26
2.2.4 Interpretation of the Kohn–Sham Energies . . . 28

3. Pseudopotentials, Full Potential, and Basis Sets 31


3.1 Pseudopotential Method . . . . . . . . . . . . . . . . . . 32
3.1.1 Generation of the Pseudopotential . . . . . . . . 33
3.1.2 Implicit Approximations . . . . . . . . . . . . . . 38
3.1.2.1 Frozen Core . . . . . . . . . . . . . . . 38
3.1.2.2 Core–Valence Linearization . . . . . . 38
3.1.2.3 Pseudoization . . . . . . . . . . . . . . 39
3.2 FP-(L)APW+lo Method . . . . . . . . . . . . . . . . . . 39
3.2.1 LAPW Basis Functions . . . . . . . . . . . . . . 42
3.2.2 APW+lo Basis Functions . . . . . . . . . . . . . 43
3.2.3 Core States . . . . . . . . . . . . . . . . . . . . . 44
3.2.4 Potential and Density . . . . . . . . . . . . . . . 45

4. Many-Body Green’s Function Theory


and the GW Approximation 47
4.1 Green’s Function Method . . . . . . . . . . . . . . . . . 49
4.1.1 The Green’s Function . . . . . . . . . . . . . . . 49
4.1.2 The Dyson Equation . . . . . . . . . . . . . . . . 52
4.1.3 Self-Energy: Hedin Equations . . . . . . . . . . . 54
4.1.4 The Quasi-Particle Concept . . . . . . . . . . . . 57
4.2 GW Approximation . . . . . . . . . . . . . . . . . . . . . 58
4.3 G0 W0 Approximation . . . . . . . . . . . . . . . . . . . . 61
4.4 Numerical Implementation of an All-Electron G0 W0 Code:
FHI-Gap . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
4.4.1 Summary of the G0 W0 Equations . . . . . . . . 68
4.4.2 The Mixed Basis . . . . . . . . . . . . . . . . . . 69
4.4.3 Matrix Form of the G0 W0 Equations . . . . . . . 71
4.4.4 Brillouin-Zone Integration of the
Polarization . . . . . . . . . . . . . . . . . . . . . 73
4.4.5 The Frequency Integration . . . . . . . . . . . . 76
4.4.6 Flowchart . . . . . . . . . . . . . . . . . . . . . . 79
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-fm page ix

Contents ix

5. Molecular Dynamics 81
5.1 Introduction to Molecular Dynamics . . . . . . . . . . . 83
5.1.1 The Verlet Algorithm . . . . . . . . . . . . . . . 84
5.1.2 The Velocity Verlet Algorithm . . . . . . . . . . 86
5.1.3 The Leap Frog Algorithm . . . . . . . . . . . . . 88
5.2 Other Ensembles . . . . . . . . . . . . . . . . . . . . . . 89
5.2.1 Andersen Thermostat . . . . . . . . . . . . . . . 90
5.2.2 Nosé–Hoover Thermostat . . . . . . . . . . . . . 92
5.2.3 Nosé–Hoover Chain . . . . . . . . . . . . . . . . 100
5.2.4 Langevin Thermostat . . . . . . . . . . . . . . . 102
5.2.5 Andersen and Parrinello–Rahman Barostats . . . 104
5.3 Examples for Practical Simulations in Real Poly-Atomic
Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . 107

6. Extension of Molecular Dynamics, Enhanced


Sampling and the Free-Energy Calculations 113
6.1 Umbrella Sampling and Adaptive Umbrella Sampling
Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
6.2 Metadynamics . . . . . . . . . . . . . . . . . . . . . . . . 124
6.3 Integrated Tempering Sampling . . . . . . . . . . . . . . 126
6.4 Thermodynamic Integration . . . . . . . . . . . . . . . . 129

7. Quantum Nuclear Effects 137


7.1 Path Integral Molecular Simulations . . . . . . . . . . . 140
7.1.1 Path Integral Representation of the
Propagator . . . . . . . . . . . . . . . . . . . . . 140
7.1.2 Path Integral Representation of the Density
Matrix . . . . . . . . . . . . . . . . . . . . . . . 143
7.1.3 Statistical Mechanics: Path Integral Molecular
Simulations . . . . . . . . . . . . . . . . . . . . . 148
7.1.4 Staging and Normal-Mode Transformations . . . 156
7.1.5 Evaluation of the Zero-Point Energy . . . . . . . 166
7.2 Extensions Beyond the Statistical Studies . . . . . . . . 175
7.2.1 Different Semiclassical Dynamical Methods . . . 176
7.2.2 Centroid Molecular Dynamics and Ring-Polymer
Molecular Dynamics . . . . . . . . . . . . . . . . 178
7.3 Free Energy with Anharmonic QNEs . . . . . . . . . . . 184
7.4 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . 188
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-fm page x

x Computer Simulations of Molecules and Condensed Matter

7.4.1 Impact of QNEs on Structures of the


Water-Hydroxyl Overlayers on Transition Metal
Surfaces . . . . . . . . . . . . . . . . . . . . . . . 188
7.4.2 Impact of Quantum Nuclear Effects on the
Strength of Hydrogen Bonds . . . . . . . . . . . 196
7.4.3 Quantum Simulation of the Low-Temperature
Metallic Liquid Hydrogen . . . . . . . . . . . . . 205
7.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . 217

Appendix A: Useful Mathematical Relations 219


A.1 Spherical Harmonics . . . . . . . . . . . . . . . . . . . . 219
A.2 Plane Waves . . . . . . . . . . . . . . . . . . . . . . . . . 220
A.3 Fourier Transform . . . . . . . . . . . . . . . . . . . . . . 220
A.4 Spherical Coordinates . . . . . . . . . . . . . . . . . . . . 221
A.5 The Step(Heaviside) Function . . . . . . . . . . . . . . . 221
Appendix B: Expansion of a Non-Local Function 223

Appendix C: The Brillouin-Zone Integration 227


C.1 The Linear Tetrahedron Method . . . . . . . . . . . . . 227
C.1.1 The Isoparametric Transfromation . . . . . . . . 229
C.1.2 Integrals in One Tetrahedron . . . . . . . . . . . 232
C.1.3 The Integration Weights . . . . . . . . . . . . . . 232
C.2 Tetrahedron Method for q-Dependent Brillouin-Zone
Integration . . . . . . . . . . . . . . . . . . . . . . . . . . 233
C.2.1 Isoparametric Transformation . . . . . . . . . . . 235
C.2.2 The Integration Region . . . . . . . . . . . . . . 236
C.2.3 Polarizability . . . . . . . . . . . . . . . . . . . . 237
C.2.3.1 Polarizability on the Real Frequency
Axis . . . . . . . . . . . . . . . . . . . 238
C.2.3.2 Polarizability on the Imaginary
Frequency Axis . . . . . . . . . . . . . 240
Appendix D: The Frequency Integration 243

References 245

Acknowledgements 261

Index 263
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-fm page xi

List of Figures

1.1 Pictorial illustration of the BO approximation . . . . . . . . . 4


1.2 Potential energy surface (PES) . . . . . . . . . . . . . . . . . 5
1.3 Illustration of the geometry optimization using traditional
Chinese painting . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.4 Illustration of the transition state using traditional Chinese
painting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

3.1 Pseudovalence wave functions in comparison with the


all-electron ones . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.2 Ionic pseudopotentials . . . . . . . . . . . . . . . . . . . . . . 36
3.3 The space partition in the augmented plane wave method . . 40

4.1 The poles of the Green function . . . . . . . . . . . . . . . . . 51


4.2 The self-consistent scheme for the exact solution of Hedin
equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
4.3 Spectral function for a discrete and a continuous spectrum . . 58
4.4 The self-consistent scheme in the GW approximation . . . . . 61
4.5 Band diagram of Si . . . . . . . . . . . . . . . . . . . . . . . . 64
4.6 The Fermi surface in the special point and tetrahedron schemes 74
4.7 The integration region in the q-dependent Brillouin-zone
integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
4.8 Polarizability of the free electron gas . . . . . . . . . . . . . . 77
4.9 Analytical structure of Σc = iGW c on the complex frequency
plane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
4.10 Flowchart of the FHI-gap G0 W0 code . . . . . . . . . . . . . 79

5.1 Illustration of an ab initio MD simulation . . . . . . . . . . . 108

xi
December 28, 2017 12:30 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-fm page xii

xii Computer Simulations of Molecules and Condensed Matter

5.2 Illustration of how the electron charge transfer is analyzed in


an ab initio MD simulation . . . . . . . . . . . . . . . . . . . 109
5.3 Illustration of the hydrogen bond dynamics related to the
proton transport in liquid water . . . . . . . . . . . . . . . . 110
5.4 Liquid hydrogen dissociation under pressure from ab initio MD
simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112

6.1 Principle underlying reconstruction of the PMF from a biased


one . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
6.2 An example when the biased potential doesn’t work . . . . . 119
6.3 Illustration of how umbrella sampling works . . . . . . . . . . 120
6.4 Illustration of the principles underlying the metadynamics
method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
6.5 Illustration of the Fλ in the thermodynamic integration
method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131

7.1 Illustration of how a propagator can be calculated using path


integral . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
7.2 Mapping from the quantum system to a classical polymer . . 151
7.3 An example of the artificial polymer in a path integral
molecular simulation . . . . . . . . . . . . . . . . . . . . . . . 152
7.4 Evolution of the internal energy’s expectation value as a
function of temperature in the MD and PIMD simulations . . 175
√ √
7.5 Static geometry optimized structure of the 3 × 3-R30◦
overlayer (with classical nuclei) that forms on the transition
metal surfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
7.6 Statistical structural information from the ab initio MD and
PIMD simulations of the water-metal interfaces, using some
selected structural properties . . . . . . . . . . . . . . . . . . 192
7.7 Free energy profile for the protons along the intermolecular
axes within the water-hydroxyl overlayers . . . . . . . . . . . 194
7.8 Probability distributions in the MD and PIMD simulations as
functions of δ and RO−O . . . . . . . . . . . . . . . . . . . . . 195
7.9 Correlation between the impact of the quantum nuclear effects
and the hydrogen bond strength . . . . . . . . . . . . . . . . 200
7.10 HF clusters as examples for detailed analysis of the quantum
nuclear effects . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-fm page xiii

List of Figures xiii

7.11 A quantification for the competition between the quantum


fluctuations on the stretching and bending modes . . . . . . . 202
7.12 Ab initio PIMD simulations of solid-liquid coexistence and
melting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
7.13 Phase diagram of hydrogen and the low-temperature metallic
liquid phase . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209
7.14 Static lattice ground state enthalpies of different crystal
structures in hydrogen . . . . . . . . . . . . . . . . . . . . . . 212
7.15 Single point total energies of snapshots from the thermalized
state of the two-phase PIMD simulations at 700 GPa . . . . . 212
7.16 Melting temperature calculated at 700 GPa using different
numbers of beads . . . . . . . . . . . . . . . . . . . . . . . . . 214
7.17 Melting temperature calculated at 700 GPa using different
number of atoms in the ab initio MD simulations . . . . . . . 214
7.18 RDF obtained from the two-phase simulation, using different
time intervals . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
7.19 Analysis of nuclear exchange effects in the PIMD simulations 216
7.20 Electron and phonon properties of the I41/amd structure of
solid hydrogen at 500 GPa . . . . . . . . . . . . . . . . . . . . 217
7.21 Dependence of Tc on the cut-off energy and the effective mass
in the superconductivity calculations . . . . . . . . . . . . . . 218
C.1 Two-dimensional sketch for the BZ in the tetrahedron method 228
C.2 The configurations for the region to be integrated . . . . . . . 237
b2530   International Strategic Relations and China’s National Security: World at the Crossroads

This page intentionally left blank

b2530_FM.indd 6 01-Sep-16 11:03:06 AM


December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-fm page xv

List of Tables

3.1 Relativistic quantum numbers. . . . . . . . . . . . . . . . . . 45

4.1 Comparison of some reported G0 W0 band gaps. . . . . . . . 65

xv
b2530   International Strategic Relations and China’s National Security: World at the Crossroads

This page intentionally left blank

b2530_FM.indd 6 01-Sep-16 11:03:06 AM


December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch01 page 1

1
Introduction to Computer
Simulations of Molecules
and Condensed Matter

Since the discovery of electron as a particle in 1896–1897, the the-


ory of electrons in matter has ranked among the great challenges in
theoretical physics. The fundamental basis for understanding mate-
rials and phenomena ultimately rests upon understanding electronic
structure [1].

It is without any hesitation that we assent to R. M. Martin’s point of


view and begin, with his quote, the introduction of this book on computer
simulations of molecules and condensed matters. The electrons, being an
interacting many-body entity, behave as a quantum glue which holds most
of the matters together. Therefore, principles underlying the behavior of
these electrons, a large extent, determine properties of the system (elec-
tronic, optical, magnetic, mechanical, etc.) we are going to investigate. As
a consequence, an introduction to the computer simulation of molecules
and condensed matters should naturally start with a discussion on theories
of electronic structures.
One point implied in this statement is that the concept of electronic
structures is polymorphous, in the sense that it covers all properties related
to the electrons in matter. For example, it can refer to the total energy
of the electrons, their density distribution, the energy needed for extract-
ing (injecting) one electron out of (into) the system, their response to an
external perturbation, etc. These properties are in principle measured by
different experiments and described using different theoretical methods.
Therefore, while discussing “electronic structures”, one must point out the

1
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch01 page 2

2 Computer Simulations of Molecules and Condensed Matter

specific properties of the electrons referred to and the theories used. Among
the various properties and theories used for describing the electronic sys-
tem, we will focus on the ones concerning depiction of the total energies and
spectroscopies of the system within the ab initio framework throughout this
book. For electronic structure theories based on model Hamiltonian, e.g. the
effective-mass envelope function and the tight-binding methods, which are
equally important in molecular simulations but will not be discussed here,
please refer to the seminal book of Martin in Ref. [1].
Besides the behavior of the electrons, the motion of the nuclei is another
aspect one must accurately address in simulating the material properties,
since a real material is composed of interacting electrons and nuclei. To
describe such a correlated motion, some basic concepts underlying our daily
research must be introduced, among which the Born–Oppenheimer (BO)
approximation and the potential energy surface are the most crucial.
Because of this, in the following, we start our discussions by introducing
these two concepts. Using these concepts, we can categorize the major-
ity of the tasks we want to fulfill in daily researches concerning simula-
tions of material properties into two different regimes, i.e. those concerning
mainly the electronic structures and those concerning mainly the nuclear
motion. The whole book is then organized on the basis of such a catego-
rization. In Chapters 2–4, we discuss different electronic structure theories
and some technical details concerning their implementation. Chapters 5–7
focus on the molecular dynamics (MD) method and its various extensions
in descriptions of the nuclear motion. With this choice of theories on both
the electronic structures and the molecular dynamics levels, we hope that
we can help the graduate students to find the proper method for tackling
the physical/chemical problems they are interested in, in their practical
researches.

1.1 Born–Oppenheimer Approximation and the


Born–Oppenheimer Potential Energy Surface
Before we start, let us first present the key equation we are going to tackle
in simulating properties of a real material. Any poly-atomic system can
be viewed as an intermixture of two coupled subsystems, constituted by
M nuclei and N electrons, respectively. In principle, the only prerequi-
site for the description of all the quantum mechanical properties of such a
system, for simplicity in the non-relativistic regime, is the solution of the
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch01 page 3

Introduction to Computer Simulations of Molecules and Condensed Matter 3

many-body Schrödinger equation

ĤΨ(r1 , r2 , . . . , rN , R1 , R2 , . . . , RM ) = EΨ(r1 , r2 , . . . , rN , R1 , R2 , . . . , RM ),
(1.1)
where ri stands for the Cartesian coordinate of the ith electron and Ri
stands for that of the ith nucleus. The Hamiltonian operator is given by
N
 M
1 1  1
Ĥ = − ∇2i + V (ri − ri ) − ∇2j
i=1
2 2  j=1
2M j
i=i

1 1
+ V (Rj − Rj  ) + V (ri − Rj ) (1.2)
2 
2 i,j
j=j

in atomic units (a.u.) with Hartree as the unit of energy. The first two terms
in Eq. (1.2) correspond to the kinetic energy and the Coulomb interaction
potential, respectively, of the electrons. The third and fourth terms repre-
sent the same physical quantities, but for the nuclei. The fifth term is the
Coulomb interaction potential between the electrons and the nuclei which
couples the two subsystems.
We note that this atomic unit will be used in all equations through-
out this book. The notation of the electron coordinate with r and that
of the nucleus with R will be used when our discussions concern both
the electrons and the nuclei. In Chapters 5 and 7, when propagations of
purely the nuclei are discussed, r is also used to denote the position of the
nucleus since in those cases, R is often used to denote other quantities in the
literature.
The BO approximation, proposed by Born and Oppenheimer in 1927
[2] and also known as the “adiabatic approximation” in descriptions of
electronic structures, makes use of the feature that the masses of the nuclei
are several orders of magnitude larger than that of the electrons and their
velocities are consequently much smaller with similar kinetic energy. Tak-
ing advantage of these extremely different dynamical regimes, this adiabatic
approximation allows us to address the dynamics of the electronic subsys-
tem separately from that of the nuclei by considering the latter as static.
Pictorially, this is similar to the case when a slow-moving person accidently
touches a hornets’ nest and gets surrounded by the hornets (Fig. 1.1). Since
the velocities of these hornets are much larger than that of the person, wher-
ever he goes in order to escape, the hornets quickly adjust their positions so
that this unlucky guy is always surrounded by the hornets and therefore has
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch01 page 4

4 Computer Simulations of Molecules and Condensed Matter

Figure 1.1 Pictorial illustration of the BO approximation analogy. The electrons are
much lighter than the nuclei and therefore with similar kinetic energy, they are much
faster. When the nuclei move, the electrons quickly adjust their positions as if the nuclei
are stationary. This is similar to the case when an unlucky slow-moving guy is surrounded
by hornets. Wherever he goes, the hornets can easily adjust their positions as if this
person is stationary. In other words, there is an “adiabatic” correlation between the
movement of this person and that of the hornets, so that this person has to suffer the
sting all the time.

to suffer being stung all the time. Similarly, in the electron–nuclei system,
since the velocities of the electrons are some orders of magnitude larger
than those of the nuclei, for a certain spatial configuration of the nuclei,
we can always focus on describing the electronic system first and allow the
electrons to relax to their eigenstates before the nuclei are ready to move
to the next step. Using this concept, from the mathematical perspective,
the poly-atomic quantum system including both electrons and nuclei as
described by Eq. (1.1) can be simplified into a system that only includes the
electrons as quantum particles. The Schrödinger equation for the electrons
then reads
Ĥe Φ(r1 , r2 , . . . , rN ) = Ee Φ(r1 , r2 , . . . , rN ), (1.3)
where the Hamiltonian
N
 1 1 1
Ĥe = − ∇2i + V (ri − ri ) + V (ri − Rj ) (1.4)
i=1
2 2  2 i,j
i=i

depends only parametrically on the nuclear configuration. As simple as it


is, the movements of the electrons and the nuclei become decoupled.
The BO approximation, as introduced above, indicates that the total
energy of the system at a certain spatial configuration of the nuclei equals
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch01 page 5

Introduction to Computer Simulations of Molecules and Condensed Matter 5

Figure 1.2 (Courtesy of Dr. Yexin Feng) PES is a plot of the system’s total energy,
defined as the total energy of the electronic system plus the classical Coulomb repulsion
potential between the nuclei as a function of the nuclear geometric parameters. In the
above case, two geometric parameters, labelled collective variables 1 and 2 (CV1 and
CV2), are chosen to characterize the spatial configuration of the nuclei, and the PESs
are a series of two-dimensional (2D) surfaces corresponding to the electronic ground and
excited states. In principle, these surfaces representing different electronic states (ground
electronic state, first excited electronic state, etc.) may intersect with each other.

the total energy of the electronic many-body system as described by


Eq. (1.3) plus that of the classical internuclear Coulomb repulsion. The
resulting quantity is a function of the nuclear geometry and it has a series
of discrete eigenvalues corresponding to the electronic quantum mechani-
cal ground and excited states. Consequently, the concept of the potential
energy surface (PES) can be introduced as a plot of this total energy as a
function of the geometry of the nuclei (Fig. 1.2). This concept of the PES
is very helpful since it provides us a simple way to visualize the relation-
ship between the geometry of the nuclei and the potential energy of the real
poly-atomic system under investigation at a specific spatial configuration of
the nuclei. Imagine that we can use two geometric parameters to describe
this spatial configuration of the nuclei. At each spatial configuration, we
then have a series of discrete total energies associated with the ground and
excited states of the electrons as a function of these two geometric parame-
ters. A gradual change in the nuclei’s geometry induces a gradual change in
these total energies. Through a continuous yet complete variation of these
geometric parameters over the whole configurational space of the system
under investigation, we can have a series of 2D PESs for this system. In fact,
the nuclei are not stationary. At finite temperatures, the system fluctuates
on these PESs and the purpose of material simulation simplifies into the
reproducing of such fluctuations of the nuclei on these electronic ground-
and excited-state PESs.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch01 page 6

6 Computer Simulations of Molecules and Condensed Matter

1.2 Categorization of the Tasks in Computer


Simulations of Molecules and Condensed
Matters
Using the concepts of the BO approximation and the PES as introduced
above, we can separate the majority of the jobs we need to do in computer
simulations of the molecules and condensed matter into two categories,
namely those concerning mainly the electronic structures and those focusing
on the motion or relaxation of the nuclei. Theories going beyond this BO
approximation are beyond the scope of this book. We direct interested
readers to see Refs. [3–5].

1.2.1 Electronic Structure Calculations

For a static spatial configuration of the nuclei, the equation we need to


solve in simulating the material properties is Eq. (1.3), which depends only
parametrically on the nuclear coordinates. However, solving such a many-
body Schrödinger equation is still much harder said than done and one
must rely on further approximations. Nowadays, there are different methods
which can be used to fulfill this task, including the traditional quantum
chemistry method, the density-functional theory (DFT), and the quantum
Monte-Carlo (QMC) method, etc. within the ab initio framework and the
effective-mass envelope function, tight-binding methods, etc. in the model
Hamiltonian scheme. These methods have achieved great success within the
last half–century as evidenced by such achievements as the Nobel Prize of
Chemistry awarded to W. Kohn and J. Pople in 1998 etc. In this book, we
will take some of them as examples to explain how the electronic structures
are calculated in real poly-atomic systems.
Our discussion starts from the traditional quantum chemistry meth-
ods and DFT in Chapter 2. In practical implementations of these meth-
ods, however, a certain basis set must be chosen for the expansion of the
electronic wave functions. These basis functions need to be both efficient
and accurate in simultaneously depicting two extremely different regimes,
i.e. the core state tightly bound to the nuclei, and the delocalized valence
state, especially in the interstitial region. In Chapter 3, we will explain how
they are designed in computer simulations in practice. In addition to this,
the many-body perturbation theory, one of the standard methods used to
depict the properties of the excited electronic states, will be described in
Chapter 4. For the time being, in this introduction, let us just assume that
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch01 page 7

Introduction to Computer Simulations of Molecules and Condensed Matter 7

we have the electronic structure of the poly-atomic system already properly


described for each spatial configuration of the nuclei.
With this assumption, we can imagine that we have mapped out the
BO PESs as introduced in Sec. 1.1 for the specific system we are interested
in. The next problem therefore transforms into describing the motion of the
nuclei on these PESs.

1.2.2 Geometry Optimization, Stationary Points on PES,


Local Minimum, and Transition State

None of the molecules or condensed matter will be happy if they remain


at an uncomfortable position on their PES. Taking the simplest case, a
positively charged ion subjected to a 2D external potential in which the
characteristic spatial scale for the variation of the potential is much larger
than the dispersion of the electrons within this ion (in other words, the ion
can be approximated as a point charge sitting on the 2D potential). As an
example, the PES for the ground state of the entity composed by the elec-
trons and the nucleus has a few local minima on the x–y plane. This is due to
interactions between the electrons, the nucleus and in this case, most impor-
tantly, the external field (Fig. 1.3). There exist some local minima on this
PES. At each local minimum, the force imposed on the nucleus is zero since
dE/dx = 0. Away from them on the slope, the nucleus feels a force which
leads it to the nearest local minimum (Fig. 1.3). In material simulations,
if we imagine that we know nothing about the structure of the molecules
or condensed matter, the first thing we need to do in order to understand
something about this system is always to construct such a relaxation using
chemical intuitions, starting from different geometries. Such a structural
relaxation is called geometry optimization. It normally helps us to find the
closest static local minimum on the PES of the system to be studied.
In real poly-atomic systems (molecules and condensed matter), since
the configurational space is much more complicated (normally 3N -3D with
N being the number of nuclei), the geometry optimization introduced above
cannot guarantee that the system reaches a minimum after it is carried
out. The only guaranteed property is that the resulting geometry satisfies
∂E/∂x = 0 for all the spatial coordinates, denoted by x, of the system.
The point on the PES corresponding to this geometry is called a stationary
point. On a multi-dimensional PES, except for the local minima where the
corresponding curvatures (namely, ∂ 2 E/∂x1 ∂x2 ) are all positive, station-
ary points also include those with negative curvatures. The most special
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch01 page 8

8 Computer Simulations of Molecules and Condensed Matter

Figure 1.3 Illustration of the geometry optimization using a traditional Chinese paint-
ing, courtesy of Ms. Yehan Feng (XZL’s wife). Imagine that one (x, y) coordinate gives
the spatial configuration of the nuclei, and the landscape represents the PES. The system
(denoted by the person) is unhappy when it is on the slope of the PES and consequently,
forces exist which drag him to a local minima (e.g. points A and B).

ones among them, of chemical interest, are the so-called first-order saddle
points, where the corresponding curvatures are all positive except for one.
These first-order saddle points are interesting because in chemistry, two
local minima on the PES can refer to the reactant and product states. The
lowest first-order saddle point between them on the PES then corresponds
to the transition state (TS) (see Fig. 1.4). Paths close to the TS allow
chemical reactions to happen with the least energy cost compared to other
paths. These terms will often be used in discussions on chemical reactions.
At these stationary points (including the local minima and the first-order
saddle point), the system does not feel the force for relaxation. Therefore,
to distinguish between them, further calculations on vibrational properties
are often needed. Close to stationary points other than those properly-
defined local minima, there are imaginary-frequency phonons (or otherwise
called soft phonons) associated with the negative curvature of the PES.
Accordingly, vibrational spectrum calculation serves as an efficient way to
distinguish a well-defined stable structure from those of other stationary
points with negative curvatures.

1.2.3 Metastable State and Transition State Searching

So far, we have introduced the concepts of the local minima on the PES,
the first-order saddle point (TS) on the PES, and geometry optimization.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch01 page 9

Introduction to Computer Simulations of Molecules and Condensed Matter 9

Figure 1.4 Similar to Fig. 1.3, this is a schematic representation of the concept of TS
using a traditional Chinese painting. Courtesy of Ms. Yehan Feng. Again, one (x, y)
coordinate is used to mean one spatial configuration of the nuclei, and the landscape
represents the PES. The stationary points on the PES refer to those where the slope is
zero (e.g. points A, B, and C in Fig. 1.3). However, the curvatures can be either positive
or negative. Among all stationary points, the one with all curvatures are positive except
for one which is called a first-order saddle point, or a TS, as indicated by the cartoon in
yellow.

Structures of the molecules and condensed matter associated with the local
minima and the first-order saddle points on the PES at the atomic level
are of primary interest in simulations of material properties. Taking the
local minima as an example, characterization of their structures can help
us understand one of the most fundamental questions in material simula-
tions, i.e. why is a material the way it is? Therefore, assuming that the
electronic structures of the system at a specific spatial configuration of the
nuclei can be satisfactorily solved, the primary task in material simulations
associated with understanding the behavior of the system on its PES then
simplifies into the characterization, on the atomic level, of the system’s
structures at these states-of-interest at the first step. The most rigorous
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch01 page 10

10 Computer Simulations of Molecules and Condensed Matter

way to identify geometries of the system related to these special points on


the PES is to carry out a complete yet static mapping of the PES over the
whole configurational space of the nuclei. However, this is not computa-
tionally doable in real materials. Imagining that we have N nuclei in the
system and the NG grid has been set on each Cartesian axis to represent
the spatial coordinate of one nucleus, we need to calculate the total energy
3 N
of the system at [NG ] spatial configurations in order to map out this BO
PES, which is an astronomical number in material simulations. Fortunately,
we do not need to know all these details in order to understand the key
properties of a material. Some simple, elegant, yet powerful computational
schemes can be designed to identify the atomic structures of the materials
at these states-of-interest in chemistry and physics, including the Crystal
structure AnaLYsis by Particle Swarm Optimization (CALYPSO) method
[7–12], the ab initio Random Structure Searching (AIRSS) method [13–15],
and the Generic Algorithm [16–18], etc. Nowadays, this is a very active
research topic in the field of computer simulations of material properties,
especially for studies of the material property under different pressures.
For people interested in research on this aspect, please see Refs. [7–18].
These methods are now very efficient in finding the stationary points on
the PES. By resorting to the phonon calculations, one can further identify
those properly defined local minima on the PES. And among them, the
one with the lowest energy (or in the most rigorous manner free energy)
is called the ground state. Other well-defined local minima are known as
metastable states.
With the local minima on the PES identified, by comparing their struc-
tures and the electronic structures associated with them to the available
experimental data, we can identify the ones consistent with experiments
and carry out further investigations from these structures. Depending on
what we want from our simulations, this includes TS searching, Monte-
Carlo sampling, and MD simulations together with their various extensions.
These concepts will be frequently used in our later discussions in this book.
Among them, TS searching is often a concept at the static level, i.e. a
concept in which we do not have to care about the thermal and quantum
nuclear effects related to the real fluctuations of the nuclei, although the
transition-state theory (TST) itself includes all these effects. In Monte-
Carlo and MD simulations, these effects must be considered. To be clear,
we start by explaining the simplest concept, i.e. TS searching, at the static
limit. As a matter of fact, such searches for the TS are of primary interest
in theoretical descriptions of the chemical reactions. What one needs to
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch01 page 11

Introduction to Computer Simulations of Molecules and Condensed Matter 11

do is to find the lowest barrier between two local minima, which represent
the reactant and product states, respectively. On a multi-dimensional PES,
searches for this TS between these two local minima, however, are still com-
putationally very difficult. Currently, there are various schemes on carrying
out such searches, including the constrained optimization (CO) [19], nudged
elastic band (NEB) [20–22], Dewar, Healy and Stewart (DHS) [23], Dimer
[24–26], activation-relaxation technique (ART) [27, 28] and one-side grow-
ing string (OGS) [29] as well as their various combinations. The purpose
of this section is to establish the concepts for material simulations related
to discussions in the later chapters. Therefore, we will not go into details
of such searches and instead, direct interested readers to the References
section.

1.2.4 Molecular Dynamics for the Thermal Effects

These TS searching methods help us to identify the structures of the TS and


the energy required for the corresponding chemical reactions to happen at
the static level. In a more realistic description of these processes, however,
statistical and dynamical effects, originating from the finite-temperature
thermal and quantum effects of the nuclei, must also be taken into account.
This leads us to the ultimate goal of material simulations as we have intro-
duced above, i.e. an accurate description of a real system’s propagation on
its BO PESs. Nowadays, a very popular method to carry out such research
is the so-called ab initio MD. In this method, prerequisite knowledge on
the starting structures of the system is needed in order for a reasonable
simulation to be carried out. It normally refers to a point on the PES close
to the region relevant to the problem of interest. During the propagation of
the nuclei, the electronic structures of the system are often described using
a certain ab initio method within either the DFT or traditional quantum
chemistry method framework. When the DFT is used, finite-temperature
contributions to the electronic free energy can often be addressed with Mer-
min’s finite-temperature version of the DFT [30–32]. With the electronic
structures determined quantum mechanically, the forces on each nucleus can
be calculated using the so-called Hellmann–Feynman theorem. The system
then propagates on the PES subjected to the forces computed “on-the-fly”
as the nuclei configuration evolves according to the Newton equation. We
note, however, that empirical potentials can also be used to describe the
interatomic interactions, which is outside the framework of the ab initio
simulation method. We focus on the ab initio methods and consequently,
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch01 page 12

12 Computer Simulations of Molecules and Condensed Matter

the ab initio MD in this book. But it is worth noting that the principles
underlying the propagation of the nuclei as introduced also apply to those
when empirical potentials are used.
There are several schemes for the ab initio MD simulations to be carried
out, depending on the ensemble used in the simulation. The simplest one is
the so-called micro-canonical ensemble. In these simulations, the total num-
ber of particles in the system (N ), the volume (V ), and the total energy
(E) of the system are kept constant. The system is completely isolated from
the environment. Therefore, this ensemble is also called N V E ensemble in
literature. Practically, however, experiments are often carried out at finite
temperatures. Therefore, the temperature (T ) of the system, instead of its
total energy (E), should be kept constant in a more realistic simulation of
such isothermal processes. This leads to the next class of the more often used
ensembles in molecular simulations, i.e. the canonical ensemble (N V T ). In
this ensemble, the control of the temperature is often achieved by coupling
the system to a heat bath (or thermostat) that imposes the desired tem-
perature on the system. Depending on how this is carried out, we have
different schemes for the ensemble to be simulated, e.g. the Andersen ther-
mostat, the Nosé–Hoover thermostat, the Nosé–Hoover chain thermostat,
or the Langevin thermostat, etc. All these different schemes help simulate
an N V T process and will be discussed in detail in Chapter 5.

1.2.5 Extensions of MD: Enhanced Sampling


and Free-Energy Calculations

The MD method as introduced above allows us to simulate the propagation


of a system using a finite-time step so that the total energy of this system is
conserved and the temperature is well in control. This means that the prop-
agation needs to be done with a time interval on the order of femtoseconds,
or even shorter. Processes of real chemical reactions, however, happen on a
much longer time-scale, e.g. seconds. This indicates that ∼1015 electronic
structures should be performed in order to see such a process happen. Plus,
integrating the equation of motion, this means an astronomical computa-
tional cost in condensed matter physics or chemistry. The physical origin
for this time-scale problem is from the morphology of PES as introduced
in the earlier discussion. In complex chemical systems, notably those in
condensed matter and in biology, the PESs typically have very complicated
features, evident in the presence of many low-energy local minima and large
barriers separating them. Transitions between these local minima are rare
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch01 page 13

Introduction to Computer Simulations of Molecules and Condensed Matter 13

events and long-time simulations are required to evidence them. Meanwhile,


complex morphology of the PES also indicates that there are shallow local
minima separated by low barriers, and transitions between them happen
much faster. Consequently, it is fair to say that one of the greatest chal-
lenges in MD simulations now is in numerically reproducing such a wide
distribution of time scales in complex systems. More practical schemes, in
which a more efficient sampling of the phase space can be guaranteed, are
highly desired.
During the last few decades, great efforts have been made to accelerate
the exploration of the multi-dimensional PES in complex systems. A num-
ber of methods, such as adaptive umbrella sampling [33], replica exchange
[34–37], metadynamics [38, 39], multi-canonical simulations [40, 41], confor-
mational flooding [42], conformational space annealing [43], and integrated
tempering sampling [44–47] etc., have been proposed, each having its own
strengths and weaknesses. For example, in the adaptive umbrella sampling,
metadynamics, and configurational flooding methods, reaction coordinates
need to be selected before the simulations can be carried out. Conversely,
in replica exchange and multi-canonical simulations, different temperature
simulations are required. This causes the trajectories obtained to lose their
dynamical information. In Chapter 6, we will take the umbrella sampling,
adaptive umbrella sampling, metadynamics, and integrated tempering sam-
pling methods as examples to show how they work in practical simulations.
Another important extension of the MD simulation technique concerns
the phase behavior of a given substance, in particular, transitions between
two competing phases. The simplest example is the melting of a solid into
liquid. In this example, the transition between the two competing phases is
first-order and their transition curve can be calculated from the principle
that at coexistence, the Gibbs free energies of the two phases are equal.
In order to describe such an equality between two free energies, a method
from which this free energy can be calculated should be available. In Chap-
ter 6, we will also use the thermodynamic integration method as an example
to show how this is done in practice [48–58].

1.2.6 Path Integral Simulations for the Quantum Nuclear


Effects

In standard MD simulations, the nuclei are often treated as classical point-


like particles and propagate according to the Newton equation. Statistically
and dynamically, the quantum nature of the nuclei is completely neglected.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch01 page 14

14 Computer Simulations of Molecules and Condensed Matter

This classical treatment is normally a good approximation since the nuclear


masses are several orders of magnitude larger than that of the electron.
However, it needs to be pointed out that there are situations when the
quantum nuclear effects (QNEs) must be accounted for, especially in sys-
tems where light elements such as hydrogen are involved. As a matter of
fact, it has long been recognized that the statistical properties of hydrogen-
bonded systems such as water heavily depend on which isotope of hydro-
gen is present. Taking the melting/boiling temperature of normal water
(composed of H2 O) and heavy water (composed of D2 O) as an example,
this value in heavy water is 3.8◦ C/1.4◦ C higher than that of normal water
under the same ambient pressure. In classical statistical mechanics, it is
well known that the statistical thermal effects do not depend on the iso-
tope. Therefore, in descriptions of such effects, the quantum nature of the
nuclei must also be taken into account.
One natural solution to describe such QNEs from the theoretical per-
spective is to construct a high-dimensional Schrödinger equation for the
many-body entity of the nuclei. By solving the eigenstate wave functions
and eigenvalues of this Schrödinger equation, one obtains both the sta-
tistical and dynamical properties of the system under finite temperature.
However, it is worth noting that this many-body Schrödinger equation has
a notorious scaling problem. Although we acknowledge that it is a rigorous
method which has been extremely successful in descriptions of the gas phase
reactions [59–62], due to the scaling problem associated with both mapping
the ab initio high-dimensional PESs and solving the Schrödinger equation,
its application is seriously limited to systems less than ∼6 atoms. When
the system gets bigger, a practical method must be resorted to.
Thanks to the development of the path integral representation of quan-
tum mechanics by Feynman [63–67], a framework where these QNEs can be
described in a practical manner was systematically presented by Feynman
and Hibbs in their seminal book in 1965 [67]. In Ref. [67, Chapter 10], when
the statistical mechanics was discussed, the partition function of a quantum
system was related to the partition function of a classical polymer using the
method of path integral. This partition function of a classical polymer can
be simulated using finite-temperature sampling methods such as MD or
Monte-Carlo. Therefore, from the statistical perspective, the framework of
such simulations, which we later called path integral molecular dynamics
(PIMD) or path integral Monte-Carlo (PIMC), were already laid out. These
PIMD and PIMC methods had experienced a golden time in the 1980s and
1990s, first purely from the statistical perspective [68–73], and then with
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch01 page 15

Introduction to Computer Simulations of Molecules and Condensed Matter 15

extensions to the dynamical regime [74–84]. After 2000, a slight revision


of the PIMD method, which was shown to be very successful in describ-
ing dynamical propagation and called ring-polymer molecular dynamics
(RPMD), was also proposed [85–93]. In Chapter 7, we will give a detailed
explanation of these methods.

1.3 Layout of the Book


Following the above introduction, we will use six chapters to introduce
some important methods for the present computer simulations of molecules
and condensed matter, covering the regimes of both electronic structures
and molecular motion. Theories underlying the ab initio electronic structure
calculations are used as the basis for these methods. Therefore, we will start
with these and introduce some traditional quantum chemistry methods and
the density-functional theory in Chapter 2. For real implementations of
these methods, pseudopotentials are often used to save the computational
cost, and a set of basis functions must be chosen to expand the electronic
wave functions. Based on this consideration, the principles underlying the
construction of a pseudopotential are explained and a frequently used basis
set for all-electron calculations, i.e. the linearized augmented plane waves
(APWS), is introduced in Chapter 3. Then, we go beyond the ground state
electronic structures and introduce the many-body perturbation theory
(or otherwise called Green’s function method) as an example of the ab initio
descriptions of the electronic excitations in Chapter 4, with a special focus
on the popular GW approximation and its all-electron implementation.
From Chapter 5, we shift our attention to the nuclear motion. The
MD method will be introduced in this chapter, with discussions covering
its original micro-canonical form and extensions to simulations of other
ensembles. One shortcoming of the standard MD method is that the multi-
timescale problem in complex systems is beyond the scope of such simula-
tions. Therefore, in Chapter 6, we will take the umbrella sampling [94–96],
adaptive umbrella sampling [33], metadynamics [38, 39], and integrated
tempering sampling [44–47] methods as examples to show how enhanced
sampling works in practice. A popular method for the calculation of the
free energy, i.e. the thermodynamic integration, will also be introduced.
The discussion in Chapters 5 and 6 assumes that the nuclei are classical
point-like particles. In practice, they do have a quantum nature. To include
the quantum features of the nuclei in the molecular simulations, the path
integral and related methods are explained in Chapter 7. In addition to
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch01 page 16

16 Computer Simulations of Molecules and Condensed Matter

the original scheme of the path integral sampling, as given in the classical
textbooks like Refs. [67, 97], a more or less complete explanation for the
computational details in the statistical PIMD simulations, as well as its
extensions to the zero-point energy calculations, dynamical calculations,
and free energy calculations are also given. We hope that this limited, yet
to a certain extent organized, introduction can provide the readers with
an idea about what molecular simulations are and how we can carry out a
molecular simulation in practice. We acknowledge that due to the limitation
of our time and knowledge, lack of clarity and mistakes may exist at several
places in the present manuscript. We sincerely welcome criticism and sug-
gestions (please email [email protected]), so that we can correct/improve in
our next version.z
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch02 page 17

2
Quantum Chemistry Methods
and Density-Functional Theory

The purpose of this chapter is to illustrate the basic principles underlying


the current standard methods to accurately solve Eq. (1.3), which repre-
sents the main task for the ab initio calculations of the electronic structure
in real poly-atomic systems. The different schemes to achieve this goal can
be classified into two main categories: the wave function-based methods,
traditionally known as quantum chemistry methods, and methods based
on the density-functional theory (DFT). Both of these two methods have
achieved great success within the last half-century, as evidenced by the
Nobel Prize of Chemistry being awarded to W. Kohn and J. Pople in
1998. Hereinafter, we will address both of these schemes, with a special
emphasis on DFT. Knowledge of their successes and limitations comprises
the required point of departure for the discussions on the Green’s function
method afterwards.

2.1 Wave Function-Based Method


The variational principle, which can be viewed as another form of the many-
body Schrödinger equation, states that any state vector of the electrons for
which the average energy, defined as
Φ|Ĥe |Φ
Ee [Φ] ≡ , (2.1)
Φ|Φ
is stationary corresponds to an eigenvector of Ĥe , with eigenvalue Ee . Fur-
thermore, for any state of the system, the corresponding average energy

17
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch02 page 18

18 Computer Simulations of Molecules and Condensed Matter

satisfies

Ee [Φ]  E0 , (2.2)

where E0 is the ground-state energy of the electronic system.


The essence of the wave function-based methods involves obtaining the
stationary solutions of Eq. (2.1) within a trial-function space. The accuracy
of the method is naturally determined by the choice of this trial-function
space. By taking more sophisticated approximations for the trial many-
body wave function of these electrons, the accuracy can be systematically
improved. Unfortunately, at the same time its computational cost could
increase. In this section, we will introduce some of these approximations in
order of increasing complexity.

2.1.1 The Hartree and Hartree–Fock Approximations

Since the many-body wave function can always be written as a linear com-
bination of products of one-particle wave functions, the simplest possi-
ble ansatz, first proposed by Hartree [98], for the many-body electronic
wave function is to assume it as the product of the single-particle wave
functions:

Φ(r1 , r2 , . . . , rN ) = ϕ1 (r1 )ϕ2 (r2 ) . . . ϕN (rN ). (2.3)

Substituting this trial wave function into Eq. (2.1), making use of the
variational principle, the many-body problem of this electronic system is
mapped onto a set of single-particle, Schrödinger-like, equations

ĥi ϕi (ri ) = i ϕi (ri ), (2.4)

with the Hamiltonian given by


 
∇2
ĥi = − + Vext (r) + V H (r) + ViSIC (r) . (2.5)
2
The solutions of Eq. (2.4) are coupled through the Hartree potential

H n(r )
V (r) = dr , (2.6)
|r − r |
which depends on the electron density, defined as
N

n(r) = |ϕi (r)|2 . (2.7)
j=1
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch02 page 19

Quantum Chemistry Methods and Density-Functional Theory 19

The last term in Eq. (2.5) corrects for the interaction of the electron with
itself, included in the Hartree potential.

|ϕi (r )|2 
ViSIC (r) = − dr . (2.8)
|r − r |

Thus, the set of Eqs. (2.4)–(2.7) has to be solved self-consistently. Solving


these equations, the ground-state energy of the electronic system can be
calculated as

N
 N N 
1  |ϕi (r)|2 |ϕj (r )|2
Ee = i − drdr . (2.9)
i=1
2 i=1 |r − r |
j=i

With these, the principle underlying a calculation within the Hartree


approximation for the many-body interactions is clear. In essence, it maps
the many-particle problem into a set of independent particles moving in the
mean Coulomb field of the other particles.
As simple and elegant as it looks, the Hartree approximation presents
one of the first practical schemes for the calculation of a many-body elec-
tronic system’s total energy. However, it needs to be pointed out that
there is a major drawback in this method; namely, the many-particle wave
function of the electrons does not obey the Pauli principle. The Pauli
principle forbids two Fermi particles from occupying the same quantum
mechanical state and therefore prevents electrons with the same spin from
getting close to each other. It is a physical property of all many-body
Fermi systems. Lacking this feature, the Hartree approximation generally
underestimates the average distances between electrons. Consequently, the
average Coulomb repulsion between them, as well as the total energy, are
overestimated.
In order to fulfill the Pauli principle, the many-particle wave function
has to be anti-symmetric among the exchange of two particles with the
same spin. The simplest ansatz for a many-particle wave function, obeying
Pauli’s principle, is obtained by the anti-symmetrized product, known as
the Slater determinant. This improved ansatz, proposed by Fock [99], is
known as the Hartree–Fock approximation. For non-spin-polarized systems
in which every electronic orbital is doubly occupied by two electrons with
opposite spins (closed shell), this method can be introduced in a very simple
form. For an N -electron system, the Slater determinant representing the
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch02 page 20

20 Computer Simulations of Molecules and Condensed Matter

many-particle wave function is written as

 
 ϕ1 (r1 , σ1 ) ϕ1 (r2 , σ2 ) · · · ϕ1 (rN , σN ) 
 
 
1  ϕ2 (r1 , σ1 ) ϕ2 (r2 , σ2 ) · · · ϕ2 (rN , σN ) 
Φ(r1 , σ1 , . . . , rN , σN ) =   ,
(N )!  
.. .. ..
. . . 
 
ϕN (r1 , σ1 ) ϕN (r2 , σ2 ) · · · ϕN (rN , σN )
(2.10)

where σi represents the spin coordinate of the ith electron.


Substituting this equation into Eq. (2.1) and making use of the orthog-
onality of the space orbitals and spin states, one arrives at a set of equations
for the single-particle orbitals:

 
∇2
− + Vext (r) + V H (r) ϕi (r, σ)
2
  ϕ∗j (r , σ  )ϕj (r, σ)
− dr |
δσ,σ ϕi (r , σ  ) = i ϕi (r, σ). (2.11)
j
|r − r

The only difference from Eq. (2.5) is the last term, representing the
exchange interaction between electrons, which is known as the Fock oper-
ator. It is non-local and affects only the dynamics of electrons with the
same spin. Note that when j = i, the exchange term equals the correspond-
ing term in the Hartree potential. This indicates that the Hartree–Fock
approximation is self-interaction-free.
In terms of the single-particle orbitals, the total energy of the system
can be written as

N
 N 
N 

Ji,j Ki,j
Ee = Φ|Ĥe |Φ = Hi,i + − , (2.12)
i i j
2 4

where
  
1
Hi,i = ϕ∗i (r, σ) − ∇2 + Vext (r) ϕi (r, σ)drdσ, (2.13)
2
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch02 page 21

Quantum Chemistry Methods and Density-Functional Theory 21


1
Ji,j = ϕ∗i (r, σ)ϕi (r, σ) ϕ∗ (r , σ  )ϕj (r , σ  )drdσdr dσ  ,
|r − r | j
(2.14)

1
Ki,j = ϕ∗i (r, σ)ϕj (r, σ) ϕ∗ (r , σ  )ϕi (r , σ  )drdσdr dσ  .
|r − r | j
(2.15)

Here, Hi,i is the non-interacting single-particle energy in the external field.


Ji,j is the “classical” Coulomb interaction between electrons in the states
i and j and Ki,j is the exchange interaction between them. The first two
terms also appear in the Hartree method; the last term is introduced by
imposing the Pauli principle on the many-body wave function and tends to
reduce the total energy.
For spin-polarized systems, the method becomes more cumbersome.
Since the number of electrons with each spin is not balanced anymore, the
contribution from the Fock operator will be different for different spins.
Consequently, the single-particle eigenvalues and eigenfunctions will also
be different [100].

2.1.2 Beyond the Hartree–Fock Approximation

In the Hartree–Fock approximation, the electron–electron interaction is


treated by means of a time-independent average potential. The fulfillment
of Pauli’s principle imposes a “static” correlation of the position of elec-
trons with the same spin (exchange hole). However, it completely neglects
dynamical effects due to the Coulomb interaction. Such dynamical effects
due to Coulomb interaction arise from the fact that the movement of a
given electron affects and is affected by, or in other words, “is correlated
with”, the movement of the other particles. Lacking this feature, many
key physical processes representing the many-body interactions of the elec-
tronic system will be absent. Physically, this is the main limitation of the
Hartree–Fock approximation. For atoms and small molecules, where these
dynamical effects are the smallest and the Hartree–Fock approximation
works best, this limitation leads to errors of around 0.5% in the total energy.
For example, in a carbon atom, where the total energy is around 1000 eV,
this corresponds to 5 eV, which already equals the order of magnitude of
a chemical single-bond energy. Thus, to obtain a reliable description of
chemical reactions, more sophisticated approximations are required.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch02 page 22

22 Computer Simulations of Molecules and Condensed Matter

Within the framework of the wave function-based methods, this can be


achieved by making more elaborate approximations for the many-body wave
function. Among the so-called post-Hartree–Fock methods, the configura-
tion interaction (CI) method [101], the Møller–Plesset (MP) perturbation
theory [102], and the coupled-cluster (CC) method [103, 104] have achieved
great success in past years. As a general shortcoming of all these methods,
their application is limited to atoms and small molecules, due to the scaling
of computational costs with the system size.

2.2 Density-Functional Theory


As discussed in the previous section, the wave function-based post-Hartree–
Fock methods can be very accurate in describing the properties of a many-
body electronic system. However, the scaling behavior with respect to the
system size is very poor. Taking the simplest post-Hartree–Fock method,
i.e. the MP2 method, as an example, the usual scaling is already N 5 , with
N representing the number of electrons. Such scaling behavior has seri-
ously limited its application to large poly-atomic systems. Starting from
the 1980s, a method stemming from a completely different origin, i.e. the
DFT, has achieved great success in the electronic structure calculations
of molecules and condensed matter, largely due to a very good balance
between computational cost and accuracy. This can be evidenced by the
fact that despite being a physicist, W. Kohn was awarded the Nobel Prize
of Chemistry in 1998. In the following, we will introduce the main principles
underlying this theory and its main limitations.
The formalism of the DFT was introduced by Hohenberg and Kohn
(HK) in 1964 [105]. In 1965, Kohn and Sham (KS) [106] presented a scheme
to approximately treat the interacting electron system within this formal-
ism. It is currently the most popular and successful method for study-
ing ground-state electronic structures. Although far from a panacea for
all physical problems in this domain, very accurate calculations can be
performed with computational costs comparable to that for the Hartree
method. In this chapter, we will present some major components of this
theory. The discussion begins with its precursor: the Thomas–Fermi theory.

2.2.1 Thomas–Fermi Theory

The Thomas–Fermi theory was independently proposed by Thomas and


Fermi in 1927 [107–109]. In its original version, the Hartree method was
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch02 page 23

Quantum Chemistry Methods and Density-Functional Theory 23

reformulated in a density-based expression for an electron gas with slowly


varying density. The kinetic energy is locally approximated by that of a
non-interacting homogeneous electron gas with the same density.
Later, Dirac introduced the exchange term into the model using the
same local approximation [110]. The total energy of the electronic system,
including the exchange term, is written as
Ee [n] = T + U ext + U H + U X
 
3
= (3π 2 )2/3 drn(r)5/3 + drVext (r)n(r)
10
 
 1/3 
1  n(r)n(r ) 3 3
+ drdr − drn(r)4/3 .
2 |r − r | 4 π
(2.16)
This expression gives the exact energy for the homogeneous electron gas in
the Hartree–Fock approximation. It is also a good approximation in systems
with slowly varying electron densities.
Based on the formulism of the total energy, within this Thomas–Fermi
theory, the density in a real system is obtained by minimizing Ee [n] under
the constraint of particle number conservation

 
δ Ee − μ n(r)dr − N = 0. (2.17)

Substituting Eq. (2.16) into Eq. (2.17), one obtains the Thomas–Fermi
equation:
1
(3π 2 )2/3 n(r)2/3 + Vext (r) + V H (r) + V x (r) − μ = 0, (2.18)
2
where the Hartree potential, V H (r) is the same as defined in Eq. (2.6), and
the exchange potential is given by
 1/3
x 3
V (r) = − n(r) . (2.19)
π
From Eq. (2.18), for a certain external potential V ext (r) and chemical
potential μ, one can obtain the electron density of this system and conse-
quently, the total energy (from Eq. (2.16)). However, in practice, as it is
based on too-crude approximations, lacks the shell character of atoms and
binding behavior, the Thomas–Fermi theory automatically fails in providing
a proper description of real systems.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch02 page 24

24 Computer Simulations of Molecules and Condensed Matter

2.2.2 Density-Functional Theory

In 1964, HK formulated two theorems, which formally justified the use of


density as the basic variable in determining the total energy of an inter-
acting many-body system [105]. The first theorem proved the existence of
a one-to-one correspondence between the external potential Vext (r), the
ground-state many-body wave function Φ, and the ground-state density
n(r). Thus, the total energy of a system, which is a functional of the many-
body wave function Φ, can also be reformulated as a functional of the
density as

Ee [n(r)] = Vext (r)n(r)dr + F [n(r)]. (2.20)

F [n(r)] contains the potential energy of the electronic interactions and the
kinetic energy of the electrons. It is a universal functional independent of
the external potential. Unfortunately, its exact form is unknown.
Since the expression for the Hartree energy as a functional of the density
is known, the functional F [n(r)] in Eq. (2.20) can be further decomposed
as

F [n(r)] = U H [n(r)] + G[n(r)], (2.21)

where the expression of the Hartree energy U H [n(r)] is already given by the
third term on the right-hand side of Eq. (2.16). Like F [n(r)], G[n(r)] is an
unknown universal functional of the density, independent of the external
potential. The total energy can then be written as
 
1 n(r)n(r )
Ee = Vext (r)n(r)dr + drdr + G[n(r)]. (2.22)
2 |r − r |

The second theorem proves that the exact ground-state energy of the
electronic system corresponds to the global minimum of Ee [n(r)], and the
density n(r) that minimizes this functional is the exact ground-state density
n0 (r).
These two theorems set up the foundation for the concept of “density
functional”. However, a practical scheme which can be used to calculate
the density was still absent. This seminal contribution was given by KS. In
1965, they proposed a scheme to calculate the G[n(r)] in Eq. (2.22) [106].
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch02 page 25

Quantum Chemistry Methods and Density-Functional Theory 25

In this scheme, one can decompose this G[n(r)] into two parts

G[n] = T s [n] + E xc [n]. (2.23)

The first term is the kinetic energy of a non-interacting system with the
same density. The second term is the exchange–correlation (XC) energy.
Minimizing the total energy in Eq. (2.20) under the constraint of par-
ticle number conservation (Eq. (2.17)), one gets



δT s [n]
δn(r) Veff (r) + − μ dr = 0, (2.24)
δn(r)

where μ is the chemical potential and


n(r )
Veff (r) = Vext (r) + dr + V xc (r). (2.25)
|r − r |

V xc (r) is called the XC potential, given by

δE xc [n]
V xc (r) = . (2.26)
δn(r)

Assuming a set of non-interacting particles with the same density,

N

n(r) = |ϕi (r)|2 , (2.27)
i=1

Eq. (2.24) is equivalent to



1
− ∇2 + Veff (r) ϕi (r) = i ϕi (r). (2.28)
2

Thus, the KS scheme maps the complex, interacting electronic systems


into a set of fictitious independent particles moving in an effective, local
potential. Since this effective potential depends on the density, Eqs. (2.25),
(2.27), and (2.28) have to be solved self-consistently.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch02 page 26

26 Computer Simulations of Molecules and Condensed Matter

The total energy in the KS scheme is given by


  
1 n(r)n(r )  xc
Ee [n(r)] = i − |
drdr + E [n(r)] − n(r)V xc (r)dr.
occ
2 |r − r
(2.29)
We note that in molecular simulations, another form of this functional is
also used:
  
1
Ee [n(r)] = drϕ∗i (r) − ∇2 ϕi (r) + n(r)Vext (r)dr
occ
2

1 n(r)n(r )
+ drdr + E xc [n(r)] . (2.30)
2 |r − r |
These two equations are equivalent, as evidenced by inputting Eq. (2.28)
to Eq. (2.29).
One limitation of the discussion above is that it is based on the orig-
inal papers [105, 106] and is therefore restricted to the zero-temperature,
non-degenerate, non-spin-polarized, and non-relativistic cases. Extensions
to the finite-temperature calculations can be found in Refs. [30–32], to the
spin-polarized system in Refs. [111–113], and to the degenerate system in
Refs. [114–118]. For a comprehensive discussion on these extensions, please
see Ref. [119, Chapter 3]. The inclusion of relativistic effect is addressed in
Ref. [120]. Further extension of the theory to superconductors may be found
in Refs. [121, 122]. Detailed discussion about the v-representability and
related questions can be found in Refs. [114, 117]. An excellent review on
the formal justification of the theory can be found in Ref. [119, Chapter 2].

2.2.3 Exchange–Correlation Energy

In the KS scheme, all the complexity of the many-body interaction is


put into the XC energy E xc [n(r)]. Unfortunately, the exact expression of
this functional is unknown. In addition, different from the wave function-
based post-Hartree–Fock method where the accuracy can be systematically
improved by taking more complicated forms of the wave functions, a sys-
tematic series of approximations converging to the exact result is absent in
the DFT. While semiempirical approaches allow one to obtain very precise
results within the sample space fitting, their physical origin can be some-
times obscure and their precision outside that space unpredictable. In order
to remain within the first-principles framework, the most universal and, to
some extent, systematic scheme is the “constraint satisfaction” approach
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch02 page 27

Quantum Chemistry Methods and Density-Functional Theory 27

[123]. These “constraints” consist of exact properties that the XC functional


can be proven to fulfill. In this approach, the approximations to the XC
functional are assigned to various rungs of the “so-called” Jacob’s ladder
[123] according to the number of ingredients they contain. The best non-
empirical functional for a given rung is constructed to satisfy as many exact
theoretical constraints as possible while providing satisfactory numerical
predictions for real systems. Increasing the number of ingredients allows the
satisfaction of more constraints, thus increasing, in principle, the accuracy.
The simplest approximation for E xc [n(r)] is the local-density approxi-
mation (LDA), proposed in the original paper of KS. It reads

xc
E [n(r)] = n(r)xc (n(r))dr, (2.31)

where xc (n(r)) is the XC energy per particle of a homogeneous electron


gas with the same density n [106]. The exchange contribution to xc can be
obtained analytically (Ref. [124]), giving
 1/3
3 3n
x (n) = − . (2.32)
4 π
The correlation contribution has to be calculated numerically. In 1980,
Ceperley and Alder performed a set of quantum Monte-Carlo (QMC) calcu-
lations for the homogeneous electron gas with different densities [125]. The
correlation term of the LDA functionals used nowadays rely on different
parametrizations of these results. One of the most used parametrizations is
that proposed by Perdew and Zunger in 1981 [126]. The LDA is exact for
the homogeneous electron gas and expected to be valid for inhomogeneous
systems with slowly varying density. A large number of calculations, how-
ever, have also shown that it works remarkably well for several real systems
with strongly inhomogeneous electron densities [113].
Naturally, are improvement to the LDA is the inclusion of the depen-
dence on the gradient of the density. This approach gave rise to the gener-
alized gradient approximations (GGAs) introduced in the late 1980s [127,
128]. The XC energy is written as

xc
E [n(r)] = n(r)xc (n(r), ∇n(r))dr. (2.33)

The so-called PBE functional [129] is the most commonly used non-
empirical GGA functional nowadays. It is an improvement to the LDA
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch02 page 28

28 Computer Simulations of Molecules and Condensed Matter

for many properties, for example, the geometries and ground-state energy
of molecules [129–132].
Further climbing Jacob’s ladder, meta-GGA’s functionals are found in
the third rung. Its description, as well as prescriptions for the fourth and
fifth rungs can be found in Ref. [133].

2.2.4 Interpretation of the Kohn–Sham Energies

The KS eigenvalues appear as formal Lagrange multipliers in Eq. (2.17)


and correspond to the eigenstates of the fictitious, non-interacting KS
particles. A crucial question is therefore whether they have any physical
meaning.
Janak’s theorem, together with Slater’s transition-state theory (TST),
provide a justification for the interpretation of the highest occupied state’s
eigenvalue as the ionization energy in extended systems [134, 135]. Later,
this justification was extended to the finite systems [136]. For the other
states, the KS eigenvalue, when calculated at half occupation, gives a good
estimation of the corresponding total energy difference [137]. When the KS
potential is continuous with respect to the electron density, these energy
differences can even be approximated by the KS eigenvalues calculated with
full occupation. However, since the exact form of E xc [n] is unknown, the
comparison between the KS eigenvalues and experiments always relies on
approximations of the XC potential.
For the LDA and GGAs, the result of such a comparison can be easily
summarized: the work function and band structures in metals are found
to be reasonably well described [138, 139], for semiconductors and insu-
lators, universal underestimations of 50–100% for the fundamental band
gaps are found. A well-known problem of the LDA is the self-interaction.
By using the exact exchange optimized effective potential (OEPx), which
is self-interaction-free, these band gaps are improved [140–144].
On the other hand, the fundamental band gap is determined from the
ground-state energy of the N − 1, N , and N + 1-electron systems. The
KS band gap is calculated as the difference between the lowest unoccu-
pied and highest occupied KS eigenvalues in an N -electron system. It was
proven that they differ by a term given by the discontinuity of the XC
potential [145–147], i.e.

