© 2020 by
Yubo “Paul” Yang
. All rights reserved.
ELECTRON-ION CORRELATION AND FINITE-SIZE EFFECTS IN
QUANTUM MONTE CARLO SIMULATIONS BY

YUBO “PAUL” YANG

DISSERTATION

Submitted in partial fulfillment of the requirements
for the degree of Doctor of Philosophy in Physics
in the Graduate College of the
University of Illinois at Urbana-Champaign, 2020 Urbana, Illinois
  Doctoral Committee:
     Assistant Professor Bryan K. Clark, Chair
Professor David M. Ceperley, Director of Research
Professor James N. Eckstein
Professor Taylor L. Hughes

Abstract

This thesis explores properties of a mixture of electrons and ions using the quantum Monte Carlo method. In many electronic structure studies, purely electronic properties are calculated on a static potential energy surface generated by “clamped” ions. This can lead to quantitative errors, for example, in the prediction of diamond carbon band gap, as well as qualitatively wrong behavior, especially when light nuclei such as protons are involved. In this thesis, we explore different ways to include effects of dynamic ions and tackle challenges that arise in the process. We benchmarked the diffusion Monte Carlo (DMC) method on electron-ion simulations consisting of small atoms and molecules. We found the method to be nearly exact once sufficiently accurate trial wave functions have been constructed. The difference between the dynamic-ion and static-ion simulations can mostly be explained by the diagonal Born-Oppenheimer correction. We applied this method to solid hydrogen at megabar pressures and tackled additional problems involving geometry optimization and finite-size effects. The phase diagram produced by our electron-ion DMC simulations differ from previous DMC studies, showing 50 GPa higher molecular-molecular transition and 150 GPa higher molecular-to-atomic transition pressures. Both aforementioned studies forego the Born-Oppenheimer approximation (BOA) at hefty computational cost. Unfortunately, this makes it more difficult to compare our results with previous studies performed within the BOA. The remainder of the thesis tackle finite-size and ionic effects within the BOA. We calculated the Compton profile of solid and liquid lithium, achieving excellent agreement with experiment. Ionic effects of the liquid were included by averaging over disorder atomic configurations. Finite-size correction was crucial for the Compton profile near the Fermi surface. Finally, we tackled the finite-size error in the calculation of band gaps and devised a higher-order correction, which allowed thermodynamic values of the band gap to be obtained from small simulation cells. These advances mark important points along the path to the exact solution of the electron-ion problem. We expect that the better understanding of both the electron-ion wave function and its relation to finite-size effects obtained in this thesis can be crucial for future simulations of electron-ion systems.

To my friends and family.

Acknowledgments

First and foremost, I would like to thank my adviser, Prof. David Ceperley, for his unwavering support and guidance over the past six years. His rigor, honesty, curiosity, and openness deeply influenced the way I view and do science. I feel especially grateful for his fair treatment of students and collaborators. I was given much freedom to pursue my own interests and always felt like a valued member of the team. From these collaborations, I would like to thank Markus Holzmann and Carlo Pierleoni for many insightful discussions. I would like to thank the QMCPACK team, especially Raymond Clay III, who guided me through some initial hurtles getting to know the code and continue to be responsive and helpful. I look forward to continued collaborations, where I can learn from everyone and hopefully contribute more to our common interests.

I want to thank the CCMS program at LLNL, especially my mentor Miguel Morales. His enthusiasm of research is infectious. It was an invigorating summer interacting with Miguel and my fellow CCMS students Marnik Bercx, Ian Bakst, and Christopher Linderälv.

Of course, I would not be here without the loving care and support of my family and the wonderful company of my friends. I will not attempt to name everyone that was important for my life at Illinois, as I will invariably miss one of you even if I go on for pages. I do have to thank my officemate Brian Busemeyer for not only being a model colleague, who is always willing to answer stupid questions and entertain crazy ideas, but also for being a kind and wonderful friend. From our shared passion for coffee, badminton, swimming, and biking to our never-ending debates on Windows vs. Linux, Python 2 vs. 3, and functions vs. classes, I know we will remain colleagues and friends for years to come.

Thanks for funding from: U.S. Department of Energy (DOE) Grant No. DE-FG02-12ER46875 as part of the Scientific Discovery through Advanced Computing (SciDAC). DOE Grant No. NA DE-NA0001789. DOE Grant No. 0002911. The Blue Waters sustained-petascalecomputing project and the Illinois Campus Cluster, supported by the National Science Foundation (Awards No. OCI-0725070 and No. ACI-1238993), the state of Illinois, the University of Illinois at Urbana-Champaign, and its National Center for Supercomputing Applications and resources of the Oak Ridge Leadership Computing Facility (OLCF) at the Oak Ridge National Laboratory, which is supported by the Office of Science DOE Contract No. DE-AC05-00OR22725.

Table of Contents

List of Tables
List of Figures
List of Abbreviations
List of Symbols
1 Introduction
 1.1 The Electron-Ion Problem
  1.1.1 The Born-Oppenheimer Approximation
 1.2 Jellium
 1.3 Hydrogen
 1.4 Lithium, Diamond, and Silicon
 1.5 Thesis Outline
2 Methods
 2.1 Classical Monte Carlo
 2.2 Quantum Monte Carlo
  2.2.1 Path Integral Monte Carlo
  2.2.2 Variation Path Integral a.k.a. Reptation Monte Carlo
  2.2.3 Diffusion Monte Carlo
  2.2.4 Variational Monte Carlo
  2.2.5 Fermion Sign Problem
 2.3 Effective One-Particle Theories
  2.3.1 Hartree-Fock
   2.3.1.1 The Hartree-Fock Equations
   2.3.1.2 Koopmans Theorem
   2.3.1.3 Basis Set Error
   2.3.1.4 Beyond Hartree-Fock
   2.3.1.5 Static and Dynamic Correlation
  2.3.2 Kohn-Sham Density Functional Theory
   2.3.2.1 The Hohenberg-Kohn theorems
   2.3.2.2 The Kohn-Sham equations
   2.3.2.3 The Band Gap Problem
   2.3.2.4 Beyond Local Density Approximation
3 Slater-Jastrow wave function
 3.1 Historic Overview
 3.2 Definitions
 3.3 Cusp Conditions
 3.4 Gaskell RPA Jastrow
 3.5 Multi-Component RPA Jastrow
  3.5.1 Local Energy of Jastrow Wavefunction
  3.5.2 Equations that define the RPA Jastrow Pair Potentials
  3.5.3 Solving for the RPA Jastrow Pair Function
 3.6 Slater Determinant in Plane Wave Basis
  3.6.1 Momentum Distribution
  3.6.2 Static Structure Factor
  3.6.3 Example: Free Fermions
 3.7 Beyond Slater-Jastrow
4 Finite-size Effects
 4.1 Correction to the potential energy
  4.1.1 General Theory
  4.1.2 Homogeneous electron gas
  4.1.3 Inhomogeneous system
 4.2 Correction to the kinetic energy
  4.2.1 From momentum distribution
  4.2.2 From wave function
   4.2.2.1 Single-particle “shell filling” effect
   4.2.2.2 Two-body size effect
   4.2.2.3 Finite-temperature correction
5 Dynamic-ion DMC Study of Solid Hydrogen at Megabar Pressures
 5.1 Introduction
  5.1.1 Experiments
  5.1.2 Calculations
 5.2 Methods
  5.2.1 Candidate Structure Optimization
  5.2.2 Supercell Construction
  5.2.3 Wavefunction Optimization
   5.2.3.1 Electronic Components
   5.2.3.2 Ionic Components
 5.3 Results and Discussion
  5.3.1 Static-Lattice Energy, Pressure, and Local Energy Variance
  5.3.2 Effect of Dynamic Ions on Energy and Pressure
  5.3.3 Effect of Dynamic Ions on Electronic Correlation Functions
  5.3.4 Proton-proton g(r) and S(k)
 5.4 Conclusion
Appendices
A Nonadiabatic Coupling
B Minimum-basis H2
C Derivation of Higher-order Band Gap Correction
References

List of Tables

5.1 vdW-DF pressure-density (expressed in rs) relation of relaxed molecular candidate structures.
5.2 Optimized 72-atom non-diagonal supercell matrices.
B.1 symmetry-inequivalent electron repulsion integrals for H2 in STO-3G.

List of Figures

2.1 RHF electronic ground-state energy of H2 in STO-3G and correlation consistent (cc) basis sets as compared to the exact values calculated by Kolos and Wolniewicz (KW) [34]
2.2 A. J. Cohen, P. Mori-Sánchez, and W. Yang [36] explains that the self-interaction error in H2+ binding curve (left panel A) is due to the presence of fractional electron (left panel B), which leads to a delocalization error when LDA is used.
4.1 Fluctuating and static contributions from valence electrons in the conventional cell of bulk silicon.
4.2 Relative error of total energy vs. number of particles with PBC (up triangles) and TABC (squares) in 2D and 3D [72]
5.1 Partial phase diagram of hydrogen on log-log scale [77].
5.2 Phase diagram of o-D2 and HD below 200 K and 200 GPa [90].
5.3 Roton and vibron changes in p-H2 across the BSP transition [87].
5.4 Tentative phase diagram of solid hydrogen below 400 K [94].
5.5 vdW-DF optimized molecular candidate structures at target pressures. (a) lattice parameters (b) molecular bond length. C2/c and Cmca-12 each have two types of H2 molecules, whereas Cmca-4 has only one.
5.6 DMC optimized atomic structure c∕a ratio as a function of density. The blue points are DFT(PBE) optimized c∕a ratios.
5.7 DFT(vdW-DF) static-lattice enthalpy of optimized structures relative to C2/c-24. Thin solid lines are enthalpies of the Cmca-4, Cmca-12 and I41/amd using our optimized structures. Dashed lines are DFT(BLYP) enthalpies of Cmca-4 and Cmca-12 from Drummond et al. [3]. Dash-dot lines are DFT(vdW-DF) enthalpies from McMinis et al. [2].
5.8 Cubic vs. rhombus supercells. The black cell is the supercell. The gray cells are periodic images. The blue line points between nearest-neighbor images, while the red line between second-nearest neighbors. The yellow circle is the inscribed circle in the Wigner-Seitz cell (not shown) of the supercell.
5.9 Supercell radius as a function of density. rs is the Wigner-Seitz radius, which is determined by the average electron density 43πrs = ρ, where ρ = NeΩ, with Ω the supercell volume. RWS is the radius of the real-space Wigner-Seitz cell of the supercell. 2RWS is the minimum distance between periodic images.
5.10 Static structure factor S0(k) and momentum distribution n0(k) calculated using the Kohn-Sham determinant wave function at rs = 1.25. The black lines show the same functions for the free Fermi (FF) gas as reference.
5.11 Jastrow and back flow pair functions optimized at static-lattice minimum at 480 GPa. Color denotes the crystal structure. The solid line is the pair function for electron-proton, the dashed line for same-spin electrons, and the dotted lines for opposite-spin electrons.
5.12 Jastrow and back flow pair functions optimized using dynamic-lattice VMC. The dash-dotted line is the pair function for proton-proton correlation. All other notations are identical to those in Fig. 5.11.
5.13 Maximum overlap optimization of the proton wave function. (a) Optimization of the C2/c-24 proton wave function at rs = 1.269. (b) Optimized proton wave function exponent Cp of all candidate structures. The markers show the Cp value at the intercept of VMC and DMC. The lines are linear fits, but shifted down slightly to error on the side of oversampling. These fits are used to determine the Cp value in the dynamic-lattice calculations.
5.14 Fitting of reference equation of state. The solid symbols are DMC data from Ref. [3], while the line is a quadratic fit to the circled points. The bottom panels shows the error in the fit.
5.15 Static-lattice DMC energy and enthalpy relative to Drummond et al. [3] reference eq. (5.9). Relative energies are shown in meV per proton (meV/p). Each solid line is obtained using a fitted energy-density EOS, which is obtained by fitting the finite-size corrected (FSC) total energy as a quadratic function of ρ. The markers are finite-size corrected simulation data without performing a fit. P0 is the pressure calculated from the reference EOS P0 = dE0∕dΩ.
5.16 QMC variance vs. pressure. The lines are guides to the eye.
5.17 Energy changes from static-lattice to dynamic-lattice simulations as functions of electron density ρ = N∕Ω. (a) zero-point energies of the candidate structures. (b) total potential energy change due to dynamic protons. (c) electronic kinetic energy change due to dynamic protons. (d) proton kinetic energy. Color and marker label candidate structures. The solid lines are guides to the eyes.
5.18 Dynamic-lattice energy and enthalpy relative to reference. Lines are quadratic fits. The crossed out points are excluded from the fit.
5.19 DMC electronic pair correlation functions at rs = 1.25. Color denotes candidate structure, while the black lines are g(r) of the unpolarized homogeneous electron gas at the same density as parametrized by P. Gori-Giorgi, F. Sacchetti and G. B. Bachelet (GSB) [61].
5.20 Electron-electron static structure factor S(k).
5.21 Upper bound on inverse dielectric constant.
5.22 Proton-proton pair correlation functions around rs = 1.25 (480 GPa). The solid lines in (b) are quadratic fits to the smooth part of S(k). (c) Bragg peaks of the p-p S(k) (d) long wavelength limit S(0). |G1|≈ 1.81 bohr1, |G2|≈ 1.87 bohr1 for all three molecular structures. |G3|≈ 2.5 for C2/c-24 and Cmca-4 and 1.93 for Cmca-12. The atomic structure’s Bragg peak can be found between 1.83 and 1.88 bohr1. The same outliers as Fig. 5.18 are excluded from the quadratic fits in (d).
5.23 Pressure-density relation. The solid lines are clamped-ion vdW-DF pressures, whereas the symbols are QMC pressures. (a) (c) show static-lattice results, while (b) (d) show dynamic lattice results. Color denotes different candidate structures. The open and filled symbols represent VMC and linearly extrapolated DMC estimator results. They overlap on the absolute scale in (a) but can be seen to differ by 1 to 5 GPa on a relative scale in (b). The reference pressure-density relation P0 = dE0∕dv is calculated from the reference EOS E0(1∕v). The same clamped-ion reference is used for both static-lattice and dynamic-lattice results.
B.1 STO-3G

List of Abbreviations

BOA

Born-Oppenheimer Approximation.

DBOC

Diagonal Born-Oppenheimer Correction.

EOS

Equation Of State.

DM

Density Matrix.

QMC

Quantum Monte Carlo.

RPIMC

Restricted Path Integral Monte Carlo.

VPI

Variational Path Integral.

RMC

Reptation Monte Carlo

FP-DMC

Fixed-Phase Diffusion Monte Carlo.

VMC

Variational Monte Carlo.

KW

Kolos and Wolniewicz.

RHF

Restricted Hartree Fock.

CISD

Configuration Interaction Singles Doubles.

CAS

Complete Active Space.

ANO

Atomic Natural Orbital.

ECG

Explicitly-Correlated Gaussian.

KS-DFT

Kohn-Sham Density Functional Theory.

HK

Hohenberg and Kohn.

LDA

Local Density Approximation.

PZ

Perdew and Zunger.

HOMO

Highest-Occupied Molecular Orbital.

LUMO

Lowest-Unoccupied Molecular Orbital.

PBE

Perdew, Burke, and Ernzerhof.

HSE

Heyd, Scuseria, and Ernzerhof.

SBJ

Slater-Backflow-Jastrow.

RPA

Random Phase Approximation.

RDM

Reduced Density Matrix.

LLT

Liquid-Liquid Transition.

DAC

Diamond Anvil Cell.

MD

Molecular Dynamics.

PES

Potential Energy Surface.

GPa

Giga-Pascal

LP

Low-pressure Phase.

BSP

Broken-Symmetry Phase.

IR

InfraRed.

EQQ

Electric Quadrupole-Quadrupole.

GCTABC

Grand-Canonical Twist Average Boundary Condition.

KZK

Kwee, Zhang, and Krakauer.

MSD

Mean-Squared Deviation.

FWHM

Full Width at Half Max.

SDF

Spectral Density Function.

BFD

Burkatzki, Filippi, and Dolg.

QE

Quantum Espresso.

PPC

PseudoPotential Correction.

FSC

Finite-Size Correction.

BF

Back Flow.

ZPE

Zero-Point Energy.

List of Symbols

h

Planck constant = -h-
2π.

Ψ

many-body electron-ion wave function.

ψ

many-body electron wave function.

ϕ

one-body electron wave function.

χ

one-body ion wave function.

ri

3D position of electron i.

rI

3D position of ion I.

R

positions of all N electrons.

RI

positions of all NI ions.

ρ

density matrix or electron density.

N

number of electrons.

NI

number of ions.

Ω

total volume.

L

linear extent of the simulation cell.

rs

Wigner-Seitz radius.

kF

Fermi wave vector.

Γ

classical coulomb coupling parameter.

β

inverse temperature.

t

imaginary time unless otherwise specified.

τ

imaginary time step unless otherwise specified.

Z

canonical partition function.

Ĥ

electron-ion hamiltonian.

electronic hamiltonian.

λ

quantumness λ = ℏ2
---
2m.

Λkj

nonadiabatic coupling between electronic states k and j.

Λ

de Broglie wavelength.

P

pressure.

V

total potential energy.

T

total kinetic energy.

E

total energy.

ET

trial total energy.

ΨT

trial wave function.

σ2

variance of the local energy.

π

probability distribution.

Δ

band gap.

g(r)

pair correlation function.

S(k)

static structure factor.

U

negative exponent of the Jastrow wave function.

u(r)

pair contribution to U.

ωp

plasmon frequency.

𝜖k

dielectric function.

n(k)

momentum distribution.

J(p)

Compton profile.

Chapter 1
Introduction

1.1 The Electron-Ion Problem

The ultimate goal of this thesis is the accurate simulation of a many-body system of charged particles in the non-relativistic limit. This goal was not achieved, but some progress has been made. For the remainder of this thesis, the ground truth is assumed to be established by the exact solution of the Schrödinger equation (1.1) for the electron-ion wave function Ψ(R,RI) of N electrons and NI ions

ĤΨ(R,RI) = i d
dtΨ(R,RI), (1.1)

where the electron-ion hamiltonian consists of non-relativistic kinetic energies and Coulomb interactions

Ĥ = I=1NI  ℏ2
----
2mII2 i=1Nℏ2
----
2mii2 I=1NI i=1N Z k e2
--I-e--
|rI − ri| + 1
-
2 i=1N j=1,jiN  ke2
---e---
|ri − rj| + 1
-
2 I=1NI J=1,JINI                     k e2
                  ---e----
                  |rI − rJ|. (1.2)

The lower-case i,j and upper-case I,J refer to the electrons and ions, respectively. The lower-case ri labels a single electron position, whereas the upper-case R denotes the positions of all electrons R ≡{ri}. rI and RI play analogous roles for the ions. ZI is the atomic number of ion I. If any eigenstate of Ĥ can be constructed to arbitrary precision in a reasonable amount of time, which grows as a polynomial in the number of particles, then the many-body electron-ion problem can be declared solved. Unfortunately, even state-of-the-art methods struggle with just the ground state [13].

For equilibrium properties at high temperature, progress can be made by considering the Bloch equation (1.3) for the thermal density matrix ρ ieβEi|Ψi⟩Ψi|

Ĥρ = d--
dβρ, (1.3)

which results from the Schrödinger eq. (1.1) after a rotation from real to imaginary time τ = it, which is also the inverse temperature τ∕= β 1(kBT). The partition function is a trace of the thermal density matrix and reduces to the classical Boltzmann distribution at high temperature

Z = Tr(ρ)⇒
β→0 Z eβV . (1.4)

Further, low-temperature properties of boltzmannons and bosons can be exactly and efficiently calculated using the path integral method (Sec.2.2.1). The exact method is no longer practical when fermions, e.g., electrons, are involved. Nevertheless, impressive results have been obtained for hydrogen when only the ground electronic state is considered [4, 5]. A complete treatment of the full electron-ion hamiltonian eq. (1.2) is rarely attempted [6, 7].

The solution of the electron-ion problem would be an important milestone in computational condensed matter, because it is considered a quantitatively accurate model for the vast majority of solids and liquids in condensed matter experiments. Further, theoretically, it is a natural extension of the jellium model to multi-component system and provides a firm foundation upon which relativistic effects can be included, for example via perturbation [8]. Finally, the laplacian in the non-relativistic kinetic energy operator can be interpreted as a generator of diffusion in imaginary time. This makes it more straightforward to develop intuitive understanding of the quantum kinetic energy as well as to deploy powerful computational techniques such as diffusion Monte Carlo (Sec. 2.2.3).

1.1.1 The Born-Oppenheimer Approximation

Suppose all eigenstates of the electronic hamiltonian {ψk} are available at any ion configuration RI

(R;RI)ψk(R;RI) = Ek(RI)ψk(R;RI). (1.5)

Then, one can attempt to separate the electron and ion problems by expanding an eigenstate of the full hamiltonian Ĥ in the complete basis of electronic states

Ψl(R,RI) = k=0χ lk(RI)ψk(R;RI), (1.6)

where χlk(RI) are expansion coefficients with no dependence on electron positions. This expansion results in a system of Schrödinger-like equations, each describing the ions moving on an electronic energy surface Ej

(                )
   ∑   ℏ2   2
 −    2MI-∇ I + Ej
    Iχj Λkj = iχj, (1.7)

and coupled by the so-called nonadiabatic operator Λjk [9]. Derivation and behavior of Λjk are discussed in Appendix A. While exact, eq. (1.7) is difficult to solve because all electronic states are coupled via Λkj. To fully separate the electron and ion problems, one must approximate Λkj.

There are two common approximations to Λkj, the first is to set the entire matrix to zero, the second is to set only the off-diagonal terms to zero. Both approximations decouple (1.7), allowing the complete separation of electronic and ionic motions. Many different and sometimes conflicting names have been given to these two approximations, including Born-Huang, Born-Oppenheimer and adiabatic approximation. To fix nomenclature, I will call the all-zero approximation, Λjk = 0, j,k, the Born-Oppenheimer approximation (BOA). The diagonal terms Λjj are considered diagonal Born-Oppenheimer correction (DBOC). Non-zero off-diagonal elements are responsible for nonadiabatic effects.

1.2 Jellium

The jellium model eq. (1.8) brings the electronic problem into focus. It replaces the material-dependent ionic potential in the electronic hamiltonian with a rigid homogeneous background of positive charge and is an important stepping stone to the electron-ion problem. In Hartree atomic units

HEG = i=1N 1
2i2 + 1
2 i=1N j=1,jiN   1
|r-−-r-|
  i   j + Eeb + Ebb, (1.8)

where Eeb and Ebb are constants due to electron-background and background-background interactions. The isotropic background eliminates potential symmetry breaking interactions that can be introduced by a crystalline arrangement of the ions. The rigidity of the background also removes electron-ion coupling effects. This model was studied in great detail in the past century and its ground-state behavior was largely understood. Much progress has even been made regarding its excitations and finite-temperature properties.

There is only one length scale in the jellium model: the average electron-electron separation a. In units of bohr, this Wigner-Seitz radius rs = a∕aB determines the density of jellium.

1
n-  Ω
N- = 4π
-3-(rsaB)3 (1.9)

in 3D, where aB is the Bohr radius. The kinetic energy scales as rs2 (due to 2) while the potential energies scales as rs1 (for 1∕r potential), so rs measures the relative strength of potential to kinetic energy. In this sense, rs is the zero-temperature analogue of the classical coulomb coupling parameter

Γ e2n1∕d-
 kBT. (1.10)

The valence electron density in alkaline metals is rs 2, meaning the kinetic energy is important, so the electrons delocalize and form a liquid to minimize kinetic energy. At sufficiently large rs (100 in 3D and 30 in 2D), the potential energy dominates, so the electrons localize to form a Wigner crystal.

1.3 Hydrogen

Hydrogen is a logical starting point for solving the electron-ion problem. It has the simplest atomic structure and no core electrons. The non-relativistic Schrödinger equation eq. (1.1) and (1.2) should work well for hydrogen. Further, the ground state of its electronic hamiltonian can be compactly and accurately represented [10]. Without core electrons, no essential modification needs to be made to the hamiltonian eq. (1.2) for a practical simulation, e.g., pseudopotential.

At sufficiently high temperatures, the hydrogen plasma, equal mixture of isotropic positive charges (protons) and negative charges (electrons), is a straightforward generalization of the jellium model to two components. However, at low temperatures, the two-component analogue of the Wigner crystal, solid hydrogen, is surprisingly complex. Since hydrogen is the lightest element, its zero-point motion has large amplitude. The ion wave function explores a sufficiently large space to invalidate the harmonic approximation for lattice vibrations. Further complicating matters, one can expect a metal-to-insulator transition as well as an atomic-to-molecular transition that may or may not coincide as temperature or pressure is decreased. On top of all that, naturally occurring isotopes, e.g., deuterium, and spin isomers, e.g., para- and ortho-H2, allow the possibility of an intriguing blend of quantum effects at low temperatures.

Hydrogen is also interesting due to its practical relevance. Being the most abundant element in the observable universe, hydrogen and its isotopes are crucial for the understanding of stars and gas giants. Consider Jupiter, which contains insulating gaseous H2 in the outer envelope and liquid metallic hydrogen deep inside. If there was a first-order liquid-liquid transition between the two phases, then there would be an interface across which density changes discontinuously. Depending on the solubility of helium in the two phases, there is the possibility of helium rain across the interface and extra heat radiation due to this condensation [11]. Further, interior models of stars and gas giants rely on numerically accurate equation-of-state (EOS) of various chemical species involving hydrogen (H+, H, H2). A few percent change to the hydrogen EOS is enough to eliminate/create a rocky core for Jupiter [12].

In addition to its relevance in astrophysics, hydrogen is also important in energy applications. Accurate understanding of hydrogen EOS at high temperatures and pressures has obvious benefits to fusion experimental design. Even at low temperatures, hydrogen-rich compounds, at sufficiently high pressures, have recently smashed the superconducting transition temperature records held by the so-called “high-temperature” superconductors [13, 14].

Finally, the 85-year-old prediction for a low-temperature insulating-to-metallic transition of solid hydrogen, the Wigner-Huntington transition, is close to being established [1517]. Experimental observations [16, 17] and theoretical calculations [2, 18] are converging, although more experimental and theoretical characterizations are needed to settle current debates.

1.4 Lithium, Diamond, and Silicon

Despite complications introduced by core electrons, the jellium model is arguably better realized in the valence of alkaline metals, e.g., lithium and sodium, than it is in hydrogen. The heavier nuclei are less quantum and the core electrons screen their interaction with the valence electrons. This allows the harmonic approximation to be more widely applicable. More importantly, they are easier to handle in experiments than hydrogen and scatter X-rays more strongly, which facilitates precise experimental determination of lattice structure along with other properties. These advantages allowed us to obtain excellent agreement with experiment using an electron-ion QMC simulation performed within the BOA in chapter ??.

In addition to alkaline metals, elemental insulators, e.g., diamond carbon and silicon, remain important testing grounds for electronic QMC methods. Accurate and practical prediction of excitation energies using QMC is still an active area of research. We make some progress in chapter ?? by reducing finite-size error in bandgap calculations.