Eg = EgKS + Δxc , (2.34)


December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch02 page 29

Quantum Chemistry Methods and Density-Functional Theory 29

where
Δxc = lim+ [V xc |N +δ − V xc |N −δ ] . (2.35)
δ→0

In both LDA and GGA, this discontinuity is zero. In OEPx, it is not


zero. Adding this term to the KS band gaps through the above equa-
tions, Grüning et al. have shown that fundamental band gaps similar to
the Hartree-Fock method can be obtained [148]. The exact XC functional
would allow the calculation of the band gap through Eq. (2.34). Neverthe-
less, for the description of the excited-state properties in general, a different
theoretical approach is required. The standard treatment nowadays is the
Green’s function method we are going to address in Chapter 4.
b2530   International Strategic Relations and China’s National Security: World at the Crossroads

This page intentionally left blank

b2530_FM.indd 6 01-Sep-16 11:03:06 AM


December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch03 page 31

3
Pseudopotentials, Full Potential,
and Basis Sets

As shown in the previous chapter, in the Kohn–Sham (KS) scheme to


the density-functional theory (DFT), the many-body electronic problem is
reduced to an independent particle problem (Eq. (2.28)) under the action
of an effective, density-dependent potential (Eqs. (2.25) and (2.26)). Any
numerical implementation of this scheme to real poly-atomic systems has
to deal with two extremely different regimes: the core states tightly bound
to the nucleus, and the delocalized valence states, especially in the inter-
stitial region. The former are represented by localized wave functions of
atomic character, and their role in the bonding of the system is mainly
concerned with the screening of the nuclear potential. The valence states
play a determining role in the bonding, which reciprocally determines the
characteristics of the wave functions of these states, going from localized
states in ionic systems to fully itinerant ones in simple metals. However,
itinerant the valence wave function may be, it also presents a fast oscillating
behavior in the region close to the nuclei. In particular for periodic systems,
where the reciprocal space representation is more efficient, plane waves are
a natural basis set for the expansion of the single-particle wave functions.
However, they are inefficient in representing both the strongly localized
core states and the rapid oscillations of the valence wave functions in the
nuclear region.
During the past decades, different strategies have been developed to
address these coexistent regimes. They can be arranged in two big groups:
all-electron and pseudopotential methods. The all-electron methods rely on
the use of more sophisticated basis functions for the expansion of the wave
functions, while keeping the potential genuine. These basis functions can
address the oscillation of the wave functions in the nuclear region, as well

31
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch03 page 32

32 Computer Simulations of Molecules and Condensed Matter

as the core states, at reasonable computational cost. The pseudopotential


methods, on the other hand, replace the strong nuclear potential and the
core contribution to the Hartree and exchange–correlation (XC) potentials
by an artificial effective ionic potential, namely the pseudopotential, which
is designed to be much softer than the full potential in the atomic region. In
this way, the valence wave functions behave smoothly in the nuclear region
and can be efficiently expanded in plane waves. Only the valence electrons
are treated explicitly.
In the following, we will give a brief introduction to the pseudopotential
method using one of its classes, the norm-conserving pseudopotentials, as
an example. Our focus is on the principles underlying the construction of
a pseudopotential. Concerning the all-electron method, we will also intro-
duce an important basis set, i.e. the full-potential-(linearized) augmented
plane waves plus local orbitals (FP-(L)APW+lo). All-electron calculation
employing this basis set is currently believed to be the most accurate form
of DFT calculation in periodic systems.

3.1 Pseudopotential Method


The main idea behind the pseudopotential method is that, as long as the
core electrons are tightly bound, they do not participate actively in the
bonding process. Thus, the strong ionic potential, including contributions
from the nucleus and the core electrons, can be replaced by an angular-
dependent pseudopotential constructed from the free atom of the corre-
sponding element [149]. In this way, we can only include the valence states
explicitly in descriptions of the chemical bonding in the poly-atomic sys-
tem, which significantly reduces the computational cost. Inside the core
region, the pseudopotential is designed to be much softer than the ionic one.
Outside the core region, it is required that the corresponding pseudowave
function equals its all-electron counterpart in order to obtain the correct
behavior over a wide range of chemical environments (transferability).
A pseudopotential fulfilling the above-mentioned prerequisites can be
generated arbitrarily in many ways. The most used one is the “norm-
conserving” scheme originally proposed by Hamann et al. [150] and later
applied to elements from H to Pu by Bachelet et al. [151]. In this scheme,
the integral of the pseudocharge-density inside the core region is required
to agree with the all-electron one. This condition guarantees that the
electrostatic potential produced outside the core radius is equal in both
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch03 page 33

Pseudopotentials, Full Potential, and Basis Sets 33

cases. Furthermore, the energy dependence of the scattering properties of


the pseudopotential is of the second order and can be ignored without affect-
ing the transferability. Nevertheless, some cases, e.g. O 2p or Ni 3d orbitals,
have been found to be impossible to construct a pseudowave function much
softer than its all-electron counterpart for [152]. This is due to the fact that
in 2p and 3d orbitals, the radial function is already nodeless. Therefore,
when applying the constraint that the normal of the radial function does
not change, there is no way to make it softer.
Such a drawback indicates that for systems containing 2p and 3d
valence states, within the framework of the norm-conserving pseudopoten-
tials, a large cut-off radius for the plane wave basis set is still required to
describe behaviors of these states in the core region. As a matter of fact, in
the 1980s and 1990s, this drawback has seriously hindered the application of
the pseudopotential method-based DFT simulations to large systems, due
to the limited computing power available at that time. In the early 1990s,
a successful attempt to circumvent this limit was proposed by Vanderbilt
et al. in Refs. [152, 153]. The corresponding method is called ultra-soft pseu-
dopotentials, in which the norm-conserving constraint is lifted. Nowadays,
due to the advances in computing power, the norm-conserving pseudpoten-
tials are still one of the most frequently used pseudoptentials in practical
simulations. In this book, we restrict ourselves to this norm-conserving
scheme for a clear illustration of the principles underlying the construction
of a pseudopotential. Readers interested in ultra-soft pseudopotentials are
directed to Refs. [152, 153].

3.1.1 Generation of the Pseudopotential

The initial step for generating a norm-conserving pseudopotential is to per-


form an all-electron DFT calculation for the free atom. This corresponds to
obtaining a self-consistent solution of the radial, Schrödinger-like equation
 
1 d2 l(l + 1)
− + + V (r) −  l ul (l , r) = 0, (3.1)
2 dr2 2r2

where V (r) is equivalent to the effective potential Veff in Eq. (2.25). It


should be obtained through a self-consistent solution of Eqs. (2.25)–(2.27),
with

ul (l , r)
ϕl,m (r) = Yl,m (r̂).
r
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch03 page 34

34 Computer Simulations of Molecules and Condensed Matter

Figure 3.1 Pseudovalence wave functions in comparison with the all-electron ones.
S atom is taken as the example. The 3s, 3p and 3d all-electron and pseudopotential
states are shown; the 3d state is an unbound state. The two wave functions agree
with each other outside the matching radius, while the pseudo one is much softer
inside this radius. Their norms are equal. The figure was generated using the fhi98PP
pseudopotential program [154], which is open access and available at http://th.fhi-
berlin.mpg.de/th/fhi98md/fhi98PP/.

After the self-consistent solution of these equations, the two full-potential


functions V (r) and ul (l , r) will be the key quantities from which the pseu-
dopotential is generated.
For each angular momentum number “l” (from now on, we call this
a “channel”), a cut-off radius (rlc ) is chosen and the pseudovalence radial
wave function ups ps
l (l , r) is derived from its all-electron counterpart ul (l , r)
with the following minimal constraints:
(i) The pseudovalence state has the same eigenvalue as the all-electron
one (psl = l ).
ps
(ii) ul (l , r) equals ul (l , r) beyond the cut-off radius (designated by
rmatch in Fig. 3.1).
(iii) ups
l (r) is nodeless. In order to obtain a continuous pseudopotential
regular at the origin, it is also required to be twice differentiable and
l+1
satisfy limr→0 ups l ∝r .
(iv) The pseudovalence radial wave function is normalized (the norm-
conserving constraint) which, together with (ii) implies
 r  r
ps ps
|ul (l ; r) | dr ≡
2
|un,l (n,l ; r) |2 dr for r  rlc . (3.2)
0 0
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch03 page 35

Pseudopotentials, Full Potential, and Basis Sets 35

Once the pseudowave function is obtained, one can construct the


screeneda pseudopotential Vlps,scr (r), which acts as the effective potential
on the pseudovalence state, by inverting the radial Schrödinger equation,
leading to

l (l + 1) 1 d2 ps
Vlps,scr (r) = ps
l − + ps u (l , r) . (3.3)
2r 2 2ul (r) dr2 l

In the last step, the pseudocharge-density ñ0v is obtained by

occ  ps


 ul (r) 2
ñ0v (r) =  
 r  , (3.4)
l

and the pseudopotential is unscreened by subtracting the Hartree and XC


potential corresponding to this pseudocharge by

Vlps (r) = Vlps,scr (r) − V H [ñ0v ; r] − V xc [ñ0v ; r]. (3.5)

So defined, the ionic pseudopotential generated from Eq. (3.5) includes


all the interactions of the valence electrons with the ion on the DFT
level and is much softer in the core region than its all-electron counter-
part (Fig. 3.2). The distinct procedures for generating a norm-conserving
pseudopotential differ only in the way the pseudovalence radial wave func-
tions (Fig. 3.1) are designed and the constraints they are required to fulfill.
In the Hamann scheme [150], an intermediate pseudopotential Vlps,i (r)
is constructed by cutting off the singularity of the full-potential V (r) at the
nucleus:
    
r r
Vlps,i (r) = V (r) 1 − f + cl f , (3.6)
rcl rcl

where f (x) is a cut-off function, which is unity at the origin, cuts off at


x ∼ 1 and decreases rapidly as x → ∞. So defined, the intermediate pseu-
dopotential Vlps,i (r) approaches its all-electron counterpart V (r) rapidly
and continuously as x increases outside the cut-off radius rcl . The free
parameter cl is then adjusted so that the nodeless solution wl of the radial

a Inthis context, the term “screened” is used in the sense that Vlps,scr (r) also contains
the interaction between valence states.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch03 page 36

36 Computer Simulations of Molecules and Condensed Matter

Figure 3.2 Ionic pseudopotentials for l = 0, 1, 2 in S atom. The pseudopotential is much


softer in the region close to the nucleus compared with the behavior of −Z/r, where Z
is the ionic charge. Similar to Fig. 3.1, the fhi98PP program is used [154].

equation
 
1 d2 l (l + 1) ps,i (i)
− + + Vl (r) − l wl (l , r) = 0 (3.7)
2 dr2 2r2
(i)
has the same eigenvalue as its all-electron counterpart (l = l ).
Since both wl and ul are solutions of the same potential outside the
core region, one can write

γl wl (r) → ul (r), r > rcl . (3.8)

The pseudowave function is now modified by adding a correction in the core


region

ups
l (r) = γl [wl (r) + δl gl (r)], (3.9)

where gl (r) must vanish as rl+1 for small r to give a regular pseudopoten-
tial at the origin, and it must vanish rapidly for r > rcl since γl wl (r) is
already the desired solution in that region. At this point, the normalization
condition is used to set the value of δl ,

2
γl2 |wl (r) + δl gl (r)| dr = 1. (3.10)
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch03 page 37

Pseudopotentials, Full Potential, and Basis Sets 37

The pseudowave functions generated by this procedure fulfill conditions


(i), (iii), and (iv). Condition (ii) is reached exponentially beyond the cut-
off radius. Even within this procedure, the pseudopotential is not unique.
It depends on the choice of states included in the core, the selection of the
cut-off functions and the core radii (see Ref. [151] for details). Later on,
Hamann [155] also extended this procedure to generate pseudopotentials
for the atomic unbound states.
In the Troullier–Martins scheme [156], the pseudovalence radial wave
function exactly equals the all-electron one outside the cut-off radius. Inside
the cut-off radius, it is assumed to have the following analytic form:
(l+1) p(r)
ups
l (r) = r e , (3.11)
where p(r) is a polynomial of sixth order in r2 . The coefficients are deter-
mined from conditions (ii)–(iv), plus the additional constraints of continuity
of the first four derivatives at rlc and zero curvature of the screened pseu-
d2 ps,scr
dopotential at the origin ( dr 2 Vl (r)|r=0 = 0). Condition (i) is accom-
plished directly by solving Eq. (3.3). As a consequence of the additional
requirements, the Troullier–Martins pseudopotentials are softer than the
Hamann ones.
The different radial dependence of the above-defined pseudopoten-
tials for each channel results in the total pseudopotential being semilocal
(i.e. non-local in the angular coordinates, but local in the radial one)
l
max m=l
 δ(r − r )
V ps (r, r ) = V loc (r)δ(r − r ) + ∗
Yl,m (r̂)δVlps (r) Yl,m (r̂ ),
r2
l=0 m=−l
(3.12)
where δVlps (r)
= Vlps (r)
− V loc (r). The l-independent term (V loc (r)) is
chosen such that the semilocal terms (δVlps ) are confined to the core region
and eventually vanish beyond some lmax .
Kleinman and Bylander (KB) [157] proposed a transformation of the
semilocal terms into a fully non-local form defining
 δVlps (r)ϕ̃l,m (r)ϕ̃∗l,m (r )δVlps (r )
δV KB (r, r ) = , (3.13)
ϕ̃l,m |δVlps |ϕ̃l,m 
l,m

ups
l (l ,r)
where ϕ̃lm (r) = r Ylm (r̂). It can be easily verified that

δV KB (r, r ) ϕ̃∗l,m (r )d3 r = δVlps (r)ϕ̃l,m (r), (3.14)
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch03 page 38

38 Computer Simulations of Molecules and Condensed Matter

that is, the KB form is equivalent to the semilocal one in the sense that it
produces the same atomic pseudoorbitals. At the expense of a more com-
plicated expression in real space, the KB form is fully separable, strongly
reducing the number of integrations necessary, e.g. in a plane wave basis
set, to calculate the Hamiltonian matrix elements. It is used in most of the
electronic structure codes nowadays.

3.1.2 Implicit Approximations

3.1.2.1 Frozen Core

The main assumption in the pseudopotential method is that the core states,
strongly bound to the nucleus and localized, are insensitive to the envi-
ronment surrounding the atom. Therefore, they can be excluded from the
self-consistent calculation in the poly-atomic system. This is the “frozen
core” approximation.
The pseudopotential is defined by the requirements that the wave func-
tions and eigenvalues are accurately reproduced, however, no conditions on
total energies are imposed. In 1980, von Barth and Gelatt [158] demon-
strated that the error in the total energy is of second order in the difference
between frozen and true-core densities. Their calculations for Mo further
confirmed this conclusion, thus validating the application of the pseudopo-
tential method in total energy calculations from this perspective.

3.1.2.2 Core–Valence Linearization

The definition of the pseudopotential in Eq. (3.5) implies that the self-
consistent total XC potential in a condensed matter system is written as

V xc [n(r)] = {V xc [n0 (r)] − V xc [ñ0v (r)]} + V xc [ñv (r)], (3.15)

where the terms in curly brackets are included in the pseudopotential. Equa-
tion (3.15) would be exact, within the frozen-core approximation, if the XC
potential were a linear functional of the density.b As it is clearly not the
case, the assumption of validity of Eq. (3.15) is known as core–valence
linearization.

b Notethat Eq. (3.15) is also exact in a non-self-consistent calculation, since in that case,
n(r) = n0 (r) and ñv (r) = ñ0v (r).
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch03 page 39

Pseudopotentials, Full Potential, and Basis Sets 39

However, the errors due to this approximation are small in most cases,
as long as the overlap between the core and valence densities are not signif-
icant. Louie et al. [159] developed a method for the generation and usage
of pseudopotentials that explicitly treats the nonlinear core–valence XC
interaction. The method consists in modifying Eq. (3.5) to
Vlps (r) = Vlps,scr (r) − V H [ñ0v ; r] − V xc [ñ0v + ñc ; r], (3.16)
where ñc is a partial core density. It reproduces the full core density in the
region where it overlaps with the valence density, outside a chosen cut-off
radius rnlc . Inside this radius, it is chosen to match the true density at
rnlc , minimize the integrated density and be easily Fourier transformed in
order to optimize its use within the plane wave basis set. This density has
to be added to the pseudovalence density in the self-consistent calculation
whenever V xc or E xc are computed.

3.1.2.3 Pseudoization

By pseudoization, we refer to the fact that the wave functions of the valence
states in the pseudopotential method are, by construction, nodeless and
much smoother than their all-electron analogue. It is only observable in
the core region, which constitutes a small portion of space. As long as
the full potential is local, the errors in the energies, within this region,
are taken care of in the pseudopotential by construction. Furthermore, the
norm-conserving constraint ensures that the Hartree potential generated by
the pseudocharge outside the core region is the same as in the all-electron
treatment. Nevertheless, whether it is also negligible in the calculation of
non-local operators is unclear, as mentioned in Ref. [160]. The fact that
pseudoization can lead to qualitative differences between PP and AE cal-
culations has been pointed out in Ref. [161], where significant discrepancies
in the electron–hole distribution function of LiF were observed. Because of
these, it is worth noting that this pseudoization is also a possible source of
error in practical ab initio electronic structure calculations, especially when
non-local operators are concerned [162, 163].

3.2 FP-(L)APW+lo Method


The FP-(L)APW+lo method is a development of the augmented plane wave
(APW) method originally proposed by Slater [164]. Thus, we start our dis-
cussions here with a short introduction of the APW method. The essential
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch03 page 40

40 Computer Simulations of Molecules and Condensed Matter

Figure 3.3 Schematic view of the space partition in the APW method. The space is
divided into the interstitial region and a series of non-overlaping muffin-tin (MT) regions.
The potential in the MT region is atomic-like, while that in the interstitial region is much
softer.

idea underlying this method is that in the region close to nuclei, the poten-
tials and wave functions are similar to those in the free atoms, strongly
varying but nearly spherical. In the space between the atoms, both of them
are smooth. In his seminal work, Slater proposed a division of the space in
the unit cell into a set of non-overlapping spheres centered at each atom
and the interstitial region between them (Fig. 3.3). The potential was taken
as spherically symmetric inside the spheres and constant outside (later on
known as the muffin-tin approximation, for obvious reasons). Accordingly,
the eigenfunctions of the Hamiltonian corresponding to each of the regions
are taken as basis functions, namely, plane waves in the interstitial and
atomic orbitals in the “muffin-tin” (MT) spheres. Adding the continuity
condition at the sphere boundary, the APWs were born

⎪ 1 i(G+k)·r

⎨Ωe , r ∈ interstitial,
k
φG (r) = 


⎩ Al,m (k + G) ul (rα , ) Yl,m (rα ) , r ∈ MT.
l,m
(3.17)
Inside each MT sphere, the radial wave function ul (rα , ) at the refer-
ence energy l , is obtained from

d2 l (l + 1)
− 2+ + V (r) −  l rul (r, l ) = 0. (3.18)
dr r2
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch03 page 41

Pseudopotentials, Full Potential, and Basis Sets 41

The augmentation coefficients ensuring the continuity of the wave function


at the sphere boundary are given by

4πil α ∗
Al,m (k + G) = j (|k + G|RMT
α , ) l
)Yl,m (k + G). (3.19)
Ω1/2 ul (RMT l

This optimized choice of basis functions in different regions is the


essence of the augmented methods and all its descendants. The wave func-
tion is expanded in terms of these APWs as

n k
ϕn,k (r) = CG φG (r) . (3.20)
G

n
The coefficients CG should be obtained by solving the eigenvalue equa-
tion

n
(HG,G − n SG,G ) CG =0 (3.21)
G

for each k where HG,G (SG,G ) are the Hamiltonian (overlap) matrix ele-
ments in the APW basis.
The major drawback of this method is that, inside the MT sphere,
the APWs are solutions of the Schrödinger equation only at the reference
energy (i.e. l = n ). Thus, the eigenvalue equation (3.21) becomes nonlinear
and its solution much more computationally demanding for each k-point.
Furthermore, it is hard, though not impossible (Refs. [165, 166]) to extend
the method to the full potential case. When the potential inside a MT
sphere is not spherical, the exact solution of the particle’s wave function
inside this MT sphere does not correspond to the solution of the radial
Schrödinger equation with the same eigenvalue.
Another shortcoming of the APW method, known as the asymptote
problem, is related to the indetermination of the augmentation coefficients
when the radial function has a node at the MT radius (ul (RMT ) in the
denominator of Eq. (3.19)). In the vicinity of this region, the relation
between Al,m and CG becomes numerically unstable.
With the aim of overcoming these limitations, Andersen [167] proposed
a modification of the APW method in which the wave functions and their
derivatives are made continuous at the MT radius by matching the intersti-
tial plane waves to the linear combination of a radial function, and its energy
derivative, calculated at a fixed reference energy. The method, known as the
linearized augmented plane waves (LAPWs) method, rapidly demonstrated
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch03 page 42

42 Computer Simulations of Molecules and Condensed Matter

its power and accuracy, becoming, during decades, the benchmark for elec-
tronic structure calculations within the KS scheme. Recently, Sjösted et al.
[168] proposed an alternative method in which the APW wave functions
are recovered, but with the radial functions calculated at a fixed energy.
The flexibility of the basis is achieved by adding a set of local orbitals
constructed as linear combinations of the same radial functions and their
energy derivatives, with the condition that the function cancels at the
sphere radius. This method, called APW plus local orbitals (APW+lo),
requires fewer plane waves for an accurate description of the electronic
structure properties, thus increasing the computational efficiency. However,
this improvement is limited by the large number of local orbitals required
for large l’s. Nowadays, the state-of-the-art method involves a combination
of both, using APW+lo’s for small l and LAPWs for the large ones [169],
known as (L)APW+lo method. The rest of the chapter is devoted to an
overview of these methods.

3.2.1 LAPW Basis Functions

The LAPW basis set is defined by

φk+G (r)

⎪ 1 i(G+k)·r

⎪ e , r ∈ interstitial,
⎪Ω

⎨ (3.22)
= [Al,m (k + G)ul (rα , l )



⎪ l,m

⎩ +Bl,m (k + G) u̇l (rα , l )]Yl,m (rα ) , r ∈ MT,

where (u̇l (r, l ) = ∂ul (r, )/∂|=l ). The augmentation coefficients Al,m and
Bl,m are obtained by requiring both the value and the slope of the basis
function to be continuous on the MT sphere boundary.
Making a Taylor expansion of the radial wave function around the
reference energy l , one has

ul (r, ) = ul (r, l ) + ( − l )u̇(r, l ) + O(( − l )2 ), (3.23)

which means that in the linearized treatment, the error in the wave function
is of the second order in −l . Taking into account the variational principle,
this leads to an error of fourth order, ( − l )4 , in the band energy. In other
words, the LAPWs form a good basis over a relatively large energy region,
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch03 page 43

Pseudopotentials, Full Potential, and Basis Sets 43

typically allowing the calculation of all the valence bands with a single set
of reference energies, i.e. by a single diagonalization of Eq. (3.21).
However, there are situations in which the use of a single set of refer-
ence energies is inadequate for all the bands of interest. Such a situation
arises, for example, when two (or more, but rarely) states with the same l
participate in the chemical bonding (semicore states), or when bands over
an unusually large energy region are required, like for high-lying excited
states. To address such cases, the local orbitals were introduced by Singh
in 1991 [170]:

φLO
LAPW (r)


⎪ 0, r ∈ interstitial,
⎨ (3.24)
α α α α
= [Al,m ul (r , l ) + Bl,m u̇l (r , l )


⎩ α
ul (rα , l )]Yl,m (rα ) , r ∈ MT.
(2)
+ Cl,m

(2)
In this way, a second set of energy parameters l is introduced to provide
the additional variational freedom required for an accurate representation
of the different states with the same l. The coefficients Al,m s, Bl,m s and
Cl,m s are determined by requiring the local orbital and its radial derivative
to be zero at the MT sphere boundary and normalized.

3.2.2 APW+lo Basis Functions

The LAPW basis set is designed to be flexible in describing the wave func-
tions in the vicinity of the reference energy. However, the requirement of
continuous derivatives at the MT radius increases the number of plane
waves needed to achieve a given level of convergence with respect to the
APW method.
Recently, Sjösted et al. [168] proposed an alternative way to linearize
the APW method in which the continuous derivative condition is released.
In this method, the eigenvalue problem of the original APW method is
linearized by choosing fixed linearization energies (l ) for the APW basis
functions in Eq. (3.17). Then, the flexibility of the basis set with respect to
the reference energy is obtained by adding a set of local orbitals (lo):

lo 0, r ∈ interstitial,
φAPW (r) = α α α
[Al,m ul (r , l ) + Bl,m u̇l (r , l )] Yl,m (r ) , r ∈ MT,
(3.25)
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch03 page 44

44 Computer Simulations of Molecules and Condensed Matter

using the same linearization energies. The coefficients are obtained by


requiring the function to be zero at the sphere boundary and normalized.
The APW basis functions keep the convergence behavior of the original
APW method while the local orbitals (lo) make it flexible with respect to
choice of the reference energy. The complete APW+lo basis set therefore
consists of two different types of basis functions, the APWs (Eq. (3.17)
at fixed linearization energies) and the lo’s (Eq. (3.25)). As in the LAPW
method, when different states with the same l (semicore states) have to be
treated, a second set of local orbitals of the form:


⎪ r ∈ interstitial,
⎨0,
LO α
φAPW (r) = [Al,m ul (r , l ) (3.26)


⎩ (2) α (2) α
+ C u (r ,  )]Y (r ) , r ∈ MT
l,m l l l,m

can be added. The coefficients are determined by matching the function to


zero at the muffin-tin radius, with no condition on the slope.

3.2.3 Core States

As already mentioned, the (L)APW+lo is an all-electron method. However,


it does not mean that core and valence states are treated in the same way.
While the latter are expanded in the previously described basis set using
the crystal potential, the former are calculated numerically by solving the
relativistic radial Schrödinger equation for the atom. The influence of the
core states on the valence is carried out by the inclusion of the core density
in the Hartree and XC potentials. Reciprocally, the core states are calcu-
lated using the spherical average of the crystal potential in the muffin-tin
sphere. Thus, both core and valence states are calculated self-consistently.
In the Wien2k code, the wave function of each core state is repre-
sented as
a
a,n,j,mj (r) = ua,n,κ (r )|jmj l ,
ϕ̃core (3.27)

where
1

2
 
1
|jmj l ≡ l ml σ|j mj Ylml (r̂a ) |σδm+σ,mj , (3.28)
1
2
σ=− 2

and (l 12 ml σ|j mj ) is the corresponding Clebsch–Gordon coefficient


(Ref. [171]).
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch03 page 45

Pseudopotentials, Full Potential, and Basis Sets 45

Table 3.1 Relativistic quantum numbers.


s
j =l+ κ Max. occupation
2
l s = −1 s = +1 s = −1 s = +1 s = −1 s = +1

s 0 1/2 –1 2
p 1 1/2 3/2 1 –2 2 4
d 2 3/2 5/2 2 –3 4 6
f 3 5/2 7/2 3 –4 6 8

The radial wave function is defined by the relativistic quantum number


κ = −s(j + 12 ) as shown in Table 3.1.

3.2.4 Potential and Density

The representation of the density and the potential has to confront the
same difficulties as the representation of the wave functions, namely, rapid
variations in the muffin-tin spheres and soft oscillations in the interstitial.
The use of a dual representation as for the wave functions, which is the
basis of the (L)APW+lo efficiency, seems the natural choice. However, an
expansion in spherical harmonics inside the spheres and plane waves in the
interstitial is clearly inefficient. The complete representation of the density
requires a basis set at least eight times larger than the basis required for
the wave functions. Since the number of augmentation functions in the MT
sphere also increases four times, the number of augmentation coefficients is
25 times larger.
This can be reduced by exploiting the symmetries of the density (poten-
tial), namely:

(i) Inside the muffin-tin sphere, they respect the symmetry of the corre-
sponding nuclear site.
(ii) In the interstitial region, they have the symmetry of the corresponding
space group.
(iii) Both are real quantities.

Inside the muffin-tin spheres, properties (i) and (iii) allow the rep-
resentation of the density in a lattice harmonic expansion [172]. For the
interstitial region, the use of stars ensures both properties (ii) and (iii) to
be fulfilled with a minimum number of coefficients. More details can be
found in Ref. [149].
b2530   International Strategic Relations and China’s National Security: World at the Crossroads

This page intentionally left blank

b2530_FM.indd 6 01-Sep-16 11:03:06 AM


December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 47

4
Many-Body Green’s Function Theory
and the GW Approximation

In Chapter 2, we pointed out that although the Kohn–Sham (KS) eigen-


values provide a good zeroth-order approximation for the single-particle
excitation energies, LDA/GGA fails to give a satisfactory description of
the fundamental band gaps in semiconductors and insulations. The many-
body Green function theory (or otherwise called many-body perturbation
theory), on the other hand, provides the formal basis for evaluating the
experimentally observed quasi-particle band structures. Nowadays, it has
become routine practice to resort to this many-body perturbation theory
(especially within its so-called “GW approximation”) for the descriptions
of the single-particle excitation in real poly-atomic systems. Because of this,
we will present an overview of the Green function method in the many-body
electronic systems, with a special emphasis on the GW approximation in
this chapter.
In order to provide a simple introduction, we begin with a general
description of the main ingredients in this theory in Sec. 4.1. This descrip-
tion starts from the definition of the single-particle Green function, the
central physical quantity in this method. In Sec. 4.1.1, we explain the cor-
respondence between the poles of this single-particle Green function on the
frequency axis and the single-particle excitation energies of a many-body
system. A numerical description of this single-particle Green function in a
many-body system, however, is much easier said than done. In practical sim-
ulations, one needs to resort to the single-particle Green function of a non-
interacting system. The equation, which relates the single-particle Green
function of the many-body system and that of the non-interacting system,
is called the Dyson equation. In Sec. 4.1.2, we show how this equation

47
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 48

48 Computer Simulations of Molecules and Condensed Matter

is deduced. In the Dyson equation, all the complexity of the many-body


interaction is put into a term which is called the “self-energy”. Then in
Sec. 4.1.3, we devote our discussions to the concept of this self-energy and
its expansion in terms of the dynamically screened Coulomb potential, i.e.
the Hedin equations. The concept of the quasi-particle is introduced in
Sec. 4.1.4.
Then we go to the practical schemes for the numerical treatment of the
Hedin equations in real poly-atomic systems. The simplest approximation
to the self-energy, which includes the dynamical screening effects, is the
so-called GW approximation. Here, G means the Green function and W
means the screening Coulomb potential. Their product presents the sim-
plest approximation for the self-energy, which, as was said, includes the
important dynamical screening feature of the many-body interactions. The
principles underlying this GW approximation will be introduced in Sec. 4.2.
This GW approximation, in its most rigorous form, means that the
Dyson equation needs to be solved self-consistently. In practical simulations
of real poly-atomic systems, however, a further approximation beyond it is
often used. This further approximation treats the self-energy, obtained in
a GW manner, only as a first-order correction to the exchange–correlation
(XC) potential of the fictitious non-interacting particles, which can corre-
spond to either the KS or the Hartree–Fock orbitals. It is nowadays often
called the G0 W0 approximation in the literature. Over the last 30 years,
it has achieved great success in describing the single-particle excitations in
a wide range of poly-atomic systems, ranging from molecules to periodic
semiconductors, to even metals. This G0 W0 approach will be described in
Sec. 4.3.
With these, we hope that we have relayed the key messages of the prin-
ciples underlying this single-particle excitation theory in a relatively clear
manner. As an example of its implementation in computer programming,
we use an all-electron GW code, i.e. the FHI-gap,a to show how some
technical details are treated in Sec. 4.4. We note that due to the heavy
computational load associated with the LAPW basis, this code might not
be the best choice for the GW calculation of large poly-atomic systems.
When large simulation cells must be used, other computational codes such
as BerkeleyGWb [173], which employs the pseudopotentials and plane wave

a For details, please see http://www.chem.pku.edu.cn/jianghgroup/codes/fhi-gap.html.


b For details, please see http://www.berkeleygw.org/.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 49

Many-Body Green’s Function Theory and the GW Approximation 49

basis set, should be a much better choice. Nevertheless, because of the


accuracy of this LAPW basis set in descriptions of the electronic states
in a full-potential (FP), all-electron manner, we would like to recommend
the FHI-gap code as the benchmark for excited state single-particle elec-
tronic structure calculations of the periodic systems. It is important to note
that most of this FHI-gap code was written by Dr. Ricardo Gómez-Abal
starting from 2002, during his postdoc years in the group of Prof. Matthias
Scheffler in the Fritz Haber Institute of the Max Planck Society, Berlin.
From 2004 to 2008, one of the authors of this book (XZL) did his Ph.D.
in this group. His thesis focused on the development of this all-electron
GW code and its applications to conventional semiconductors. Dr. Ricardo
Gómez-Abal supervised him directly during this time. Starting from 2006,
Hong Jiang (who is currently a Professor in the College of Chemistry and
Molecular Engineering, Peking University, China) joined this project and
significantly improved its completeness and efficiency. After 2008, this code
was maintained and further developed mainly by Prof. Hong Jiang.

4.1 Green Function Method


4.1.1 The Green Function

The single-particle Green function is defined as

G (r, t; r , t ) = −iN |T̂ {ψ̂ (r, t) ψ̂ † (r , t )}|N , (4.1)

where ψ̂(r, t) (ψ̂ † (r , t )) is the quantum field operator describing the anni-
hilation (creation) of one electron at position r and time t (position r and
time t ). The operator T̂ is the time-ordering operator, which reorders the
field operators in ascending time order from right to left. |N  is the ground-
state eigenfunction of the N -electron system. Making use of the Heaviside
function (Appendix A), and the commutation relations for fermionic oper-
ators, Eq. (4.1) can be rewritten as

G (r, t; r , t ) = − iN |ψ̂ (r, t) ψ̂ † (r , t ) |N Θ (t − t )


(4.2)
+ iN |ψ̂ † (r , t ) ψ̂ (r, t) |N Θ (t − t) ,

making evident that for t > t (t < t ), the Green function describes the
propagation of an added electron (hole) in the system.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 50

50 Computer Simulations of Molecules and Condensed Matter

In the Heisenberg representation, the field operator is written as

ψ̂ (r, t) = eiĤt ψ̂ (r) e−iĤt , (4.3)

where Ĥ is the Hamiltonian operator and ψ̂ (r) is the field operator in the
Schrödinger representation.
Inserting Eq. (4.3) into Eq. (4.2) and making use of the completeness
relation in the Fock space:

∞ 

1= |n, s n, s| , (4.4)
n=0 s

where |n, s corresponds to the sth eigenstate of the the n-electron system,
we can transform Eq. (4.2) into
 s 
G (r, t; r , t ) = − i N |ψ̂ (r) |N + 1, se−i(EN +1 −EN )(t−t )
s

· N + 1, s|ψ̂ † (r ) |N Θ (t − t ) + i N |ψ̂ † (r ) |N − 1, s
s
s 
−i(EN −1 −EN )(t −t)
·e N − 1, s|ψ̂ (r) |N Θ (t − t) .
(4.5)
Here, EN stands for the ground-state energy of the N -electron system, and
s
EN ±1 for the sth excited-state energy of the N ± 1 electronic system.
Using the excitation energy s and amplitude ψs (r) defined by

s
s = EN +1 − EN , ψs (r) = N |ψ̂ (r) |N + 1, s, for s  μ,
(4.6)
s
s = EN − EN −1 , ψs (r) = N − 1, s|ψ̂ (r) |N , for s < μ,

where μ is the chemical potential of the N -electron system (μ = EN +1 −


EN ),c we can further simplify Eq. (4.5) into the form
 
G (r, r ; t − t ) = −i ψs (r) ψs∗ (r ) e−is (t−t )
s

[Θ (t − t ) Θ (s − μ) − Θ (t − t) Θ (μ − s )] . (4.7)

c InFig. 4.1, this corresponds to the energy of the lowest unoccupied state when T → 0 K;
for its derivation, see Ref. [174].
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 51

Many-Body Green’s Function Theory and the GW Approximation 51

Figure 4.1 Position of the poles of the Green function (Eq. (4.8)) in the complex
frequency plane. Those corresponding to the unoccupied states are slightly below
the real frequency axis while those corresponding to the occupied states are slightly
above it.

Performing a Fourier transform to the frequency axis, we obtain the


spectral, or Lehmann [175], representation

  
Θ (s − μ) Θ (μ − s )
G (r, r , ω) = lim+ ψs (r) ψs∗ (r ) + .
η→0
s
ω − (s − iη) ω − (s + iη)
(4.8)
The key feature of Eq. (4.8) is that the Green function has single poles
corresponding to the exact excitation energies of the many-body system.
For excitation energies larger (smaller) than the chemical potential, these
singularities lie slightly below (above) the real axis in the complex frequency
plane (Fig. 4.1).
It can be easily shown that in the non-interacting case, Eq. (4.8)
reduces to

  
 Θ (n − F ) Θ (F − n )
G0 (r, r , ω) = lim ϕn (r) ϕ∗n
(r ) 
+ ,
η→0+
n
ω − (n − iη) ω − (n + iη)
(4.9)
where n (ϕn ) is the eigenvalue (eigenfunction) of the single-particle Hamil-
tonian and F is the Fermi energy.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 52

52 Computer Simulations of Molecules and Condensed Matter

4.1.2 The Dyson Equation

The time evolution of the field operator in the Heisenberg representation is


given by the equation of motion:

∂  
i ψ̂ (r, t) = ψ̂ (r, t) , Ĥ (4.10)
∂t

with the Hamiltonian operator given by


  
1 2

Ĥ = drdtψ̂ (r, t) − ∇ + Vext (r) ψ̂ (r, t)
2

1
+ drdtdr dt ψ̂ † (r, t) ψ̂ † (r , t ) v (r, t; r , t ) ψ̂ (r , t ) ψ̂ (r, t) ,
2
(4.11)

where v(r, t; r , t ) = δ(t − t )/|r − r | is the Coulomb interaction. By


evaluating the commutator in Eq. (4.10), the equation of motion for the
single-particle Green function can be obtained:
 
∂ 1
i + ∇2 − Vext (r) G (r, t; r , t )
∂t 2

1
+i dr1 N |T [ψ̂ † (r1 , t) ψ̂ (r1 , t) ψ̂ (r, t) ψ̂ † (r , t )]|N 
|r − r1 |
= δ (r − r ) δ (t − t ) . (4.12)

The quantity in the integrand of the second term is the two-particle Green
function. Following the same procedure to obtain the equation of motion
for the two-particle Green function will give a term depending on the three-
particle Green function, and so on.
To break this hierarchy, a mass operator can be introduced, defined by

dr1 dt1 M (r, t; r1 , t1 ) G (r1 , t1 ; r , t )

= −i dr1 v (r − r1 ) N |T [ψ̂ † (r1 , t) ψ̂ (r1 , t) ψ̂ (r, t) ψ̂ † (r , t )]|N .

(4.13)
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 53

Many-Body Green’s Function Theory and the GW Approximation 53

With this operator, Eq. (4.12) can be rewritten as


 
∂ 1
i + ∇2 − Vext (r) G (r, t; r , t )
∂t 2

− dr1 dt1 M (r, t; r1 , t1 ) G (r1 , t1 ; r , t ) = δ (r − r ) δ (t − t ) .
(4.14)
Since the Hartree interaction is a one-particle operator, it is usually sep-
arated from the mass operator M to define the self-energy, Σ = M − VH .
Replacing the mass operator in Eq. (4.14), we arrive at
 

i − H0 (r) G (r, t; r , t )
∂t

− dr1 dt1 Σ (r, t; r1 , t1 ) G (r1 , t1 ; r , t ) = δ (r − r ) δ (t − t ) ,

(4.15)
where
1
H0 (r) = − ∇2 + Vext (r) + VH (r) . (4.16)
2
In the Hartree approximation, Eq. (4.15) becomes
 

i − H0 (r) G0 (r, t; r , t ) = δ (r − r ) δ (t − t ) . (4.17)
∂t
Multiplying Eq. (4.15) by G0 on the left, using the hermiticity of the single-
particle operator together with Eq. (4.17) and then integrating yields the
well-known Dyson equation:
G (r, t; r , t ) = G0 (r, t; r , t )

+ dr1 dt1 dr2 dt2 G0 (r, t; r2 , t2 ) Σ (r2 , t2 ; r1 , t1 )

×G (r1 , t1 ; r , t ) . (4.18)
Recurrently replacing G on the right-hand side by G0 + G0 ΣGd leads to
the series expansion:
G = G0 + G0 ΣG0 + G0 ΣG0 ΣG0 + · · · , (4.19)

d Inthis symbolic notation, products imply an integration, as a product of matrices with


R
continuous indices, i.e. AB = A(1, 3)B(3, 2)d3.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 54

54 Computer Simulations of Molecules and Condensed Matter

which shows that the single-particle propagator G(r, t; r , t ) is equal to


the “free” particle propagator G0 (r, t; r , t ) plus the sum of the probabil-
ity amplitudes propagating from (r, t) to (r , t ) after single, double, etc.,
scattering processes, with Σ playing the role of the scattering potential.
Diagrammatically, this relation is shown as

where the double plain arrow represents the interacting Green function and
the plain arrow represents the non-interacting one.

4.1.3 Self-Energy: Hedin Equations

For an electron propagating in a solid or condensed phase, the origin of


the scattering processes lies in the Coulomb interaction with the Fermi sea.
Thus, it is natural to expand the self-energy in terms of the bare Coulomb
interaction. In the diagrams below, we show examples of some simple (low-
order) scattering processes. Diagram (a) is a first-order scattering process
that describes the propagating electron instantaneously exchanging its posi-
tion with one electron from the Fermi sea via the Coulomb interaction. It
corresponds to the exchange interaction. Solving the Dyson equation (4.15)
including only this term in the self-energy and updating the Green function
self-consistently yields the Hartree–Fock approximation. In Diagram (b),
the interaction of the probe electron with the Fermi sea excites an electron
out of it, generating an electron–hole pair, which annihilate each other
at a later time, interacting again with the probe electron. This second-
order scattering process, depicted in a “bubble” diagram, represents an
electron repelling another from its neighborhood, thus generating a positive
charge cloud around it. It is one of the simplest dynamical screening pro-
cesses. In Diagram (c), the excited electron in the electron–hole pair of
Diag. (b) further excites another electron–hole pair from the Fermi sea,
changing the positive charge cloud around the probe electron again. Nev-
ertheless, the long range of the bare Coulomb interaction results in a poor
convergence of this expansion for the self-energy; in fact, it diverges for
metals.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 55

Many-Body Green’s Function Theory and the GW Approximation 55

To solve this convergence problem related to the expansion of the self-


energy in terms of the bare Coulomb interaction, in 1965, Hedin [176] pro-
posed a different approach for obtaining this self-energy. In this approach,
the self-energy is obtained by expanding it in terms of a dynamically
screened Coulomb potential instead of the bare one. The derivation using
the functional derivative technique can be found in Refs. [176–178]. Here,
we just present the resulting set of equations:

δΣ (1, 2)
Γ (1, 2, 3) = δ (1, 2) δ (2, 3) + d (4, 5, 6, 7) G (4, 6)
δG (4, 5)

× G (7, 5) Γ (6, 7, 3) , (4.20a)



P (1, 2) = −i G (2, 3) G (4, 2) Γ (3, 4, 1) d (3, 4) , (4.20b)

W (1, 2) = v (1, 2) + W (1, 3) P (3, 4) v (4, 2) d (3, 4) , (4.20c)


Σ (1, 2) = i d (3, 4) G 1, 3+ W (1, 4) Γ (3, 2, 4) . (4.20d)

In these equations, we have used the Indo-Arabic numbers to denote the


events and simplify the notation, e.g. 1 = (r1 , t1 ). Γ is a vertex function, P
is the polarizability and W is the dynamically screened Coulomb potential.
In Eq. (4.20a), the vertex function is written in terms of a four-point kernel
(given by the functional derivative of the self-energy). Replacing the self-
energy by the expression in Eq. (4.20d) would allow the expansion of the
vertex function in terms of the screened Coulomb potential. For the purpose
of this book, it will nevertheless be sufficient to represent it by a filled
triangle:
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 56

56 Computer Simulations of Molecules and Condensed Matter

Equations (4.20b)–(4.20d) can then be represented diagrammatically as

and

where the double wiggly line represents the screened Coulomb potential.

Figure 4.2 Schematic representation of the self-consistent solution of the Hedin equa-
tions in conjunction with the Dyson equation for the determination of the Green function
(G) and the self-energy (Σ). Entries in boxes symbolize the mathematical relations that
link Σ, G, Γ, P and W .

The set of Eq. (4.20), together with the Dyson Equation (4.18), consti-
tute the definitive solution of the quantum mechanical many-body problem.
One just needs to solve them self-consistently to obtain the single-particle
Green function of the interacting system (see Fig. 4.2). However, one needs
to note that a direct numerical solution is prevented by the functional
derivative in Eq. (4.20a), and, as usual, one has to rely on approximations.
This will be the subject of the second part of this chapter.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 57

Many-Body Green’s Function Theory and the GW Approximation 57

4.1.4 The Quasi-Particle Concept

Defining the excitation energies s and amplitudes ψs (r) (Eq. (4.6)) allowed
us to write the Green function of the interacting system in the spectral rep-
resentation (Eq. (4.8)). The expression obtained has the same form as the
Green function of the non-interacting system (Eq. (4.9)), with the excita-
tion energies (amplitudes) playing the role of the single-particle eigenvalues
(eigenfunctions). We may ask, under which condition can the object defined
by ψs (r) and s be interpreted as a “particle” that can be measured exper-
imentally.
The experimentally obtained quantity in photoemission experiments is
the spectral function,e i.e. the density of the excited states that contribute
to the spectrum. For a finite system, it is defined by (Fig. 4.3(a))


A(r, r ; ω) = ψs (r) ψs∗ (r ) δ(ω − s ), (4.21)
s

and the Green function can be rewritten as



A(r, r ; ω  )
G (r, r , ω) = lim dω  . (4.22)
η→0+ ω− ω + i sgn(ω  − μ)η

Therefore, the interpretation of the excitation as a “particle” presents


no difficulty in a finite system. Furthermore, inserting the expression for
the Green function in terms of s and ψs (r) (Eq. (4.7)) in Eq. (4.15), it can
be shown that they are solutions of the quasi-particle equationf
  
1 2
− ∇ + Vext (r) + VH (r) ψs (r) + dr Σ (r, r ; s ) ψs (r ) = s ψs (r) .
2
(4.23)
In an extended system, the delta functions in Eq. (4.21) form a con-
tinuous spectrum (Fig. 4.3(b)). However, if in a given energy window, the
spectrum can be described by a series of Lorenzian peaks with finite widths,

e Assuming the cross-section of the perturbation to be independent of the energy and


neglecting experimental errors.
f This equation was first derived by Schwinger in Ref. [179]. It was applied to the many-

body electronic system by Pratt in Refs. [180, 181] and later systematically by Hedin in
Refs. [176, 182].
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 58

58 Computer Simulations of Molecules and Condensed Matter

(a) (b)

Figure 4.3 Spectral function for a discrete (a) and a continuous (b) spectrum.

so that the spectral density function can be written as


 Γs
A(r, r ; ω) = ψs (r) ψs∗ (r ) , (4.24)
s
(ω − s )2 + Γ2s

where s is the center of the peak and Γs the width, Eq. (4.22) can be inte-
grated analytically. Thus, the results of Eqs. (4.7) and (4.8) are recovered,
provided one redefines s = s + iΓs . In this case, the object defined by
ψs (r) and the complex s is called a “quasi-particle”. It describes the group
behavior of a set of excitations with continuous excitation energies. The real
part of s corresponds to the average energy of these related excitations.
The imaginary part leads to a decaying factor e−Γs t , i.e. the excitation has
a finite lifetime given by τ = Γ−1s . That the quasi-particle “disappears” can
be physically understood, taking into account that one is dealing with an
infinite system. In other words, the quasi-particle can decay to the “infi-
nite” reservoir. The quasi-particle equation (4.23) remains valid, provided
one performs an analytic continuation of the self-energy to the complex
frequency plane. A more detailed discussion on the subject can be found in
Ref. [183].

4.2 GW Approximation
In the above discussions, we have shown that in descriptions of the single-
particle excitation of a many-body system, the key quantity is the self-
energy and the key equation is the Dyson equation. The self-energy contains
all the complexity of the many-body interactions and in practical simula-
tions, one must resort to approximations of this quantity. The simplest
approximation of this quantity, which contains the dynamic feature of the
screened Coulomb interaction, is the so-called GW approximation. It was
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 59

Many-Body Green’s Function Theory and the GW Approximation 59

first proposed by Hedin in 1965 [176]. Mathematically, it amounts to taking


the zeroth-order expansion of the vertex function in terms of W . Thus, we
are left with

Γ (1, 2, 3) = δ (1, 2) δ (2, 3) , (4.25a)

P (1, 2) = −iG (1, 2) G (2, 1) , (4.25b)



W (1, 2) = v (1, 2) + d (3, 4) W (1, 3) P (3, 4) v (4, 2) , (4.25c)

Σ (1, 2) = iG (1, 2) W 1+ , 2 . (4.25d)

Diagrammatically, the three-point Γ function is collapsed into a point. The


elementary unit in this set of equations is the bubble diagram of the polar-
izability operator:

This approximation for the polarizability is known as the random phase


approximation (RPA). Physically, it represents the polarization generated
by the creation and annihilation of a dressed electron–hole pair, while the
interaction between the (dressed) electron and hole is neglected. In other
words, scattering processes where the electron or the hole in the electron–
hole pair interacts with the medium are taken into account. For example,
the process represented by the following diagram:

can be included. However, processes like

where the electron and the hole interact with each other are neglected.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 60

60 Computer Simulations of Molecules and Condensed Matter

The screened Coulomb interaction resulting from Eq. (4.25c) is the


same as in Eq. (4.20c):

except that now the polarizability is represented in the RPA.


In Eq. (4.25d), the self-energy is written as a product of the Green
function and the screened Coulomb interaction diagrammatically as

The shape of this diagram is similar to the Hartree–Fock approximation,


with the instantaneous bare Coulomb potential replaced by the dynamically
screened Coulomb one. This approximation to the self-energy includes pro-
cesses represented, for example, by the diagrams:

through the screened Coulomb potential. Also, processes like

are included through the interacting Green function. However, diagrams


like

where the added electron interacts with that of the electron–hole pair are
neglected.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 61

Many-Body Green’s Function Theory and the GW Approximation 61

Figure 4.4 Schematic representation of the self-consistent solution of Hedin equations


in GW approximation. Entries in boxes symbolize the mathematical relations that link
Σ, G, P and W .

In Fig. 4.4, we show a sketch of the self-consistent procedure required to


solve the GW equations (4.25) together with the Dyson equation. However,
this procedure is still extremely computationally demanding and rarely
carried out.

4.3 G0 W0 Approximation
A popular method in avoiding the self-consistent solution of the Dyson
equation, which contains the essence of the many-body interactions at the
GW level in descriptions of the self-energy, is the so-called G0 W0 approx-
imation. The key point of this approximation is the assumption that one
can count on an effective single-particle potential V xc (r), which contains
some of the XC effects in a many-body system and approximates the self-
energy reasonably well. In other words, the solutions of the single-particle
equation

Ĥeff (r)ϕi (r) = i ϕi (r) (4.26)

with
1
Ĥeff (r) = − ∇2 + Vext (r) + V H (r) + V xc (r) (4.27)
2
are such that ϕi (r) ≈ ψs (r), i ≈ (s ) (ψs (r) and s are the solutions of
Eq. (4.23)).
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 62

62 Computer Simulations of Molecules and Condensed Matter

For a convenient understanding of the perturbative treatment underly-


ing this approximation, we can rewrite the quasi-particle equation (4.23) as
 
1 2 H xc
− ∇ + Vext (r) + V (r) + V (r) ψs (r)
2

+ dr ΔΣ (r, r ; s ) ψs (r ) = s ψs (r) , (4.28)

where

ΔΣ (r, r ; s ) = Σ (r, r ; s ) − V xc (r ) δ(r − r ). (4.29)

Since, according to our assumption, the correction due to ΔΣ is small, one


can obtain the quasi-particle energies by applying the first-order perturba-
tion theory through

ψi (r) = ϕi (r),

qp qp
i = i + ϕi (r1 ) | [ΔΣ (r1 , r2 ; i )] |ϕi (r2 ) . (4.30)

Taking the self-energy in the GW approximation, and further assuming


that the non-interacting Green function G0 corresponding to Ĥeff is a good
approximation to the interacting one, the self-energy can be calculated
through

P0 (1, 2) = −iG0 (1, 2) G0 (2, 1) , (4.31a)



W0 (1, 2) = v (1, 2) + d (3, 4) W0 (1, 3) P (3, 4) v (4, 2) , (4.31b)

Σ (1, 2) = iG0 (1, 2) W0 1+ , 2 . (4.31c)

With these, one can easily calculate the quasi-particle corrections to the
single-particle excitation energy in Eq. (4.30).
We note that this treatment of the quasi-particle excitation energies
is obviously not done in a self-consistent manner, as illustrated pictorially
by the diagram in Fig. 4.4. Due to the fact that the self-energy is repre-
sented by the product of G0 and W0 , this further approximation beyond the
self-consistent treatment of the GW approximation is known as the G0 W0
approximation in the literature.
With this framework of the G0 W0 approximation set up, the next
thing we need to do is to find a good zeroth-order approximation for
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 63

Many-Body Green’s Function Theory and the GW Approximation 63

the quasi-particle energies and wave functions. The KS scheme to the


density-functional theory naturally provides such a good zeroth-order
approximation since its single-particle effective potential includes quite
some XC effects. Thus, one usually starts from the KS Green function:
 KS
G0 (1, 2) = −i ϕKS
j (r1 ) ϕj
KS∗
(r2 ) e−ij (t1 −t2 )
j

 
· Θ (t1 − t2 ) Θ KS KS
j − μ − Θ (t2 − t1 ) Θ μ − j . (4.32)

We note that other fictitious single-particle states can also be used as the
starting guess of the quasi-particle states. The results depend on this start-
ing guess and normally correct them in the right direction. For example,
the LDA-based KS orbitals give smaller band gaps in molecules and semi-
conductors (in the case of a molecule, it is the difference between the lowest
unoccupied molecular state energy and the highest occupied molecular state
energy). After the G0 W0 correction, these band gaps increase. On the other
hand, when the Hartree–Fock band gaps are used, the starting point has
larger band gaps than those in experiments. Then when the G0 W0 cor-
rection is used, these values decrease toward the experimental results. In
spite of this dependence, the KS orbitals are still the most popular choice
in calculations of the G0 W0 corrections. As an example, in Fig. 4.5, we
show how the LDA-based KS band diagram is compared with the G0 W0
band diagram in Si. It is clear that after the G0 W0 correction is added,
the agreement between the theoretically calculated band diagram (in an ab
initio manner) and experiment significantly improves.
For the completeness of this introduction to the GW method, in the
following, we use three more paragraphs to summarize the history behind
the development of this method and its present status.
The first application of the G0 W0 method to real materials was carried
out by Hybertsen and Louie in 1985 [185]. Since then, it has achieved an
impressive success in reproducing the experimental single-particle excita-
tion spectra for a wide range of systems from simple metals [186, 187]
to weakly correlated semiconductors and insulators [188–190]. Nowadays,
it has become the standard method used to calculate the single-electron
excitations of condensed matter systems. For these early implementa-
tions, one notes that the pseudopotential method and its associated plane
waves for the expansion of the wave functions were often resorted to
due to their simplicity of implementation and computational efficiency.
In the last 10 years, all-electron method-based implementations of the
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 64

64 Computer Simulations of Molecules and Condensed Matter

Figure 4.5 Band diagram of Si (in eV, referenced to the valence band maximum)
obtained from the all-electron LDA (labels as AE-LDA in solid lines) and G0 W0 (labelled
as AE-G0 W0 in dash lines) calculations. The experimental results (circles) are taken from
Ref. [184].

GW approximation also appeared [160–163, 178, 191–193]. The earliest


results for simple semiconductors such as Si and GaAs show that when
the all-electron method is used, the G0 W0 band gaps of these semiconduc-
tors become smaller compared with the pseudopotential-based results [160,
191] (see Table 4.1), and their agreement with the experimental values gets
worse.
For a better understanding of this discrepancy, we take Si as the guid-
ing example and explain how its fundamental G0 W0 band gap evolves
with time in the last 30 years. Early pseudopotential-based G0 W0 calcula-
tions give values between 1.18 eV and 1.29 eV in the 1980s and 1990s [188,
194–196]. This is within 0.12 eV accuracy compared with the experimen-
tal value of 1.17 eV. However, after 2000, all-electron G0 W0 calculations
obtained 0.85 eV [160] and 0.90 eV [191]. These results triggered a debate
about the reliability of both pseudopotential and all-electron based G0 W0
results [198, 199]. In Ref. [160], this discrepancy was assigned mainly to
the exclusion of the core electrons in the calculation of the self-energy,
and the possible role of pseudoization was tangentially mentioned. On
the other hand, the main criticism to the value of 0.85 eV in Ref. [160],
put forward in Ref. [198], was the lack of convergence with respect to
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 65

Many-Body Green’s Function Theory and the GW Approximation 65

Table 4.1 Comparison of band gaps from some reported all-electron


and pseudopotential-based G0 W0 calculations.
(Unit: eV)

Expt All-electron-G0 W0 a Traditional PP-G0 W0

C 5.50 5.49 5.54c


Si 1.17 0.90, 0.85b 1.18c , 1.19d ,
1.24e , 1.42f
AlAs(Γ − X) 2.23 1.68 2.08f
AlAs(Γ − Γ) 3.13 2.69 2.75f
AlP 2.50 2.15 2.59g
AlSb 1.69 1.32 1.64g
Ge(Γ − L) 0.74 0.47, 0.51a 0.75f
Ge(Γ − Γ) 0.89 0.79, 1.11a 0.71f
GaAs 1.63 1.42 1.29f
GaP(Γ − X) 2.35 1.90 2.55h
GaP(Γ − Γ) 2.86 2.53 2.93h
GaSb 0.82 0.49 0.62g
InP 1.42 1.25 1.44g
InAs 0.42 0.32 0.40g
InSb 0.24 0.32 0.18g
ZnS 3.80 3.22 3.98i
ZnSe 2.80 2.21 2.84i
CdTe 1.61 1.07 1.76i

Notes: a Ref. [191], b Ref. [160], c Ref. [194], d Ref. [195], e Ref. [196],
f Ref. [188], g Ref. [190], h Ref. [197], i Ref. [189].

the number of unoccupied states included in the calculation.g The validity


of this criticism has been confirmed by the all-electron G0 W0 result from
Friedrich et al. [192], who obtained, after careful convergence with respect
to the number of unoccupied states, a fundamental band gap of 1.05 eV
in Si. This trend was later supported by another LAPW-based G0 W0 cal-
culation in Refs. [162, 163], which gives a value of 1.00 eV. Meanwhile, the
value of 0.90 eV in Ref. [191], based on the LMTO method, has been further
increased to 0.95 eV in Ref. [200] after including local orbitals. From the
pseudopotential perspective, in 2004, Tiago et al. [201] performed a set of
pseudopotential calculations for Si in which only the 1s orbitals are treated
as core, obtaining a fundamental band gap of 1.04 eV for Si. With these, it

g Related to other arguments put forward in this discussion, namely the cancellation of
errors between lack of self-consistency and absence of vertex correction, we consider them
pertinent to support the G0 W0 approach itself, but irrelevant to an explanation of the
differences between all-electron and pseudopotential-based calculations.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 66

66 Computer Simulations of Molecules and Condensed Matter

seems clear that the earlier reported all-electron results in Refs. [160, 191]
should have some problems related to its convergence with respect to the
number of unoccupied states included in the calculations of the self-energy.
When this issue is taken care of, larger band gaps and smaller discrepancy
with the early pseudopotential-based results should be expected. However,
one notes that differences on the order of 0.1 eV still remain in many of
the simple semiconductors. Physically, there is no other reason to explain
such a discrepancy, except for the intrinsic approximations underlying the
pseudopotential method. In Refs. [162, 163, 193], a systematic analysis on
how these approximations underlying the pseudopotential-based GW cal-
culations are given.
In spite of this clarification, we note that there is still no perfect method
for calculating the single-particle excitation energy from the theoretical
perspective due to the different approximations we have already made in
this chapter for the simplification of the many-body perturbation theory
into a practical scheme. These approximations include, most seriously, the
neglect of the vertex correction. Concerning the self-consistent solution of
the Dyson equation within GW approximation, we acknowledge that great
progress has been made in the last few years [202–211]. From a practical
perspective, the pseudopotential-based G0 W0 method prevails nowadays
due to its efficiency and well-defined physical meaning for the interpretation
of the quasi-particle energies, especially in calculations of large systems.
In some solids, when the all-electron method is applicable and its error
is controllable, the all-electron method can be used to set a benchmark
for the pseudopotential method-based calculations. However, application of
this GW method to real poly-atomic systems and real problems in physics,
to a large extent, still relies on a good combination of these different kinds
of implementation.

4.4 Numerical Implementation of an All-Electron


G0 W0 Code: FHI-Gap
In the earlier discussions, we have explained the principles underlying the
current state-of-the-art single-particle excitation theory. In the following,
we take FHI-gap, an implementation of the G0 W0 approximation in which
one of the authors of this book (XZL) has participated, as an example
to show how these methods are implemented in a computer code. To be
simple, this introduction will focus on explaining some features which are
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 67

Many-Body Green’s Function Theory and the GW Approximation 67

common in an all-electron implementation of the G0 W0 method, even with


other basis sets. Therefore, the explanations of the technical details are far
from being complete. For these technical details, we strongly recommend
the readers to Ref. [212], where a very complete explanation of algorithms
is given.h
The basis set with which the KS wave functions are expanded is
the so-called (Linearized) Augmented Plane Wave plus local orbitals
((L)APW+lo) [167, 168], as explained in the previous chapter. Due to its
combined abilities to provide the most reliable results for periodic systems
within DFT and to address the widest range of materials, this method is
a FP method currently considered to be a benchmark in the DFT calcu-
lations. Consequently, the development of a G0 W0 code has constituted a
demanding task. The whole process has been imbued with the compromise
between computational efficiency and the numerical precision necessary to
achieve the ambitious goals of reliability and wide applicability complying
with the FP-(L)APW+lo standards.
Among the different existing implementations of the method, the
Wien2k code [213] is used as the basis on which the G0 W0 correction is
applied, although we acknowledge that nowadays it has been extended so
that other LAPW-based DFT codes such as Exciting [214–216] can also be
used to generate the input KS orbitals. In Sec. 4.4.1, we present a sum-
mary of the G0 W0 equations to be implemented. The representation of
the non-local operators (polarization, dielectric function, bare and screened
Coulomb potentials) requires an efficient basis set, able to address extended
as well as localized valence states and core states. An optimized set of func-
tions consisting of plane waves in the interstitial region and a spherical
harmonics expansion within the MT spheres can be used [178], which will
be introduced in Sec. 4.4.2. In Sec. 4.4.3, we summarize the matrix form of
the G0 W0 equations after expansion in the mixed basis.
Calculating the polarization for a given wave vector q requires an inte-
gration over all possible transitions from occupied to unoccupied states
and vice versa which conserve the total wave number. In other words, a
precise q-dependent Brillouin-zone integration is required. The efficiency of
the linear tetrahedron method is comparable to the special points methods

h Reference [212] is the Ph.D. thesis of XZL which is available online as “All-electron
G0 W0 code based on FP-(L)APW+lo and applications, Ph.D. Thesis, Free University of
Berlin, 2008”. Readers may contact XZL via [email protected] if the thesis is inaccessible.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 68

68 Computer Simulations of Molecules and Condensed Matter

for semiconductors and insulators, while it is clearly superior for metallic


systems. To be able to treat the widest possible range of materials, the
linear tetrahedron method is extended to the q-dependent case. A descrip-
tion of this development, together with the special requirements for the
polarization, are described in Sec. 4.4.4. The frequency convolution of the
correlation self-energy is presented in Sec. 4.4.5. This chapter concludes
with a flowchart of this FHI-gap code in Sec. 4.4.6.