1.5 Thesis Outline

The remainder of this thesis is organized as follows. In chapter 2, I start by introducing the Monte Carlo methods that we use to accurately treat electron correlation in the electronic problem as well as to solve the full electron-ion problem at times. I then introduce the effective one-particle electronic structure methods used to generate trial wave functions for the aforementioned QMC methods. Chapter 3 displays the form and properties of the Slater-Jastrow wave function in detail, while chapter 4 discusses many-body finite-size correction, sometimes based on the properties of the many-body wave function. Both wave function form and finite-size correction are crucial for accurate and practical QMC simulations. The next four chapters display QMC results for a few simple electron-ion systems. Chapter ?? benchmarks the QMC method as a complete solver for the electron-ion problem without invoking the BOA on small atoms and molecules. Chapter 5 applies this dynamic-ion QMC method to solve for the ground state of solid hydrogen. Chapter ?? considers the effect of the ions on the momentum distribution of the valence electrons in lithium within the BOA. Finally, chapter ?? presents an improved finite-size correction to the (purely electronic) bandgap calculated in QMC.

Chapter 2
Methods

The main method used throughout this thesis is ground-state quantum Monte Carlo (QMC). This chapter provides a physically motivated introduction to this method. In the first part, I start from the familiar finite-temperature classical Monte Carlo method, then describe its generalization to a quantum system, finally take the zero-temperature limit. The second part of this chapter describes practical methods for constructing a many-body trial wave function, which is a crucial ingredient in many QMC methods.

2.1 Classical Monte Carlo

The Monte Carlo methods mentioned in this thesis perform high-dimensional integrals by using random numbers to sample probability distributions. These distributions must be non-negative in the entire domain of “states” over which they are defined. In classical mechanics, a “state” of N particles in 3 dimensions is labeled by the positions R ≡{ri} and momenta P ≡{pi} of the particles i = 1,,N. The classical partition function for the canonical ensemble

Z Tr(eH∕kBT) =   1
N!h3N- d3NRd3Np eH(R,P)∕kBT, (2.1)

where H(R,P) is the hamiltonian, kB is the Boltzmann constant, h is the Planck constant, and T is temperature. For N distinguishable non-relativistic particles with mass m, the kinetic contribution -p2
2m to eq. (2.1) can be integrated analytically, giving

Z = --1-3
N !Λ d3NReβV (R), (2.2)

where V (R) is the potential energy of the N-particle system, β kBT, and Λ the de Broglie wavelength

Λ = (         )
  ---h2---
  2πmkBT12. (2.3)

All equilibrium statistical mechanics properties can be calculated from the partition function, so the entirety of classical equilibrium statistical mechanics reduces to the problem of evaluating the 3N-dimensional integral in eq. (2.2) and its derivatives. Monte Carlo methods are ideally suited to evaluating high-dimensional integrals, because the amount of computation does not increase as an exponential in the number of dimensions as in a brute-force quadrature approach.

To calculate a property in the canonical ensemble, ones takes the trace

⟨O⟩ =   (  −βH )
Tr-Oe-----
 Tr(e−βH), (2.4)

where ⟨⟩ denotes ensemble average. Limiting to local observables that can be evaluated on the particle coordinates, e.g., total potential energy V (r) and pair correlation function g(r)

⟨O ⟩ = d3NR∫---e−βV(R)----
  d3NR ′e− βV (R ′)O(R) = d3NRπ(R)O(R), (2.5)

where π(R) is the Bolzmann distribution

π(R)      −βV(r)
∫--3eN--′−βV(R′)
  d  R e eβV (R). (2.6)

Monte Carlo estimation of ⟨O⟩ works by sampling particle configurations from the Bolzmann distribution π(R) and accumulating the average O(R)

⟨O⟩ = limNs→∞1--
Ns i=1Ns O(Ri). (2.7)

How does one sample a generic multi-dimensional probability distribution such as π(R)? An excellent answer was given a 1953 paper authored by Metropolis et al. [19]. The Metropolis algorithm, designed by the Rosenbluths supervised by the Tellers, works by constructing a Markov chain having π(R) as its stationary state. This is achieved by a rejection method that maintains detailed balance

π(R)P(R R) = π(R)P(R′→ R), (2.8)

where P(R R) is the Markov chain transition probability from state R to R. The Metropolis algorithm breaks P into two steps: proposal and acceptance

P(R R) = T(R R)A(R R), (2.9)

with the following accept/reject criteria: For any transition probability used to propose the state change T(R R), accept the change with probability

A(R R) = min(                 )
 1, π(R-′)T-(R-′ →-R-)
    π(R )T(R →  R′). (2.10)

Using eq. (2.9) and (2.10) to prove eq. (2.8) is a good way to appreciate the design of this acceptance probability.

Mathematically, π(R) is the unique stationary state of the Markov chain constructed by the Metropolis method so long as P(R R) is ergodic. That is, there is finite probability of reaching any state Rstarting from any state R using the transition rule P. In practice, however, a simulation can be stuck in a meta-stable state for its entire duration, for example, due to a bad initial condition. Careful monitoring and checking of convergence is a must in any serious Monte Carlo simulation.

2.2 Quantum Monte Carlo

I will start with the general, albeit somewhat complicated, path integral Monte Carlo (PIMC) method, because it rigorously takes temperature into account and connects well with classical Monte Carlo. Then, I will describe ground-state methods as limits and efficiency tricks to specialize the path integral method to the ground state. While contrary to the historic progression of these methods, I find this perspective helpful for relating the methods and visualizing them in their respective niches.

2.2.1 Path Integral Monte Carlo

The quantum partition function for the canonical ensemble needs to trace over discrete N-particle eigenstates, rather than 2N 3-dimensional variables as in eq. (2.1)

Z Tr(eĤ∕kBT) = Tr(                   )
 ∑∞  − E∕k T
     e  i B  |Ψi⟩ ⟨Ψi|
  i=1, (2.11)

where Ei and |Ψi⟩ are the eigenvalues and eigenstates of the hamiltonian H. To make contact with classical mechanics, we put the density matrix (DM) for distinguishable particles in position basis (first quantization)

ρD(R,R;β) ⟨         ⟩
 R|e−βH|R ′, (2.12)

where β kB1T--. Then the partition function becomes

Z(β) = d3NRρ D(R,R;β), (2.13)

which can be exactly factorized into two pieces

Z(β) = d3NRd3NRρ D(R,R;τ)ρD(R,R;β τ). (2.14)

This factorization can be repeated until the temperature becomes high enough (τ 0) that a semi-classical approximation to ρD(R,R;τ) is accurate. Given translation symmetry along imaginary time, β is typically broken down into M equal-length pieces, i.e., τ = β∕M. For N non-relativistic particles each having mass m, ρD can be calculated as an integral over a discretized path of particle coordinates {Rm}

ρD(R0,RM;β) = limM→∞ dR1dRM1 exp(∑M                 )
     ln ρpD(Rm −1,Rm )
 m=1. (2.15)

The primitive approximation for the high-temperature density matrix ρDp satisfies

dN
---
2ln(4πλτ) lnρDp(R m 1,Rm) = (R     − R  )2
---m−1----m--
     4λτ + τV (Rm), (2.16)

where λ  ℏ2
---
2m is the quantumness of the particle and d is the number of spatial dimensions. Thus,

ρD(R0,RM;β) = limM→∞(4πλτ)dNM∕2 exp(   M∑  [            2          ])
  −     (Rm-−1 −-Rm-) + τV(Rm )
   m=1       4λτ. (2.17)

The main advantage of eq. (2.15) is that it turns the partition function eq. (2.13) into an integral of a product of high-temperature DMs over all closed paths

Z = limM→∞ dRdR1dRM1ρD(R,R1;τ)ρD(R1,R2;τ)ρD(RM1,R;τ). (2.18)

Each closed path can be visualized as a collection of ring polymers, one for each particle. The linear extension of a ring polymer is proportional to the particle’s de Broglie wavelength Λ = √ -----
  4πλβ eq. (2.3). For distinguishable particles, the integral needed to evaluate the quantum partition function eq. (2.18) poses no essential difficulty to a Monte Carlo method when compared with its classical counterpart eq. (2.2). One simply has to integrate M classical systems, which are coupled by the spring-like kinetic energy term in eq. (2.16). Each classical system is typically referred to as a slice of imaginary time or a bead on the ring polymer. Converged results is obtained in the zero time step τ 0, equivalently the infinite slice M →∞ limit. The the primitive approximation eq. (2.16) to the exact density matrix is correct only to O(τ), so a large number of beads is needed, resulting in slow simulations. Better approximations can be constructed to be correct to O(τ2), for example the pair-product form in Ref. [20]. However, if the particles are identical bosons, then one has to consider particle permutations along the path (fermions pose an additional essential problem, see Sec. 2.2.5)

ρB(R0,RM;β) = -1-
N ! 𝒫ρD(R0,𝒫RM;β), (2.19)

where 𝒫RM contains the same N coordinates as RM, but with the particles relabeled. This permutation can happen via any number of 1-, 2-, and up to N-particle exchanges between adjacent time slices along the path. Thus, the state space of bosonic path integral is much larger than that of boltzmannic path integral. Efficient sampling of permuting paths is a significant technical challenge. Fortunately, no uncontrolled approximation has been introduced and exact simulations are possible for both bolzmannons and bosons via the Monte Carlo method [20, 21].

2.2.2 Variation Path Integral a.k.a. Reptation Monte Carlo

Even with accurate approximation to the high-temperature density matrix, the ground state is still costly to study using the path integral formalism presented so far, because a large number of time slices have to be included to approximate β →∞. Fortunately, one can still efficiently study this zero-temperature limit with the help of a trial wave function |ΨT ⟩, so long as it is non-negative. The ground-state “partition” function has only one term

Z0 = limβ→∞⟨Ψ |e−βH|Ψ ⟩
   0      0, (2.20)

where |Ψ0⟩ is the ground state of the hamiltonian H. For sufficiently large β, any trial wave function not orthogonal to the ground state ⟨ΨT|Ψ0⟩0 will be projected to the ground state by eβH, so

Z0 = limβ→∞⟨    − βH − βH   ⟩
 ΨT|e 2  e 2 |ΨT⟨    −βeH  −βeH   ⟩
 ΨT|e    e    |ΨT, (2.21)

for some βe large enough to be considered “equilibrated”. Performing path discretization as before

Z0 = limM→∞ dRMdR0dRMΨT(R M)ρ(RM,RM+1;τ)ρ(RM1,RM;τT(RM), (2.22)

where τ = βe∕M. βe can be small if |ΨT⟩ is a good approximate to the ground state |Ψ0⟩. In this sense, |ΨT ⟩ΨT| plays the role of a low-temperature density matrix to quickly close a long path, although its temperature is ill-defined. No permutation needs to be sampled because quantum statistics are encoded in the trial wave function. However, translation symmetry along imaginary time is broken. The 2M + 1 time slices each sample a different probability distribution. Observables that do not commute with the hamiltonian are unbiased only when evaluated in the middle section Rm, where |m| is small. This is because Rm needs to be sufficiently separated from the trail wave function slices RM and RM to be considered the zero-temperature limit. The trial wave functions at the ends and the DMs in the middle of eq. (2.22) must all be non-negative for the integrand to be interpreted as a probability distribution for the path {RM,,RM}. This path is open in general (RMRM) and can be visualized as a “reptile”. This method was first mentioned as variational path integral (VPI) [20], but later popularized as reptation Monte Carlo (RMC) [22]. While the RMC method can be efficient, it still requires all M classical systems to be stored in memory at one time and intelligent Monte Carlo moves to change the reptile without ergodicity problems. This makes RMC more troublesome to implement than classical Monte Carlo or molecular dynamics, because the entire reptile, containing O(M) classical systems have to be updated to generate O(1) new decorrelated sample for the pure estimator eq. (2.23).

2.2.3 Diffusion Monte Carlo

The diffusion Monte Carlo (DMC) method can be viewed as a simplification of RMC. When calculating a ground-state observable using eq. (2.21), the pure estimator

⟨  ⟩
 Oˆp ⟨                   ⟩
 ΨT |e− βeH ˆOe−βeH|ΨT⟨       ⟩
 Ψ0|ˆO|Ψ0 (2.23)

is an unbiased ground-state estimate of Ô whether it commutes with the hamiltonian or not. We can forgo the pure estimator for a simpler algorithm. Consider the mixed estimator

⟨  ⟩
 Oˆm ⟨              ⟩
  ΨT|ˆOe−βeH|ΨT⟨       ⟩
 ΨT |ˆO |Ψ0. (2.24)

Equation (2.24) has the advantage that the operator can be immediately applied to a trial wave function, which is known at the beginning of the calculation. Further, the Ψ0 on the r.h.s. can be interpreted as being propagated from ΨT using the imaginary-time propagator

Û(t) = etH. (2.25)

For any t > βe, the mean of the observable eq. (2.24) should be stationary. The algorithm as described so far is similar to classical molecular dynamics. One starts with a trial wave function at t = 0 and propagates it along imaginary time. After some initial equilibration period, the mixed estimator fluctuates around some stationary mean. One then runs for “longer” and accumulate statistics.

For identical non-relativistic particles having mass m, Û in coordinate basis is the Green function for imaginary-time Schrödinger equation

G(R′← R;t) = ⟨ ′ ˆ    ⟩
 R |U(t)|R = ⟨  ′ −t(λ∇2+V )  ⟩
  R |e         |R
limτ0G(R′← R;τ) = (                       )
               (R-−-R′)2
|( (4πλτ)−dN∕2e−   4λτ   |)(       )
 e−τV (R). (2.26)

The two terms in eq. (2.26) are the Green function for diffusion and weight accumulation. The quantumness λ =  2
ℏ2m- of the particle determines its diffusion constant in imaginary time. Lighter particles diffuse “faster”. One can, in principle, start with any classical system R, apply the Green functions repeatedly to update R, and eventually end up sampling the mixed distribution ⟨ΨT |Ψ0 ⟩.

While eq. (2.24)-(2.26) contain the main idea behind the DMC method, they do not result in a practical algorithm. The weight of the classical system due to the potential term goes to zero or infinity exponentially fast, especially when two charged particles coalesce. For a stable algorithm, one can modify the Green function eq. (2.26) to more directly sample the mixed distribution

f(R,t) ΨT(R)et(HET)Ψ T(R)  =
t→ ∞ ΨT(R 0(R), (2.27)

where a trial energy ET is introduced to stablize the potential term. Substitute ΨT1f in place of Ψ into the imaginary-time Schrödinger equation tΨ = HΨ, where H = λ2 + V , we obtain

tf = ΨTHΨT1f ⇒− tf = λ(v)f + (EL ET)f, (2.28)

where the local energy EL H-ΨT-
 ΨT and the drift vector

v T1Ψ T = lnΨT2. (2.29)

After this importance-sampling transformation, the Green function eq. (2.26) is now modified to have three contributing processes: diffusion, drift, and weighting. The pure diffusion process becomes a drift-diffusion process guided by the trial wave function. Further, the weighting by the bare potential energy becomes weighting by the local energy. Given suitably designed trial wave function, EL can be made continuous even if the original potential V contains divergences, e.g., in the coulomb interaction. In practice, the mixed distribution is approximated by an ensemble of walkers

f(R,t) m=1Nw δ(RRm), (2.30)

and the trial energy ET is adjusted every so often to keep the population of walkers Nw from explosion and extinction. Equation (2.28) defines the DMC algorithm. While not strictly necessary, one typically adds a Metropolis rejection step using the Green function G as transition probability T in eq. (2.10). This ensures that the algorithm samples the desired probability distribution at any finite time step, where the Green function is approximate [23].

For bosons and Bolzmannons, DMC gives the exact ground-state energy of H in the limit of infinitesimal time step and uncontrolled walker population. Observables that do not commute with the hamiltonian will suffer a mixed-estimator error, which vanishes as the trial wave function approaches the ground state.

2.2.4 Variational Monte Carlo

Variational Monte Carlo (VMC) can be viewed as a limit of DMC at zero projection time, i.e., no branching. I define VMC as a Monte Carlo algorithm that calculates the expectation value of an operator using a fixed trial wave function ΨT

⟨  ⟩
  ˆO = ⟨        ⟩
 ΨT |ˆO|ΨT. (2.31)

When VMC is used to calculate the expectation value of the hamiltonian (Ô = ), the resultant energy is an upper bound to the true ground-state energy by the variational principle

EV ⟨    ˆ    ⟩
 ΨT |ˆℋ|ΨT = ⟨ΨT |EL (R;ΨT )|ΨT ⟩E0, (2.32)

where the local energy serves as a local estimator for the total energy

EL(RT) ˆℋ Ψ
---T-
 ΨT(R). (2.33)

The variational energy EV is often taken as a measure of the quality of the trial wave function. In variational optimization, one changes parameters in ΨT to lower EV . However, EV is only one number and is far from a complete descriptor of the 3N-dimensional many-body wave function. Another, arguably more powerful, measure of the quality of ΨT is the variance of the local energy eq. (2.33)

σ2 T] ⟨ΨT |(EL(R; ΨT)− EV )2|ΨT ⟩0. (2.34)

This variance will be zero if ΨT is any eigenstate of . Further, for systems with a gap Eg, σ2 and EV also provide a lower bound for the ground-state energy by eq. (6.16) in Ref. [24]

EV σ2∕E g E0 ET. (2.35)

The VMC algorithm is particularly simple for a local observable. In this case, eq. (2.31) becomes an integral over the 3N-dimensional particle coordinates

⟨O⟩ = drO (r)|ΨT (r)|2
--⟨Ψ--|Ψ--⟩--
    T   T, (2.36)

which is easily evaluated by sampling r from the probability distribution

P(r) = |Ψ  (r)|2
--T-----
⟨ΨT |ΨT ⟩, (2.37)

and accumulating O(r). How does one sample a generic probability distribution like P(r)? One can use the Metropolis algorithm to devise a Markov chain with P(r) being the stationary distribution. Alternatively, P(r) can be set to be the stationary distribution of a dynamical process, e.g., Fokker-Planck dynamics [25].

Suppose we wish a drift-diffusion process governed by

∂f-
 ∂t = i=1Nλ-∂--
∂xi(           )
  -∂--− vi(r)
  ∂xif (2.38)

to have P(r) as its stationary distribution

limt→∞f(r,t) = P(r) ∝|ΨT|2. (2.39)

We can set each term in sum of the r.h.s. of eq. (2.38) to vanish

i2P = P i vi + vi iP, (2.40)

which gives the drift vector

vi = ∇ f
-i--
 f = 2∇  Ψ
--i-T-
 ΨT. (2.41)

v pushes a walker towards peaks of P(r), making the sampling process more efficient than a random move. In fact, no accept/reject procedure is necessary. The correct stationary distribution P(r) will be reached so long as the time step is small enough to accurately approximate the Green function for each step. The Langevin equation needed to solve eq. (2.38) is

∂r(t)
 ∂t = λv(r(t)) + η, (2.42)

where η is a multidimensional Gaussian with a mean of zero and a variance of 2λ. In this light, the VMC algorithm is more akin to stochastic classical molecular dynamics than Monte Carlo. However, the most efficient algorithm is obtained when the Fokker-Planck formulation is combined with Metropolis Monte Carlo. By introducing a metropolis accept/reject procedure at each step of the Fokker-Planck dynamics, we eliminate the time step bias because detailed balance is enforced to sample |ΨT|2. Another way to view this is that Fokker-Planck dynamics provides efficient drift-diffusion moves for an exact Monte Carlo method.

Equation (2.38) defines an efficient implementation of the VMC algorithm. Interestingly, the governing equation of DMC eq. (2.28) without the branching term is identical to that of VMC eq. (2.38) when the same trial wave function is used for drift eq. (2.41). This implies that the drift-diffusion term in the DMC Green function performs sampling of the trial wave function only. The local energy term in eq. (2.28), when accumulated over the drift-diffusion process, is responsible for imaginary-time projection.

2.2.5 Fermion Sign Problem

An essential difficult arises when one applies the path integral formalism to fermions. Even- and odd-permutations contribute to the fermion DM with opposite signs

ρF(r0,rM;β) =  1
N-! 𝒫(1)𝒫ρ D(r0,𝒫rM;β). (2.43)

ρF is no longer positive definite and cannot be interpreted as a probability distribution to be sampled by Monte Carlo. The canonical workaround is to sample the absolute value of the fermionic DM, which is the bosonic DM, and keep the sign as an observable. In this way, a fermionic observable is calculated as the ratio between a signful bosonic observable and the bosonic average of the permutation sign

⟨ ⟩
 ˆOF   (      )
Tr Oˆe −βH
----------F-
 Tr(e−βH)F =    (       )
Tr  ˆOσˆe− βH
------------B
  Tr(ˆσe−βH)B =    (       )
Tr  ˆOˆσe− βH   ∕ZB
------------B----
  Tr(ˆσe−βH)B ∕ZB = ⟨   ⟩
 Oˆˆσ
-----B-
 ⟨ˆσ⟩B. (2.44)

While mathematically exact, this leads to the well-known fermion sign problem, where the denominator in eq. (2.44), the average sign, goes to zero exponentially fast as system size N and inverse temperature β increase. To see this, consider the total free energies of N bosons vs. N fermions governed by the same hamiltonian H. The average sign can be written as an exponential of the free energy difference

⟨ˆσ⟩B = ZF-
ZB = e−βFF-
e−βFB = eβ(FFFB). (2.45)

Since the total free energy is extensive, the exponent in eq. (2.45) is proportional to βN. At zero Kelvin, all permutations are equally likely and the average sign is exactly zero. At the degeneracy temperature and above, signful PIMC is often possible, but comes at a hefty computational cost even for very small systems, e.g., O(10) particles. A practical workaround for the sign problem in path integral is the restricted path approximation [2628]. This approximation uses a trial density matrix to restrict the space of paths that can be sampled. Any Monte Carlo move that constructs a path which crosses the node of the trial density matrix is rejected. This restricted path integral Monte Carlo (RPIMC) method was proved to be exact if the node of the trial density matrix is exact [27].

The sign problem manifests rather differently in DMC than it does in PIMC. If one runs the bosonic DMC algorithm eq. (2.28) using the absolute value of a fermionic trial wave function |ΨT| as guiding function, then the drift vector eq. (2.29) will diverge as a walker approaches the node of ΨT, pushing the walker away. This trapping effect greatly diminishes the chance that a walker can cross the node. Thus, the walker distribution will approach the bosonic solution in this positive nodal pocket. However, once a walker does cross the node, it quickly gets pushed to regions with high |ΨT|2 and branches to solve a similar bosonic problem in the newly found negative nodal pocket. Walkers in the positive nodal pocket contribute to estimators with a positive sign, whereas walkers in the negative pocket contribute with a negative sign. While the final average is the exact fermionic estimator, it is the difference between two large values and its noise diverges exponentially fast with system size and projection time. Exact fermionic simulation is possible using the release-node method [29], but a highly-accurate trial wavefunction is required and much care must be taken to efficiently converge the calculation before noise takes over.

The practical workaround for the sign problem in DMC is the fixed-node approximation (for real-valued wave functions). In this approximation, no walker is allowed to cross the node of the trial wave function. This effectively adds to the hamiltonian an infinite potential barrier at the node of ΨT. The bosonic problem is exactly solved in the nodal pocket within this barrier, while the rest of Hilbert space is constructed via antisymmetry of the wave function. This restricted random walk effectively constructs a fixed-node wave function ΨFN as an approximation to the true ground state, so the fixed-node DMC (FN-DMC) total energy is variational EFN E0. The exact ground-state energy can be obtained if the node of the trial wave function is exact. When generalized to systems with complex-valued trial wave function, an analogous work around is the fixed-phase DMC method (FP-DMC), which forbids walkers to move past phase-change boundaries of the trial wave function. FP-DMC is so similar to FN-DMC that they are often not distinguished as different methods in the literature.

For a general fermionic system, the sign problem is present in all known QMC methods in some form. However, the sign problem can be absent in a particular system, for example due to particle-hole symmetry in a half-filled Hubbard model, and can be alleviated at finite-temperature. Unfortunately, no known polynomial-scaling algorithm can solve the sign problem in all cases.

2.3 Effective One-Particle Theories

The most widely used ground-state QMC method, FN-DMC, requires a trial wave function ΨT to start the imaginary time projection process. The node of ΨT directly controls the one uncontrolled error, the fixed-node error, in the DMC total energy. Even bosonic details of ΨT can matter to observables that do not commute with the hamiltonian due to the mixed-estimator error. Further, the complexity of ΨT affects the efficiency of a DMC run, because ΨT and many its derivatives need to be evaluated at every step of the algorithm. Therefore, accurate and compact trial wave functions are crucial for practical high-accuracy DMC simulations.

While it is possible to construct such trial wave functions analytically [10, 30, 31], the process is rather involved and requires deep understanding of the particular system being simulated. For a generic material, it is much easier to base the many-body trial wave function on existing mean-field theory or effective one-electron theory that approximately include some effects of electron correlation. This chapter introduces one theory in each of these two categories.

2.3.1 Hartree-Fock

2.3.1.1 The Hartree-Fock Equations

The Hartree-Fock (HF) equations are a set of equations that couple spin orbitals in a determinant wave function. They can be obtained by minimizing the energy of a Slater determinant with the constraint that spin orbitals remain orthonormal. If each molecular orbital a is written as a linear combination of a set of basis functions {ϕμ}

ψa = μ=1KC μaϕμ, (2.46)

then the constraint optimization problem can be converted to a set of linear equations, resulting in an eigenvalue problem. However, the eigenvectors of this linear problem are needed to construct the problem to be solved. This self-consistency requirement makes the HF equations non-linear, thus requiring an iterative solver. Once an initial guess for the eigenvectors Cμa have been chosen, one can construct the linear problem to be solved via a Fock matrix. For a spin-unpolarized system N = N = N∕2, the restricted Hartree-Fock (RHF) solution is defined by the first N∕2 eigenvectors of the Fock matrix

Fμν = Hcoreμν + λσPλσ[         1       ]
 (μν|σλ )− -(μλ|σν)
          2, (2.47)

where Pλσ = 2 a=1N∕2CλaCσa is the density matrix of the trial states. The Coulomb integral notation

(μν|λσ) = ∫∫dr1dr2ψμ(r 1)ψν(r1)---1----
|r1 − r2|ψλ(r 2)ψσ(r2). (2.48)

Hcoreμν is the one-electron part of the Hamiltonian expressed in the given basis

Hcoreμν = dr1ϕμ(r 1)(  1     ∑      Z    )
  −- ∇21 −    ----A----
   2      A  |r1 − RA |ϕν(r1). (2.49)

The RHF total energy is the expectation of the electronic hamiltonian in the determinant wave function

ΨRHF = D({ψa})D({ψa}), (2.50)
E0RHF ⟨              ⟩
 ΨRHF  |ℋ ˆ|ΨRHF = 2 ⟨ψa|Hcore|ψa⟩
     μν + a=1N∕2 b=1N∕2⟨ab||ab⟩, (2.51)

where the exchange integral is defined as

⟨ab||ab⟩ dr1dr2ψa(r 1)ψb(r 2)r121(1 −𝒫 12)ψc(r1)ψd(r2). (2.52)

Importantly, E0RHF is not the sum of the eigenvalues of the Fock operator

2 a=1N∕2𝜖 a = 2 ⟨ψ |Hcore|ψ ⟩
  a  μν   a + 2 a=1N∕2 b=1N∕2⟨ab||ab⟩ = E 0RHF a=1N∕2 b=1N∕2⟨ab||ab⟩, (2.53)

because it double counts Coulomb interaction. The root cause if that both 𝜖a and 𝜖b include the Coulomb interaction energy between orbitals a and b. Not being able to sum the eigenvalues to obtain the total energy is a minor inconvenience for the fact that these eigenvalues have the physical meaning of electron/hole excitation energies. Since the HF method constructs a determinant as trial wave function for the exact electronic hamiltonian, the HF energy is variational E0RHF E0.

The RHF equations can be generalized to treat open-shell systems. If NN, but the spatial part of each occupied orbital is required to be identical for the and electrons. Then, the method is known as restricted open-shell HF (ROHF). If the occupied orbitals are further allowed to differ, then the method is unrestricted (UHF). An application of RHF to the isolated H2 molecule is detailed in Appendix B.

2.3.1.2 Koopmans Theorem

HF theory can be used to study excitations from the ground state. Consider removing an electron from orbital δ N, thus creating a hole (h). The wave function for the (N 1)-electron system is [32]

     ⟩
|Ψ (hH,δF) = âδ|ΨHF ⟩. (2.54)

In the frozen orbital approximation, where the orbitals of the remaining electrons cannot respond to the removed one, the energy of this wave function can be shown to differ from the ground state by 𝜖δ, the HF eigenvalue of the orbital being emptied

EHF(h,δ) ⟨      ⟩         ⟩⟩
  |Ψ (hH,Fδ) |ℋ ˆ| |Ψ (Hh,Fδ) = EHF 𝜖δ. (2.55)

T. Koopmans [33] first proved this for the highest occupied molecular orbital (HOMO) as an approximation to the ionization energy, although the above derivation is general for any orbital. Following Chapter 2.2.3 in Ref. [32], one can similarly calculate the energy of an N-electron system that differs from the HF ground state by an electron-hole excitation

       ⟩
|Ψ (eH,γF;h,δ) = âγâ δ|ΨHF ⟩, γ > N, δ N. (2.56)
EHF(e,γ;h,δ) = E HF + 𝜖γ 𝜖δ Δγδ, (2.57)

where Δγδ > 0. For a stable HF solution, eq. (2.57) leads to a coulomb gap in the HF density of states

N(e) < 2d1d( 𝜖)
 -2
 ed(e e F)d1, (2.58)

where d is the number of spatial dimensions, and 𝜖 is the dielectric constant of the insulator. That is, the HF density of states must vanish at least as fast as (e eF)d1 around the Fermi energy eF.

Koopmans theorem is only applicable relative to the HF ground state. For example, one cannot repeatedly apply Koopmans theorem to reconstruct the total energy of the system by stripping one electron at a time. As shown in eq. (2.53), the sum of HF eigenvalues double counts the Coulomb interaction energy.

2.3.1.3 Basis Set Error

To obtain a converged HF solution, the basis set used to represent the spin orbitals, i.e. {ϕμ} in eq. (2.46), must be complete. In practice, one can approach this complete basis set limit by using a sequence of basis sets increasing in size. The correlation-consistent (cc) basis sets are a widely used standard for this purpose. The convergence of the total energy of an H2 molecule is shown in Fig. 2.1 and compared to the exact references obtained by Kolos and Wolniewicz (KW) [34]. The total energy is converged on the scale of the plot using a triple-zeta (TZ) basis, which contains three basis functions per atom, and above. All RHF energy curves are at least 40 mha above the exact solution of the Schrödinger equation including electron-electron correlation, in accordance with the variational principle. Further, even after basis-set convergence, the RHF energy is still quite different from the exact values. Remarkably, besides an overall shift, the HF curve agrees well with the KW curve at equilibrium bond length 1.4 bohr and below. However, at larger bond lengths the HF energy increases faster than the KW curve and is above the exact total energy at infinite separation. This is because the two electrons are forced into the same orbital, when they should each reside close to a different proton. This correlation between unlike-spin electrons is completely absent from the HF method. The difference between the exact and the RHF energies defines the correlation energy.

pict

Figure 2.1:RHF electronic ground-state energy of H2 in STO-3G and correlation consistent (cc) basis sets as compared to the exact values calculated by Kolos and Wolniewicz (KW) [34]
2.3.1.4 Beyond Hartree-Fock

Even when the complete basis set limit is reached, the HF solution is still not the exact electronic ground state due to its neglect of electron correlation. One way to account for correlation effects is to perform a determinant expansion eq. (2.59). The unoccupied virtual orbitals can be used to construct N-electron determinants that differ from the HF ground state by changing orbital occupation. These determinants form a many-body basis, in which any wave function can be expressed as a linear combination. This leads to the configuration interaction (CI) expansion, where the exact electronic ground state is expanded as a sum of determinants

ψ0 = limM→∞ i=0Mc i|Di⟩. (2.59)

If all determinants that differ by one particle-hole excitation from reference are considered, then we obtain a CI singles (CIS) expansion. If these and all determinants with two particle-hole excitations are considered, then the expansion is CI singles and doubles (CISD), etc.. If all excitations among a set of “active” orbitals are considered, then the expansion is said to involve the complete active space (CAS).

2.3.1.5 Static and Dynamic Correlation

The ground state is said to have static correlation if one or more determinants in the exact expansion eq. (2.59) are nearly degenerate with the reference determinant. This will happen if there are virtual orbitals nearly degenerate with the highest occupied molecular orbital. In contrast, the system has dynamic correlation if the ci coefficients are small but non-zero for many determinants with high levels of excitation. Dynamic correlation is often attributed to strong local correlation such as the electron-electron cusp condition. The current definition of static and dynamic correlations is not precise [35]. I introduce the above working definitions, because static correlation can be interpreted as a delocalization error due to fractional electron [36], and is related to the self-interaction error in density functional theory (DFT). This bridges the languages used in quantum chemistry and condensed matter as well as points to a solution of the infamous “bandgap problem”, to be introduced in Sec. 2.3.2.3.

While the HF method enjoys much success in the study of atoms, its complete neglect of (Coulomb) electron correlation is woefully inadequate for many solids. The HF total energy is dominated by inner shell contributions, which overshadows the valence contributions important for correlated excitations in solids. The HF energy eigenvalues show vanishing density of states at the Fermi level in metals and unphysically large band gaps in insulators [37].

2.3.2 Kohn-Sham Density Functional Theory

Using a mapping from electron density to total energy, density functional theory (DFT) is a method that can, in principle, exactly include electron correlation effects. Though the exact density functional is unknown, even approximate functionals can lead to useful results. DFT uses the three-dimensional total electron density n(r) as the basic variable rather than the 3N-dimensional many-body wave function Ψ(r1,,rN). This is a dramatic simplification that likely lead to its dominance in modern electronic structure theory of solids and material science.

2.3.2.1 The Hohenberg-Kohn theorems

While having roots in Thomas-Fermi theory [38], DFT was put on firm theoretical foundation by P. Hohenberg and W. Kohn (HK) in 1964 [39], where they calculate the total electronic energy E from an external potential v(r) and a functional of the ground-state electron density

E drn(r)v(r) + F[n(r)]. (2.60)

Two theorems are often attributed to this work:

Definition 2.1. V-representable density A density n(r) is V-representable if it is the ground-state density of some Hamiltonian H in an external potential v(r).

Theorem 2.1. Assuming non-degenerate ground state, any V-representable ground-state density n(r) uniquely determines its external potential v(r).

Proof. by contradiction: Suppose there are two distinct external potentials v and vthat give rise to the same density n via different hamiltonians H, Hand wave functions Ψ and Ψ, respectively. By the variational principle ⟨Ψ′|H ′|Ψ ′⟩+⟨Ψ′|v′ − v|Ψ ′⟩ =

⟨Ψ′|H |Ψ′⟩ > ⟨Ψ|H |Ψ ⟩ =     ′
⟨Ψ |H  |Ψ ⟩+   ′
⟨Ψ |v − v|Ψ⟩ >   ′  ′ ′
⟨Ψ |H |Ψ ⟩+    ′
⟨Ψ|v − v|Ψ ⟩. Since Ψ and Ψgive the same density, the local term ⟨v′ − v⟩ cancels to give ⟨Ψ ′|H ′|Ψ′⟩ > ⟨Ψ′|H′|Ψ ′⟩, which is a contradiction. □

Theorem 2.2. Assuming number-conserving density variations that retain V-representability, the energy functional has a unique minimum at the ground-state density.

Proof. Consider an external potential v, its hamiltonian H, and its unique ground state Ψ and density n. After a number-conserving variation, the new wave function Ψcan be used with the original hamiltonian and ⟨Ψ′|H|Ψ′⟩ > ⟨Ψ |H |Ψ ⟩ by the variational principle. □

These initial proofs by HK have two important assumptions: 1. the ground-state is non-degenerate and 2. the electron density n(r) is V-representable. The latter is especially sever because reasonable densities were shown not to be V-representable [40, 41]. Fortunately, M. Levy proved that both assumptions can be weakened [42]. The HK theorems hold for N-representable densities regardless of ground-state degeneracy.

Definition 2.2. N representable density A density n(r) is N-representable if it can be obtained from some many-body wave function of N particles.

While less publicized, HK also pointed out that the exact density functional less the direct/Column contribution can be calculated from a local energy-density functional gr[n]

G[n] F[n] 1
-
2 drdrn(r)n(r′)
-----′--
|r− r | = rgr[n], (2.61)
gr[n] 1
2rrn1(r,r)|r=r + 1
2 dr       ′       ′
C2(r-− r-∕2;r+-r-∕2)
        |r ′|, (2.62)

which is constructed from one- and two-body reduced density matrices n1 and C2. They even went as far as to relate the leading-order behavior of the density functional to polarizability

G[n] = G[n0] + drdrK(r r)ñ(r)ñ(r) + h.o., (2.63)

where ñ is a small number-conserving density variation and the kernel K is related to the polarizability in reciprocal space

K(r r) = 1
--
Ω qK(q)eiq(rr), (2.64)
K(q) = 2π2-
q[       ]
 --1- − 1
 α(q) = 2π2-
q-1--
𝜖(q), (2.65)

where α(q) and 𝜖(q) are the polarizability and dielectric constant, respectively. The HK theorems and limits provide some checks for practical parametrization of the exact density functional. Unfortunately, they provide no guidance on how one might start to construct numerical approximations to the exact functional.

2.3.2.2 The Kohn-Sham equations

One year after HK, W. Kohn and L. J. Sham (KS) [43] worked out practical guidelines for constructing approximations to the exact density functional. KS first partitioned the total energy to highlight the least-understood “exchange-correlation” term Exc[n].

E = drn(r)v(r) + 1
-
2 drdrn(r)n(r′)
------′-
 |r− r | + T[n] + Exc[n], (2.66)

where T is a kinetic energy functional. KS then approximated Exc by the corresponding contribution in the homogeneous electron gas, the local density approximation (LDA)

Exc[n] = drn(r)𝜖xc(n(r)). (2.67)

Next, by minimizing eq. (2.66) with respect to number-conserving density variation, they obtained the stationary condition for the ground-state density

δn[δT[n]        ∫     n(r′)    d(n 𝜖 (n))]
 -----+ v(r)+   dr′-----′-+ ---xc----
  δn               |r − r|      dn = 0. (2.68)

To solve eq. (2.68), KS assumed that the ground-state density n came from an auxiliary system of non-interacting electrons, i.e., a Slater determinant. This KS ansatz turns eq. (2.68) into a system of one-particle equations of non-interacting particles in some effective potential vKSeff determined by the density n(r). Most practical implementations of DFT use the KS auxiliary-system formulation.

Practical success of the DFT LDA method was not realized until 1981, when J. P. Perdew and A. Zunger (PZ) [37] parametrized exact quantum Monte Carlo data of the homogeneous electron gas, obtained by D. M. Ceperley and B. J. Alder [44] the year prior. PZ’s eq. (13-17) define the KS-DFT method and the LDA we use today. The electron density for spin σ is a sum of occupied spin orbitals squared

nσ = σ a=1Nσ f|ψ(r)|2, (2.69)

where the occupation numbers f [0,1], and the kinetic energy is the sum of contributions from all occupied spin orbitals of the non-interacting system

Ts[n] = σ a=1Nσ f⟨       1      ⟩
  ψaσ|− 2∇2 |ψaσ. (2.70)

Minimizing eq. (2.66) with the constraint that the spin orbitals remain orthonormal, PZ obtains

--δ--
δψaσ[      ∫          ∫   ∫                          ]
 T [n ]+   drv(r )+   dr  dr′-n(r′)-+ -δ-E  (n ,n )
  s                        |r− r′|  δnσ  xc ↑  ↓. (2.71)

2.3.2.3 The Band Gap Problem

The Kohn-Sham eigenvalues do not have the same physical meaning as the Hartree-Fock eigenvalues as given by Koopmans theorem. Thus, a band gap “problem” arises when one compares the HOMO-LUMO gap from KS-DFT to experimental measurements of the fundamental gap Eg. The eigenvalue of the Kohn-Sham HOMO can be identified with the ionization energy of an isolated molecule if the exchange-correlation potential vanish at infinity. This is a special case, because the ionization energy is dominated by the long-range asymptotic behavior of the 1RDM, which is determined by the long-range tail of the electronic density with no contribution from the short-sighted exchange correlation potential.

When extended to handle systems with fractional electron number, the derivative discontinuity of the exact density functional is Eg. It has contribution from both the non-interacting kinetic functional Ts[n] and the exact exchange-correlation functional Exc[n]. The KS HOMO-LUMO gap measures the derivative discontinuity in Ts[n]. However, as shown in Fig. 2.2, the Exc[n] under LDA is smooth, so its contribution to the gap is missing. As a result, the LDA gap is 50% that of experiment, showing that the derivative discontinuity in Exc[n] can be of similar magnitude as the gap and cannot be ignored.

PIC PIC

Figure 2.2:A. J. Cohen, P. Mori-Sánchez, and W. Yang [36] explains that the self-interaction error in H2+ binding curve (left panel A) is due to the presence of fractional electron (left panel B), which leads to a delocalization error when LDA is used.
2.3.2.4 Beyond Local Density Approximation

While extremely successful, the LDA has well-documented deficiencies. As discussed in Sec. 2.3.2.3, LDA underestimates the band gap due to a lack of derivative discontinuity. Further, it tends to overestimate the binding energy of molecular systems. This is attributed to the logarithmic divergence of the LDA correlation functional as density tends to infinity (see eq. (7.55) in Ref. [32]).

The logarithmic divergence of the LDA is largely corrected by the generalized gradient approximation (GGA). This modification is not as straight-forward as adding an extra term that depends on the gradient of the electron density n to the exchange-correlation function. An arbitrary choice of the gradient term can distort the xc hole and break the sum rule that controls its global strength

rn(r)[g(r,r) 1] = 1. (2.72)

J. Perdew and co-workers overcame this difficulty by introducing a real-space cutoff to the exchange hole. While the details are technical, the final form of the GGA exchange functional is elegant

ExGGA[n] = drn(r)𝜖x(n(r))Fx(s(r)), (2.73)

where 𝜖x is the LDA exchange function and the scaled gradient

s(r) = --∇n-(r)--
2kF(r)n(r). (2.74)

A simple form for the exchange enhancement factor Fx was given by J. Perdew, K. Burke, and E. Ernzerhof [45] in 1996. After another careful cutoff on the correlation functional, the immensely popular PBE functional was constructed. The PBE xc functional takes the same form as eq. (2.73), but with a different enhancement factor: Fxc instead of Fx. PBE softens molecular bonds relative to the LDA and predicts an order of magnitude more accurate dissociation energies of many molecules [45].

Unfortunately, both LDA and PBE suffer from a well-known failure of any local and semi-local density functional theory: the absence of Van der Waals interaction. This interaction is entirely due to the correlation of density fluctuations and is dominant for two neutral objects with non-overlapping electron densities. When fluctuation creates an instantaneous dipole moment in one electron distribution, it induces an anti-aligned dipole moment in the other. The van der Waals interaction is thus attractive and decays as dipole-dipole interaction strength, 1r6-, in the separation distance r between the centers of the two distributions. While KS-DFT relies on the average electron density of a non-interacting system, it is still possible to include the contribution of Van der Waals interaction in the density functional. Consider the Coulomb interaction as a perturbation to two widely separated atoms, one can show that the interaction energy is proportional to a convolution of the density-density response functions of the isolated atoms

⟨    ⟩
  ˆH
   12 = -ℏ-
2π dr1dr1 dr2dr2---e2---
|r1 − r2|---e2---
|r′1 − r′2|
× lim0χ 1(r1,r1)χ2(r2,r2) + h.o., (2.75)

as written in eq. (7.116) in Ref. [32]. One can then proceed to approximate the density-density response function using the static electron density, e.g. using the polarizability of the homogeneous electron gas.

Finally, exact-exchange functionals such as PBE0 and HSE [46] have been developed for atoms having localized valence electrons, e.g., with d or f angular momentum. These functionals are crucial in the study of magnetism, as the PBE functional overly favor delocalized electrons and often predict qualitatively wrong magnetic momentum and spin orderings of transition metals. However, these exact-exchange functionals were never a focus throughout this thesis, so I will not go into details here.

Chapter 3
Slater-Jastrow wave function

In the previous chapter, I introduced the FN-DMC method, which calculates ground-state properties of a many-body system starting from a trial wave function ΨT. The accuracy and efficiency of the method depend on the choice of ΨT. Understanding of the many-body wave function and its connection to physical properties of particular systems can help us make educated guesses at high-quality trial wave function and perform accurate simulations. In this chapter, I will describe the most well-understood many-body wave function for electronic structure, the Slater-Jastrow wave function, and discuss what behavior of electrons we can learn from it.

The many-body wave function is also interesting in its own right. One goal of studying the many-body wave function is to understand electron correlation [47]. As P.A.M. Dirac pointed out, knowing the Dirac/Schrödinger equation and the hamiltonian of the system does not constitute an understanding, because it “leads to equations much too complicated to be soluble” [48]. Even a simple hamiltonian that contains only pair interactions, e.g., the Coulomb hamiltonian, can create complex many-body correlations and phase diagrams. Thus, direct studies of experimental observables, the many-body wave function, and perhaps properties the exact density functional will be more informative.

3.1 Historic Overview

In condensed matter, the development of many-body wave function took off in the study of homogeneous quantum liquids, e.g., liquid helium and the homogeneous electron gas, a.k.a. jellium. Most studies made use of the variational principle eq. (2.32), which states that given a Hamiltonian , any normalized trial wave function ΨT will have an energy value no less than that of the true ground state ⟨         ⟩
 ΨT | ˆℋ|ΨT⟨       ⟩
 Ψ0| ˆℋ|Ψ0. This principle allowed the pioneers to make educated guesses, then check their quality using the expectation value of .

As early as 1934, E. Wigner [49] noted that by introducing a “hole” in the correlation function of opposite-spin elections, one can improve the Slater determinant and lower its energy value in the homogeneous electron gas. Although, this hint was not acted on until much later. In 1940, A. Bijl [50] found that the logarithm of the wave function of many-interacting particles is size-extensive. This logarithm can be expressed as a perturbation expansion involving one- and two-electron terms and is convergent in the thermodynamic limit. Unfortunately, this work went unnoticed for 30 years while others independently developed similar ideas. Expanding the logarithm of the wave function is a general idea that later became known as the Bijl-Dingle-Jastrow-Feenberg expansion [51] for historical reasons, which I will now describe.

In 1949, R. B. Dingle [52], while estimating the zero-point energy of hard spheres, came up with a product of exponential functions as a variational wave function by considering symmetries and limits. In 1955, R. Jastrow [53] generalized the Dingle pair-product wave functions to indistinguishable particles with Bose and Fermi statistics. To generalize to fermions, he multiplied by a Slater determinant to enforce antisymmetric permutation symmetry. Thus, the Slater-Jastrow wave function was born. While this thesis is mostly focused on the Slater-Jastrow wave function for electronic matter, the idea of separating particle statistics from correlation is general. The Slater-Jastrow wave function is the fermion variant of eq. (3.1)

Ψ = (|
|||{ I({ϕi})exp(− U)  distinguishabe
  P ({ϕi})exp(− U ) bosons
|||
|( D ({ϕi})exp(− U)  fermions, (3.1)

where I, P, D are identity, permanent, and determinant, respectively. {ϕi} is a set of single-particle wave functions and U consists of pair terms only

U = i<ju(rij). (3.2)

The minus sign in the definition of eq. (3.1) is intentional. At high temperature, |Ψ|2 for distinguishable particles becomes the Bolzmann distribution eq. (2.6). When only pair interaction is present, the pair contribution to U has the same sign as the potential (see Chapter 6.6 in Ref. [24])

limβ0u(rij) = 1
2βv(rij). (3.3)

Some important improvements to eq. (3.1) came from the study of bosons rather than fermions.

In 1956, R.P. Feynman and M. Cohen [54] found it crucial to include the effect of back flow to accurately describe rotons in liquid helium. A roton is the quantum analog of a microscopic vortex ring. When an atom moves through the ring, it triggers a returning flow far from the ring, a.k.a. back flow. R.P. Feynman first estimated the energy-momentum curve of liquid helium using a permanent of plane wave orbitals in 1954, but found the roton energies severely over-estimated [55]. It is only after the introduction of back flow into the trial wave function did the roton energy reduce to a reasonable value, in qualitative agreement with the phenomenological theory of Landau and with experiment [54]. In 1961, F.Y. Wu and E. Feenberg [56] related the Jastrow pair function, u(r) in eq. (3.2), to the pair correlation function of liquid helium

u(r) = g(r)   1
8π3ρ dkeikr  [S(k)− 1]2
1-+-ξ[S(k)−-1], (3.4)

where ξ 0.97 gave accurate results under the superposition approximation. Direct relation between experimentally measurable correlation functions and the wave function is an important avenue to glean understanding from the many-body wave function. Similar ideas arose in the study of fermions in the same year.

Also in 1961, T. Gaskell [57] derived a Jastrow pair function for homogeneous electron gas from perturbation calculation of its pair correlation function. He expressed the Jastrow pair function in collective coordinates and derived an analytical formula using the random phase approximation (RPA). By minimizing the total energy in the long wavelength limit, Gaskell found an accurate Jastrow pair function, having correct limits at both short and long wavelengths. This Gaskell RPA wave function proved particularly useful in the study of the homogeneous electron gas and will be discussed in Sec. 3.4, then extended in Sec. 3.5. Remarkably, this wave function is accurate in both 2D and 3D. Using the Gaskell RPA wave function as trial function, in 1980, D. M. Ceperley and B. J. Alder [44] found the ground state of the homogeneous electron gas in 3D using exact QMC simulations.

A common theme of these early successes is guessing and checking of correlation. One guesses that a many-body correlation is important, incorporates said correlation into the wave function, then checks if the energy value is lowered and/or correlation functions get closer to experiment. In contrast, recent improvements to the many-body wave function rely heavily on numerical optimization of general wave function forms with many parameters. Examples include orbital rotation in a single determinant, multi-determinant expansion, iterative back flow [58], and neural network. As we move towards these complicated wave functions, it will likely become increasingly difficult to extract physical understanding directly from the wave function. This chapter will serve as a summary of some physical insights we have been able to grasp from the Slater-Jastrow wave function. I hope some of these will remain useful for more complex wave function forms. After some definitions, I will first discuss the short-range asymptotic behavior of the two-body contribution, i.e., the cusp condition in Sec. 3.3. Second, I discuss the two-body long wavelength behavior by studying the Gaskell RPA Jastrow in Sec. 3.4 for one-component system, then extend it to multi-component system in Sec. 3.5. Third, in Sec. 3.6, I show observables that can be calculated from the Slater determinant in plane wave basis, namely the momentum distribution from one-particle reduced density matrix (1RDM) in Sec. 3.6.1 and the static structure factor from two-particle reduced density matrix (2RDM) in Sec. 3.6.2.

3.2 Definitions

Definition 3.1. The Fourier Transform of a 3D function in coordinate space is defined as

f(k) = d3reikrf(r). (3.5)

The above Fourier transform convention eq. (3.5) defines its inverse

f(r) = -d3k-
(2π)3eikrf(k). (3.6)

The consistency of eq. (3.5) and (3.6) can be checked using the Coulomb potential v(r) = 1
r and v(k) = 4π2
k.

In a finite cell of volume Ω, momentum states are discretized. Each state takes up    3
(2Ωπ)- in reciprocal space. Therefore the inverse Fourier transform (3.6) becomes

f(r) = 1
--
Ω keikrf k. (3.7)

Definition 3.2. The collective coordinates of N particles of species α is the Fourier transform of their instantaneous number density

ρα(k) = d3reikrρα(r) = d3reikr j=1Nα δ(r rjα) = j=1Nα eikrjα . (3.8)

The collective coordinates provide a fixed basis for many-body functions in reciprocal space. Consider N particles in a cell of volume Ω interacting via an isotropic pair potential v(r). The potential energy

V = i<jv(rij) = 1
2 ijv(rij)
=  1
2Ω- kvk ijeik(rirj)
=  1
---
2Ω kvk(ρkρ−k − N ). (3.9)

When generalized to multiple species, eq. (3.9) becomes

V = 1
2 α,β i=1Nα j=1Nβ,(j,β)(i,α)vαβ(|r iα r jβ|)
= -1-
2Ω kvkαβ(               )
 ρβ  ρα − δαβN α
  kαβ −k. (3.10)

For particles interacting via the Coulomb pair potential

vαβ(r) = Q αQβ
--r---, (3.11)

where Qα and Qβ are the charges of species α and β, respectively.

Definition 3.3. Jastrow Pair Function: The general form of a Jastrow wavefunction containing two-body terms is

Ψ = exp(− U ), (3.12)

where

U = 1
2 α,β i=1Nα j=1Nβ,(j,β)(i,α)uαβ(|r iα r jβ|)
= -1-
2Ω kαβukαβ(             )
 ρβkρα−k − δαβN α. (3.13)

uαβ(r) is the Jastrow pair function. In the high temperature limit, uαβ(r) = vαβ(r)
2kBT. ukαβ is the Fourier transform of uαβ(r) in the unit cell having volume Ω as defined by eq. (3.7).

Definition 3.4. A Slater determinant is a many-body wavefunction ansatz for the ground state of a collection of same-spin fermions. It is the anti-symmetrized version of a product wavefunction ansatz for distinguishable particles.

Ψ = √-1--
  N ! 𝒫(1)𝒫(∏N       )
    ϕ𝒫i(ri)
 i=1, (3.14)

where N is the number of fermions, r1,r2,,rN are their spatial coordinates. 𝒫 is a permutation of the particle indices 1,2,,N. ϕ12,N are a set of one-body wave functions (a.k.a. orbitals).