4.4.1 Summary of the G0 W0 Equations

In the G0 W0 approach, the quasi-particle energy qpn,k is obtained from a


first-order correction to the KS energy eigenvalue n,k through

qp qp
n,k = n,k + ϕn,k (r1 ) |[Σ(r1 , r2 ; n,k )] − V
xc
(r1 ) δ (r1 − r2 ) |ϕn,k (r2 ),
(4.33)
where ϕn,k (r) and V xc are the KS eigenfunctions and XC potential, respec-
tively. The self-energy Σ(r1 , r2 ; ω) is obtained from the Fourier transform
of Eq. (4.31c):

i
Σ(r1 , r2 ; ω) = G0 (r1 , r2 ; ω + ω  )W0 (r2 , r1 ; ω  )dω  , (4.34)

where G0 is the Green’s function in the KS scheme defined by


 ϕn,k (r1 )ϕ∗n,k (r2 )
G0 (r1 , r2 ; ω) = , (4.35)
ω − n,k ± iη
n,k

and the dynamically screened Coulomb potential W0 (r2 , r1 ; ω) is given by



W0 (r1 , r2 ; ω) = ε−1 (r1 , r3 ; ω)v(r3 , r2 )dr3 . (4.36)

ε(r1 , r2 ; ω) is the dielectric function, it can be calculated from



ε(r1 , r2 ; ω) = 1 − v(r1 , r3 )P (r3 , r2 ; ω)dr3 , (4.37)

where the polarizability P (r1 , r2 ; ω), in the random phase approximation


(RPA), is written as

i
P (r1 , r2 ; ω) = − G0 (r1 , r2 ; ω + ω  )G0 (r2 , r1 ; ω  )dω  . (4.38)

December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 69

Many-Body Green’s Function Theory and the GW Approximation 69

The self-energy can be separated into the exchange and correlation


terms. If we define
W0c (r1 , r2 ; ω) = W0 (r1 , r2 ; ω) − v(r1 , r2 ), (4.39)
1
where v(r1 , r2 ) = |r1 −r 2|
is the bare Coulomb potential, the exchange and
correlation parts of the self-energy can be calculated from

x i
Σ (r1 , r2 ) = G0 (r1 , r2 ; ω  )v(r2 , r1 )dω 

occ

= ϕn,k (r1 )v(r2 , r1 )ϕ∗n,k (r2 ) (4.40)
n,k

and

c i
Σ (r1 , r2 ; ω) = G0 (r1 , r2 ; ω + ω  )W0c (r2 , r1 ; ω  )dω  (4.41)

separately.
The required input for solving this set of equations are the eigenfunc-
tions, (ϕn,k (r)), eigenvalues (n,k ) and the exchange–correlation potential
(Vxc (r)) of the KS orbitals, and these data can be obtained from a self-
consistent DFT calculation, and in this particular case, using the Wien2k
code [213].

4.4.2 The Mixed Basis

For periodic systems, the reciprocal space representation improves the effi-
ciency by exploiting explicitly the translational symmetry of the Bravais
lattice. However, a direct Fourier transform of the operators, which implies
taking plane waves as a basis set, is computationally inefficient for their
representation in a FP, all-electron implementation (see Chapter 3). Anal-
ogous to the proposal of Kotani and van Schilfgaarde [191], an optimized
set of functions satisfying Blöch’s theorem can be used for the expansion
of the operators as summarized in Sec. 4.4.1. This “mixed” basis set uses
the space partition in the MT and interstitial regions following the APW
philosophy. In this section, we introduce how this mixed basis set is defined.
By replacing the expression of Eq. (4.35) for Green’s function in
Eq. (4.38), one can see that the spatial dependence of the non-local polariz-
ability on each coordinate is a product of two KS wave functions. Thus, the
basis set one chooses to expand this polarizability should be efficient in rep-
resenting those products of two KS wave functions. From Sec. 3.2, we know
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 70

70 Computer Simulations of Molecules and Condensed Matter

that the KS wave functions are linear combinations of (L)APW+lo basis. In


other words, they are expanded in terms of spherical harmonics in the MT
spheres and plane waves in the interstitial region. The product of two plane
waves is still a plane wave and the product of two spherical harmonics can
be expanded in spherical harmonics using the Clebsch–Gordan coefficients.
Therefore, the same kind of space partition can be used. As a matter of
fact, this method is exactly what has been taken in FHI-gap to define the
basis set for the non-local operators used in the G0 W0 calculations. Inside
the MT sphere of atom α, the basis functions are defined as

γα,N,L,M (rα ) = vα,N,L (rα ) YL,M (rα ) . (4.42)

To obtain an optimal set of radial functions vα,N,L (rα ), one proceeds as


follows:

• u̇l (r)’s are not taken into account because |u̇l (r)|2 r2 dr is typically less
than 10% of |ul (r)|2 r2 dr. Possible errors will be taken care of by the
other basis functions.
• To truncate the expansion, one takes a maximum lmax for the choice of
ul (rα )’s.
• For each L in vα,N,L (rα ), one takes all the products of two radial functions
ul (rα )ul (rα ) which fulfill the triangular condition |l − l |  L  l + l .
• The overlap matrix between this set of product radial functions is calcu-
lated by

 Rα
MT
O (l,l );(l1 ,l1 ) = uα,l (rα )uα,l (rα )uα,l1 (rα )uα,l1 (rα )(rα )2 drα .
0
(4.43)

• The matrix O(l,l );(l1 ,l1 ) is diagonalized, obtaining the corresponding set
of eigenvalues λN and eigenvectors {cN l,l }.
• Eigenvectors corresponding to eigenvalues (λN ) smaller than a certain
tolerance λmin are assumed to be linear-dependent and discarded.
• The remaining eigenvectors, after normalization, constitute the radial

basis set: vα,N,L (rα ) = l,l cN α α
l,l ul (r )ul (r ).

So defined, the functions {γα,N,L,M } constitute an orthonormal basis


set. The translational symmetry of the lattice is imposed by taking the
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 71

Many-Body Green’s Function Theory and the GW Approximation 71

Blöch summation:
q 1  iq·(R+rα )
γα,N,L,M (r) = √ e γα,N,L,M (rα ), (4.44)
Nc R

where rα is the position of atom α in the unit cell, and R is a Bravais lattice
vector.
Since the interstitial plane waves are not orthogonal, one can diagonal-
ize the overlap matrix by solving the eigenvalue equation:

OG,G SG ,i = εi SG,i , (4.45)
G

where OqG,G ≡ PG


q q
|PG q
  and r|PG  = √
1 i(G+q)·r
V
e .
The orthogonal basis set in the interstitial region is defined by

P̃iq (r) ≡ q
S̃G,i PG (r). (4.46)
G

where S̃G,i = √1εi SG,i so that the orthogonal interstitial plane waves are
normalized. The plane wave expansion can be truncated at a certain Gmax .
In fact, a parameter Q that defines Gmax in units of GLAPW max (the plane
wave cut-off of the LAPW basis functions) can be introduced. Finally, the
orthonormal mixed basis set is
 
χqj (r) ≡ {γα,N,L,M
q q
(r), P̃G (r)}. (4.47)

4.4.3 Matrix Form of the G0 W0 Equations

The basis set as introduced above was derived from the requirement of
efficiency to expand products of KS eigenfunctions. The principal quantity
for such expansion, and also a central quantity for the whole implementation
are the matrix elements:


i
Mn,m (k, q) ≡ [χ̃qi (r)ϕm,k−q (r)] ϕn,k (r)dr. (4.48)
Ω

Following the expansion of non-local operators in periodic systems, we can


write the Coulomb potential matrix elements in the mixed basis as
  
vi,j (q) = (χqi (r1 ))∗ v(r1 , r2 − R)e−iq·R χqj (r2 )dr2 dr1 . (4.49)
Ω Ω R
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 72

72 Computer Simulations of Molecules and Condensed Matter

Using the matrix element defined in Eq. (4.48), the polarizability can
be calculated by
 occ unocc
BZ  
i j ∗
Pi,j (q, ω) = Mn,n (k, q)[Mn,n (k, q)]

k n n
 
1 1
· − .
ω − n ,k−q + n,k + iη ω − n,k + n ,k−q − iη
(4.50)
To avoid the divergence at q = 0 of the dielectric function as defined
in Eq. (4.37), one resorts to the symmetrized dielectric functioni defined as
 1 1
ε̃i,j (q, ω) = δi,j − 2
vi,l 2
(q)Pl,m (q, ω)vm,j (q). (4.51)
l,m

Using this equation, the correlation term of the screened Coulomb interac-
tion can be calculated through
 1 1
c
Wi,j (q, ω) = 2
vi,l (q)[ε̃−1
l,m − δl,m ](q, ω)vm,j .
2
(4.52)
l,m

The diagonal matrix element of the self-energy in the basis of the KS


states is
Σcn,k (ω) = ϕn,k |Σc (r1 , r2 ; ω)|ϕn,k 
BZ
i 
i ∗ j
= Mn,n (k, q) Mn,n (k, q)
2π q i,j 
n
 c
∞ Wi,j (q, ω  )
· dω  (4.53)
−∞ ω + ω  − n ,k−q ± iη
for the correlation term, and
Σxn,k = ϕn,k |Σx (r1 , r2 )|ϕn,k 
BZ 
 occ

i
∗ j
=− vi,j (q) Mn,n (k, q) Mn,n (k, q) (4.54)
q i,j n

for the exchange term.

i Formore details about this symmetrized dielectric function and how the divergence of
the related screened Coulomb potential is treated for its contribution to the self-energy,
please refer to Ref. [212].
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 73

Many-Body Green’s Function Theory and the GW Approximation 73

The quasi-particle energies should be obtained by solving Eq. (4.33)


self-consistently, since the self-energy depends on the quasi-particle energy
qp KS
n,k . Usually, a first-order Taylor expansion of the self-energy around n,k
is used instead. The quasi-particle energies are then given by
⎡ ⎤
c
  
KS  KS
 
QP KS ⎣ KS x  KS ⎦ ,
n,k = n,k + Zn,k n,k + Σn,k − ϕn,k VXC (r) ϕn,k
n,k

(4.55)

where
⎡ ⎛ ⎞ ⎤−1
c

⎢ ∂ ⎥
Zn,k = ⎣1 − ⎝ (ω)⎠ ⎦ . (4.56)
∂ω
n,k
KS
n,k

In this implementation of the G0 W0 method in FHI-gap, these two equa-


tions are used to get a first set of quasi-particle energies, which are then
taken as the starting point to solve Eq. (4.33) iteratively with respect to
qp
n,k . To ensure the conservation of the quasi-particle number, the quasi-
particle Fermi level is aligned to the KS one after each iteration for
Eq. (4.33). The iteration is stopped when this shift is smaller than a chosen
tolerance.

4.4.4 Brillouin-Zone Integration of the Polarization

Brillouin-zone integration is an important ingredient of any reciprocal


space method and has been a subject of intense interest since the earli-
est implementation of electronic structure codes. Fundamental quantities
like the total energy or the density of states require an integration over
the Brillouin zone of a certain operator, e.g. the eigenvalues weighted by
the Fermi distribution function for the former, the energy derivative of the
Fermi distribution for the latter.
In the 1970s, a large number of studies were carried out for solving
these problems, among which the special points [217–220] and the linear
tetrahedron method [221–223] are the ones most used nowadays. These
two methods perform equally well for insulators and semiconductors. For
metals, the Brillouin-zone integration becomes more cumbersome due to
the presence of the Fermi surface, which defines the integration region in
the Brillouin zone. The linear tetrahedron method becomes advantageous
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 74

74 Computer Simulations of Molecules and Condensed Matter

Figure 4.6 Two-dimensional sketches of the description of the Fermi surface in the
special points (a) and the linear tetrahedron method (b). The red dotted line shows the
exact Fermi surface, and the blue solid line the approximated one. The k-points grid
is represented by dots. The tetrahedron method gives better description of the Fermi
surface.

in these systems due to its better description of the Fermi surface (Fig. 4.6)
and, therefore, of the integration region [224].
In the linear tetrahedron method, first proposed by Jepsen and
Andersen [221] and Lehmann et al. [222], the Brillouin zone is divided into
a set of tetrahedra. The energy eigenvalues (n,k ) and the integrand are
calculated on the vertices of these tetrahedra and, through the procedure
known as isoparametrization, linearly interpolated inside each of them. The
values of the integrand can be factorized out of the integral. The remain-
ing integrals, independent of the values at the vertices, can be integrated
analytically and added to obtain integration weights dependent only on
the k-point and the band index (Appendix C). In metallic systems, the
Fermi surface is approximated, through the isoparametrization, by a plane
that limits the integration region inside the tetrahedra it intersects. The
occupied region of the Brillouin zone can thus be described much better
than in any of the special points methods (Fig. 4.6).
The calculation of quantities like the polarizability (Eq. (4.50)) or mag-
netic susceptibility presents particular characteristics that require a differ-
ent treatment. The integral depends on a second vector q, it is weighted
by two Fermi functions, so that the states at k are occupied while those at
k − q are unoccupied, and finally, the eigenvalues appear in the denomi-
nator of the integrand. The grid of k-points for this integration is chosen
as usual in the tetrahedron method. On the other hand, the calculation of
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 75

Many-Body Green’s Function Theory and the GW Approximation 75

Figure 4.7 The integration region in the tetrahedron method for q-dependent Brillouin-
zone integration. The two tetrahedra on the left side are connected by the vector q
(arrow). The right zone corresponds to the occupied region for the state (n, k), the
middle one to the unoccupied region for the state (n , k − q). The resulting integration
region, determined by superimposing the two tetrahedra on the left and taking the
intersection of the right and middle zones, is the left region in the tetrahedron on the
right-hand side.

the self-energy (Eqs. (4.53) and (4.54)) requires a grid of q-points to be


suitable also for integration. To avoid the repeated generation of eigenval-
ues and eigenvectors at several different grids, the set of q-points should be
such that {k} = {k − q}. For this equality to hold, the set of q-points must
include the Γ point. In FHI-gap, the same mesh for k and q is taken.
Due to the presence of the eigenvalues in the denominator of the inte-
grand, a simultaneous isoparametrization of both, the integrand and the
eigenvalues, becomes inappropriate. In 1975, Rath and Freeman proposed
a solution to this problem for the calculation of the static magnetic suscep-
tibilities in metals [223]. They approximated the numerator of the integrand
by its mean value in each tetrahedron, while the denominator was included
in the analytic integration to obtain the weights. In FHI-gap, two further
steps were taken. First, the isoparametrization was applied not only to
the eigenvalues but also to the numerator of the integrand, which signifi-
cantly improves the accuracy. Besides this, we also extended the method to
include the frequency dependence. In doing so, the integration inside each
tetrahedron can finally be performed analytically (see Appendix C).
Since the integration runs simultaneously on two tetrahedra (at k and
k − q), there will be situations, in metallic systems, where both tetrahedra
are intersected by the Fermi surface. In this case, the integration region
inside the tetrahedron is delimited by the two Fermi “planes” under the
condition nk < F < mk−q , as shown in Fig. 4.7. The complexity of the
integration region is such that the integration cannot be performed ana-
lytically on the whole tetrahedron as in the standard tetrahedron method.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 76

76 Computer Simulations of Molecules and Condensed Matter

However, as pointed out in Ref. [223], the integration region can always be
subdivided into, at most six, tetrahedra. The integration can be performed
analytically inside each of these tetrahedra and then projected onto the
vertices of the original tetrahedron to obtain the weights for each k-point.
These different configurations of the distinct integration regions determined
by two Fermi “planes” were then analyzed and categorized so that computer
programming can be applied (see Fig. C.2 in Appendix C).
To test the accuracy and stability of this implementation, the static
polarizability of the free electron gas can be calculated and compared to
its well-known analytical solution (the Lindhard function). The results are
shown in Fig. 4.8. This tetrahedron method performs really well for the free
electron gas, which is one of the most demanding examples for the Brillouin-
zone integration. Comparison of Fig. 4.8(a) with Fig. 4 in Ref. [223] shows
that this implementation achieves a comparable accuracy with a coarser
mesh (a 13 × 13 × 13 mesh in our calculation compared to the 24 × 24 × 24
mesh used in Ref. [223]).

4.4.5 The Frequency Integration

Then we discuss the frequency convolution for the correlation term of the
self-energy in Eq. (4.53). Due to the poles of both Green’s function and
W c , infinitesimally close to the real axis (Fig. 4.9), this integral is difficult
to converge numerically, requiring a large number of frequencies.
Several schemes have been proposed to improve the computational effi-
ciency in the evaluation of this convolution. One of the first methods was
proposed by Godby, Schlüter, and Sham [225]. Using the idea of the well-
known Matsubara summation [124, 126], which analytically continuates the
integrand into the complex frequency plane and calculates the integral over
the real frequency axis from the integral over the imaginary axis plus the
sum of the residues corresponding to the poles of Green’s function between
the given frequency and the Fermi energy, they have shown that the simple
form of the integrand in the imaginary axis allows a precise calculation of
the integral with few frequencies only. In a different approach, proposed by
Rieger et al. [227], the screened Coulomb potential is Fourier transformed to
the imaginary-time axis. The self-energy is then obtained by direct product,
according to Eq. (4.31c) and transformed back to the imaginary frequency
axis. Afterwards, it is fitted by an analytic function and continued to the
complex plane to obtain its dependence on the real frequency axis.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 77

Many-Body Green’s Function Theory and the GW Approximation 77

(a)

(b)

Figure 4.8 Comparison of the numerical (points) and analytical (line) results for the
polarizability of the free electron gas (Lindhard function) as function of q (a) on the
(100) direction and (b) on other directions.

In FHI-gap, the screened Coulomb potential, Green’s function and the


self-energy were calculated directly on the imaginary frequency axis. Equa-
tion (4.53) in this case becomes
BZ
1 
i ∗
Σcn,k (iω) = Mn,n (k, q)
π q i,j 
n
 c
∞ (n ,k−q − iω)Wi,j (q, iω  )  j
· dω Mn,n (k, q), (4.57)
0 (iω − n ,k−q )2 + ω 2
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 78

78 Computer Simulations of Molecules and Condensed Matter

Figure 4.9 The analytic structure of Σc = iGW c for ω > μ (a) and ω < μ (b). Frequency
integration of the self-energy along the real axis is equivalent to the integration along
the imaginary axis including the path C.

where we have made use of the inversion symmetry of W c on the imaginary


frequency axis:

c c
Wi,j (q, iω) = Wi,j (q, −iω). (4.58)

The integrand in Eq. (4.57) is singular when ω = ω  and n ,k−q = 0. There-
fore, a direct numerical integration becomes unstable for small eigenvalues.
The numerical details, as well as the method to avoid this instability, are
shown in Appendix 7.5. In the end, each matrix element of the self-energy
is fitted with a function of the form:
m j
j=0 an,k,j (iω)
Σcn,k (iω) = m+1 , (4.59)
j
j=0 bn,k,j (iω)

which is then analytically continued onto the real frequency axis.


December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 79

Many-Body Green’s Function Theory and the GW Approximation 79

Figure 4.10 Flowchart of the FHI-gap G0 W0 code.

4.4.6 Flowchart

We conclude this chapter with a short summary of the procedure carried


out by the implementation of the G0 W0 method as introduced above.
The flowchart is shown in Fig. 4.10. The KS eigenvalues and eigenfunc-
tions, as well as the XC potential are taken from Wien2k. The eigenfunc-
tions are used to obtain the mixed basis as was described in Sec. 4.4.2.
Having defined the basis functions, one can calculate the bare Coulomb
i
matrix and the matrix elements Mn,n  (k, q). Afterwards, the KS eigen-

values are required for calculating the Brillouin-zone integration weights


December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch04 page 80

80 Computer Simulations of Molecules and Condensed Matter

as described in Sec. 4.4.4 and Appendix C. Together with the matrix ele-
i
ments, Mn,n  (k, q), these weights are used to obtain the polarization matrix

(Eq. (4.50)). The latter, together with the bare Coulomb matrix, is the input
required to obtain the dielectric matrix and then the screened Coulomb
potential. With these, the matrix elements of the exchange and correla-
tion terms of the self-energy can be calculated separately. The coefficients
of the expansion described in Sec. 3.2.4 of the XC potential are obtained
from the Wien2k code and used to obtain its diagonal matrix elements
for each eigenstate. Finally, a first set of G0 W0 quasi-particle energies are
obtained by solving Eq. (4.55), which are used as the starting point for
solving Eq. (4.33) self-consistently.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 81

5
Molecular Dynamics

So far, we have discussed the electronic structures; from now on, we shift
our attention to descriptions of the nuclear motion.
As introduced in Chapter 1, molecular dynamics (MD) is a technique
which allows us to investigate the statistical and dynamical properties
of a real poly-atomic system at the molecular level. When the Born–
Oppenheimer (BO) approximation is further imposed, this simplifies into
reproducing the propagation of the nuclei on their BO potential-energy
surfaces (PESs). In standard treatment of such propagations, the classi-
cal equations of motion governing the microscopic evolution of a many-
body entity (composed by nuclei) are often solved numerically, subject to
the boundary conditions appropriate for the geometry and/or symmetry
of the system. Therefore, the method is, in principle, classical and these
nuclei are classical point-like particles in this treatment. We note that there
are extensions of this method to the quantum (in the statistical perspec-
tive) or semiclassical (in the dynamical perspective) regime, where some
of the quantum nuclear effects (QNEs) can be explicitly addressed. In this
chapter, to be clear, we restrict ourselves to the classical picture that the
nuclei are point-like particles for the illustration of the basic underlying
principles. Its extension to the descriptions of QNEs will be described in
Chapter 7.
Depending on how the interatomic interactions are calculated, the MD
simulations can be separated into two categories: the traditional classi-
cal MD simulations, where the interactions between different nuclei are
described by empirical potentials, and the more recent ab initio MD simu-
lations, where the interactions between the nuclei are calculated on-the-fly
using the Hellmann–Feynman theorem after the electronic structures of
the system at each specific spatial configuration of the nuclei were obtained

81
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 82

82 Computer Simulations of Molecules and Condensed Matter

quantum mechanically in an ab initio manner. Throughout this book, we


focus on the ab initio methods and consequently the ab initio MD, but
we note that the principles underlying the propagation of the nuclei, as
introduced here, also apply to those simulations where empirical potentials
are used.
Within the framework of ab initio MD, it is also worthwhile to note
that there is a scheme when the BO approximation is not rigorously done,
but rather, an adiabatic separation between the electronic and the nuclear
degrees of freedom has been assumed using a special treatment for the
movement of the electrons and the nuclei, like the Car–Parrinello (CP)
MD [228]. There is a scheme where the electronic structure optimization is
rigorously performed at each spatial configuration of the nuclei along the
trajectory, like the BO MD. The former treatment is an approximation to
the latter one. Due to the lower computation cost which originates from the
fact that the self-consistent electronic structure optimization does not have
to be done at each MD step — rather, a fictitious small mass is assigned
to the electronic degrees of freedom which enables an adiabatic separation
between the movements of the electrons and the nuclei — the former treat-
ment has been very popular from the 1980s till now. The BO MD method
only became a standard routine after the late 1990s when the improvement
of computational algorithms and the development of greater computing
power made it feasible to carry out a self-consistent electronic structure
optimization at each MD step. Because of this, the development of some
methods within the framework of ab initio MD, especially those related
to the thermostats, has been largely influenced by this historical trend. In
later discussions, we will point this out in detail when those methods are
introduced.
Our discussions in this chapter are categorized mainly into two parts.
Section 5.1 presents an introduction to the ideas behind MD. With these
concepts, our molecular simulation is able to sample a micro-canonical
ensemble and its phase space on a constant-energy shell. In real experi-
ments, however, it is the temperature that is kept constant. Section 5.2 then
gives us an introduction to some methods which go beyond the simulations
of the micro-canonical ensemble. Techniques underlying numerical controls
of the molecular temperature (thermostats) are the basis for such simula-
tions. Then, more sophisticated situations when other ensembles — such
as the constant-pressure and constant-temperature (isothermal–isobaric)
ones — must be simulated will also be discussed. With these, in principle,
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 83

Molecular Dynamics 83

one should be able to do a proper simulation for a system under the situation
to be studied and compare the results with experiments.

5.1 Introduction to Molecular Dynamics


The foundation of modern MD technique is the so-called Hamiltonian
mechanics, which is often used in descriptions of a micro-canonical ensem-
ble in Gibbs’ language. Hamiltonian mechanics is, in principle, equivalent
to the simplest Newton’s equation of motion for the description of the
nuclear propagations, with the only difference being that the equations of
motion can be derived easily in non-Cartesian coordinates. Because of this
advantage, properties in the phase space can be discussed in an easier way.
Therefore, although it provides nothing more than what can be obtained
by solving the Newton’s equation of motion, we prefer using Hamiltonian
mechanics for discussions in this chapter.
Within this framework of Hamiltonian mechanics, where the Hamilto-
nian’s equation of motion for the nuclei are numerically solved, the simula-
tions provide descriptions for the statistical properties of the system within
a certain constant-energy shell in the phase space. There is neither particle
exchange, energy exchange between the system under investigation and
its environment, nor any change of the system’s volume. In other words,
most essentially, the total energy (Hamiltonian) is a conserved quantity.
Later, extensions of this Hamiltonian mechanics to the non-Hamiltonian
ones were proposed, where the total energy of the system is not conserved
and other ensembles such as the canonical or isothermal–isobaric ones can
be simulated. These different ensembles serve different purposes in simu-
lations of the real materials. To be clear, we start our discussions from
the simplest Hamiltonian simulations of the micro-canonical ensemble (or
otherwise N V E ensemble, where N stands for the number of the parti-
cles, V for the volume of the cell under simulation, and E for the total
energy) and then extend these techniques to the non-Hamiltonian scheme
and simulations of other ensembles.
Intuitively, the three key quantities in this MD simulation of a
micro-canonical ensemble within the framework of Hamiltonian mechanics
should be the spatial coordinate of each nucleus (ri ), the velocity of each
nucleus (vi ), and the force imposed by the quantum glue of the electrons
and the classical internuclear Coulomb interactions on each nucleus (Fi ) at
time zero. These three quantities, together with the Hamiltonian’s equation
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 84

84 Computer Simulations of Molecules and Condensed Matter

of motion, determine the nuclear evolution of the system and consequently,


the equilibrium and dynamical properties of the system under investigation.
The entity of the interacting nuclei can be viewed as a classical many-
body system, with interactions between the nuclei changing as the system
evolves. For such a many-body entity, it is impossible to solve this equation
of motion analytically. Therefore, one needs to resort to numerical integra-
tions for the equation of motion.

5.1.1 The Verlet Algorithm

As mentioned above, the first step in carrying out a numerical integration


for the equation of motion is to set up a proper initial state of the system.
This includes setting the initial spatial configuration of the nuclei, their
initial velocities, and the initial forces imposed on them. Then the trajec-
tory of the nuclei can be integrated out numerically in principle. Seemingly
simple, the integration of such an equation of motion is highly non-trivial
due to the fact that a finite-time step must be used. Error accumulation
ultimately results in two initially-close trajectories diverging exponentially
as the dynamic evolves.
Fortunately, the MD simulations differ in principle from the trajec-
tory predictions in classical dynamics, such as in the study of the comet’s
orbital (see e.g. Ref. [48, Sec. 4.3]). What we are primarily aiming at in the
MD simulations is some knowledge on the statistical property of the system
whose initial state we already know something about, e.g. the energy. In
the language of the Gibbs’ ensemble (the micro-canonical ensemble in this
case), many individual microscopic configurations of a large system lead to
the same macroscopic behaviors. This statement implies that it is unnec-
essary to know the precise propagation of every nucleus in the system in
order to predict its macroscopic properties. Rather, a simple average over a
large number of states visited during the simulation will give us properly-
formulated macroscopic observables. As long as our integrator respects the
fact that the Newton’s equation of motion is time-reversible and the cor-
responding Hamiltonian dynamics leaves the volume of the phase space
explored in the MD simulation unchanged, the energy conservation can be
guaranteed if the finite-time step between movements are small enough in
practical simulations [48]. This indicates that a micro-canonical ensemble
is already being simulated. As it is one of the successful schemes on imple-
menting such a propagation when the total energy and the time reversibil-
ity of the dynamics can be preserved, we take the Verlet algorithm as an
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 85

Molecular Dynamics 85

example and explain the underlying principles by going through the key
equations in the following [229].
In this algorithm, the spatial configuration of the system at time t + Δt
is determined by
Fi (t) 2
ri (t + Δt) = ri (t) + vi (t)Δt + Δt + O(Δt3 ) + O(Δt4 ), (5.1)
2mi
when the Taylor expansion is used. Here, O(Δt3 ) (O(Δt4 )) represents error
 
on the order of Δt3 (Δt4 ), which equals r (t)Δt3 /3! (r (t)Δt4 /4!). Fur-
ther expansions beyond O(Δt4 ) are neglected. We note that the quantities
calculated along the trajectories are only ri , vi , and Fi ; the terms O(Δt3 )
and O(Δt4 ) will not be calculated. But it is important to write them down
in Eq. (5.1) in order to show the order of accuracy for ri and vi calculated
along the trajectory.
From Eq. (5.1), it is easy to see that in reverse time, the spatial con-
figuration respects
Fi (t) 2
ri (t − Δt) = ri (t) − vi (t)Δt + Δt − O((Δt)3 ) + O((Δt)4 ).
2mi
(5.2)

If we sum up Eqs. (5.1) and (5.2), we can have


Fi (t) 2
ri (t + Δt) = 2ri (t) − r(t − Δt) + Δt + 2O(Δt4 ). (5.3)
mi
The O(Δt3 ) terms were cancelled out. Therefore, the evolution of the
nuclear position obtained from this integrator contains an error on the
order of Δt4 (quantities related to the lower-order expansions needed in
the analytical expression in Eq. (5.3), i.e. ri and Fi , are all calculated
along the trajectory).
To obtain the new velocity, we subtract Eqs. (5.1) by (5.2). In doing
so, we arrive at

ri (t + Δt) − ri (t − Δt) = 2vi (t)Δt + 2O(Δt3 ). (5.4)

The velocity is then updated by


ri (t + Δt) − ri (t − Δt) O(Δt3 )
vi (t) = − . (5.5)
2Δt Δt
From this equation, it is clear that the velocity contains an error on the
order of Δt2 .
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 86

86 Computer Simulations of Molecules and Condensed Matter

It is worthwhile to note that Eqs. (5.1) and (5.2) are the bases of the
Verlet algorithm, which is obviously time-reversible. Therefore, trajectories
obtained from the Verlet algorithm respect the time reversibility of the
Newton’s equation of motion. Disregarding the error terms which are not
calculated, in this Verlet algorithm, the trajectory is propagated through
Fi (t) 2
ri (t + Δt) = 2ri (t) − ri (t − Δt) + Δt , (5.6)
mi
and
ri (t + Δt) − ri (t − Δt)
vi (t) = . (5.7)
2Δt
The positions and velocities contain errors on the orders of Δt4 and Δt2 ,
respectively.

5.1.2 The Velocity Verlet Algorithm

We note that in Eq. (5.6), the velocity is not used when the new position is
generated. There is an algorithm equivalent to the Verlet one in which the
velocity is used in the calculation of the new position. This is the so-called
velocity Verlet algorithm [230].
In this algorithm, the new position is calculated from
Fi (t) 2
ri (t + Δt) = ri (t) + vi (t)Δt + Δt . (5.8)
2mi
The velocity is updated by
Fi (t + Δt) + Fi (t)
vi (t + Δt) = vi (t) + Δt. (5.9)
2mi
We note that a key difference between this algorithm and the Euler scheme,
which updates the velocity by
Fi (t)
vi (t + Δt) = vi (t) + Δt, (5.10)
mi
is that the new velocity can only be calculated after the new position and
its corresponding forces are generated in the velocity Verlet algorithm. This
fine numerical difference actually results in the fundamental distinction that
the trajectory obtained from the Euler algorithm is not time-reversible and
the position contains an error on the order of Δt3 , whereas in the velocity
Verlet algorithm, the trajectory is time-reservable and the updated position
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 87

Molecular Dynamics 87

contains an error on the order of Δt4 , although this is not clear in Eqs. (5.8)
and (5.9).
To prove the equality between the velocity Verlet algorithm and the
original Verlet algorithm, we start from the original Verlet algorithm in
Eqs. (5.6) and (5.7) and try to deduce the velocity Verlet algorithm in
Eqs. (5.8) and (5.9). To this end, we first rewrite Eq. (5.6) into the following
form:
ri (t + Δt) − ri (t − Δt) Fi (t) 2
ri (t + Δt) = ri (t) + Δt + Δt . (5.11)
2Δt 2mi
Then, by replacing the (ri (t + Δt) − ri (t − Δt))/(2Δt) with vi (t) through
Eq. (5.7), we arrive at Eq. (5.8). Therefore, the updating of the position is
completely equivalent in these two algorithms.
Concerning the deduction of Eq. (5.9), from Eq. (5.8), we see that
ri (t + Δt) − ri (t) Fi (t)
vi (t) = − Δt. (5.12)
Δt 2mi
This equation is equivalent to
ri (t + 2Δt) − ri (t + Δt) Fi (t + Δt)
vi (t + Δt) = − Δt, (5.13)
Δt 2mi
which further equals
ri (t + 2Δt) − ri (t) ri (t + Δt) − ri (t) Fi (t + Δt)
vi (t + Δt) = − − Δt.
Δt Δt 2mi
(5.14)
Using Eq. (5.7), it can be reduced to
ri (t + Δt) − ri (t) Fi (t + Δt)
vi (t + Δt) = 2vi (t + Δt) − − Δt,
Δt 2mi
(5.15)
which equals
ri (t + Δt) − ri (t) Fi (t + Δt)
vi (t + Δt) = + Δt. (5.16)
Δt 2mi
From Eq. (5.12), we see that (ri (t + Δt) − ri (t))/Δt equals vi (t) +
Fi (t)Δt/(2mi ). Replacing this (ri (t + Δt) − ri (t))/Δt term with vi (t) +
Fi (t)Δt/(2mi ) in Eq. (5.16), one arrives at Eq. (5.9).
The above deduction means that starting from Eqs. (5.6) and (5.7),
one easily arrives at Eqs. (5.8) and (5.9). In other words, Eqs. (5.8) and
(5.9) propagate the trajectory in the same way as Eqs. (5.6) and (5.7), i.e.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 88

88 Computer Simulations of Molecules and Condensed Matter

the velocity Verlet and the Verlet algorithms are completely equivalent. The
position (velocity) contains error on the order of Δt4 (Δt2 ). They both differ
from the Euler algorithm in Eq. (5.10) in the sense that their trajectories
are time-reversible and preserve a higher order of accuracy in terms of Δt.

5.1.3 The Leap Frog Algorithm

The second frequently used algorithm for the propagation of the MD trajec-
tory, which is also equivalent to the Verlet algorithm, is the so-called Leap
Frog algorithm [48, 231]. In this algorithm, the velocities and positions are
updated by
Fi (t)
vi (t + Δt/2) = vi (t − Δt/2) + Δt, (5.17)
mi
and
ri (t + Δt) = ri (t) + vi (t + Δt/2)Δt. (5.18)
The propagation is equivalent to the Verlet algorithm with the only differ-
ence that the velocities are updated at half-time step.
For the completeness of this introduction, we use one more paragraph
to explain how these equations can be derived from Eqs. (5.6) and (5.7).
From Eq. (5.7), we have
ri (t) − ri (t − Δt)
vi (t − Δt/2) = , (5.19)
Δt
and
ri (t + Δt) − ri (t)
vi (t + Δt/2) = . (5.20)
Δt
From Eq. (5.20), it is easy to obtain Eq. (5.18) for the update of the posi-
tions. Concerning the deduction of Eq. (5.17), one can rewrite Eq. (5.6)
into the following form:
ri (t + Δt) − ri (t) ri (t) − ri (t − Δt) Fi (t)
= + Δt. (5.21)
Δt Δt mi
Then by inputting Eqs. (5.19) and (5.20) into Eq. (5.21), one gets
Fi (t)
vi (t + Δt/2) = vi (t − Δt/2) + Δt, (5.22)
mi
which is the same as Eq. (5.17). Therefore, the trajectory is equivalent to
that of the Verlet algorithm. However, due to the fact that the velocities
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 89

Molecular Dynamics 89

are not calculated at the same time slides as the positions, the total energy
cannot be rigorously calculated.
The Verlet, velocity Verlet, and the Leap Frog algorithms are the main
frequently used simple algorithms in standard implementations of the MD
method. For most MD simulations, they are sufficiently accurate. From
Eq. (5.1), it is easy to see that if higher orders of expansion for the updating
of the position in terms of Δt are included, one may use a larger time
step for the propagation of the trajectory and consequently, obtain higher
efficiency of the simulations. The predictor–corrector algorithm is one such
example [232]. However, these extensions may not respect the time-reversal
symmetry of the Newton’s equation and consequently may not lead to more
reliable sampling of the corresponding phase space. In this chapter, we
restrict ourselves to the simplest ones introduced and refer the readers
interested in more complicated ones to Refs. [48, 233].

5.2 Other Ensembles


When a proper initial state of the system under investigation is set up
and the trajectory is generated using the algorithms as introduced above,
if the time step is sufficiently small, the total energy E will be conserved
within a certain tolerance. On the basis of this, if the ergodic hypothesis
is further imposed — which means that given an infinite amount of time
the trajectory will go through the entire constant-energy hypersurface in
the phase space, a proper micro-canonical ensemble will be simulated. This
micro-canonical ensemble provides an accurate description for the statis-
tical/dynamical behavior of the system within a certain constant-energy
shell in the phase space, when there is neither particle exchange, energy
exchange between the system under investigation and its environment, nor
any change in the system’s volume [234–238]. Most experiments, however,
are performed in situations where these quantities are not conserved. The
simplest situation is when the experiment is carried out at a constant-
temperature T instead of the total energy E, due to the energy exchange
between the system under investigation and its environment. In this case,
the statistical behavior of the system at a finite temperature must be accu-
rately taken into account, and the states associated with the system in
the phase space visited during the time when the experiment is carried
out are not within a certain constant-energy shell anymore. Canonical,
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 90

90 Computer Simulations of Molecules and Condensed Matter

or in other words, the constant-volume and constant-temperature (N V T )


ensemble therefore needs to be simulated [239–246].
Theoretically, this can be done by introducing a proper thermostat in
the simulation, through which the energy exchange between the system
under investigation and the environment can be properly described. How-
ever, other thermodynamic variables such as the volume of the simulation
cell at finite pressure P may also change due to fluctuations. To describe
situations like this, a barostat needs to be further introduced which keeps
the tensor of the system constant during the simulations. This ensem-
ble with constant-temperature and constant-pressure is called an N P T ,
or isothermal–isobaric ensemble in Refs. [247–249]. In ascending order
of complexity, other ensembles which simulate experiments at different
situations, such as the isoenthalpic–isobaric (N P H) [238, 250–252], grand-
micro-canonical (μVL, with μ standing for the chemical potential) [238,
253], etc., also exist. In this chapter, we aim at providing an introduc-
tion to the principles underlying theoretical simulations of these ensembles.
Therefore, we will use the simplest case, i.e. the canonical (N V T ) ensem-
ble, as an example to explain how these principles work. Its extension to
the isothermal–isobaric (NPT ) ensemble will also be briefly discussed. For
readers interested in those methods for simulating more complicated ensem-
bles, please refer to their original papers [238, 250–253]. We note that in
these simulations, since the total energy is not conserved and the equation
of motion cannot be obtained from the Hamiltonian of the system under
simulations, they are often called the non-Hamiltonian MD simulations in
the literature.

5.2.1 Andersen Thermostat

There are several schemes for the temperature to be controlled. The under-
lying principle is that at a finite temperature T , the probability of finding
a classical particle, e.g. the ith atom in the system, with kinetic energy
p2i /(2mi ), which is labeled as P(pi ) throughout our discussions, respects
the Maxwell–Boltzmann distribution:
  32 p2
β −β i
P(pi ) = e 2mi . (5.23)
2πmi
Here, β = 1/(kB T ). The simplest implementation of this restriction on
simulation of real systems naturally is achieved through a stochastic colli-
sion of nuclei with a thermostat that imposes such a Maxwell–Boltzmann
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 91

Molecular Dynamics 91

distribution on the velocities. The corresponding thermostat is the so-called


Andersen thermostat [239]. When it is used, with a certain initial state,
the propagator as introduced in Sec. 5.1 ensures that the system propa-
gates on a certain constant-energy shell in the phase space between col-
lisions. The collisions themselves, on the other hand, randomly act on a
selected particle and reset its new velocity from the Maxwell–Boltzmann
distribution function at the desired temperature, which effectively reduces
correlation between events along the trajectory for increased sampling effi-
ciency, at the cost of accuracy in short-time dynamics. With a reasonable
choice of the collision frequency, the canonical ensemble with velocities
respecting the Maxwell–Boltzmann distribution can be conveniently simu-
lated. Reference [48, Chapter 6] gives details of how this is implemented,
including how the program should be written, which we do not cover. For
such details, please refer to their explanation.
Here, we note that although it is efficient and simple to implement
while imposing a canonical ensemble to be sampled, the stochastic collision
results in non-deterministic and discontinuous trajectories, which lead to
all real-time information being lost at collision. Related physical quanti-
ties, such as the vibrational frequencies, diffusion coefficients etc., therefore
cannot be calculated. In addition to this, more seriously, in spite of the
fact that a fairly ergodic sampling can often be guaranteed on all degrees
of freedom in such a stochastic simulation, when an adiabatic separation
between the electronic and ionic degrees of freedom must be assumed, like
in CP MD, a breakdown of such adiabatic separation often happens [254,
255]. As mentioned in the introduction of this chapter, ab initio MD has
been dominated by the CP MD treatment before the late 1990s due to
its lower computational cost. Therefore, deterministic methods such as the
Nosé–Hoover thermostat which avoid such a numerical problem have been
widely adopted since the 1980s [240–244]. It is only after the wide use of BO
MD that stochastic methods such as the Andersen or Langevin thermostat
became popular in ab initio MD simulations. Because of this historical
reason, when one goes through the literature, one may realize a change of
the taste in attempts to control the temperature influenced by this trend.
In this chapter, to present a general introduction of these methods and be
consistent with this historical trend, we will give a detailed explanation
of the principles underlying the Nosé–Hoover thermostat and its related
Nosé–Hoover chain. After this, we come back to another stochastic method,
i.e. the Langevin thermostat, which is widely used in ab initio BO MD
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 92

92 Computer Simulations of Molecules and Condensed Matter

simulations nowadays. We note that extensions of this method were also


designed to be compatible with CP MD recently [254].

5.2.2 Nosé–Hoover Thermostat

As one of the first successful attempts to incorporate deterministic tra-


jectories in the MD simulation at finite temperature, Nosé and Hoover
have separately used an extended system with an additional and artificial
degree of freedom to control the temperature [240–243]. The corresponding
thermostat is called the Nosé–Hoover thermostat. We note that this idea
of extracting a canonical ensemble out of a micro-canonical ensemble on
an extended system is similar to how the concept of canonical ensemble
is derived from the concept of micro-canonical ensemble in standard text-
books on statistics. In that case, it is required that the degree of freedom
in the thermostat is much larger than that in the canonical ensemble. Here,
a very clever choice of the potential, to which the added degree of free-
dom particle is subjected, ensures that one single extra degree of freedom
imposes a canonical ensemble on the system to be studied. In the follow-
ing, we will show how this works by going through the underlying key
equations.
The main assumption of this extended system method is that an addi-
tional coordinate s needs to be introduced to the classical N -nuclei system,
with the Lagrangian for the extended system

N
 mi Q 2 g
LNosé = s2 ṙ2i − V (r1 , . . . , rN ) + ṡ − lns. (5.24)
i=1
2 2 β

Here, we use “Nosé” to denote the extended system. Q is the effective mass
associated with the additional degree of freedom, g is a parameter to be
fixed, and [glns]/β is the external potential acting on this extra degree of
freedom. The corresponding Hamiltonian of this “Nosé” system can then
be written as

N
 1 1 2 g
HNosé = p2 + V (r1 , . . . , rN ) + p + lns, (5.25)
i=1
2mi s2 i 2Q s β

where s, ps , ri and pi are variables in the phase space and pi = ∂LNosé /


∂ ṙi = mi s2 ṙi .
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 93

Molecular Dynamics 93

From this Hamiltonian, the partition function for the micro-canonical


(Nosé) ensemble is defined as

1
ZNosé = dps dsdp1 · · · dpN dr1 · · · drN δ(HNosé − E), (5.26)
N!
Now, if we scale the momentum pi with the additional degree of
freedom by
pi = pi /s, (5.27)
the partition function in Eq. (5.26) simplifies into

1
ZNosé = dps dsdp1 · · · dpN dr1 · · · drN s3N
N!
 N  
 1 1 g
·δ p2 + V (r1 , . . . , rN ) + p2 + lns − E .
i=1
2mi i 2Q s β
(5.28)
Further simplification of this equation requires use of the following equation:
δ(s − s0 )
δ [h(s)] = , (5.29)
h (s0 )
where h(s) is a function of s with a single root for h(s) = 0 at s0 and h (s0 )
is its derivative at this point. Now, if we take
N 
 1 1 2 g
2
h(s) = p + V (r1 , . . . , rN ) + p + lns − E (5.30)
i=1
2mi i 2Q s β

in Eq. (5.28) and make use of Eq. (5.29), Eq. (5.28) can be simplified into

1 βs
ZNosé = dps dsdp1 · · · dpN dr1 · · · drN s3N
N! g
⎛ ⎡ N ⎤⎞
1 2 p2s
p
i=1 2mi i + V (r1 , . . . , rN ) + 2Q − E
· δ ⎝s − exp ⎣−β ⎦⎠
g


1 βs(3N +1)
= dps dsdp1 · · · dpN dr1 · · · drN
N! g
⎛ ⎡ N ⎤⎞
1 2 p2s
i=1 2mi pi + V (r1 , . . . , rN ) + 2Q −E
· δ ⎝s − exp ⎣−β ⎦⎠ .
g
(5.31)
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 94

94 Computer Simulations of Molecules and Condensed Matter

We integrate s out first. In doing so, we can get the partition function of
the “Nosé” system as

1 β
ZNosé = dps dp1 · · · dpN dr1 · · · drN
N! g
⎡ ⎛ ⎞ ⎤
N 1 2 p2s
i=1 2mi i p + V (r1 , . . . , rN ) + 2Q − E
· exp ⎣−β ⎝ ⎠ (3N + 1)⎦
g
  
1 βexp[Eβ(3N + 1)/g] 3N + 1 p2s
= dps exp −β
N! g g 2Q

· dp1 · · · dpN dr1 · · · drN exp
  N

3N + 1  1 2
· −β p + V (r1 , . . . , rN ) . (5.32)
g i=1
2mi i

This equation indicates that in the extended system with a Hamiltonian


as shown in Eq. (5.25), its partition function can be separated into a product
of two parts, i.e. that of a subsystem with variables pi and ri in the phase
space and that of a constant C, as
  
3N + 1
ZNosé = C dp1 · · · dpN dr1 · · · drN exp −β H(p , r) . (5.33)
g
Here,
  
1 βexp(Eβ(3N + 1)/g) 3N + 1 p2s
C= dps exp −β , (5.34)
N! g g 2Q
and
N
1 2
H(p , r) = p + V (r1 , . . . , rN ). (5.35)
i=1
2mi i

The equation of motion for this extended system, rigorously, is deter-


mined by the Hamiltonian in Eq. (5.25) through
dri ∂HNosé pi
= = , (5.36a)
dt ∂pi m i s2
dpi ∂HNosé ∂V (r1 , . . . , rN )
=− =− , (5.36b)
dt ∂ri ∂ri
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 95

Molecular Dynamics 95

ds ∂HNosé ps
= = , (5.36c)
dt ∂s Q
N 
dps ∂HNosé  p2 g
i
=− = − s. (5.36d)
dt ∂s i=1
m i s2 β

With the quasi-ergodic hypothesis that the time average along the trajec-
tory equals the ensemble average when the simulation time is sufficiently
long, the ensemble average of a certain physical quantity A that depends
on p and r in the “Nosé” system equals

 1 τ
A(p , r) = lim A(p , r)dt. (5.37)
τ →∞ τ 0

Here, · · ·  indicates the micro-canonical ensemble average of the extended


(Nosé) system, and the right-hand side is the time average over the trajec-
tory generated by Eq. (5.36). If we put in the partition function as simplified
in Eq. (5.33) for the “Nosé” system to the left-hand side of Eq. (5.37), we
can have
  
dp1 · · · dpN dr1 · · · drN A(p , r)exp −β 3Ng+1 H(p , r)
A(p , r) =    .
dp1 · · · dpN dr1 · · · drN exp −β 3Ng+1 H(p , r)
(5.38)

Now, if we take g = 3N + 1, this further simplifies into


1

dp · · · dp dr1 · · · drN A(p , r)exp [−βH(p , r)]
A(p , r) = N ! 1 1  N  .
N! dp1 · · · dpN dr1 · · · drN exp [−βH(p , r)]
(5.39)

We label the right-hand side of the above equation as A(p , r)NVT , since
it is equivalent to the canonical ensemble average of the operator A(p , r)
in a system with Hamiltonian H(p , r) at T . In other words,

A(p , r) = A(p , r)NVT . (5.40)

From this equality, it is clear that A(p , r)NVT , the canonical ensemble
average of A(p , r) in the subsystem, can be evaluated using the micro-
canonical ensemble average of the extended system. Numerical, if the quasi-
ergodic hypothesis is further imposed, by combining Eqs. (5.37) and (5.40),
this canonical ensemble average further equals the time average of this
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 96

96 Computer Simulations of Molecules and Condensed Matter

physical quantity along the deterministic trajectory of the “Nosé” system,


through
 τ
1
A(p , r)NVT = lim A(p , r)dt. (5.41)
τ →∞ τ 0

In other words, imagining that the real system we want to investigate is


subjected to the Hamiltonian in Eq. (5.35), Eq. (5.40) means that the
canonical ensemble of a certain physical quantity in this real system can
be evaluated using the micro-canonical ensemble average of an artificial
extended system, in which a deterministic trajectory is generated. When
the quasi-ergodic hypothesis is imposed on the MD simulation of the micro-
canonical extended system, the equality in Eq. (5.41) further means that
this canonical ensemble average of the physical quantity in the subsystem
is completely equivalent to the time average of this physical quantity on
the deterministic trajectory of the extended system.
Till now, the principle underlying the Nosé–Hoover thermostat is clear.
The micro-canonical distribution of the extended (“Nosé”) system with an
augmented variable s and Hamiltonian as shown in Eq. (5.25) is equivalent
to a canonical distribution of the subsystem with variables pi and ri , where
pi is the scaled momentum pi /s. This feature is obtained by artificially
introducing an additional degree of freedom s, which is subjected to an
external potential of a special form, and the property that the canonical
ensemble can be sampled this way is guaranteed by the above equations. We
note that the trajectory in Eq. (5.41) needs to be generated by Eq. (5.36),
using an even distribution of the time interval dt.
From Eqs. (5.27), (5.36) and (5.39), one can also see that the scaled
momentum pi is the real momentum of the canonical system under inves-
tigation and the momentum pi before scaling is only for the artificial
extended “Nosé” system. Because of this difference, from now on, we call
the variables related to pi the real variables and those related to pi the
virtual ones. Since Eq. (5.36) is the equation of motion from which the
trajectory is determined, the ensemble average in Eq. (5.39) must follow an
even distribution of the virtual time (t). Following this rigorous route that
the trajectory is generated for the extended system with an even distribu-
tion of the virtual time and the propagation generated from Eq. (5.36) is on
the virtual variables, while the ensemble average is done for the subsystem
with real momentum pi , there exists some obvious inconsistency which is
cumbersome to handle.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 97

Molecular Dynamics 97

To avoid such an inconsistency where the trajectory is calculated and


the ensemble averages are performed in different ways, we can rewrite the
equations of motion in Eq. (5.36) by scaling the key variables in the virtual
system to the real one through: ri = ri , pi = pi /s, s = s, ps = ps /s,
and dt = dt/s. In terms of these real variables, the equation of motion in
Eq. (5.36) can be rewritten as

dri dri pi pi pi


= s = s = = , (5.42a)
dt dt m i s2 mi s mi
dpi d(pi /s) dpi 1 ds ∂V (r1 , . . . , rN ) ds 

=s = − pi = − − pi
dt dt dt s dt ∂ri dt
∂V (r1 , . . . , rN ) s ps 
=− − p, (5.42b)
∂ri Q i
ds ds sps s2 ps
= s = = , (5.42c)
dt dt Q Q
N 
dps d(ps /s) dps 1 ds 1  p2 i g s p2
s
= s = − p s = − − . (5.42d)
dt dt dt s dt s i=1 mi β Q

The deterministic trajectory of a system involving variables ri = ri , pi =


pi /s, s = s, ps = ps /s, and dt = dt/s can be determined from Eq. (5.42).
Along this trajectory, the conserved quantity is

N
 1 2 s2 p2
s g
HNosé = pi + V (r1 , . . . , rN ) + + lns . (5.43)
i=1
2mi 2Q β

But we note that this is not the Hamiltonian of the system since the equa-
tions of motion cannot be obtained from it. The MD simulation performed
in this manner, accordingly, is called non-Hamiltonian MD in Ref. [246].
Now, we can generate a trajectory using Eq. (5.42) which satisfies the
equation of motion defined by the Hamiltonian in Eq. (5.25). Accordingly,
the expectation value of a physical quantity A(p , r ) should be evalu-
ated with it. However, we note that the application of the quasi-ergodic
hypothesis as shown in Eq. (5.41), which relates the ensemble average of
the extended micro-canonical ensemble to the real canonical subsystem, is
achieved by using the trajectory generated with Eq. (5.36) for the extended
system, where an even distribution of dt is imposed. In order for this quasi-
ergodic hypothesis to be applied to the new trajectory of the extended
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 98

98 Computer Simulations of Molecules and Condensed Matter

system, where an even distribution of dt is used, a mathematical transfor-


mation and justification must be made, which we will show in the following
paragraphs.
Along the new trajectory, the time average of a certain physical quan-
tity, e.g. A(p(t )/s(t ), r(t )), can be evaluated by
 τ
1
lim A(p(t )/s(t ), r(t ))dt , (5.44)
τ →∞ τ 

0

according to the quasi-ergodic hypothesis, where an even distribution on


dt is imposed. Since dt = dt/s and the trajectory is determined for the
real subsystem, the virtual time interval dt varies in each time step during
the MD simulation. The τ  in Eq. (5.44) is related to the virtual time τ in
Eq. (5.37) through
 τ
1
τ = dt. (5.45)
0 s(t)

We note that till now, along the trajectory generated by Eq. (5.42), the
time average of the physical quantity as given in Eq. (5.44) is not related
to the ensemble average of any kind. In order to relate it to an ensemble
average of some kind, we can use Eq. (5.45) and rewrite Eq. (5.44) in terms
of the trajectory in the virtual time as
 τ  τ
1 τ 1 A(p(t)/s(t), r(t))
lim A(p(t )/s(t ), r(t ))dt = lim dt
τ  τ  →∞ 0 τ  τ τ →∞ 0 s(t)
1

A(p(t)/s(t),r(t))
τ limτ →∞ 0 s(t) dt
=
τ  /τ
1
τ A(p(t)/s(t),r(t))
τ limτ →∞ 0 s(t) dt
= 1
τ 1 . (5.46)
τ limτ →∞ 0 s(t) dt

Here, τ1 limτ →∞ 0 · · · dt is the time average of a physical quantity along
the trajectory in the virtual time (that generated by Eq. (5.36)).
Then, making use of Eq. (5.37), these quantities related to the time
average of the virtual trajectory can be further simplified into
 
A(p(t)/s(t),r(t))
s(t)
  , (5.47)
1
s(t)
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 99

Molecular Dynamics 99

where · · ·  denotes the ensemble average of the extended (Nosé) system.


Now, we can make use of the partition function in Eq. (5.31) again and
further simplify Eq. (5.47) into
 
A(p(t)/s(t),r(t))
s(t)
 
1
s(t)
R 
dp1 ···dpN dr1 ···drN A(p ,r)exp[−β 3N 
g H(p ,r)]
R   3N +1
dp1 ···dpN dr1 ···drN exp[−β g H(p ,r)] 

=  R 
dp1 ···dpN dr1 ···drN exp[−β 3N 
g H(p ,r)]
R 3N +1
dp1 ···dpN dr1 ···drN exp[−β g H(p ,r)]


dp1 · · · dpN dr1 · · · drN A(p , r)exp[−β 3N 
g H(p , r)]
= 
dp1 · · · dpN dr1 · · · drN exp[−β 3N 
g H(p , r)]

1
    3N 
N ! dp1 · · · dpN dr1 · · · drN A(p , r)exp[−β g H(p , r)]
= 1
 . (5.48)
N! dp1 · · · dpN dr1 · · · drN exp[−β 3N 
g H(p , r)]

From this equation, it is clear that if we take g = 3N , this quantity equals


the ensemble average of a canonical system simulated by Eq. (5.42), which
we denote as · · · c , and by linking Eqs. (5.44), (5.46), (5.47), with (5.48),
we can get
 τ
1
lim A(p(t )/s(t ), r(t ))dt = A(p , r)c . (5.49)
τ  τ  →∞ 0

It is worthwhile to note that in Eq. (5.46), the quantity to be integrated has


a denominator s(t), which cancels one s(t) in the s(t)3N +1 in Eq. (5.31). It
is this tiny difference that results in the fundamentally different ways for
the canonical ensemble to be sampled by the trajectories from Eqs. (5.36)
and (5.42). In the former case, the “Nosé” thermostat simulates a canoni-
cal ensemble using trajectory generated with an evenly distributed virtual
time, and the parameter g needs to be set as 3N + 1. In the latter case,
the refined form of the “Nosé” thermostat simulates a canonical ensemble
using trajectory generated with an evenly distributed real time, and the
parameter g is set as 3N . In most of the current implementations of this
thermostat, the latter form is used.
We note that in real implementations of Eq. (5.42), it can be further
simplified by replacing the variable s by ξ  and ps by pξ , where ξ  = lns
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 100

100 Computer Simulations of Molecules and Condensed Matter

and pξ = s ps . In doing so, Eq. (5.42) can be rewritten as


dri p

= i, (5.50a)
dt mi
dpi 
∂V (r1 , . . . , rN ) pξ 
= − − p, (5.50b)
dt ∂ri Q i
dξ  pξ
= , (5.50c)
dt Q
N
dpξ
= p2
i /mi − g/β. (5.50d)
dt i=1

Here, Q is still a parameter we need to set in our simulations which is


associated with the mass of the particle in the extra degree of freedom.
From Eq. (5.50b), it is clear that pξ is associated with the friction coefficient
in controlling the temperature. Since the variables ri , pi , dt are the real
variables, g equals 3N . The energy which is conserved (Eq. (5.43)) is now
reformed to
N
 p2
i
p2
ξ g
HNosé = + V (r1 , . . . , rN ) + + ξ. (5.51)
i=1
2mi 2Q β

Again, we note that this is not a real Hamiltonian since the equations
of motion cannot be generated from it. It is a quantity which should be
conserved during the MD simulation.

5.2.3 Nosé–Hoover Chain

As simple as it is, the Nosé–Hoover thermostat succeeded in sampling


the canonical ensemble in a series of systems. There are, however, many
situations when the quasi-ergodic hypothesis fails. A 1D system when the
external potential is a harmonic oscillator is such an example. In this
case, the Andersen thermostat still works well on sampling the Maxwell–
Boltzmann distribution of the velocities. The Nosé–Hoover thermostat,
however, only samples a micro-canonical ensemble [242]. Now, it has been
realized that the reason for this breakdown is due to the fact that in the
deduction of the equations from Eq. (5.28) to Eq. (5.33), it is essential to
identify all the conserved quantities [246].
In the 1990s, a scheme in which the Nosé–Hoover thermostat is coupled
to other thermostats to take these additional conservation laws into account
has been proposed by Martyna et al. to get over this deficiency [245, 246]. In
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 101

Molecular Dynamics 101

this improved version of the Nosé–Hoover thermostat, a system of N nuclei


is coupled to a chain of thermostats (say, the chain number is M ) through

dri pi
= , (5.52a)
dt mi

dpi ∂V (r1 , . . . , rN ) pξ1 
= − − p, (5.52b)
dt ∂ri Q1 i
dξ1 pξ
1
= , (5.52c)
dt Q1
 
dpξ N
p2 pξ
i
1
= − g/β − 2 pξ1 , (5.52d)
dt i=1
mi Q2
dξj pξ
j
= .......................2  j  M − 1, (5.52e)
dt Qj
 2 
dpξ pξ  pξ
j
= j−1
− 1/β − j+1
p  , (5.52f)
dt Qj−1 Qj+1 ξj
dξM pξ
M
= , (5.52g)
dt QM
 2 
dpξ pξ 
M
= M −1
− 1/β , (5.52h)
dt QM−1

where g = 3N . The conserved total energy out of these equations of


motion is
N M p2 M
 p2
i ξj g 1 
HNHC = + V (r1 , . . . , rN ) + + ξ1 + ξj .
i=1
2m i j=1
2Q j β j=2
β
(5.53)

Now, it has been shown that the Nosé–Hoover chain is very efficient
in sampling ergodically the statistical property of many systems. It is also
of practical use that dynamic properties such as the vibrational frequency
can be extracted from their trajectories, although a rigorous justification of
such applications is absent. As mentioned at the end of Sec. 5.2.1, after the
BO MD method became a standard routine in practical ab initio MD simu-
lations since the late 1990s, other stochastic methods such as the Andersen
and Langevin thermostats were also widely used. In Sec. 5.2.1, we have
introduced the Andersen thermostat. In the following, we will give a brief
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 102

102 Computer Simulations of Molecules and Condensed Matter

introduction to the method of temperature control using Langevin dynam-


ics.