3.3 Cusp Conditions

Derivation guided by Ex. 6.6 in Ref. [24]

Consider two non-relativistic distinguishable particles having masses m1 and m2 interacting via a pair potential v(r). The Schrödinger equation in the center-of-mass coordinate is

[           ]
 − λ∇2 +v(r)ψ = Eψ, (3.15)

where λ =  2
2ℏμ, and μ = (m11 + m21)1. The ground-state wave function

ψ = exp(u(r)) (3.16)

should have a stationary local energy

EL ℋˆψ
---
 ψ = v(r) + λ2u(r) (u(r))2
= v(r) + λ(u′′ +       ′
(d-−-1)u--
   r) λu2 = const., (3.17)

where d is the number of spatial dimensions. We see that the laplacian term in the kinetic energy has a potentially divergent term at r = 0. This term can respond to the potential and keep EL stationary, even if v(r) has a divergence at r = 0, e.g., the Coulomb potential. Suppose the two particles have charges q1, q2, and v(r) = q1q2∕r, the condition for stationary EL is

limr01
r(q1q2 + λ(d 1)u) = 0 u(0) = --q1q2--
λ(d− 1) (3.18)

For electron-electron interaction in Hartree atomic units m1 = m2 = 1, so λ = 1 and u(0) = 1
2 in 3D. This is the cusp condition for unlike-spin electron pair. For same-spin pair, the two particles are indistinguishable and the laplacian for each particle contributes a copy of the divergent term, thus u(0) = 14. For an electron-ion pair in the clamped-ion approximation (m2 →∞)

u(0) =  2Z
d−-1-, (3.19)

where Z is the atomic number of the ion. Imposing the cusp conditions on a trial wave function greatly reduces the variance of the local energy and improves the efficiency of a QMC calculation. The electron-ion cusp eq. (3.19) is the most important one to maintain, because the wave function amplitude around an ion is high and many samples from the MC algorithm will have some electron close to an ion. In contrast, one rarely samples a configuration with two electrons close together due to strong electron-electron repulsion at density relevant for materials science, e.g., bulk silicon. However, at high density, electron-electron correlation is weak relative to kinetic energy, so the electron-electron cusp condition is important to maintain. Nevertheless, the effect of imposing the electron-electron cusp condition is typically less pronounced than that of the electron-ion one.

3.4 Gaskell RPA Jastrow

The RPA Jastrow potential electron gas in 3D given by T. Gaskell [29, 57, 59, 60] is

2ρukRPA = [(               )−1∕2]
  S0(k)−2 +2ρvk∕𝜖k1 S 0(k)1, (3.20)

where the νk is the Coulomb potential in reciprocal space and 𝜖k is the energy-momentum dispersion relation. For non-relativistic electrons in 3D, νk = 4π
k2 and 𝜖k = k2
2 using Hartree atomic units. S0(k) is the static structure factor of the free Fermi gas  [61]

S0(k) = {  (   )      (   )3}
  3  k-- − -1  -k-
  4  kF    16  kFΘ(2kF k) + Θ(k 2kF). (3.21)

Gaskell [57] used an integral identity to obtain an approximate relation between the Jastrow potential and the static structure factor

2ρuk S1(k) S 01(k). (3.22)

Therefore, the RPA structure factor can be read off of ukRPA in eq. (3.20) via eq. (3.22)

SRPA(k) = (              )
 S−02(k)+ 2ρvk∕𝜖k12, (3.23)

where kF is the Fermi k-vector. For unpolarized electrons kF = 3π2ρ = (   )
  94πr3s-13.

Equation (3.23) is exact in the long wavelength k 0 limit. Taylor expanding eq. (3.23) around k = 0

S(k) = k2
2ω--
  p k2k4
9Fω3--
  p + O(  )
 k6, (3.24)

where the plasmon frequency ωp = √4πρ- =   ----
∘ 3∕r3s.

While Gaskell originally derived eq. (3.20) using perturbation theory, one can derive the same form by minimizing the variance of the local energy in the long wavelength limit, as shown in the next Sec. 3.5.

3.5 Multi-Component RPA Jastrow

Based on notes from D. M. Ceperley dated Sep. 1980

Given Jastrow wavefunction Ψ = exp(U), where

U = i<ju(rij) = 1
-
2 α,β i=1Nα j=1Nβ,(j,β)(i,α)uαβ(|r iα r jβ|), (3.25)

and non-relativistic Coulomb hamiltonian

H = T + V = α j=1Nα λαjα2 + 1
2 α,β i=1Nα j=1Nβ,(j,β)(i,α)vαβ(|r iα r jβ|), (3.26)

where α,β label particle species, i,j label particle positions. λα = -ℏ2--
2mα, vαβ(r) = QαQβ-
 r. In terms of pair potentials and collective coordinates (see Fourier convention eq. (3.5) and its corollaries eq. (3.6-3.8))

U =  1
---
2Ω kαβukαβ(             )
 ρβkρα−k − δαβN α, (3.27)
V = -1-
2Ω kαβvkαβ(             )
 ρβkρα−k − δαβN α. (3.28)

The goal of this section is to obtain good Jastrow pair potentials ukαβ. The strategy is to minimize the variance the local energy EL Ψ1HΨ = T + V , where

T = γ λγ l=1Nγ (                  )
 ∇ γlU ⋅∇ γlU − ∇γl2U. (3.29)

In the following, I will detail the few steps needed to obtain the RPA Jastrow potentials. First, we express the local energy in terms of the collective coordinates eq. (3.8). Second, we find the equations that make the local energy invariant to changes in the collective coordinates. Third and finally, we solve these equations for one and two component systems. Assume uαβ = uβα and uk = uk.

3.5.1 Local Energy of Jastrow Wavefunction

The gradient, laplacian, and gradient squared of eq. (3.27) are

lγU =  1
---
2Ω kα(ik)ukγα(    γ            γ)
 eik⋅rl ρα−k − ραke−ik⋅rl (3.30)
lγ lγU = 1--
2Ω kαk2u kγα(    γ            γ      )
 eik⋅rl ρα−k + ραke−k⋅rl − 2δαγ (3.31)
lγU lγU = -12-
4Ω kqαβk qukγαu qγβ ×(
ei(k+q)rρ kαρ qβ ei(kq)rρ kαρ qβ ei(qk)rρ kαρ qβ + ei(k+q)rρ kαρ qβ
). (3.32)

Summing over l turns eikrlγ into ρkγ in eq. (3.31) and (3.32). Thus

l=1Nγ lγ lγU =  1
---
2Ω kαk2u kγα(                     )
ργkρα− k + ραkργ−k − 2Nγδαγ (3.33)
l=1Nγ lγU lγU =  1
---2
4Ω kqαβk qukγαu qγβ ×(
ρk+qγρ kαρ qβ ρ kqγρ kαρ qβ ρqkγρ kαρ qβ + ρ (k+q)γρ kαρ qβ
). (3.34)

Equation (3.34) contains terms that couple three wave vectors, i.e. O(ρ3). In the spirit of RPA, we will drop all such mode coupling terms. Note ρ0γ = Nγ, and use uk = uk

l=1Nγ lγU lγU = N-γ-
2Ω2 kαβk2u kγαu kγβ( α  β    α β )
 ρ−kρk + ρkρ−k. (3.35)

Finally, sum over γ with λγ to obtain terms in the kinetic energy. To simplify later assembly of the local energy, rename dummy variables α,β,γ such that every O(ρ2) term contains ρkαρkβ (use uαβ = uβα)

γ λγ l=1Nγ lγU lγU = 1
Ω- kαβγλγNγ
Ω--k2u kγαu kγβρ kαρ kβ, (3.36)
γ λγ l=1Nγ lγ lγU = 1-
Ω(                              )
( ∑  λ-α +-λβ αβ α β           )
        2   uk ρk ρ−k − N αλαδα,β
  kαβ. (3.37)

Finally, the local energy can be assembled

EL = k[      λ +λ    αβ                      ]
 vαkβ-− -α2-βk2uk--− ∑  (Nγ∕Ω)λγk2uαkγuβkγ-
 2Ω        Ω         γ        Ωρkαρ kβ + Nα2(λα-+-12)
    Ω. (3.38)

3.5.2 Equations that define the RPA Jastrow Pair Potentials

Variance of EL can be minimized by setting the ρkαρkβ term to zero. Define 𝜖kα λαk2

vαβ
-k--
 2 𝜖α+ 𝜖β
-k---k
   2ukαβ γN
--γ
 Ω𝜖kγu kαγu kβγ = 0. (3.39)

Equation (3.39) can be solved for each k independently. We no longer need the collective coordinates or the label k. It is now safe to recycle the symbol ργ Nγ-
Ω to mean the number density of species γ. Simplify eq. (3.39) to

vαβ
 2 1
2(𝜖α + 𝜖β)uαβ γργ𝜖γuαγuβγ = 0. (3.40)

3.5.3 Solving for the RPA Jastrow Pair Function

One Component

For a one-component system, eq. (3.40) becomes a quadratic equation of one variable u11

v11
 2 𝜖1u11 ρ1𝜖1u112 = 0. (3.41)

The solution is

2ρ1u11 = 1 + ∘------------
 1 + 2ρ1v11∕𝜖1, (3.42)

which agrees with Gaskell’s solution eq. (3.20), except S0(k) is replaced by 1. Notice, if one uses a different Fourier convention, replacing volume Ω with number of particles N in eq. (3.27)

U =  1
2N- kαβũkαβ( β α         )
 ρkρ−k − δαβN α, (3.43)

then the density ρ drops from the expression for ũ, e.g., eq. (8) in Ref. [29] and eq. (3) in Ref. [6]

2ũ = 1 + (1 + 2vk∕ek). (3.44)

Two Components

For two-component system, eq. (3.40) becomes a set of 3 coupled quadratic equations

(
||  v11− 𝜖 u  − ρ𝜖 u2 − ρ 𝜖 u2 = 0
||{   2    1 11    11 11   2 2 12
   v122− 12(𝜖1 + 𝜖2)u12 − ρ1𝜖1u11u12 − ρ2𝜖2u12u22 = 0
||||  v22             2        2
(  -2-− 𝜖2u22 − ρ1𝜖1u12 − ρ2𝜖2u22 = 0. (3.45)

Suppose species 2 has infinite mass λ2 0, thus no dispersion 𝜖2 = 0. Then we should ignore the last equation (α = β = 2), which determines u22 (when u12 = 0). The remaining equations allow us to solve for the Jastrow pair potentials u11 and u12

(
|{  v11-             2
   2 − 𝜖1u11 − ρ1𝜖1u11 = 0
|(  v12−  𝜖1u12 − ρ1𝜖1u11u12 = 0
   2    2. (3.46)

The first equation provides the same Jastrow potential as in the one-component case eq. (3.42). The second equation can be used to solve for u12

(1 + 2ρu11)𝜖1u12 = v12
u12 = ∘---v12∕𝜖1----
  1 +2ρ1v11∕𝜖1. (3.47)

For completeness, the exact solutions are (by Mathematica)

2ρ1u11 = 1 + 𝜖(1 +a  )± 𝜖 √A-
-1--√--11-√--2---
     𝜖1𝜖2 B, (3.48)
u12 = --±v12√---
√ 𝜖1𝜖2- B, (3.49)
2ρ2u22 = 1 + 𝜖(1 +a  )± 𝜖 √A-
-2--√--22-√--1---
     𝜖1𝜖2 B, (3.50)

where

A = (1 + a11)(1 + a22) a122, (3.51)
B = 𝜖2-
𝜖1(1 + a22) + 𝜖1-
𝜖2(1 + a11) ± 2√ --
  A, (3.52)

with

(
||        2ρ1v11-
||{  a11 = √𝜖1---
   a12 = 2√ρ1ρ2v11-
||||          𝜖1𝜖2
(  a22 = 2ρ2𝜖v222- . (3.53)

3.6 Slater Determinant in Plane Wave Basis

Based on notes from D. M. Ceperley dated Aug. 1 2018

When the orbitals are expressed in plane wave basis

ϕi(r) = kcikeikr. (3.54)

We require the orbitals to be orthonormal

Ωdrϕi(r)ϕ j(r) = δij Ω kcikc jk = δij. (3.55)

We can verify that the determinant written in eq (3.14) is normalized

⟨⟩ dr1drNΨ Ψ
=  1
---
N ! 𝒫,𝒫′(1)𝒫(1)𝒫′(∏N ∫                )
      drlϕ∗𝒫l(rl)ϕ𝒫′l(rl)
 l=1
=  1
---
N ! 𝒫,𝒫′(1)𝒫(1)𝒫′(∏N      )
    δ𝒫l,𝒫′l
 l=1
= -1-
N ! 𝒫 = 1. (3.56)

The key step in eq (3.56) is to separate and distribute the many-body integrals into the product.

Many properties of the slater determinant can be evaluated analytically. Here we focus on reciprocal-space properties accessible by scattering experiments: the momentum distribution n(k) and the static structure factor S(k).

3.6.1 Momentum Distribution

The momentum distribution is the Fourier transform of the 1RDM (eq. (5.9) in Ref. [24]). The 1RDM can be calculated from the many-body wavefunction

ρ(x,x) = N dr2rNΨ(x,r 2,)Ψ(x,r2,). (3.57)

Given a Slater determinant wavefunction eq (3.14), all the dr integrals can be done analytically

ρ(x,x) = N dr2drN(                                             )
     ∑                         ∏N
( 1--   (− 1)𝒫(− 1)𝒫′ϕ∗𝒫1(x)ϕ𝒫′1(x ′) ϕ𝒫∗l (rl)ϕ𝒫l′(rl))
  N! 𝒫,𝒫′                       l=2
= N--
N ! 𝒫,𝒫′(1)𝒫(1)𝒫′(             N∏      )
 ϕ∗𝒫 (x)ϕ𝒫′(x′)   δ𝒫 ,𝒫′
   1     1    l=2  l l
= N
N-! 𝒫ϕ𝒫1(x)ϕ 𝒫1(x)
= 𝒫1=1Nϕ 𝒫1(x)ϕ 𝒫1(x). (3.58)

Notice that the diagonal (x = x) of the 1RDM is particle density. Given PW orbitals eq (3.54)

n(k) = --13---
(2π )N drdr′′eikr′′ρ(r,r r′′)
= --1----
(2π )3N i,g,gcigc ig drdr′′eigreig′⋅(rr′′)
= --1----
(2π )3N i,g,gcigc igΩδg,gΩδg,k
=  Ω
(2π-)3-Ω
N- i=1N|c i,k|2. (3.59)

Given the current definitions, dkn(k) = 1 for an infinite system. In practice, one bins the Fourier coefficient squared of all occupied orbitals at allowed momenta of the supercell.

3.6.2 Static Structure Factor

The static structure factor is the density-density correlation function in reciprocal space

Sq 1-
N⟨ρ ρ  ⟩
  q −q1-
N⟨  N        N       ⟩
 (∑  eiq⋅ri)(∑  e−iq⋅rj)
  i=1      j=1
= 1-
N ij⟨        ⟩
 eiq⋅(ri−rj) = 1 + 1-
N ij⟨         ⟩
  eiq⋅(ri− rj)
= 1 + (N 1)⟨        ⟩
 eiq⋅(r1−r2). (3.60)

Focus on the many-body integral

⟨         ⟩
 eiq⋅(r1−r2) = 1
---
N! 𝒫,𝒫′(1)𝒫(1)𝒫′ dr1drNeiq(r1r2) l=1Nϕ 𝒫l(r l)ϕ𝒫l(rl). (3.61)

Similar to eq (3.56) and eq (3.59), 𝒫l = 𝒫′l,l1,2. Define 𝒫1 = i, 𝒫2 = j, then 𝒫1,2= i,j contributes a positive term, and 𝒫1,2= j,i contributes a negative term. Thus eq (3.61) simplifies

⟨         ⟩
  eiq⋅(r1−r2) =     1
---------
N (N − 1) i,j dr1dr2eiq(r1r2)×
[                                             ]
 ϕ∗i(r1)ϕi(r1)ϕ∗j(r2)ϕj(r2)− ϕ∗i(r1)ϕj(r1)ϕ∗j(r2)ϕi(r2)
=     1
N-(N-−-1) ij[∫
   dr1eiq⋅r1ϕ∗i(r1)ϕi(r1) dr1eiqr2 ϕj(r 2)ϕj(r2)
 ∫                    ∫                    ]
−   dr1eiq⋅r1ϕ∗i(r1)ϕj(r1)  dr1e−iq⋅r2ϕ ∗j(r2)ϕi(r2)
= ----1----
N (N − 1) ij[mii(q)mjj(− q)− mij(q)mji(− q)], (3.62)

where we have defined the matrix of integrals

mij(q) drϕi(r)ϕ j(r)eiqr. (3.63)

Notice mij(q) = mji(q), thus

Sq = 1 + 1-
N ij[mii(q)m ∗(q)− mij(q)m∗ (q)]
        jj           ij
= 1 + 1
N-⌊                        ⌋
  ∑            ∑
⌈ |   mii(q)|2 −    |mij(q)|2⌉
    i          i,j. (3.64)

3.6.3 Example: Free Fermions

The ground-state wavefunction of non-interacting fermions is a determinant of plane waves. The first N plane-wave orbitals with the lowest momenta are filled. In case of degeneracy, the wavefunction will have a non-zero net momentum.

Simply stated, the free fermion wavefunction is a determinant eq (3.14) whose orbitals each have a single Fourier component

cik = -1-
√Ω-δk,ki. (3.65)

We see from eq (3.59) that the momentum distribution n(k) of the free fermions is a step function, which is constant within the Fermi surface and zero outside. As for the static structure factor S(k), first note that the matrix of integrals mij(q) eq (3.63) is sparse

mij(q) = cikic jkjΩδq,kikj = δq,kikj. (3.66)

Plug (3.66) into (3.64) and

Sq = (|
{  N                 q = 0
|(  1− 1-∑ |δ     |2  q ⁄= 0
      N i,j  q,ki− kj. (3.67)

Eq (3.66) has a simple geometric interpretation. Namely, mij(q) is non-zero only if q connects two occupied plane wave orbitals. In the thermodynamic limit, the geometric interpretation allows S(k) to be calculated from a simple integral for the overlapping volume of two spheres

S(k0) = 1 (4πk3 )
 ----F
   312 q∕2kF dkπ(kF2 k2)
= (   (  )     (  )
|{ 3  -q  − 1- -q  3 q < 2kF
  4  kF    16 kF
|( 1                 q ≥ 2kF . (3.68)

3.7 Beyond Slater-Jastrow

One way to systematically obtain more accurate wave functions than Slater-Jastrow is to expand the many-body wave function in a basis of determinants. This configuration interaction (CI) method is routinely used in quantum chemistry to study molecules. The CI method is especially effective when the correlation in the ground state is mostly static in nature, e.g., in benzene molecule where the ground-state wave function is approximately the equal superposition of six non-interacting determinants. This method is less successful in bulk calculations due to difficulties in performing finite-size extrapolation. Specifically, a fixed-length determinant expansion is more accurate for a small system than a large one. Thus, properties such as the total energy are not size-extensive and cannot be extrapolated to the thermodynamic limit using known asymptotic formula.

Another method for systematically improving the many-body wave function is to use the generalized Feynman-Kac formula (eq. (6.12) in Ref. [24]), which is intimately related to the DMC method (eq. (24.23) in Ref. [24])

Ψ0 =   Ψ
----T---
⟨ΨT |Ψ0 ⟩⟨   (   ∫ ∞                  )⟩
 exp  −    dt[EL(Rt;ΨT )− E0]
         0ΨT2, (3.69)

where ⟨⟩ΨT2 denotes average over drift-diffusion random walks guided by ΨT eq. (2.28). EL(RtT) is the local energy of the trial wave function at walker position Rt during the random walk. If ΨT is a good approximation for the ground state Ψ0, then imaginary time propagation can be cut off at a small equilibrium “time” τ and eq. (3.69) has a compelling physical interpretation eq. (6.13) in Ref. [24]

lnΨ0(R0) lnΨT(R0) τ[⟨EL⟩ΨT2(R0) E0]. (3.70)

Equation (3.70) says the correction to the exponent of ΨT at R0 is proportional to the average of its local energy over a small region around R0 having size O(√ ----
  dλτ) in dN-dimensional space. This observation is quite useful for practitioners of the DMC method, because it implies longer projection time is needed for:

  1. larger system, because the space over which EL needs to be averaged increase with N.
  2. heavier particles, because they have smaller quantumness λ, thus smaller diffusion constant.
  3. long-range properties, because the effect of projection is local and expands slowly with projection time.

Finally, the Feynman-Kac formula was shown to be capable of suggesting compact wave function that accurately capture electron correlation starting from a simple initial guess [10].

Chapter 4
Finite-size Effects

In the previous two chapters, I introduced the ground-state QMC method and the trial wave function needed to make it accurate for a well-defined electron-ion Hamiltonian. However, practical QMC simulations cannot handle the hamiltonian governing O(1023) particles found in typical condensed matter experiments. While we use periodic boundary condition to eliminate surfaces from our simulations, there is significant remaining difference between the properties of this model system and those from experiment.

In this chapter, I will describe common types of finite size effects encountered in electronic structure simulations and a few correction schemes. I will begin discussion from the most general correction that works for quantum and classical systems at zero and finite temperature: the static structure factor correction to the potential energy. From there, the method is extended to correct kinetic energy using the momentum distribution, and in the quantum case, using the Jastrow pair function.

4.1 Correction to the potential energy

4.1.1 General Theory

Given a simulation using periodic boundary conditions, the accessible momenta are quantized. As the simulation cell is enlarged, the grid of accessible momenta becomes finer until it grants access to the full continuum of momentum space in the thermodynamic limit. By considering the difference between this infinite system and the finite simulation cell, we can understand finite-size effects and attempt to correct them. Consider an orthorhombic box with side length Lx along the x direction. The momentum along the x direction must take discrete values kx = 2π-
Lxnx, where nx . Similarly discretization exist along the other directions. In a simulation where particles interact via the Coulomb pair potential v(r) = 1
r, the total potential energy of the system

V V background + 1
-
2 i=1N L j=1Nv(|x i xj L|), (4.1)

where L loop over the supercell lattice. The sums loop over all pairs of particles in the infinite system. This equation can be equivalently written in reciprocal space as

V N V∕N = vM +  1
2Ω- k0vkS(k), (4.2)

where vk = 2π(d − 1)
--kd−1-- is the Fourier transform of the Coulomb pair potential in d spatial dimensions. vM is the Madelung constant, which combines the electrostatic energy of an infinite periodic array of charges on L with a neutralizing background. Equation (4.2) looks different than its original proposal in Ref. [62] due to a difference in Fourier transform convention. Here, I follow the definitions eq. (6) and (7) in Ref. [60], which are reiterated in Sec. 3.1.

Suppose S(k) is converged, i.e., does not change with system size, then the only difference between the infinite-system and the finite-size potential energies is the replacement of the sum in eq. (4.2) by an integral

ΔV N V V N = [               ]
 ∫  ddk    1 ∑
    (2π)d-− Ω-
              kvk
-2S(k). (4.3)

Equation (4.3) is not practical, because limk→∞S(k) = 1 and both the sum and the integral diverge. Fortunately, large k corresponds to short-range interaction, so its contribution to finite-size error vanish rapidly with system size. Thus, we can truncate the large-k part of eq. (4.3) with little effect on its value. This can be achieved either using an explicit suppression factor e𝜖k2 as done in eq. (24) of Ref. [63] by Drummond et al., or splitting out the long-range part of the Coulomb potential as done in eq. (30) of Ref. [60] by Holzmann et al..

While the Madelung term in eq. (4.2) is specific to charged systems, the idea of finite-size error as a quadrature error eq. (4.3) applies to any pair potential v(r). This error can be accurately corrected given converged pair correlation functions in real g(r) and reciprocal S(k) spaces.

4.1.2 Homogeneous electron gas

In the case of the electron gas, more progress can be made due to known long-wavelength behavior of the wave function. The dominant contribution to eq. (4.3) comes from the volume element around k = 0, because vk diverges there.

ΔV N (2π)d-
 Ωddk
(2π)d-vk
-2S(k). (4.4)

Bohm and Pines [64] discovered that the many-body wave function of the electron gas can be factored into short-range and long-range contributions, where the long-range part describes weakly coupled collective modes (plasmons) [60, 65]

Ψ = Ψs.r. exp(                                             )
    1 ∑            1  ∑
(− 2Ω-   ukρkρ−k + Ω2-   w(k,q)ρk+qρ−kρ−q + ...)
       k              k,q. (4.5)

In the random phase approximation (RPA), we ignore the mode coupling terms, such as w(k,q), and find the Gaskell RPA static structure factor eq. (3.23). Using the leading-order approximation of S(k) eq. (3.24) in the finite-size correction formula eq. (4.4), we obtain the main correction to the potential energy

ΔV Nl.o. = ωp-
 4. (4.6)

Similarly, in 2D [66]

S0(k) = ({   ⌊      (          ∘ ---(----)2) ⌋)}
  2-⌈arcsin (--k-+ --k-  1−   -k-- ) ⌉
( π         2kF   2kF        2kF     )Θ(2kF k) + Θ(k 2kF), (4.7)
S(k) = (k∕kF-)3∕2-
23∕4rs1∕2 + O(k2), (4.8)
ΔV Nl.o. =    C2D
π5∕4(2rs)3∕2-   1
N-5∕4- + O(N32), (4.9)

where kF √ -
  2∕rs, C2D = 3.9852 and 3.9590 for square and hexagonal cells, respectively [63]. Equation (4.9) differs from eq. (60) in Ref. [63], because Ref. [63] erroneously used the dimensionless form of the RPA structure factor rather than its Hartree atomic unit form eq. (4.8).

4.1.3 Inhomogeneous system

In a real crystal, valence electrons interact with a periodic arrangement of localized ionic cores rather than a homogeneous neutralizing background of positive charge. In such inhomogeneous systems, it is instructive to separate the static and fluctuating contributions to the static structure factor

S(k) 1-
N⟨ρ ρ  ⟩
  k −k = -1
N⟨(ρ  − ⟨ρ ⟩)(ρ − ⟨ρ ⟩)⟩
  k    k   k    k + -1
Nρk⟩⟨ρk. (4.10)

PIC

Figure 4.1:Fluctuating and static contributions from valence electrons in the conventional cell of bulk silicon.

As shown in Fig. 4.1, the static part due to charge density ρk is non-zero only at reciprocal lattice of the primitive cell of the underlying crystal structure, whereas the fluctuating part varies smoothly from 0 to 1. Its value within the missing k = 0 region (red area in Fig. 4.1) can be used to correct the potential energy to leading order following eq. (4.4). Changes in the charge density with system size is a higher-order effect and is small compared to the total potential energy [67]. However, as shown in Ref. [68], this static contribution becomes important when energy differences are taken, such as in the calculation of the fundamental gap.

Even in an insulator, eq. (3.23) still works well at k kF, but the short-range contribution S0(k) no longer comes from a simple determinant of plane waves. Dielectric screening suppresses long-range fluctuation and changes the leading-order behavior of S(k) from eq. (3.24) to eq. (7) in Ref. [68]

S(k)  k2
2ωp-(1 𝜖k1)12, (4.11)

so long as only one excitation dominates the dynamic structure factor.

4.2 Correction to the kinetic energy

4.2.1 From momentum distribution

The kinetic energy can be calculated as the second moment of the momentum distribution

TN T∕N = 1
-
ρ  ddk
----d
(2π)e(k)nN(k), (4.12)

where the dispersion of non-relativistic particles

e(k) = -ℏ2
2mk2, (4.13)

and nN(k) is the momentum distribution of these particles at the given system size N. It is crucial to distinguish the finite-size nN(k) from its thermodynamic limit n(k), because it is a nonlocal quantity that converges slowly with system size [59]. Similar to eq. (4.3), the finite-size correction to the kinetic energy

ΔTN = [1 ∫  ddk          ]
 ρ   (2π)de(k)n ∞(k)[              ]
 1 ∑
 N-   e(k)nN(k)
    k. (4.14)

The key difficulty in using eq. (4.14) is finding a reasonable approximation to n(k). Some progress can be made by analyzing the Monte Carlo estimator for the Fourier transform of the momentum distribution, i.e. the off-diagonal one-particle density matrix [69]

n(r) = ⟨ Ψ(R : r → r)⟩
  ------i-----
     Ψ (R )|Ψ|2, (4.15)

where R denotes the positions of all N particles, and the notation “: ri r” means that particle i is moved from ri to r. As noted by W.R. Magro and D.M. Ceperley [70], direct application of eq. (4.15) with periodic boundary condition can result in superfluous contributions, because all periodic images of particle i are moved. The images will contribute to the ratio in eq. (4.15) if Ψ has long-range components, e.g., eq. (4.5). Chiesa et al. [65] and Holzmann et al. [59] later used this observation to design a finite-size correction to the momentum distribution and the kinetic energy. To leading-order [59]

n(k) nN(k) (2π)Ωd ddq
(2π)d[                  2    ]
 uq(1− S(q))− ρu(q) S(q)(nN(k + q)− nN(k)), (4.16)

where uq is the Jastrow pair function in the wave function eq. (4.5). As shown in Fig. 8 of Ref. [71], the leading-order correction to n(k) works well for lithium. Further, Table III in the Supplemental materials of Ref. [71] shows that the corrected nN(k) can be used to accurately correct the finite-size error in the kinetic energy using eq. (4.14). To achieve this good result for a metal with a sharp Fermi surface such as lithium, it is crucial to densely sample momentum space while preserving the Fermi surface using grand-canonical twist averaging.

4.2.2 From wave function

Instead of using the relation between kinetic energy and the momentum distribution eq. (4.12), one can directly analyze the QMC estimator for kinetic energy to find its finite-size correction. The VMC estimator for the kinetic energy of a Slater-Jastrow wave function Ψ = DeU is (from eq. (14) of Ref. [60])

TNV MC = 1-
N⟨   N     [             ]⟩
  −∑   ℏ2- ∇2iD--− (∇  U)2
   i=1 2m   D       iTND + T NU. (4.17)

The dominant finite-size correction in the determinant term is due to one-particle “shell filling” effects, whereas the dominant correction in the Jastrow term is due to long-range two-particle correlation. I will now discuss these two effects in turn.

4.2.2.1 Single-particle “shell filling” effect

If the orbitals in the determinant are from some effective one-particle theory such as HF or KS-DFT, then they are solutions of some effective one-particle potential veff

[  ℏ2∇2      ]
 − -2m--+ veffϕn(r) = 𝜖nϕn(r). (4.18)

Thus, one can work out the determinant contribution to the kinetic energy

TND ≡− 2
ℏ--
2m i=1N  2
∇-iD-
 D = [∑N      N∑       ]
     𝜖n −    veff(ri)
 n=1     i=1, (4.19)

by eq. (19) in Ref. [60]. As the number of electrons is increased, the kinetic contribution eq. (4.19) increases by the energies of the new orbitals being occupied. Consider the ideal Fermi gas in a 2D square box. As the number of same-spin particles increase from N = 2 to 5, the first shell of states in reciprocal space become filled. They all have the same single-particle energy, so the total kinetic energy increases linearly with N. However, when one adds a 6th particle, it increases the total energy by twice the amount as one in the first shell. As shown in Fig. 4.2, this shell filling effect causes oscillation in the kinetic energy as a function of the number of particles, making size extrapolation difficult. It can be drastically reduced by adopting canonical twist averaged boundary conditions (TABC), where the number of particles is the same across all twists. Further, grand-canonical twist averaged boundary condition (GC-TABC), where the number of particles change according to the exact Fermi surface, can exactly remove this single-particle finite-size effect. Finally, there is another “pocket” method which reduces the number of twists needed. Within a pocket in reciprocal space, the orbitals in the determinant smoothly acquire a phase as the twist is varied. Once these pockets are mapped out, one can perform one calculation per pocket and weigh it by the volume of the pocket to exactly remove the one-particle finite-size error [60].

PIC

Figure 4.2:Relative error of total energy vs. number of particles with PBC (up triangles) and TABC (squares) in 2D and 3D [72]
4.2.2.2 Two-body size effect

Using the RPA Jastrow pair potential from eq. (4.5), its contribution to eq. (4.17) becomes

TNU -1
N⟨ ∑N   2      ⟩
     -ℏ-(∇iU )2
  i=12m1-
Ω k0 2
ℏ--
2mρukukS(k), (4.20)

by eq. (33) in Ref. [60]. Recall the general relation between S(k) and u(k) eq. (3.22)

2ρuk S1(k) S 01(k). (4.21)

Since S1(k) diverges faster than S01(k) as k 0, eventually 2ρuk S1(k). Thus, to leading-order eq. (4.20) becomes

TNU -1-
2Ω k0e(k)uk, (4.22)

which should be compared with eq. (4.2) for potential energy. Thus, the procedure to correct the two-body finite-size error in the kinetic energy is analogous to the potential correction scheme eq. (4.3)

ΔTN TTN = [∫   d         ]
   -d-k- − 1-∑
   (2π)d   Ω  ke(k)
 2u(k), (4.23)

where u(k) is an interpolation of the converged Jastrow pair potential, which can be approximated using the converged static structure factor in the long wavelength limit.

4.2.2.3 Finite-temperature correction

Based on notes from D. M. Ceperley
At finite temperature and under the RPA, the long-range part of the action can be optimized to minimize the variance of the local energy, resulting in

2ρuk = Q(k)1 tanh(         )
  β--e(k)--
  2 Q(k,β)S0(k)1, (4.24)

where β is inverse temperature, e(k) = λk2 is the dispersion of the particle, ρ = N∕Ω is density, and S0(k) is the non-interacting structure factors.

Q(k) (            2ρv )
  S0(k,β )− 2 +---k
             e(k)12. (4.25)

Q(k) reduces to Gaskell RPA S(k) as β →∞. Since limx→∞tanh(x) = 1, eq. (4.24) becomes eq. (3.22) in this limit. Assuming the relation between S(k) and uk eq. (3.22) holds, the finite-temperature RPA structure factor is

S(k) = Q(k)tanh( β  e(k) )
  2-Q(k,β)-. (4.26)

Equations (4.24) and (4.26) can be used to derived leading-order finite-size corrections to the kinetic and potential energies using eq. (4.23) and (4.3), respectively.

Chapter 5
Dynamic-ion DMC Study of Solid Hydrogen at Megabar Pressures

5.1 Introduction

The properties and phase transitions of hydrogen under megabar pressures are important in diverse fields of study. For astronomy, models of the interior of gas giants such as Jupiter and Saturn depend critically on the nature of the molecular liquid to atomic liquid transition (LLT), namely whether it is first-order or continuous [12, 73]. For condensed matter, metallic hydrogen holds promise for a room temperature conventional (BCS) superconductor [74, 75]. For computational physics, hydrogen remains an important benchmark for both electronic structure [76] and ion dynamics methods. With no need for a pseudopotential, simulations of hydrogen avoid a significant source of bias. However, the low mass of the nuclei necessitates quantum treatment of the lattice degree of freedom, often beyond the harmonic approximation.

PIC

Figure 5.1:Partial phase diagram of hydrogen on log-log scale [77].

Established experimental results on high-pressure hydrogen are limited. At room temperature and below, diamond anvil cell (DAC) is the dominant apparatus to achieve such high pressures. Small size of the cell and fragility of the sample limit experimental probes to low-power optics such as infrared and Raman spectroscopy[78]. Hydrogen is a weak scatterer of X-Rays [79], thus excluding this excellent tool for structural determination in most experiments. Only recently has X-ray analysis been performed up to 254 GPa [80, 81]. At high temperatures, shock wave compression is the main method to achieve megabar pressures. Due to the transient nature of theses experiments, acquiring and analyzing shock-wave data is challenging. Most notably, one cannot directly measure temperature, which may cause misinterpretation of raw data [5, 82, 83]. Given the experimental difficulties, predictive simulations are highly desirable as they can inform and verify experiments [4].

Simulation of high-pressure hydrogen is also challenging. Without experimental structural information from X-ray, many theoretical calculations have been performed on structures found in density functional theory (DFT) random structure searches [84]. Constrained by computational cost, these searches are limited to classical protons, causing the methods to miss, for example, saddle-point structures that can be stabilized by nuclear quantum effect [85]. Predictive simulations of hydrogen require accurate methods both in the description of the electronic ground-state Born-Oppenheimer (BO) potential energy surface (PES) and in the inclusion of nuclear quantum effect beyond the quasi-harmonic approximation. The popular Perdew-Burke-Ernzerhof (PBE) density functional in DFT erroneously predicts some molecular structures to be metallic [3]. However, its use in conjunction with Classical molecular dynamics (MD) results in reasonable transition pressure for the LLT at certain temperatures due to error cancellation [86]. This and other fortuitous cancellations of error has led many to believe that the PBE functional provides a good description of solid hydrogen and caused much confusion in the community. PBE predicts a conductive molecular structure above 200 GPa, a molecular-to-atomic transition around 300 GPa [2], and low-temperature superconducting liquid. All these predictions contradict experimental evidence. Systematic benchmark of the PES from various DFT functionals against QMC found the vdW-DF1 functional to be the most accurate for molecular hydrogen at megabar pressures [67]. However, this functional has yet to gain widespread adoption due to its higher computational cost and lower popularity compared to PBE.

In this chapter, I will focus on the solid phases of hydrogen. Sec. 5.1.1 summarizes experimental observations, Sec. 5.1.2 summarizes relevant computational studies, Sec. 5.2 details the approach taken in this study, and Sec. 5.3 presents the computational results.

5.1.1 Experiments

As element number one with the simplest atomic structure, hydrogen has surprisingly complex phases at megabar pressures. Further complicating matters, the phase diagram depends on the isotopes, e.g., hydrogen H, deuterium D, and spin isomers of molecular hydrogen. The proton spins anti-align to form a singlet in para-hydrogen (p-H2), whereas they align to form a triplet in ortho-hydrogen (o-H2). To clarify the narrative, I will first introduce the well-established phases in pure samples, then discuss changes due to isotopic and ortho-para conversion.

For pure p-H2 at low temperature (510 K), three solid phases are well-established. The low-pressure phase (LP) below 100 GPa is a molecular crystal having spherically symmetric H2 molecules on hcp lattice sites. Above 110 GPa, hydrogen enters a broken-symmetry phase (BSP), where anisotropic intermolecular interactions favor the J = 2 v = 1 vibrational state of the H2 molecules rather than the spherically symmetric J = 0 v = 1 state [87]. Above 160 GPa, after crossing a first-order transition, one finds an orientationally-ordered phase known as the A phase (H-A) [88].

As shown in Fig. 5.2, the transition from LP to BSP phase is sensitive to isotope and nuclear spin. o-D2, HD, and p-H2 enters the BSP at 28 GPa [89], 70 GPa [90], and 110 GPA [87], respectively. In contrast, the transition to the A phase is fairly robust across isotope and spin isomer variants. HD, o-D2, and p-H2 all enter the A phase between 150 and 160 GPa [88, 9092]. The phase lines for o-D2 and HD are shown in Fig. 5.2. The p-H2 LP-BSP phase line near 100 GPa is not shown. The size and shape of the BSP is the only difference between the phase diagrams of p-H2, o-D2, and HD. These orientation transitions are not relevant to o-H2 and p-D2, which have a hcp to fcc transition at ambient pressure.

PIC

Figure 5.2:Phase diagram of o-D2 and HD below 200 K and 200 GPa [90].

Transitions to these orientationally ordered phases are detected by changes in Raman and IR spectra. As shown in Fig. 5.3, during the LP to BSP transition, one can observe clear broadening and weakening of the low-frequency roton bands around 350 cm1 and an associated small (15 cm1) discontinuity in the position of the vibron peak, which is about 4150 cm1 near the transition pressure 110 GPa [87]. Upon further increase of pressure past 160 GPa, a much larger discontinuity of the Raman vibron (100 cm1) signals the onset of the A phase [88, 91]. A direct transition from H-A back to the LP phase can be achieved by raising temperature. Across this transition, the intensities of the libron bands decrease discontinuously [87].

PIC (a) rotons

PIC (b) vibron
Figure 5.3:Roton and vibron changes in p-H2 across the BSP transition [87].

The optical signatures for the LP to BSP transition in o-D2 are qualitatively similar to those in p-H2. The vibron decreases discontinuously by 3 cm1 rather than 15 cm1, while the roton bands broaden and weaken near the transition pressure of 28 GPa rather than 110 GPa [89]. Further confirmation of these two phase transitions were later obtained from IR absorption spectra [92]. Three absorption peaks appear around 3150 cm1 upon entering the BSP phase and are replaced by a single broad peak at the same frequency range when the A phase is reached. The same signatures were used to identify the BSP and A phases of HD at 70 and 160 GPa, respectively [90]. In the A phase, the rather broad and pressure-independent roton band weakens, disappears, and is replaced by a few sharp and strongly pressure-dependent peaks in the frequency range 100700 cm1 [93]. These new modes are considered to be lattice libration modes due to their pressure dependence.

Phases with mixed ortho-para concentrations of H2 are labeled I, II, and III [16, 94], which correspond to the LP, BSP, and H-A phases of pure p-H2, respectively. As shown in Fig. 5.4, at 300 K and above 220 GPa, we enter yet another solid phase IV, characterized by a splitting of the vibron peak [95]. Both theory and experiment suggest that phase IV consists of alternating layers having rather different in-plane structures, possibly with two types of molecules. Below 100 K and above 350 GPa, molecular hydrogen becomes semi-metallic, possibly due to the closure of an indirect band gap [96]. Then, above 425 GPa, all IR radiation is absorbed indicating a closure of the direct band gap [17]. Finally, at sufficiently high pressures, the hydrogen molecules will dissociate to form an atomic solid, reportedly at 495 GPa [97], although consensus has yet to be reached.

PIC

Figure 5.4:Tentative phase diagram of solid hydrogen below 400 K [94].

While the phase boundaries of solid hydrogen are reasonably well-established below 400 K and 400 GPa by diamond-anvil cell (DAC) experiments, characterizations of the solid structures are limited. Due to the small scattering cross section and small sample size in DAC experiments, only a handful of X-ray [80, 81, 98102] and only one neutron [102] scattering experiments have been published over the past 40 years. Most of our understanding of solid hydrogen is built upon IR and Raman spectra, which provide partial information on the microscopic details of the solid structures. This lack of definitive structural information poses significant difficulty for both theoretical and experimental understanding of solid hydrogen. Experimentally, this has lead to the misidentification of a triple point as a critical point [87, 92], subtlety in the detection of a new phase [103, 104], among many debates over interpretation of optical data.

5.1.2 Calculations

Early computational studies of solid hydrogen rely on assumed crystal structures from known high-pressure phases of other materials or simple symmetry and energetic arguments. Even before the observation of the oriented A phase of solid hydrogen [91], S. Raynor [105] used Hartree-Fock and perturbation theory to show that the molecular hexagonal closed packed structure with H2 molecules aligned along the c-axis (mhcp-c) is more energetically favorable than previous considered cubic structures. While a promising candidate for phase III [106], the mhcp-c structure has an early band overlap, rendering it metallic below 150 GPa, resists compression along the c-axis, and has no IR-active vibron [107, 108], all in contradiction with experimental evidence. Thus, E. Kaxiras et al. [107] explored different orientations of H2 molecules in the 2-atom hcp unit cell and found a more energetically favorable insulating structure with molecules oriented 60 from the c-axis. This static-lattice LDA study was later validated by a dynamic-lattice QMC calculation [7], and the structure named mhcp-o. In addition to the hcp structures, H. Nagara and T. Nakamura [109] proposed various rutile structures by minimizing the static-lattice electric quadrupole-quadrupole (EQQ) interactions, while B. Edwards, N. W. Ashcroft, and T. Lenosky [110] proposed an orthorhombic layered structure of Cmca symmetry, which turned out to be metallic at pressures relevant to phase III [111]. This Cmca crystal structure also appeared spontaneously in path integral simulation [112]. These theoretical calculations drove much debate about the fate of phase III at pressures over 300 GPa. Does it become a metallic molecular solid or does it dissociate into an atomic solid without the band gap closing?

In 2007, the advent of random structure searching algorithms produced new candidate crystal structures that have lower enthalpy than previous proposals [84]. The insulating layered structure having C2/c symmetry became the main candidates for phase III. Three diffusion Monte Carlo studies followed to characterize the candidate structures: Azadi et al. [1], McMinis et al. [2], and Drummond et al. [3]. Azadi et al. used PBE-optimized geometries and included anharmonic phonon zero-point energy, leading to a molecular dissociation at 374 GPa, from Cmca-12 to I41/amd. In contrast, McMinis et al. used vdW-DF-optimized geometries and harmonic phonon zero-point energy to predict a dissociation pressure of 447(3) GPa. In hind sight, the prediction by McMinis et al. is in better agreement with subsequent experiments.

On the low pressure side, a new hexagonal candidate structure for phase III was proposed by Monserrat et al. [85] in 2016, then calculated to be more stable than C2/c below 210 GPa [113]. Band gap of the C2/c structure shows closure around 460 GPa, when extrapolated using IR measurements up to 420 GPa [17]. This gap closure pressure agrees with the most recent DMC calculation [18], which is at variance with the previous prediction by Azadi et al. [113], presumably due to different treatments of finite-size effects. Finally, a recent coupled cluster calculation of the molecular candidate structures show good agreement with DMC results [114] at the static lattice level, although lattice zero-point energy has yet to be included.

In this chapter, we examine the most promising candidate structures of solid hydrogen using dynamic-lattice DMC. This method treats the electrons and ions on the same footing while harnessing the accuracy of DMC. Lattice vibrations are included beyond the harmonic approximation. Further, nonadiabatic effects can be captured. The goal is to provide the most accurate properties of the solid hydrogen phases.

5.2 Methods

5.2.1 Candidate Structure Optimization

We consider three candidates C2/c-24 [84], Cmca-4 [111], Cmca-12 [84] for the molecular phase and one candidate I41/amd [74] for the atomic phase at T = 0K over the pressure range 350 GPa to 700 GPa. The static-lattice structures in the molecular phase are optimized using the vdW-DF functional at constant pressure. As shown in Table 5.1, all three molecular structures optimize to similar density at each pressure.

Table 5.1:vdW-DF pressure-density (expressed in rs) relation of relaxed molecular candidate structures.




















vdW-DF P(GPa) 360 400 440 480 520 560 650 700 780










Cmca-4 1.303 1.283 1.265 1.250 1.235 1.222 1.196 1.183 1.164
Cmca-12 1.306 1.286 1.268 1.252 1.237 1.224
C2/c-24 1.307 1.287 1.269 1.253 1.239 1.225 1.198 1.185




















In contrast, the atomic structure is optimized using DMC at constant volume. While both constant-volume and constant-pressure optimizations are valid ways to find the minimum energy structure, the final density and pressure differ in general. The DMC optimizations for the atomic structure have been carried out at evenly spaced rs values: 1.31, 1.29, 1.27, , 1.17. At a given pressure, the density difference between the atomic and molecular structure is appreciably larger than that among molecular-phase candidate structures. Thus, when calculating energy and enthalpy differences, interpolation is needed further from data for the atomic structure than for the molecular structures.

All three molecular structures are monoclinic having a = bc, α = β = 90 + η, and γ = 120 + δ. The slight distortions differ for each structure: η = 0, δ ≈−0.5 for Cmca-4, η = 0, δ +3.5 for Cmca-12, and η 0.1, δ ≈−0.1 for C2/c-24. The evolution of the lattice parameters as a function of pressure are shown in Fig. 5.5(a). Both a and c decrease with increasing pressure. However, the c∕a ratio remains roughly constant at 1.062 ± 0.003 and 1.771 ± 0.003 for Cmca-12 and C2c-24, respectively. In contrast, the c∕a ratio of the Cmca-4 structure decreases from 1.562 at 350 GPa to 1.530 at 560 GPa linearly with pressure. Besides having a slightly different unit cell, the Cmca-4 structure has only one type of H2 molecule, whereas Cmca-12 and C2/c each have two. The bond length of the H2 molecules in the optimized geometry is shown as a function of pressure in Fig. 5.5(b).

The vdW-DF optimized H2 bond lengths of all three molecular structures are shown in Fig. 5.5(b). The bond length in Cmca-4 is comparable to its isolated value of 1.4 Bohr, whereas in C2/c-24 it is 3 to 4% compressed. One type of the H2 molecules in Cmca-12 has pressure-sensitive bond length, increasing from 1.38 Bohr at 360 GPa to 1.4 Bohr at 560 GPa, while the other type has 3% compressed bond length irrespective of pressure.

pict (a) a and c lattice parameters

pict (b) H2 bond length
Figure 5.5:vdW-DF optimized molecular candidate structures at target pressures. (a) lattice parameters (b) molecular bond length. C2/c and Cmca-12 each have two types of H2 molecules, whereas Cmca-4 has only one.

The atomic candidate structure I41/amd has only one free parameter, the c∕a ratio, at each density. Therefore, we can afford to optimize its geometry using DMC within the clamped-ion approximation. As shown in Fig. 5.6, the DMC-optimized c∕a ratio is 4 to 8% below the PBE-optimized ones. It increases continuously as density increases, whereas the PBE-optimized c∕a ratio exhibits discontinuities around rs = 1.21 and 1.14. We fit c∕a as a linear function of rs to smooth out noise from the optimization process.

pict

Figure 5.6:DMC optimized atomic structure c∕a ratio as a function of density. The blue points are DFT(PBE) optimized c∕a ratios.

While the vdW-DF optimized structures in Ref. [2] are not published, we can infer from the enthalpy-pressure relations that the same structures as in Ref. [2] have been reproduced in this study. Figure 5.7 shows the enthalpy of each candidate structure relative to C2/c-24 at the vdW-DF static-lattice minimum. The results agree well with those from McMinis et al. [2] where available. In comparison to predictions by Drummond et al. [3], our C2/c structure is slightly more stable and Cmca-4 slightly less stable. They used BLYP rather than vdW-DF functional, so differences are expected. In fact, it is encouraging to see two different functionals give similar results (within a few meV/p).

pict

Figure 5.7:DFT(vdW-DF) static-lattice enthalpy of optimized structures relative to C2/c-24. Thin solid lines are enthalpies of the Cmca-4, Cmca-12 and I41/amd using our optimized structures. Dashed lines are DFT(BLYP) enthalpies of Cmca-4 and Cmca-12 from Drummond et al. [3]. Dash-dot lines are DFT(vdW-DF) enthalpies from McMinis et al. [2].

5.2.2 Supercell Construction

To reliably obtain QMC energies in the thermodynamic limit, we need to tile the optimized primitive cells to sufficiently large supercells so that pair correlation functions are converged. The remaining finite-size error can be removed using methods discussed in Chap. 4. The supercells also need to be small enough for dynamic-ion QMC to be practical. In the end, we use 72-atom simulation cells for all QMC calculations.

Each simulation cell is tiled from the optimized unit cell using a non-diagonal supercell matrix [115], which is optimized to maximize the distance between minimum images under periodic boundary conditions. A supercell matrix in 3D is a 3 × 3 matrix of integers that map primitive lattice vectors a,b,c to supercell lattice vectors as,bs,cs

(    )
  a
||   s||
||  bs||
(    )
   cs = (               )
   S   S    S
||   11   12   13 ||
||  S21  S22  S23 ||
(               )
   S31  S32  S33(    )
   a
||    ||
||  b ||
(    )
   c. (5.1)

Once a supercell is chosen, the crystal structure can be created using the following cropping method: first, tile the atoms from the primitive cell a large number of times along each lattice vector, then crop out only the atoms that fall inside the supercell. The total number of atoms in the supercell should be det(S) times that in the primitive cell.

PIC
(a) square a = 2 Å

PIC
(b) rhombus a = 2.149 Å
Figure 5.8:Cubic vs. rhombus supercells. The black cell is the supercell. The gray cells are periodic images. The blue line points between nearest-neighbor images, while the red line between second-nearest neighbors. The yellow circle is the inscribed circle in the Wigner-Seitz cell (not shown) of the supercell.

Non-diagonal supercell matrices can be used to maximize the minimum image radius, radius of the real-space Wigner-Seitz cell RWS. As shown in Fig. 5.8, a rhombus supercell provides a larger RWS than a square having the same area. This is because the periodic images form a closed-packed lattice given a rhombus supercell. If the primitive cell is square, then all diagonal supercell matrices result in square supercells. In contract, one can construct a rhombus-like supercell using a non-diagonal matrix. A more useful application of non-diagonal supercell matrix is for accessing a particular momentum to address a certain excitation [115], but that is beyond the scope of this study.

The chosen supercell matrices and their resulting image radii are shown in Table 5.2 and Fig. 5.9, respectively. The inscribing radius of each supercell Rsc are also shown in Fig. 5.9 to give a sense of how far each supercell is from being orthorhombic. An orthorhombic cell has RWS = Rsc.

Table 5.2:Optimized 72-atom non-diagonal supercell matrices.








Cmca-4 Cmca-12 C2/c-24 I41/amd




(      )
|− 1 21|
||      ||
|(  2− 11|)
   3 30 (       )
|   21− 1|
||       ||
|(   21 1|)
  − 11 0 (    )
| 210|
||    ||
|( 120|)
  001 (       )
|  2− 21|
||       ||
|(  2  30|)
  − 2 11








pict

Figure 5.9:Supercell radius as a function of density. rs is the Wigner-Seitz radius, which is determined by the average electron density 4π3rs = ρ, where ρ = NeΩ, with Ω the supercell volume. RWS is the radius of the real-space Wigner-Seitz cell of the supercell. 2RWS is the minimum distance between periodic images.

5.2.3 Wavefunction Optimization

5.2.3.1 Electronic Components

For the electronic wave function, we use the standard Slater-Backflow-Jastrow (SBJ) form

Ψ(R;RI) = det({ϕi(rj + Δrj;RI )})eU(R,RI), (5.2)
U = 1
-
2 α,β i=1Nα j=1Nβ,(j,β)(i,α)u αβ(ri,rj), (5.3)

where α,β denote unique particle species, including up-electron, down-electron, and proton, although the proton-proton contribution is always set to zero in static-lattice calculations. rj + Δrj is the quasi-particle coordinate of particle j as determined by the back flow function η as follows

Δrj = 1
-
2 α,β i=1Nα j=1Nβ,(j,β)(i,α)η αβ(riα,r jβ)(r i rj). (5.4)

Isotropic Jastrow u(r) and back flow η(r) functions are optimized in VMC to lower the variational energy. The single-particle orbitals in the determinant are taken as the Kohn-Sham orbitals with the lowest eigenvalues. These orbitals are generated using the vdW-DF functional for the molecular candidates and PBE for the atomic structure. We remove the approximate cusp of each orbital by dividing out the RPA e-p Jastrow following Refs. [7, 116, 117]. This allows the exact e-p Jastrow to be re-introduced in the e-p Jastrow and is always maintained as the protons move. Before back flow transformation, the static structure factor of the Slater determinant is quite similar to that of the unpolarized homogeneous electron gas, as shown in Fig. 5.10(a). However, as shown in Fig. 5.10(a), the momentum distribution of the metallic atomic structure is different from the rest, because the molecular structures are insulating.

pict
(a) structure factor

pict
(b) momentum distribution
Figure 5.10:Static structure factor S0(k) and momentum distribution n0(k) calculated using the Kohn-Sham determinant wave function at rs = 1.25. The black lines show the same functions for the free Fermi (FF) gas as reference.