5.2.4 Langevin Thermostat

The Langevin dynamics is a simple and efficient scheme to control the


temperature of a canonical ensemble, which is often used in ab initio MD
simulations nowadays. The equations of motion for the particles are

ṙi = pi /mi , (5.54a)

ṗi = Fi − γpi + Ri . (5.54b)

Here, ri and pi are the position and velocity of the ith particle in the simu-
lation respectively. Fi is the force imposed by the other particles explicitly
treated in the simulation. The two parameters which are essential to control
the temperature in this scheme are γ and Ri . The first one (γ) represents
a viscous damping due to fictitious “heat bath” particles, while the second
one (Ri ) represents the effect of collisions with these particles. A balance
between them, which decelerates or accelerates the particles to be simu-
lated, respectively, keeps the average kinetic energy and consequently, the
temperature of the system constant.
In MD simulations where all atoms are explicitly treated, these “heat
bath” particles are purely fictitious, originating from the thermal fluctu-
ations. In computer simulations of large molecules in a solvent composed
of smaller ones (though these solvent molecules will not be explicitly sim-
ulated), this “heat bath” represents thermal effects originating from real
collisions between the smaller molecules in the solvent which are implicitly
treated and the larger ones which are explicitly treated under simulation.
Therefore, in the limiting case for very large friction coefficients, the Brow-
nian movement of the large molecules in solvent can also be described by
this equation, which simplifies into

ṙi = γ −1 m−1
i [Fi + Ri ] (5.55)

because ṗi = 0. In this book, we focus on the MD simulations where all


nuclei are explicitly treated. Therefore, the Brownian movement of the large
molecules in solvent is unrelated to the main problem we want to solve. But
it is worthwhile to note that the principles underlying this dynamics also
apply to these problems.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 103

Molecular Dynamics 103

There are four properties for the stochastic forces Ri to comply with in
order to control the temperature in simulations using Langevin dynamics:
(i) they are uncorrelated with ri and pi , (ii) the autocorrelation function of
Ri (t) has a form of δ-function, (iii) their mean values are zero, and (iv) their
distribution
 is Gaussian-like and the unit variance of this Gaussian function
is 2kB T γm/Δt. For a clarification of these four properties, we note that
the first one is easy to understand since it just implies that the random
force imposed by the collisions of the nuclei and the fictitious “heat bath”
particles is uncorrelated with the position and momentum of the nuclei. The
second point needs some explanation. It originates from the fact that the
timescale of the collisions between the particles under simulations and the
fictitious “heat bath” ones is much smaller than that of the ionic movement.
Therefore, the autocorrelation function of the random force Ri reads

Ri (t)Ri (t ) = 2kB T γmδ(t − t ). (5.56)

These t and t only take values with integer time step along the trajec-
tory. Besides this, it is also assumed that during an MD movement, there
are many collisions between the particles under simulation and the “heat
bath”. Therefore, one can impose a Gaussian distribution on Ri based on
the central limit theorem [256]. Then the Stokes–Einstein relation for the
diffusion coefficient can be used to show that the unit variance associated
with this Gaussian distribution is 2kB T γm/Δt and its mean value is zero.
With this, the principles underlying the choice of Ri for the simulation
of a canonical ensemble using the Langevin dynamics is clear. We note
that this choice of Ri limits this thermostat to the equilibrium simula-
tions only. For its extension to the non-equilibrium state simulations, one
may refer to Ref. [257]. The last point which needs to be explained for
the simulation of a canonical ensemble using Langevin dynamics is that
the equations of motion in Eq. (5.54) are equivalent to a Fokker–Planck
equation in description of the phase space probability density distribution
function ρ(r1 , . . . , rN , p1 , . . . , pN ), of the form
N   N
∂ρ  pi   
+ · ∇ri ρ + Fi · ∇pi ρ = γ ∇pi · pi ρ + mi kB T ∇pi ρ ,
∂t i=1 mi i=1
(5.57)
which has the canonical phase space probability distribution function as a
stationary solution. Therefore, this method is justified for the simulation
of the canonical ensemble. From these principles, it is clear that the choice
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 104

104 Computer Simulations of Molecules and Condensed Matter

of γ is the key issue in this simulation, which actually determines a com-


promise between the statistical sampling efficiency of the preservation of
the accuracy in short-term dynamics. As a stochastic method, one should
not expect the dynamics to be comparable to those from the simulations
using deterministic trajectories, like those of Nosé–Hoover thermostat sim-
ulations. But the stochastic feature normally guarantees the ergodicity to
be sampled in a more efficient way.
In the end, it is worthwhile to note that although traditional stochastic
thermostats often cause breakdown of the adiabatic separation in the CP
MD simulations, a recent development of the Langevin dynamics actually
showed that using correlated noise, it is possible to tune the coupling of a
stochastic thermostat with the various degrees of freedom for the nuclear
motion [254]. This development not only enables the use of the Langevin
dynamics in the CP MD simulations, it also improves the canonical sam-
pling efficiency because this thermostat can be tailored in a predictable and
controlled fashion. For more details about this recent development, please
see Ref. [254].

5.2.5 Andersen and Parrinello–Rahman Barostats

In simulations of an N V T ensemble, a non-flexible unit cell is used.


Therefore, although the thermal fluctuation of the nuclei can be properly
described using thermostats, the fluctuation of the simulation cell is com-
pletely absent. This fluctuation of the unit cell is a consequence of the
mechanical contact of the system with its environment, which is inevitable
in most experiments. Therefore, similar to the control of temperatures using
thermostats in the N V T simulations, it is also highly desirable to introduce
a barostat to account for this fluctuation of the unit cell, so that the envi-
ronment under simulation is as close to real experiments as possible.
Based on this consideration, there have been many methods proposed
to control this quantity in the MD simulations from the earliest Andersen
scheme [239, 258] and Parrinello–Rahman scheme [247, 248], to the more
recent work using Langevin dynamics [256, 259]. The corresponding simula-
tions are called N P T , or isobaric–isothermal simulations in the literature.
In this section, we will use the Andersen and Parrinello–Rahman schemes
as examples to illustrate how this is done. For principles underlying the
scheme using Langevin dynamics, please see Refs. [256, 259].
In the Andersen scheme [239], a constant-pressure simulation method
for a fluid system is proposed by introducing the instantaneous volume of
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 105

Molecular Dynamics 105

the simulation cell as an environmental variable in the effective Lagrangian.


Due to the isotropic feature of the system (fluid) to be simulated, a cubic cell
is used with volume Ω. The side length of this cubic cell equals Ω1/3 . Then,
the Cartesian coordinates of the nuclei (ri ) can be scaled with respect to the
instantaneous size of the simulation cell by xi = ri /(Ω1/3 ). In terms of these
scaled variables and the instantaneous Ω, the Lagrangian can be written as
N
 
mi M 2
LAndersen = Ω2/3 ẋ2i − V (Ω1/3 xi,j ) + Ω̇ − P0 Ω. (5.58)
i=1
2 i>j
2

Here, M is a parameter which can be viewed as an effective inertia for the


expansion and contraction of the volume. P0 is the external potential. xi,j
is the scaled distance (with no unit) between the ith atom and the jth atom
in the contracted or expanded cell.
There is a physical interpretation of the P0 Ω term in Eq. (5.58), in
the original words of Andersen: “Suppose the fluid to be simulated is in a
container of variable volume. The fluid can be compressed by a piston. Thus,
Ω, whose value is the volume, is the coordinate of the piston, P0 Ω is a P V
potential derived from an external pressure P0 acting on the piston, and M
is the mass of the piston. The piston is not of the usual cylindrical type that
expands or contracts the system along only one direction; instead, a change
in Ω causes an isotropic expansion or contraction.” But we note that this
interpretation is not entirely consistent with the Lagrangian in Eq. (5.58),
since from xi = ri /(Ω1/3 ), one easily derives
1
ṙi = Ω1/3 ẋi + Ω−2/3 Ω̇xi . (5.59)
3
This equation means that the kinetic energy of the atoms should also
contain contributions from Ω̇, which are absent in Eq. (5.58). In spite
of this, we acknowledge that Eq. (5.58) still presents a well-defined
Lagrangian.
Now, we analyze the dynamics generated from this Lagrangian. The
momentum conjugate to xi is denoted as πi , which equals
∂LAndersen
πi = = mi Ω2/3 ẋi . (5.60)
∂ ẋi
The momentum conjugate to Ω is denoted as Π, which equals
∂LAndersen
Π= = M Ω̇. (5.61)
∂ Ω̇
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 106

106 Computer Simulations of Molecules and Condensed Matter

The Hamiltonian corresponding to the Lagrangian in Eq. (5.58) is then


written as
N

HAndersen(x1 , . . . , xN , π1 , . . . , πN , Ω, Π) = ẋi · πi + ΩΠ − LAndersen
i=1
N 
1
= (2Ω2/3 )−1 πi · πi + V (Ω1/3 xi,j ) + (2M )−1 Π2 + P0 Ω.
i=1
m i i>j
(5.62)

With this Hamiltonian, one can derive the equations of motion as


dxi HAndersen πi
= = ,
dt ∂πi mi Ω2/3
dπi HAndersen  xi,j ∂V (Ω1/3 xi,j )
=− = −Ω1/3 ,
dt ∂xi i>j
|xi,j | ∂xi

dΩ HAndersen Π
= = ,
dt ∂Π M
dΠ HAndersen
=
dt ∂Ω
 N
 1
= −(3Ω)−1 −2(2Ω2/3 )−1 πi · πi +Ω1/3
i=1
m i

 ∂V (Ω1/3 xi,j )
× xi,j +3P0 Ω⎠ . (5.63)
i<j
∂x i

With these equations of motion, one can use the MD method as


introduced in Sec. 5.1 for the micro-canonical system to generate a
trajectory for x1 , . . . , xN , π1 , . . . , πN , Ω, and Π. We note that the volume is
not a conserved quantity in real space. However, the scaled extended system
with variables x1 , . . . , xN , π1 , . . . , πN , Ω, and Π can still be regarded as a
micro-canonical ensemble in which the volume is unity since the coordi-
nates xi are all scaled to lie within a dimensionless unit cell. Under the
ergodic hypothesis, the time average over this trajectory is equivalent to
the ensemble average of the micro-canonical extended system. Therefore,
for any function of x1 , . . . , xN , π1 , . . . , πN , Ω, and Π, the time average over
this trajectory will be regarded as equivalent to the ensemble average of this
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 107

Molecular Dynamics 107

quantity on the extended system, and the N P T ensemble average of the


real subsystem can be naturally evaluated using this trajectory too. We note
that this method is one of the first methods which employed an extended
Lagrangian in an MD simulation. Later development of many ensemble
simulation methods, e.g. the Nosé–Hoover thermostat, has benefited a lot
from this idea.
One shortcoming of the Andersen barostat is that an isotropic change
of the simulation cell must be assumed. This assumption is reasonable in
simulations of fluids. However, when it comes to solids, especially in the
case when the unit cell is not cubic, in addition to the volume, one must
also allow the shape of the simulation cell to change. This generalization
is essential in studies of structural phase transitions. Based on this con-
sideration, Parrinello and Rahman soon designed a scheme which allows
the change of the simulation cell’s shape in MD simulations [247, 248].
In this method, instead of using only one variable Ω, nine variables asso-
ciated with three vectors in the Cartesian coordinate (a, b, and c) are
used, which allow both the shape and the volume of the simulation cell
to be changed. These nine variables are Ai,j , where i (and j) goes from
1 to 3. A is the matrix representation of (a, b, c). The volume (Ω) of
the simulation cell equals a · (b × c). The Lagrangian for this extended
system is
N 3
1  M 
LPR = mi (Aẋi )T Aẋi − V (A−1 xi,j A) + Ȧi,j − P0 Ω.
2 i=1 i>j
2 i,j=1
(5.64)
Using the same trick that we have applied to the Andersen barostat (and
Nosé–Hoover thermostat), one can derive the equations of motion for xi
and Ai,j . We do not go through this deduction here; for such details, please
see Refs. [247, 248].

5.3 Examples for Practical Simulations


in Real Poly-Atomic Systems
After understanding the principles of MD as introduced above, one can
easily carry out a practical MD simulation using a computer code which
is accessible in either a commercial or an open-source manner. Taking one
of the most frequently used settings, an ab initio BO MD simulation using
a Nosé–Hoover chain thermostat for the N V T ensemble as an example, in
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 108

108 Computer Simulations of Molecules and Condensed Matter

Figure 5.1 Illustration of an ab initio MD simulation. We show one snapshot of an


ab initio MD simulation on a system composed of 32 water molecules and one extra
proton. The extra proton is denoted by the sphere in green. The oxygen nuclei and other
hydrogen nuclei are shown by spheres in red and white, respectively. The silver contour
indicates the electron density distribution, which is computed in an ab initio manner
“on-the-fly” as the system evolves.

Fig. 5.1, we show one snapshot of a system composed of 32 water molecules


and one extra proton. This simulation is based on the implementation of
the MD method in the VASP code [260–264]. As mentioned in the intro-
duction, the nuclei are treated as classical point-like particles. Therefore,
the ball-and-stick model is still used for the representation of the nuclei in
this snapshot. The electronic structures, however, are computed “on-the-
fly” in an ab initio manner as the system evolves. This is reflected by the
silver contour representing the density distribution of the electrons at this
snapshot, which describes the bond making and breaking processes during
the MD simulation in a seamless manner.
From such snapshots and statistical (as well as dynamical) analysis, one
can get quite some information concerning the evolution of the electronic
structures and associated nuclear motion at the molecular level in such
systems. For example, one can analyze the electron charge transfer between
the extra proton and the bulk water solution explicitly using the electronic
structure of such a snapshot. Specifically, this can be done by calculating the
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 109

Molecular Dynamics 109

Figure 5.2 Illustration of how the electron charge transfer is analyzed in an ab initio MD
simulation. The same simulation cell and notation of the nuclei are used in Fig. 5.1. The
contour in brown shows electron charge accumulation and the contour in blue indicate
electron charge depletion. This analysis shows that an extraordinarily strong hydro-
gen bond is associated with the extra proton. Besides this, significant electron charge
redistribution also exists on the neighboring hydrogen bonds, enhancing most of them.
Therefore, for a reasonable characterization of the transfer process of this extra proton,
besides the dynamics of this extra proton, the hydrogen bond dynamics of the “shell” is
also crucial.

electron density distribution of the system composed of the water molecules


and the extra proton first, and then that for each of them (solely the water
molecules and solely the proton) separately. Taking the difference between
the first one and the sum over the latter two, one obtains how the electron
charge density is redistributed by the interactions between the extra proton
and the water solution in a very clear manner. In Fig. 5.2, we show how this
electron charge transfer behaves in the neighborhood of the excess proton.
Besides the electron accumulation and depletion on the water molecules and
hydrogen bond associated with the extra proton, obvious electron charge
redistribution exists in the neighboring region, indicating that the hydrogen
bond strengths in the second shell are also significantly influenced by the
presence of this extra proton.
The above analysis also indicates that the behaviors of the hydro-
gen bonds in the neighboring shells and the excess proton are correlated.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 110

110 Computer Simulations of Molecules and Condensed Matter

Because of this, for a better understanding of the proton transport mecha-


nism in liquid water, one can investigate in more detail the hydrogen bond
dynamics in the presence of this excess proton. As a matter of fact, this
issue has been intensively studied from both the experimental and theoret-
ical perspectives in the last two decades [264–274]. Now, it has been widely
accepted that the behavior of this extra proton is basically determined by
the dynamics of the hydrogen bonds within the neighboring shells [264–271].
This story can be understood pictorially by taking some sequential snap-
shots of the MD simulation, as shown in Fig. 5.3. In these snapshots, the

Figure 5.3 Illustration of the hydrogen bond dynamics related to the proton transport
in liquid water obtained from an ab initio MD simulation using classical nuclei. Red
spheres are oxygen nuclei and white spheres are hydrogen nuclei. The excess proton is in
green and the pivot oxygen (Op ) is in blue. Hydrogen bonds (except for the most active
one) are indicated by green dashed lines. Os is the oxygen which shares the most active
proton with the pivot oxygen. The first three panels show that the most active proton
switches identity between the three hydrogen bonds related to Op . Going from the third
to the fourth, the coordination number of Os reduces from 4 to 3. Then proton transport
happens going from the fourth to the fifth panel. After the proton transport, the most
active proton begins to exchange sites about a new pivot oxygen (the last two panels).
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 111

Molecular Dynamics 111

pivot oxygen (Op ) is defined as the oxygen nucleus which owns three hydro-
gen atoms simultaneously, and Os is defined as the oxygen nucleus which
shares the most active proton with the pivot one. The pair (hydrogen bond)
between Op and Os is called “the special pair”. In the MD simulation, the
pivot oxygen normally stays on one site, i.e. the pivot oxygen’s identity does
not change, peacefully for a certain time interval (∼ 1 ps). During this time,
the special pair switches identity between the hydrogen bonds in which the
pivot oxygen is involved frequently, which is often called the special pair
dancing in Refs. [267–271]. The coordination number of Os is normally 4
during this dancing process (the first three panels in Fig. 5.3). Upon going
from the third panel to the fourth one, the coordination number of Os
reduces from 4 to 3. Then proton transport happens (the fifth panel). After
the proton transport, the most active proton begins to exchange sites about
a new pivot oxygen (the last two panels). In short, the proton transport
happens through the well-known Grotthuss mechanism [275, 276], with the
hydrogen bond dynamics determining the interconversion between different
hydration states of the excess proton.
The second example we want to show briefly here for practical simu-
lations using the ab initio MD method is the phase diagram of condensed
matter under extreme conditions, e.g. high pressures. It is well known that
experimental realization of pressures above 100 GPa is extremely difficult.
For theoreticians, however, this is readily doable since an ab initio MD
simulation using an N P T ensemble easily reproduces the atomic level evo-
lution of the system under investigation if (i) the electronic structures are
accurate, (ii) the simulation cell is large enough, and (iii) the simulation
time is long enough. As a matter of fact, this simulation technique has
already been widely used in studies of the high-pressure phase diagrams for
a series of systems [277–280, 291]. One of the most prominent examples is
the high-pressure phase diagram of hydrogen, as shown in Fig. 5.4. Below
300 GPa, it is well known that three regions exist, consisting of the molec-
ular solid, molecular liquid, and the atomic liquid. The dissociation line
between the molecular liquid and the atomic liquid as shown in the figure
was just determined by the ab initio MD simulations in Refs. [277, 278].
We note that the accuracy of the electronic structures obtained from
the present ab initio method, which is affordable for the MD simulations, is
still a problem which determines the accuracy of the theoretical predictions
of this phase diagram [282–285]. As a matter of fact, this is a problem
which is vital for the accuracy of the MD simulations in general. It is also
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch05 page 112

112 Computer Simulations of Molecules and Condensed Matter

Figure 5.4 Illustration of hydrogen phase diagram under pressure. The dissociation line
between the molecular liquid phase and the atomic liquid phase was obtained using the
ab initio MD simulations in Refs. [277, 278].

why when chemical bond breaking and forming happen, non-reactive force
field-based method is to be abandoned in descriptions of the interatomic
interactions and ab initio MD method prevails in recent years. Therefore,
to conclude the chapter with a reminder, when the MD simulation tech-
nique is used to interpret the experiments or make predictions, one must
pay special attention to the accuracy of the interatomic interactions for all
spatial configurations covered during the simulation.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch06 page 113

6
Extension of Molecular Dynamics,
Enhanced Sampling and the
Free-Energy Calculations

The atomistic level MD method as introduced in the previous chapter allows


us to simulate the propagation of a system using a finite-time step so that
the total energy of this system is conserved and the temperature is under
control. This means that the propagation needs to be done with a time
interval on the order of femtoseconds, or even shorter. Processes of real
chemical interest, however, happen at much longer time scales. Taking the
ab initio MD simulation of some molecules being adsorbed on a metal
surface as an example, currently, it requires a powerful computer cluster
with a few hundred processors roughly one month to simulate the MD
propagation of 100 ps [286]. Imagining that we want to see a rare-event
chemical reaction happening at a time scale of one millisecond, we need
to continue this simulation on this cluster for about 106 years in order to
see this event happening once. Even if the ab initio method is renounced
in descriptions of the interatomic interactions, the empirical potentials are
used so that the MD simulation can propagate for, say 100 nanoseconds a
day, reproducing the propagation of 1 millisecond still requires a computer
simulation of about 30 years. Plus, this event needs to happen several times
in order for the statistics to work. Therefore, from a statistical point of
view, the brute force MD technique is completely unacceptable in practical
simulations involving such rare events. A better scheme, in which rare events
like this can be simulated, is highly desired.
During the past years, great effort has been made in order for this
purpose to be fulfilled. One route is to give up the all-atom description
and use a coarse-grained model [287, 288]. In doing so, the time step used

113
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch06 page 114

114 Computer Simulations of Molecules and Condensed Matter

for the MD simulation can be significantly increased so that the time scale
accessible to computer simulations increases, sometimes to a value large
enough to complement biological or chemical experiments. However, it is
worth noting that an a priori detailed knowledge of the system to be sim-
ulated is required, which unfortunately is often not available.
A different route, which keeps the atomistic feature for the description
of the propagation, is to develop a method so that the frequency with which
a rare event happens can be significantly increased. Or, in other words,
an enhanced sampling efficiency is guaranteed. Over the last few decades
(almost four), great success has been achieved in this direction. A num-
ber of methods, e.g. the umbrella sampling [94, 95], the adaptive umbrella
sampling [33], the replica exchange [34, 35], the metadynamics [38, 39],
the multi-canonical simulations [40, 41], the conformational flooding [42],
the conformational space annealing [43], the integrated tempering sampling
[44–47] methods, etc. had been proposed, each having its own strength and
weakness. For example, in the umbrella sampling, adaptive umbrella sam-
pling, metadynamics, and configurational flooding methods, predetermined
reaction coordinates are required before the simulations were carried out.
While in replica exchange and multi-canonical simulations, different tem-
perature simulations are needed, which make the trajectories obtained lose
their dynamical information. In this chapter, we will take the umbrella sam-
pling, the adaptive umbrella sampling, and the metadynamics methods as
examples of the former class to show how they work in practice. In addition
to this, the integrated tempering sampling method from the second class
will also be discussed.
The enhanced sampling methods as introduced above focus on explor-
ing the free energy profile, or otherwise often called potential of mean force
(PMF), throughout the entire conformational phase space of the poly-
atomic system under investigation. In some applications of the molecu-
lar simulations, e.g. phase transitions, this entire PMF exploration can be
avoided if the two competing phases are well-defined metastable states in
standard MD simulations. In such cases, the free energy of the system at
each phase can be calculated using methods like thermodynamic integration
at a certain temperature and pressure. Then, by monitoring the competi-
tion of these two free energies at different temperatures and pressures, one
can find out which phase is more stable at a certain condition and conse-
quently determine the phase diagram of this substance under investigation.
Since phase transition is also an important topic in molecular simulations,
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch06 page 115

Extension of Molecular Dynamics, Enhanced Sampling 115

principles underlying the thermodynamic integration method will also be


explained in this chapter. We note that although this is the motivation for
our explanation of the thermodynamic integration method, the range of its
applications is much wider. As a matter of fact, the principles underlying
the idea of thermodynamic integration can be applied to the mapping of
the PMF in a much more general sense, as long as a certain parameter can
be defined to link states of interest whose free energies are to be calculated.
This chapter is organized as follows. In Sec. 6.1, we introduce prin-
ciples underlying PMF exploration and the umbrella sampling, as well as
the adaptive umbrella sampling methods. Then, we will explain, in brief,
the metadynamics method in Sec. 6.2. The integrated tempering sampling
method, as an example of enhanced sampling in which an a priori defini-
tion of transition coordinate is not required, is explained in Sec. 6.3. The
thermodynamic integration method for the calculation of the free energy of
a certain phase is introduced in Sec. 6.4. We hope this introduction can help
graduate students who have started working on molecular simulations to set
up some concepts before carrying out their enhanced sampling simulations.

6.1 Umbrella Sampling and Adaptive Umbrella


Sampling Methods
As mentioned in the introduction, a key concept in our understanding of the
enhanced sampling method is the PMF, which was first given by Kirkwood
in 1935 [289]. In this definition, a reaction coordinate ξ is used, which is a
function of the poly-atomic system’s nuclear configuration in the Cartesian
space. Here, we denote this spatial configuration of the nuclei as x. It is a
3N -dimensional vector composed of xi , with i going through 1 and N . xi
represents the Cartesian coordinate of the ith nucleus and ξ can be written
as ξ(x). This nomenclature for the nuclear configuration will also be used
for our discussions of the path integral method in Chapter 7.
From the principles of statistical mechanics, one can first write down
the free energy profile of the poly-atomic system as a function of ξ, with
the variable denoted by ξ0 , as

1
F (ξ0 ) = − ln P (ξ0 )
β
 −βV (x)δ(ξ(x)−ξ )
0
1 e dx
= − ln . (6.1)
β Q
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch06 page 116

116 Computer Simulations of Molecules and Condensed Matter

Here, β equals 1/(kB T ) and Q is the canonical partition function of the


system:

Q = e−βV (x) dx. (6.2)

ξ0 is the variable used to construct the PMF. In the following discussion,


we use a 1D case as an example, for simplicity. The extension of the equa-
tions to higher-dimensional situations is straightforward. This PMF is a key
function in descriptions of the configurational (or conformational) equilib-
rium properties as well as the transition rate of the dynamically activated
processes in computer simulations for the behaviors of poly-atomic systems
under finite temperatures.
From this definition of the PMF, one knows that it is closely related
to the BO PES since the latter, together with the thermal and quantum
fluctuations of the nuclei, determine the probability distribution function
P (ξ0 ) at a finite temperature. In other words, one can think of the PMF
as a revised version of the PES, with the thermal and quantum effects
included when addressing the electronic and nuclear degrees of freedom. In
the introduction of this chapter, one can see that in cases when there exist
multiple deep local minima on the PES, the probability distribution func-
tion P (ξ0 ) is far from being ergodic and one needs the enhanced sampling
method to map out the PMF over the entire conformational space relevant
to the problem of interest. Among the various efforts carried out in the
last, close-to-four decades, the umbrella sampling method first proposed by
Torrie and Valleau is now believed to be one of the most influential methods
[94, 95]. The mathematical basis for a treatment like this is that in order
for an ergodic sampling of the conformational space relevant to the problem
of interest, an additional biased potential Vb as a function ξ can be added
to the real interatomic potential V (x) in the molecular simulations. With
this biased additional potential, a biased PMF can be constructed using
1
Fb (ξ0 ) = − ln Pb (ξ0 )
β
 −β[V (x)+V (ξ(x))]δ(ξ(x)−ξ )
b 0
1 e dx
= − ln , (6.3)
β Qb
where

Qb = e−β[V (x)+Vb (ξ(x))] dx. (6.4)
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch06 page 117

Extension of Molecular Dynamics, Enhanced Sampling 117

Figure 6.1 Taking the simplest case, i.e. the system is 1D and the ξ is the same as the x,
as an example, the original PES V (x) has a very deep valley so that the MD simulation
cannot go through the whole space in practical MD simulations. However, if an additional
biased potential Vb (x) which roughly equals −V (x) is added to the original potential,
the new potential Vb (x) + V (x) will be very flat so that in an MD simulation using this
potential, the particle can go through the entire x axis freely and Fb (x) can be sampled
very well. Based on this Fb (x), the original free energy profile can be reconstructed using
Eq. (6.7). This principle also applies to the system with higher dimensions for the nuclear
degree of freedom and more complex form of the reaction coordinate ξ.

Due to the fact that the biased potential can be chosen in a way that
V (x)+Vb (ξ(x)) is rather flat as a function of ξ (see Fig. 6.1), the probability
distribution Pb (ξ0 ) can be much better-sampled over the whole conforma-
tional space relevant to the problem of interest compared with P (ξ0 ) in the
MD simulations so that the biased PMF can be calculated efficiently using
MD statistics.
Then, one relates the biased PMF Fb (ξ0 ) with the unbiased one F (ξ0 )
using the following equation:

1 e−β[V (x)+Vb (ξ(x))]δ(ξ(x)−ξ0 ) dx
Fb (ξ0 ) = − ln
β Qb
 −βV (x)
1 e δ(ξ(x) − ξ0 )dx
= − ln e−βVb (ξ0 )
β Qb
 
1 e−βV (x) δ(ξ(x) − ξ0 )dx Q
= Vb (ξ0 ) − ln
β Q Qb
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch06 page 118

118 Computer Simulations of Molecules and Condensed Matter


1 e−βV (x) δ(ξ(x) − ξ0 )dx 1 Q
= Vb (ξ0 ) − ln − ln
β Q β Qb
1 Q
= Vb (ξ0 ) + F (ξ0 ) − ln . (6.5)
β Qb
Therefore, from the biased PMF Fb (ξ0 ), which can be sampled well in
practical MD simulations over the whole conformational space, one can
reconstruct the unbiased PMF through
1 Qb
F (ξ0 ) = Fb (ξ0 ) − Vb (ξ0 ) + ln . (6.6)
β Q
We note that this (1/β) ln(Qb /Q) is a constant, which we denote as F . It
is independent of ξ0 , but determined by the choice of Vb (ξ), the conforma-
tional space of the poly-atomic system relevant to the problem of interest,
the original interatomic potential, and the temperature. Rewriting Eq. (6.6)
also gives us
F (ξ0 ) = Fb (ξ0 ) − Vb (ξ0 ) + F. (6.7)
As simple as it looks, the addition of a biased potential easily solves
the problem for the ergodic exploration of the conformational space for
the mapping of the PMF in principle, in the MD simulations. However,
we note that in practical applications, the high-dimensional PES function
V (x) is always unknown. Therefore, it is impossible to define an additional
potential in its negative form in an a priori manner so that the system
can go through the entire conformational space relevant to the problem of
interest efficiently, using this biased potential. When this is the case and the
chosen additional potential does not compensate for the original one, the
system will still be trapped in one local minimum in the MD simulations
under the biased potential (see e.g. Fig. 6.2).
One way to circumvent this problem and ensure that the whole confor-
mational space relevant to the problem of interest is sufficiently (or even
uniformly) sampled is to separate this conformational space into some boxes
using the chosen reaction coordinates. When there is only one reaction coor-
dinate, this is simplified into taking some values of ξ in the region relevant
to the problem of interest, say ξ i . Then, a series of MD simulations can be
carried out using these biased potentials which constrain the system in the
neighborhood of these ξ i s so that the conformational spaces of the system
with reaction coordinates in the neighborhood of all these ξ i s are all well
sampled. Taking the 1D PMF associated with two stable states (reactant
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch06 page 119

Extension of Molecular Dynamics, Enhanced Sampling 119

Figure 6.2 The situation is similar to Fig. 6.1. However, one does not know the form of
the original potential and therefore the additional one does not compensate for it well.
In this case, the sampling of the system with the biased potential will not be ergodic on
the conformational space either.

and product) separated by a high energy barrier as an example (Fig. 6.3),


in the language of Torrie and Valleau [94, 95], a series of bias potentials can
be used along the reaction path between the reactant and the product. At
each point along this path, a bias potential is used to confine the system to
its neighborhood and the corresponding simulation is often called a “biased
window simulation”. From this simulation, one can reconstruct the PMF
in this neighborhood from the MD simulation using the biased potential.
Then, by linking the PMF in each region, one obtains the PMF along
the reaction coordinate all over the conformational space concerned. We
note that there can be many choices for this additional constraint poten-
tial. One of the most often used is the harmonic function with the form
Vbi (ξ0 ) = K(ξ0 − ξ i )2 /2, as shown by the red curves in Fig. 6.3. Due to
the fact that this additional bias potential looks like an umbrella (Fig. 6.3),
this method is called the umbrella sampling in the literature. With this
treatment, a uniform series of sample points in the region relevant to the
problem of interest will ensure that the conformational space is sufficiently
sampled and consequently the PMF well reconstructed.
The main technical problem related to the reconstruction of the PMF
from umbrella sampling as mentioned above originates from the fact that in
Eq. (6.7), there is a constant which depends on the choice of the additional
potential and the region it explores. This indicates that this constant is
different in each of the biased MD simulations. Consequently, one needs to
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch06 page 120

120 Computer Simulations of Molecules and Condensed Matter

Figure 6.3 Illustration of how umbrella sampling is used in simulations of the transition
between a reactant state and a product, one separated by high energy barrier. Along the
reaction coordinate, a series of points were taken. At each point ξi , a bias potential with
the form Vb (ξ) = K(ξ − ξi )2 /2 is used to constrain the system to its neighborhood.
From this biased simulation, the PMF at this neighborhood can be obtained from the
corresponding probability distribution (indicated by the blue curve). These series of
additional biasing potentials ensure that the whole conformation space relevant to the
problem of interest is sufficiently sampled.

align the PMF reconstructed in each region of the simulation window so


that the relative values of the PMF obtained from different MD simulations
make sense. Traditionally, this is done by adjusting F (ξ0 ) of the adjacent
boxes (windows) in which they overlap so that they match [94–96], using,
e.g. least-square method [290]. In doing so, the PMF over the whole region
of interest can be constructed. However, we note that there are serious lim-
itations for this scheme. First, in matching the overlap region between the
PMF of the neighboring windows, a significant overlap is required to ensure
statistical accuracy. This indicates that a lot of sampling in the MD sim-
ulations are superfluous and consequently wasted, and more importantly,
when more than one reaction coordinate is used for the construction of the
PMF, the value of the constant F in one simulation window allowing the
best fit with its adjacent region in one direction may not ensure the best for
others [291]. Therefore, this scheme is of limited use in practical simulations
of complex systems when analysis of high-dimensional PMF is required.
Among the various attempts to solve these problems [290, 292–294], the
weighted histogram analysis method (WHAM) proposed by Kumar et al. is
now the most popular, mainly due to its numerical stability and convenience
in addressing PMF with multiple variables [295]. The central equation for
the WHAM method includes the optimization of the unbiased distribution
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch06 page 121

Extension of Molecular Dynamics, Enhanced Sampling 121

function as a weighted sum over its expression in different windows as

Nw
 
 i
ni e−β(Vb (ξ0 )−Fi )
i
P (ξ0 ) = P (ξ0 ) N j , (6.8)
i=1
w
j=1 nj e−β(Vb (ξ0 )−Fj )

where Nw is the number of windows and P i (ξ0 ) represents the unbiased


probability distribution reconstructed from the biased MD simulation in
the ith window using Vbi (ξ0 ). Similar to discussions above, ξ0 denotes the
variable and Vbi (ξ0 ) = K(ξ0 − ξ i )2 /2. P (ξ0 ) is the optimized overall distri-
bution function obtained from the Nw biased simulations. This definition of
P (ξ0 ) (instead of a direct sum over P i (ξ0 )) is very reasonable since not only
the relative weights between the number of independent data points within
each window (i.e. ni ) are taken into account (more data, better statistics
and consequently larger weights in the summation), but also, the small
weights of P i (ξ0 ) in the region with large Vbi (ξ0 ) − Fi means that the error
originating from worse statistics of P i (ξ0 ) in this region becomes numeri-
cally well under control. This expression, together with the right values of
the Fi s (i goes from 1 to Nw ), gives the unbiased distribution function of
P (ξ0 ) throughout the regions on the ξ axes relevant to the conformational
space of interest. What one needs to do next is to optimize the values of
Fi s in Eq. (6.8).
To understand how this is done in practice, we go back to Eq. (6.7) for
the biased MD simulation in each window. This equation changes to

F i (ξ0 ) = Fbi (ξ0 ) − Vbi (ξ0 ) + F i , (6.9)

where F i (ξ0 ) = − ln P i (ξ0 )/β and Fbi (ξ0 ) = − ln Pbi (ξ0 )/β. Pbi (ξ0 ) repre-
sents the biased probability distribution sampled with the biased potential
around ξ i , which is directly obtained from the ith MD simulations. Sta-
tistically, it samples the region around ξ i rather well. By inputting the
expressions of F i (ξ0 ) and Fbi (ξ0 ) into Eq. (6.9), one gets
i i
P i (ξ0 ) = Pbi (ξ0 )eVb (ξ0 )−F . (6.10)

Therefore, Eq. (6.8) evolves into

Nw
 ni Pbi (ξ0 )
P (ξ0 ) = Nw j . (6.11)
i=1 j=1 nj e−β(Vb (ξ0 )−Fj )
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch06 page 122

122 Computer Simulations of Molecules and Condensed Matter

The free energies FI s in Eq. (6.11), on the other hand, satisfy the following
equation:

1 Qib
Fi = − ln
β Q
 −β[V (x)+V i (ξ(x))]
1 e b dx
= − ln
β Q

1 i
= − ln e−βVb (ξ0 ) P (ξ0 )dξ0 . (6.12)
β

Therefore, Eqs. (6.11) and (6.12) compose a set of equations which define
the relationship between the Fi s and P (ξ0 ). By solving these two equations
self-consistently, one can obtain a much better and more efficient estimate of
P (ξ0 ) compared with the traditional PMF matching method from the sta-
tistical point of view. Once again, we note that the essence of this method,
in the language of Roux [96], is “constructing an optimal estimate of the
unbiased distribution function as a weighted sum over the data extracted
from all the simulations and determining the functional form of the weight
factors that minimizes the statistical error”.
So far we have discussed the umbrella sampling method. From this
introduction, we know that a key point on reconstructing the PMF using
the biased potential is that a sufficient sampling, in the best case a uniform
sampling, of the conformational space relevant to the problem of interest is
guaranteed. To this end, an a priori set of umbrella potentials are needed
in order for this uniform sampling to be carried out in a practical manner.
For simulations of relatively simple systems, in which the variable related to
the umbrella potential is easy to define, this method works well. However,
for complex systems where this a priori definition of the umbrella potential
is unlikely, which unfortunately is usually true in reactions with more than
one degree of freedom, this method often fails.
An alternative method within a similar scheme, which is more often
used in present explorations of the PMF in complex systems, is the so-called
adaptive umbrella sampling method [296–298]. A key difference between the
adaptive umbrella sampling method and the umbrella sampling method is
that in the former case, the umbrella potentials are chosen and updated
in the simulations, while in the latter one, the umbrella potentials do not
change. Plus, in the umbrella sampling method, the conformational space is
separated into a series of regions and the MD simulation with bias potential
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch06 page 123

Extension of Molecular Dynamics, Enhanced Sampling 123

centered at each region have different Fi . A key point in combining the


PMFs constructed from the biased MD simulations in umbrella sampling is
to align these different Fi s. Conversely, in the adaptive umbrella sampling
method, the umbrella potential does not focus a specific region, but rather
tries to compensate the original PES. Consequently, the problem related
to the treatment of the constant in Eq. (6.7) transforms into the shift of
this value during the iterations so that in the end, it equals the free energy
density of the system over the whole conformation space relevant to the
study.
The central idea behind the adaptive umbrella sampling method is
that if an umbrella potential with the form Vb (ξ0 ) results in a uniform
biased probability distribution Pb (ξ0 ), the corresponding umbrella potential
satisfies the following equation:
1
Vb (ξ0 ) = −V (ξ0 ) = ln P (ξ0 ), (6.13)
β
where P (ξ0 ) stands for the unbiased probability distribution we want to
simulate. Based on this equation, the biasing potential can be adapted
to the PMF which is determined using the information from the previous
simulations in an iterative manner [299]. The analysis of the PMF is often
carried out using WHAM [295]. To be more specific, an initial guess of the
umbrella potential, which is often taken as Vb1 (ξ0 ) = 0, can be used for the
MD simulation and a biased probability distribution Pb1 (ξ0 ) is obtained.
From this biased probability distribution, one can construct the unbiased
one (P 1 (ξ0 )). Due to the ergodic problem related to the original PES, this
P 1 (ξ0 ) will be very different from the final P (ξ0 ). However, one can input
this unbiased probability distribution into Eq. (6.13) and get a new umbrella
potential Vb2 (ξ0 ). Adding this biasing potential to the original potential,
one can perform the second biased MD simulation and generate a new
Pb2 (ξ0 ). From this biased probability distribution, a normalized unbiased
P 2 (ξ0 ) can be achieved. This unbiased potential will be summed up with
the unbiased one in the first round subjected to some normalized weighting
factors to generate the new input for Eq. (6.13). From Eq. (6.13), again, one
obtains a new umbrella potential. Continue this iteration and pay attention
to the fact that the input for Eq. (6.13) is always generated using weighted
sum over unbiased probability distributions obtained from earlier iterations,
until the conformational space relevant to the problem of interest has been
sampled adequately. One can then get a well-converged unbiased PMF. This
unbiased PMF can then be used for statistical or dynamical studies in the
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch06 page 124

124 Computer Simulations of Molecules and Condensed Matter

future. We note that this is just a general description of how the adaptive
umbrella sampling works. There are many numerical details concerning real
implementation. For these details, please see Ref. [299, 300].

6.2 Metadynamics
In the above discussions on the umbrella sampling and adaptive umbrella
sampling methods, Eq. (6.7) is the basis for the analysis to be carried out
and rigorous justification for the construction of the unbiased PMF exists
behind the biased MD simulations. Parallel to this scheme, however, there
is another algorithm which (as pointed by A. Laio, the main founder of
this method) did not follow from any ordinary thermodynamic identity but
was rather postulated on a heuristic basis and later verified empirically to
be very successful in the enhanced sampling simulations of many complex
systems [301–307]. This method is the so-called metadynamics [38, 39].
The principles underlying this metadynamics method can be under-
stood pictorially using Fig. 6.4. The first step is to choose a sensible collec-
tive variable (labelled as CV, corresponding to the reaction coordinate used
in the earlier discussion) which, in principle, should be able to distinguish
the initial, intermediate, and final states and describe the slow processes
relevant to the problem of interest. This is similar to the umbrella sampling
and adaptive umbrella sampling methods introduced before. The difference
appears afterwards. Imagine that only one CV is chosen and the BO PES
looks like the black curve in Fig. 6.4, starting from one deep valley, the sys-
tem will take a time which is unacceptably long for atomic simulations to
escape from it using the standard MD method. In the language of metady-
namics, what one can do in order to impose an efficient enhanced sampling
on the PES is to add in some additional potentials so that the deep valley
can be filled up quickly (Fig. 6.4(a)). This potential can be written in terms
of Gaussian functions as
 − (ξ(x)−ξ(t
 ))2

VG (ξ(x), t) = w e 2(δs)2 , (6.14)


t =τG ,2τG ,... and t <t

where ξ(x) stands for the CV coordinate associated with the spatial con-
figuration of the nuclei x, and ξ(t ) stands for its instantaneous value at
t . It is obvious that the Gaussian height w, Gaussian width δs, and the
time interval τG at which the Gaussians are added control the form of this
additional potential and consequently the accuracy and efficiency of the
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch06 page 125

Extension of Molecular Dynamics, Enhanced Sampling 125

Figure 6.4 Illustration of the principles underlying the metadynamics method, taking
the 1D case as an example. Imagine that the PES has multiple deep valleys. In meta-
dynamics, some artificial potentials will be added along the chosen collective variable
coordinates as the system evolves. At a certain valley, this is similar to the case that a
man falls into a deep well. To get out of it, simply by jumping is unlikely, but imagine
that he has an infinite amount of stones in his “magic” pocket, he can throw a stone
every few steps he walks. With time going on, these stones will fill up the well and he
can get out of it easily. By remembering the geometric shape of the space filled up by
the stone, intuitively the PMF can be reconstructed. It is easy to understand that the
size of the stone determines the accuracy and efficiency of the reconstructed PMF. As
mentioned earlier, this rationalization is postulated on a heuristic basis. However, we
note that it has been verified empirically to be very successful in the enhanced sampling
simulations of many complex systems [301–307].

enhanced sampling. Imagine that the Gaussian potentials are like small
stones (size of the stone controlled by w and δs) which fill up the valley
gradually, the difference between the curve the stick man is standing on
and the curve below then naturally represents the sum over the Gaussian
functions as shown in Eq. (6.14), and the size of the stones determines
the efficiency and accuracy for the construction of this difference between
the two curves. After one valley was filled up, the walker which describes
the evolution of the system in the enhanced sampling process starts to
explore the neighboring one (Fig. 6.4(b)), until all the valleys are filled.
After this, the system is allowed to travel between the valleys in a free
manner (Fig. 6.4(c)). When the exploration of the walker is controlled
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch06 page 126

126 Computer Simulations of Molecules and Condensed Matter

using an ensemble-based method, such as the MD or the Monte-Carlo (MC)


algorithm, the random walking above all the barriers will naturally include
thermal fluctuations and entropy. Consequently, if a canonical ensemble is
used, the difference between the curve the man treads on (which is not flat)
and the curve below intuitively reflects the profile for the PMF. As a result,
the following equality exists:

F (ξ) = − lim VG (ξ, t). (6.15)


t→∞

This equation serves as the basic assumption of metadynamics. In prac-


tice, many details exist concerning the implementation of the algorithm
described above. Readers are directed to Ref. [39] for a detailed explana-
tion. We note that at first glance, this metadynamics looks similar to the
adaptive umbrella sampling, due to the fact that the sum of the Gaussian
functions gradually gives us the PMF using Eq. (6.15). In the adaptive
umbrella sampling method, a gradual history-dependent improvement on
the assessment of the PMF also exists by going through the iterations.
However, we note that the philosophy is very different. In the adaptive
sampling method, the update of the PMF using Eq. (6.13) has a rigorous
justification. On the other hand, in the metadynamics, the construction of
the PMF using Eq. (6.15) is postulated on a heuristic basis. This postulation
was later verified empirically in several systems with increasing complexity,
as can be seen, for example in Refs. [38, 39, 301–307].

6.3 Integrated Tempering Sampling


In all methods introduced above, one or more reaction coordinates are
required in order to describe the atomistic level evolution of the system in
its conformational space. By increasing the complexity of the poly-atomic
system, however, it can easily happen that the definition of such reaction
coordinates becomes difficult or even impossible. In these cases, another
kind of methods, which effectively avoid an a priori selection of the reaction
coordinates, serve as an alternative. In these methods, a commonly used
technical trick is to alter the potential energy landscape using the tempera-
ture or energy itself, so that the exploration of the conformational space can
be accelerated and the preselection of the transition coordinates is avoided.
One of the earliest examples is the Tsallis statistical method proposed in
the late 1980s [308]. In this method, a high temperature sampling, which
can be viewed as an exploration of the “potential energy surface scaled
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch06 page 127

Extension of Molecular Dynamics, Enhanced Sampling 127

by temperature”, does enforce an enhanced sampling of the rare events.


However, due to the relatively primitive algorithm, the high temperature
regions, which do not play an important role at the targeting temperature
of the canonical ensemble under investigation, can be easily oversampled
[309]. To improve on this, a series of methods were developed in the last
two decades, with the Wang–Landau [310, 311], replica exchange [34–37],
integrated tempering sampling (ITS) [44–46], and selective integrated tem-
pering sampling (SITS) [312] methods being the most prominent successful
examples. Among them, the Wang–Landau method possesses a uniform
distribution on the energy scale, which is very suitable for MC simulations.
The replica exchange, ITS, and SITS are presently very frequently used
methods in the MD simulations. A thorough description for all of them is
beyond the scope of this chapter. Here, we will take the ITS method as
an example to show how such enhanced samplings are realized when the
exploration of the PMF focuses on the energy space in the MD simulations.
The ITS method is intrinsically temperature-based, in which a gener-
alized non-Boltzmann ensemble is used. This generalized non-Boltzmann
ensemble allows enhanced sampling in a desired broad energy and temper-
ature range. The key quantity is a distribution function of the interatomic
potential. It is defined as an integral or summation of the Boltzmann terms
over temperature through
M

p(U ) = nk e−βk U , (6.16)
k=1

where U stands for the physical interatomic interaction potential and βk =


1/(kB Tk ). M is the number of temperatures used for the summation over the
Boltzmann distribution. Tk means the temperature used in each of them,
which increases with k from T1 to TM . The highest temperature used (TM )
is determined by the height of the barrier and the associated time scale one
wants to simulate. nk is a weight. It is determined through the requirement
that each term in the summation in Eq. (6.16) contributes a desired fraction
of the system’s Boltzmann distribution (at a given temperature Tk ) to the
total non-Boltzmann one.
In order to allow an MD-based method to sample the distribution
function in Eq. (6.16), one must impose an equality between a Boltzmann-
like distribution function and the non-Boltzmann one in Eq. (6.16), at a
targeting temperature T (with the corresponding β equals 1/(kB T )). To
this end, the interatomic potential must be revised. Let us assume that this
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch06 page 128

128 Computer Simulations of Molecules and Condensed Matter

revised form of the interatomic potential is U  . Then, the requirement that


the distribution function p(U ) in Eq. (6.16) can be reproduced by a finite
temperature MD simulation imposes


M

e−βU = p(U ) = nk e−βk U . (6.17)
k=1

From this equality, one easily obtains



 1 M
−βk U
U = − ln nk e , (6.18)
β k=1

and consequently, the biased force Fb on the ith atom can be calculated
from
M −βk U
k=1 nk βk e
Fib = M Fi , (6.19)
β k=1 nk e −β k U

where Fi stands for the Hellmann–Feynman force on the ith nuclei when
the inter-atomic potential is the physical one (namely U ).
From Eqs. (6.17)–(6.19), it is clear that an MD simulation using a
Boltzmann ensemble (but with an “artificial” interatomic potential) can be
employed to reproduce the distribution function associated with the non-
Boltzmann ensemble in Eq. (6.16). Due to the fact that a wide temperature
range is included in the summation over the Boltzmann distribution func-
tion in Eq. (6.16) for the non-Boltzmann ensemble, it is expected that the
probability of the rare event, which hardly happens in a normal MD sim-
ulation at the targeting physical temperature, can be increased. However,
one notes that there are still some weighting factors to be determined in
Eq. (6.19). These factors are determined through the requirement that each
term in the summation of Eq. (6.16) contributes to the total distribution
with a desired fraction. In other words, if one defines

Pk = nk e−βk U(x) dx, (6.20)

where x stands for a 3N -dimensional vector in the conformation space


(N means the number of nuclei in the poly-atomic system), each term in
the summation of Eq. (6.16) will contribute to the total distribution with
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch06 page 129

Extension of Molecular Dynamics, Enhanced Sampling 129

a fraction
Pk
pk =  M . (6.21)
k=1 Pk

These pk s should be aimed at some predetermined quantities with which


we want our finite temperature (Tk ) Boltzmann distribution to contribute
to the non-Boltzmann one. Let us label such a fraction as p0k . In practice,
the fraction that all these p0k s equal, 1/M is often used.
With these predetermined p0k s, an initial guess for the numbers nk will
be employed in the first, a certain number of, say Nτ , MD steps using the
forces generated by Eq. (6.19). The targeting temperature is T . At the end
of these Nτ steps, the values of pk will be calculated using Eqs. (6.20) and
(6.21). It is often true that these pk s are different from their targeting values
p0k s. To ensure that they fluctuate around and approach their targeting
values with the simulation going on, in the next Nτ steps, the value of nk
will be changed into its original value multiplied by p0k /pk . At the end of
these Nτ steps, the values of pk will be calculated again and the values
of nk are updated by the same relation. With the simulation going on, a
sufficient sampling of the non-Boltzmann in Eq. (6.16) will be guaranteed.
We note that this algorithm reflects only the principles underlying the ITS
method. In practical applications, there are more reliable algorithms used.
For these technical details, please refer to Gao, Yang, Fan, and Shao’s works
in Refs. [44, 46].

6.4 Thermodynamic Integration


In the earlier sections, we have introduced some extensions of the MD
method as presented in Chapter 5 by mainly focusing on the mapping of
the PMF. Another problem which can be studied using the MD simulation
technique concerns the phase behavior of a given substance, in particular,
transitions between two competing phases obtained from either random
structure searching or enhanced sampling molecular dynamics simulations.
Melting from a solid to a liquid phase is one example, and evaluation of
the relative stability between two competing solid phases is another one.
In these examples, the transition between the two competing phases is
first-order and their transition curve can be calculated from the principle
that at coexistence, the Gibbs free energies of the two phases are equal.
Here, we will introduce a method to calculate the free energy of a given
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch06 page 130

130 Computer Simulations of Molecules and Condensed Matter

phase at a finite temperature and pressure so that such phase transitions


can be studied in terms of molecular simulations. We note that continuous
transitions, which will not be discussed here, are by no means simpler.
Interested readers are directed to Ref. [313] for more details.
The method we want to introduce here is called the thermodynamic
integration method [48–58, 280]. To be clear, we start by interpreting the
principles underlying this method and then explain some technical details.
The key point of this method is that it is a general method used to determine
the free energy difference F1 − F0 between two systems, whose total energy
functions are known, denoted as U1 and U0 . By total energy function, we
mean that for a specific spatial configuration of the nuclei (R1 , . . . , RN ), the
total energy of the first system is U1 (R1 , . . . , RN ) while the total energy for
the second system is U0 (R1 , . . . , RN ). As per its definition, the Helmholtz
free energy is determined by these total energy functions through


1 −βU1 (R1 ,...,RN )
F1 = −kB T ln dR1 · · · dRN e ,
N !Λ3N

 (6.22)
1 −βU0 (R1 ,...,RN )
F0 = −kB T ln dR1 · · · dRN e ,
N !Λ3N

where Λ = h/(2πM kB T )1/2 is the thermal wavelength and M is the nuclear


mass. For a simple nomenclature, we have assumed that all the nuclei have
the same mass, but we note that the extension of Eq. (6.22) to systems
with nuclei of different kinds is straightforward for the illustration of the
principles to be discussed below.
By thermodynamic integration, we mean that if one imposes a series
of fictitious systems between the ones with free energy F1 and F0 , with the
total energy function

Uλ (R1 , . . . , RN) = U0 (R1 , . . . , RN)+λ(U1 (R1 , . . . , RN)−U0 (R1 , . . . , RN )),


(6.23)
F1 can be calculated with a thermodynamic integration treatment if F0
is known. The mathematical foundation underlying this thermodynamic
integration treatment is that for one specific value of λ, there is a total
energy function Uλ (R1 , . . . , RN ) and a free energy Fλ . It is clear from the
definition of Uλ (R1 , . . . , RN ) in Eq. (6.23) that Fλ equals F0 if λ equals
zero, and it equals F1 if λ equals one. Therefore, one can plot the evolution
of Fλ as a function of λ between zero and one, with the value of Fλ starting
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch06 page 131

Extension of Molecular Dynamics, Enhanced Sampling 131

Figure 6.5 Illustration of the evolution of Fλ in the thermodynamic integration method.


The x axis corresponds to the variable λ, which goes from 0 to 1. The y axis corresponds
R Fλλ, which goes from F0 to F1 . F0 is known while F1 is unknown, which
to the function
equals F0 + 01 dF

dλ.

from F0 and ending at F1 . Using this evolution, it is clear that


 1
dFλ
F1 − F0 = dλ. (6.24)
0 dλ
This is shown in Fig. 6.5.
Now, we look at dFλ /dλ. Similar to Eq. (6.22), Fλ is defined as


1 −βUλ (R1 ,...,RN )
Fλ = −kB T ln dR 1 · · · dR N e . (6.25)
N !Λ3N
Therefore, dFλ /dλ equals
 1 ,...,RN ) −βUλ (R1 ,...,RN )
dFλ dR1 · · · dRN dUλ (Rdλ e
=  . (6.26)
dλ dR1 · · · dRN e −βU λ (R1 , . . . , RN )
If one resorts to the definition of Uλ in Eq. (6.23), it further equals

dR1 · · · dRN [U1 (R1 , . . . , RN ) − U0 (R1 , . . . , Rn )]
dFλ e−βUλ (R1 ,...,RN )
=  . (6.27)
dλ dR1 · · · RN e−βUλ (R1 , . . . , RN )
In other words, by putting Eqs. (6.27) and (6.24) together, one arrives at
 1
F1 − F0 = dλU1 (R1 , . . . , RN ) − U0 (R1 , . . . , RN )λ . (6.28)
0

By λ , we mean that the thermal average of the quantity to be evaluated


is calculated using an ensemble average of this quantity in a system
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch06 page 132

132 Computer Simulations of Molecules and Condensed Matter

with atomic interactions Uλ (R1 , . . . , RN ). Based on this deduction, it


is clear that if the free energy of a system with total energy function
U0 (R1 , . . . , RN ) is known, one can calculate the free energy of another
system with total energy function U1 (R1 , . . . , RN ) by thermodynamically
integrating the free energy difference between F1 and F0 .
Now, we go to real poly-atomic systems. The definition of the
thermodynamic integration as defined in Eq. (6.28) is robust, which means
that as long as one has a good reference energy F0 , the free energy of the real
system F1 can be calculated through a continuous and isothermal switch-
ing of the total energy function from U0 (R1 , . . . , RN ) to U1 (R1 , . . . , RN ).
However, we note that the efficiency of this thermodynamic integration
sensitively depends on the similarity between the reference system and the
system to be calculated. Mathematically, one can relate this similarity to the
value of U1 (R1 , . . . , RN ) − U0 (R1 , . . . , RN ) being evaluated in Eq. (6.27).
The smaller this value, the fewer steps one needs to integrate for the value
of λ between zero and one in Eq. (6.28).
Traditionally, this thermodynamic integration method has been exten-
sively used in molecular simulations with empirical interatomic potentials
in the 1980s and 1990s [48–54]. In studies of liquids, this reference sys-
tem is often taken as the ideal gas with Lennard–Jones potential, while
in studies of solids, this reference system is often chosen as the harmonic
lattice. Starting from the 1990s, molecular simulations based on an ab ini-
tio treatment of the electronic structures have overtaken the traditional
method in calculations of such free energies, due to its power in describing
interatomic interactions under complex chemical environments [55–58, 280].
Since these ab initio calculations are much more expensive than the empir-
ical potential calculations, it is highly recommended that one introduces an
intermediate state between the idealized system with a rigorous analytical
expression for the free energy and the real system [55, 280]. Due to the
fact that a well-designed intermediate state held together by a well-defined
empirical potential can be more similar to the real system compared with
the idealized model, the thermodynamic integration between this interme-
diate state and the real one needs a smaller value of λ in Eq. (6.28). To
compensate, more thermodynamic integration steps will be needed from
the idealized model to the intermediate state. However, since the computa-
tional load for empirical potential-based MD simulations is much smaller,
this treatment can help us save a lot of computation time, with errors well in
control.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch06 page 133

Extension of Molecular Dynamics, Enhanced Sampling 133

Up to now, the principles underlying the thermodynamic integrations


are clear. In the following, we use the free energy of a crystal, with electronic
structures determined by the density-functional theory (DFT) calculations,
as an example to show how this is done in practice [55, 280]. We note
that the treatment of the nuclear motion in this thermodynamic integra-
tion is classical. Its extensions to the quantum nuclei will be discussed
in Sec. 7.3 by resorting to the path integral method as will be intro-
duced in Chapter 7. In the ab initio MD-based thermodynamic-integration
calculations to be discussed here, the first key quantity is the electronic
free energy U (R1 , . . . , RN ; Tel ). Compared with the total energy function
U (R1 , . . . , RN ) we have used in the earlier discussions, one may note two
differences. The first one is that we prefer the word “free energy” here while
in the earlier discussions, we use the word “total energy”. The second dif-
ference is that there is one more parameter dependence of this total energy,
on Tel , which denotes the temperature of the electrons. The physical origin
of these two differences is the thermal electronic excitations in calculations
of the electronic structures, which might be important in some situations
[55, 280]. In many practical calculations of the DFT, a smearing factor is
used to generate partially occupied orbitals so that the electronic structures
can be converged faster. We note that the smearing factor used there is not
necessarily related to the real electron temperature. But here, by Tel , we
really mean the electron temperature which is determined by the environ-
ment. The corresponding electronic structure calculations should resort to
the finite-temperature DFT developed by Mermin in the 1960s [30–32].
Therefore, the total energy function U (R1 , . . . , RN ) becomes the electronic
free energy U (R1 , . . . , RN ; Tel ), since the electronic entropy effects are natu-
rally included in this finite-temperature DFT. Accordingly, Tel -dependency
of this quantity enters. We note that this electronic free energy equals

U (R1 , . . . , RN ; Tel ) = E(R1 , . . . , RN ; Tel ) − Tel S, (6.29)

where
  
1 2
E(R1 , . . . , RN ; Tel ) = fi drψi∗ (r) − ∇ ψi (r) + n(r)Vext (r)dr
i
2

1 n(r)n(r )
+ drdr
2 |r − r |
+ E xc [n(r)] + Eion-ion (R1 , . . . , RN ), (6.30)
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch06 page 134

134 Computer Simulations of Molecules and Condensed Matter

and

S = −2kB T [fi ln fi + (1 − fi ) ln(1 − fi )] . (6.31)
i

In Eqs. (6.30) and (6.31), fi is the Fermi–Dirac occupation number



i fi |ψi (r)| .
2
of the Kohn–Sham (KS) orbital ψi at Tel and n(r) =
Equation (6.30) can be thought of as a finite-temperature version of
the total energy functional for the electronic system as introduced in
Eq. (2.30), with nuclei interactions added. Such a treatment ensures that
U (R1 , . . . , RN ; Tel ) includes the electronic entropy effects and corresponds
to the free energy of the whole system with static nuclei at a certain spatial
configuration (R1 , . . . , RN ). Therefore, for the nuclear system, it is just the
total energy. Since this energy depends on Tel , we use U (R1 , . . . , RN ; Tel )
to label this total energy function of the nuclear configuration, with the
free energy of the electronic system completely included at the finite-
temperature DFT level.
With this total energy function in hand, we can calculate the Helmholtz
free energy. The first thing we do is to separate the total energy func-
tion into two parts, i.e. the static energy at the perfect-lattice positions
U (R01 , . . . , R0N ; Tel ) and the remainder, through

U (R1 , . . . , RN ; Tel ) = U (R01 , . . . , R0N ; Tel ) + U vib (R1 , . . . , RN ; Tel ).


(6.32)

U (R01 , . . . , R0N ; Tel ) is a constant which does not depend on (R1 , . . . , RN ).