The Jastrow and back flow functions are variationally optimized using the electronic hamiltonian at vdW-DF optimized clamped-ion geometries. The optimized functions at 480 GPa (rs 1.25) are shown in Fig. 5.11. These functions remain quantitatively similar across all densities explored. The electronic components of the Jastrow and back flow functions are nearly identical for all three molecular candidates. The largest contribution is the u-d term, which introduces correlation between opposite-spin electrons to keep them apart. However, the electron-ion components show some variation among candidate structures, with Cmca-4 being the standout among molecular candidates. Its optimized e-p Jastrow and back flow have much in common with those of the atomic phase. This is likely an artifact of either the clamped-ion approximation or under-converged optimization. Once these functions are reoptimized with dynamic ions, all Jastrow and back flow components involving the electrons become essentially identical across all candidate structures, as shown in Fig. 5.12.

pict
(a) Jastrow pair functions

pict
(b) Back flow pair functions
Figure 5.11:Jastrow and back flow pair functions optimized at static-lattice minimum at 480 GPa. Color denotes the crystal structure. The solid line is the pair function for electron-proton, the dashed line for same-spin electrons, and the dotted lines for opposite-spin electrons.
5.2.3.2 Ionic Components

For an electron-ion simulation, the SBJ wavefunction has an additional Hartree product for distinguishable ions as well as an ion-ion Jastrow

Ψei(R,RI) = det(                )
 {ϕi(rj + Δrj; R0I)}eU(R,RI) I=1Np χ(rI,rI0), (5.5)

where the proton orbital is an isotropic gaussian centered around its optimized geometry

χ(ri,rI0) = (    )
  2Cp-
   π34 exp(           )
 − Cp |rI − r0I|. (5.6)

The expressions for the Jastrow and back flow terms remain unchanged, i.e. eq. (5.3) and (5.4), although the proton coordinates are now dynamic variables rather than parameters. We take the electronic orbitals ϕi from the clamped-ion DFT calculation, performed at the vdW-DF optimized geometry, for practical purposes. This is a severe approximation as the ideal orbitals for a system with small nonadiabatic effects should come from a Born-Oppenheimer calculation, i.e. rerun DFT at every ion position update. Fortunately, the back flow transformation effectively changes the electronic orbitals as the ions move, because the quasi-particle position of an electron is influenced by the protons around it.

We re-optimize the Jastrow and back flow functions in dynamic-lattice VMC, using the optimized functions from static-lattice calculations as starting points. The electron-ion Jastrow of the Cmca-4 Jastrow and back flow functions fall into good agreement with those from other candidate structures. As shown in Fig. 5.12, after re-optimization, the only noticeable difference among the three structures lie in the proton-proton Jastrow.

pict
(a) Jastrow pair functions

pict
(b) Back flow pair functions
Figure 5.12:Jastrow and back flow pair functions optimized using dynamic-lattice VMC. The dash-dotted line is the pair function for proton-proton correlation. All other notations are identical to those in Fig. 5.11.

The width of the gaussians in the proton orbital Cp are optimized using maximum overlap [7, 117, 118]. In this method, we measure the mean-squared deviation (MSD) of each proton from its ideal site, as determined by vdW-DF, and adjust the gaussian exponent in the proton wave function Cp until the VMC and DMC estimates of the MSD agree. This method works by maximizing the overlap between the trial and ground-state wave function along the dimension controlled by Cp. Since proton MSD is a rather direct measure of the shape of the wave function and DMC pushes the VMC MSD towards its ground-state value, the optimal Cp is obtained when DMC and VMC MSDs agree. Consider an isolated proton having a normalized gaussian wave function

χ(x,y,z) = (    )
  2Cp-
   π34 exp(− C (x2 + y2 + z2))
    p. (5.7)

The mean squared deviation of this wave function is

⟨  ⟩
 x2 = ⟨  ⟩
 y2 = ⟨  ⟩
 z2 =  1
----
4Cp. (5.8)

We expect some deviation from eq. (5.8) due to electron-proton and proton-proton correlations. When Cp is large, the confining effect of the gaussian wave function dominates over correlation due to Jastrow and back-flow components, making each proton approximately independent of all other particles. However, as Cp decreases, the effects of correlation become more important. As shown in Fig. 5.13(a), at large values of Cp (localized protons) VMC MSDs are accurately approximated by eq. (5.8), whereas the DMC MSDs and VMC ones at small Cp differ from this independent-particle approximation. Note that the DMC runs here are short and likely under-converged. Fortunately, they are sufficient for determining the point of maximum overlap between the VMC and DMC wave functions. As such, the computational cost of maximum overlap optimization is more similar to VMC optimization than to a converged DMC run. We further note that optimal Cp is the same along x, y, and z directions, so an isotropic Gaussian wave function is flexible enough to optimally sample the proton wave function. Finally, the exact value of Cp will not bias the DMC result so long as the projection time is sufficiently converged. This is because the protons are treated as distinguishable particles for which DMC is exact at infinite projection time.

pict
(a) Lattice deviations vs. prediction from isolated gaussian wave function with exponent Cp

pict
(b) optimized exponents as a function of density
Figure 5.13:Maximum overlap optimization of the proton wave function. (a) Optimization of the C2/c-24 proton wave function at rs = 1.269. (b) Optimized proton wave function exponent Cp of all candidate structures. The markers show the Cp value at the intercept of VMC and DMC. The lines are linear fits, but shifted down slightly to error on the side of oversampling. These fits are used to determine the Cp value in the dynamic-lattice calculations.

In practice, the choice of Cp is important. Since the proton-electron mass ratio is more than 1836, the RMSD of a proton during the diffusion process is roughly 143 that of an electron in the same amount of imaginary time. Thus, it takes more simulation time to fix errors in the proton wave function than those in the electron wave function. Fortunately, the proton wave function is highly localized and alleviates the problem to some extend. As shown in Fig. 5.13(a), the standard deviation of the proton orbital of C2/c-24 is < 0.5 bohr at rs = 1.27. Nevertheless, it pays to choose an optimal value for Cp. If Cp is too large, then the proton wave function will be too localized. This results in undersampling of the ground-state wave function and is a dangerous scenario. During the DMC run, the kinetic energy of the protons will likely have mostly small fluctuation in this case. However, occasional spikes, which are important for an unbiased mean, will occur infrequently. If one misses or discards these spikes, then the results will be biased. In contrast, oversampling will increase the variance of the wave function without biasing the mean. The resultant calculation will require more time to reach a certain accuracy target, so a balance needs to be sought. As shown by the lines in Fig. 5.13(b), we choose isotropic gaussian exponent that are close to optimal but error slightly on the side of oversampling (smaller Cp than optimal).

5.3 Results and Discussion

5.3.1 Static-Lattice Energy, Pressure, and Local Energy Variance

The dominant variation of the total energy as pressure increases is due to the isotropic compression of the electron liquid and the crystal lattice. This variation is on the order of 1 eV/p, which is roughly 50 times the energy difference among the candidate structures. To better observe these small differences, we plot energy and pressure relative to a reference equation of state (EOS). As a definition of zero energy at each density, this reference EOS need not be overly precise. Nevertheless, we use a fit to high-quality DMC data for the C2/c-24 molecular crystal structure obtained by Drummond et al. [3], because they are freely accessible online [119] and are well approximated by a quadratic polynomial in density. The data file E_SJDMC_dt0_KZKcorr_v_V.dat in directory DMC/N096/C2c-24/ contains the size-corrected DMC energy-density relation needed to fit an EOS. As shown in Fig. 5.14, we fit the high-density region of the reference data to obtain:

E0(ρ) = 1.605872ρ2 + 0.699501ρ 0.610902, (5.9)

where ρ = N∕Ω is the electron density in bohr3, and E0 is in hartree. We choose to fit the size-corrected 96-proton data obtained using a Slater-Jastrow wave function, because the results are easier to reproduce than its large-cell and back-flow counterparts. The LDA-based finite-size correction scheme designed by H. Kwee, S. Zhang, and H. Krakauer (KZK) [120] is not general in principle. However, it was shown to work well for solid hydrogen comparing 768-proton and 96-proton results [3]. Equation (5.9) reproduces the QMC data to well within 1 meV/p while interpolating between 212 and 413 GPa, but deviates by 1.5 meV/p when extrapolated by 50 GPa to 163 GPa. Thus, we can expect the reference EOS to be accurate over the pressure range 150 (rs = 1.48) to 460 GPa (rs = 1.26), but fail when extrapolated much beyond these limits.

pict

Figure 5.14:Fitting of reference equation of state. The solid symbols are DMC data from Ref. [3], while the line is a quadratic fit to the circled points. The bottom panels shows the error in the fit.

The static-lattice DMC energy is plotted relative to reference in Fig. 5.15(a). At the same density, we obtain lower total energy than Drummond et al. [3] partly due to our use of vdW-DF, rather than PBE, to optimize the crystal structure; and partly due to in the inclusion of back flow transformations. At the same geometry, DMC energy is unchanged to within 3 meV/p whether PBE or vdW-DF is used to generate the orbitals in the determinant wave function. Our C2/c curve is relatively flat at densities lower than rs = 1.25 and starts to bend downwards as density increases. Thus, our C2/c pressure-density relations will be roughly identical to that of the reference EOS in the low-density region. This is encouraging, because it implies that the fixed-node error on the pressure is small. Even with improved geometry and wave function, we get back very similar pressure as PBE geometry and Slater-Jastrow wave function at the same density.

pict (a) energy-density relation

pict (b) enthalpy-pressure relation
Figure 5.15:Static-lattice DMC energy and enthalpy relative to Drummond et al. [3] reference eq. (5.9). Relative energies are shown in meV per proton (meV/p). Each solid line is obtained using a fitted energy-density EOS, which is obtained by fitting the finite-size corrected (FSC) total energy as a quadratic function of ρ. The markers are finite-size corrected simulation data without performing a fit. P0 is the pressure calculated from the reference EOS P0 = dE0∕dΩ.

QMC enthalpy can be calculated from virial pressure and plotted relative to the reference EOS as shown in Fig. 5.15(b). We observe a transition from C2/c-24 to Cmca-4 around 640 GPa and no transition to the atomic structure up to 800 GPa. This disagrees with the previous QMC study [2], showing a 450 GPa transition from C2/c-24 to Cmca-4, which then transitions to the atomic structure around 700 GPa. The reason for this disagreement is unknown.

The QMC virial pressure, calculated from kinetic energy T, potential energy V , and volume Ω

P = 2T-− V-
  3Ω (5.10)

is shown as a function of density in Fig. 5.23(a). We first note that the QMC pressure is consistently lower than vdW-DF pressure at the same density. Plotting the same data on a relative scale in Fig. 5.23(c), we see that the QMC virial pressure is 4.2% and 3.2% lower than vdW-DF pressure at 340 GPa (rs 1.21) and 460 GPa (rs 1.25), respectively. This agrees well with the findings of R. C. Clay III et al., i.e. Fig. 4 in Ref. [67]. In this plot, we also see that the mixed estimator error is 1 to 5 GPa, so the linearly-extrapolated static-lattice DMC pressure should be accurate to 1 GPa.

pict (a) static lattice

pict (a) dynamic lattice
Figure 5.16:QMC variance vs. pressure. The lines are guides to the eye.

Finally, the variance of the optimized static-lattice wave functions are shown in Fig. 5.16(a) as a function of pressure. The variance of the molecular structures are 7.5 8.5 eV2/p below 400 GPa and increases to 8.5 9.5 eV2/p at 550 GPa. The increase in variance may be due to the closing band gaps of the molecular structures around 400 GPa. If true, then it may be better to generate orbitals above 400 GPa using the PBE functional, because it prefers metallic rather than insulating states when compared to vdW-DF. The variance of the atomic structure is about 50% higher than that in the molecular phase, ranging from 11 eV2/p at rs = 1.31 to 15 eV2/p at high densities. For comparison, the variance of the best analytical wave function for bcc hydrogen is 13.2(1) eV2/p at rs = 1.31 (Table III in Ref. [10]) and 7.40(3) eV2/p at rs = 1.40 (Table 2 in Ref. [31]). Small mixed-estimator error and low energy variance are hallmarks of a well-optimized wave function. We demonstrate both in Fig. 5.23 and Fig. 5.16 for the static-lattice electronic wave function. The same cannot be said about the electron-ion wave function. As shown in Fig. 5.16, the local energy variance of all structures in the dynamic-lattice calculations are twice as high as their static-lattice counterparts. The atomic-phase variance is still roughly 3 eV2/p above the molecular structures. This large increase in variance is due, in large part, to the use of static-lattice orbitals eq. (5.5) at all proton configurations. They can be lowered by re-optimizing the electronic wave function upon every proton-position update, albeit at a significantly increased computational cost.

5.3.2 Effect of Dynamic Ions on Energy and Pressure

When clamped protons are allowed to move, the energy of the electron-ion system changes in two ways. First, proton kinetic energy contributes a sizable positive term to the total kinetic energy of the system. As shown in Fig. 5.17(d), the kinetic energy of the protons increases from 230 meV/p at 360 GPa (rs 1.31) to 265 meV/p at 600 GPa (rs 1.20), increasing by roughly 7.3 meV/p every 50 GPa. Second, the electronic wave function changes as the protons move. Thus, the electronic potential and kinetic energies change as well. As shown in Fig. 5.17(b) and (c), the increase of the total potential energy is roughly twice that of the electron kinetic energy, which is around 80 ± 20 meV/p for all candidate structures. Unlike the increase of proton kinetic energy, the electron kinetic and total potential increases are insensitive to the electron density. The increase of electron kinetic energy for C2/c-24 and i41/amd is 20 meV/p higher than that of the Cmca structures, while the increase in their potential energies is 40 meV/p lower. The sum of all three contributions make up the total energy change from static- to dynamic-lattice, i.e. the zero-point energy (ZPE). As shown in Fig. 5.17(a), the ZPE increases from 470 meV/p at 360 GPa to 510 meV/p at 600 GPa, i.e. 8.3 meV/p every 50 GPa. Most of the density dependence of the ZPE comes from the proton kinetic energy.

pict
(a) total energy change (zero-point energy)

pict
(b) potential energy change
pict
(c) electronic kinetic energy change
pict
(d) proton kinetic energy
Figure 5.17:Energy changes from static-lattice to dynamic-lattice simulations as functions of electron density ρ = N∕Ω. (a) zero-point energies of the candidate structures. (b) total potential energy change due to dynamic protons. (c) electronic kinetic energy change due to dynamic protons. (d) proton kinetic energy. Color and marker label candidate structures. The solid lines are guides to the eyes.

An immediate consequence of the higher kinetic and potentials energies in the dynamic-lattice simulation is higher virial pressure, about 20 GPa higher than its static-lattice counterpart. Interestingly, this brings the dynamic-lattice QMC pressures into better agreement with static-lattice vdW-DF pressures, as we saw in Fig. 5.23 that the static-lattice QMC pressures are systematically lower by 15 GPa.

The energy-density and enthalpy-pressure relations from the dynamic-lattice simulations are shown relative to the same reference EOS eq. 5.9 in Fig.5.18. Figure 5.18(a) shows the total energy as a function of density. First, due to ZPE, the window of energy is shifted up by roughly 500 meV/p from its static-lattice counterpart Fig. 5.15(a). Second, the relative energy among molecular structures remain similar. The C2/c-24 and Cmca-12 structures remain almost degenerate at all densities, whereas the Cmca-4 transitions from being 20 meV/p higher in energy at rs = 1.31 to being the lowest-energy structure at sufficiently high density. The transition density for Cmca-4 does decrease significantly, from rs = 1.19 in the static-lattice case to rs = 1.25 when ZPE is included. In contrast, the overall trend of the molecular-phase total energy is noticeably changed by the inclusion of ZPE. Instead of staying flat relative to the reference EOS, the dynamic-lattice energies of the molecular structures all trend upwards. Similar effects can be observed in the atomic-phase energy-density relation, although the overall upward shift due to ZPE is about 40 meV/p less. In summary, ZPE stabilizes the atomic phase by roughly 40 meV/p and the Cmca-4 phase by 10 meV/p relative the C2/c-24 and Cmca-12 without changing the relative angles at which their EOSs cross.

pict (a) energy-density relation

pict (a) enthalpy-pressure relation
Figure 5.18:Dynamic-lattice energy and enthalpy relative to reference. Lines are quadratic fits. The crossed out points are excluded from the fit.

The enthalpy-pressure relation in Fig. 5.18(b) tells the same story and we can read off the transition pressures among the candidate structures. We see a transition from C2/c-24 to Cmca-4 slightly below 500 GPa, then a molecular-to-atomic transition around 660 GPa from Cmca-4 to I41/amd. The calculated transition pressure to the atomic phase is 160 GPa higher than the previous QMC calculation [2] and a reported experimental observation [97]. However, the previous calculation used the harmonic approximation to calculate proton ZPE, while the experiment made little characterization of the final sample and has yet to be reproduced. As such, a 660 GPa transition pressure is not outside the realm of possibilities. Admittedly, the current approach has its own drawbacks. The most concerning approximation is the use of static-lattice orbitals in the electron-ion wave function eq. (5.5). This “frozen orbital” approach incurs around 80 meV/p more fixed-node error than a Born-Oppenheimer wave function fully-optimized at each new geometry, although this error cancels to a large extent between molecular and atomic phases.

5.3.3 Effect of Dynamic Ions on Electronic Correlation Functions

pict
(a) static lattice

pict
(b) dynamic lattice
Figure 5.19:DMC electronic pair correlation functions at rs = 1.25. Color denotes candidate structure, while the black lines are g(r) of the unpolarized homogeneous electron gas at the same density as parametrized by P. Gori-Giorgi, F. Sacchetti and G. B. Bachelet (GSB) [61].

The electron-electron pair correlation function in real space, g(r), are shown in Fig. 5.19. The correlation functions of the 3D unpolarized homogeneous electron gas are also shown as reference. The bottom panel of each plot shows the difference between the hydrogen g(r)s from those of the electron gas. This difference highlights the modification of the electron liquid due to the ionic lattice. The dynamic-lattice electronic g(r)s are identical to those from the static-lattice calculations on the scale of the difference panel (0.1). For the molecular structures, both the same-spin (u-u) and opposite-spin (u-d) g(r)s are the same over the entire range from zero (r = 0) to the supercell image distance (r = RWS). The most notable deviation from the electron gas is an enhanced correlation between opposite-spin electrons at small separation. This is likely due to the presence of covalent bonds, formed by a pair of opposite-spin electrons at the center of each H2 molecules. A more subtle difference is the presence of small periodic modulations of the electronic g(r) that appear to persist at long distances. The atomic structure shows no enhancement of opposite-spin correlation function at r = 0, whereas it shows the same periodic modulations as the molecular structures at long range.

pict
(a) static

pict
(b) dynamic
Figure 5.20:Electron-electron static structure factor S(k).

Pair correlation function in reciprocal space, the electron-electron static structure factor S(k), is shown in Fig. 5.20. At all densities and for all structures, both static-lattice and dynamic-lattice results are in excellent agreement with RPA S(k) for the electron gas. However, small increase in correlation can be seen at intermediate range k∕kF 1.5. Further, subtle differences at long-range (k 0) can make the electronic properties of the candidate structures qualitatively different from those of the electron gas and from each other.

pict
(a) static lattice

pict
(b) dynamic lattice
Figure 5.21:Upper bound on inverse dielectric constant.

The inverse dielectric function is bounded from above by 𝜖1(k) 1 Γk2, where

Γk = 2mS (k )
-ℏ2k2--. (5.11)

The inequality becomes an equality if there is only a single excitation, such as plasmon, at long wavelength. The limit limk01 Γk2 can be a good approximation to the inverse dielectric constant of an isotropic system. Importantly, if this limit reaches zero, then the material is metallic. In Fig. 5.21, the isotropic limit is shown as a function of density. There are sizable systematic errors in these results due to a lack of data at small k and the choice of fitting range. These systematic errors are not precisely estimated, but its size should be comparable to the oscillation of the data points. The inaccessible region around k = 0 increases with density, so the high-density results are less reliable than the low-density ones. In the static-lattice results Fig. 5.21(a), the atomic structure is metallic at all densities as expected. The Cmca-4 and C2/c-24 structures become metallic at roughly rs = 1.22 and rs = 1.17, respectively. These densities correspond to 550 GPa and 800 GPa, which are higher than experiment [17] and calculation [18]. It is possible that the molecular structures first start to conduct in-plane while being insulating across the planes. In this case, the isotropic dielectric bound derived from spherically averaged S(k) will overestimate the metalization pressure. The dynamic-lattice results Fig. 5.21(b) show all structures as metallic at densities higher than rs = 1.26 (460 GPa) except for C2/c-24, which remains borderline insulating.

5.3.4 Proton-proton g(r) and S(k)

pict
(a) pair correlation function g(r)

pict
(b) static structure factor S(k)
pict
(c) Bragg peaks of p-p S(k)
pict
(d) long wavelength limit
Figure 5.22:Proton-proton pair correlation functions around rs = 1.25 (480 GPa). The solid lines in (b) are quadratic fits to the smooth part of S(k). (c) Bragg peaks of the p-p S(k) (d) long wavelength limit S(0). |G1| ≈ 1.81 bohr1, |G2| ≈ 1.87 bohr1 for all three molecular structures. |G3| ≈ 2.5 for C2/c-24 and Cmca-4 and 1.93 for Cmca-12. The atomic structure’s Bragg peak can be found between 1.83 and 1.88 bohr1. The same outliers as Fig. 5.18 are excluded from the quadratic fits in (d).

The proton-proton pair correlation function of all candidate structures at rs 1.25 are shown in Fig. 5.22(a). Every molecular structure has a peak at the molecular bond length, in agreement with the DFT bond lengths shown in Fig. 5.5(b). C2/c-24 bond length is around 1.35 bohr, Cmca-4 around 1.40 bohr, and the Cmca-12 structure is in between. The C2/c-24 and Cmca-12 structures have remarkably similar p-p g(r). Since they are almost degenerate over the entire density range explored, we can infer that the slight monoclinic distortions of their unit cells and the small differences in their molecular bond lengths do not contribute significantly to the total energy. The less distorted C2/c-24 structure is more stable. The Cmca-4 structure is significantly different from C2/c-24. Cmca-4 has second- and third-nearest neighbor peaks at 2 and 3 bohr, respectively, rather than C2/c-24’s 2.5 and 3.2. The size and shape of these peaks are more similar to those in the atomic structure than to the other molecular candidates.

The long-range fluctuation and order of the protons can be observed in the proton-proton static structure factor. Its fluctuating part is shown in Fig. 5.22(b). All three molecular candidates have very similar levels of long-range fluctuations, whereas the atomic structure has noticeably more.

The molecular structures have more crystalline signatures than the atomic phase. As shown in Fig. 5.22(c), after spherical average, each molecular structure has at least three Bragg peaks for k < 4 bohr1, while the atomic structure has only one. The C2/c-24 and Cmca-12 Bragg peaks have very similar intensities and remain relatively unchanged by pressure. However, the Cmca-4 structure’s most intense Bragg peak is comparable with that in the atomic structure and decreases in intensity as pressure increases. Finally, the long wavelength limit of the static structure factor can be related to the isothermal compressibility κ n(∂P
∂n-)|T

nkB= S(0), (5.12)

where n is the density. We can extract the long wavelength limit of the p-p S(k) by fitting its fluctuating part to a quadratic polynomial as shown in Fig. 5.22(a). The value of the fit extrapolated to k = 0 is shown as a function of pressure for all candidates in Fig. 5.22(d). The compressibility of C2/c-24 is approximately independent of pressure, whereas that of the Cmca structures and the atomic structure increase with pressure at a similar rate. The atomic structure is twice as compressible as the molecular structures at all pressures.

pict
(a) static-lattice absolute scale

pict
(b) dynamic-lattice absolute scale
pict
(c) static-lattice relative to P0
pict
(d) dynamic-lattice relative to P0
Figure 5.23:Pressure-density relation. The solid lines are clamped-ion vdW-DF pressures, whereas the symbols are QMC pressures. (a) (c) show static-lattice results, while (b) (d) show dynamic lattice results. Color denotes different candidate structures. The open and filled symbols represent VMC and linearly extrapolated DMC estimator results. They overlap on the absolute scale in (a) but can be seen to differ by 1 to 5 GPa on a relative scale in (b). The reference pressure-density relation P0 = dE0∕dv is calculated from the reference EOS E0(1∕v). The same clamped-ion reference is used for both static-lattice and dynamic-lattice results.

5.4 Conclusion

The ordering of the candidate structures from this dynamic-lattice study show similar features as the previous QMC study using vdW-DF geometries and harmonic ZPE [2]. The Cmca-12 structure is nearly degenerate with C2/c-24 but slightly less stable at low pressures. There is a transition from C2/c-24 to Cmca-4 around 500 GPa. Then, a Cmca structure transitions to the I41/amd atomic structure at a higher pressure. However, we predict no stability range for the Cmca-12 structure, whereas in Ref. [2], C2/c-24 transitions to it at 424(3) GPa. The most significant disagreement lies in the atomic phase. We see the atomic phase becomes more stable relative to C2/c-24 around 610 GPa, whereas Ref. [2] shows this crossing around 450 GPa. Our final prediction for the molecular-to-atomic transition pressure is 660 GPa, which is significantly higher than the previous experimental result [97]. However, as already mentioned in Sec. 5.3.2, this prediction is not impossible. The dynamic-lattice DMC method is a conceptually simple way to directly simulate an electron-ion system and offers more ionic observables as well as direct comparisons with the static-lattice system than a single clamped-ion calculation. A Born-Oppenheimer study using the DMC energy surface is needed to bridge the previous static-lattice and harmonic ZPE study with the current dynamic-lattice nonadiabatic calculation. A detailed comparison among the three variants will hopefully resolve some of the disagreements we currently have.

Appendices

Chapter A
Nonadiabatic Coupling

The electron-ion problem can be made more tractable at the cost of a physically motivated Born-Oppenheimer approximation (BOA). I define the BOA at the end of this section, but first lay out the exact formulation in eq. (A.4) so that the content of the approximation is clear. Ions move much slower than electrons due to their heavy mass (mI 103 to 105mi), so it is sensible to isolate the ionic degrees of freedom and consider the electronic part separately. The coupling between the ionic and electronic problems is presumably weak because of the separation of timescales.

Ĥ = I ℏ2
2mI-I2 + (R;R I), (A.1)

where is the clamped-ion or electronic hamiltonian, which typically defines the ultimate goal of an electronic structure method. The semicolon in (R;RI) indicates that the electronic hamiltonian is only parametrically dependent on the ion positions RI. M. Born and R. Oppenheimer (BO) [121] first utilized this separation of timescales to study diatomic molecules in 1927. As explained around eq. (27) and (28) in Ref. [121], BO expressed the electronic hamiltonian as a Taylor expansion around the equilibrium positions of the ions. They discussed results using the first four leading order terms in the vibration amplitude of the ions. Thus, what we define as “the” BOA can be ambiguous. Here, I follow the interpretation by G. A. Worth and L. S. Cederbaum [9], which is equivalent to assuming a product ansatz eq. (A.7), but without a “diagonal correction”.

If one can obtain the eigenstates of the electronic hamiltonian {ψk} at any ion configuration RI

(R;RI)ψk(R;RI) = Ek(RI)ψk(R;RI), (A.2)

then one can expand an eigenstate of the full hamiltonian Ĥ in the basis of electronic eigenstates

Ψl(R,RI) = k=0χ lk(RI)ψk(R;RI), (A.3)

where the expansion coefficients χlk(RI) will later be identified with the ionic wave function in the Born-Oppenheimer approximation. l runs over the full electron-ion hamiltonian’s eigenstates, which can have both ionic (vibrational) and electronic characters. The coefficient for one of these vibronic states cannot be determined separately for each electronic level k in general. To see this, substitute the l = 0 expansion eq. (A.3) into the time-dependent Schrödinger equation for the full electron-ion hamiltonian (drop l for simplicity)

(              )
     ∑    ℏ2
  ˆℋ −    2M--∇2I
      I     I(        )
  ∑
     χkψk
   k = id
dt(        )
  ∑
     χkψk
   k apply operators
kEkχkψk I-ℏ2-
2MI( 2                         2  )
∇ Iχkψk + 2∇I χk ⋅∇Iψk + χk∇ Iψk = kiχkψk apply ψj
kEkχkδjk I ℏ2
2M--
  I(                           )
 ∇2Iχkδjk + 2∇I χk ⋅Fjk + χkGjk = kiχkδjk perform k
(        2        )
  − ∑  -ℏ--∇2+ Ej
    I  2MI  Iχj (         2 (              )   )
 ∑   ∑  -ℏ-- 2F jk ⋅∇I + Gjk χk
  k   I 2MI    I         I = iχj, (A.4)

where the matrix elements for gradient (derivative-coupling terms) and laplacian (scalar-coupling terms) in the electronic eigenstates basis are

(
|{ F jk= ∫ drψ∗(r;R)∇  ψ (r;R )
    I   ∫    j       I k
|( GjkI =  dr ψ∗j(r;R )∇2Iψk(r;R ). (A.5)

The matrix elements that couple different electronic states in eq. (A.4) are named nonadibatic coupling operators by Worth and Cederbaum [9]

Λjk = I-ℏ2-
2MI(              )
 2F jk ⋅∇  + Gjk
   I    I    I. (A.6)

Every term in Λjk has an inverse ion mass prefactor -ℏ2
2MI, so they are expected to be small in most cases. There are two common approximations of Λkj, the first is to set the entire matrix to zero, the second is to set only the off-diagonal terms to zero. Both approximations decouple (A.4), allowing the complete separation of electronic and ionic motions. Many different and sometimes conflicting names have been given to these two approximations including Born-Huang, Born-Oppenheimer and adiabatic approximation. To fix nomenclature, I will call the all-zero approximation, Λjk = 0, j,k, the Born-Oppenheimer approximation (BOA). The diagonal terms Λjj are considered diagonal Born-Oppenheimer correction (DBOC). Non-zero off-diagonal elements are responsible for nonadiabatic effects.

The ground state in the BOA is a product of an ionic and an electronic component

ΨlkBO(R,R I) = χlk(RI)ψk(R;RI), (A.7)

where a set of vibrational states labeled by l can be defined over a particular electronic state k. χl(RI) obeys its own Schrödinger equation on an effective potential energy surface provided by an eigenvalue of the electronic hamiltonian Ek(RI) = ⟨       ⟩
 ψk| ˆℋ|ψk, a.k.a. the Born-Oppenheimer potential energy surface (BO-PES)

(   ∑    2        )
  −    -ℏ--∇2I + Ek
     I 2MIχl = iχl. (A.8)

Once the ionic eigenstates are obtained by diagonalizing eq. (A.8), the total energy of the electron-ion system is finally obtained as

ElkBO ⟨       ∑   ℏ2      ⟩
 χl|Ek −    ----∇2I|χl
         I 2MI. (A.9)

E00BO differs in two ways from the electronic ground-state energy

E0 ⟨       ⟩
 ψ0|ℋ ˆ|ψ0(RI), (A.10)

which is a function of the positions of the ions RI. First, in E00BO the electronic energy is averaged over a distribution of ion configurations |χ0|2(RI) rather than evaluated at one fixed configuration RI. This quantum delocalization effect raises the total energy from the bottom of the BO-PES RIe = argmin
  RI E0(RI), which would have been the electron-ion ground state if the ions were classical. Second, the ions have kinetic energy even at absolute zero, which also contributes a positive term to the total energy. The difference between the electron-ion ground-state energy and the electronic one is the zero-point energy (ZPE). In the BOA, ZPE contains only two terms from delocalization and kinetic energy of the ions.

Within the BOA framework, solving the electron-ion problem for a particular combination of vibrational l and electronic state k involves finding the kth eigenvalue of the clamped-ion electronic problem (R;RI) for many ion configuration RI. There are established first-principle molecular dynamics and Monte Carlo methods for achieving this, but they are not practical for even moderately sized system, e.g., O(1000) atoms, because the computational cost of electronic structure methods generally scale as N3 or worse.

The main short-fall of the BOA is its lack of pathways for the ions to transfer energy to the electrons. This is critical in the study of radiation damage, where a fast moving ion can transfer energy to both the electrons and the ions in a material. Further, for chemical reactions involving vibration-assisted bond breaking, the BOA reduces the number of pathways dissociation can happen, thereby resulting in an incomplete description. The BOA can also break down if the electrons interact with a particle much lighter than an atomic nucleus such as a positron or a muon. Finally, the nonadiabatic coupling terms can diverge when two electronic states cross, e.g., at a conical intersection. Thus, it is sometimes important to go beyond the BOA.

There are two small parameters that control the scale of nonadiabatic coupling eq. (A.6). One is clearly the inverse ionic mass -1-
MI, while the other is the difference between electronic energy levels 𝜖j 𝜖k. This can be seen from an explicit form of FIjk ⟨ψj|∇I |ψk⟩ in the derivative-coupling term. Consider the effect of ion motion on the electronic problem, i.e., take I of the time-independent electronic Schrödinger equation

I(ψk) = I(𝜖kψk) ψkI + Iψk = ψkI𝜖k + 𝜖kIψk apply ψj
(∫          )
    ψ∗jψk∇I ˆℋ + 𝜖jFIjk = I𝜖kδjk + 𝜖kFIjk solve for F Ijk
FIjk = ⟨         ⟩
 ψj|∇I ℋˆ|ψk  + ∇I 𝜖kδjk
----------------------
       𝜖k − 𝜖j. (A.11)

If the electronic eigenstates are defined to be orthonormal, then the real part of the derivative-coupling vectors vanish (FIkj) + FIjk = 0 Re[FIjk] = 0 as derived in eq. (A.12).

(|    jk
{  FI  = ⟨ψj|∇I ψk⟩
|(  ⟨ψ |ψ  ⟩ = δ
     j  k    jk I⟨ψ |ψ  ⟩
  j  k = 0
⟨∇I ψj|ψk ⟩ + ⟨ψj|∇I ψk⟩ = (FIkj) + F Ijk = 0. (A.12)

FIkj can be interpreted as follows: the motion of the ions apply an imaginary “force” ⟨ψ |∇  |ψ ⟩
  j  I  k that drives an electronic transition from state k to state j. This interpretation has led to surface hopping methods for classical and quantum ions [122, 123], which have been applied successfully to describe proton transfer and proton-coupled electron transfer reactions [124]. Further, this “force” is inversely proportional to the energy separation between the two eigenstates. The derivative-coupling term is considered more interesting than the scalar-coupling term due to its potential divergence as 𝜖k 𝜖j.

If the wave function is real, then there is no derivative coupling within the same electronic state FIjj = ⟨ψj|∇I ψj⟩ = ⟨∇Iψj|ψj⟩ = (FIjj)FIjj = 0. In this case, the DBOC is simply the expectation value of the ion kinetic operator in the electronic state

Λjj = I ℏ2
----
2MIGIjj = ψj I ℏ2
----
2MII2ψ j = ⟨  ⟩
 TˆIj. (A.13)

The diagonal correction for the hydrogen molecule was studied extensively by Kolos and Wolniewicz [34, 125, 126]. For the atomization energy of H2, the DBOC was found to be 4.947 cm1, which is only a 0.0129% correction of its clamped-ion value of 38292.7 cm1. The nonadiabatic contribution to ionization and atomization energies of a few atoms and small molecules are explored in Ref. [127] and references therein.

Chapter B
Minimum-basis H2

To derive the Hartree-Fock equations, we use a Slater determinant to evaluate the total energy, then minimize it. Consider N spinless fermions, labeled using i,j,k,, in N orbitals χab,N. Given determinant wavefunction |Ψ0⟩ = |χa χb...χN ⟩ and electronic Hamiltonian made up of only one- and two-electron terms = i=1Nh(i) + i=1N j>iNv(i,j). The total energy is

E = a=1N[a|h|a] + 1
2 a,b=1N[aa|bb] [ab|ba], (B.1)

where [a|h|a] denotes, and [aa|bb] denotes [128].

Constraint minimization of eq. (B.1) with the extra requirement that each spin orbital is doubly occupied leads to the restricted Hartree-Fock (RHF) Fock operator. Its first N∕2 eigenvectors are the spin orbitals in the lowest-energy Slater determinant. The lowest energy value can be obtained by a weighted sum of its eigenvalues according to the occupation of the spin orbitals.

Instead of starting with the tedious derivation of the Fock operator and its iterative numerical solver, I will first show a concrete application of RHF to minimum-basis hydrogen molecule (H2). On p. 140 of A. Szabo and N. S. Ostlund, the restricted Fock operator in any basis {ϕμ} is written as

Fμν = Hcoreμν + a=1N∕22(μν|aa) (μa|), (B.2)

where a labels molecular orbitals, which are eigenstates of the Fock operator. We immediately note that the Fock operator is a peculiar one-electron operator that depends on its own eigenstates. A self-consistent solution to Fμν typically involves guessing, checking and iterating.

Hcoreμν is the one-electron part of the Hamiltonian expressed in the given basis

Hcoreμν = dr1ϕμ(r 1)(        ∑           )
  − 1 ∇21 −   ---ZA----
   2      A  |r1 − RA |ϕν(r1). (B.3)

The two-electron integral notation (μν|λσ) is defined by eq. (3.155) in Szabo

(μν|λσ) = ∫∫dr1dr2ϕμ(r 1)ϕν(r1)   1
--------
|r1 − r2|ϕλ(r 2)ϕσ(r2). (B.4)

The first term in the sum of eq. (B.9)

Jμν (μν|aa) (B.5)

is often called the Coulomb or direct operator, because it describes the Classical density-density interaction of charged particles having density distribution ϕa(r2)ϕa(r2). The second term

Kμν (μa|νa) (B.6)

is the exchange operator and has no classical interaction. The exchange-correlation contribution to the Fock matrix is sometimes called the Hartree-Fock effective potential operator

V effμν 2Jμν Kμν. (B.7)

Suppose each molecular orbital a is written as a linear combination of the basis functions

ψa = μ=1KC μaϕμ, (B.8)

then the Fock operator can be written as (from Roothaan equations)

Fμν = Hcoreμν + λσPλσ[                 ]
 (μν|σλ )− 1(μλ|σν)
          2, (B.9)

where Pλσ = 2 a=1N∕2CλaCσa is the density matrix of the trial states.

Conceptually, the simplest approach would be to use the ground-state wavefunctions of the two hydrogen atoms as the basis for the hydrogen molecule. We can guess the ground-state wavefunction of the hydrogen molecule. First, the spins of the two electrons anti-align, so they are distinguishable particles. Second, due to symmetries imposed by the two protons, the ground state must be equal superposition of the two basis functions. Third, the lowest-energy solution has no node. Therefore, the ground state of H2 in the minimum basis is

ψ1 = [2(1+ S12)]12(ϕ1 +ϕ2), (B.10)

where S12 = ⟨ϕ1|ϕ2⟩. That is C11 = C21 = [2(1+ S12)]12

Pλσ = [2(1+ S  )]
      1212(       )
|  1  1 |
(  1  1 ). (B.11)

This guess was obtained as early as 1927 by V. W. Heitler and F. London [129]. Unfortunately, the multi-center integrals eq. (B.3) and (B.4), needed to evaluate the total energy, have no analyical form in the basis of Slater type orbitals (STOs) (see thesis of Michał Lesiuk). Thus, Heitler-London used an upper bound to approximate the two-electron integral and obtain a bond length of 1.5 bohr and binding energy of 2.5 eV, noticeably different from the experimental values of 1.4 bohr and 4.5 eV.

In modern quantum chemistry, instead of directly approximating the integrals, we analytically evaluate the integrals by approximating each basis function as a sum of Gaussians. This reduces the multi-center integrals to single-center integrals, because a product of Gaussians centered on different atoms is also Gaussian but with a different center. The so-called STO-3g basis expresses a STO as a sum of 3 “primitive Gaussians” (see eq. (3.225) of Szabo). Using this basis, the bond length and binding energy become 1.35 bohr and 3.2 eV, having roughly half the discrepancy with experiment when compared to the Heitler-London values.

Figure B.1 shows the STO-3g basis compared to the exact STO it approximates

χ(r) (   )
  ζ3
  π12eζr ϕ(r) = i=13c i(    )
  2αi
  π34eαir2 , (B.12)

where the exponents α and coefficients c are given below:





α c


3.42525091 0.15432897
0.62391373 0.53532814
0.16885540 0.44463454




pict
Figure B.1:STO-3G

For H2, the STO-3G basis consists of only two 1s functions, each centered around a nucleus.

(             (    )
|||         ∑3    2αi 3∕4 −αi|r−R1|2
{  ϕ1(r) = i=1ci   π     e
||         ∑3  ( 2αi)3∕4         2
|(  ϕ2(r) =   ci  -π-    e−αi|r−R2|
          i=1 , (B.13)

where R1 = 0, and R2 = 1.4  bohr near equilibrium. At 1.4 bohr separation, the kinetic and the electron-ion interaction matrices evaluate to

Tμν ⟨            ⟩
 ϕ  |−  1∇2|ϕ
   μ   2    ν = (                       )

|(  0.76003188  0.23645465 |)
   0.23645465  0.76003188, (B.14)
V μν ⟨    ∑2           ⟩
  ϕμ|   ---1---|ϕ ν
     i=1 |r − Ri| = (                          )
   − 1.88044088 − 1.19483461
|(                          |)
   − 1.19483461 − 1.88044088, (B.15)

which sum to the 1-electron hamiltonian Hcoreμν by eq. (B.3)

Eigenvectors of Hcoreμν are typically used to construct the initial density matrix to start a self-consist solution of the Hartree-Fock equations. However, in the case of H2, these eigenvectors coincide with the final solution, so we obtain the converged density matrix with no iteration from eq. (B.11)

Cμ1 = (              )
|   0.70710678  |
(              )
   − 0.70710678; P μν = (                      )
| 0.60265716 0.60265716|
(                      )
  0.60265716 0.60265716. (B.16)

Finally, we can evaluate the so-called electron repulsion integrals (eris) and the Fock matrix eq. (B.9)

Table B.1:symmetry-inequivalent electron repulsion integrals for H2 in STO-3G.










μ ν λ σ (μν|λσ)





1 1 1 1 0.774605930
1 1 1 2 0.444107650
1 1 2 2 0.569675915
1 2 1 2 0.297028535










Jμν = (                       )
|  0.67271523  0.44665109 |
(                       )
   0.44665109  0.67271523;  K μν = (                       )
|  0.59055879  0.52880753 |
(                       )
   0.52880753  0.59055879; (B.17)
Fμν = (                          )
|  − 0.36553735  − 0.59388537 |
(                          )
   − 0.59388537  − 0.36553735. (B.18)

The total energy is 1.11671432 ha, while the electronic contribution is 1.831 ha, before adding the ion-ion repulsion V ii = 11.4 ha. Interested reader should reproduce the Fock matrix for STO-3G H2 at 1.4 bohr separation, i.e. eq. (B.18), to consolidate a practical understanding of RHF.

In the PZ formulation of KS-DFT, the only difference between the LDA and the RHF calculations lies in the “exchange” matrix

Kμν= (                      )

|( 0.38980073 0.25499926|)
  0.25499926 0.38980073, (B.19)

which now contains an approximation to both exchange and correlation effects rather than exact exchange in the case of HF. The Fock matrix is

Fμν = (                         )
| − 0.16477935  − 0.32007715 |
(                         )
  − 0.32007715  − 0.16477935, (B.20)

and the LDA electronic contribution is 1.73929592 ha.

Chapter C
Derivation of Higher-order Band Gap Correction

Total potential energy

V = Ne-
Ω k01
2vkSk, (C.1)

where Sk 1-
Ne⟨ρ ρ  ⟩
  k −k. ρk j=1Neeikrj is the electronic density in reciprocal space. Ω is system volume. ⟨⟩ denotes average over walker ensemble. The potential energy V can be written as a sum of static (density) contribution and fluctuating contribution

V = (                  )
    ∑
( 1-   1vk⟨ρk⟩⟨ρ−k⟩)
  Ω k⁄=02 + (                               )
    ∑
( 1-    1vk(ρk − ⟨ρk⟩)(ρ−k − ⟨ρ−k⟩))
  Ω k⁄=0 2. (C.2)

The fluctuating part can be used to calculate leading-order finite-size correction to the band gap, where the static part leads to the next-to-leading-order correction.

V s 1-
Ω k01
2vk⟨ρ ⟩
  k ⟨ρ ⟩
  −k. (C.3)

The goal of this Appendix is to find the FSC formula for V s. Assuming the finite-size error in the electron density can be fully recovered by twist-averaging, the infinite-system potential energy can be computed using the twist-averaged density ⟨ρk⟩, where twist average is denoted by overline

V sN→∞ = 1
--
Ω k01
-
2vk⟨ρk⟩  ⟨ρ−k⟩. (C.4)

Therefore, the FSC of V s is

δV s V sN→∞V s = 1
Ω- k01
2vk[---- -----          ]
 ⟨ρk⟩ ⟨ρ−k⟩− ⟨ρk⟩⟨ρ−k⟩. (C.5)

⟨ρk⟩ differs from twist to twist, so there is one such correction for each twist. Define

Ck [---- -----           ]
 ⟨ρk⟩ ⟨ρ−k⟩− ⟨ρk⟩⟨ρ−k⟩, (C.6)

then the FSC of the potential energy is

δV s = 2π
-Ω- k0Ck∕k2. (C.7)

In simulation, we assemble the MNe + 1 system with varying density at each twist

ρkMNe±1 = 𝜃(                   )
 ⟨ρk⟩Ne±1,ϕ − ⟨ρk ⟩Ne,ϕδ𝜃,ϕ + ⟨ρk⟩Ne,𝜃. (C.8)

The GCTA corrected charge density varies only at the target twist ϕ

ρkMNe±1 = 𝜃(                   )
 ⟨ρk⟩Ne±1,ϕ − ⟨ρk⟩Ne,ϕδ𝜃,ϕ + ρkNe . (C.9)

The neutral mean density ρkNe has no dependence on the twist 𝜃

ρkNe -1-
M 𝜃⟨ρ⟩Ne,𝜃. (C.10)

For simplicity, define the charge density of the particle/hole as

Πk ⟨ρk⟩Ne±1,ϕ ⟨ρk⟩Ne,ϕ. (C.11)

Now, the GCTA correction can be clearly seen as the replacement of neutral twist density ⟨ρk⟩Ne,𝜃 with neutral mean density ρkNe

(                              (               )
||  ρMNe±1 ≡    ∑ ρNe±1(𝜃) =  ∑  Π  δ   + ⟨ρ ⟩
{   k           𝜃 k          𝜃    k 𝜃,ϕ    k Ne,𝜃
||  ρ-MNe±1 ≡  ∑  ρ-Ne±1(𝜃) = ∑ (Π  δ  + ρ-Ne)
(   k          𝜃 k           𝜃    k𝜃,ϕ   k. (C.12)

The charged twist density ρkNe±1(𝜃) and the charged mean density ρkNe±1(𝜃) are defined by the summand on each line. A likely cause of confusion here is that the charged mean density depends on the twist 𝜃. Further, it is not the mean of the charged twist density

ρkNe±1(𝜃) 1
---
M 𝜃 𝜃ρkNe±1(𝜃). (C.13)

This correction does not affect properties linear in ρk, but does change the potential energy, which is quadratic in ρk. If the MNe system is constructed from M independent simulations each containing Ne electrons, then the charge density is different at each twist. The Hartree contribution to the total potential energy is

V sMNe+1 = 1-
Ω 𝜃 kG1
2vk ρkNe+1(𝜃)ρ kNe+1(𝜃). (C.14)

One can speed up the convergence of eq. (C.14) to the thermodynamic limit by replacing twist density with mean density

limN→∞V sMNe+1 = 1
--
Ω 𝜃 kG1
-
2vk ρkNe+1(𝜃)ρ kNe+1(𝜃). (C.15)

Define the correction factor

CkNe+1(𝜃) ρkNe+1(𝜃)ρ kNe+1(𝜃) ρ kNe+1(𝜃)ρ kNe+1(𝜃), (C.16)

then the GCTA finite-size correction to the Hartree contribution to the potential is

δV sMNe+1 = 1
--
Ω kG1
-
2vk  𝜃CkNe+1(𝜃). (C.17)

Eq. (C.16) can be much simplified when calculated relative to the neutral state

CkNe+1(𝜃) = kδ𝜃,ϕ + ρkNe )(Πkδ𝜃,ϕ + ρkNe ) kδ𝜃,ϕ + ⟨ρk⟩Ne,𝜃)(Πkδ𝜃,ϕ + ⟨ρk⟩Ne,𝜃)
= Πkδ𝜃,ϕ(ρkNe ⟨ρ−k⟩Ne,𝜃) + (ρkNe ⟨ρk⟩Ne,𝜃kδ𝜃,ϕ
+ [                         ]
 ρkNeρ−kNe − ⟨ρk⟩Ne,𝜃 ⟨ρk⟩Ne,𝜃
= Πkδ𝜃,ϕ(ρkNe ⟨ρ−k⟩Ne,𝜃) + (ρkNe ⟨ρk⟩Ne,𝜃kδ𝜃,ϕ + CkNe (𝜃). (C.18)

Therefore, the correction to electron addition energy is

δμs+ = 1
--
Ω kG1
-
2vk[∑                   ]
    CNek+1 (𝜃 )− CNke(𝜃)
  𝜃
= 1-
Ω kG1
2vk{[--Ne          ]         [---Ne          ]}
  ρk   − ⟨ρk⟩Ne,ϕ  Π−k + Πk ρ− k  − ⟨ρ−k⟩Ne,ϕ. (C.19)

References

[1]

Sam Azadi et al. “Dissociation of High-Pressure Solid Molecular Hydrogen: A Quantum Monte Carlo and Anharmonic Vibrational Study”. In: Phys. Rev. Lett. 112.16 (2014), p. 165501. issn: 0031-9007. doi: 10.1103/PhysRevLett.112.165501. url: https://link.aps.org/doi/10.1103/PhysRevLett.112.165501.

[2]

Jeremy McMinis et al. “Molecular to Atomic Phase Transition in Hydrogen under High Pressure”. In: Phys. Rev. Lett. 114.10 (2015), pp. 1–6. issn: 10797114. doi: 10.1103/PhysRevLett.114.105305.

[3]

N. D. Drummond et al. “Quantum Monte Carlo study of the phase diagram of solid molecular hydrogen at extreme pressures”. In: Nat. Commun. 6 (2015), p. 7794. issn: 2041-1723. doi: 10.1038/ncomms8794. arXiv: arXiv:1508.02313v1. url: http://www.nature.com/doifinder/10.1038/ncomms8794.

[4]

Carlo Pierleoni et al. “Liquid–liquid phase transition in hydrogen by coupled electron–ion Monte Carlo simulations”. In: Proc. Natl. Acad. Sci. 113.18 (2016), pp. 4953–4957. issn: 0027-8424. doi: 10.1073/pnas.1603853113. url: http://www.pnas.org/lookup/doi/10.1073/pnas.1603853113.

[5]

Peter M Celliers et al. “Insulator-metal transition in dense fluid deuterium”. In: Science 361.6403 (2018), pp. 677–682. issn: 0036-8075. doi: 10.1126/science.aat0970. url: https://www.sciencemag.org/lookup/doi/10.1126/science.aat0970.

[6]

D.M. Ceperley and B.J. Alder. “The calculation of the properties of metallic hydrogen using Monte Carlo”. In: Phys. B+C 108.1-3 (1981), pp. 875–876. issn: 03784363. doi: 10.1016/0378-4363(81)90742-7. url: https://linkinghub.elsevier.com/retrieve/pii/0378436381907427.

[7]

Vincent Natoli, Richard M. Martin, and David Ceperley. “Crystal Structure of Molecular Hydrogen at High Pressure”. In: Phys. Rev. Lett. 74.9 (1995), pp. 1601–1604. issn: 0031-9007. doi: 10.1103/PhysRevLett.74.1601. url: https://link.aps.org/doi/10.1103/PhysRevLett.74.1601.

[8]

Yueqing Chang and Lucas K. Wagner. “Effective spin-orbit models using correlated first-principles wave functions”. In: Phys. Rev. Res. 2.1 (2020), p. 013195. issn: 2643-1564. doi: 10.1103/PhysRevResearch.2.013195. url: https://link.aps.org/doi/10.1103/PhysRevResearch.2.013195.

[9]

Graham A. Worth and Lorenz S. Cederbaum. “BEYOND BORN-OPPENHEIMER: Molecular Dynamics Through a Conical Intersection”. In: Annu. Rev. Phys. Chem. 55.1 (2004), pp. 127–158. issn: 0066-426X. doi: 10.1146/annurev.physchem.55.091602.094335. url: http://www.annualreviews.org/doi/10.1146/annurev.physchem.55.091602.094335.

[10]

M Holzmann et al. “Backflow correlations for the electron gas and metallic hydrogen.” In: Phys. Rev. E. 68.4 Pt 2 (2003), p. 046707. issn: 1063-651X. doi: 10.1103/PhysRevE.68.046707. arXiv: 0304165 [cond-mat].

[11]

Burkhard Militzer. “Path Integral Monte Carlo Simulations of Hot Dense Hydrogen”. PhD thesis. University of Illinois at Urbana-Champaign, 2000.

[12]

W. B. Hubbard and B. Militzer. “A PRELIMINARY JUPITER MODEL”. In: Astrophys. J. 820.1 (2016), p. 80. issn: 1538-4357. doi: 10.3847/0004-637X/820/1/80. arXiv: 1602.05143. url: http://dx.doi.org/10.3847/0004-637X/820/1/80http://stacks.iop.org/0004-637X/820/i=1/a=80?key=crossref.acea06e37f7f7dc99ca975ec8290c997.

[13]

A. P. Drozdov et al. “Conventional superconductivity at 203 kelvin at high pressures in the sulfur hydride system”. In: Nature 525.7567 (2015), pp. 73–76. issn: 0028-0836. doi: 10.1038/nature14964. url: http://www.nature.com/articles/nature14964.

[14]

Ion Errea et al. “Quantum crystal structure in the 250-kelvin superconducting lanthanum hydride”. In: Nature 578.7793 (2020), pp. 66–69. issn: 0028-0836. doi: 10.1038/s41586-020-1955-z. url: http://www.nature.com/articles/s41586-020-1955-z.

[15]

E. Wigner and H. B. Huntington. “On the Possibility of a Metallic Modification of Hydrogen”. In: J. Chem. Phys. 3.12 (1935), pp. 764–770. issn: 0021-9606. doi: 10.1063/1.1749590. url: http://aip.scitation.org/doi/10.1063/1.1749590.

[16]

Ranga P Dias, Ori Noked, and Isaac F Silvera. “New Phases and Dissociation-Recombination of Hydrogen Deuteride to 3.4 Mbar”. In: Phys. Rev. Lett. 116.14 (2016), p. 145501. issn: 10797114. doi: 10.1103/PhysRevLett.116.145501. arXiv: 1510.05686. url: https://link.aps.org/doi/10.1103/PhysRevLett.116.145501.