Therefore, the free energy of the crystal, as determined by

F = −kB T ln


1 −β(U(R01 ,...,R0N ;Tel )+U vib (R1 ,...,RN ;Tel ))
× dR1 · · · dRN e ,
N !Λ3N
(6.33)

equals

F = U (R01 , . . . , R0N ; Tel ) − kB T ln




1 −βU vib (R1 ,...,RN ;Tel )
× dR1 · · · dRN e . (6.34)
N !Λ3N
From this equation, it is clear that the key issue relates to determining the
second term in Eq. (6.34), which we label as F vib . This quantity, again, can
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch06 page 135

Extension of Molecular Dynamics, Enhanced Sampling 135

be separated into two terms, i.e. F harm which represents the free energy of
a harmonic lattice and the deviation of F vib from it. In terms of thermo-
dynamic integration, the reference state is the harmonic lattice, whose free
energy can be calculated analytically using
3kB T   1 βωk,s 1

F harm = ln e 2 − e− 2 βωk,s , (6.35)
Nk,s
k,s

where ωk,s means the phonon frequency of branch s at reciprocal space


point k. The deviation of F vib from F harm , which we label as F anharm , needs
to be calculated using thermodynamic integration. From our discussions,
this term equals
 1
F anharm
= dλU vib (R1 , . . . , RN ) − U harm (R1 , . . . , RN )λ . (6.36)
0

Here, U (R1 , . . . , RN ) refers to the harmonic potential associated with


harm

the Hessian matrix of the crystal. In some cases, to decrease computational


load, an intermediate state can also be used, but the principles underlying
these calculations for the free energy of a crystal is already there. For more
details, please see Ref. [55].
b2530   International Strategic Relations and China’s National Security: World at the Crossroads

This page intentionally left blank

b2530_FM.indd 6 01-Sep-16 11:03:06 AM


December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 137

7
Quantum Nuclear Effects

In Chapter 5, we have introduced the principles underlying standard molec-


ular dynamics (MD) simulations nowadays. It was clearly pointed out in
this introduction that the nuclei are treated as classical point-like particles.
This classical treatment of the nuclei is normally a good approximation
since the nuclear masses are much larger than that of the electron. However,
it needs to be pointed out that these values are still far from being large
enough so that the classical treatment is rigorous, especially for the lightest
element hydrogen, where it is only ∼1836 times of that of the electron. The
quantum nuclear effects (QNEs) might still be important in reality. As a
matter of fact, it has long been realized that the statistical properties of
hydrogen-bonded systems, such as water, heavily depend on the isotope of
hydrogen. Taking the melting/boiling temperatures of normal water (com-
posed of H2 O) and heavy water (composed of D2 O) as an example, this
value in heavy water is ∼3.8◦ C/1.4◦ C higher than that of normal water
under the same ambient pressure. In statistical mechanics, it is well known
that the classical thermal effects of different isotopes are rigorously the same
at the same temperature. Therefore, the difference between these statistical
properties in the two materials must originate from the QNEs.
Another example where the quantum feature of the nuclei plays an
important role is derived from the studies of proton tunneling, which has
been well characterized in both hydrogen diffusion on metal surface [314,
315] and proton transfer in biosystems [316]. The phenomena in the latter
system have an important influence on enzyme catalysis [316–318]. In order
to be able to account for such QNEs at the atomic level in the simulations, it
is highly desired that one can have a scheme in statistical mechanics where,
in addition to the thermal effects, the QNEs can be equally addressed. As a
matter of fact, due to the development of the path integral representation of

137
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 138

138 Computer Simulations of Molecules and Condensed Matter

the quantum mechanics starting from the late 1940s, the foundation of such
a scheme has been rigorously and systematically presented by Feynman and
Hibbs in Ref. [67]. Based on this foundation, the MD simulation technique
as introduced in Chapter 5 was combined with the framework of this path
integral representation of the statistical mechanics and a series of path
integral molecular dynamics (PIMD) simulations had been performed in
the 1980s (see, e.g., Refs. [68, 69, 74]). Parallel to these PIMD simulations,
the Monte-Carlo (MC) sampling technique was also used and properties of
liquid helium including its superfluidity were systematically studied using
this path integral Monte-Carlo (PIMC) method (see, e.g., Refs. [70–73]).
In these PIMD/PIMC simulations, the QNEs are addressed on the same
footing as the thermal ones when the statistical properties of the system
to be simulated are evaluated, as will be explained in detail soon in this
chapter. Therefore, when comparing their results with the ones where the
nuclei are treated as classical particles, such as the MD simulation method
we have introduced in Chapter 5, the differences account for the impact of
the QNEs on the statistical properties in a very clear manner.
This comparison sets up a rigorous framework for the QNEs to be ana-
lyzed, which is still used nowadays. However, it needs to be pointed out that
in both these early PIMD and PIMC simulations, the empirical potentials
were used to account for the interatomic interactions in the simulation.
These potentials are simple and very good in describing the statistical
property of many solids and liquids. However, they can easily fail when
chemical reaction happens, due to a serious reconstruction of the electronic
structures, which needs to be addressed “on-the-fly” as the dynamics of
the system evolves. To address problems like this, where the impact of
the QNEs is often more interesting, people started trying a combination
of the PIMD and PIMC simulation techniques with the ab initio method
for the description of the electronic structures after the 1990s, first within
the framework of Car–Parrinello (CP) MD (see e.g. Refs. [319–321]) and
then directly on the Born–Oppenheimer (BO) MD or MC schemes [282, 286,
322, 323]. These methods really allow the bond making and bond breaking
events to happen, as well as the thermal and quantum nuclear effects to
be accounted for in a seamless manner based on the forces computed “on-
the-fly” as the dynamics of the system evolve. Now, it is fair to say that
they have come to such a mature stage that not only different functionals
with the density-functional theory (DFT) can be used in descriptions of
the electronic structures (see, e.g., Refs. [324, 325]), but also traditional
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 139

Quantum Nuclear Effects 139

quantum chemistry methods such as the MP2 method can be used (see,
e.g., Refs. [326, 327]). With these choices of the electronic structures, when
the interatomic interactions are accurate enough and the sampling over
the high-dimensional phase space is complete (ergodicity is satisfied, in the
language of statistics), one can safely rely on results obtained from such
simulations, even under low temperatures when the classical description of
the nuclei fails. Therefore, on the statistical level, a scheme in which the
thermal nuclear effects and QNEs are accounted for on the same footing
in the atomic simulations is already there. The only thing we need to be
attentive to is in its use the choice of the electronic structures for the
description of the interatomic interactions and the ergodicity issue in the
PIMD/PIMC sampling (which, however, is non-trivial at all).
Besides these statistical properties, another kind of property where the
QNEs may also play an important role relates to the dynamics, especially
when the chemical reaction rate is evaluated. This chemical reaction rate is
a key parameter in chemistry which is very hard to simulate rigorously. One
theory underlying descriptions of this quantity is the so-called transition-
state theory (TST) [328–331]. Since the probability of finding the system
close to the transition state (TS) is much smaller than that of the reactant
or product state, this theory is intrinsically both statistical and dynamical.
The term “statistical” means that this chemical reaction rate is propor-
tional to the ratio of the equilibrium density of the system at the TS to
its value at the reaction state, which is a statistical property. Conversely,
the term “dynamical” indicates that since the TS is defined by a divid-
ing surface separating the reaction and product states, after the system
at the TS falls into either the reactant or the product region, it stays at
this state for a longer time than it spends at the TS. Therefore, what
happens dynamically at the TS is of crucial importance to its behavior in
the future, and theories underlying descriptions of such processes should be
dynamical. Following the principles of scientific research, i.e. from the easy
and idealized models to the more difficult and realistic ones, the earliest
methods within the TST usually assume a classical treatment of the nuclei.
Later, when events like quantum tunneling were found to be crucial in
describing the chemical reaction behaviors, a quantum version of this was
also proposed. This development is also associated with the development of
the purely statistical PIMD method to the dynamical regime. Two of the
schemes most often used in descriptions of this dynamics within the scheme
of path integral molecular dynamics are the centroid molecular dynamics
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 140

140 Computer Simulations of Molecules and Condensed Matter

(CMD) and the ring-polymer molecular dynamics (RPMD). This exten-


sion of the statistical PIMD method to the dynamical regime will also be
discussed.
In Chapter 6, a scheme in which the free energy with anharmonic con-
tributions from the nuclei is calculated using the sampling method of MD
was introduced. In our discussions there, the assumption that the nuclei
are classical particles results in the fact that although the anharmonic con-
tribution from the thermal fluctuations of the nuclei are accounted for in
the thermodynamic integrations with MD sampling, the anharmonic effect
associated with the quantum feature of the nuclei is completely neglected.
This anharmonic contribution originating from the QNEs is often believed
to be unimportant at moderate and high temperatures. However, when the
mass of the particle is small, this assumption might fail, starting from the
moderate temperature regime to low temperatures. As a matter of fact, this
anharmonic correction is recently found to be very important in describing
the phase diagram of hydrogen and neon (see e.g. Refs. [332, 333]). There-
fore, from both the methodology and the practical simulation point of view,
it is highly desired that a scheme in which the thermodynamic integration
is combined with the ab initio PIMD method is developed to address such
problems. This treatment of the anharmonic quantum nuclear correction
to the free energy will also be discussed [333].
The chapter is organized as follows. In Sec. 7.1, we introduce the PIMD
and related methods for statistics, where it is rigorously justified. Then,
some extensions beyond these statistical studies will be briefly discussed in
Sec. 7.2. After these, an introduction to how this PIMD method is combined
with the thermodynamic integration method for the calculation of the free
energy will be presented in Sec. 7.3. We end this chapter with some examples
in Sec. 7.4 and a brief summary in Sec. 7.5.

7.1 Path Integral Molecular Simulations


7.1.1 Path Integral Representation of the Propagator

For a theoretical description of the QNEs, it is crucial to start our discus-


sions from their origin, i.e. the intrinsic quantum nature of the nuclei. The
development of quantum mechanics in the last century shows us that the
fundamental difference between the classical world and the quantum world
lies in the fact that in the quantum world, things must be described in
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 141

Quantum Nuclear Effects 141

terms of “probability”. In understanding this principle, we can make use of


a scene that many of us might have experienced during our primary school
time, as an analogy. We remember that when a naughty boy makes a loud
noise during the class and irritates his teacher, the teacher will throw a piece
of chalk he is holding toward this guy to remind him to be quiet. In most
cases, it works. In the language of classical/quantum mechanics, we would
like to say that it works because the piece of chalk is heavy enough so that
it behaves like a classical particle, and the classical particles move according
to their trajectories. Therefore, as long as you control the trajectory, you
control the consequence.
Now, imagine that this chalk is a particle small enough so that its
quantum feature is important in descriptions of the phenomena related to
it, what happens after the chalk is thrown out will then be completely out
of control due to the principle of quantum mechanics. Every person in the
classroom might be hit at a later time. From our textbook of quantum
mechanics, we know that in order to quantitatively describe behaviors like
these, a propagator needs to be used. This propagator is a function of two
events, with each event representing something happening at a certain time
and a certain position. Still taking the classroom with a naughty boy and
his teacher as an example, we can label the event when the teacher throws
the piece of chalk on the stage (position labeled as xa ) at a certain time ta
as event a, and the event when “someone” got hit by the chalk at his/her
position (labeled as xb ) at a later time tb as event b.
If we forget about the person who did this and only focus on the parti-
cle, these two events can be rephrased as “the generation of a particle (piece
of chalk) at a and the annihilation of this particle at b”. As mentioned in
Chapter 4, in a many-body quantum entity, the correlation between these
two events should be described using the propagator, which is also known as
Green’s function. After event a happens, the probability of event b happen-
ing equals the square of the propagator’s absolute value. In the Schrödinger
representation of the quantum mechanics, this propagator is written as

G (xb , tb ; xa , ta ) = ψj (xb )ψj∗ (xa )e(−i/)Ej (tb −ta ) , (7.1)
j

where j runs over all eigenstates of the quantum system. This equation
indicates that if one knows the eigenstate wave functions and eigenvalues
of this quantum system, this propagator can be expressed analytically
and consequently, the correlation between any two events is accurately
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 142

142 Computer Simulations of Molecules and Condensed Matter

described. However, it is well known that the many-body Schrödinger


equation is difficult to solve and it has a notorious scaling problem.
Nowadays, many studies on the QNEs still resort to such a method, where
the high-dimensional ab initio potential energy surfaces (PESs) are mapped
out first and then the Schrödinger equation is solved directly [59–62]. This
method is rigorous because not only nuclear exchange but also real-time
propagation can be described rigorously in this framework [60]. But due
to the scaling problem associated with both mapping the ab initio high-
dimensional PESs and solving the Schrödinger equation, its application is
seriously limited to systems less than ∼6 atoms. When the system gets
bigger, a practical scheme for descriptions of this quantity must be used.
Thanks to the development of the path integral representation of the
quantum mechanics starting from the late 1940s by Feynman [63–67], a
framework where this propagator is calculated using a method we would
like to use in this chapter was systematically presented by Feynman and
Hibbs in their seminal book in 1965 (Ref. [67]). In this book, it was clearly
explained that this propagator can be calculated not only from Eq. (7.1)
but also in terms of a numerical path integral, in which contributions from
all paths between events a and b are taken into account. The trick is to
divide the time interval between these two events, i.e. tb − ta , into P slides.
Draw a line between ta and tb which intersects with ti on xi . Then, on each
time slide ti , move the spatial coordinates xi through the whole Cartesian
space. When P equals infinity, all paths between these two events will be
taken into account. The propagator is calculated by adding contributions
from all these paths into one quantity through
  
1 dx1 dx2 dxP −1
G (xb , tb ; xa , ta ) = lim ··· e(i/)S[b,a] ··· .
P →∞ A V V V A A A
(7.2)
1
Here, A is a renormalization factor which equals [2πi(tb − ta )/(P m)] 2 .
S[b, a] is the action of the path linking events a and b, defined by the spatial
coordinates x1 , x2 , . . . , and xP −1 on t1 , t2 , . . . , tP −1 . For one specific path,
as shown in Eq. (7.2), its contribution to the propagator is determined by
the action of this path, which is calculated from
 tb
S[b, a] = L(ẋ, x, t)dt. (7.3)
ta

To be more precise, taking the choice of path x1 , x2 , . . . , xP −1 happening


at times t1 , t2 , . . . , tP −1 , as an example, the action of this path as defined
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 143

Quantum Nuclear Effects 143

by Eq. (7.3) can be written as


  2 
m x1 − xa 1
S[b, a] = − (V (x1 ) + V (xa )) (t1 − ta )
2 t1 − ta 2

P −1
  2 
 m xi − xi−1 1
+ − (V (xi ) + V (xi−1 )) (ti − ti−1 )
i=2
2 ti − ti−1 2
  2 
m xb − xP −1 1
+ − (V (xb ) + V (xP −1 )) (tb − tP −1 ) .
2 tb − tP −1 2
(7.4)

With this definition, it is clear that the integration in Eq. (7.2) can be
calculated numerically through such a procedure, as shown in Fig. 7.1 for
the 1D case. In practice, a finite number of time intervals must be chosen,
one often tests the convergence of the quantity to be calculated with respect
to the number of slides till a reasonable accuracy can be obtained.
With this, we hope that we have made our point clear. To put it simply,
the same quantity, i.e. the propagator, can be obtained by using either (i)
all eigenstate wave functions of the quantum system, or (ii) a sum over
contributions from all paths between the events to be investigated. The
first option looks elegant, but it is difficult to handle for large many-body
systems. In cases when it is not feasible, the path integral approach provides
a numerically simple alternative, which we will make use of in studies of
QNEs to be discussed in the following sections.

7.1.2 Path Integral Representation of the Density Matrix

So far we have discussed the propagator; in statistical mechanics, the key


quantity we are interested in is actually not the propagator, but the density
matrix. Therefore, in order to understand how the concept of the path
integral is used in studies on statistical mechanics, the key point is to
understand how this quantity (density matrix) is expressed in terms of
path integral. As the basis, the first thing we explain here is how this key
quantity is expressed in quantum mechanics in general.
Imagine that the quantum system we want to study has exact eigenstate
wave functions ψi , eigenvalues Ei , and a Hamiltonian Ĥ. In the operator
notation, the density operator is e−β Ĥ , where β = 1/(kB T ) and T is the
temperature. The trace of this operator is the so-called partition function
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 144

144 Computer Simulations of Molecules and Condensed Matter

Figure 7.1 Illustration of how a propagator can be calculated using path integral. The
first step is to divide the time interval into some slides. At the starting and ending points
(ta and tb ), the spatial coordinates (xa and xb in this 1D case) must be fixed since they
represent the events whose correlation is to be investigated. With the spatial coordinates
on the time slices between ta and tb chosen, we have one specific “path”. Contribution
from this path to the propagator is then calculated using Eq. (7.4). Then one further
moves the spatial coordinates on each time slide between ta and tb through the whole
space (x axis in this 1D case). In doing so, we can take account of contributions from
all paths between a and b. If all these contributions are calculated numerically using
Eq. (7.4) and the time interval number approaches infinity, we have the propagator
rigorously defined which equals the number it gives through Eq. (7.1). Calculations of
the eigenstate wave functions are avoided accordingly.

Z Q . The expectation value of any physical observable Ô equals

Ô = (Z Q )−1 Tr[Ôe−β Ĥ ]. (7.5)

Now, we discuss how these quantities are represented in two spaces, i.e.
the system’s eigenstate wave functions’ Hilbert spaces and the position
space. Such a comparison can help us in setting up a link between the
density matrix and the propagator we have discussed in Sec. 7.1.1, which
you will see soon. We first look at the system’s eigenstate wave functions’
Hilbert space, where the probability of finding the quantum system at its
ith eigenstate equals e−βEi at thermal equilibrium. From this probability,
it is clear that for an operator Ô, its expectation value at the thermal
equilibrium equals

Ô = (Z Q )−1 ψi |Ô|ψi e−βEi . (7.6)
i
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 145

Quantum Nuclear Effects 145

The partition function Z Q itself has the simple form



ZQ = e−βEi . (7.7)
i

Then we come to the position space. In this space, the density matrix
is written as
ρ (xa , xb , β) = xb |e−β Ĥ |xa , (7.8)
whose diagonal part is the so-called density function. To arrive at the wave-
function-based expression of this density matrix, one can insert an identity

matrix Iˆ = j |ψj ψj | into the right-hand side of this equation. In doing
so, we have
⎛ ⎞ 

−β Ĥ ⎝
ρ (xa , xb , β) = xb |e |ψj ψj |⎠ |xa
j

= xb |ψj e−βEj ψj |xa 
j

= ψj (xb )ψj∗ (xa )e−βEj . (7.9)
j

Similar to the propagator, to obtain this quantity, in the Schrödinger repre-


sentation of the quantum mechanics, we need to know the eigenstate wave
functions. However, as mentioned, calculating the wave function of nuclei is
not feasible for systems with more than ∼6 atoms. An alternative approach
must be adopted.
Now, we apply the concept of the path integral as introduced in
Sec. 7.1.1 to the description of this density matrix, the non-local func-
tion as shown in Eq. (7.9). The trick is to compare it with Eq. (7.1) for the
propagator. From this comparison, it is easy to see that these two equations
for the propagator and the density matrix share a very strong similarity.
The only difference is that in Eq. (7.1) for the propagator, the index of the
exponential function is imaginary and it is determined by the time interval
tb −ta , while in Eq. (7.9) for the density matrix, the index of the exponential
function is real and it is determined by the temperature T . In Sec. 7.1.1,
we have shown that for the propagator in Eq. (7.1), we can avoid using the
wave functions by resorting to integrations over paths between the events
to be described. Similarly, with this tiny difference between Eq. (7.9) and
Eq. (7.1) in mind, we can also resort to the path integral representation of
the quantum mechanics and rewrite the density matrix as shown in Eq. (7.9)
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 146

146 Computer Simulations of Molecules and Condensed Matter

in terms of the path integral. The only thing we need to do is to treat the
temperature-dependent factor β as an imaginary-time interval and replace
the i(tb −ta )/ term in Eq. (7.1) by this temperature-dependent imaginary-
time interval. To be more precise, we should replace the time interval tb − ta
as used in Eq. (7.1) by −iβ. Then, using the same trick as that mentioned
for the numerical representation of the propagator in Eqs. (7.2)–(7.4), we
can rewrite the density matrix in terms of path integral without resorting
to the nuclear wave functions.
Following this routine, the first thing we need to do is to divide the
time interval −iβ into P slices. Such a treatment results in a time step
of −iβ/P along the imaginary time path to be integrated in the density
matrix. Now, imagine that one path is defined by a certain choice of x1 ,
x2 , . . . , xP −1 between xa and xb , the action which determines the weight
of this path in the path integral scheme then equals
 
mP 2 2 1 −iβ
S[b, a] = − 2 2
(x1 − x a ) + (V (x 1 ) + V (xa ))
2β  2 P
P
 −1  
mP 2 2 1 −iβ
− 2 2
(xi − xi−1 ) + (V (xi ) + V (xi−1 ))
i=2
2β  2 P
 
mP 2 2 1 −iβ
− (xb − xP −1 ) + (V (xb ) + V (xP −1 ))
2β 2 2 2 P
(7.10)

from Eq. (7.4). We note that because imaginary time is used, the positive
sign of the kinetic energy becomes negative, resulting in the action (which
is originally defined as an integral over the Lagrangian) as a term above
which looks like an integral over a Hamiltonian with a minus sign. The
(i/)S[b, a] term on the index of the exponential functional in Eq. (7.2)
consequently becomes
 
mP 2 β
(i/)S[b, a] = − (x1 − xa ) + (V (x1 ) + V (xa ))
2β2 2P
P
 −1  
mP 2 β
− 2
(xi − xi−1 ) + (V (xi ) + V (xi−1 ))
i=2
2β 2P
 
mP 2 β
− (x b − xP −1 ) + (V (xb ) + V (x P −1 )) .
2β2 2P
(7.11)
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 147

Quantum Nuclear Effects 147

To clarify the labeling in the above equations further, we can relabel the
xa and xb in the density matrix as x0 and xP , respectively. The imaginary-
time interval is still divided into P slices and the path is determined by
a consequence of spatial coordinates from x1 to xP −1 , with x0 and xP
keeping fixed. Then, Eq. (7.11) can be rewritten in a simple form as

P  
mP 2 β
(i/)S[b, a] = − (xi − x i−1 ) + (V (xi ) + V (xi−1 ))
i=1
2β2 2P

P 
 
1 1
= −β mωP2 (xi − xi−1 )2 + (V (xi ) + V (xi−1 )) ,
i=1
2 2P
(7.12)

where ωP = P /(β). If we put this exponential index to the path integral
representation of propagator as shown in Eq. (7.2), but using the imaginary
time for the density matrix, the equation this density matrix ends up with
will be

ρ (x0 , xP , β)
   PP
1 1 2 2 1
= lim ··· e−β i=1 2 mωP (xi −xi−1 ) + 2P
[ (V (xi )+V (xi−1 ))]
P →∞ A V V V

dx1 dx2 dxP −1


· ··· . (7.13)
A A A
In the case of the real-time propagator, we have mentioned that A is a
renormalization factor which equals [2πi(tb − ta )/(P m)]1/2 . Here, for the
imaginary-time density matrix, tb − ta should be replaced by −iβ. Accord-
ingly, A becomes [2πβ2 /(P m)]1/2 . Replacing the A in Eq. (7.13) with this
value, we finally arrive at

ρ (x0 , xP , β)
  P2   
mP
= lim ···
P →∞ 2βπ2 V V V
PP 1 2 2 1
· e−β i=1 2 mωP (xi −xi−1 ) + 2P
[ (V (xi )+V (xi−1 ))]
dx1 dx2 · · · dxP −1 .
(7.14)

Till now, the numerical representation of the density matrix in terms of


the path integral is already clear. However, there is still a key concept whose
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 148

148 Computer Simulations of Molecules and Condensed Matter

physical meaning needs some further explanation. In the above discussions,


the variables x0 to xP are defined as points in the Cartesian space for the
nucleus to be studied for simplicity. However, we note that they can also be
used to represent a poly-atomic system’s spatial configuration of the nuclei
under investigation. Suppose that this poly-atomic system is a molecule
containing N nuclei, this xi is then a 3N -dimensional vector, representing
a spatial configuration of this molecule’s nuclei. To understand this, one just
needs to imagine that the wave functions as used in Eqs. (7.1) and (7.9) are
many-body wave functions of this N -nuclei system whose square magnitude
represents the probability of finding the poly-atomic system at this specific
spatial configuration. Accordingly, the path from x0 , through x1 , . . . , xP −1 ,
to xP represents a path in the 3N -dimensional configuration space linking
x0 and xP , two spatial configurations of the nuclei. All discussions above
about the position space extend to this 3N -dimensional configuration space.
The only impact on the equations above is that in Eq. (7.14), an extra
iteration over the atoms needs to be included. With this extra iteration
included, Eq. (7.14) becomes
ρ (x0 , xP , β)
⎡ ⎤
N   P2   
mj P
= lim ⎣ ⎦ · · ·
P →∞
j=1
2βπ2 V V V

PP hP 2
i
·e
−β i=1
N 1 2
j=1 2 mj ωP (xji −xji−1 ) 1
+ 2P (V (x1i ,...,xN 1 N
i )+V (xi−1 ,...,xi−1 ))

· dx1 dx2 · · · dxP −1 . (7.15)


Here, xji means the 3D position space associated with the jth atom’s ith
bead. We note that this understanding of xi and Eq. (7.15) associated are
the frequently used theoretical foundations for discussions of the density
matrix in the path integral molecular simulations. Again, P is a parameter
which represents the number of slices sampled along the path. The conver-
gence of the property under investigation with respect to this parameter
must be tested in practical simulations.

7.1.3 Statistical Mechanics: Path Integral Molecular


Simulations

Equation (7.15) gives an expression for the density matrix in terms of the
path integral in the 3N -dimensional configuration space, where N is the
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 149

Quantum Nuclear Effects 149

number of nuclei in the poly-atomic system. The diagonal part of this


density matrix is the density function, which describes the probability of
finding the poly-atomic system under investigation at a certain spatial
configuration. This function can be obtained from Eq. (7.15) by setting
x0 = xP as

ρ (xP , xP , β)
⎡ ⎤
N   P2   
mj P
= lim ⎣ ⎦ · · ·
P →∞
j=1
2βπ2 V V V

PP hP i
·e
−β i=1
N 1 2
j=1 2 mj ωP (xji −xji−1 )2 + 2P
1
(V (x1i ,...,xN 1 N
i )+V (xi−1 ,...,xi−1 ))

· dx1 dx2 · · · dxP −1 . (7.16)

In Sec. 7.1.2, we stated that the density matrix is e−β Ĥ in the operator
notation, whose trace gives us the partition function. Therefore, in the
configuration space, this partition function is easily obtainable from the
density matrix. What we need to do is to perform an extra integration on
xP over the configuration space in Eq. (7.16). This integration gives us the
partition function Z Q for the quantum canonical system in the limit of P
approaching infinity through

⎡ ⎤
N 
  P2   
⎣ m j P ⎦
ZQ = lim ···
P →∞
j=1
2βπ2 V V V

PP hP i
·e
−β i=1
N 1 2
j=1 2 mj ωP (xji −xji−1 )2 + 2P
1
(V (x1i ,...,xN 1 N
i )+V (xi−1 ,...,xi−1 ))

· dx1 dx2 · · · dxP . (7.17)

This quantum canonical partition function is a function of temperature.


When it is known, in principle, all thermodynamic quantities of the quan-
tum system under investigation are obtainable.
Now, we look at the partition function in Eq. (7.17). In practice, a finite
P is always used. We label the partition function represented in Eq. (7.17)
for a finite P as ZP . We note that this ZP can be understood as the “con-
figurational” partition function for a fictitious 3N × P -particle system in
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 150

150 Computer Simulations of Molecules and Condensed Matter

an effective potential V eff , in the form of


V eff (x1 , x2 , . . . , xP )

P N
⎣ 1 1
= mj ωP2 (xji −xji−1 )2 + (V (x1i , . . . , xN
i )
i=1 j=1
2 2P

+V (x1i−1 , . . . , xN
i−1 )) , (7.18)

where x0 = xP . Again, xi is a 3N -dimensional vector representing the


spatial configuration of the nuclei in the ith image and xji is a 3D vector
representing the position of the jth nucleus in this image. P is the number
of slices we have chosen for the sampling of the path integral along the
imaginary-time interval. Because of the similarity between the cyclic path
and a necklace, these sampling points are also called “beads”, “images”, or
“replicas” in the literature. In terms of this V eff , ZP in Eq. (7.17) can be
rewritten as
⎡ ⎤
N   P2   
mj P eff
ZP = ⎣ 2
⎦ · · · e−βV (x1 ,x2 ,...,xP ) dx1 dx2 · · · dxP .
j=1
2βπ V V V

(7.19)
The relationship between this configuration partition function (ZP ) over a
3N × P configuration space and the quantum canonical partition function
(Z Q ) of the real poly-atomic system is that ZP equals Z Q when P goes to
infinity.
Pictorially, this relationship can be understood from the comparison
as shown in Fig. 7.2. We use H2 , the simplest molecule, as an example.
The canonical partition function of the quantum system is what we want
to simulate. From Eq. (7.9), we know that one needs the eigenstate wave
functions and the eigenvalues of the nuclei, which is not feasible in the
studies of most poly-atomic systems, except for descriptions of some simple
gas phase small molecules [59–62]. As an alternative, one can resort to
Eq. (7.17) and construct a fictitious polymer for the real system under
investigation. This polymer is composed of P replicas of the real poly-
atomic system. In between the replicas, the same atoms are linked by spring
interactions, whose spring
√ constant (defined as mj ωP2 ) is determined by mj
and ωP , with ωP = P /(β). Within one replica, the interatomic potential
is calculated by either force fields or ab initio methods upon which the
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 151

Quantum Nuclear Effects 151

Figure 7.2 Illustration of how the mapping from the canonical quantum system to a
classical polymer is done in path integral statistical mechanics. The simplest molecule,
H2 , is taken as an example. In principle, one needs to calculate the nuclear wave functions,
as shown on the left. Using path integral, this calculation of the wave function can be
avoided. What one needs to do is to set up a fictitious polymer. This polymer is composed
by P replicas of the real molecule. In each replica, the potential is determined by the real
potential of the system at the specific spatial configuration of this replica. In between the
replicas, the neighboring images (beads) of the same atoms are linked √ by springs. The
spring constant is determined by mj and ωP as mj ωP 2 , where ω = P /(β). Therefore,
P
the higher the temperature, the heavier the nucleus, the stronger the interaction between
the beads. In the limit of T → ∞ and mj → ∞, one arrives at the classical limit when all
images overlap with each other. The partition function of the quantum system as shown
on the left equals the configurational partition function of the polymer on the right as
P → ∞.

molecular simulation is based. These two terms, i.e. the spring interaction
and the intra-replica potential, correspond to the kinetic and the potential
energies of the path integral, respectively. From the form of the spring
constant as shown in Eq. (7.17), it is clear that as the temperature and
mj go to infinity, the spring constant becomes so large that all the replicas
overlap in configuration space, resulting in a simulation in the classical limit.
When the temperature is low and mj is reasonably small, it is reasonable
to expect that a molecular simulation based on this polymer gives results
very different from the one with P = 1, i.e. the classical simulation. In
other words, from P = 1 (the classical simulation), when P approaches
infinity, one approaches the quantum limit of this canonical ensemble. In
practice, a finite P must be taken. The statistical results obtained from
these simulations with a finite P should always be converged with respect to
this P . The difference between results obtained from the P = 1 simulation
and this path integral converged simulation tells us the impact of QNEs on
the statistical results.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 152

152 Computer Simulations of Molecules and Condensed Matter

Figure 7.3 An example of the artificial polymer in a real path integral molecular sim-
ulation, courtesy of Dr. Brent Walker from our joint paper [322]. The system is a layer
of squaric acid, a molecular crystal held together with intermolecular hydrogen bonds.
The quantum nature of the nuclei is addressed using the path integral treatment by
generating a series of images for the real poly-atomic system and connecting the same
atom in neighboring images with spring potential. The electronic structures are calcu-
lated quantum mechanically for each image, and the blue contour denotes the density
distribution of the electrons in one image. For more details, please see Ref. [322].

In order to provide a more vivid explanation of this artificial polymer in


path integral simulation (with electronic structures calculated “on-the-fly”
in an ab initio manner as the dynamics of the system evolves), we take
a real system, i.e. a layer of squaric acid, a hydrogen-bonded molecular
crystal, as an example and show a schematic scheme in Fig. 7.3. P is set
as 16 and the temperature is 100 K. Around 16 images of the real system
are generated and the same atoms of the neighboring images are connected
by artificial spring interactions. As said, for the electronic structures of the
system in each image, they are calculated quantum mechanically using the
ab initio method. This is shown by blue contours designating the density
distribution of the electrons. Since the mass of the hydrogen is small, the
spring interaction with which the hydrogen nuclei is connected is weak and
consequently the images of the hydrogen nuclei are delocalized in real space.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 153

Quantum Nuclear Effects 153

The mass of the oxygen nucleus is larger and consequently, their dispersions
are smaller, but still observable. If we compare results of simulations using
this polymer and the classical system (P = 1), it is obvious that the QNEs
are small for O and large for H. With the increase of the temperature, it
is also easy to expect that the difference between simulations with P = 1
and larger P decreases. For more details on this path integral simulation,
readers are directed to Ref. [322].
Then we come back to the mathematics. With this relationship between
the “configurational” partition function of the polymer and the quan-
tum partition function as introduced in the early paragraphs in mind,
we now shift our attention to this “configurational” partition function.
Equation (7.19) for this “configurational” partition function of a polymer
may look a little complicated to be handled numerically at first sight. How-
ever, for people working on molecular simulations, this is an equation which
cannot be more simple. To understand what we mean by this statement,
one just needs to rewrite ZP using a proper partition function in a fictitious
phase space, composed of variables x1 , . . . , xP , p1 , . . . , pP , as
 
ZP = C · · · e−βH(x1 ,...,xP ,p1 ,...,pP ) dx1 · · · dxP dp1 · · · dpP . (7.20)

The Hamiltonian in this equation is designed using the effective potential


V eff through
P 
 N
(pj )2
H(x1 , . . . , xP , p1 , . . . , pP ) = i
j + V eff (x1 , x2 , . . . , xP ).
i=1 j=1 2Mi
(7.21)
Since the kinetic energy and the potential energy terms are separable in
this Hamiltonian, one can easily replace the P -dependent constant
N   P2
mj P
j=1
2βπ2

in Eq. (7.19) by a product of N × P uncoupled Gaussian integrals, which


originate from the integrals over momentum pji , as in Eq. (7.20). The con-
stant C in front of the integration over configurational and momentum
space in Eq. (7.20) ensures that ZP is unchanged compared to Eq. (7.19).
It is determined by the choice of the artificial mass Mij associated with the
beads of each nucleus.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 154

154 Computer Simulations of Molecules and Condensed Matter

We note that such an equality means that for a finite P , the path inte-
gral of a quantum system is isomorphic to a classical polymer composed of
P replicas of the real poly-atomic system under investigation, subject to a
classical Hamiltonian given by Eq. (7.21) [68]. In doing so, the molecular
simulation techniques as introduced in the earlier chapters can be directly
used. These molecular simulation techniques include both MD and MC
methods, with which the partition function as given in Eq. (7.20) can be
evaluated. In the following, we discuss how these path integral-based molec-
ular simulation techniques are used, taking the MD-based implementations
as the guiding example. The only thing we need to take care in is to design
the fictitious polymer in a proper way, so that the contributions from the
quantum nature of the nuclei to the statistical properties of the system are
not miscounted. Readers interested in the PIMC method are directed to
peruse Refs. [70–73].
Besides the partition function and the density matrix, other quantities,
such as the expectation values of any physical observable Ô, can also be
described using this path integral. The definition of such an expectation
value is already given in Eq. (7.5). From this definition, one can obtain a
path integral-based representation for the expectation value of this observ-
able using the Trotter factorization. To understand this statement, we first
go back to the path integral representation of the propagator as shown
in Eq. (7.2). As a matter of fact, its form originates and is equivalent to
a Trotter factorization-based equation. What one does in such a Trotter
factorization is to insert P − 1 identity matrices dxi |xi xi | = I on the
time slices between ta and tb . With this treatment, one arrives at such an
equation for the propagator

G (xb , tb ; xa , ta)
  
1
= lim ··· xa |e(i/)L̂Δt |x1 x1 |e(i/)L̂Δt |x2  · · ·
P →∞ A V V V

dx1 dx2 dxP −1


·xP −1 |e(i/)L̂Δt |xb  ··· . (7.22)
A A A

We note that this equation is equivalent to the path integral treatment in


terms of action in Eqs. (7.2)–(7.4).
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 155

Quantum Nuclear Effects 155

With this equality in mind, one can rewrite the expectation value of Ô
in the configuration space as
 
Ô = (Z Q )−1 Tr Ôe−β Ĥ
  
1 1 1
= lim ··· x0 |Ôe− P β Ĥ |x1 x1 |e− P β Ĥ |x2  · · ·
P →∞ ZP V V V

− P1β Ĥ
·xP −1 |e |xP dx1 dx2 · · · dxP
  
1 1 1
= lim ··· O(x1 )x0 |e− P β Ĥ |x1 x1 |e− P β Ĥ |x2  · · ·
P →∞ ZP V V V

− P1β Ĥ
·xP −1 |e |xP dx1 dx2 · · · dxP , (7.23)

where x0 = xP and O(x1 ) denotes the expectation value of Ô at x1 . Due


to the fact that Ĥ/(P β) is an infinitely small Trotter factor, the same
treatment also applies to x2 , as
  
1 1 1
Ô = lim ··· x0 |e− P β Ĥ |x1 x1 |Ôe− P β Ĥ |x2  · · ·
P →∞ ZP V V V

− P1β Ĥ
·xP −1 |e |xP dx1 dx2 · · · dxP
  
1 1 1
= lim ··· O(x2 )x0 |e− P β Ĥ |x1 x1 |e− P β Ĥ |x2  · · ·
P →∞ ZP V V V

− P1β Ĥ
·xP −1 |e |xP dx1 dx2 · · · dxP . (7.24)

Then we can continue such a cycling to xP and make an average over all
the images. Such a treatment gives us an expression for the expectation
value of Ô as

    P

1 1 
Ô = lim ··· O(xi )
P →∞ ZP V V V P i=1
1 1
×x0 |e− P β Ĥ |x1 x1 |e− P β Ĥ |x2  · · ·
1
·xP −1 |e− P β Ĥ |xP dx1 dx2 · · · dxP . (7.25)
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 156

156 Computer Simulations of Molecules and Condensed Matter

In doing so, the equation from which the expectation value of any
physical observable is calculated in the frequently used path integral scheme
is arrived at.
We note that most of the studies in this field focus on the expec-
tation value of local physical quantities [282, 286, 321, 322], such as the
density function. In recent years, however, simulations for some non-local
physical quantities such as the momentum distribution of a proton in water
have also attracted much attention from the theoretical perspective, espe-
cially after the deep inelastic neutron scattering (DINS) experiment became
available [334–336]. In these cases, rigorously speaking, the so-called open-
path integral molecular dynamics method should be resorted to (see e.g.
Morrone, Lin and Car’s work in Refs. [337–339]), where the constraint
x0 = xP is released. This open path allows the density matrix itself to
be simulated. However, we note that compared to the normal PIMD simu-
lations with close path, special care must be taken in the open-path simula-
tions concerning its stability. To deal with the problem, some new methods
on simulating the momentum distribution using the conventional close-path
PIMD method were also proposed; see e.g. Refs. [340–342]. Readers inter-
ested in theoretical simulations concerning this quantity may refer to two
sets of works in particular: Morrone, Lin, Car, and Parrinello’s in Refs. [337–
341], and Ceriotti and Manolopoulos’, in Ref. [342].

7.1.4 Staging and Normal-Mode Transformations

In terms of molecular dynamics, the canonical sampling associated with the


partition function in Eq. (7.20), the density function in Eq. (7.15), and the
expectation value of a physical observable in Eq. (7.25) can be obtained
using the equations of motion of the fictitious polymer resulting from the
Hamiltonian in Eq. (7.21).
A number of well-known numerical difficulties, however, exist in such
a straightforward implementation of the PIMD method. These difficulties
mainly√arise from three aspects. First, from the definition of ωP , which
equals P /(β), it is clear that the stiffness of the spring constant increases
with P . Therefore, with the increase in the number of beads, if the masses
of the beads are independent of it, the spring interaction requires smaller
and smaller time steps for the PIMD simulations to be carried out in order
to characterize the interbead vibrations. Second, the stiffness of this spring
interaction results in the external potential generated from the ab initio
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 157

Quantum Nuclear Effects 157

(or force field) calculations within each image serving only as a small per-
turbation to the spring interaction. Consequently, the trajectories for the
beads of the artificial polymer will remain close to “invariant tori” in the
Cartesian space and efficient sampling of its entire configuration space is
seriously hindered. Third, even if this multi-time scale and the “invariant
tori” problems are solved, a sufficient number of thermostats should still
be incorporated into the dynamical scheme to ensure ergodicity of a simple
polymer in the PIMD simulations. Therefore, the ergodicity for the config-
uration space sampling of the fictitious polymer in the PIMD simulation is
technically much trickier than a simple implementation of the equation of
motion generated from Eq. (7.21). To the best of our knowledge, this non-
ergodic sampling in direct implementations of the PIMD method was first
pointed out by Hall and Berne in 1984 [343], using water and liquid neon as
examples. To a certain extent, this is also why the MC methods are prefer-
entially used in the path integral molecular simulations in the 1980s [70–73].
An MD-based method, in which an efficient sampling over the configuration
space of the artificial polymer was guaranteed, was highly desired.
These difficulties had been largely solved in the late 1980s and the
early 1990s. In order to explain these efficient sampling methods in a clear
manner, we first rewrite the Hamiltonian in Eq. (7.21) in the following form:

H(x1 , . . . , xP , p1 , . . . , pP )
N 
 P
(pj )2
N 
 P
1  2 P
1
= i
+ mj ωP2 xji − xji−1 + V (x1i , . . . , xN
i ),
j=1 i=1 2Mij j=1 i=1
2 i=1
P
(7.26)

where the close-path feature x0 = xP is used to simplify the V (x1i , . . . , xN


i )
term, as compared to Eqs. (7.21) and (7.18). For a transparent nomencla-
ture in the following discussions, between the summation over beads and
nuclei, we take the sum over nuclei (N ) as the outer loop since the form
of coordinate transformation, which serves as the key step in solving this
ergodicity problem, does not depend on the nuclei. In addition to this sim-
plification of the potential energy term, since the first term in Eqs. (7.21)
and (7.26) is just introduced in the molecular dynamics simulations to
sample the configuration space, the mass Mij in this term, in principle,
is arbitrary for calculations of the statistical properties. It is allowed to
have one value for each artificial particle in the polymer (in total N × P
artificial particles), as long as the spring constant is physical. With this in
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 158

158 Computer Simulations of Molecules and Condensed Matter

mind, Eq. (7.26) can be further reformed into

H(x1 , . . . , xP , p1 , . . . , pP )
 P
N  N
 P
(pj )2 1 1 
= i
+ mj ωP2 xj Axj + V (x1i , . . . , xN
i ),
j=1 i=1 2Mij j=1
2 P i=1
(7.27)

where
⎛ ⎞
2 −1 0 0 ··· 0 0 0 −1
⎜ ⎟
⎜−1 2 −1 0 ··· 0 0 0 0 ⎟
⎜ ⎟
⎜ 0 −1 2 −1 · · · 0 0 0 0 ⎟
⎜ ⎟
⎜ . ⎟
A=⎜
⎜ ..
..
.
..
.
..
.
..
.
..
.
..
.
..
.
..
.
⎟.
⎟ (7.28)
⎜ ⎟
⎜ 0 0 0 0 · · · −1 2 −1 0 ⎟
⎜ ⎟
⎜ ⎟
⎝ 0 0 0 0 ··· 0 −1 2 −1 ⎠
−1 0 0 0 ··· 0 0 −1 2
Here, A is a P × P matrix. For the matrix multiplication term
(1/2)mj ωP2 xj Axj in Eq. (7.27), xj can be viewed as a P -dimensional vector
composed of (xj1 , xj2 , . . . , xjP ), with xji representing the position of the ith
image of the jth nucleus. When this position is a 3D vector in the Cartesian
space, this matrix multiplication term goes through the coordinate of x, y,
and z one by one. We note that Mij is the artificial mass we set for the ith
image of the jth nucleus, while mj is the physical mass of the jth nucleus.
From Eq. (7.27), it is clear that the smallest time step required in
the PIMD simulations to describe the motion of the polymer, which origi-
nates from the spring interaction between neighboring beads, is associated
with the largest eigenvalue of the matrix A. However, we note that an
artificial mass can be attributed to each fictitious particle in the polymer.
Based on this advantage which is intrinsic in the principles of the PIMD
method, a coordinate transformation can be employed to solve the infa-
mous multi-time scale problem originating from the interbead vibrations.
This coordinate transformation first decouples the harmonic interactions
between neighboring beads in Eq. (7.27). Then, the mass associated with
the different renormalized degrees of freedom (after coordinate transfor-
mation) can be artificially chosen so that the spring interaction results in
vibrations of the same frequency, and it is advisable to choose those masses
so that the resulting frequency does not depend on P . In doing so, the
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 159

Quantum Nuclear Effects 159

problems associated with the interimage vibrations can be avoided so that


an efficient sampling of the polymer’s configuration space can be carried
out in a much easier manner.
In order to carry out such a decoupling of the spring interactions
between neighboring beads, there are currently two popular schemes for
the PIMD simulations to follow. The first one is the so-called “staging”
method. It was originally proposed in a PIMC algorithm [71] and then
employed in the PIMD simulations [319]. In this method, taking the jth
nucleus as an example, the coordinate transformation from the original ones
(xj1 , xj2 , . . . , xjP ) to the transformed ones (uj1 , uj2 , . . . , ujP ) is

uj1 = xj1 ,
(7.29)
(i − 1)xji+1 + xj1
uji = xji − , for i  2,
i
whose form does not depend on the nuclear index j. The inverse of this
relation is

xj1 = uj1 ,
P
 (7.30)
(i − 1)
xji = uj1 + ujl , for i  2,
l−1
l=i

which can also be obtained recursively by

xj1 = uj1 ,
(7.31)
(i − 1) j 1
xji = uji + xi+1 + xj1 , for i  2.
i i

Since xjP +1 = xj1 , one often carries out the recursion in Eq. (7.31) in the
order xj1 , xjP , xjP −1 , . . . , xj2 .
One advantage of such transformed coordinates is that the interbead
spring interaction can be normalized through
P
 P
 i
(xji − xji+1 )2 = (uj )2 . (7.32)
i=1 i=2
i−1 i

To understand such a relationship from a practical perspective, we take the


P = 4 case as an example. The relationship between the original coordinates
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 160

160 Computer Simulations of Molecules and Condensed Matter

and the transformed ones in Eq. (7.29) indicates

uj1 = xj1 ,
3xj1 + xj1
uj4 = xj4 − = xj4 − xj1 ,
4
2xj + xj1 (7.33)
uj3 = xj3 − 4 ,
3
xj + xj1
uj2 = xj2 − 3 .
2

Putting this relation into 4i=2 i−1i
(uji )2 , one easily obtains
4
 i 3 4
(uj )2 = 2(uj2 )2 + (uj3 )2 + (uj4 )2
i=2
i−1 i 2 3
 2  2
j xj3 + xj1 3 j 2xj4 + xj1 4
= 2 x2 − + x3 − + (xj4 − xj1 )2
2 2 3 3
= 2(xj1 )2 + 2(xj2 )2 + 2(xj3 )2 + 2(xj1 )2 − 2xj1 xj2 − 2xj2 xj3
−2xj3 xj4 − 2xj4 xj1
4

= (xji − xji+1 )2 . (7.34)
i=1

Therefore, the coordinate transformation decouples the relative motion of


the beads in the position space.
Another point which is implied in Eqs. (7.29) and (7.30) is that there is
a one-to-one correspondence between (xj1 , xj2 , . . . , xjP ) and (uj1 , uj2 , . . . , ujP ).
Therefore, by setting
mj1 = 0,
i (7.35)
mji = mj , for i  2.
i−1
Equation (7.27) can be rewritten as

H(x1 , . . . , xP , p1 , . . . , pP )

= H(u1 , . . . , uP , P1 , . . . , PP )
N  P
  P
 (Pji )2 1 j 2 j 2  1
= + mi ωP (ui ) + V (u1i , . . . , uN
i ). (7.36)
2M j 2 P
j=1 i=1 i i=1
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 161

Quantum Nuclear Effects 161

Here, Pi is assumed to be conjugate to ui in the transformed coordinates,


j
and M  i is the associated mass which is related to the Mij in Eq. (7.26)
by the coordinate transition in Eq. (7.29). We do not need to care about
the specific form of this relation. There is no potential originating from the
harmonic spring interaction for the first artificial particle in the transformed
coordinate of the polymer (mj1 = 0). For the artificial particles associated
with the other degrees of freedom, this potential is determined by mji (i  2)
j
in Eq. (7.35). Therefore, as long as we keep M  i (i  2) as a constant multi-
ple of mji , the vibrations originating from the spring interactions associated
with the other renormalized degrees of freedom will share the same time
j j
scale. In practice, these M  i s are often chosen as M  1 = mj (physical mass
j j
of the jth nucleus) and M  i = mi for i  2. The equations of motion in the
transformed coordinates obtained from the Hamiltonian in Eq. (7.36) is
Pji
u̇ji = ,
M  ji
(7.37)
j 1 ∂V
Ṗi = −mji ωP2 uji − ,
P ∂uji
where the only term which needs to be obtained from calculations in the
original Cartesian space is ∂V /∂uji .
In order to obtain such forces in the transformed coordinate space,
one needs to go back to the original Cartesian space and calculate the
Hellmann–Feynman forces of the corresponding nuclear configuration,
namely, ∂V /∂xji . Then a transformation must be made to link these
∂V /∂xji s with the ∂V /∂uji s to be used for the simulation of the propagation
in Eq. (7.37). Such a transformation is defined from Eqs. (7.29) and (7.30).
The starting point is the following equation:
P
1 ∂V 1  ∂V ∂xji
= ,
P ∂ujk P i=1 ∂xji ∂ujk
(7.38)
P
1 ∂V 1  ∂V ∂uji
=
P ∂xjk P i=1 ∂uji ∂xjk

for a certain k. When k = 1, from Eq. (7.30), it is clear that ∂xji /∂ujk equals
one for any i. Therefore, using the first equation in Eq. (7.38), one easily
obtains
P
1 ∂V 1  ∂V
= . (7.39)
P ∂uj1 P i=1 ∂xji
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 162

162 Computer Simulations of Molecules and Condensed Matter

When k  2, from Eq. (7.29), it is clear that ∂uj1 /∂xjk = 0 and ∂uji /∂xjk = 0
for i  2. Therefore, the second equation in Eq. (7.38) becomes
P P
1 ∂V 1  ∂V ∂uji 1  ∂V ∂uji
= = . (7.40)
P ∂xjk P i=1 ∂uji ∂xjk P i=2 ∂uji ∂xjk

Then, we can make use of the second equation in Eq. (7.29) and rewrite it
into
P
1 ∂V 1  ∂V ∂uji
=
P ∂xjk P i=2 ∂uji xjk

P
 
1  ∂V ∂xji
j
i − 1 ∂xi+1
= −
P i=2 ∂uji ∂xjk i ∂xjk
P  
1  ∂V i−1
= δ i,k − δ i+1,k
P i=2 ∂uji i

1 ∂V 1 k − 2 ∂V
= j
− . (7.41)
P ∂uk P k − 1 ∂ujk−1

Therefore, the ∂V /∂ujk can be calculated recursively from


1 ∂V 1 ∂V 1 k − 2 ∂V
= + . (7.42)
P ∂ujk P ∂xjk P k − 1 ∂ujk−1

Combining Eqs. (7.39) and (7.42), the ∂V /∂ujk s to be used in the numerical
simulation of the propagation in Eq. (7.37) should be calculated recursively
from
P
1 ∂V 1  ∂V
= ,
P ∂uj1 P i=1 ∂xji
(7.43)
1 ∂V 1 ∂V 1 k − 2 ∂V
= + , for k  2.
P ∂ujk P ∂xjk P k − 1 ∂ujk−1
It is worth noting that, different from the recursion for the transformation
of the coordinates between xji and uji in Eq. (7.31), the force transformation
here starts from k = 2, after ∂V /∂uj1 is obtained using the first expression
in Eq. (7.43).
With these, the coordinate transformation which decouples the move-
ment of the neighboring images and imposes one single frequency for all
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 163

Quantum Nuclear Effects 163

these interbead vibrations has been introduced. The next thing one needs
to do, in order to carry out an ergodic sampling of the polymer naturally, is
a massive thermostat sampling for all the degrees of freedom for the artifi-
cial polymer in the transformed coordinates. This thermostat sampling can
be performed using the thermostats introduced in Chapter 5. Taking the
Andersen thermostat as an example, in this case, the momentum Pji for
1  i  P and 1  j  N should be rescaled for the desired temperature
during each collision between the artificial particle and the thermostat. In
the case when Nosé–Hoover chain thermostat of a certain length is used, the
equation of motion in Eq. (7.37) for the ith bead of the jth nucleus should
be linked with this Nosé–Hoover chain according to Eq. (5.52). One notes
that the computational cost of these thermostats is much lower than that of
the force calculation in real poly-atomic systems, especially when ab initio
methods for the electronic structures are used. Therefore, picking up a
thermostat which is efficient in pushing the polymer into its equilibrium
state is essential in practical PIMD simulations.
So far, we have discussed the “staging” transformation. As mentioned,
the coordinate transformation aims at decoupling the relative motion
between neighboring beads in the transformed coordinates. Therefore, as
an alternative to the “staging” method, the most direct way to fulfill such a
task is to diagonalize the matrix A in Eq. (7.28) and then use its eigenvec-
tors to perform the coordinate transformation. In doing so, the Hamiltonian
in Eq. (7.27) will have diagonal spring interactions in its second term in the
transformed coordinates, with the constant of these diagonal spring inter-
actions determined by the eigenvalues of the matrix A. This method is the
so-called “normal-mode” method in Refs. [81, 319]. Parallel to the “staging”
method, it is the other frequently used coordinate transformation method
in practical PIMD simulations.
Now, we go into the details of how such a “normal-mode” transforma-
tion is performed. Since the matrix A is the same for all the nuclei, a single
unitary orthogonal matrix U can be used to transform coordinates of all
nuclei from the Cartesian to the normal-mode ones. This transformation
reads
1 √
uj = √ Uxj and xj = P UT uj , (7.44)
P
where j goes through all the nuclear index from 1 to N . From our knowledge
of linear algebra, we know that the matrix U can also be used to diagonalize
the matrix A through P UAUT , and the resulting matrix Γ is diagonal. If
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 164

164 Computer Simulations of Molecules and Condensed Matter

we choose the number of P to be even (which equals 2n + 2) and align Γ as


⎛ ⎞
λ−n 0 0 0 0 0
⎜ ⎟
⎜ 0 λ−n+1 0 0 0 0 ⎟
⎜ ⎟
⎜ 0 0 ··· 0 0 0 ⎟
Γ=⎜

⎟,
⎟ (7.45)
⎜ 0 0 0 λ0 0 0 ⎟
⎜ ⎟
⎝ 0 0 0 0 ··· 0 ⎠
0 0 0 0 0 λn+1

the unitary orthogonal matrix U will have the following simple form:
⎧#  

⎪ 2/P sin 2πki
, −n  k < 0,

⎪ 2n+2


⎪#
⎨ 1/P , k = 0,
Uk,i = #   (7.46)

⎪ 2πki
0 < k  n,

⎪ 2/P cos 2n+2 ,



⎩ √
(−1)i / P , k = n + 1.

Here, the row index k goes from −n to n + 1 and the column index l goes
from 1 to P [344]. In other words, we label the beads from 1 to P in the
Cartesian coordinates and from −n to n + 1 in the normal-mode coor-
dinates. The diagonal elements of Γ are λ0 = 0, λ±i = 4P sin2 (iπ/P ) for
0 < i  n, and λn+1 = 4P . In the transformed “normal-mode” coordinates,
the Hamiltonian in Eq. (7.27) then reads

H(u−n , . . . , un+1 , P−n , . . . , Pn+1 )


N n+1
  n+1
  (pj )2 1 j
 1
= i
j + mj ωP2 λi (ui )2 + V (u1i , . . . , uN
i ).
2M  2 P
j=1 i=−n i i=−n
(7.47)
From the expression of the transformation matrix U in Eq. (7.46),
it is clear that in principle, it is equivalent to a Fourier transform of a
periodic path. For the eigenvector associated with the eigenvalue λ0 = 0,
from Eqs. (7.44) and (7.46), we see that the transformed coordinate
should be
P P
1  1  j
uj0 = √ U0,i xji = x ; (7.48)
P i=1 P i=1 i
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 165

Quantum Nuclear Effects 165

which represents the centroid of the path of the jth nucleus. Therefore, in
the normal-mode coordinate, the propagation of this specific mode natu-
rally describes the evolution of the centroid. Analogous to the “staging”
method, from the other eigenvalues of the matrix Γ and the fact that an
j
artificial mass M  i can be set for each artificial particle, one can easily
choose
M0j = mj ,
(7.49)
Mij = cλi mj , for − n  i  −1 and 1  i  n + 1,

with c representing a constant. With these, the different modes correspond-


ing to the interbead vibrations naturally move on the same time scale.
From the Hamiltonian in Eq. (7.48), the equation of motion is easily
obtainable from
Pji
u̇ji = ,
M  ji
(7.50)
j 1 ∂V
Ṗi = −mji ωP2 λi uji − .
P ∂uji

The next thing one needs to do is the same as the above descriptions for the
“staging” method, i.e. calculating ∂V /∂uji from ∂V /∂xji . Compared with
the “staging” method, here, due to the fact that the coordinate transfor-
mation is performed with a constant matrix determined by P only, from
Eqs. (7.38) and (7.44), this transformation between ∂V /∂uji from ∂V /∂xji
can be carried out in a much simpler manner.
In Eq. (7.44), we can see that the transformation between the Carte-
sian and the normal-mode coordinates can be carried out as follows. For
a specific nucleus, e.g. the jth, we go through the indices x, y, and z one
by one. For each index, e.g. the x index, the Cartesian coordinates of the
P images were organized as a P -dimensional vector. Then the coordinate
transformation to uj is carried out using Eq. (7.44). In the force transfor-
mations, we follow the same routine. For the jth nucleus, we go through the
indices x, y, and z one by one. For each index, e.g. x, the forces along the x
axis on the P images of the jth nucleus were organized as a P -dimensional
vector, which we label as F j . It is composed of
 $ $ 
∂V $$ ∂V $$
$ ,..., $ ,
∂xj1 $ ∂xjP $
x x
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 166

166 Computer Simulations of Molecules and Condensed Matter

$
∂V $
where $
∂xji x
means the Hellmann–Feynman force on the ith image of the
jth nucleus along the x axis. We label F j as the vector which represents
the transformed forces, composed of
 $ $ 
∂V $$ ∂V $$
$ ,..., $ .
∂uj−n $x ∂ujn+1 $x

Then, from Eqs. (7.38) and (7.44), one easily obtains


1 j 1
F = √ UT F j . (7.51)
P P
These forces are then used to propagate the equation of motion in
Eq. (7.50). Analogous to what we say in the “staging” method, massive
thermostats must again be used, with an efficient one imposed on each
degree of freedom. With these, an ergodic sampling can also be realized
using this “normal-mode” method. We note that due to the advantage
that the zeroth (according to our labeling from −n to n + 1) normal
mode naturally describes the propagation of the centroid, which does
have a physical meaning, the “normal-mode” method is getting more and
more popular nowadays, especially when extensions to real-time propaga-
tion are concerned. We will give a brief discussion to this extension in
Sec. 7.2.

7.1.5 Evaluation of the Zero-Point Energy

In the above discussions, we have introduced how the PIMD method should
be implemented. From the corresponding PIMD simulations, one can esti-
mate the statistical expectation value of a physical quantity at a finite
temperature using Eq. (7.25). The result of such an evaluation is the expec-
tation value of this quantity in the poly-atomic system at a finite T , with
the QNEs rigorously addressed on the same footing as the thermal ones.
From our discussions in Chapter 5, we know that the expectation value of
the same physical quantity can also be calculated in an MD simulation, with
only the thermal nuclear effects taken into account. Therefore, by compar-
ing the results obtained from these two simulations, one can evaluate the
impact of QNEs on this physical quantity in a very clear manner.
Besides these quantities which can be evaluated using Eq. (7.25),
e.g. the radial distribution function [323, 338, 345], the intra and inter-
molecular bond length distributions [286, 321], etc., there are also some
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 167

Quantum Nuclear Effects 167

physical quantities in which the evaluation of their expectation values from


the PIMD simulations contain some subtleties. The internal energy is such
an example. As a matter of fact, this internal energy at finite temper-
atures is among the most relevant quantities in molecular simulations.
A comparison between this quantity from the PIMD and MD simulations
at different temperatures and then an extrapolation of their differences
toward 0 K can give us the nuclear zero-point energy in a real poly-atomic
system, beyond the frequently used harmonic approximation. This zero-
point energy is of primary concern in studies of many problems, whose
value otherwise must be calculated from very expensive quantum MC sim-
ulations (normally with a force field treatment of the interatomic nuclear
interactions; see e.g. Refs. [346–349]). As mentioned before, for large poly-
atomic systems, such quantum MC simulations might not be applicable.
Therefore, the PIMD simulation, in principle, gives us a useful estima-
tor for this key physical quantity in practical calculations of poly-atomic
systems.
To understand how this purpose is fulfilled, we first go back to the
original definition of the internal energy in statistical mechanics, which is

1 ∂Z Q
E = − . (7.52)
Z Q ∂β

Here, Z Q is the partition function of the quantum poly-atomic system,


whose expression can be given in different ways, e.g. Eq. (7.20) in the limit
of P → ∞ and Eq. (7.17), etc. We take its expression in Eq. (7.20) and
rewrite it as
 
Z Q = lim C · · ·
P →∞
j » j – ff
PN PP (p )2 2 PP
−β j=1 i=1
i
j
2
+ 12 mj ωP (xji −xji−1 ) + 1
i=1 P V (x1i ,...,xN
i )
×e 2M
i

· dx1 · · · dxP dp1 · · · dpP


 
= lim C · · ·
P →∞
» j –
PN PP (p )2 2 P
j=1 i=1 −β i
j
1
−β P
22
mj (xji −xji−1 ) − P 1 1 N
i=1 β P V (xi ,...,xi )
×e 2M
i

· dx1 · · · dxP dp1 · · · dpP . (7.53)


December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 168

168 Computer Simulations of Molecules and Condensed Matter

Using this expression of the quantum partition function, an expansion of


Eq. (7.52) easily gives us
⎧  
  ⎨ N  P
C (pji )2 1 P j j 2
E = lim Q ··· − 2 2 mj (xi − xi−1 )
P →∞ Z ⎩
j=1 i=1 2Mij β 2

P
%
 1 1 N
+ V (xi , . . . , xi )
i=1
P
» j –
PN PP (p )2 1 2 P
−β j=1 i=1
i + m ω2
j 2 j P (xji−xji−1 ) − P 1 1 N
i=1 β P V (xi ,...,xi )
·e 2M
i

× dx1 · · · dxP dp1 · · · dpP


⎧  
  ⎨ N  P
C (pji )2 1 2 j j 2
= lim Q ··· − mj ωP (xi − xi−1 )
P →∞ Z ⎩
j=1 i=1 2Mij 2

P
%
 1 1 N
+ V (xi , . . . , xi )
i=1
P
» j –
PN PP (p )2 1 P
−β j=1 i=1
i + m ω2
j 2 j P (xji−xji−1 )2 − P 1 1 N
i=1 β P V (xi ,...,xi )
·e 2M
i

× dx1 · · · dxP dp1 · · · dpP .


(7.54)
Now, if we define
N  P
  P
 (pji )2 1 j j
 1
2 2
E= j − m j ω P (xi − xi−1 ) + V (x1i , . . . , xN
i ),
j=1 i=1 2M i
2 i=1
P
(7.55)

Eq. (7.54) can be further rewritten as


" #
j
PN PP (p )2 j j 2 PP
R R −β j=1 i=1
i
j +1
2
2
mj ωP (xi −xi−1 ) − i=1
1 V (x1 ,...,xN )
βP i i
· · · Ee 2M
i

· dx1 · · · dxP dp1 · · · dpP


E = lim "
j
# .
P →∞ PN PP (p )2 2 j j 2 PP 1 V (x1 ,...,xN )
R R −β j=1 i=1
i
j +1
2
mj ωP (xi −xi−1 ) − i=1 βP i i
··· e 2M
i

· dx1 · · · dxP dp1 · · · dpP .


(7.56)
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 169

Quantum Nuclear Effects 169

Therefore, the instantaneous quantity E as defined in Eq. (7.55) becomes


the quantity whose ensemble average should be evaluated in the PIMD
simulation. From Eq. (7.56), we see that this ensemble average gives the
internal energy of the poly-atomic system, in which all QNEs are included.
The minus sign in front of the second term in the quantity in Eq. (7.55) to be
averaged during the simulation originates from the temperature dependence
of the spring constant in the Hamiltonian of the fictitious polymer. In a
simpler form of the ensemble averages at a finite T in molecular dynamics,
the expectation value of the internal energy can be further reformed as
N P  P 
3N P  1  2  1
2 j j 1 N
E = − mj ωP xi − xi−1 + V (xi , . . . , xi ) .
2β j=1 i=1
2 i=1
P
(7.57)
From this estimator, it is easy to see that the kinetic energy (first term)
and the spring potential (second term) both scale linearly with the number
of beads P in a finite-temperature PIMD simulation. Therefore, with the
increase of P , these two quantities go to large values. Fortunately, due to
the minus sign in front of the second term, the difference between them
converges with P and the zero-point energy of a poly-atomic system can be
evaluated with Eq. (7.57) in practical PIMD simulations. However, we note
that large fluctuations on these two quantities still remain at large P . As a
consequence, for highly quantum systems in which a large P must be used
in order to arrive at an accurate simulation of the QNEs, a loss of precision
might exist [97, 350]. An alternative estimator, in which all terms involved
converge with P , is highly desired.
To circumvent this problem, a path integral version of the virial the-
orem was introduced by Herman et al. in 1982, where an estimator of the
internal energy which suffers much less from these fluctuations was pro-
posed [350, 351]. For a clear explanation of how this works, we first go back
to the original quantum partition function in Eq. (7.17), where no artificial
kinetic energy for the molecular dynamics simulations is introduced and
the integration goes only through the configuration space (composed of the
xji s) instead of the phase space (composed of the xji s and pji s). Using the
periodic boundary condition of the path integral sampling, this equation
can be rewritten as
 P2   
mP eff
Z Q = lim · · · e−βV (x1 ,x2 ,...,xP ) dx1 dx2 · · · dxP ,
P →∞ 2βπ2 V V V
(7.58)
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 170

170 Computer Simulations of Molecules and Condensed Matter

where

N 
P 
   P
1 1
V eff
(x1 , x2 , . . . , xP ) = mj ωP2 (xji − xji−1 )2 + V (x1i , . . . , xN
i ).
i=1 j=1
2 i=1
P
(7.59)
Similar to what we have used for the nomenclature before, (x1 , x2 , . . . , xP )
altogether represents the spatial configuration of the artificial polymer. xi
is composed of (x1i , . . . , xN
i ). It means the spatial configuration of the poly-
atomic system at its ith image. The key point of the virial internal energy
estimator is that the first two terms in Eq. (7.57) is replaced by the mean
kinetic energy of the quantum system as

N P  N P

3N P  1 1   j ∂V (x1i , . . . , xN
2 j j 2 i )
− mj ωP (xi − xi−1 ) = x · .
2β j=1 i=1
2 2P j=1 i=1 i ∂xji
(7.60)
With this treatment, the expectation value of the internal energy in
Eq. (7.57) reforms into

N P
 P

1   j ∂V (x1i , . . . , xN
i )
 1 1 N
E = x · + V (xi , . . . , xi ) .
2P j=1 i=1 i ∂xji i=1
P
(7.61)
Here, you can see that neither of the two terms being evaluated in the PIMD
simulation scales with P . Therefore, a smaller fluctuation of the internal
energy to be evaluated in the PIMD simulations should be expected.
To understand how the equality in Eq. (7.60) exists, we use the parti-
tion function in Eq. (7.58) to evaluate the quantity on the right-hand side
of Eq. (7.60). From this partition function, the ensemble average of this
quantity equals

N P

1   j ∂V (x1i , . . . , xN
i )
xi · j
2P j=1 i=1 ∂xi
   1 N P j ∂ P1 V (x1i ,...,xN
i )

· · · x · (7.62)
V V 2 j=1 i=1 i ∂x j
i

eff
e−βV (x1 ,x2 ,...,xP ) dx1 dx2 · · · dxP
=  
−βV eff (x1 ,x2 ,...,xP ) dx dx · · · dx
.
V ··· V e 1 2 P
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 171

Quantum Nuclear Effects 171

Here, the relation between V (x1i , . . . , xN


i )/P and V
eff
(x1 , . . . , xP ) is given
by Eq. (7.59). Now, if we label

P 
 N
1
α(x1 , x2 , . . . , xP ) = mj ωP2 (xji − xji−1 )2 , (7.63)
i=1 j=1
2

and

P
1
λ(x1 , x2 , . . . , xP ) = V (x1i , x2i , . . . , xN
i ), (7.64)
i=1
P

then Eq. (7.59) will be reformed into

V eff (x1 , x2 , . . . , xP ) = α(x1 , x2 , . . . , xP ) + λ(x1 , x2 , . . . , xP ), (7.65)

and Eq. (7.62) reforms to


N P

1   j ∂V (x1i , . . . , xN
i )
x ·
2P j=1 i=1 i ∂xi j

   N P eff

V
· · · V 12 j=1 i=1 xji · ∂V (x 1 ,...,xP )
∂xj i

−βV eff (x1 ,x2 ,...,xP )


e dx1 dx2 · · · dxP
=   (7.66)
V
··· V
e−βV eff (x1 ,x2 ,...,xP ) dx 1 dx2 · · · dxP
  1 N P j ∂α(x1 ,...,xP ) 
V
· · · V 2 j=1 i=1 xi · ∂xj i

−βV eff (x1 ,x2 ,...,xP )


e dx1 dx2 · · · dxP
−  .
V ··· V e−βV eff (x1 ,x2 ,...,xP ) dx 1 dx2 · · · dxP

For a further evaluation of this quantity, we first make use of an important


property of α(x1 , . . . , xP ), that it is a homogeneous function of (x1 , . . . , xP )
of degree 2, so that the following equation exists:

N 
 P
∂α(x1 , . . . , xP )
xji · = 2α(x1 , . . . , xP ). (7.67)
j=1 i=1 ∂xji
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 172

172 Computer Simulations of Molecules and Condensed Matter

Therefore, the second term on the right-hand side of Eq. (7.66) equals

   1 N P 
V ··· V 2 j=1 i=1 xji · ∂α(x1 ,...,xP )
∂xji
eff
e−βV (x1 ,x2 ,...,xP ) dx1 dx2 · · · dxP
− 
V
· · · V e−βV eff (x1 ,x2 ,...,xP ) dx1 dx2 · · · dxP
  eff

V · · · V α(x1 , . . . , xP )e−βV (x1 ,x2 ,...,xP ) dx1 dx2 · · · dxP


=− 
−βV eff (x1 ,x2 ,...,xP ) dx dx · · · dx
V ··· V e 1 2 P

= α(x1 , . . . , xP )
N P 
 1 j j
2 2
=− mj ωP (xi − xi−1 ) , (7.68)
j=1 i=1
2

and Eq. (7.66) further evolves to

N P

1   j ∂V (x1i , . . . , xN
i )
xi · j
2P j=1 i=1 ∂xi
  1 N P j ∂V eff (x1 ,...,xP )
V
· · · V 2 j=1 i=1 xi · ∂xj i

eff (7.69)
e−βV (x1 ,x2 ,...,xP ) dx1 dx2 · · · dxP
= 
−βV eff (x1 ,x2 ,...,xP ) dx dx · · · dx
V ··· V e 1 2 P
N P 
 1
− mj ωP2 (xji − xji−1 )2 .
j=1 i=1
2

Comparing Eq. (7.69) with Eq. (7.60), the only equality we need to prove
becomes

   1 N P j ∂V eff (x1 ,...,xP )



V
··· V 2 j=1 i=1 xi · ∂xji
eff
e−βV (x1 ,x2 ,...,xP ) dx1 dx2 · · · dxP
 
−βV eff (x1 ,x2 ,...,xP ) dx dx · · · dx
V ··· V e 1 2 P

3N P
= . (7.70)

December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 173

Quantum Nuclear Effects 173

This proof is doable if we reform the left-hand side of Eq. (7.70) in the
following manner:
   1 N P j ∂V eff (x1 ,...,xP )

V
··· V 2 j=1 i=1 xi · ∂xji
eff
e−βV (x1 ,x2 ,...,xP ) dx1 dx2 · · · dxP
 
V
· · · V e−βV eff (x1 ,x2 ,...,xP ) dx1 dx2 · · · dxP
  (7.71)
  1 N P j ∂ −βV eff (x1 ,x2 ,...,xP )
V
· · · V 2 j=1 x
i=1 i · ∂xj e
i

1 dx1 dx2 · · · dxP


=−   .
β V
· · · V e−βV eff (x1 ,x2 ,...,xP ) dx1 dx2 · · · dxP

Then, making use of an integration by parts, this quantity further evolves


into
  1 N P  
j ∂ −βV eff (x1 ,x2 ,...,xP )
1 V · · · V 2 j=1 x
i=1 i · ∂xij e dx1 dx2 · · · dxP
−   eff
β V
· · · V e−βV (x1 ,x2 ,...,xP ) dx1 dx2 · · · dxP
  1 N P −βV eff (x ,x ,...,x )  ∂ j

V
· · · V 2 j=1 i=1 e 1 2 P
∂x j · xi
i

1 dx1 dx2 · · · dxP


=  
β V
· · · V e−βV eff (x1 ,x2 ,...,xP ) dx1 dx2 · · · dxP
1 3N P
= . (7.72)
β 2

With these, the equality in Eq. (7.60) is proven and one can use the
virial estimator as given in Eq. (7.61) to calculate the finite-temperature
internal energy of the poly-atomic system under investigation. In practice,
a slight variation of Eq. (7.61) is often used. This variation is based on the
following equation:
N P

1   j ∂V (x1i , . . . , xN
i )
xi · j
2P j=1 i=1 ∂xi
N P

1   j ∂V (x1i , . . . , xN
i )
= xc · j
2P j=1 i=1 ∂xi
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 174

174 Computer Simulations of Molecules and Condensed Matter

N P

1  j ∂V (x1
, . . . , xN
)
+ (x − xjc ) · i i
2P j=1 i=1 i ∂xji
N  N P

1 j 1  ∂V (x 1
, . . . , xN
)
= x · Fjc + (xj − xjc ) · i i
2 j=1 c 2P j=1 i=1 i ∂xji
N P

3N 1  j ∂V (x1i , . . . , xN
i )
= + (xi − xjc ) · j , (7.73)
2β 2P j=1 i=1 ∂xi

where Fjc stands for the effective force imposed on the centroid of the jth
atom and xjc stands for its centroid position, and the estimator in Eq. (7.61)
further changes into
N P

3N 1  j j ∂V (x1i , . . . , xN
i )
E = + (x − xc ) ·
2β 2P j=1 i=1 i ∂xji
P 
 1 & '
1 N
+ V xi , . . . , xi . (7.74)
i=1
P

One can use either the estimator in Eq. (7.74) or the one in Eq. (7.57) to
evaluate the internal energy in real poly-atomic systems.
We note that the internal energy calculated this way includes contribu-
tion from the QNEs. As mentioned above, an MD simulation, in which only
the thermal nuclear effects are included, can also give us an expectation
value of this quantity. Therefore, a comparison between results obtained
from these two simulations in principle can give us the zero-point energy of a
real poly-atomic system. To understand how this works in practice, we show
a sketch for the evolution of the internal energy in the MD and PIMD (using
different P ) simulations as a function of temperature in Fig. 7.4. At 0 K,
the internal energy equals the static geometry-optimized total energy in the
MD simulation. With the increase in the temperature, it increases linearly
due to the classical virial theorem. In the PIMD simulations, this internal
energy evolves differently from the one obtained from the MD simulations,
and this difference originates from the QNEs. Its value increases with the
number of beads P till convergence. At higher T s, a small value of P is
good energy to describe this difference. At lower T s, larger P is needed. At
0 K, since an infinite P is needed for the path integral sampling, the PIMD
simulation loses its precision too. However, an extrapolation of the QNEs
from finite T still presents a good estimator for the zero-point energy.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 175

Quantum Nuclear Effects 175

Figure 7.4 Illustration of how the internal energy’s expectation value changes with
temperature in the MD and PIMD simulations. In the MD simulation, as T approaches
0 K, the internal energy approaches the static geometry-optimized total energy E0 . With
the increase in T , it increases linearly because of thermal fluctuations. In the PIMD sim-
ulations, since the QNEs are included, at finite T , there is a difference in this value from
its classical result. This difference increases with the number of beads P till convergence.
At higher temperatures, a small P is already good energy to describe this difference. At
lower temperatures, larger P s are needed. At 0 K, due to the fact that an infinite P is
needed for the path integral sampling, the PIMD method also loses its power. However,
an extrapolation of the difference between the internal energy obtained from the con-
verged PIMD simulations and MD simulations at finite T (indicated by the dotted line)
still presents a good estimator for the zero-point energy.

7.2 Extensions Beyond the Statistical Studies


So far, all our discussions have been restricted to the statistics. The associ-
ated time-averaged quantities can be used to study the impact of the QNEs
on the equilibrium statistical properties of a poly-atomic system under
investigation at finite temperatures. Another aspect of the real quantum
world, i.e. the dynamics, however, has never been touched. We note that
descriptions of such nuclear dynamics, with relevant electronic structures
computed accurately “on-the-fly”, poses a “grand challenge” to both theo-
retical physics and chemistry. As a matter of fact, illustrations of many key
physical/chemical properties in the real world, e.g. the transport properties,
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 176

176 Computer Simulations of Molecules and Condensed Matter

the chemical reaction rates, and the neutron or light scattering spectra,
etc., require their descriptions. In a many-body (poly-atomic) system, we
know that the key quantity in describing such dynamics is the so-called
time-correlation function. Taking the infrared absorption spectrum as an
example, it is directly related to the dipole–dipole time-correlation function
of the system. The translational diffusion coefficient, on the other hand, can
be understood using the velocity autocorrelation function, etc. Because of
these, inclusion of the QNEs in descriptions of such time-correlation func-
tions becomes an issue of considerable interest in present studies in both
theoretical physics and chemistry [74, 79, 91, 352–357].
A natural choice for the calculation of such time-correlation functions is
to solve the time-dependent Schrödinger equation of the nuclei in real poly-
atomic systems. This implies propagating the nuclear quantum dynamics
on the Born–Oppenheimer potential energy surfaces (BO-PES) which are
precomputed with very accurate electronic structure theories. The multi-
configuration time-dependent Hartree method (MCTDH) is such an exam-
ple [358, 359]. In the past years, it has been very successful in describing
some gas phase chemical reactions and the dynamical properties of small
molecules [60, 61, 360–362]. However, one notes that the scaling of their
computational cost with system size makes it unfeasible to many practi-
cal systems quickly as the nuclear degree of freedom increases. Alternative
methods in describing such nuclear dynamics, where scaling behavior is
much better so that simulations can be performed in systems of relevance
in practical studies, must be used.

7.2.1 Different Semiclassical Dynamical Methods

Similar to the route we have chosen in studies of the statistical mechanics,


we now resort to the path integral representation of the quantum mechanics.
Within this picture, the most rigorous method in quantifying such a time-
correlation function naturally involves using a complex-time path integral
sampling technique [67, 74]. In this method, the thermal effects are treated
as imaginary time in the complex-time space and the real-time axis takes
care of the dynamics. Extension of the path integral equations from statis-
tics to dynamics is straightforward [74]. However, different from the suc-
cess of PIMD/PIMC techniques in addressing the statistical properties in
simulations of real poly-atomic systems, practical simulations based on such
extension of the path integral sampling methods to the complex-time space
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 177

Quantum Nuclear Effects 177

has continued to be problematic due to an extensive phase cancellation orig-


inating from the paths with weights that are non-positive in character [78,
79]. In spite of these difficulties, by defining a symmetrized time-correlation
function that lends itself to PIMD/PIMC simulations, significant progress
has still been made over the last 30 years within this complex-time path
integral scheme [75–77]. As a prominent example, in a recent development
of this method by Nakayama and Makri for studies of the subcritical liquid
para-hydrogen, the authors have shown that accurate quantum mechanical
results for the initial 0.2 ps segment of the symmetrized velocity autocorre-
lation function, as well as the incoherent dynamic structure factor at certain
momentum transfer values at moderate temperatures and densities, can be
obtained [352]. But for more general problems involving longer time dynam-
ics and more complicated systems with higher density, which is clearly of
more practical use, this method becomes less practical and one needs to
resort to less accurate quantum dynamical methods.
Following the summary by Braams and Manolopoulos [91], here we
categorize these less accurate quantum dynamical methods that have been
applied to condensed matter systems, essentially into three classes, not-
ing that each method has its own strengths and weaknesses. The first
class of methods simulate the system using an imaginary-time propagator
only. Then, an inverse Wick rotation is used to infer the thermodynami-
cally averaged real-time-correlation function from the imaginary-time one
based on the Baym–Mermin theorem [363]. This trick is similar to the
analytical continuation of the self-energy from the imaginary frequency
axis to the real frequency one for the calculation of the quasi-particle
energies as introduced in Sec. 4.4.5. There are two practical schemes for
the time-correlation function to be calculated, i.e. the numerical analytic
continuation (NAC) [75–77, 364–367] and the quantum mode-coupling the-
ory (QMCT) [368–373]. The advantage of these methods is that the short-
time behavior of the time-correlation function can normally be described
accurately and the imaginary-time treatment makes the correlation function
easy to compute. However, we note that the numerical stability of the ana-
lytical continuation is much worse than its counterparts in the calculation
of the self-energies. Some standard methods, such as the Padé technique,
are of limited use in practice [374, 375], and these methods are not exact
at the classical limit.
The second class of methods combine an exact treatment of the
quantum Boltzmann operator with an approximate treatment of the
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 178

178 Computer Simulations of Molecules and Condensed Matter

real-time evolution based on classical dynamics. These methods include the


linearized semiclassical-initial value representation (LSC-IVR) method (see
e.g. Refs. [355, 356, 376–379], etc.), the Feynman–Kleinert linearized path-
integral (FK-LPI) method (see e.g. Refs. [380–383], etc.), and the forward–
backward semiclassical dynamics (FB–SD) (see e.g. Refs. [353, 354, 357],
etc.). Their main advantage is that they are exact in three important limits,
i.e., the short-time limit, the limit of a harmonic potential, and the classi-
cal limit, while the main disadvantage is that the classical trajectories do
not, in general, conserve the quantum mechanical equilibrium distribution
functions [91].
The third class of methods, which is currently of more practical use in
condensed matter, includes the so-called CMD by Cao and Voth [80–84, 384]
and the RPMD by Manolopoulos and his coworkers [85–93]. A key differ-
ence between these two methods and those in the earlier two classes is that
mathematically there are only slight modifications in the standard PIMD
method as we have introduced in Sec. 7.1, although conceptually these dif-
ferences are fundamental, and they all originate from generalizations of the
physical concepts implied in the PIMD simulations, with the CMD method
that appears earlier. Therefore, in the following, we will give a detailed
explanation of these two methods in a chronological order with a special
focus on their similarities and differences with the normal statistical PIMD
simulations, starting from a precursor of the CMD method, i.e. Gillan’s
generalization of Feynman’s path centroid concept to its applications in
the TST.

7.2.2 Centroid Molecular Dynamics and Ring-Polymer


Molecular Dynamics

In the seminal book by Feynman and Hibbs [67], it has been shown that the
concept of path centroid can be used to interpret the impact of quantum
effects on the effective potential the particle under investigation feels. Later,
this concept was extended by Feynman and Kleinert so that we have an
effective centroid potential [385], which is of practical use in molecular
simulations. As one example, Gillan has employed it in calculating the
probability of finding a quantum particle at its transition state during a
rare-event transition process, which could be used later to describe the
transition rate between two stable states at the quantum mechanical level,
see e.g. Refs. [386, 387]. We note that theories behind characterizing such
transition rates of slow processes, such as the chemical reactions and the
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 179

Quantum Nuclear Effects 179

diffusion events, are the so-called TSTs. It is both statistical and dynamical
in the sense that a statistical property (the probability of finding the system
at its TS) is assumed to be proportional to a dynamical property (the TS).
Due to this fundamental assumption, theories behind characterizing this
probability of finding the system in its transition state lie at the heart of
the TST.
Already at the static level, searching for this TS is non-trivial in a
poly-atomic system, due to the high-dimensional feature of the BO PES
associated with the nuclear degrees of freedom. Currently, there are sev-
eral schemes in which such a hunting can be carried out, including the
constrained optimization (CO) [19], nudged elastic band (NEB) [20–22],
Dewar, Healy and Stewart (DHS) [23], Dimer [24–26], activation–relaxation
technique (ART) [27, 28] and one-side growing string (OGS) [29] as well as
their various combinations. Here, we suppose that a reasonable estimation
of the TS and its associated hyperplane separating the reactant and product
states is already carried out. Our discussions only concern further thermal
and quantum nuclear effects on the static energies based on this knowledge.
At the classical level, suppose that we have a well-defined reaction
coordinate s, which properly separates the reactant and product states.
According to the TST, the transition rate is
v⊥  Z T δ v⊥  Z T
kTST = = , (7.75)
2δ Z 2 Z
where δ is defined as the interval for the reaction coordinate s with which we
think that the system is at the TS. With this definition, Z T δ/Z stands for
the probability of finding the system at the TS while v⊥ /(2δ) stands for
the escape rate from the TS. Putting these two factors together, one obtains
a transition rate through the TS from the reactant to the product, which
is sensitive to the choice of the “TS”. One notes that this estimation of
the transition rate with the TST in Eq. (7.75) sets the upper bound for the
real transition rate [331], since moving the “TS” used here away from the
real one will clearly increase Z T /Z. Therefore, it is highly recommended
that one searches for the TS and its associated hyperplane dividing the
reactant and product states so that the estimated rate in Eq. (7.75) reaches
a minimum. In doing so, one obtains the best evaluation of the transition
rate. Such a variational method is called variational transition-state theory
(VTST) [331, 388, 389].
Then one tries to include the QNEs. There are several extensions of
this TST to its quantum version [390–395], with the simplest ones only
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 180

180 Computer Simulations of Molecules and Condensed Matter

replacing the classical statistical averaging with the quantum one. In these
simplest methods, we note that the extension of the centroid’s concept from
the traditional statistical PIMD method is already used in calculating the
free energy difference between the reactant and the TSs. One prominent
example is the work by Gillan in the 1980s [386, 387]. In this example, the
probability of finding the system in the TS compared to that in the reactant
was calculated using the reversible work. In particular, a series of locations
along the transition path were selected and a PIMD simulation is carried out
by fixing the centroid of the path integral chain of the transition particle
at each point. In doing so, the average over all the remaining quantum
degrees of freedom at this point can be properly evaluated and one obtains
an effective mean force on the centroid of the transition particle along
the transition path. By integrating this effective force along the transition
path using reversible work, one gets the free energy difference between the
reactant and the TSs. The relative density of the centroid at this TS is
then calculated using this free energy difference and by assuming that the
transition rate is proportional to this relative density, one finally gets the
transition rate.
Following the idea that effective forces on the centroids of the quan-
tum particles can be calculated by fixing the centroid positions and doing
statistics over all other quantum degrees of freedom, the CMD method was
introduced by Cao and Voth in 1993 as an approximate method to compute
the real-time quantum correlation function for the dynamical properties of
a real poly-atomic system to be described [384]. The central point of this
CMD method is the assumption that the real-time evolution of the centroid
positions on their potential of mean force (PMF) surface can be used to
generate approximate quantum dynamical properties of real poly-atomic
systems. Their evolution respects the Newton equations:
pjc
ẋjc = ,
m (7.76)
j j
ṗc = F (xc ),
where j again runs through the atomic index, xc represents a spatial config-
uration of the centroids and xjc means the centroid position of the jth atom.
F j (xc ) is the derivative of potential of mean force surface with respect to
xjc , i.e. the mean field centroid force at xc . It is mathematically defined as

dx1 · · · dxP δ(x0 − xc )F0j (xc )e−βV (x1 ,...,xP )
eff
j 
F (x0 ) = . (7.77)
dx1 · · · dxP δ(x0 − xc )e−βV eff (x1 ,...,xP )
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 181

Quantum Nuclear Effects 181

Here, V eff (x1 , . . . , xP ) is the effective potential defined in Eq. (7.18) and


F0j (xc ) is the instantaneous Hellmann–Feynman force imposed on the cen-
troid, given by
P
1  ∂V (x1 , . . . , xP )
F0j (xc ) = , (7.78)
P i=1 ∂xji

and x0 represents the instantaneous centroid configuration to which xc


should be restricted. Due to the use of the Newton equations in describ-
ing the centroid propagation, this method is intrinsically a semiclassical
method, with the QNEs rigorously described only at the statistical level
when the mean field centroid force is calculated. However, we note that the
quantum correction to the effective potential sometimes already incorpo-
rates the dominant elements of QNEs in descriptions of the nuclear dynam-
ics and it is currently used as a standard routine to investigate the impact
of QNEs on the dynamical property of condensed matter systems.
One point implied in the procedure described above for the CMD
method is that a fully statistical PIMD (or PIMC) simulation should be
carried out at each centroid configuration before it propagates to the next
centroid configuration. For complex poly-atomic systems, however, this is
inapplicable due to the computational cost associated. As a simplified ver-
sion of this method, the adiabatic approximation can be used [396]. For a
better explanation of this idea, we go back to the normal-mode coordinate as
explained in Sec. 7.1.4. Mathematically, the fundamental difference between
a PIMD simulation in the primitive Cartesian coordinate and the normal-
mode coordinate is that a coordinate transformation should be made at
each PIMD step in order to convert the forces and the spatial configura-
tions of the polymer, so that the forces can be calculated in the Cartesian
space and the equation of motion can be propagated in the normal-mode
one. The first normal mode describes the propagation of the centroid, while
the other modes describe interbead vibrations. When the masses used in
the PIMD simulation are chosen according to Eq. (7.49), one ensures that
all the “artificial” interbead vibrations have the same frequency, whose
value is determined by the constant c. Frequency associated with the cen-
troid vibration is determined by the interatomic potential which is real
and physical. Therefore, intuitively, one can set this constant c to a very
small value so that the masses associated with the “artificial” interbead
vibrations are small and they can adiabatically react to the motion of the
centroid. Since the centroid mode moves much slower than the other ones,
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 182

182 Computer Simulations of Molecules and Condensed Matter

during a characteristic time for its vibration which is determined by the


interatomic potential, the other interbead vibrational modes can already
perform a very good statistical averaging over their degrees of freedom. In
doing so, the potential each centroid feels upon characterizing the physical
vibrations can be calculated “on-the-fly”. As a cost of not doing a PIMD
simulation at each centroid configuration, a much smaller time step should
be used in order to address the fast interbead vibrations originating from
the small masses associated with them.
We note that this “on-the-fly” simplified calculation of the mean field
centroid force was first proposed in Ref. [83], where different time inter-
vals were suggested for the propagation of the centroid and the much
faster interbead motions. In between the centroid propagation time steps, a
series of interbead propagations should be made, subject to the constraint
in Eq. (7.77) so that a statistical averaging over the centroid force can
be obtained before propagating the centroid. Therefore, different from a
rigorous implementation of the CMD method, there is only one trajectory.
This simplified version of the CMD method is called the adiabatic centroid
molecular dynamics (ACMD) method [83]. In 2006, it is further simplified
so that a single small time interval is used for the propagation of both the
centroid and interbead vibrations [397] and the corresponding simulation is
called partially adiabatic centroid molecular dynamics (PACMD) simula-
tions [397]. We note that nowadays this PACMD method is the frequently
used method in practical simulations of complex systems, and in practice,
this distinction between PACMD, ACMD, and CMD is often obviated and
one simply refers to PACMD as CMD [398]. Here, we follow such a tradition
and refer to PACMD as CMD in later discussions.
Now, we look at the differences between the statistical PIMD and the
dynamical CMD simulations. For this comparison to be as simple as possi-
ble, we use the normal-mode coordinate for the PIMD simulations. In sta-
tistical normal-mode PIMD simulations, the constant c in Eq. (7.49) is set
as one and the time step is determined by the frequency of real interatomic
vibrations. Instead, in a CMD simulation, since the interbead vibrations
have much higher frequency due to their small artificial masses, this c takes
a small value between zero and one, and a much smaller time step than
that of the interatomic vibrations should be used to ensure the adiabatic
approximation works. For the thermostating strategy, each mode should be
coupled to an efficient thermostat in both cases. Therefore, mathematically
these two simulations are very similar, although conceptually they are
fundamentally different.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 183

Quantum Nuclear Effects 183

Then we compare the CMD method with the more recently proposed
RPMD [85–93]. The differences are mainly located in three aspects. First,
the RPMD method chooses the kinetic mass Mij in Eq. (7.26) as Mij =
mj /P , where mj is the mass of the jth nucleus, if the dynamics is done
at real temperature T . Such a setting ensures that the mass of each bead
associated with its potential V (x1i , . . . , xN
i )/P in the case of Eq. (7.26) gives
the physical interatomic vibration frequency when interbead interactions
are neglected. If the dynamics is performed at P T , then the potential part
in Eq. (7.26) will be
P 
N 
 
1
P mj ωP2 (xji − xji−1 )2 + V (x1i , . . . , xN
i ) , (7.79)
j=1 i=1
2

and the mass will be chosen as Mij = mj .


The second difference between the CMD and RPMD methods is that
in CMD, the centroid dynamics is used to calculate the time-correlation
function, while in the RPMD method, the dynamics in each image is cal-
culated separately and then the time-correlation function for the whole
system is an average over all images. Because of this difference, in the
RPMD method, the interbead vibrations are also accounted for when the
vibrational spectrum of the quantum system is calculated, and these arti-
ficial vibrational frequencies are evenly distributed on the frequency axis,
which often pollute the real physical vibrational frequencies associated with
the interatomic motion. In the CMD method, on the other hand, the time-
correlation function is calculated from the propagation of the centroid. In
doing so, the vibration of the centroid will not be polluted by the inter-
bead vibrations. However, at low temperatures, taking the OH stretching
mode as an example, the centroid of the H atom often falls much closer
to the oxygen atom compared with its physical value within each image.
This induces the so-called “curvature problem” in CMD simulations, which
artificially softens the covalent bond stretching frequencies [399].
The third difference between the CMD and RPMD methods is that
the RPMD method needs to be carried out in a Hamiltonian manner when
the time-correlation function is calculated. In other words, no thermostat
should be added when the trajectory under construction will be used in the
calculation of the time-correlation function. The temperature effect should
be included during the thermal equilibrium process when the canonical
distribution of the snapshots starting from which the micro-canonical sim-
ulations are carried out is generated. In the CMD method, on the other
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 184

184 Computer Simulations of Molecules and Condensed Matter

hand, the canonical ensemble is used for the single trajectory to be gener-
ated. Because of this difference, although RPMD does not need a very small
time step, many trajectories are needed in order for the thermal averaging
on the time-correlation function to be sufficiently sampled.
In recent years, there are several studies aiming at comparing the per-
formance of CMD and RPMD in some models and real poly-atomic sys-
tems [91, 397, 398]. In Ref. [91], Braams and Manolopoulos showed that
the Kubo-transformed autocorrelation functions obtained from the RPMD
simulations are accurate on the time scale up to the sixth order for the
position and the fourth for the velocity; that of the CMD method leads to
an accuracy of the fourth order and second order for these two quantities,
respectively. Hone et al., on the other hand, showed results clearly in favor
of CMD [397], where simulations on para-hydrogen demonstrates that the
CMD method gives better agreement with experiments. Later, Perez et al.
pointed out that when such a comparison is made, the differences in the
setting of the simulations as mentioned above must be kept in mind [398].
Furthermore, in complex systems when the accuracy of the inter-atomic
potential is unclear, comparison with experimental results cannot be used
to judge which one is more accurate, since it is impossible to discern how
much of the discrepancy with experiment is due to the accuracy of quantum
dynamics and how much is due to the interatomic potential. Rather, an
alternative method for such a comparison should be used. In this paper, it
is suggested that one uses the same numerical treatment to infer the time-
correlation function from the real-time axis as obtained from the CMD and
RPMD methods to the imaginary-time axis. Then, these results can be
compared with the numerically exact results from imaginary-time PIMD
or PIMC simulations. In doing so, the performance of these two methods
on quantum dynamics is compared solely. We highly recommend such a
choice of criterion for future studies in this direction.

7.3 Free Energy with Anharmonic QNEs


In the previous chapter, we have introduced the thermodynamic integra-
tion method. Using this method, the anharmonic contribution from the
nuclear thermal fluctuations to the free energy can be calculated in real
poly-atomic systems, as long as a well-defined reference state exists. The
QNEs, however, stay on the level of the harmonic approximation. In real-
ity, we know that these QNEs also have anharmonic contributions to their
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 185

Quantum Nuclear Effects 185

vibrations/rotations and consequently the free energy. To include such


effects, one needs to extend the thermodynamic integration method as
introduced in Chapter 6 so that the QNEs on the free energy beyond the
harmonic approximation are also accounted for. In practice, this can be
done through an extension of the thermodynamic integration method in
the framework of PIMD/PIMC, as will be introduced below.
The starting point for this discussion is the quantum partition function
for the real poly-atomic system, defined as
⎡ ⎤
N 
  P2
mj P
ZQ = lim ZP = lim ⎣ 2

P →∞ P →∞ 2βπ
j=1
  
eff
· ··· e−βV (x1 ,x2 ,...,xP )
dx1 dx2 · · · dxP , (7.80)
V V V

where

N 
P 
   P
1 1
V eff (x1 , x2 , . . . , xP ) = mj ωP2 (xji − xji−1 )2 + V (x1i , . . . , xN
i ).
i=1 j=1
2 i=1
P
(7.81)
It is clear from our earlier discussion that the free energy associated with
this partition function is the free energy of the quantum poly-atomic system,
given by

1
FQ = − ln Z Q . (7.82)
β

Now, we look at the effective potential in Eq. (7.82), we note that we


can replace the second term in it by the effective potential on the centroid
as

N 
P 
   P
1 1
V eff (x1 , x2 , . . . , xP ) = mj ωP2 (xji − xji−1 )2 + V (x1c , . . . , xN
c ).
i=1 j=1
2 i=1
P
(7.83)
Here, xjc means the centroid position of the jth nucleus, which does not
depend on the bead index i. By inputting this equation into Eq. (7.80), we
can see that the partition function Z becomes
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 186

186 Computer Simulations of Molecules and Condensed Matter

⎡ ⎤
N 
  P2   PN
mj P 1 N
Z C = lim ⎣ 2
⎦ · · · e−β j=1 V (xc ,...,xc )
P →∞ 2βπ V V
j=1

PN PP 2
· e−β j=1
1 2
i=1 2 mj ωP (xji −xji−1 ) dx1 · · · dxP . (7.84)

At this time, if we resort to the normal-mode coordinate as defined in


Eqs. (7.47) and (7.48), Eq. (7.84) can be rewritten as

⎡ ⎤
N  P2 
m j P 1 N
Z C = lim ⎣ 2
⎦ du0 e−βV (u0 ,...,u0 )
P →∞ 2βπ V
j=1


−1 n+1
 PN 1 2 j 2
· e−β j=1 2 mj λi ωP (ui ) dui
i=−n i=1 V
⎡ ⎤
N  P2 
mj P 1 N
= lim ⎣ 2
⎦ du0 e−βV (u0 ,...,u0 )
P →∞ 2βπ V
j=1

 
−1 n+1
− 12
PN mj P
(2 sin iπ 2 j 2
P ) (ui )
· e j=1 β2
dui . (7.85)
i=−n i=1 V

Here, the n is related to P by P = 2n + 2 and the normal modes are aligned


as uj−n , . . . , uj−1 , uj0 , uj1 , . . . , ujn+1 , as introduced in Sec. 7.1.4. Making use
−1 n+1
  iπ
of the property that 2 sin = P , the Z C in Eq. (7.85) can be
i=−n i=1
P
further rewritten into
⎡ ⎤⎧  −1 n+1 
N   P2 ⎨  N   2πβ2 − 2
1
m j P
Z = lim ⎣
C ⎦
P →∞ 2βπ2 ⎩ mj P
j=1 j=1 i=−n i=1

 −1 % 
iπ 1 N
× 2 sin · du0 e−βV (u0 ,...,u0 )
P V
⎡ ⎤⎧  ⎫
N   P2 ⎨  N  P −1 ⎬
2 − 2
mj P 2πβ 1
= lim ⎣ ⎦
P →∞ 2βπ 2 ⎩ m j P P ⎭
j=1 j=1
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 187

Quantum Nuclear Effects 187


1 N
× du0 e−βV (u0 ,...,u0 )
⎡ V ⎤
N 
  12 
m j 1 N
=⎣ 2
⎦ dxc e−βV (xc ,...,xc ) , (7.86)
j=1
2βπ V

which is simply the classical partition function. In other words, the partition
function defined by Eq. (7.80) evolves into a classical partition function if
one sets the effective potential V eff (x1 , x2 , . . . , xP ) as

N 
P 
   P
1 1
mj ωP2 (xji − xji−1 )2 + V (x1c , . . . , xN
c ). (7.87)
i=1 j=1
2 i=1
P

From Sec. 6.4, we know that the free energies of two systems (F1 and F0 )
with potentials (U1 and U0 ) can be linked by a thermodynamic integral.
In the above introduction, we also understand that the free energies of
the “artificial” polymer corresponds to the free energies of the quantum
and classical systems, respectively, if one takes the effective potential as
Eqs. (7.81) and (7.83). Based on this analysis, one can easily introduce an
artificial effective potential between the classical and quantum systems as

P 
 N
1  2
V eff (x1 , x2 , . . . , xP ; λ) = mj ωP2 xji − xji−1
i=1 j=1
2

P
1 + ,
+ λV (x1i , . . . , xN 1 N
i ) + (1 − λ)V (xc , . . . , xc ) .
i=1
P
(7.88)

Using this effective potential, one can calculate the free energy of the “arti-
ficial system” between the classical and quantum ones through

1
F (λ) = − ln [Z(λ)] , (7.89)
β

with F (1) giving the quantum free energy and F (0) giving the classical one.
The difference between them, in terms of thermodynamic integration, can
be calculated using
 1

ΔF = F (1) − F (0) = dλF (λ), (7.90)
0
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 188

188 Computer Simulations of Molecules and Condensed Matter

where
P

 1 + 1 N 1 N
,
F (λ) = V (xi , . . . , xi ) − V (xc , . . . , xc ) . (7.91)
P i=1
V eff (λ)

Similar to Sec. 6.4, the symbol · · · V eff (λ) means that ensemble is generated
using the effective potential in Eq. (7.88). From Eqs. (7.90) and (7.91), the
free energy difference between the classical and quantum systems can be
rigorously evaluated. If the free energy of the classical system is known,
the free energy of the quantum system will be obtainable. We note that
this introduction follows the algorithm by Morales and Singer in Ref. [400].
The only thing one needs to take care in is the numerical stability, especially
in the strong quantum case when V (x1i , . . . , xN 1 N
i ) and V (xc , . . . , xc ) differ
significantly, as pointed out in Ref. [333]. In such cases, nonlinear interpo-
lation of the effective potential can be used. For details of this extension,
please see Ref. [333].

7.4 Examples
For a better understanding of the principles underlying the path integral
molecular simulations, similar to Chapter 5, we also use some examples to
show how they work in practice.

7.4.1 Impact of QNEs on Structures of the


Water–Hydroxyl Overlayers on Transition
Metal Surfaces

The first example concerns the problem on how the impact of the QNEs is
like that on the structure of the water–metal interface. This problem was
investigated by Li et al. in Ref. [286]. Here, we use some of their results to
show how the results of PIMD simulations are analyzed. The system chosen
is composed of a transition metal substrate and a hexagonal water–hydroxyl
overlayer.
Concerning the importance of such interfaces, it was already well known
in the study of surface physics/chemistry that under ambient conditions,
most surfaces are covered in a film of water [401]. These wet surfaces are of
pervasive and fundamental importance in processes like corrosion, friction,
and ice nucleation. On many such surfaces, it was also well known that the
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 189

Quantum Nuclear Effects 189

first contact layer of water does not comprise pure water, but instead a mix-
ture of water and hydroxyl molecules [401–411]. Physically, these overlayers
form because they provide the optimal balance of the hydrogen bonding
within the overlayer and the bonding of these overlayers to the surface, and
now they have been observed on several oxide, semiconductor, and metal
surfaces.
For the characterization of such overlayer structures, from the experi-
mental perspective, it is fair to say that these water–hydroxyl wetting layers
are now most well characterized on close-packed metal surfaces under ultra-
high vacuum (UHV) conditions [401]. In these experiments, it is widely
accepted that the molecules in the overlayer are “pinned” in registry with
the substrate, with the oxygen atoms sitting right above individual metal
atoms in the hydrogen bonded network (see Fig. 7.5). Because of this fea-
ture, we can say that the distances between the adjacent molecules are
determined mainly by the substrate, being relatively large on a metal with
a large lattice constant (e.g. ∼2.83 Å on average on Pt(111)) and small on
a metal with a relatively small lattice constant (e.g. ∼2.50 Å on average on
Ni(111)). From earlier studies of water in other environments, e.g. certain
phases of bulk ice, it was known that the behavior of the shared proton in
intermolecular hydrogen bonds varies dramatically over such a large range
of O–O distances. For example, under ambient pressures, bulk ice is a con-
ventional molecular crystal, with O–O separations of ∼2.8 Å. At very high
pressures ( 70 GPa), however, the O–O separations can decrease to ∼2.3 Å.
In the meantime, ice loses its integrity as a molecular crystal and the protons
become delocalized between the O nuclei (see, e.g., Refs. [339, 412, 413]).
Now, if we make a direct comparison between the behavior of proton in ice
under pressure and that of the water–hydroxyl overlayer on metal surfaces,
it is reasonable to expect that in the latter system, pronounced substrate
dependence of QNEs might exist.
To describe the influence of the QNEs on the structure of such over-
layers as well as its substrate dependence, as mentioned, one can take a
series of systems and perform both ab initio MD and PIMD simulations.
A comparison between the MD and PIMD results illustrates in a clear
manner how such an influence of the QNEs will be, and analysis on the
differences between the impact of QNEs on different substrates shows us
the substrate dependence. Based on this consideration, we choose three
substrates, i.e. Pt(111), Ru(0001), and Ni(111), and perform ab initio MD
and PIMD simulations at 160 K. These three substrates, in descending order
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 190

190 Computer Simulations of Molecules and Condensed Matter

√ √
Figure 7.5 Static geometry optimized structure of the 3 × 3-R30◦ overlayer (with
classical nuclei) that forms on the transition metal surfaces. Side views on right show
the cases when the proton is donated from water to hydroxyl (upper, labeled “short”)
and from hydroxyl to water (lower, labeled “long”). The short and long hydrogen bond
lengths (denoted by the dashed lines) are ∼1.7 and ∼2.1 Å on Pt, ∼1.6 and ∼1.9 Å on
Ru, and ∼1.4 and ∼1.6 Å on Ni, respectively. The coordinate for proton transfer δ is
defined as ROa H −ROb H , where ROa H and ROb H are the instantaneous O–H distances
between Oa and H and Ob and H, respectively. For a proton equidistant from its two
neighbors, δ = 0 and upon transfer from one O to another δ changes its sign.

of magnitude of the lattice constant, give an average O–O distance of ∼2.8,


∼2.7, and ∼2.5 Å. Ni was chosen here because of its relatively small lat-
tice constant, although we acknowledge that water–hydroxyl films have not
yet been characterized on it [401]. The simulation package chosen is the
famous Cambridge Sequential Total Energy Package (CASTEP) [414]. The
Perdew–Burke–Ernzerhof (PBE) exchange–correlation (XC) functional is
used for the descriptions
√ √ of the electronic interactions within the DFT [129],

together with a 3× 3-R30 water–hydroxyl overlayer (see e.g. Refs. [401–
407, 415, 416]).
Our discussions start from the properties of these overlayers at the clas-
sical level. These overlayers are composed of hexagonal hydrogen bonded
networks
√ of √ water and hydroxyl bonded above metal atoms of the substrate
in a 3 × 3-R30◦ periodicity. Both types of the molecules lie almost
parallel to the surface, forming a perfect extended 2D network. Because
OH is a better acceptor of hydrogen bonds than a donor, there is an
asymmetry in the overlayer with each molecule involved in two short and
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 191

Quantum Nuclear Effects 191

one long hydrogen bond at the classical static ground state (Fig. 7.5). At
finite temperature, ab initio MD simulations with classical nuclei show that
this asymmetry is still kept, although thermal fluctuations cause the peaks
associated with the long and short hydrogen bonds to overlap, particularly
on Ni (which has the smallest lattice constant). This asymmetry is illus-
trated in Fig. 7.6 where we show the probability distributions of O–H and
O–O distances on Pt, Ru, and Ni. In addition to this asymmetric feature,
the probability distribution of O–H distances also shows that the overlayer
comprises individual H2 O and OH molecules of hydrogen bonded to each
other. This is reflected by the sharp peak at ∼1.0 Å, characteristic of the
covalent bonds of water and hydroxyl, and broader peaks at ∼1.7–2.1 Å,
∼1.6–1.9 Å, and ∼1.5 Å, characteristic of the hydrogen bonds on Pt, Ru,
and Ni, respectively. The probability distributions of the O–H bond length
between these peaks characteristic of the covalent and hydrogen bonds are
negligible.
Then we turn on the QNEs and see what happens in the PIMD simula-
tions. A key result is that there is no longer a clear division between short
covalent and longer hydrogen bonds. This is explicitly shown in Fig. 7.6.
On Pt, the population of covalent O–H bonds is reduced by one-third and
replaced with a clearly non-zero probability distribution over the entire
range of 1–1.5 Å (Fig. 7.6(a)). Likewise, the proportion of the short O–O
distances is reduced from two-thirds to one-third, and the center of the
peak associated with the short O–O distances moves from ∼2.7 Å to ∼2.5 Å
(Fig. 7.6(b)). These changes are associated with one-third of the shared
protons being delocalized between the two oxygen atoms to which they
are bonded. In turn, this delocalization proton further “drags” the oxygen
atoms sharing it closer and in doing so creates an “H3 O2 ” complex. We note
that in this complex, the shared proton belongs to neither of the two oxygen
atoms. A typical snapshot from the ab initio PIMD simulation is shown in
Fig. 7.6(g) with the H3 O2 complex located along one particular O–O axis.
This snapshot also shows how when the two oxygen atoms on either side of
the shared proton are drawn close, the distances to their other oxygen neigh-
bors increase. It is this effect that leads to a larger proportion of the long O–
O distances than that was observed in the classical simulation (Fig. 7.6(b)).
On Ru, similarly, delocalization of the proton was observed and again
the structure contained H3 O2 complexes (Fig. 7.6(h)). The smaller lat-
tice constant of Ru also means that only a small variation in the propor-
tion of the short O–O separation (∼2.5 Å) is required to enable proton
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 192

192 Computer Simulations of Molecules and Condensed Matter

Figure 7.6 Statistical structural information from the ab initio MD and PIMD sim-
ulations of the water–metal interfaces, using some selected structural properties. More
specific, probability distributions of the O–H ((a), (c), (e)) and O–O distances ((b), (d),
(f)) on Pt(111), Ru(0001), and Ni(111) are chosen. Results obtained from ab initio MD
simulations with classical nuclei were labeled “classical” and shown by solid lines. Those
from the ab initio PIMD simulations with quantum nuclei at the statistical level were
labeled “quantum” and shown by dashed lines. A key difference between the MD and
PIMD results is that in the PIMD simulations, a non-negligible distribution of the O–H
distance between the covalent and hydrogen bond peaks was observed. This feature is
absent in the MD simulations with classical nuclei and it originates from some spatial
configurations of the system during the simulation in which one proton is equally shared
by two oxygen atoms. In panels (g)–(i), we show some snapshots for typical spatial
configurations of the overlayer on Pt, Ru, and Ni obtained from the PIMD simulations
(using 16 beads). On Pt and Ru, at any given snapshot, one proton is equally shared by
two of the oxygen atoms yielding an intermediate “H3 O2 ” complex. On Ni at any given
snapshot, several protons can simultaneously be shared between the oxygens.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 193

Quantum Nuclear Effects 193

delocalization (Fig. 7.6(d)). Upon moving to Ni, the influence of the QNEs
on these structural properties becomes even larger. This is shown by the
larger magnitude for the distribution of the O–H distances between the
peaks characterizing the covalent and hydrogen bonds (Fig. 7.6(e)). Because
of Ni’s smaller lattice constant, the quantum delocalization of the proton
within the overlayer becomes possible without any major rearrangement of
the oxygen nuclear “skeleton”. A snapshot from the PIMD simulation on
Ni, in which several protons are delocalized simultaneously and the distinc-
tion between covalent and hydrogen bonds is completely lost, is shown in
Fig. 7.6(i).
The obviously different probability distributions observed in the MD
and PIMD simulations mean that the QNEs significantly change the struc-
tures of the water–hydroxyl overlayer on the transition metal surfaces stud-
ied. For a more rigorous characterization of difference from a statistical
perspective, we further calculated the free energy profiles for the protons
along the intermolecular axes. This free energy profile is calculated using
ΔF (δ) = −kB T ln P (δ), where P (δ) is the probability distribution of δ and
δ is the proton transfer reaction coordinate as defined in Fig. 7.5. kB is
the Boltzmann constant. For an unbiased analysis, we take all inequiva-
lent hydrogen bonds in the system into account. In other words, the free
energy profile calculated here is an average over all hydrogen bonds in the
overlayer. The results are shown in Fig. 7.7. In the MD simulations with
classical nuclei, the free energy profiles are characterized by two partially
overlapping valleys. On Pt (Ru), they are located at δ ∼ 0.7 (0.6) and
δ ∼ 1.1 (0.9) Å. On Ni, these two valleys almost completely overlap at
∼0.5 Å, since, as we have said, thermal broadening obscures the distinction
between short and long hydrogen bonds on this surface. Concerning proton
transfer, it is a rare event, as reflected by the presence of large classical free
energy barriers on all substrates, at δ = 0.
Then we move to the free energy profiles obtained from the PIMD
simulation, in which the QNEs are included in the theoretical descriptions.
It is clear in Fig. 7.7 that they differ significantly from the MD ones. On the
Pt and Ru substrates, the minima for the long hydrogen bond remain. But
we note that those associated with the short hydrogen bonds completely
disappear due to the formation of the intermediate H3 O2 complexes as
mentioned before. On Ni, the single valley feature was kept. However, it was
softened and its position shifted from δ ∼ 0.5 Å to δ ∼ 0.4 Å. We note that
the key difference between the quantum and classical free energy profiles
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 194

194 Computer Simulations of Molecules and Condensed Matter

(a) (b) (c)


Figure 7.7 Free energy profile (denoted as ΔF ) for the protons along the intermolecular
axes within the water–hydroxyl overlayers on Pt (a), Ru (b), and Ni (c) from the ab
initio MD and PIMD simulations at 160 K. The MD results were labeled classical and
shown by solid lines and the PIMD results were labeled quantum and shown by dashed
lines.

is that in the quantum simulations, the proton transfer energy barriers are
significantly smaller than the classical ones. Upon going from Pt through Ru
to Ni, the height of the barrier and the area beneath it decreases, indicating
that proton transfer probability increases as the lattice constant is reduced,
and a plateau appears on all three quantum free energy profiles.
To understand how this plateau appears, we correlate the location of
the proton along the intermolecular axes (δ) with the corresponding O–O
distances (RO−O ) and plot the probability distribution as a function of
these two variables in Fig. 7.8. In the MD simulations (Figs. 7.8(a), 7.8(d),
and 7.8(g)), these functions are characterized by negligible distributions at
δ = 0, consistent with the fact that the proton transfer is a rare event
and the protons hop from one side of the hydrogen bond to the other. The
O–O distribution has two peaks for the short and long hydrogen bonds,
respectively, on Pt and Ru, but they merge on Ni. When the QNEs are taken
into account, finite distributions at δ = 0 appear on all three substrates.
These distributions correspond to the delocalized protons, as shown by the
snapshots in Fig. 7.6. To understand the behavior of this “delocalized”
proton from a more rigorous perspective, one can focus on the most active
proton, i.e. the proton which at any given snapshot in the PIMD simulations
has the smallest magnitude of δ. On Pt and Ru, this is the proton located
along the hydrogen bond with the smallest O–O distance. On Ni, due to
the fact that the average O–O distance is only ∼2.5 Å, the most active
proton need not necessarily be the one with the shortest O–O distance.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 195

Quantum Nuclear Effects 195

(a) (b) (c)

(d) (e) (f)

(g) (h) (i)

Figure 7.8 Probability distribution in the MD and PIMD simulations as a function of δ


and RO−O . Similar to the earlier figures, the MD simulations are labeled as classical and
the PIMD simulations are labeled as quantum. The left and middle columns show results
obtained from all hydrogen bonds in the overlayer. On the right column, only data from
the most active hydrogen bond are chosen in the PIMD simulations. The most active
proton is defined as the one with the smallest δ. All MD and PIMD distribution functions
have been symmetrized with respect to δ.

The results are shown in Figs. 7.8(c), 7.8(f), and 7.8(i). The key feature is
that different from the panels on the left and middle columns in Fig. 7.8,
where the mean peak is located at δ with large magnitude of the absolute
value, the distribution peaks on the right column clearly are located around
δ = 0. Therefore, the corresponding free energy barrier for the transfer of
the most active proton is zero and the classical proton transfer energy
barrier is wiped out by the QNEs.
Another mean feature of the distribution functions in Fig. 7.8 is that a
“horseshoe” shape exists. On Pt and Ru, this indicates that the covalently
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 196

196 Computer Simulations of Molecules and Condensed Matter

bonded proton requires the oxygen atoms to move close first. When this
O–O distance is smaller than a certain value, the classical proton transfer
energy barrier for the most active proton will become so small that it can
be easily wiped out by its zero-point energy. In this case, the quantum
nature of the proton results in an “adiabatic” response to the movement
of the oxygen atoms and the proton quickly becomes delocalized along
this short hydrogen bond. In this case, the H3 O2 complex as shown in the
earlier discussions appears, and it persists till the thermal fluctuations of
the oxygen force them to move apart. When this O–O distance is larger
than a certain value, the quantum zero-point energy fails to wipe out the
classical proton transfer energy barrier and consequently it falls to either
side and becomes covalently bonded to one of the oxygen atoms. Therefore,
the mechanism for proton transfer on Pt and Ru is the so-called “adiabatic
proton transfer” [417], as predicted for the diffusion of the excess proton in
water and ice at certain pressures [321, 412].
For more details of these simulations, please see Ref. [286].

7.4.2 Impact of Quantum Nuclear Effects on the Strength


of Hydrogen Bonds

The second example we want to show here, in which the quantum nature
of the nuclei is explicitly addressed, concerns a fundamental problem in
physics and chemistry, i.e. what will be the impact of QNEs on the strength
of hydrogen bonds.
We all know that hydrogen bonds are weak intermolecular interactions
which hold much of soft matter together, as well as the condensed phases
of water, network liquids, and many ferroelectric crystals. The small mass
of hydrogen, as shown already in the above example, means that they are
inherently quantum mechanical in nature, and effects such as zero-point
motion and tunneling must be taken into account in descriptions of the
properties related to it. As a prominent example, from the statistical point
of view, it is well known that by replacing H by D, the hydrogen bond
strength changes. However, as direct as it looks, a simple picture, in which
the impact of QNEs on the strength of hydrogen bonds and consequently
the structure of the hydrogen bonded systems, can be rationalized, has been
absent for a long time.
As a matter of fact, this problem concerning the influence of QNEs on
the strength of hydrogen bonds is a fundamental problem in physics and
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 197

Quantum Nuclear Effects 197

chemistry. Already in the 1950s, it was observed experimentally that in


some hydrogen bonded molecular crystals, by replacing the hydrogen with
deuterium, the heavy atom (e.g. O–O) distances change [418, 419]. This phe-
nomenon is known as the Ubbelohde effect. The conventional Ubbelohde
effect causes an elongation of the O–O distance upon replacing H with
D, indicating that the QNEs strengthen the hydrogen bond, although a
negative Ubbelohde effect has also been observed in several systems [418,
419]. From the molecular simulation perspective, starting from the early
1980s, when PIMD and PIMC simulations became a conventional routine
to investigate the influence of QNEs on real poly-atomic systems, computer
simulations on this issue had been carried out in a wide range of sample sys-
tems. A general conclusion is that the result is strongly system-dependent.
In liquid hydrogen fluoride (HF), for example, ab initio MD and PIMD
simulations using DFT for the description of the electronic structures have
shown that when the QNEs are accounted for, the first peak in the F–F
radial distribution function (RDF) sharpens and shifts to a shorter F–F dis-
tance [420]. The implication of this increase in the structuring of the RDF
in the liquid is that the hydrogen bond is strengthened upon including the
QNEs. In contrast, similar simulations for liquid water show that the O–O
radial distribution function is less peaked when simulations with quantum
nuclei are compared with those with classical nuclei [338], suggesting a
decrease of the overall hydrogen bond strength. We note, however, that
although this conclusion is probably correct, it is the opposite of what was
observed in an earlier ab initio study [345].
Besides these discussions concerning hydrogen bonded crystals and liq-
uids, the influence of the QNEs on the hydrogen bonds has also been widely
discussed in studies of gas phase clusters [349, 421, 422]. Specifically, in
water clusters up to hexamer, it is predicted that the QNEs weaken the
hydrogen bonds, whereas in simulations of the HF clusters, both strength-
ening and weakening is predicted depending on the size of the cluster [349,
421, 423]. For clusters smaller than tetramer, a weakening of intermolec-
ular hydrogen bond is predicted upon including the QNEs. For clusters
larger than tetramer, a strengthening of the hydrogen bond is expected.
In tetramer, the influence is negligible. Clearly, it would be very useful to
rationalize these various results within a single conceptual framework and
identify the underlying factors that dictate the influence of the QNEs on
hydrogen bond strength for a broad class of materials. In a recent study
[322], Li et al. gave a simple picture to rationalize these different results
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 198

198 Computer Simulations of Molecules and Condensed Matter

using analysis based on ab initio MD and PIMD simulations. Here, we use


some of their key results to show how it is done in practice.
First of all, a broad range of hydrogen bonded systems are chosen,
including HF clusters (dimer to hexamer), H2 O clusters (dimer, pentamer,
and octamer), charged, protonated, and hdyroxylated water and ammonia
clusters (H9 O− + − −
5 , H9 O4 , H7 O4 , and N2 H5 ), organic dimers (formic acid
and formamide), and solids (HF, HCl, and squaric acid C4 H2 O4 ). For each
system, both conventional ab initio MD simulations, in which the nuclei are
treated as classical point-like particles, and more state-of-the-art ab initio
PIMD simulations, in which the QNEs are accounted for, were performed.
With these two complementary sets of simulations, one can identify in a very
clear manner the precise influence of the QNEs on the statistical properties
of interest at finite temperatures.
Before we start, let us first make the following points clear. First, the
quantities we focus on when characterizing H-bonds are: (i) the heavy-atom
(X–X, where X is either O, Cl, C, N, or F) distances, which characterize
the intermolecular separations, (ii) the H-bond angles (X–H–· · · –X), which
are associated with H-bond bending (libration) modes, and (iii) the X–H
covalent bond lengths, characteristic of the covalent bond stretching in the
H-bond donor molecules. In later discussion, it will become clear that these
quantities provide an indication of H-bond strength. However, as the main
measure of H-bond strength, we still use a standard estimate based on
the computed red-shift (softening) in the X–H stretching frequency of the
H-bond donor molecule. We note that there is no perfect measure for H-
bond strength [424], however the red-shift of the stretching frequency is a
widely used measure [see e.g. Refs. [425, 426]). This measure is particularly
useful here because it allows us to discriminate between different types of
H-bond in the same complex and can be used for both neutral and charged
systems. In Fig. 7.9(b), it is shown that this estimator correlates well with
the computed binding energy per H-bond in the neutral systems we study.
This binding energy is defined as the difference between the total energy
of the system and the sum over its unrelaxed components, as in Ref. [424].
When the red-shift of the stretching frequency (measured as the ratio of
the X–H stretching frequency in the H-bonded cluster to that in the free
monomer) gets larger, the H-bond becomes stronger.
With the definition of the above-mentioned quantities in mind, we first
look at the results for the impact of the QNEs on the strength of hydro-
gen bonds. Upon comparing these results for the various hydrogen bonded
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 199

Quantum Nuclear Effects 199

systems, an interesting correlation can be established between the H-bond


strength and the change in intermolecular separations. This correlation is
shown in Fig. 7.9(a) where we see that as the H-bond gets stronger, the
heavy-atom separations in the PIMD simulations with quantum nuclei go
from being longer than those in the MD simulations with classical nuclei
(positive Δ(X–X)) to being shorter (negative Δ(X–X)). Thus, the QNEs
result in longer hydrogen bonds in weak hydrogen-bonded systems and
shorter hydrogen bonds in relatively strong hydrogen-bonded systems. We
note that the hydrogen-bond strength increases upon going from small to
large clusters and from water to HF. The trend reported in Fig. 7.9 is a
key finding and in the following, we explain why it emerges and discuss the
implications it has for H-bonded materials in general.
To understand the reason for this correlation between the impact of
the QNEs on the strength of hydrogen bond and the strength of hydro-
gen bond itself, it is useful to look at the HF clusters. These provide the
ideal series because upon increasing the cluster size, the hydrogen-bond
strength increases, and the influence of the QNEs switches from a tendency
to lengthen to a tendency to shorten the intermolecular separations (as seen
in Ref. [423]). Our analysis is summarized in Fig. 7.10, where we plot the
distance and angle distributions from MD and PIMD simulations for these
three HF clusters separately. The left column shows the final results, where
one can see that in the dimer, the averaged F–F distance is increased by
including the QNEs; in the tetramer, there is no difference between the
averaged quantum and classical F–F distances; in the pentamer, the F–F
distance is clearly shortened by including the QNEs. The key to understand-
ing this variation of the heavy-atom distances is in recognizing that there
are also related differences between MD and PIMD in the covalent F–H
bond lengths (center) and H-bond angles (right). Because of anharmonic
quantum fluctuations, these two geometric properties also show systematic
changes. First of all, the F–H bonds are longer in the quantum compared
to the classical simulations, and this elongation becomes more pronounced
as the H-bonds get stronger. Second, the hydrogen bonds are more bent in
the quantum than in the classical simulations, and this bending generally
becomes less pronounced as the hydrogen bonds get stronger. In order to
understand the influence of these variations in structure, analysis of various
dimer configurations was performed. This analysis reveals that the covalent
bond stretching increases the intermolecular interaction whereas hydrogen
bond bending decreases it. Taking the HF dimer as an example, a 0.04 Å
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 200

200 Computer Simulations of Molecules and Condensed Matter

(a) (b)

(c)

Figure 7.9 Correlation between the impact of the QNEs and the hydrogen bond
strength. In panel (a), the differences between the shortest heavy-atom distances obtained
from the PIMD and MD simulations (X–X)PIMD MD
average –(X–X)average , denoted by Δ(X–
X), is chosen as the y axis. It characterizes the impact of the QNEs on the strength
of the hydrogen bonds. This influence is drawn as a function of the hydrogen-bond
strength. As mentioned in the prose, this hydrogen bond strength is defined as the
ratio of the X–H stretching frequency in the hydrogen-bonded system to that in the
free monomer. In panel (b), the correlation between this hydrogen-bond strength index
and the binding energy per hydrogen bond in the neutral systems is given. In panel
(c), simplified schematic illustration of the expected isotope (Ubbelohde) effect on the
differences in heavy-atom distances. We suggest that three regimes of positive, negli-
gible, and negative Ubbelohde effect depending on the hydrogen-bond strength exist.
For the HF clusters, labels (1)–(5) denote the hydrogen bonds in the dimer to the
hexamer. For the water clusters, labels (1), (2), (3a), and (3b) refer to the hydrogen
bonds in the dimer, pentamer, and the long (short) hydrogen bond in the octamer. For
the charged clusters, labels 1–4 refer to H9 O− + − −
5 , H9 O4 , H7 O4 , and N2 H5 , respectively.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 201

Quantum Nuclear Effects 201

←−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−
Figure 7.9 (continued) For the organic dimers, labels (1a), (1b), and (2) refer to the
red-shifted and blue-shifted hydrogen bond in the formamide and the red-shifted hydro-
gen bond in the formic acid. For the solids, labels (1)–(3) refer to the hydrogen bonds
in HCl, HF, and squaric acid. The same labels are applied in Fig. 7.11. For the water
cluster in panel (b), the octamer is not included since there are two kinds of hydrogen
bonds. Results for the trimer and tetramer are added to further test the correlation.

(a) (b) (c)

(d) (e) (f)

(g) (h) (i)

Figure 7.10 HF clusters as examples for detailed analysis of the QNEs. Distributions of
the F–F distances (left), the F–H bond lengths (center), and the intermolecular bending
(F–H· · · F angle, right) from the MD (solid lines) and PIMD (dashed grey lines) for
a selection of systems: the HF dimer (top), the HF tetramer (middle), and the HF
pentamer (bottom). The MD and PIMD averages are shown in black and grey vertical
dashes, respectively.

increase in the F–H bond length of the donor leads to a 40 meV increase in
interaction energy within the dimer, whereas in contrast, a 21◦ reduction in
H-bond angle leads to a 16 meV decrease in interaction energy. This analysis
provides a qualitative understanding of the trend observed. In short, the
F–F distance increases in the dimer as a result of a large decrease in hydro-
gen bond angle, but only a small increase in the covalent F–H bond length.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 202

202 Computer Simulations of Molecules and Condensed Matter

Figure 7.11 A quantification for the competition between the quantum fluctuations on
the stretching and bending modes. Differences in average shortest heavy-atom distances
between PIMD and MD simulations (Δ(X–X), vertical axis) vs. the ratio of the projec-
tion of the donor X–H covalent bond along the intermolecular axis from PIMD and MD
simulations (horizontal axis). For the meaning of the labels, please refer to the caption of
Fig. 7.9. x larger (smaller) than 1 indicates a dominant contribution from the stretching
(bending) mode when the QNEs are included. Negative values of Δ(X–X) indicate that
quantum nuclear effects decrease the intermolecular separation. An almost linear corre-
lation between the two variables can be observed: When the contribution from stretching
becomes more dominant, the QNEs turn from weakening to strengthening the H-bonds.
The inset illustrates the geometry used for projecting the donor covalent X–H bond onto
the intermolecular axis. The curved arrow represents the intermolecular bending and the
straight arrow represents the intra-molecular stretching.

While in the tetramer, the F–H stretching is sufficiently pronounced to com-


pensate for the increase in hydrogen-bond bending, leaving the overall F–F
distance unchanged, in the pentamer, the F–F distance decreases because
the F–H covalent bond stretching dominates over the H-bond bending.
For a rigorous examination of this picture and a quantitative description
of this competition for all systems studied, one can further calculate the pro-
jection (X–H|| ) of the donor molecule’s covalent bond along the intermolec-
ular axis (see inset of Fig. 7.11). Since X–H|| increases upon intra-molecular
stretching but decreases upon intermolecular bending, this quantity itself
allows the balance between stretching and bending to be evaluated, to a
certain extent. The influence of the QNEs is quantified by the ratio of
the PIMD and MD projections, i.e. x = (X–H|| )PIMD /(X–H|| )MD . When
this value is clearly greater than one, it indicates that when the QNEs are
included, the main influence is on the stretching of the covalent bond, and
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 203

Quantum Nuclear Effects 203

when this value is clearly smaller than 1, it indicates that when the QNEs
are included, the main influence is on the bending of the hydrogen bond.
When one plots this ratio against the variations in intermolecular separa-
tions, y = Δ(X–X) (which we used to quantify the impact of the QNEs). A
striking (almost linear) correlation is observed (Fig. 7.11). For all systems
where hydrogen-bond bending dominates (x clearly smaller than 1), the
heavy-atom distances are longer in PIMD than in MD (y > 0). In cases
where covalent bond stretching is dominant (x clearly larger than 1), the
heavy-atom distances are shorter in PIMD than in MD (y < 0). With
the increase of x, quantum fluctuations on the stretching mode become
more dominant and the QNEs turn from weakening the hydrogen bonds
to strengthening them. Thus, the overall influence of the QNEs on the
hydrogen bonding interaction quantitatively comes down to this delicate
interplay between covalent bond stretching and intermolecular bond bend-
ing. One notes that this explanation arrived at here for the general case
is consistent with what Manolopoulos and coworkers have elegantly shown
for liquid water in Ref. [427].
Given the ubiquity of the hydrogen bonds in the physical, chemical, and
biological sciences, there are a number of implications in this finding. Con-
sidering that liquid HF comprises long polymer chains and rings whereas
liquid water is widely considered to be made up of small clusters, these
results shed light on why the QNEs strengthen the structure of liquid HF
but weaken that of liquid water [338, 420]. More generally, one can use the
trend observed in Fig. 7.9 as a simple rule of thumb to estimate the impact
of the QNEs on hydrogen-bonded systems without performing expensive
PIMD simulations. All that is required is an estimate of hydrogen-bond
strength, which can be obtained from the red-shift in the covalent stretch-
ing frequency or from other commonly used measures of hydrogen-bond
strength such as hydrogen-bond length. Thus, the trend may be particularly
useful to biological systems such as α-helixes and β-sheets for which many
crystal structures have been determined and where cooperative effects lead
to particularly strong H bonds [428].
In addition to the implications mentioned, this trend also allows one
to rationalize the Ubbelohde effect over a broad range of H-bond regimes
(Fig.7.9(c)). Specifically speaking, traditional Ubbelohde ferroelectrics such
as potassium dihydrogen phosphate fall in the relatively strong H-bond
regime where a positive Ubbelohde effect (i.e., an increase of the X–X dis-
tance upon replacing H with D) is observed in the experiment and also
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 204

204 Computer Simulations of Molecules and Condensed Matter

in recent PIMD studies [419, 429], and in this context, the squaric acid,
the solid HF, and the larger HF clusters are expected to exhibit a tradi-
tional Ubbelohde effect upon replacing H with D. In contrast, the smaller
hydrogen-bonded clusters studied here and solid HCl are expected to exhibit
a negative Ubbelohde effect (a decrease of the X–X distance upon replac-
ing H with D). Hydrogen-bonded materials of intermediate strength such
as large water clusters and ice at ambient pressure are predicted to exhibit
a negligible Ubbelohde effect because in this regime, the QNEs have little
influence on the intermolecular separations. Indeed, this observation is con-
sistent with experimental and theoretical observations for the ferroelectric
hydrogen-bonded crystals, ice, and gas phase dimers [418, 419, 422].
A further prediction stemming from this work is that ice under pressure
will exhibit the traditional Ubbelohde effect. However, one cautions that at
very high pressure, ice possesses such strong hydrogen bonds, with shared
symmetric protons [412], that the picture sketched in Fig. 7.11 is not likely
to apply. Indeed, this note of caution applies to all ultra-strong hydrogen
bonds, where the proton is shared symmetrically by the two heavy atoms
already in the classical perspective. In this case, the distinction between a
relatively short covalent bond and a relatively long hydrogen bond is lost
and bond stretching along the X–X axis does not lead to any strengthening
of the intermolecular interactions. The gas phase Zundel complex, H5 O+ 2,
is an example of one such ultra-strong hydrogen bond and the calculations
in Ref. [322] show an approximate increase of 0.016 Å in the O–O distance,
which is consistent with the previous studies [326, 430].
Another class of very strong H-bonded systems are the so-called “low-
barrier” H-bonds, e.g. H3 O− + −
2 , N2 H7 , and N2 H5 . In these systems, there
remains a clear distinction between covalent and hydrogen bonds, and the
picture we have presented still holds. This fact can be seen from our data for
N2 H−5 in Fig. 7.9. We caution, however, that in these very strong H-bonded
systems, errors associated with the underlying XC functional can have a
qualitative impact on the results and that the accuracy of the underlying
PES is of critical importance. For example, in H3 O− +
2 and N2 H7 , using the
PBE XC functional yields a shared symmetric proton already in the classical
MD simulations. But earlier studies with the more accurate second-order
Møller–Plesset perturbation theory and also the Becke–Lee–Yang–Parr XC
functional [326, 430, 431] show that protons actually feel a double-well
potential and in this case, quantum nuclear effects strengthen the hydrogen
bond, consistent with the model presented here. In addition to this, since
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 205

Quantum Nuclear Effects 205

both inter- and intramolecular vibrations are relevant to the QNEs, this
work also highlights the need for flexible anharmonic monomers in force
field simulations of the quantum nuclear effects. Specifically, if this feature
is absent, only hydrogen-bond bending will be present in the simulation and
consequently the intermolecular interaction will be “artificially” weakened.
For more details concerning such discussions and the numerical details of
the calculations, please see Ref. [322].

7.4.3 Quantum Simulation of the Low-Temperature


Metallic Liquid Hydrogen

The third example we show here, in which the ab initio PIMD simulation
is used to study the fundamental properties of condensed matter, concerns
the existence of a low-temperature quantum metallic liquid, which exists in
high-pressure hydrogen.
Concerning the importance for the existence of this low-temperature
metallic liquid phase, one can track back to a very famous conjecture about
hydrogen under pressure. This conjecture was first proposed by Wigner
and Huntington in 1935, and states that solid molecular hydrogen would
dissociate and form an atomic metallic phase at high pressures [432]. Ever
since this prediction, the phase diagram of hydrogen has been the focus
of intense experimental and theoretical studies in condensed matter and
high-pressure physics [278, 281, 433–436].
Due to the advancement of many experimental techniques, notably
diamond anvil cell approaches, it is possible nowadays to explore hydro-
gen at pressures up to about 360 GPa [433, 437–439], and one notes that
new types of diamond anvil cell may be able to access even higher pres-
sures [440]. These experiments, together with numerous theoretical stud-
ies, have revealed a remarkably rich and interesting phase diagram com-
prising regions of stability for a molecular solid, a molecular liquid and
an atomic liquid, and within the solid region, four distinct phases have
been detected [438, 439, 441]. In high-temperature shock-wave experiments,
metallic liquid hydrogen has also been observed [442, 443]. It is accepted
to be a major component of gas giant planets, such as Jupiter and
Saturn [443]. Despite the tremendous and rapid progress, important gaps
in our understanding of the phase diagram of high-pressure hydrogen
still remain, with arguably the least well-understood issue being the solid
to liquid melting transition at very high pressures. Indeed, the melting
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 206

206 Computer Simulations of Molecules and Condensed Matter

curve is only established experimentally and theoretically up to around


200 GPa [278, 436]. From 65 GPa to about 200 GPa, the slope of the melting
curve is negative (that is, the melting point drops with increasing pressure),
which suggests that at yet higher pressures, a low-temperature liquid state
of hydrogen might exist or, as suggested by Ashcroft [444], perhaps even a
metallic liquid state at 0 K. Further interest in hydrogen at pressures well
above 200 GPa stems from other remarkable suggestions, such as superflu-
idity [435] and superconductivity at room temperature [445, 446], all of
which imply that hydrogen at extreme pressures could be one of the most
interesting and exotic materials among all condensed matter types.
In Ref. [323], Chen et al. used computer simulation techniques to
probe the low-temperature phase diagram of hydrogen in the ultra-high
500–1200 GPa regime to try and find this potential low-temperature liq-
uid state of hydrogen. Concerning the proton motion in this condensed
phase, ab initio PIMD as introduced earlier in this chapter has been used.
To compute the melting curve, the solid and liquid phases in coexistence
were simulated [278, 447, 448]. This coexistence approach, otherwise called
the two-phase simulation method, minimizes hysteresis effects arising from
superheating or supercooling during the phase transition. With this combi-
nation of approaches, they have found a low-temperature metallic atomic
liquid phase at pressures of 900 GPa and above, down to the lowest tempera-
ture they can simulate reliably, 50 K. The existence of this low-temperature
metallic atomic liquid is associated with a negative slope of the melting
curve between atomic liquid and solid phases at pressures between 500 and
800 GPa. This low-temperature metallic atomic liquid is strongly quantum
in nature, as treating the nuclei as classical particles using the ab initio MD
method significantly raises the melting curve of the atomic solid to ∼300 K
over the whole pressure range. The classical treatment of the nuclei does not
reproduce a notable negative slope of the melting curve and consequently
does not predict a low-temperature liquid phase.
For a clear explanation of these results, we go through the detailed
procedures of their study here. One problem which is essential in a the-
oretical description of the hydrogen phase diagram is that in its solid
phase, many local minima on the potential energy surface exist. There-
fore, in order for the ab initio MD and PIMD simulations to make sense,
extensive computational searches for low-enthalpy solid structures of hydro-
gen must be performed. From earlier studies using DFT methods [15, 449–
452], a metallic phase of I41/amd space group symmetry has been widely
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 207

Quantum Nuclear Effects 207

reported to be stable from about 500 to 1200 GPa, when quasi-harmonic


proton zero-point motion was included. Accordingly, they have used this
phase as the starting point for their finite-temperature exploration of the
phase diagram and melting curve [323]. With the coexistence method, they
have performed a series of two-phase solid–liquid simulations at different
temperatures (from 50 K to 300 K), which are then used to bracket the
melting temperature from above and below. We begin by considering the
500–800 GPa pressure regime and show an example of the data they have
obtained from the coexistence simulations at 700 GPa in Fig. 7.12. At this
pressure, one can see that for T  125 K, the system transforms into a
liquid state, whereas for T  100 K, it ends up as a solid. To characterize
these states, they have used a pair-distribution function and averaged out
its angular dependence. The result is a function of interatomic separations,
which is denoted by g(r) throughout this section. In a liquid, this is the
so-called radial distribution function. As can be seen in Fig. 7.12(d), upon
moving from 100 to 125 K, the system clearly possesses less structure, indi-
cating that a transition from solid to liquid occurs. These phases were also
characterized by the variations in the mean-square displacement (MSD) of
the nuclei of the particles over time. As PIMD rigorously provides only ther-
mally averaged information, they have used the partially adiabatic centroid
MD (PACMD) approach within the path integral scheme to obtain real-time
quantum dynamical information [397]. Again, as shown in Fig. 7.12(e), the
distinction between the solid phase at 100 K and liquid phase at 125 K is
clear.
The same coexistence procedure was used to locate the melting point at
500 GPa and 800 GPa, leading to the melting curve shown in Fig. 7.13. The
up (down) triangles indicate the highest (lowest) temperatures at which
the solid (liquid) phases are stable, bracketing the melting temperatures
within a 25 K window. From this, we see that the melting temperature
is between 150 K and 175 K at 500 GPa, and that it drops rapidly with
increasing pressure, yielding a melting temperature that is only between
75 K and 100 K at 800 GPa. Thus, the melting curve has a substantial neg-
ative slope (dP/dT < 0) in this pressure range. Across this entire pressure
range, the molten liquid state is atomic, and the solid phase, which grows, is
the original atomic I41/amd phase that was used as the starting structure.
Given that molecular phases have been observed at pressures lower than
360 GPa in both experimental and theoretical studies [437, 439], they have
suggested that a molecular-to-atomic solid–solid phase transition should
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 208

208 Computer Simulations of Molecules and Condensed Matter

Figure 7.12 Ab initio PIMD simulations of solid–liquid coexistence and melting. Snap-
shots of the PIMD simulations at 700 GPa showing (a) the starting structure, (b) the
final state at 100 K, and (c) the final state at 125 K. Around 32 beads (pictured in the
larger squares) were used to represent the imaginary-time path integral for each atom.
The balls in the insets of (b) and (c) correspond to the centroid of each atom. (d) The
angularly averaged pair-distribution function g(r) for the same two simulations at 100 K
and 125 K. At 100 K, the solid state persists (solid line) as indicated by the relatively
sharp peaks. At 125 K (dashed line), these peaks are much broader and the g(r) is
characteristic of a liquid. This is further supported by the data in panel (e), where the
MSDs as a function of time from separate adiabatic CMD simulations within the path
integral framework are shown. The MSD for the 100 K solid-phase saturates rapidly,
whereas for the liquid phase at 125 K, it rises approximately linearly with time, resulting
in a finite-diffusion coefficient.
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 209

Quantum Nuclear Effects 209

Figure 7.13 Phase diagram of hydrogen and the low-temperature metallic liquid phase.
Regions of stability for the molecular solid, molecular liquid, atomic solid, and atomic
liquid are indicated by the various shadings. The dashed line separating the molecular
and atomic liquid phases is taken from quantum MC calculations [282]. The solid line
separating the molecular solid and molecular liquid phases is taken from ab initio MD
simulations [284], whose negative slope has been confirmed by experiment [436]. The
thick black line is the melting curve obtained in this study from the ab initio PIMD
coexistence simulations. The solid lines separating phases I, II, III, and IV are from
Refs. [441, 453]. The inset shows how the high-pressure melting curve (dashed lines) are
established here. The black and grey triangles (inset) correspond to the PIMD and MD
results, respectively. The solid up triangles give the highest temperatures for solidification
and the solid down triangles show the lowest temperatures for liquefaction. At 900 GPa
and 1200 GPa, the so-called degeneracy temperature is ∼40 K, below which the exchange
of nuclei will be important. Accordingly, 50 K was the lowest temperature examined in
our PIMD simulations. At this temperature, each simulation yields a liquid state, and so
the two open triangles at 900 GPa and 1200 GPa indicate upper bounds for the melting
temperature.

occur between 360 GPa and 500 GPa (the lowest pressure they have con-
sidered in their simulations of the melting).
The negative slope of the melting curve up to 800 GPa suggests that at
even higher pressures, a lower-temperature liquid phase might exist. Moti-
vated by this, they also carry out simulations at 900 GPa and 1200 GPa.
However, in this pressure range, one needs to consider nuclear exchange
effects, which are neglected in the PIMD simulations, but could potentially
become significant. Indeed, analysis of their simulations reveals that at these
pressures, the dispersion of the beads in the path integral ring polymer
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 210

210 Computer Simulations of Molecules and Condensed Matter

becomes comparable to the smallest interatomic separations when the


temperature is below ∼40 K. This is the so-called quantum degeneracy
temperature below which the exchange of nuclei will be important, and
consequently simulations with a (standard) PIMD approach are expected
to be inaccurate. With this in mind, they have performed all simulations in
this very high-pressure regime at T  50 K. Interestingly, they find that at
50 K, at both 900 GPa and 1200 GPa, the systems are already in the liquid
state, revealing that the melting temperature at these pressures is below
50 K. Whether the liquid phase is the 0 K ground state of hydrogen at these
pressures is not something one can establish at this stage. However, the large
negative slope of the melting curve at lower pressures and the observation
of a liquid phase at temperatures as low as 50 K provide strong support
for Ashcroft’s low-temperature liquid metallic state of hydrogen [444], and
it implies that any room temperature superconductor in this regime would
have to be a liquid.
In order to understand the role the QNEs play in inducing the prop-
erties discussed, it is informative to compare the results of the ab initio
PIMD simulations with those obtained from the ab initio MD approach in
which the nuclei are approximated by classical point-like particles. To this
end, they have performed a second complete set of coexistence simulations
with ab initio MD across the entire 500–1200 GPa range. The ab initio
MD melting curve is shown by the grey data in the inset of Fig. 7.13,
where it can be seen that the melting temperatures obtained from the MD
simulations are much higher than those from the fully quantum PIMD sim-
ulations. The ab initio MD melting temperature is well above 200 K across
the pressure range 500–1200 GPa, and the slope of the melting curve is
small. A melting curve with a negative slope was also found above 90 GPa
in the ab initio MD simulations of hydrogen by Bonev et al. [278] and
above 10 GPa in lithium [454]. In Ref. [323], ab initio MD simulations with
classical nuclei exhibit considerably higher melting temperatures than the
ab initio PIMD ones at pressures above 500 GPa, which shows that the
quantum description of the protons strongly depresses the melting point.
The entropy arising from the greater delocalization of the protons in the
quantum description has a crucial role in stabilizing the low-temperature
liquid.
Before we end our discussion, it is worthwhile to note that serious
analysis of the accuracy of the simulations should always be carried out
in molecular simulations in general. Taking the low-temperature metallic
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 211

Quantum Nuclear Effects 211

liquid phase we discussed above as an example, the main conclusion is that


the melting line of solid hydrogen has a negative slope, and that the quan-
tum fluctuations of the nuclei lead to a low temperature (<50 K) metallic
liquid phase at pressures higher than 900 GPa. We use the remaining part
of this section to discuss the accuracy of the simulations from which these
conclusions are drawn. This analysis includes: (i) the accuracy of the elec-
tronic structures, and (ii) the convergence of the ab initio MD and PIMD
simulations with respect to the number of beads used in representing the
finite-temperature imaginary-time path integral of the nuclei, the simula-
tion cell size and simulation time, as well as the significance of nuclear
exchange effects, a factor which is not accounted for in the PIMD method.
Besides these, the superconducting properties of the solid atomic phase will
also be discussed.
We start with the accuracy of the Brillouin-zone integrations and
plane wave basis set cut-off energies used in our MD and PIMD calcu-
lations. In the main manuscript of Ref. [323], a Monkhorst–Pack k-point
mesh of spacing 2π × 0.05 Å−1 was used for the Brillouin-zone integra-
tion and a 500 eV cut-off was used for the expansion of the electronic
wave functions. Figure 7.14 shows the variation of the relative static lat-
tice enthalpies of various relevant structures over the pressure range 500–
1200 GPa. These results are in very good agreement with those reported
in an even earlier study [451]. The molecular Cmca phase is found to
be the most stable one at 500 GPa, and the phase transition from Cmca
to the atomic I41/amd phase occurs at about 500 GPa. The I41/amd
phase has the lowest static lattice enthalpy from about 500 GPa to over
1200 GPa.
Then one investigates how well the above k-point mesh spacing and cut-
off energy perform when thermal and quantum fluctuations of the nuclei are
included in the calculations. For this purpose, six snapshots (three from sim-
ulations for solids at low temperatures and three from liquid phases) were
chosen at random from the PIMD simulations at 700 GPa. The centroid of
each atom is used, and they have performed single point calculations for
the total energy of these structures using a higher energy cut-off (600 eV)
and a denser k-point mesh (8 × 8 × 8, which corresponds to a grid of spac-
ing 2π × 0.025 Å−1 in the Monkhorst–Pack k-point mesh). The differences
between the results obtained with these settings and those used in the MD
and PIMD simulations (500 eV and a 4 × 4 × 4 k-point mesh) are smaller
than 1 meV, see Fig. 7.15(a). These errors are negligible compared with the
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 212

212 Computer Simulations of Molecules and Condensed Matter

Figure 7.14 Static lattice ground state enthalpies of different crystal structures relative
to FCC in solid hydrogen as a function of pressure.

Figure 7.15 Single point total energies of snapshots from the thermalized state of the
two-phase PIMD simulations at 700 GPa. The centroid position is used for simplicity. s-1,
s-2, and s-3 correspond to snapshots at low temperature (100 K) with hydrogen in the
solid I41/amd phase. l-1, l-2, and l-3 correspond to snapshots of the liquid phase at high
temperature (150 K).

energy differences of several tens of meV between the internal energy of the
liquid and solid phases.
One notes that the PBE XC functional was used in MD and
PIMD simulations reported. This functional suffers from self-interaction
errors which can be significant in systems containing hydrogen. One can
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 213

Quantum Nuclear Effects 213

investigate the potential role of the self-interaction errors by comparing


the total energies obtained with PBE and PBE0. PBE0 is a hybrid func-
tional containing 25% Hartree–Fock exact exchange [455]. Therefore, the
self-interaction error arising from the PBE0 functional is expected to be
smaller than that of the PBE functional. In addition to this, comparisons
with other functionals such as LDA and optB88-vdW [456, 457] within the
van der Waals density functional (vdW-DF) scheme [458, 459] should also
give some insights. Based on this consideration, in Fig. 7.15(b), we compare
the relative energies of the six snapshots using LDA, PBE, optB88-vdW and
PBE0. LDA and optB88-vdW give very similar results to the PBE ones.
PBE0 gives lower total energies for the liquid phase than for the solid, and
consequently, it favors melting of the solid. It is therefore likely that using
a more accurate density functional than PBE would lead to stabilization of
the liquid phase at even lower temperatures.
In PIMD simulations, the number of beads used to sample the
imaginary-time path integral is a very important parameter in the descrip-
tion of the quantum nuclear effects. A series of tests at 700 GPa were there-
fore performed and the melting temperature was calculated using 1 (MD), 4,
8, 16, 24, 32, 48, and 64 beads to check if the melting temperature converges
with respect to the number of beads. The results are shown in Fig. 7.16. We
find that 32 beads are required to ensure that the melting temperature is
converged within a window of 25 K. In Ref. [323], they have therefore used
32 beads for the main calculations reported in the main manuscript.
The results of the two-phase simulations also depend on the size and
shape of the simulation cell. In Fig. 7.17, one checks the dependence of the
results on the cell size in the MD simulations. Using a cell containing 200
atoms gives results identical to those from 432 to 576 atoms. In the PIMD
simulation reported in Ref. [323], they have also compared results using 200
and 300 atoms and found that these simulations gave essentially identical
results. Therefore, they believe that using a cell containing 200 atoms in
the simulations is accurate (at least) for a qualitative description of the
phenomena reported.
In the two-phase PIMD simulations, solidification and melting hap-
pen on a time scale of 1 ps. To ensure that the systems have equilibrated,
Chen et al. have run all simulations for 10 ps and calculated the angularly
averaged pair-distribution function g(r), as explained above, using different
time intervals. They already found very good convergence with respect to
the simulation time at 3 ps (Fig. 7.18).
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 214

214 Computer Simulations of Molecules and Condensed Matter

Figure 7.16 Melting temperatures calculated at 700 GPa using different numbers of
beads. A bead number of 1 means an MD simulation. The upper and lower limits of
the melting temperatures from the two-phase simulations are indicated by down and
up triangles, respectively. The dashed line indicates the middle of the upper and lower
limits.

Figure 7.17 Melting temperatures calculated at 700 GPa using different numbers of
atoms in the ab initio MD simulations. Upper and lower limits of melting temperatures
from two-phase simulations are plotted with down and up triangles. The dashed line
indicates the middle of the upper and lower limits.

In standard PIMD simulations, the exchange of nuclei is neglected. To


estimate if the neglect of nuclear exchange effects has a significant effect
on the accuracy of the simulations, Chen et al. have also examined the
distributions of distances between beads in the same nucleus and between
beads in neighboring nuclei. Results at 1200 GPa (900 GPa) and 50 K are
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 215

Quantum Nuclear Effects 215

Figure 7.18 The angularly averaged pair-distribution function g(r), as explained above,
calculated using different intervals during a two-phase PIMD simulation of hydrogen at
700 GPa and different temperatures. Black (from 2 ps to 3 ps) and red solid (from 4 ps
to 5 ps) lines give g(r) from simulation at 100 K when the system solidifies. Green and
blue dashed lines are results at 125 K when the hydrogen melts.

reported in Fig. 7.19, where the distribution of distances between bead 1


and bead N /2+1 of the ring polymer in the 32-bead simulation (N = 32)
was compared with the distribution of H–H distances for the same bead.
The distances between bead 1 and bead N /2+1 of the ring polymer give the
solid curves and the H–H distances of the same bead give the dashed curves.
The absence of any significant overlap between the peaks of these two curves
in this highest pressure and lowest temperature simulation suggests that
exchange effects are unlikely to be an issue in the simulations reported. As
the pressure decreases from 1200 GPa to 900 GPa, the overlap of the two
curves becomes even smaller.
The last property of interest in the simulations is the superconductivity
feature of the hydrogen at this region of the phase diagram. It has been
widely reported that metallic hydrogen formed at these pressures could be
a high Tc superconductor [445, 446, 460]. Therefore, Chen et al. have also
calculated Tc for the I41/amd solid phase using the Allen–Dynes equa-
tion [461] and the QUANTUM-Espresso code [462]. They found that over
the entire pressure range examined (500–1200 GPa), Tc is predicted to be
around room temperature or above, which is consistent with previous pre-
dictions for I41/amd at these pressures in Ref. [446]. At 500 GPa, for exam-
ple, the conservative estimate of Tc is 358 K. To understand the physical
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 216

216 Computer Simulations of Molecules and Condensed Matter

Figure 7.19 Probability distribution of the distances between the first and
N /2+1(th)beads in the same atom (solid lines scale on left) and probability distribution
of the distances between the first bead in two neighboring atoms of different molecules
(dashed lines scale on right). Distributions are reported from a 32-bead two-phase PIMD
simulation at 1200 GPa and 50 K, the highest pressure and lowest temperature case
investigated, and for comparison, a simulation at 900 GPa and 50 K.

origin of this high Tc phase, we show details of its electronic and vibrational
properties at 500 GPa in Fig. 7.20. This reveals a high electronic density
of states (DOS) at the Fermi level (Fig. 7.20(a)), strong electron–phonon
coupling (Fig. 7.20(d)), and consequently a high value of λ (2.15 as shown
in Fig. 7.21) which leads to a high value of Tc within Bardeen–Cooper–
Schrieffer (BCS) theory [463].
These results were obtained using a dense q-point mesh (8×8×8). Their
convergence with the energy cut-off for the PAW pseudopotential is shown
in Fig. 7.21. A cut-off energy of 80 Ryd gives good convergence for both Tc
and the electron–phonon interaction parameter λ. In Fig. 7.21(b), we also
plot Tc versus μ∗ for the LDA and PBE functionals. We found that the LDA
and PBE results for Tc are similar, and both of them give Tc values which are
much higher than the melting temperature of the solid phase. In Ref. [323],
we have used μ∗ = 0.1 to obtain the value of Tc = 358 K reported. This value
of μ∗ is close to the value of 0.085 obtained from the Bennemann–Garland
formula [464] and larger than the value of 0.089 used in Ref. [446]. From
Fig. 7.21(b), it is clear that Tc decreases with increasing μ∗ . Considering
the fact that they have chosen a large value of μ∗ and that their value of Tc
is still much higher than the melting temperature of the solid phase, it is
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 217

Quantum Nuclear Effects 217

(a) (b)

(c) (d)

Figure 7.20 Electron and phonon properties of the I41/amd structure of solid hydrogen
at 500 GPa, with a volume of 2.28 Å3 for its primitive cell: (a) electronic band structure
and DOS, (b) Fermi surface in the Brillouin-zone, (c) phonon dispersion curves, and (d)
phonon DOS (dashed line) and α2 F (ω) (solid line).

reasonable to suppose that the atomic solid phase under the melting line is
superconducting. As the crystal melts well below room temperature, their
results also rule out a room temperature superconducting phase of solid
hydrogen at the pressures considered here and concluded that any room
temperature superconductor in this regime would have to be a liquid. For
more details concerning this study, please see Ref. [323].

7.5 Summary
In summary, we have discussed some extensions of the molecular simulation
methods as introduced in the earlier chapters to descriptions of the QNEs in
this chapter. The language we have used is the path integral representation
December 28, 2017 10:19 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ch07 page 218

218 Computer Simulations of Molecules and Condensed Matter

(a) (b)
Figure 7.21 Superconductivity of the I41/amd phase at 500 GPa: (a) superconduct-
ing critical temperatures Tc (circles) and the electron–phonon interaction parameter λ
(squares) as a function of the plane wave cut-off energy using the PBE functional, and
(b) superconducting critical temperatures Tc as a function of the effective Coulomb
interaction parameter μ∗ using the PBE (circle) and LDA (square) functionals.

of the quantum mechanics. Based on this language, the general theory


behind the statistical PIMD and PIMC methods as well as their extensions
to the dynamical regime were explained. A combination between the ther-
modynamic integration and PIMD methods was also presented, and some
examples for the practical simulations of these computational methods were
shown. These introductions, together with the computational methods for
the calculation of the electronic structures and simulations of the molecular
dynamics as presented in the earlier chapters, aim to set up a framework of
concepts concerning molecular simulations of molecules and condensed mat-
ters. We sincerely hope this framework can help graduate students working
on computer simulations of molecules and condensed matter find the proper
method for tackling the problems of interest.
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-appa page 219

Appendix A: Useful Mathematical


Relations

In this appendix, we provide a summary of useful mathematical formulae


that are used throughout the text.

A.1 Spherical Harmonics


Definition: In the Condon Shortley phase convention, the spherical harmon-
ics are defined as

m 2l + 1 (l − m)! m
Yl,m (θ, φ) ≡ (−1) P (cos θ)eimφ , (A.1)
4π (l + m)! l

where Plm (x) is the corresponding Legendre polynomial (see Ref. [465]).
Recurrence relations

1
Y0,0 (θ, φ) = , (A.2a)


3
Y1,0 (θ, φ) = cos(θ), (A.2b)


3
Y1,1 (θ, φ) = − sin(θ)eiφ , (A.2c)

Y1,−1 (θ, φ) = −Y1,1 (θ, φ), (A.2d)

2l + 1
Yl,l (θ, φ) = − sin(θ)eiφ Yl−1,l−1 , (A.2e)
2l

219
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-appa page 220

220 Computer Simulations of Molecules and Condensed Matter


(2l − 1)(2l + 1)
Yl,m (θ, φ) = cos(θ)Yl−1,m (θ, φ)
(l − m)(l + m)

(l − 1 + m)(l − 1 − m)(2l + 1)
− Yl−2,m (θ, φ). (A.2f)
(2l − 3)(l − m)(l + m)

Conjugation

Yl,−m (θ, φ) = (−1)m Yl,m



(θ, φ). (A.3)

Inversion

Yl,m (r̂) = (−1)l Yl,m (−r̂). (A.4)

A.2 Plane Waves


Rayleigh expansion

∞ 
 +λ
eiG·r = 4π iλ jλ (Gr)Yλ,μ

(T −1 Ĝ)Yλ,μ (T −1 r̂)
λ=0 μ=−λ
(A.5)
∞ 
 +λ
= 4π iλ jλ (Gr)Yλ,μ (T −1 Ĝ)Yλ,μ

(T −1 r̂).
λ=0 μ=−λ

A.3 Fourier Transform


Definition: We use the following convention for the time-frequency Fourier
Transform
 ∞
F (ω) = F (t)eiωt dt,
−∞
 ∞
(A.6)
1 −iωt
F (t) = F (ω)e dω.
2π −∞

Imaginary axes: The Fourier transform between imaginary axes work


like its counterpart on the real axes, except that additional factors of ±i
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-appa page 221

Appendix A: Useful Mathematical Relations 221

have to be included
 ∞
F (iω) = −i F (iτ )e−iωτ dτ ,
−∞
 ∞
(A.7)
i iωτ
F (iτ ) = F (iω)e dω.
2π −∞

A.4 Spherical Coordinates


Derivatives
 
±iφ ∂ e±iφ ∂ i ∂
∂x ± i∂y = sin θe + cos θ ± ,
∂r r ∂θ sin θ ∂φ
(A.8)
∂ 1 ∂
∂z = cos θ − sin θ .
∂r r ∂θ

A.5 The Step(Heaviside) Function


Definition:

1, r ∈ interstitial,
Θ(r) = (A.9)
0, r∈/ interstitial.
Since the step function Θ(r) has the periodicity of the lattice, we may
expand it in a Fourier series as

Θ(r) = Θ̃G eiG·r , (A.10)
G

where Θ̃G can be calculated analytically, giving


 4πr3
1 − a 3Ωa , G = 0,
Θ̃G = (A.11)
4π 2 iG·ra
− ΩG a j1 (Gra ) ra e , G = 0.
b2530   International Strategic Relations and China’s National Security: World at the Crossroads

This page intentionally left blank

b2530_FM.indd 6 01-Sep-16 11:03:06 AM


December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-appb page 223

Appendix B: Expansion
of a Non-Local Function

In calculations of the non-local self-energy, many other non-local operators,


such as the bare Coulomb potential, the polarizability, and the screened
Coulomb potential, etc. need to be calculated in a matrix form for the basis
set chosen. In this appendix, we discuss how these non-local operators are
expanded.
We assume f (r1 , r2 , τ ) is the general form of this non-local operator,
where τ can be either a time or a frequency coordinate, or nothing. This
non-local operator possesses a lattice translational symmetry with respect
to R (e.g. v, W , P , ε, etc.)

f (r1 + R, r2 + R, τ ) = f (r1 , r2 , τ ). (B.1)

To calculate a function of this type, we use the expansion in a complete


set of Bloch functions {χqi (r)} in the following way:
⎧  BZ 
⎨f (r1 , r2 , τ ) = BZ q  q ∗
q q i,j χi (r1 )fi,j (q, q , τ )(χj (r1 )) ,
(B.2)
⎩f (q, q , τ ) =   (χq (r ))∗ f (r , r , τ )χq (r )dr dr .
i,j V V i 1 1 2 j 2 2 1

Since χqi (r1 ) is a Bloch function (χqi (r − R) = e−iq·R χqi (r)) normalized to
unity in the crystal with volume V , the matrix element fi,j (q, q , τ ) can be
evaluated by
 

fi,j (q, q , τ ) = (χqi (r1 ))∗ f (r1 , r2 , τ )χqj (r2 )dr2 dr1
V V
 
= (χqi (r1 − R))∗ f (r1 − R, r2 − R − R , τ )
R,R Ω Ω

223
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-appb page 224

224 Computer Simulations of Molecules and Condensed Matter


×χqj (r2 − R − R )dr2 dr1
    
= eiq·R (χqi (r1 ))∗ f (r1 , r2 − R , τ )e−iq ·R e−iq ·R
R Ω R Ω


×χqj (r2 )dr2 dr1
  

= ei(q−q )·R (χqi (r1 ))∗f (r1 , r2 − R , τ )
R Ω R Ω

  
×e−iq ·R χqj (r2 )dr2 dr1
  
= Nc δq,q (χqi (r1 ))∗ f (r1 , r2 − R , τ )
Ω Ω R

×e−iq·R χqj (r2 )dr2 dr1 , (B.3)

where we have made use of this relation for the Bravais lattice
 
e−i(q−q )·R = Nc δq,q . (B.4)
R

Nc is the number of cells in the crystal.


With these treatments, the expansion of Eq. (B.2) is written as
⎧  
⎨f (r1 , r2 , τ ) = BZ
q
q q ∗
i,j χi (r1 )fi,j (q, τ )(χj (r2 )) ,
(B.5)
⎩f (q, τ ) =   (χq (r ))∗ f (r , r , τ )χq (r )dr dr ,
i,j V V i 1 1 2 j 2 2 1

where the integration must be done on the whole volume of the crystal, or
⎧  
⎨f (r1 , r2 , τ ) = BZ
q
q q ∗
i,j χi (r1 )fi,j (q, τ )(χj (r2 )) ,
⎩f (q, τ ) = N   (χq (r ))∗  f (r , r − R, τ )e−iq·R χq (r )dr dr .
i,j c Ω Ω i 1 R 1 2 j 2 2 1
(B.6)
The integration must be performed only on the Wigner–Seitz cell.
If we have a product of operators, say

h(r1 , r2 , τ ) = f (r1 , r3 , τ )g(r3 , r2 , τ )dr3 , (B.7)
V
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-appb page 225

Appendix B: Expansion of a Non-Local Function 225

then, according to Eq. (B.5), the expansion of h in the set of functions


{χqi (r)} is
 

hi,j (q, τ ) = [χqi (r1 )] h(r1 , r2 , τ )χqj (r2 )dr2 dr1
V V
  


= [χqi (r1 )] f (r1 , r3 , τ )g(r3 , r2 , τ )dr3 χqj (r2 )dr2 dr1 .
V V V
(B.8)
We can now use the second line of Eq. (B.5) for f and g and the
orthogonality of the basis to get
  

q ∗
hi,j (q, τ ) = [χi (r1 )] f (r1 , r3 , τ )g(r3 , r2 , τ )dr3 χqj (r2 )dr2 dr1
V V V
   BZ 

 q
= [χqi (r1 )] χl 1 (r1 )fl,m (q1 , τ ) [χqm1 (r3 )]∗
V V V q l,m
1

BZ 

 ∗
× χqn2 (r3 )gn,p (q2 , τ ) χqp 2 (r2 ) dr3 χqj (r2 )dr2 dr1
q2 n,p

BZ   
BZ 


∗ q
= [χqi (r1 )] χl 1 (r1 )dr1 fl,m (q1 , τ )
q1 q2 l,m n,p V


× [χqm1 (r3 )] χqn2 (r3 )dr3 gn,p (q2 , τ )
V


q ∗
× χp 2 (r2 ) χqj (r2 )dr2
V

BZ 
 BZ  
= δ(q, q1 )δi,l fl,m (q1 , τ )
q1 q2 l,m n,p

×δ(q1 , q2 )δm,n gn,p (q2 , τ )δ(q2 , q)δpj . (B.9)


With this, one arrives at the expected expression

hi,j (q, τ ) = fi,l (q, τ )gl,j (q, τ ). (B.10)
l
b2530   International Strategic Relations and China’s National Security: World at the Crossroads

This page intentionally left blank

b2530_FM.indd 6 01-Sep-16 11:03:06 AM


December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-appc page 227

Appendix C: The Brillouin-Zone


Integration

In Sec. 4.4.4, we introduced the ideas of the q-dependent linear tetrahedron


method for the calculation of the polarization matrix. Compared with the
one introduced by Rath and Freeman in Ref. [223], the frequency depen-
dence and the variation of the operator to be integrated within each tetra-
hedron are included. For a clear illustration, we begin with the formula
of the traditional linear tetrahedron method and the idea of isoparametric
transformation. Then we extend these ideas to the q-dependent case and
illustrate the different configurations of the possible integration region. The
frequency dependence is discussed in the end.

C.1 The Linear Tetrahedron Method


The task of this Brillouin-zone integration is to calculate the average expec-
tation value of an operator satisfying the form

1 
X = Xn (k)f (n (k))dk, (C.1)
VG n VG

where
Xn (k) = ϕn (k)|X|ϕn (k). (C.2)
This Xn (k) is the expectation value of this operator on the state (n, k). VG
is the volume of the reciprocal unit cell. f () is the Fermi function. An
exact evaluation of Eq. (C.1) requires calculating the expectation value of
this operator over all its occupied states, including an infinite number of
k points in the Brillouin zone. In practice, this average expectation value
is determined from a set of sample points in the Brillouin zone; each has

227
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-appc page 228

228 Computer Simulations of Molecules and Condensed Matter

Figure C.1 The 2D sketch of the BZ in the tetrahedron method. In this case, the space
is divided into a list of triangles. One triangle is related to another by a vector q.

a certain weight addressing the integration of Eq. (C.1) over the region
around it.
In the tetrahedron method, this is obtained by dividing the Brillouin
zone into a set of tetrahedra using a grid (as shown in Fig. C.1 for the 2D
case). The values of Xn (k) are calculated on the discrete set of vectors ki
at the vertices of all these tetrahedra, namely the grid points. A function
X̄n (k) obtained by linearly interpolating the function Xn (k) within the
tetrahedra using its expectation values on the vertices can be written as a
superposition of functions wi (k), such that

X̄n (k) = Xn (ki )wi (k), (C.3)
i

where wi (kj ) = δij and it is linear within the corresponding tetrahedron


and zero outside of it. Now, replacing Xn (k) in Eq. (C.1) by its linear
approximation, one has

∼ 1 
X = X̄n (k)f (n (k))dk
VG n VG
 
1 
= Xn (ki )wi (k)f (n (k))dk
VG n VG i
 
1
= Xn (ki ) wi (k)f (n (k))dk.
n i
VG VG
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-appc page 229

Appendix C: The Brillouin-Zone Integration 229

Defining

1
wn,i = wi (k)f (n (k))dk, (C.4)
VG VG

one can write the average expectation value of X in Eq. (C.1) as a weighted
sum over the discrete set of k points

X = Xn (ki )wn,i . (C.5)
i,n

Since wi (k) is zero for all {kj } except ki , we can rewrite the weights as

1  
wn,i = wi (k)f (n (k))d3 k = 1T
wn,i , (C.6)
VG
Ti VT Ti

where Ti means that the sum runs only over those tetrahedra containing
ki as one of its vertices, and one has defined

1T 1
wn,i = wi (k)f (n (k))d3 k.
VG
VT

With this, it is clear that the integration in Eq. (C.1) can be approx-
imated by a sum of the form in Eq. (C.5) where wn,i can be calculated
by summing its contribution from each tetrahedron containing this ki as a
vertex. The next job is to define the function wi (k) in order to calculate
these wn,i . For this, one needs the isoparametric transformation.

C.1.1 The Isoparametric Transfromation

In Eq. (C.3), the function behavior is approximated inside each tetrahedron


by a linear interpolation between the function values at the vertices. Let F
be such a function, and x, y and z be the coordinates, then

F = Ax + By + Cz + D, (C.7)

where the constants A, B, C and D are to be determined. Substituting


x = xi , y = yi and z = zi where i = 0, 1, 2, 3 label the vertices, the values
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-appc page 230

230 Computer Simulations of Molecules and Condensed Matter

of Fi at the vertices (which are known) can be written as

Fi = Axi + Byi + Czi + D. (C.8)

Clearly, Eq. (C.8) for i = 0 can be used to eliminate the constant D.


Then we have

F − F0 = A(x − x0 ) + B(y − y0 ) + C(z − z0 ). (C.9)

The constants A, B and C are determined by solving the system of


equations:

F1 − F0 =A(x1 − x0 ) + B(y1 − y0 ) + C(z1 − z0 ),


F2 − F0 =A(x2 − x0 ) + B(y2 − y0 ) + C(z2 − z0 ), (C.10)
F3 − F0 =A(x3 − x0 ) + B(y3 − y0 ) + C(z3 − z0 ),

with solution
⎛ ⎞ ⎛ ⎞−1 ⎛ ⎞
A x1 − x0 y1 − y0 z1 − z0 F1 − F0
⎜ ⎟ ⎜ ⎟ ⎜F − F ⎟
⎝B ⎠ = ⎝x2 − x0 y2 − y0 z2 − z0 ⎠ ⎜ ⎝ 2 0⎟ .
⎠ (C.11)
C x3 − x0 y3 − y0 z3 − z0 F3 − F0

If one defines a coordinate (ξ, η, ζ) inside this tetrahedron, with each


vertex 0, 1, 2, 3 having coordinates (0,0,0), (1,0,0), (0,1,0), (0,0,1), respec-
tively, the function F can be linearly interpolated as

F − F0 = ξ(F1 − F0 ) + η(F2 − F0 ) + ζ(F3 − F0 ). (C.12)

Putting Eq. (C.10) into the above equation, we have


⎛ ⎞
F1 − F0

⎜ ⎟
F − F0 = ξ η ζ ⎝F2 − F0 ⎠
F3 − F0
⎛ ⎞⎛ ⎞
x1 − x0 y1 − y0 z1 − z0 A

⎜ ⎟⎜ ⎟
= ξ η ζ ⎝x2 − x0 y2 − y0 z2 − z0 ⎠ ⎝B ⎠ .
x3 − x0 y3 − y0 z3 − z0 C
(C.13)
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-appc page 231

Appendix C: The Brillouin-Zone Integration 231

On the other hand, Eq. (C.9) can be written as

⎛ ⎞
A

⎜ ⎟
F − F0 = x − x0 y − y0 z − z0 ⎜ ⎟
⎝B ⎠ . (C.14)
C

Comparing Eq. (C.14) with Eq. (C.13), we have

x − x0 = ξ(x1 − x0 ) + η(x2 − x0 ) + ζ(x3 − x0 ),

y − y0 = ξ(y1 − y0 ) + η(y2 − y0 ) + ζ(y3 − y0 ), (C.15)

z − z0 = ξ(z1 − z0 ) + η(z2 − z0 ) + ζ(z3 − z0 ).

Combining Eq. (C.15) with Eq. (C.12), we see that the same expression
holds for the function F as well as for the coordinates x, y, and z. This
coordinate transition from outside the tetrahedron to inside the tetrahedron
is called as an isoparametric transformation. The functions wi (k) used in
Eq. (C.3) can be simply written as

w0 (ξ, η, ζ) = 1 − ξ − η − ζ,

w1 (ξ, η, ζ) = ξ,
(C.16)
w2 (ξ, η, ζ) = η,

w3 (ξ, η, ζ) = ζ

in terms of this internal coordinates. The energy eigenvalue of the state


(n, k) with the coordinate (ξ, η, ζ) inside this tetrahedron is linearly inter-
polated as

n (ξ, η, ζ) = (n,1 − n,0 )ξ + (n,2 − n,0 )η + (n,2 − n,0 )ζ + n,0 ,


(C.17)

where n,i is the energy eigenvalue on the vertex i.


December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-appc page 232

232 Computer Simulations of Molecules and Condensed Matter

C.1.2 Integrals in One Tetrahedron

The integral of any function F inside one tetrahedron, after applying the
isoparametric transformation, is given by

F (x, y, z)f (n (x, y, z))dxdydz
VT
 1 1−ζ  1−ζ−η
= ξ(F1 − F0 ) + η(F2 − F0 )
0 0 0
 
 ∂(xyz) 
+ζ(F3 − F0 ) + F0   f (n (ξ, η, ζ))dξdηdζ,
∂(ξηζ) 
(C.18)
 
 ∂(xyz) 
where VT is the volume of the tetrahedron and   is the Jacobian
∂(ξηζ) 
determinant given by
   
 ∂x ∂x ∂x   
   ∂ξ ∂η ∂ζ  x1 − x0 y1 − y0 z1 − z0 
 ∂(xyz)   ∂y ∂y ∂y   
     
 ∂(ξηζ)  =  ∂ξ ∂η ∂ζ  = x2 − x0 y2 − y0 z2 − z0  . (C.19)
 ∂z ∂z ∂z   
 ∂ξ ∂η ∂ζ  x3 − x0 y3 − y0 z3 − z0 

This is just the volume of a parallelepiped whose sides are given by those
of the tetrahedron, clearly
 
 ∂(xyz) 
 
 ∂(ξηζ)  = 6VT . (C.20)

Then Eq. (C.18) is just



F (x, y, z)dxdydz
VT
 1 1−ζ 1−ζ−η
= 6VT ξ(F1 − F0 ) + η(F2 − F0 ) + ζ(F3 − F0 ) + F0
0 0 0

·f (n (ξ, η, ζ))dξdηdζ. (C.21)

C.1.3 The Integration Weights

Let us take one of the tetrahedra with its four vertices denoted as 0, 1, 2,
and 3. Using the wi (k) and n (k) defined in Eq. (C.16) and Eq. (C.17), one
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-appc page 233

Appendix C: The Brillouin-Zone Integration 233

can calculate the integration weights on these vertices. If the four energies
are below the Fermi energy, the occupation is identically one and we have
    
1T 6VT 1 1−ζ 1−ζ−η 6VT 1 1−ζ
wn,i = ζdξdηdζ = ζ(1 − ζ − η)dηdζ
VG 0 0 0 VG 0 0
  
6VT 1 1 2 3VT 1 2 1 VT
= ζ(1 − ζ) dζ = − + = .
VG 0 2 VG 2 3 4 4VG
(C.22)
Let us now take the case where only n,0 < F and, for the sake of
simplicity, n,3 > n,2 > n,1 > n,0 , then the integration limits are changed,
and one gets
 F −n,0  F −n,0 −ζ(n,3 −n,0 )  F −n,0 −ζ(n,3 −n,0 )−η(n,2 −n,0 )

1T 6VT n,3 −n,0 n,2 −n,0 n,1 −n,0


wn,3 =
VG 0 0 0

× ζdξdηdζ
VT (F − n,0 )4
= . (C.23)
4VG (n,1 − n,0 )(n,2 − n,0 )(n,3 − n,0 )2
A similar calculation for the rest of the vertices leads to
1T VT (F − n,0 )4
wn,2 = ,
4VG (n,1 − n,0 )(n,2 − n,0 )2 (n,3 − n,0 )

1T VT (F − n,0 )4
wn,1 = 2
,
4VG (n,1 − n,0 ) (n,2 − n,0 )(n,3 − n,0 )

1T VT (F − n,0 )3 1T 1T 1T
wn,0 = − wn,1 − wn,2 − wn,3 .
VG (n,1 − n,0 )(n,2 − n,0 )(n,3 − n,0 )
(C.24)
The last line in Eq. (C.24) can be calculated using w0 (k) = 1−ξ−η−ζ =
wt − w1 (k) − w2 (k) − w3 (k), where wt means the total weight over this
tetrahedron. Expressions for the remaining cases can be found in Ref. [224].
Since these vertices are also sample points in the grid mesh, the integration
weight on each grid point can be calculated from Eq. (C.6).

C.2 Tetrahedron Method for q-Dependent


Brillouin-Zone Integration
If one wants to calculate the mean value of a q-dependent operator, the
situation becomes more complicated. In this section, we discuss the case
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-appc page 234

234 Computer Simulations of Molecules and Condensed Matter

when the expectation value of this operator satisfies



1 
X(q) = Xnn (k, q)f [n (k)] (1 − f [n (k − q)]) d3 k,
VG  VGn,n
(C.25)

where

Xn,n (k, q) = ϕn (k)|X(q)|ϕn (k − q). (C.26)

To evaluate this operator, one needs to know Xnn (k, q) on each k point
in the Brillouin zone in principle. In practice, again, this is obtained by
calculating the expectation value of this operator on a set of sample points
weighted by a certain factor. In addition to n (ki ) and ϕn (ki ) on the set of
sample points {ki }, one also needs to know n (ki − q) and ϕn (ki − q) on
another set of sample points {ki − q}.
In Ref. [212], an even division of the Brillouin zone along each axis is
made. Then, one takes the q vector from this mesh. With this treatment,
the meshes of ki and ki − q overlap totally with each other. One just needs
to know the eigenwavefunctions and the energy eigenvalues in one mesh. A
2D sketch for the k-mesh is shown in Fig. C.1.
Using this grid, the Brillouin zone is divided into a set of tetrahedra.
The expectation values of the function Xn,n (k, q) are calculated on the
vertices of these tetrahedra, namely, the grid points, giving Xn,n (ki , q).
Following the same procedure as in the previous section, we interpolate
the function Xn,n (k, q) linearly within each tetrahedron using

X̄n,n (k, q) = Xn,n (ki , q)wi (k, q), (C.27)
i

where wi (kj , q) = δi,j , and it is a linear function. Since the integration is


over the vector k and this wi (kj , q) is only a function of the coordinates of
k for a fixed q, it is easy to see that we can get rid of the q dependence.
Eq. (C.27) becomes

X̄n,n (k, q) = Xn,n (ki , q)wi (k). (C.28)
i

For the expectation value, we get



X(q) = Xn,n (ki , q)wn,n ,i (q), (C.29)
i,n,n
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-appc page 235

Appendix C: The Brillouin-Zone Integration 235

with

1
wn,n ,i (q) = wi (k)f [n (k)] (1 − f [n (k − q)]) d3 k. (C.30)
VG VG

To calculate the weights, following the steps as in the previous section,


one has

1T
wn,n ,i (q) = wn,n  ,i (q), (C.31)
Ti

where

1T 1
wn,n ,i (q) = wi (k)f [n (k)] (1 − f [n (k − q)]) d3 k. (C.32)
VG
VT

Ti runs over all the tetrahedra in which the sample point ki serves as a
vertex.

C.2.1 Isoparametric Transformation

Now, we perform the isoparametic transformation to calculate the inte-


gration of Eq. (C.32) in one tetrahedron. If we denote the vertices of this
tetrahedron as 0, 1, 2, and 3, respectively, we have

w0 (k) = w0 (ξ, η, ζ) = 1 − ξ − η − ζ,
w1 (k) = w1 (ξ, η, ζ) = ξ,
w2 (k) = w2 (ξ, η, ζ) = η,
w3 (k) = w3 (ξ, η, ζ) = ζ,
(C.33)
n (k) = n (ξ, η, ζ) = ξ(n,1 − n,0 ) + η(n,2 − n,0 )

+ ζ(n,3 − n,0 ) + n,0 ,


n (k − q) = n (ξ, η, ζ) = ξ(n ,1 − n ,0 )
+ η(n ,2 − n ,0 ) + ζ(n ,3 − n ,0 ) + n ,0 ,

where we have used the shorthand notation n,i and n ,i to represent the
energy eigenvalues of the state (n, k) and (n , k − q) on the vertices of this
tetrahedron.
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-appc page 236

236 Computer Simulations of Molecules and Condensed Matter

Then, the general formula for the contribution of one tetrahedron to


the weight is
1T
wn,n ,i (q)

 1 1−ζ  1−ζ−η
6VT
= wi (ξ, η, ζ)Θ[F −ξ(n,1 −n,0 )
VG 0 0 0
(C.34)
− η(n,2 −n,0 )−ζ(n,3 −n,0 )−n,0]
× Θ[ξ(n ,1 − n ,0 ) + η(n ,2 − n ,0 ) + ζ(n ,3 − n ,0 )
+ n ,0 − F ]dξdηdζ,

where Θ is the step function to address the Fermi function in Eq. (C.32).

C.2.2 The Integration Region

From Eq. (C.34), we see that the Θ functions determine the integration
region within this tetrahedron. For insulators and semiconductors, this
region is either the full tetrahedron or zero. For metals, the situation
becomes more complicated. If the integration region does not include the
full tetrahedron, that is, the n,i s are smaller or bigger than F , the Fermi
surface represented by the first Θ function in Eq. (C.34) will intersect with
this tetrahedron, leading to only part of it satisfying the condition, the
first Θ function equals one. Another case is when the n ,i are smaller or
bigger than F — the Fermi surface represented by the second Θ function
in Eq. (C.34) will intersect with this tetrahedron, leading to only part of it
satisfying the condition, and the second Θ function equals one. If neither
of these cases happen, the integration region is either the full tetrahedron
or zero. Otherwise, the integration region is determined by the intersection
of these Fermi surfaces with this tetrahedron (Fig. 4.7 shows one example
when both of them intersect with this tetrahedron).
There are in total nine different configurations for this region. They
are shown in Fig. C.2 except for the simplest case of a tetrahedron. All
of them can be subdivided into smaller tetrahedra. Then, we perform one
further isoparametric transformation inside each of these small tetrahedra.
The weight on each of its vertices is
 1 1−z 1−y−z
6VT VST VST
w0 = (1 − x − y − z)dxdydz = ,
VG VT 0 0 0 4VG (C.35)
w1 = w2 = w3 = w0 ,
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-appc page 237

Appendix C: The Brillouin-Zone Integration 237

Figure C.2 The configurations for the region to be integrated. How these regions are
decomposed into the principle units of the small tetrahedra is shown by the red lines in
the graph.

where VST is the volume of the small tetrahedron, and wi (i = 0, 1, 2, 3)


represents the weight on each vertex. We further distribute these weights
linearly into the vertices of the big tetrahedron. Assuming the coordinates
of one vertex of this small tetrahedron is (ξ1 , η1 , ζ1 ) in the big tetrahedron
before the second parametric transformation, the integration weight on this
point will be distributed with the ration 1 − ξ1 − η1 − ζ1 , ξ1 , η1 , and ζ1 to
the vertices 0, 1, 2, and 3 of the big tetrahedron.

C.2.3 Polarizability

As has already been mentioned in Sec. 4.4.4, for the polarizability, we can-
not assume both the energies and the integrand to be simultaneously linear
in the coordinates of the tetrahedron. In this case, we have to include the
energy-dependent factor of Eq. (4.50) into the analytical integration. In
Sec. 4.4.5, we have discussed the frequency integrations in the GW calcula-
tions, where we pointed out that we calculate all the frequency-dependent
properties on the imaginary frequency axis. The polarizability is such a
property. In this section, we will discuss the integration weight of the polar-
izability on both the real and imaginary frequency axes. The latter is the
one used in the GW calculation. The former can be used to calculate the
macroscopic dielectric constant.
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-appc page 238

238 Computer Simulations of Molecules and Condensed Matter

C.2.3.1 Polarizability on the Real Frequency Axis

On the real frequency axis, the polarization matrix is

BZ occ unocc
Nc    i i ∗
Pi,j (q, ω) = Mn,n (k, q)[Mn,n (k, q)]
 n 
k n
 
1 1
· − .
ω − n ,k−q + n,k + iη ω − n,k + n ,k−q − iη


(C.36)

We define the weight as



1T
wn,n ,i (q, ω) = wn,n  ,i (q, ω), (C.37)
Ti

where

1T 1
wn,n ,i (q, ω) = wi (k)f [n (k)](1 − f [n (k − q)])
VG VT
 
1 1
× − d3 k.
ω − n (k − q) + n (k) + iη ω − n (k) + n (k − q) − iη
(C.38)

Following the procedures in Sec. C.2.2, the weight on each vertex of the
small tetrahedron is calculated by
 1 1−z  1−y−z
6VST 2(1 − x − y − z)
w0 = dxdydz,
VG 0 0 0 ω 2 − (xΔ1,0 + yΔ2,0 + zΔ3,0 + Δ0 )2
 1 1−z  1−y−z
6VST 2x
w1 = dxdydz,
VG 0 0 0 ω2 − (xΔ1,0 + yΔ2,0 + zΔ3,0 + Δ0 )2
 1 1−z  1−y−z
6VST 2y
w2 = dxdydz,
VG 0 0 0 ω 2 − (xΔ1,0 + yΔ2,0 + zΔ3,0 + Δ0 )2
 1 1−z  1−y−z
6VST 2z
w3 = 2 − (xΔ 2
dxdydz.
VG
0 0 0 ω 1,0 + yΔ 2,0 + zΔ3,0 + Δ0 )
(C.39)
Here, Δi = n ,i − n,i and Δi,j = Δi − Δj .
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-appc page 239

Appendix C: The Brillouin-Zone Integration 239

It is more complicated to solve the first equation in Eq. (C.39) analyt-


ically compared to the other three due to the presence of three variables in
the numerator. So, we solve the other three respectively and then calculate
the total integration weight over this tetrahedron with

 1 1−z  1−y−z
6VST 2
wt = dxdydz.
VG 0 0 0 ω2
− (xΔ1,0 + yΔ2,0 + zΔ3,0 + Δ0 )2
(C.40)
The correponding w0 is then calculated from w0 = wt − w1 − w2 − w3 .
Even with this treatment, solving this analytical integration is very
complicated. We use Mathematica to treat it. There exists a general solu-
tion. To restrict the size of this appendix, we just list that of wt here,
which is the simplest case due to the absence of variables in the numerator
in Eq. (C.39):

f (ω) = (ω − Δ3 )3 Δ21,0 Δ2,0 Δ22,1 ln[|ω − Δ3 |]

−(ω − Δ2 )3 Δ21,0 Δ3,0 Δ23,1 ln[|ω − Δ2 |],

f (ω) = f (ω) + [Δ1,0 Δ2,1 (ω − Δ3 )

−(ω − Δ0 )Δ2,1 Δ3,1 + (ω − Δ2 )Δ1,0 Δ3,1 ]

×Δ2,0 Δ3,0 Δ3,2 (ω − Δ1 )2 ln[|ω − Δ1 |],

f (ω) = f (ω) + (ω − Δ0 )3 Δ22,1 Δ3,2 Δ23,1 ln[|ω − Δ0 |]

−(ω − Δ1 )2 Δ1,0 Δ2,0 × Δ2,1 Δ3,0 Δ3,1 Δ3,2 ,

f (ω)
f (ω) = ,
6Δ21,0 Δ2,0 Δ22,1 Δ3,0 Δ23,1 Δ3,2

6VST
wt = [f (ω) + f (−ω)] (C.41)
VG

(this equation is written following the programming rules because it is too


long). In this equation, it is required that ω = Δi and Δi,j = 0. Δi,j = 0 is
required because of this analytical solution. ω = Δi is required because the
denominators in Eq. (C.39) and Eq. (C.40) cannot be zero. When these con-
ditions are not fulfilled, we use Mathematica to get the analytical solution
of that specific case, respectively.
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-appc page 240

240 Computer Simulations of Molecules and Condensed Matter

C.2.3.2 Polarizability on the Imaginary Frequency Axis

The polarization matrix of our calculation on the imaginary frequency


axis is
Pi,j (q, ω)
BZ occ unocc
Nc    i i ∗ −2(n ,k−q − n,k )
= Mn,n (k, q)[Mn,n (k, q)] .
 n 
ω 2 + (n ,k−q − n,k )2
k n
(C.42)
In this case, the procedure is essentially the same as the above, except
for the fact that the weight on the vertices of each small tetrahedron is
calculated with
  
6VST 1 1−z 1−y−z 2(1 − x − y − z)
w0 = 2 + (xΔ 2
dxdydz,
VG 0 0 0 ω 1,0 + yΔ2,0 + zΔ3,0 + Δ0 )
  
6VST 1 1−z 1−y−z 2x
w1 = dxdydz,
VG 0 0 0 ω + (xΔ1,0 + yΔ2,0 + zΔ3,0 + Δ0 )2
2

  
6VST 1 1−z 1−y−z 2y
w2 = 2 + (xΔ 2
dxdydz,
VG 0 0 0 ω 1,0 + yΔ 2,0 + zΔ3,0 + Δ0 )
  
6VST 1 1−z 1−y−z 2z
w3 = 2 + (xΔ 2
dxdydz.
VG 0 0 0 ω 1,0 + yΔ 2,0 + zΔ3,0 + Δ0 )
(C.43)
Again, we introduce wt as
  
6VST 1 1−z 1−y−z 2
wt = 2 + (xΔ 2
dxdydz
VG 0 0 0 ω 1,0 + yΔ 2,0 + zΔ3,0 + Δ0 )
(C.44)
to avoid solving the first equation of Eqs. (C.43) directly. Its general solu-
tion is
f (ω) = 2(ω − Δ20 )Δ0,1 Δ0,2 Δ0,3 Δ1,2 Δ1,3 Δ2,3 ,

f (ω) = f (ω) + 2ω[3Δ40 − ω 2 (Δ1 Δ2 + Δ2 Δ3 + Δ3 Δ1 )

−3Δ20 (ω 2 + Δ2 Δ3 + Δ1 Δ2 + Δ1 Δ3 )
+2Δ0 (ω 2 Δ2 + ω 2 Δ3 + Δ1 ω 2 + 3Δ1 Δ2 Δ3 )]

×Δ1,2 Δ1,3 Δ2,3 arctan[Δ0 /ω],


December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-appc page 241

Appendix C: The Brillouin-Zone Integration 241

f (ω) = f (ω) + 2ω(ω 2 − 3Δ21 )Δ20,2 Δ20,3 Δ2,3 arctan[Δ1 /ω],

f (ω) = f (ω) − 2ω(ω 2 − 3Δ22 )Δ20,1 Δ20,3 Δ1,3 arctan[Δ2 /ω],

f (ω) = f (ω) + 2ω(ω 2 − 3Δ23 )Δ20,1 Δ20,2 Δ1,2 arctan[Δ3 /ω],

f (ω) = f (ω) + [Δ40 (Δ1 + Δ2 + Δ3 ) − 3ω 2 Δ1 Δ2 Δ3

−2Δ30 (3ω 2 + Δ2 Δ3 + Δ1 Δ2 + Δ1 Δ3 )

+3Δ20 (ω 2 Δ1 + ω 2 Δ2 + ω 2 Δ3 + Δ1 Δ2 Δ3 )]

×Δ1,2 Δ1,3 Δ2,3 ln[ω 2 + Δ20 ],

f (ω) = f (ω) + Δ1 (3ω 2 − Δ21 )Δ20,2 Δ20,3 Δ2,3 ln[ω 2 + Δ21 ],

f (ω) = f (ω) + Δ2 (Δ22 − 3ω 2 )Δ20,1 Δ20,3 Δ1,3 ln[ω 2 + Δ22 ],

f (ω) = f (ω) − Δ3 (Δ23 − 3ω 2 )Δ20,1 Δ20,2 Δ1,2 ln[ω 2 + Δ23 ],

6VST f (ω)
wt = . (C.45)
VG 6Δ0,1 Δ0,2 Δ20,3 Δ1,2 Δ1,3 Δ2,3
2 2

Again, in this equation, it is required that Δi,j = 0. When this condition


is not fulfilled, we use Mathematica to get the analytical solution of that
case again specifically, the same way we do in the above section.
b2530   International Strategic Relations and China’s National Security: World at the Crossroads

This page intentionally left blank

b2530_FM.indd 6 01-Sep-16 11:03:06 AM


December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-appd page 243

Appendix D: The Frequency


Integration

In this section, we show how to perform the integration on the frequency


axis in Eq. (4.57).
This integration is peaked around ω  = ω when nk+q is small. To
handle this problem, one can add and subtract the term
 c
1 ∞ (n ,k+q − iω)Wi,j (q, iω)  1 c
2 2
dω = sgn(n ,k+q )Wi,j (q, iω). (D.1)
π 0 (iω − n ,k+q ) + ω 2
Then we have
  c 
1 ∞(
n ,k+q −iω) Wi,j (q, iω  )−Wi,jc
(q, iω)
I=
π 0 (iω − n ,k+q )2 + ω 2
1
× dω  + sgn(n ,k+q )Wi,j
c
(q, iω). (D.2)
2
The integrand is now smooth and a Gaussian quadrature may be used.
To solve the semi-infinite integral of Eq. (D.2) which has the form
 ∞
I= f (ω)dω, (D.3)
0

we split it into (following Ref. [225])

I = I1 + I2 , (D.4a)
 ω0
I1 = f (ω)dω, (D.4b)
0
 ∞
I2 = f (ω)dω. (D.4c)
ω0

243
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-appd page 244

244 Computer Simulations of Molecules and Condensed Matter

ω0
For I1 , we make the change of variables u = 2ω/ω0 −1 and thus, dω = du.
2
Then we have
 ω0 
ω0 1
I1 = f (ω)dω = f [(u + 1)ω0 /2]du, (D.5)
0 2 −1
which can be solved by standard Gauss–Legendre quadrature. For I2 ,
we make the change of variables u = 2ω0 /ω − 1 and thus, dω =
2ω0
− du. Then we have
(u + 1)2
 ∞  1  
2ω0
I2 = f (ω)dω = 2ω0 f (u + 1)−2 du, (D.6)
ω0 −1 u+1
which can also be solved by standard Gauss–Legendre quadrature.
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-bib page 245

References

1. R. M. Martin, Electronic Structure, Basic Theory and Practical Methods


(Univ. Princeton, Cambridge, 2004).
2. M. Born and J. R. Oppenheimer, Ann. Phys. 84, 457 (1927).
3. P. Ehrenfest, Z. Phys. 45, 455 (1927).
4. H. D. Meyer and W. H. Miller, J. Chem. Phys. 70, 3214 (1979).
5. J. C. Tully, Faraday Discussions 110, 407 (1998).
6. X. S. Li, J. C. Tully, H. B. Schlegel, and M. J. Frisch, J. Chem. Phys. 123,
084106 (2005).
7. Y. C. Wang, J. Lv, L. Zhu, and Y. M. Ma, Phys. Rev. B 82, 094116 (2010).
8. Y. C. Wang, J. Lv, L. Zhu, and Y. M. Ma, Comput. Phys. Commun. 183,
2063 (2012).
9. J. Lv, Y. C. Wang, L. Zhu, and Y. M. Ma, J. Chem. Phys. 137, 084104
(2012).
10. Y. C. Wang, M. S. Miao, J. Lv, L. Zhu, K. T. Yin, H. Y. Liu, and Y. M.
Ma, J. Chem. Phys. 137, 224108 (2012).
11. X. Y. Luo, J. H. Yang, H. Y. Liu, X. J. Wu, Y. C. Wang, Y. M. Ma, S.-H.
Wei, X. G. Gong, and H. J. Xiang, J. Am. Chem. Soc. 133, 16285 (2011).
12. X. X. Zhang, Y. C. Wang, J. Lv, C. Y. Zhu, Q. Li, M. Zhang, Q. Li, and
Y. M. Ma, J. Chem. Phys. 138, 114101 (2013).
13. C. J. Pickard and R. J. Needs, J. Phys.: Condens. Matter 23, 053201 (2011).
14. C. J. Pickard and R. J. Needs, Physica Status Solidi (b) 246, 536 (2009).
15. C. J. Pickard and R. J. Needs, Nat. Phys. 3, 473 (2007).
16. D. J. Jacobs, A. J. Rader, L. A. Kuhn, and M. F. Thorpe, Proteins–Struct.
Funct. Genet. 44, 150 (2001).
17. E. S. Huang, R. Samudrala, and J. W. Ponder, J. Mol. Biol. 290, 267 (1999).
18. R. J. Trabanino, S. E. Hall, N. Vaidehi, W. B. Floriano, V. W. T. Kam,
and W. A. Goddard, Biophys. J. 86, 1904 (2004).
19. Y. Abashkin and N. Russo, J. Chem. Phys. 100, 4477 (1994).
20. G. Mills, H. Jónsson, and G. K. Schenter, Surf. Sci. 324, 305 (1995).
21. G. Henkelman and H. Jónsson, J. Chem. Phys. 113, 9978 (2000).
22. S. A. Trygubenko and D. J. Wales, J. Chem. Phys. 120, 2082 (2004).

245
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-bib page 246

246 Computer Simulations of Molecules and Condensed Matter

23. M. J. S. Dewar, E. F. Healy, and J. J. P. Stewart, J. Chem. Soc. Faraday


Trans. II. 80, 227 (1984).
24. G. Henkelman and H. Jónsson, J. Chem. Phys. 111, 7010 (1999).
25. B. Peters and A. Heyden, J. Chem. Phys. 120, 7877 (2004).
26. A. Heyden, A. Bell, and F. J. Keil, J. Chem. Phys. 123, 224101 (2005).
27. Y. Tateyama, T. Ogitsu, K. Kusakabe, and S. Tsuneyuki, Phys. Rev. B 54,
14994 (1996).
28. G. T. Barkema and N. Mousseau, Phys. Rev. Lett. 77, 4358 (1996).
29. J. Klimes̆, D. R. Bowler, and A. Michaelides, J. Phys.: Condens. Matter 22,
074203 (2010).
30. N. D. Mermin, Phys. Rev. 137, A1441 (1965).
31. M. J. Gillan, J. Phys.: Condens. Matter 1, 689 (1989).
32. R. M. Wentzcovitch, J. L. Martins, and P. B. Allen, Phys. Rev. B 45, 11372
(1992).
33. C. Bartels and M. Karplus, J. Phys. Chem. B 102, 865 (1998).
34. Y. Sugita and Y. Okamoto, Chem. Phys. Lett. 314, 141 (1999).
35. Y. Sugita, A. Kitao, and Y. Okamoto, J. Chem. Phys. 113, 6042 (2000).
36. P. Liu, B. Kim, R. A. Friesner, and B. J. Berne, Proc. Natl. Acad. Sci.
U.S.A. 102, 13749 (2005).
37. X. H. Huang, M. Hagen, B. Kim, R. A. Friesner, R. H. Zhou, and B. J.
Berne, J. Phys. Chem. B 111, 5405 (2007).
38. A. Laio and M. Parrinello, Proc. Natl. Acad. Sci. U.S.A. 99, 12562 (2002).
39. A. Laio and F. L. Gervasio, Rep. Prog. Phys. 71, 126601 (2008).
40. B. A. Berg and T. Neuhaus, Phys. Lett. B 267, 249 (1991).
41. S. G. Itoh and Y. Okamoto, J. Chem. Phys. 124, 104103 (2006).
42. H. Grubmuller, Phys. Rev. E 52, 2893 (1995).
43. J. Lee, H. A. Scheraga, and S. Rackovsky, J. Comput. Chem. 18, 1222
(1997).
44. Y. Q. Gao, L. J. Yang, Y. B. Fan, and Q. Shao, Int. Rev. Phys. Chem. 27,
201 (2008).
45. Y. Q. Gao, J. Chem. Phys. 128, 064105 (2008).
46. Y. Q. Gao, J. Chem. Phys. 128, 134111 (2008).
47. L. J. Yang, Q. Shao, and Y. Q. Gao, Prog. Chem. 24, 1199 (2012).
48. D. Frenkel and B. Smit, Understanding Molecular Simulation: From Algo-
rithms to Applications (Academic Press; 2nd Ed., 2001).
49. D. Frenkel and B. M. Mulder, Mol. Phys. 55, 1171 (1985).
50. A. Stroobants, H. N. W. Lekkerkerker, and D. Frenkel, Phys. Rev. A 36,
2929 (1987).
51. D. Frenkel, H. N. W. Lekkerkerker, and A. Stroobants, Nature 332, 822
(1988).
52. E. J. Meijer, D. Frenkel, R. A. LeSar, and A. J. C. Ladd, J. Chem. Phys.
92, 7570 (1990).
53. E. J. Meijer and D. Frenkel, J. Chem. Phys. 94, 2269 (1991).
54. M. D. Eldridge, P. A. Madden, and D. Frenkel, Nature 365, 35 (1993).
55. D. Alfè, G. D. Price, and M. J. Gillan, Phys. Rev. B 64, 045123 (2001).
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-bib page 247

References 247

56. O. Sugino and R. Car, Phys. Rev. Lett. 74, 1823 (1995).
57. G. A. de Wijs, G. Kresse, and M. J. Gillan, Phys. Rev. B 57, 8223 (1998).
58. D. Alfè, G. A. de Wijs, G. Kresse, and M. J. Gillan, Int. J. Quantum Chem.
77, 871 (2000).
59. A. Kupperman, Potential Energy Surfaces and Dynamical Calculations,
D. Truhlar, ed. (Plenum, New York, 1981), pp. 375–420.
60. K. Liu, Ann. Rev. Phys. Chem. 52, 139 (2001).
61. M. H. Qiu and Z. F. Ren et al., Science 311, 1440 (2006).
62. Y. T. Lee, Science 236, 793 (1987).
63. R. P. Feynman, Phys. Rev. 76, 769 (1949).
64. R. P. Feynman, Phys. Rev. 90, 1116 (1953).
65. R. P. Feynman, Phys. Rev. 91, 1291 (1953).
66. R. P. Feynman, Phys. Rev. 91, 1301 (1953).
67. R. P. Feynman and A. R. Hibbs, Quantum Mechanics and Path Integrals
(McGraw-Hill Inc., 1965).
68. D. Chandler and P. G. Wolynes, J. Chem. Phys. 74, 4078 (1981).
69. M. Parrinello and A. Rahman, J. Chem. Phys. 80, 860 (1984).
70. D. M. Ceperley, Rev. Mod. Phys. 67, 279 (1995).
71. E. L. Pollock and D. M. Ceperley, Phys. Rev. B 30, 2555 (1984).
72. D. M. Ceperley and E. L. Pollock, Phys. Rev. Lett. 56, 351 (1986).
73. E. L. Pollock and D. M. Ceperley, Phys. Rev. B 36, 8343 (1987).
74. B. J. Berne and D. Thirumalai, Annu. Rev. Phys. Chem. 37, 401 (1986).
75. D. Thirumalai and B. J. Berne, J. Chem. Phys. 81, 2512 (1984).
76. D. Thirumalai and B. J. Berne, Chem. Phys. Lett. 116, 471 (1985).
77. R. D. Coalson, J. Chem. Phys. 83, 688 (1985).
78. J. D. Doll, D. L. Freeman, and T. L. Beck, Adv. Chem. Phys. 78, 61 (1990).
79. N. Makri, Comput. Phys. Commun. 63, 389 (1991).
80. J. S. Cao and G. A. Voth, J. Chem. Phys. 100, 5093 (1994).
81. J. S. Cao and G. A. Voth, J. Chem. Phys. 100, 5106 (1994).
82. J. S. Cao and G. A. Voth, J. Chem. Phys. 101, 6157 (1994).
83. J. S. Cao and G. A. Voth, J. Chem. Phys. 101, 6168 (1994).
84. G. A. Voth, Adv. Chem. Phys. 93, 135 (1996).
85. I. R. Craig and D. E. Manolopoulos, J. Chem. Phys. 121, 3368 (2004).
86. I. R. Craig and D. E. Manolopoulos, J. Chem. Phys. 122, 084106 (2005).
87. T. F. Miller III and D. E. Manolopoulos, J. Chem. Phys. 122, 184503 (2005).
88. I. R. Craig and D. E. Manolopoulos, J. Chem. Phys. 123, 034102 (2005).
89. T. F. Miller III and D. E. Manolopoulos, J. Chem. Phys. 123, 154504 (2005).
90. I. R. Craig and D. E. Manolopoulos, Chem. Phys. 322, 236 (2006).
91. B. J. Braams and D. E. Manolopoulos, J. Chem. Phys. 125, 124105 (2006).
92. S. Habershon, B. J. Braams, and D. E. Manolopoulos, J. Chem. Phys. 127,
174108 (2007).
93. T. E. Markland and D. E. Manolopoulos, J. Chem. Phys. 129, 024105
(2008).
94. G. M. Torrie and J. P. Valleau, Chem. Phys. Lett. 28, 578 (1974).
95. G. M. Torrie and J. P. Valleau, J. Comput. Phys. 77, 187 (1977).
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-bib page 248

248 Computer Simulations of Molecules and Condensed Matter

96. B. Roux, Comput. Phys. Commun. 91, 275 (1995).


97. M. P. Allen and D. J. Tildesley, Computer Simulation of Liquids (Oxford
University Press, USA, 1989).
98. D. R. Hartree, Proc. Cambridge Phil. Soc. 24, 89 (1928).
99. V. Fock, Z. Physik 61, 126 (1930).
100. J. A. Pople and P. K. Nesbet, J. Chem. Phys. 22, 571 (1954).
101. A. Szabo and N. S. Ostlund, Modern Quantum Chemistry: Introduction to
Advanced Electronic Structure Theory (Dover Pub. Inc., New York, 1996).
102. C. Mller and M. S. Plesset, Phys. Rev. 46, 618 (1934).
103. F. Coester, Nucl. Phys. 7, 421 (1958).
104. F. Coester and H. Kummel, Nucl. Phys. 17, 477 (1960).
105. P. Hohenberg and W. Kohn, Phys. Rev. 136, B864 (1964).
106. W. Kohn and L. J. Sham, Phys. Rev. 140, A1133 (1965).
107. L. H. Thomas, Proc. Cambridge Phil. Soc. 23, 542 (1927).
108. E. Fermi, Rend. Accad. Naz. Lincei 6, 602 (1927).
109. E. Fermi, Z. Physik 48, 73 (1928).
110. P. A. M. Dirac, Proc. Cambridge Phil. Soc. 26, 361 (1930).
111. U. von Barth and L. Hedin, J. Phys.: Condens. Maths (c.f. [13]) 5, 1629
(1972).
112. O. Gunnarsson, B. I. Lundqvist, and J. W. Wilkins, Phys. Rev. B 10, 1319
(1974).
113. R. O. Jones and O. Gunnarsson, Rev. Mod. Phys. 61, 689 (1989).
114. M. Levy, Proc. Natl. Acad. Sci. U.S.A. 76, 6062 (1979).
115. M. Levy, Phys. Rev. A 26, 1200 (1982).
116. A. Shimony and H. Feshbach, Physics as Natural Philosophy (MIT Press,
Cambridge, 1982), p. 111.
117. E. H. Lieb, Int. J. Quant. Chem. 24, 243 (1983).
118. R. M. Dreizler and J. da Providencia, Density Functional Methods in Physics
(Plenum, New York, 1985), p. 31.
119. R. M. Dreizler and E. K. U. Gross, Density Functional Theory (Springer-
Verlag, Berlin, Heidelberg, 1990).
120. A. K. Rajagopal and J. Callaway, Phys. Rev. B 7, 1912 (1973).
121. L. N. Oliveira, E. K. U. Gross, and W. Kohn, Phys. Rev. Lett. 60, 2430
(1988).
122. S. Kurth, M. Marques, M. Lüders, and E. K. U. Gross, Phys. Rev. Lett. 83,
2628 (1999).
123. J. P. Perdew and K. Schmidt, Density Functional Theory and Its Applica-
tions to Materials, V. Van Doren et al., eds. (American Institute of Physics,
New York, 2001).
124. G. D. Mahan, Many-Particle Physics (Plenum Press, New York, 1990).
125. D. M. Ceperley and B. J. Alder, Phys. Rev. Lett. 45, 566 (1980).
126. J. P. Perdew and A. Zunger, Phys. Rev. B 23, 5048 (1981).
127. J. P. Perdew and Y. Wang, Phys. Rev. B 33, 8800 (1986).
128. J. P. Perdew and Y. Wang, Phys. Rev. B 40, 3399 (1989).
129. J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett. 77, 3865 (1996).
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-bib page 249

References 249

130. J. P. Perdew, K. Burke, and Y. Wang, Phys. Rev. B 54, 16533 (1996).
131. W. Kohn, A. D. Becke, and R. G. Parr, J. Chem. Phys. 100, 12974
(1996).
132. A. D. Becke, J. Chem. Phys. 102, 8554 (1997).
133. J. P. Perdew, A. Ruzsinszky, J. M. Tao, V. N. Staroverov, G. E. Scuseria,
and G. I. Csonka, J. Chem. Phys. 123, 062201 (2005).
134. J. C. Slater and K. H. Johnson, Phys. Rev. B 5844 (1972).
135. J. F. Janak, Phys. Rev. B 18, 7165 (1978).
136. C. O. Almbladh and U. von Barth, Phys. Rev. B 31, 3231 (1985).
137. S. Lizzit, A. Baraldi, A. Groso, K. Reuter, M. V. Ganduglia-Pirovano,
C. Stamp, M. Scheer, M. Stichler, C. Keller, W. Wurth et al., Phys. Rev. B
63, 205419 (2001).
138. J. P. Perdew, D. C. Langreth, and V. Sahni, Phys. Rev. Lett. 38, 1030
(1977).
139. V. L. Moruzzi, J. F. Janak, and A. R. Williams, Calculated Electronic Prop-
erties of Metals (Pergamon Pr., Oxford, 1978).
140. A. Görling and M. Levy, Phys. Rev. A 50, 196 (1994).
141. A. Görling, Phys. Rev. B 53, 7024 (1996).
142. S. Sharma, J. K. Dewhurst, and C. Ambrosch-Draxl, Phys. Rev. Lett. 95,
136402 (2005).
143. M. Städele, J. A. Majewski, P. Vogl, and A. Göorling, Phys. Rev. Lett. 79,
2089 (1997).
144. T. Kotani, Phys. Rev. Lett. 74, 2989 (1995).
145. J. P. Perdew and M. Levy, Phys. Rev. Lett. 51, 1884 (1983).
146. L. J. Sham and M. Schlüter, Phys. Rev. Lett. 51, 1888 (1983).
147. L. J. Sham and M. Schlüter, Phys. Rev. B 32, 3883 (1985).
148. M. Grüning, A. Marini, and A. Rubio, J. Chem. Phys. 124, 154108 (2006).
149. D. J. Singh, Planewaves, Pseudopotential and the LAPW method (Kluwer
Academic Publisher, Norwell, Massachusetts, 1994).
150. D. R. Hamann, M. Schlüter, and C. Chiang, Phys. Rev. Lett. 43, 1494
(1979).
151. G. B. Bachelet, D. R. Hamann, and M. Schlüter, Phys. Rev. B 26, 4199
(1982).
152. D. Vanderbilt, Phys. Rev. B 41, 7892 (1990).
153. K. Laasonen, A. Pasquarello, R. Car, C. Lee, and D. Vanderbilt, Phys. Rev.
B 47, 10142 (1993).
154. M. Fuchs and M. Scheer, Comput. Phys. Commun. 119, 67 (1999).
155. D. R. Hamann, Phys. Rev. B 40, 2980 (1989).
156. N. Troullier and J. L. Martins, Phys. Rev. B 43, 1993 (1991).
157. L. Kleinman and D. M. Bylander, Phys. Rev. Lett. 48, 1425 (1982).
158. U. von Barth and C. D. Gelatt, Phys. Rev. B 21, 2222 (1980).
159. S. G. Louie, S. Froyen, and M. L. Cohen, Phys. Rev. B 26, 1738 (1982).
160. W. Ku and A. G. Eguiluz, Phys. Rev. Lett. 89, 126401 (2002).
161. P. Puschnig and C. Ambrosch-Draxl, Phys. Rev. B 66, 165105 (2002).
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-bib page 250

250 Computer Simulations of Molecules and Condensed Matter

162. R. Gomez-Abal, X. Z. Li, M. Scheer, and C. Ambrosch-Draxl, Phys. Rev.


Lett. 101, 106404 (2008).
163. X. Z. Li, R. Gomez-Abal, H. Jiang, C. Ambrosch-Draxl, and M. Scheer,
New J. Phys. 14, 023006 (2012).
164. J. C. Slater, Phys. Rev. 51, 846 (1937).
165. P. de Ciccio, Phys. Rev. 153, 931 (1967).
166. N. Elyashar and D. D. Koelling, Phys. Rev. B 13, 5362 (1976).
167. O. K. Andersen, Phys. Rev. B 12, 3060 (1975).
168. E. Sjöstedt, L. Nordström, and D. J. Singh, Solid State Commun. 114, 15
(2000).
169. G. K. H. Madsen, P. Blaha, K. Schwarz, E. Sjöstedt, and L. Nordström,
Phys. Rev. B 64, 195134 (2001).
170. D. Singh, Phys. Rev. B 43, 6388 (1991).
171. M. E. Rose, Elementary Theory of Angular Momentum (John Wiley and
Sons, 1957).
172. M. Kara and K. Kurki-Suonio, Acta Crystallograca A 37, 201 (1981).
173. J. Deslippe, G. Samsonidze, D. A. Strubbe, M. Jain, M. L. Cohen, and
S. G. Louie, Comput. Phys. Commun. 183, 1269 (2012).
174. M. R. A. Shegelski, Am. J. Phys. 72, 676 (2004).
175. A. L. Fetter and J. D. Walecka, Quantum Theory of Many-Particle Systems
(McGraw-Hill Inc., 1971).
176. L. Hedin, Phys. Rev. 139, A796 (1965).
177. L. Hedin and S. Lundqvist, Solid State Phys.: Advances in Research and
Applications 23, 1 (1969).
178. F. Aryasetiawan and O. Gunnarsson, Rep. Prog. Phys. 61, 237 (1998).
179. J. Schwinger, Proc. Natl. Acad. Sci. U.S.A. 37, 452 (1951).
180. G. Pratt, Phys. Rev. 118, 462 (1960).
181. G. Pratt, Rev. Mod. Phys. 35, 502 (1963).
182. L. Hedin, Bull. Am. Phys. Soc. 8, 535 (1963).
183. E. K. U. Gross, R. Runge, and O. Heinonen, Many-Particle Theory (Adam
Hilger, 1991).
184. D. Straub, L. Ley, and F. J. Himpsel, Phys. Rev. Lett. 54, 142 (1985).
185. M. S. Hybertsen and S. G. Louie, Phys. Rev. Lett. 55, 1418 (1985).
186. J. E. Northrup, M. S. Hybertsen, and S. G. Louie, Phys. Rev. Lett. 59, 819
(1987).
187. C. Petrillo and F. Sacchetti, Phys. Rev. B 38, 3834 (1988).
188. E. L. Shirley, Z. J. Zhu, and S. G. Louie, Phys. Rev. B 56, 6648 (1997).
189. O. Zakharov, A. Rubio, X. Blase, M. L. Cohen, and S. G. Louie, Phys. Rev.
B 50, 10780 (1994).
190. X. J. Zhu and S. G. Louie, Phys. Rev. B 43, 14142 (1991).
191. T. Kotani and M. van Schilfgaarde, Solid State Commun. 121, 461 (2002).
192. C. Friedrich, A. Schindlmayr, S. Blügel, and T. Kotani, Phys. Rev. B 74,
045104 (2006).
193. H. Jiang, R. Gomez-Abal, X. Z. Li, C. Meisenbichler, C. Ambrosch-Draxl,
and M. Scheer, Comput. Phys. Commun. 184, 348 (2013).
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-bib page 251

References 251

194. F. Gygi and A. Baldereschi, Phys. Rev. Lett. 62, 2160 (1989).
195. A. Fleszar and W. Hanke, Phys. Rev. B 56, 10228 (1997).
196. R. W. Godby, M. Schlüter, and L. J. Sham, Phys. Rev. B 36, 6497 (1987).
197. J. Q. Wang, Z. Q. Gu, and M. F. Li, Phys. Rev. B 44, 8707 (1991).
198. K. Delaney, P. Garcia-Gonzalez, A. Rubio, P. Rinke, and R. W. Godby,
Phys. Rev. Lett. 93, 249701 (2004).
199. W. Ku and A. G. Eguiluz, Phys. Rev. Lett. 93, 249702 (2004).
200. M. van Schilfgaarde, T. Kotani, and S. V. Faleev, Phys. Rev. B 74, 245125
(2006).
201. M. L. Tiago, S. Ismail-Beigi, and S. G. Louie, Phys. Rev. B 69, 125212
(2004).
202. M. van Schilfgaarde, T. Kotani, and S. Faleev, Phys. Rev. Lett. 96, 226402
(2006).
203. S. Faleev, M. van Schilfgaarde, and T. Kotani, Phys. Rev. Lett. 93, 126406
(2004).
204. M. Shishkin and G. Kresse, Phys. Rev. B 75, 235102 (2007).
205. M. Shishkin, M. Marsman, and G. Kresse, Phys. Rev. Lett. 99, 246403
(2007).
206. T. Kotani and M. van Schilfgaarde, Phys. Rev. B 76, 165106 (2007).
207. K. S. Thygesen and A. Rubio, Phys. Rev. B 77, 115333 (2008).
208. F. Bruneval, N. Vast, and L. Reining, Phys. Rev. B 74, 045102 (2006).
209. W. D. Schöne and A. G. Eguiluz, Phys. Rev. Lett. 81, 1662 (1998).
210. C. Rostgaard, K. W. Jacobsen, and K. S. Thygesen, Phys. Rev. B 81, 085103
(2010).
211. A. Stan, N. E. Dahlen, and R. van Leeuwen, Europhys. Lett. 76, 298 (2006).
212. X. Z. Li, All-electron G0W0 code based on FP-(L)APW+lo and applications,
Ph.D. Thesis (Free University of Berlin, 2008).
213. P. Blaha, K. Schwarz, G. K. H. Madsen, D. Kvasnicka, and J. Luitz,
WIEN2k, An Augmented Plane Wave Plus Local Orbitals Program for Cal-
culating Crystal Properties (Tchn. Universität Wien, Austria, 2002), ISBN
3-9501031-1-2.
214. URL http://exciting-code.org.
215. S. Sagmeister and C. Ambrosch-Draxl, Phys. Chem. Chem. Phys. 11, 4451
(2009).
216. S. Sharma, J. K. Dewhurst, and C. Ambrosch-Draxl, Phys. Rev. Lett. 95,
136402 (2005).
217. A. Baldereschi, Phys. Rev. B 7, 5212 (1973).
218. D. J. Chadi and M. L. Cohen, Phys. Rev. B 7, 692 (1973).
219. D. J. Chadi and M. L. Cohen, Phys. Rev. B 8, 5747 (1973).
220. H. J. Monkhorst and J. D. Pack, Phys. Rev. B 13, 5188 (1976).
221. O. Jepsen and O. K. Andersen, Solid State Commun. 9, 1763 (1971).
222. G. Lehmann, P. Rennert, M. Taut, and H. Wonn, Phys. Status Solidi 37,
K27 (1970).
223. J. Rath and A. J. Freeman, Phys. Rev. B 11, 2109 (1975).
224. P. E. Blöchl, O. Jepsen, and O. K. Andersen, Phys. Rev. B 49, 16223 (1994).
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-bib page 252

252 Computer Simulations of Molecules and Condensed Matter

225. R. W. Godby, M. Schlüter, and L. J. Sham, Phys. Rev. B 37, 10159 (1988).
226. T. Matsubara, Prog. Theor. Phys. 14, 351 (1955).
227. M. M. Rieger, L. Steinbeck, I. D. White, H. N. Rojas, and R. W. Godby,
Comput. Phys. Commun. 117, 211 (1999).
228. R. Car and M. Parrinello, Phys. Rev. Lett. 55, 2471 (1985).
229. L. Verlet, Phys. Rev. 159, 98 (1967).
230. W. C. Swope, H. C. Andersen, P. H. Berens, and K. R. Wilson, J. Chem.
Phys. 76, 637 (1982).
231. R. W. Hockney and J. W. Eastwood, Computer Simulations Using Particles
(McGraw-Hill, New York, 1981).
232. A. Rahman, Phys. Rev. 136, A405 (1964).
233. W. F. Vangunsteren and H. J. C. Berendsen, Mol. Phys. 34, 1311 (1977).
234. J. L. Lebowitz, J. K. Percus, and J. Verlet, Phys. Rev. 153, 250 (1967).
235. P. S. Y. Cheung, Mol. Phys. 33, 519 (1967).
236. J. R. Ray and H. W. Graben, Mol. Phys. 43, 1293 (1981).
237. J. R. Ray and H. W. Graben, Phys. Rev. A 44, 6905 (1991).
238. H. W. Graben and J. R. Ray, Phys. Rev. A 43, 4100 (1991).
239. H. C. Andersen, J. Chem. Phys. 72, 2384 (1980).
240. S. Nose, J. Chem. Phys. 81, 511 (1984).
241. S. Nose, Mol. Phys. 52, 255 (1984).
242. W. G. Hoover, Phys. Rev. A 31, 1695 (1985).
243. W. G. Hoover, Phys. Rev. A 34, 2499 (1986).
244. S. Nose, Mol. Phys. 57, 187 (1986).
245. G. J. Martyna, M. L. Klein, and M. E. Tuckerman, J. Chem. Phys. 97, 2635
(1992).
246. M. E. Tuckerman and G. J. Martyna, J. Phys. Chem. B 104, 159 (2000).
247. M. Parrinello and A. Rahman, Phys. Rev. Lett. 45, 1196 (1980).
248. M. Parrinello and A. Rahman, J. Appl. Phys. 52, 7182 (1981).
249. G. J. Martyna, D. J. Tobias, and M. L. Klein, J. Chem. Phys. 101, 4177
(1994).
250. J. M. Haile and H. W. Graben, J. Chem. Phys. 73, 2412 (1980).
251. J. M. Haile and H. W. Graben, Mol. Phys. 40, 1433 (1980).
252. J. R. Ray and H. W. Graben, Phys. Rev. A 34, 2517 (1986).
253. J. R. Ray and H. W. Graben, J. Chem. Phys. 75, 4077 (1981).
254. M. Ceriotti, G. Bussi, and M. Parrinello, Phys. Rev. Lett. 102, 020601
(2009).
255. P. E. Blöchl and M. Parrinello, Phys. Rev. B 45, 9413 (1992).
256. D. Quigley and M. I. J. Probert, J. Chem. Phys. 120, 11432 (2004).
257. W. G. Hoover, K. Aoki, C. G. Hoover, and S. V. de Groot, Physica
D-Nonlinear Phenomena 187, 253 (2004).
258. W. G. Hoover, Phys. Rev. A 34, 2499 (1986).
259. A. Kolb and B. Dunweg, J. Chem. Phys. 111, 4453 (1999).
260. G. Kresse and J. Hafner, Phys. Rev. B 47, 558 (1993).
261. G. Kresse and J. Hafner, Phys. Rev. B 49, 14251 (1994).
262. G. Kresse and J. Furthmüller, Comput. Mat. Sci. 6, 15 (1996).
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-bib page 253

References 253

263. G. Kresse and J. Furthmüller, Phys. Rev. B 54, 11169 (1996).


264. J. Chen, X. Z. Li, Q. F. Zhang, A. Michaelides, and E. G. Wang, Phys.
Chem. Chem. Phys. 15, 6344 (2013).
265. M. Tuckerman, K. Laasonen, M. Sprik, and M. Parrinello, J. Phys. Chem.
99, 5749 (1995).
266. M. Tuckerman, K. Laasonen, M. Sprik, and M. Parrinello, J. Chem. Phys.
103, 150 (1995).
267. R. Vuilleumier and D. Borgis, J. Phys. Chem. B 102, 4261 (1998).
268. R. Vuilleumier and D. Borgis, J. Chem. Phys. 111, 4251 (1999).
269. U. W. Schmitt and G. A. Voth, J. Phys. Chem. B 102, 5547 (1998).
270. U. W. Schmitt and G. A. Voth, J. Chem. Phys. 111, 9361 (1999).
271. G. A. Voth, Acc. Chem. Res. 39, 143 (2006).
272. O. Markovitch, H. N. Chen, S. Izvekov, F. Paesani, G. A. Voth, and
N. Agmon, J. Phys. Chem. B 112, 9456 (2008).
273. S. Woutersen and H. J. Bakker, Phys. Rev. Lett. 96, 138305 (2006).
274. J. M. Headrick, E. G. Diken, R. S. Walters, N. I. Hammer, R. A. Christie,
J. Cui, E. M. Myshakin, M. A. Duncan, M. A. Johnson, and K. D. Jordan,
Science 308, 1765 (2005).
275. C. J. T. de Grotthuss, Ann. Chim. 58, 54 (1806).
276. N. Agmon, Chem. Phys. Lett. 244, 456 (1995).
277. S. Scandolo, Proc. Natl. Acad. Sci. U.S.A. 10, 3051 (2003).
278. S. A. Bonev, E. Schwegler, T. Ogitsu, and G. Galli, Nature 431, 669 (2004).
279. H. Y. Liu and Y. M. Ma, Phys. Rev. Lett. 110, 025903 (2013).
280. D. Alfé, M. J. Gillan, and G. D. Price, Nature 405, 172 (2000).
281. X. Z. Li, B. Walker, M. I. J. Probert, C. J. Pickard, R. J. Needs, and
A. Michaelides, J. Phys.: Condens. Matter 25, 085402 (2013).
282. M. A. Morales, C. Pierleoni, E. Schwegler, and D. M. Ceperley, Proc. Natl.
Acad. Sci. U.S.A. 107, 12799 (2010).
283. M. A. Morales, J. M. McMahon, C. Pierleoni, and D. M. Ceperley, Phys.
Rev. B 87, 184107 (2013).
284. M. A. Morales, J. M. McMahon, C. Pierleoni, and D. M. Ceperley, Phys.
Rev. Lett. 110, 065702 (2013).
285. S. Azadi and W. M. C. Foulkes, Phys. Rev. B 88, 014115 (2013).
286. X. Z. Li, M. I. J. Probert, A. Alavi, and A. Michaelides, Phys. Rev. Lett.
104, 066102 (2010).
287. V. Tozzini, Curr. Opin. Struct. Biol. 15, 144 (2005).
288. M. Christen et al., J. Comput. Chem. 26, 1719 (2005).
289. J. G. Kirkwood, J. Chem. Phys. 3, 300 (1935).
290. J. Shen and J. A. McCammon, Chem. Phys. 158, 191 (1991).
291. M. Mezei, P. K. Mehorotra, and D. L. Beveridge, J. Am. Chem. Soc. 107,
2239 (1985).
292. C. Haydock, J. C. Sharp, and F. G. Prendergast, Biophys. J. 57, 1269
(1990).
293. T. B. Woolf and B. Roux, J. Am. Chem. Soc. 116, 5916 (1994).
294. T. B. Woolf, S. Crouzy, and B. Roux, Biophys. J. 67, 1370 (1994).
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-bib page 254

254 Computer Simulations of Molecules and Condensed Matter

295. S. Kumar, D. Bouzida, R. H. Swendsen, P. A. Kollman, and J. M. Rosen-


berg, J. Comput. Chem. 13, 1011 (1992).
296. M. Mezei, J. Comput. Phys. 68, 237 (1987).
297. G. H. Paine and H. A. Scheraga, Biopolymers 24, 1391 (1985).
298. R. W. W. Hooft, B. P. van Eijck, and J. Kroon, J. Chem. Phys. 97, 6690
(1992).
299. R. Rajamani, K. J. Naidoo, and J. L. Gao, J. Comput. Chem. 24, 1175
(2003).
300. C. Bartels and M. Karplus, J. Comput. Chem. 18, 1450 (1997).
301. M. Iannuzzi and M. Parrinello, Phys. Rev. Lett. 93, 025901 (2004).
302. A. Barducci, R. Chelli, P. Procacci, V. Schettino, F. Gervasio, and M. Par-
rinello, J. Am. Chem. Soc. 128, 2705 (2006).
303. A. Laio, A. Rodriguez-Fortea, and F. L. Gervasio, J. Phys. Chem. B 109,
6714 (2005).
304. R. Martonak, A. Laio, and M. Parrinello, Phys. Rev. Lett. 90, 075503 (2003).
305. B. Ensing, M. De Vivo, Z. W. Liu, P. Moore, and M. L. Klein, Acc. Chem.
Res. 39, 73 (2006).
306. F. L. Gervasio, A. Laio, and M. Parrinello, J. Am. Chem. Soc. 127, 2600
(2005).
307. S. Piana and A. Laio, J. Phys. Chem. B 111, 4553 (2007).
308. C. Tsallis, J. Stat. Phys. 52, 479 (1988).
309. I. Fukuda and H. Nakamura, Phys. Rev. E 71, 046708 (2005).
310. F. G. Wang and D. P. Landau, Phys. Rev. E 64, 056101 (2001).
311. F. G. Wang and D. P. Landau, Phys. Rev. Lett. 86, 2050 (2001).
312. L. J. Yang and Y. Q. Gao, J. Chem. Phys. 131, 214109 (2009).
313. O. G. Mouritsen, Computer Studies of Phase Transitions and Critical Phe-
nomena (Springer, Berlin, 1984).
314. L. J. Lauhon and W. Ho, Phys. Rev. Lett. 85, 4566 (2000).
315. Y. V. Suleimanov, J. Phys. Chem. C 116, 11141 (2012).
316. L. Masgrau et al., Science 312, 237 (2006).
317. C. R. Pudney et al., J. Am. Chem. Soc. 135, 2512 (2013).
318. J. R. Rommel et al., J. Phys. Chem. B 116, 13682 (2012).
319. M. E. Tuckerman, D. Marx, M. L. Klein, and M. Parrinello, J. Chem. Phys.
104, 5579 (1996).
320. D. Marx and M. Parrinello, J. Chem. Phys. 104, 4077 (1996).
321. D. Marx, M. E. Tuckerman, J. Hutter, and M. Parrinello, Nature 397, 601
(1999).
322. X. Z. Li, B. Walker, and A. Michaelides, Proc. Natl. Acad. Sci. U.S.A. 108,
6369 (2011).
323. J. Chen, X. Z. Li, and Q. F. Zhang et al., Nat. Commun. 4, 2064 (2013).
324. M. A. Morales, J. M. McMahon, C. Pierleoni, and D. M. Ceperley, Phys.
Rev. Lett. 110, 065702 (2013).
325. M. A. Morales, J. M. McMahon, C. Pierleoni, and D. M. Ceperley, Phys.
Rev. B 87, 184107 (2013).
326. M. Tachikawa and M. Shiga, J. Am. Chem. Soc. 127, 11908 (2005).
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-bib page 255

References 255

327. A. Kaczmarek, M. Shiga, and D. Marx, J. Phys. Chem. A 113, 1985 (2009).
328. P. Pechukas, Ann. Rev. Phys. Chem. 32, 159 (1981).
329. W. H. Miller, Acc. Chem. Res. 9, 306 (2005).
330. J. D. Doll and A. F. Voter, Ann. Rev. Phys. Chem. 38, 413 (1987).
331. G. Mills, H. Jonsson, and G. K. Schenter, Surf. Sci. 324, 305 (1995).
332. R. Ramrez, C. P. Herrero, A. Antonelli, and E. R. Hernandez, J. Chem.
Phys. 129, 064110 (2008).
333. S. Habershon and D. E. Manolopoulos, J. Chem. Phys. 135, 224111 (2011).
334. G. F. Reiter, J. Mayers, and P. Platzman, Phys. Rev. Lett. 89, 135505
(2002).
335. G. F. Reiter, J. C. Li, J. Mayers, T. Abdul-Redah, and P. Platzman, Braz.
J. Phys. 34, 142 (2004).
336. C. Andreani, D. Colognesi, J. Mayers, G. F. Reiter, and R. Senesi, Adv.
Phys. 54, 377 (2005).
337. J. A. Morrone, V. Srinivasan, D. Sebastiani, and R. Car, J. Chem. Phys.
126, 234504 (2007).
338. J. A. Morrone and R. Car, Phys. Rev. Lett. 101, 017801 (2008).
339. J. A. Morrone, L. Lin, and R. Car, J. Chem. Phys. 130, 204511 (2009).
340. L. Lin, J. A. Morrone, R. Car, and M. Parrinello, Phys. Rev. Lett. 105,
110602 (2010).
341. L. Lin, J. A. Morrone, and R. Car, J. Stat. Phys. 145, 365 (2011).
342. M. Ceriotti and D. E. Manolopoulos, Phys. Rev. Lett. 109, 100604 (2012).
343. R. W. Hall and B. J. Berne, J. Chem. Phys. 81, 3641 (1984).
344. S. Habershon, G. S. Fanourgakis, and D. E. Manolopoulos, J. Chem. Phys.
129, 074501 (2008).
345. B. Chen, I. Ivanov, M. Klein, and M. Parrinello, Phys. Rev. Lett. 91, 215503
(2003).
346. V. Buch, J. Chem. Phys. 97, 726 (1992).
347. P. Sandler, J. O. Jung, M. M. Szczesniak, and V. Buch, J. Chem. Phys.
101, 1378 (1994).
348. J. K. Gregory and D. C. Clary, Chem. Phys. Lett. 228, 547 (1994).
349. J. K. Gregory and D. C. Clary, J. Phys. Chem. 100, 18014 (1996).
350. M. F. Herman, E. J. Bruskin, and B. J. Berne, J. Chem. Phys. 76, 5150
(1982).
351. M. E. Tuckerman, Statistical Mechanics: Theory and Molecular Simulations
(Oxford University Press, USA, 2010).
352. A. Nakayama and N. Makri, J. Chem. Phys. 125, 024503 (2006).
353. J. S. Shao and N. Makri, J. Phys. Chem. A 103, 7753 (1999).
354. J. S. Shao and N. Makri, J. Phys. Chem. A 103, 9479 (1999).
355. Q. Shi and E. Geva, J. Phys. Chem. A 107, 9059 (2003).
356. Q. Shi and E. Geva, J. Phys. Chem. A 107, 9070 (2003).
357. J. Liu and N. Makri, Chem. Phys. 322, 23 (2006).
358. M. H. Beck, A. Jackle, G. A. Worth, and H. D. Meyer, Phys. Rep. 324, 1
(2000).
359. S. C. Althorpe and D. C. Clary, Annu. Rev. Phys. Chem. 54, 493 (2003).
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-bib page 256

256 Computer Simulations of Molecules and Condensed Matter

360. D. H. Zhang, M. A. Collins, and S. Y. Lee, Science 290, 961 (2000).


361. M. A. Collins, Theo. Chem. Acc. 108, 313 (2002).
362. O. Vendrell, F. Gatti, and H. D. Meyer, Angew. Chem. Int. Ed. 46, 6918
(2007).
363. G. Baym and N. D. Mermin, J. Math. Phys. 2, 232 (1961).
364. E. Rabani, G. Krilov, and B. J. Berne, J. Chem. Phys. 112, 2605 (2000).
365. G. Krilov, E. Sim, and B. J. Berne, Chem. Phys. 268, 21 (2001).
366. E. Sim, G. Krilov, and B. J. Berne, J. Phys. Chem. A 105, 2824 (2001).
367. E. Rabani, D. R. Reichman, G. Krilov, and B. J. Berne, Proc. Natl. Acad.
Sci. U.S.A. 99, 1129 (2002).
368. D. R. Reichman and E. Rabani, Phys. Rev. Lett. 87, 265702 (2001).
369. E. Rabani and D. R. Reichman, Phys. Rev. E 65, 036111 (2002).
370. E. Rabani and D. R. Reichman, J. Chem. Phys. 116, 6271 (2002).
371. D. R. Reichman and E. Rabani, J. Chem. Phys. 116, 6279 (2002).
372. E. Rabani and D. R. Reichman, J. Chem. Phys. 120, 1458 (2004).
373. E. Rabani and D. R. Reichman, Annu. Rev. Phys. Chem. 56, 157 (2005).
374. D. Kim, J. D. Doll, and J. E. Gubernatis, J. Chem. Phys. 106, 1641 (1997).
375. B. J. Berne, J. Stat. Phys. 43, 911 (1986).
376. H. Wang, X. Sun, and W. H. Miller, J. Chem. Phys. 108, 9726 (1998).
377. X. Sun, H. Wang, and W. H. Miller, J. Chem. Phys. 109, 7064 (1998).
378. J. Liu and W. H. Miller, J. Chem. Phys. 125, 224104 (2006).
379. J. Liu and W. H. Miller, J. Chem. Phys. 127, 114506 (2007).
380. J. A. Poulsen, G. Nyman, and P. J. Rossky, J. Chem. Phys. 119, 12179
(2003).
381. J. A. Poulsen, G. Nyman, and P. J. Rossky, J. Phys. Chem. A 108, 8743
(2004).
382. J. A. Poulsen, G. Nyman, and P. J. Rossky, J. Phys. Chem. B 108, 19799
(2004).
383. J. A. Poulsen, G. Nyman, and P. J. Rossky, Proc. Natl. Acad. Sci. U.S.A.
102, 6709 (2005).
384. J. S. Cao and G. A. Voth, J. Chem. Phys. 99, 10070 (1993).
385. R. P. Feynman and H. Kleinert, Phys. Rev. A 34, 5080 (1986).
386. M. J. Gillan, Phys. Rev. Lett. 58, 563 (1987).
387. M. J. Gillan, J. Phys. C: Solid State Phys. 20, 3621 (1987).
388. B. C. Garrett and D. G. Truhlar, J. Phys. Chem. 83, 1052 (1979).
389. B. C. Garrett and D. G. Truhlar, J. Phys. Chem. 87, 4553 (1983).
390. W. H. Miller, J. Chem. Phys. 61, 1823 (1974).
391. B. C. Garrett, D. G. Truhlar, R. S. Grev, and A. W. Magnuson, J. Phys.
Chem. 84, 1730 (1980).
392. J. W. Tromp and W. H. Miller, J. Phys. Chem. 90, 3482 (1986).
393. G. A. Voth, D. Chandler, and W. H. Miller, J. Chem. Phys. 91, 7749 (1989).
394. M. Messina, G. K. Schenter, and B. C. Garrett, J. Chem. Phys. 98, 8525
(1993).
395. N. F. Hansen and H. C. Andersen, J. Chem. Phys. 101, 6032 (1994).
396. J. Cao and G. J. Martyna, J. Chem. Phys. 104, 2028 (1996).
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-bib page 257

References 257

397. T. D. Hone, P. J. Rossky, and G. A. Voth, J. Chem. Phys. 124, 154103


(2006).
398. A. Pérez, M. E. Tuckerman, and M. H. Müser, J. Chem. Phys. 130, 184105
(2009).
399. S. D. Ivanov, A. Witt, M. Shiga, and D. Marx, J. Chem. Phys. 132, 031101
(2010).
400. J. Morales and K. Singer, Mol. Phys. 73, 873 (1991).
401. A. Hodgson and S. Haq, Surf. Sci. Rep. 64, 381 (2009).
402. G. Held and D. Menzel, Surf. Sci. 316, 92 (1994).
403. A. Michaelides and P. Hu, J. Am. Chem. Soc. 123, 4235 (2001).
404. P. J. Feibelman, Science 295, 99 (2002).
405. S. Völkening, K. Bedürftig, K. Jacobi, J. Wintterlin, and G. Ertl, Phys.
Rev. Lett. 83, 2672 (1999).
406. C. Clay, S. Haq, and A. Hodgson, Phys. Rev. Lett. 92, 046102 (2004).
407. T. Schiros, L. A. Näslund, K. Andersson, J. Gyllenpalm, G. S. Karlberg,
M. Odelius, H. Ogasawara, L. G. M. Pettersson, and A. Nilsson, J. Phys.
Chem. C 111, 15003 (2007).
408. L. Giordano, J. Goniakowski, and J. Suzanne, Phys. Rev. Lett. 81, 1271
(1998).
409. Y. D. Kim, R. M. Lynden-Bell, A. Alavi, J. Stultz, and D. W. Goodman,
Chem. Phys. Lett. 352, 318 (2002).
410. B. Meyer, D. Marx, O. Dulub, U. Diebold, M. Kunat, D. Langenberg, and
C. Wöll, Angew. Chem. Int. Ed. 43, 6641 (2004).
411. A. Michaelides, A. Alavi, and D. A. King, Phys. Rev. B. 69, 113404
(2004).
412. M. Benoit, D. Marx, and M. Parrinello, Nature 392, 258 (1998).
413. E. Schwegler, M. Sharma, F. Gygi, and G. Galli, Proc. Natl. Acad. Sci.
U.S.A. 105, 14779 (2008).
414. S. J. Clark, M. D. Segall, C. J. Pickard, P. J. Hasnip, M. J. Probert,
K. Refson, and M. C. Payne, Z. Kristallogr. 220, 567 (2005).
415. C. Sachs, S. Völkening, J. Wintterlin, and G. Ertl, Science 293, 1635 (2001).
416. K. Bedürftig, S. Völkening, Y. Wang, J. Wintterlin, K. Jocobi, and G. Ertl,
J. Chem. Phys. 123, 064711 (2005).
417. K. Ando and J. T. Hynes, J. Phys. Chem. B 101, 10464 (1997).
418. A. R. Ubbelohde and K. J. Gallagher, Acta Crystallogr. 8, 71 (1955).
419. E. Matsushita and T. Matsubara, Prog. Theor. Phys. 67, 1 (1982).
420. S. Raugei and M. L. Klein, J. Am. Chem. Soc. 125, 8992 (2003).
421. D. C. Clary, D. M. Benoit, and T. van Mourik, Acc. Chem. Res. 33, 441
(2000).
422. A. C. Legon and D. J. Millen, Chem. Phys. Lett. 147, 484 (1988).
423. C. Swalina, Q. Wang, A. Chakraborty, and S. Hammes-Schier, J. Phys.
Chem. A 111, 2206 (2007).
424. K. Wendler, J. Thar, S. Zahn, and B. Kirchner, J. Phys. Chem. A 114,
9529 (2010).
425. S. S. Xantheas and T. H. Dunning, J. Chem. Phys. 99, 8774 (1993).
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-bib page 258

258 Computer Simulations of Molecules and Condensed Matter

426. E. Cubero, M. Orozco, P. Hobza, and F. J. Luque, J. Phys. Chem. A 103,


6394 (1999).
427. S. Habershon, T. E. Markland, and D. E. Manolopoulos, J. Chem. Phys.
131, 024501 (2009).
428. J. Ireta, J. Neugebauer, M. Scheer, A. Rojo, and M. J. Galvan, J. Phys.
Chem. B 107, 1432 (2003).
429. V. Srinivasan and D. Sebastiani, J. Phys. Chem. C 115, 12631 (2011).
430. M. E. Tuckerman, D. Marx, M. L. Klein, and M. Parrinello, Science 275,
817 (1997).
431. H. Ishibashi, A. Hayashi, M. Shiga, and M. Tachikawa, ChemPhysChem 9,
383 (2008).
432. E. Wigner and H. B. Huntington, J. Chem. Phys. 3, 764 (1935).
433. H. K. Mao and R. J. Hemley, Rev. Mod. Phys. 66, 671 (1994).
434. K. A. Johnson and N. W. Ashcroft, Nature 403, 632 (2000).
435. E. Babaev, A. Sudbo, and N. W. Ashcroft, Nature 431, 666 (2004).
436. S. Deemyad and I. F. Silvera, Phys. Rev. Lett. 100, 155701 (2008).
437. P. Loubeyre, F. Occelli, and R. LeToullec, Nature 416, 613 (2002).
438. M. I. Eremets and I. A. Troyan, Nat. Mater. 10, 927 (2011).
439. C. S. Zha, Z. X. Liu, and R. J. Hemley, Phys. Rev. Lett. 108, 146402 (2012).
440. L. Dubrovinsky, N. Dubrovinskaia, V. B. Prakapenka, and A. M. Abakumov,
Nat. Commun. 3, 1163 (2012).
441. A. F. Goncharov, R. J. Hemley, and H. K. Mao, J. Chem. Phys. 134, 174501
(2011).
442. V. E. Fortov, R. I. Ilkaev, V. A. Arinin, V. V. Burtzev, V. A. Golubev,
I. L. Iosilevskiy, V. V. Khrustalev, A. L. Mikhailov, M. A. Mochalov, V. Y.
Ternovoi et al., Phys. Rev. Lett. 99, 185001 (2007).
443. W. J. Nellis, S. T. Weir, and A. C. Mitchell, Science 273, 936 (1996).
444. N. W. Ashcroft, J. Phys.: Condens. Matter 12, A129 (2000).
445. P. Cudazzo, G. Profeta, A. Sanna, A. Floris, A. Continenza, S. Massidda,
and E. K. U. Gross, Phys. Rev. Lett. 100, 257001 (2008).
446. J. M. McMahon and D. M. Ceperley, Phys. Rev. B 84, 144515 (2011).
447. T. Ogitsu, E. Schwegler, F. Gygi, and G. Galli, Phys. Rev. Lett. 91, 175502
(2003).
448. D. Alfè, Phys. Rev. B 68, 064423 (2003).
449. C. J. Pickard, M. Martinez-Canales, and R. J. Needs, Phys. Rev. B 85,
214114 (2012).
450. C. J. Pickard, M. Martinez-Canales, and R. J. Needs, Phys. Rev. B 86,
059902 (2012).
451. J. M. McMahon and D. M. Ceperley, Phys. Rev. Lett. 106, 165302 (2011).
452. H. Y. Liu, H. Wang, and Y. M. Ma, J. Phys. Chem. C 116, 9221 (2012).
453. R. T. Howie, C. L. Guillaume, T. Scheler, A. F. Goncharov, and E. Grego-
ryanz, Phys. Rev. Lett. 108, 125501 (2012).
454. E. R. Hernandez, A. Rodriguez-Prieto, A. Bergara, and D. Alfè, Phys. Rev.
Lett. 104, 185701 (2010).
455. V. Adamo and V. Barone, J. Chem. Phys. 110, 6158 (1999).
December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-bib page 259

References 259

456. J. Klimes, D. R. Bowler, and A. Michaelides, J. Phys.: Condens. Matter 22,


022201 (2010).
457. J. Klimes, D. R. Bowler, and A. Michaelides, Phys. Rev. B 83, 195131
(2011).
458. M. Dion, H. Rydberg, E. Schroder, D. C. Langreth, and B. I. Lundqvist,
Phys. Rev. Lett. 92, 246401 (2004).
459. M. Dion, H. Rydberg, E. Schroder, D. C. Langreth, and B. I. Lundqvist,
Phys. Rev. Lett. 95, 109902 (2005).
460. N. W. Ashcroft, Phys. Rev. Lett. 21, 1748 (1968).
461. P. B. Allen and R. C. Dynes, Phys. Rev. B 12, 905 (1975).
462. URL http://www.quantum-expresso.org.
463. J. Bardeen, L. N. Cooper, and J. R. Schrieer, Phys. Rev. 108, 1175 (1957).
464. R. Szczesniak and M. W. Jarosik, Solid State Commun. 149, 2053 (2009).
465. M. Abramowitz and I. A. Stegun, Handbook of Mathematical Functions
(Dover Pub. Inc., New York, 1972).
b2530   International Strategic Relations and China’s National Security: World at the Crossroads

This page intentionally left blank

b2530_FM.indd 6 01-Sep-16 11:03:06 AM


December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-ack page 261

Acknowledgements

The authors would like to acknowledge a number of people, who, from


the time of their education, have been of great help in the preparation
of this book. We are grateful to J. B. Xia for his patience in reading
a large part of the book and giving comments/suggestions. We are also
grateful to a number of people for their help during the education of each
of the authors. Specifically, XZL is deeply indebted to J. B. Xia, M. Schef-
fler, A. Michaelides, R. Gomez-Abal, C. Draxl, H. Jiang, E. K. U. Gross,
R. J. Needs, D. Manolopoulos, D. Alfè, M. J. Gillan, C. J. Pickard, M. I.
J. Probert, A. Alavi, P. J. Hu, X. G. Ren, W. X. Li, P. Rinke, V. Blum,
K. Reuter, Q. M. Hu, H. Wu, X. L. Hu, L. M. Liu, and all his colleagues in
FHI (Berlin) and UCL (London). EGW wants to thank G. Allan (ISEN) and
D. S. Wang (IOP) for their kind help. During the writing of this manuscript
in Peking University, we are also grateful to Y. Q. Gao, L. J. Yang, W. J.
Xie, G. Sun, J. Liu, J. S. Shao, Q. Shi, H. T. Quan, J. Chen, Y. X. Feng,
W. Fang, W. An, L. M. Xu, J. Feng, and all our colleagues/friends in China
for the illuminating discussions. Finally, XZL wants to thank his wife, his
late parents, his parents-in-law, his daughter, his sister and her family for
all the support in the past years. The book would not have been completed
without the support from them.

Xin-Zheng Li & Enge Wang

261
b2530   International Strategic Relations and China’s National Security: World at the Crossroads

This page intentionally left blank

b2530_FM.indd 6 01-Sep-16 11:03:06 AM


December 28, 2017 10:20 Computer Simulations of Molecules and Condensed Matter 9in x 6in b3021-index page 263

Index

computational condensed matter GW approximation, 15, 48


physics, v
many-body perturbation theory, v, 6,
density-functional theory, v, 6, 15, 17, 15
31, 133, 138 molecular dynamics, v, vi, 2, 14, 15,
81, 129, 137–140, 156, 157, 169,
electronic excitations, 15, 133 182, 218
electronic structures, v, vi, 1–3, 6, 7,
9–12, 15, 17, 22, 38, 39, 42, 49, 73, path-integral, 178
81, 82, 108, 111, 132, 133, 138, 139,
152, 163, 175, 176, 197, 211, 218 quantum nuclear effects, v, 10, 14, 81,
enhanced sampling, v, 15, 114–116, 137, 138, 179, 202, 204, 205, 213
124, 125, 127, 129
single-particle excitation, 47, 48, 58,
first-principle, 26 62, 63, 66

263

You might also like