[17]

Paul Loubeyre, Florent Occelli, and Paul Dumas. “Synchrotron infrared spectroscopic evidence of the probable transition to metal hydrogen”. In: Nature 577.7792 (2020), pp. 631–635. issn: 0028-0836. doi: 10.1038/s41586-019-1927-3. url: http://dx.doi.org/10.1038/s41586-019-1927-3http://www.nature.com/articles/s41586-019-1927-3.

[18]

Vitaly Gorelov et al. “Energy gap closure of crystalline molecular hydrogen with pressure”. In: Phys. Rev. Lett. 124.11 (2019), p. 116401. issn: 1079-7114. doi: 10.1103/PhysRevLett.124.116401. arXiv: 1911.06135. url: http://arxiv.org/abs/1911.06135.

[19]

Nicholas Metropolis et al. “Equation of State Calculations by Fast Computing Machines”. In: J. Chem. Phys. 21.6 (1953), pp. 1087–1092. issn: 0021-9606. doi: 10.1063/1.1699114. url: http://aip.scitation.org/doi/10.1063/1.1699114.

[20]

D. M. Ceperley. “Path integrals in the theory of condensed helium”. In: Rev. Mod. Phys. 67.2 (1995), pp. 279–355. issn: 0034-6861. doi: 10.1103/RevModPhys.67.279. url: https://link.aps.org/doi/10.1103/RevModPhys.67.279.

[21]

Carolyn Meldgin et al. “Probing the Bose glass–superfluid transition using quantum quenches of disorder”. In: Nat. Phys. 12.7 (2016), pp. 646–649. issn: 1745-2473. doi: 10.1038/nphys3695. arXiv: 1502.02333. url: http://www.nature.com/articles/nphys3695.

[22]

Stefano Baroni and Saverio Moroni. “Reptation Quantum Monte Carlo: A Method for Unbiased Ground-State Averages and Imaginary-Time Correlations”. In: Phys. Rev. Lett. 82.24 (1999), pp. 4745–4748. issn: 0031-9007. doi: 10.1103/PhysRevLett.82.4745. url: https://link.aps.org/doi/10.1103/PhysRevLett.82.4745.

[23]

Peter J. Reynolds et al. “Fixed‐node quantum Monte Carlo for molecules a) b)”. In: J. Chem. Phys. 77.11 (1982), pp. 5593–5603. issn: 0021-9606. doi: 10.1063/1.443766. url: http://aip.scitation.org/doi/10.1063/1.443766.

[24]

R. M. Martin, Reining L., and Ceperley D. M. Interacting Electrons : Theory and Computational Approaches. Cambridge University Press, 2016.

[25]

B. L. Hammond, W. A. Lester, and Reynolds P. J. Monte Carlo Methods in Ab Initio Quantum Chemistry. World Scientific Publishing Co., 1994.

[26]

D. M. Ceperley. “Fermion nodes”. In: J. Stat. Phys. 63.5-6 (1991), pp. 1237–1267. issn: 0022-4715. doi: 10.1007/BF01030009. url: http://link.springer.com/10.1007/BF01030009.

[27]

D. M. Ceperley. “Path-integral calculations of normal liquid He3”. In: Phys. Rev. Lett. 69.2 (1992), pp. 331–334. issn: 0031-9007. doi: 10.1103/PhysRevLett.69.331. url: http://link.aps.org/doi/10.1103/PhysRevLett.69.331https://link.aps.org/doi/10.1103/PhysRevLett.69.331.

[28]

D. M. Ceperley. “Path integral Monte Carlo methods for fermions”. In: Monte Carlo and Molecular Dynamics of Condensed Matter Systems. Ed. by K. Binder and G. Ciccotti. Italy: Editrice Compository, 1996.

[29]

D. M. Ceperley. “Ground state of the fermion one-component plasma: A Monte Carlo study in two and three dimensions”. In: Phys. Rev. B 18.7 (1978), pp. 3126–3138. issn: 0163-1829. doi: 10.1103/PhysRevB.18.3126. url: https://link.aps.org/doi/10.1103/PhysRevB.18.3126.

[30]

Markus Holzmann, Bernard Bernu, and D. M. Ceperley. “Many-body wavefunctions for normal liquid He3”. In: Phys. Rev. B 74.10 (2006), p. 104510. issn: 1098-0121. doi: 10.1103/PhysRevB.74.104510. url: https://link.aps.org/doi/10.1103/PhysRevB.74.104510.

[31]

Carlo Pierleoni et al. “Trial wave functions for high-pressure metallic hydrogen”. In: Comput. Phys. Commun. 179.1-3 (2008), pp. 89–97. issn: 00104655. doi: 10.1016/j.cpc.2008.01.041. arXiv: 0712.0161. url: http://www.sciencedirect.com/science/article/pii/S0010465508000210.

[32]

G. F Giuliani and G. Vignale. Quantum Theory of the Electron Liquid. Cambridge University Press, 2005.

[33]

T. Koopmans. “Über die Zuordnung von Wellenfunktionen und Eigenwerten zu den Einzelnen Elektronen Eines Atoms”. In: Physica 1.1-6 (1934), pp. 104–113. issn: 00318914. doi: 10.1016/S0031-8914(34)90011-2. url: https://linkinghub.elsevier.com/retrieve/pii/S0031891434900112.

[34]

W. Kolos and L. Wolniewicz. “Accurate Adiabatic Treatment of the Ground State of the Hydrogen Molecule”. In: J. Chem. Phys. 41.12 (1964), pp. 3663–3673. issn: 0021-9606. doi: 10.1063/1.1725796. url: http://aip.scitation.org/doi/10.1063/1.1725796.

[35]

Carlos L. Benavides-Riveros, Nektarios N. Lathiotakis, and Miguel A. L. Marques. “Towards a formal definition of static and dynamic electronic correlations”. In: Phys. Chem. Chem. Phys. 19.20 (2017), pp. 12655–12664. issn: 1463-9076. doi: 10.1039/C7CP01137G. arXiv: 1705.00238. url: http://xlink.rsc.org/?DOI=C7CP01137G.

[36]

Aron J. Cohen, P. Mori-Sanchez, and Weitao Yang. “Insights into Current Limitations of Density Functional Theory”. In: Science 321.5890 (2008), pp. 792–794. issn: 0036-8075. doi: 10.1126/science.1158722. url: https://www.sciencemag.org/lookup/doi/10.1126/science.1158722.

[37]

J. P. Perdew and Alex Zunger. “Self-interaction correction to density-functional approximations for many-electron systems”. In: Phys. Rev. B 23.10 (1981), pp. 5048–5079. issn: 0163-1829. doi: 10.1103/PhysRevB.23.5048. url: https://link.aps.org/doi/10.1103/PhysRevB.23.5048.

[38]

R. G. Parr and W. Yang. Density-Functional Theory of Atoms and Molecules. New York, NY: Oxford University Press, 1989.

[39]

P. Hohenberg and W. Kohn. “Inhomogeneous Electron Gas”. In: Phys. Rev. 136.3B (1964), B864–B871. issn: 0031-899X. doi: 10.1103/PhysRev.136.B864. url: https://link.aps.org/doi/10.1103/PhysRev.136.B864.

[40]

Mel Levy. “Electron densities in search of Hamiltonians”. In: Phys. Rev. A 26.3 (1982), pp. 1200–1208. issn: 0556-2791. doi: 10.1103/PhysRevA.26.1200. url: https://link.aps.org/doi/10.1103/PhysRevA.26.1200.

[41]

Elliott H. Lieb. “Density functionals for coulomb systems”. In: Int. J. Quantum Chem. 24.3 (1983), pp. 243–277. issn: 0020-7608. doi: 10.1002/qua.560240302. url: http://doi.wiley.com/10.1002/qua.560240302.

[42]

M. Levy. “Universal variational functionals of electron densities, first-order density matrices, and natural spin-orbitals and solution of the v-representability problem”. In: Proc. Natl. Acad. Sci. 76.12 (1979), pp. 6062–6065. issn: 0027-8424. doi: 10.1073/pnas.76.12.6062. url: http://www.pnas.org/cgi/doi/10.1073/pnas.76.12.6062.

[43]

W. Kohn and L. J. Sham. “Self-Consistent Equations Including Exchange and Correlation Effects”. In: Phys. Rev. 140.4A (1965), A1133–A1138. issn: 0031-899X. doi: 10.1103/PhysRev.140.A1133. url: https://link.aps.org/doi/10.1103/PhysRev.140.A1133.

[44]

D. M. Ceperley and B. J. Alder. “Ground State of the Electron Gas by a Stochastic Method”. In: Phys. Rev. Lett. 45.7 (1980), pp. 566–569. issn: 0031-9007. doi: 10.1103/PhysRevLett.45.566. url: https://link.aps.org/doi/10.1103/PhysRevLett.45.566.

[45]

John P. Perdew, Kieron Burke, and Matthias Ernzerhof. “Generalized Gradient Approximation Made Simple”. In: Phys. Rev. Lett. 77.18 (1996), pp. 3865–3868. issn: 0031-9007. doi: 10.1103/PhysRevLett.77.3865. url: https://link.aps.org/doi/10.1103/PhysRevLett.77.3865.

[46]

Jochen Heyd, Gustavo E. Scuseria, and Matthias Ernzerhof. “Hybrid functionals based on a screened Coulomb potential”. In: J. Chem. Phys. 118.18 (2003), pp. 8207–8215. issn: 0021-9606. doi: 10.1063/1.1564060. url: http://aip.scitation.org/doi/10.1063/1.1564060.

[47]

D. M. Ceperley. “Understanding Electronic Wave Functions”. In: Electron Correlations and Materials Properties. Ed. by A. Gonis, N. Kioussis, and M. Ciftan. Springer US, 1999, pp. 429–438.

[48]

Paul Adrien Maurice Dirac and Ralph Howard Fowler. “Quantum mechanics of many-electron systems”. In: Proceedings of the Royal Society of London. Series A, Containing Papers of a Mathematical and Physical Character 123.792 (1929), pp. 714–733. doi: 10.1098/rspa.1929.0094. eprint: https://royalsocietypublishing.org/doi/pdf/10.1098/rspa.1929.0094. url: https://royalsocietypublishing.org/doi/abs/10.1098/rspa.1929.0094.

[49]

E. Wigner. “On the Interaction of Electrons in Metals”. In: Phys. Rev. 46.11 (1934), pp. 1002–1011. issn: 0031-899X. doi: 10.1103/PhysRev.46.1002. url: https://link.aps.org/doi/10.1103/PhysRev.46.1002.

[50]

A. Bijl. “The lowest wave function of the symmetrical many particles system”. In: Physica 7.9 (1940), pp. 869–886. issn: 00318914. doi: 10.1016/0031-8914(40)90166-5. url: https://linkinghub.elsevier.com/retrieve/pii/0031891440901665.

[51]

Giuseppe Carleo et al. “Unitary Dynamics of Strongly Interacting Bose Gases with the Time-Dependent Variational Monte Carlo Method in Continuous Space”. In: Phys. Rev. X 7.3 (2017), p. 031026. issn: 2160-3308. doi: 10.1103/PhysRevX.7.031026. url: http://link.aps.org/doi/10.1103/PhysRevX.7.031026.

[52]

R.B. Dingle. “LI. The zero-point energy of a system of particles”. In: London, Edinburgh, Dublin Philos. Mag. J. Sci. 40.304 (1949), pp. 573–578. issn: 1941-5982. doi: 10.1080/14786444908521743. url: http://www.tandfonline.com/doi/abs/10.1080/14786444908521743.

[53]

Robert Jastrow. “Many-Body Problem with Strong Forces”. In: Phys. Rev. 98.5 (1955), pp. 1479–1484. issn: 0031-899X. doi: 10.1103/PhysRev.98.1479. url: https://link.aps.org/doi/10.1103/PhysRev.98.1479.

[54]

R. P. Feynman and Michael Cohen. “Energy Spectrum of the Excitations in Liquid Helium”. In: Phys. Rev. 102.5 (1956), pp. 1189–1204. issn: 0031-899X. doi: 10.1103/PhysRev.102.1189. url: https://link.aps.org/doi/10.1103/PhysRev.102.1189.

[55]

R. P. Feynman. “Atomic Theory of the Two-Fluid Model of Liquid Helium”. In: Phys. Rev. 94.2 (1954), pp. 262–277. issn: 0031-899X. doi: 10.1103/PhysRev.94.262. url: https://link.aps.org/doi/10.1103/PhysRev.94.262.

[56]

Fa Yueh Wu and Eugene Feenberg. “Ground State of Liquid Helium (Mass 4)”. In: Phys. Rev. 122.3 (1961), pp. 739–742. issn: 0031-899X. doi: 10.1103/PhysRev.122.739. url: https://link.aps.org/doi/10.1103/PhysRev.122.739.

[57]

T Gaskell. “The Collective Treatment of a Fermi Gas: II”. In: Proc. Phys. Soc. 77.6 (1961), pp. 1182–1192. issn: 0370-1328. doi: 10.1088/0370-1328/77/6/312. url: http://stacks.iop.org/0370-1328/72/i=5/a=301?key=crossref.a03e8e8c89696490a53671d5515ea8dahttp://stacks.iop.org/0370-1328/77/i=6/a=312?key=crossref.a131ff4ffa841443ea0142ac2f764057.

[58]

Michele Taddei et al. “Iterative backflow renormalization procedure for many-body ground-state wave functions of strongly interacting normal Fermi liquids”. In: Phys. Rev. B 91 (11 2015), p. 115106. doi: 10.1103/PhysRevB.91.115106. url: https://link.aps.org/doi/10.1103/PhysRevB.91.115106.

[59]

Markus Holzmann et al. “Renormalization factor and effective mass of the two-dimensional electron gas”. In: Phys. Rev. B 79.4 (2009), p. 041308. issn: 1098-0121. doi: 10.1103/PhysRevB.79.041308. arXiv: 0810.2450. url: https://link.aps.org/doi/10.1103/PhysRevB.79.041308.

[60]

Markus Holzmann et al. “Theory of finite size effects for electronic quantum Monte Carlo calculations of liquids and solids”. In: Phys. Rev. B 94.3 (2016), p. 035126. issn: 2469-9950. doi: 10.1103/PhysRevB.94.035126. arXiv: 1603.03957. url: http://arxiv.org/abs/1603.03957https://link.aps.org/doi/10.1103/PhysRevB.94.035126.

[61]

Paola Gori-Giorgi, Francesco Sacchetti, and Giovanni Bachelet. “Analytic static structure factors and pair-correlation functions for the unpolarized homogeneous electron gas”. In: Phys. Rev. B 61.11 (2000), pp. 7353–7363. doi: 10.1063/1.1390176.

[62]

Simone Chiesa et al. “Finite-Size Error in Many-Body Simulations with Long-Range Interactions”. In: Phys. Rev. Lett. 97.7 (2006), p. 076404. issn: 0031-9007. doi: 10.1103/PhysRevLett.97.076404. url: https://link.aps.org/doi/10.1103/PhysRevLett.97.076404.

[63]

N. D. Drummond et al. “Finite-size errors in continuum quantum Monte Carlo calculations”. In: Phys. Rev. B 78.12 (2008), p. 125106. issn: 1098-0121. doi: 10.1103/PhysRevB.78.125106. arXiv: 0806.0957. url: https://link.aps.org/doi/10.1103/PhysRevB.78.125106.

[64]

David Bohm and David Pines. “A Collective Description of Electron Interactions: III. Coulomb Interactions in a Degenerate Electron Gas”. In: Phys. Rev. 92.3 (1953), pp. 609–625. issn: 0031-899X. doi: 10.1103/PhysRev.92.609. url: https://link.aps.org/doi/10.1103/PhysRev.92.609.

[65]

Simone Chiesa et al. “Random phase approximation and the finite size errors in many body simulations”. In: AIP Conf. Proc. Vol. 918. 2007. AIP, 2007, pp. 284–288. isbn: 0735404283. doi: 10.1063/1.2751996. url: http://aip.scitation.org/doi/abs/10.1063/1.2751996.

[66]

Paola Gori-Giorgi, Saverio Moroni, and Giovanni B. Bachelet. “Pair-distribution functions of the two-dimensional electron gas”. In: Phys. Rev. B 70.11 (2004), pp. 1–12. issn: 01631829. doi: 10.1103/PhysRevB.70.115102.

[67]

Raymond C. Clay et al. “Benchmarking density functionals for hydrogen-helium mixtures with quantum Monte Carlo: Energetics, pressures, and forces”. In: Phys. Rev. B 93.3 (2016), p. 035121. issn: 2469-9950. doi: 10.1103/PhysRevB.93.035121. url: https://link.aps.org/doi/10.1103/PhysRevB.93.035121.

[68]

Yubo Yang et al. “Electronic band gaps from quantum Monte Carlo methods”. In: Phys. Rev. B 101.8 (2020), p. 085115. issn: 2469-9950. doi: 10.1103/PhysRevB.101.085115. arXiv: 1910.07531. url: http://arxiv.org/abs/1910.07531https://link.aps.org/doi/10.1103/PhysRevB.101.085115.

[69]

W.L.McMillan. “Ground State of Liquid He4”. In: Phys. Rev. 138.2 (1965).

[70]

W. R. Magro and D. M. Ceperley. “Ground-State Properties of the Two-Dimensional Bose Coulomb Liquid”. In: Phys. Rev. Lett. 73.6 (1994), pp. 826–829. issn: 0031-9007. doi: 10.1103/PhysRevLett.73.826. url: https://link.aps.org/doi/10.1103/PhysRevLett.73.826.

[71]

Yubo Yang et al. “Quantum Monte Carlo Compton profiles of solid and liquid lithium”. In: Phys. Rev. B 101.16 (2020), p. 165125. issn: 2469-9950. doi: 10.1103/PhysRevB.101.165125. url: https://link.aps.org/doi/10.1103/PhysRevB.101.165125.

[72]

C Lin, F H Zong, and D M Ceperley. “Twist-averaged boundary conditions in continuum quantum Monte Carlo algorithms”. In: Phys. Rev. E 64.1 (2001), p. 016702. issn: 1063-651X. doi: 10.1103/PhysRevE.64.016702. arXiv: 0101339 [cond-mat]. url: http://journals.aps.org/pre/abstract/10.1103/PhysRevE.64.016702https://link.aps.org/doi/10.1103/PhysRevE.64.016702.

[73]

S. M. Wahl et al. “Comparing Jupiter interior structure models to Juno gravity measurements and the role of a dilute core”. In: Geophys. Res. Lett. 44.10 (2017), pp. 4649–4659. issn: 19448007. doi: 10.1002/2017GL073160. arXiv: 1707.01997.

[74]

Jeffrey M. McMahon and David M. Ceperley. “Ground-State Structures of Atomic Metallic Hydrogen”. In: Phys. Rev. Lett. 106 (16 2011), p. 165302. doi: 10.1103/PhysRevLett.106.165302. url: https://link.aps.org/doi/10.1103/PhysRevLett.106.165302.

[75]

Jeffrey M. McMahon and David M. Ceperley. “Erratum: High-temperature superconductivity in atomic metallic hydrogen [Phys. Rev. B 84 , 144515 (2011)]”. In: Phys. Rev. B 85.21 (2012), p. 219902. issn: 1098-0121. doi: 10.1103/PhysRevB.85.219902. arXiv: 0504077 [arXiv:cond-mat]. url: https://link.aps.org/doi/10.1103/PhysRevB.85.219902.

[76]

Mario Motta et al. “Towards the Solution of the Many-Electron Problem in Real Materials: Equation of State of the Hydrogen Chain with State-of-the-Art Many-Body Methods”. In: Phys. Rev. X 7.3 (2017), p. 031059. issn: 2160-3308. doi: 10.1103/PhysRevX.7.031059. arXiv: 1705.01608. url: https://link.aps.org/doi/10.1103/PhysRevX.7.031059.

[77]

Jeffrey M. McMahon et al. “The properties of hydrogen and helium under extreme conditions”. In: Rev. Mod. Phys. 84 (2012), pp. 1607–1653. doi: 10.1103/RevModPhys.84.1607. url: https://link.aps.org/doi/10.1103/RevModPhys.84.1607.

[78]

Ranga P. Dias and Isaac F. Silvera. “Observation of the Wigner-Huntington transition to metallic hydrogen”. In: Science 355.6326 (2017), pp. 715–718. issn: 0036-8075. doi: 10.1126/science.aal1579. arXiv: 1702.04246. url: https://www.sciencemag.org/lookup/doi/10.1126/science.aal1579.

[79]

Chang-Sheng Zha et al. “Raman measurements of phase transitions in dense solid hydrogen and deuterium to 325 GPa.” In: Proc. Natl. Acad. Sci. U. S. A. 69.7 (2014), pp. 1129–1132. issn: 1091-6490. doi: 10.1073/pnas.1402737111. url: http://www.ncbi.nlm.nih.gov/pubmed/24639543.

[80]

Yuichi Akahama et al. “Evidence from x-ray diffraction of orientational ordering in phase III of solid hydrogen at pressures up to 183 GPa”. In: Phys. Rev. B 82.6 (2010), p. 060101. issn: 1098-0121. doi: 10.1103/PhysRevB.82.060101. url: https://link.aps.org/doi/10.1103/PhysRevB.82.060101.

[81]

Cheng Ji et al. “Ultrahigh-pressure isostructural electronic transitions in hydrogen”. In: Nature 573.7775 (2019), pp. 558–562. issn: 0028-0836. doi: 10.1038/s41586-019-1565-9. url: http://dx.doi.org/10.1038/s41586-019-1565-9http://www.nature.com/articles/s41586-019-1565-9.

[82]

M. D. Knudson et al. “Principal Hugoniot, reverberating wave, and mechanical reshock measurements of liquid deuterium to 400 GPa using plate impact techniques”. In: Phys. Rev. B 69.14 (2004), p. 144209. issn: 1098-0121. doi: 10.1103/PhysRevB.69.144209. url: https://link.aps.org/doi/10.1103/PhysRevB.69.144209.

[83]

M. D. Knudson and M. P. Desjarlais. “High-Precision Shock Wave Measurements of Deuterium: Evaluation of Exchange-Correlation Functionals at the Molecular-to-Atomic Transition”. In: Phys. Rev. Lett. 118.3 (2017), p. 035501. issn: 0031-9007. doi: 10.1103/PhysRevLett.118.035501. url: https://link.aps.org/doi/10.1103/PhysRevLett.118.035501.

[84]

Chris J. Pickard and Richard J. Needs. “Structure of phase III of solid hydrogen”. In: Nat. Phys. 3.7 (2007), pp. 473–476. issn: 1745-2473. doi: 10.1038/nphys625. url: http://www.nature.com/nphys/journal/v3/n7/abs/nphys625.htmlhttp://www.nature.com/articles/nphys625.

[85]

Bartomeu Monserrat et al. “Hexagonal structure of phase III of solid hydrogen”. In: Phys. Rev. B 94.13 (2016), p. 134101. issn: 2469-9950. doi: 10.1103/PhysRevB.94.134101. arXiv: 1609.07486. url: http://link.aps.org/doi/10.1103/PhysRevB.94.134101https://link.aps.org/doi/10.1103/PhysRevB.94.134101.

[86]

Miguel A. Morales et al. “Nuclear quantum effects and nonlocal exchange-correlation functionals applied to liquid hydrogen at high pressure”. In: Phys. Rev. Lett. 110.6 (2013), pp. 1–6. issn: 00319007. doi: 10.1103/PhysRevLett.110.065702. arXiv: arXiv:1303.6671v1.

[87]

Hector E. Lorenzana, Isaac F. Silvera, and Kenneth A. Goettel. “Orientational phase transitions in hydrogen at megabar pressures”. In: Phys. Rev. Lett. 64.16 (1990), pp. 1939–1942. issn: 00319007. doi: 10.1103/PhysRevLett.64.1939.

[88]

Hector E. Lorenzana, Isaac F. Silvera, and Kenneth A. Goettel. “Evidence for a structural phase transition in solid hydrogen at megabar pressures”. In: Phys. Rev. Lett. 63.19 (1989), pp. 2080–2083. issn: 0031-9007. doi: 10.1103/PhysRevLett.63.2080. url: https://link.aps.org/doi/10.1103/PhysRevLett.63.2080.

[89]

Isaac F. Silvera and Rinke J. Wijngaarden. “New Low-Temperature Phase of Molecular Deuterium at Ultrahigh Pressure”. In: Phys. Rev. Lett. 47.1 (1981), pp. 39–42. issn: 0031-9007. doi: 10.1103/PhysRevLett.47.39. url: https://link.aps.org/doi/10.1103/PhysRevLett.47.39.

[90]

Akobuije Chijioke and Isaac F. Silvera. “Megabar-Pressure Infrared Study of Hydrogen Deuteride”. In: Phys. Rev. Lett. 97.25 (2006), p. 255701. issn: 0031-9007. doi: 10.1103/PhysRevLett.97.255701. url: https://link.aps.org/doi/10.1103/PhysRevLett.97.255701.

[91]

R. J. Hemley and H. K. Mao. “Phase Transition in Solid Molecular Hydrogen at Ultrahigh Pressures”. In: Phys. Rev. Lett. 61.7 (1988), pp. 857–860. issn: 0031-9007. doi: 10.1103/PhysRevLett.61.857. url: https://link.aps.org/doi/10.1103/PhysRevLett.61.857.

[92]

Lijing Cui et al. “Megabar pressure triple point in solid deuterium”. In: Phys. Rev. Lett. 72.19 (1994), pp. 3048–3051. issn: 0031-9007. doi: 10.1103/PhysRevLett.72.3048. url: https://link.aps.org/doi/10.1103/PhysRevLett.72.3048.

[93]

Alexander F. Goncharov et al. “New High-Pressure Excitations in Parahydrogen”. In: Phys. Rev. Lett. 80.1 (1998), pp. 101–104. issn: 0031-9007. doi: 10.1103/PhysRevLett.80.101. url: https://link.aps.org/doi/10.1103/PhysRevLett.80.101.

[94]

Ranga P. Dias, Ori Noked, and Isaac F. Silvera. “Quantum phase transition in solid hydrogen at high pressure”. In: Phys. Rev. B 100.18 (2019), p. 184112. issn: 2469-9950. doi: 10.1103/PhysRevB.100.184112. url: https://link.aps.org/doi/10.1103/PhysRevB.100.184112.

[95]

Chang-sheng Zha et al. “High-Pressure Measurements of Hydrogen Phase IV Using Synchrotron Infrared Spectroscopy”. In: Phys. Rev. Lett. 110.21 (2013), p. 217402. issn: 0031-9007. doi: 10.1103/PhysRevLett.110.217402. url: https://link.aps.org/doi/10.1103/PhysRevLett.110.217402.

[96]

M. I. Eremets et al. “Semimetallic molecular hydrogen at pressure above 350 GPa”. In: Nat. Phys. 15.12 (2019), pp. 1246–1249. issn: 17452481. doi: 10.1038/s41567-019-0646-x. arXiv: 1708.05217. url: http://dx.doi.org/10.1038/s41567-019-0646-x.

[97]

Isaac F. Silvera, Mohamed Zaghoo, and Ashkan Salamat. “Reply to Comment on Evidence of a first-order phase transition to metallic hydrogen’ ”. In: Phys. Rev. B 96.23 (2017), p. 237102. issn: 2469-9950. doi: 10.1103/PhysRevB.96.237102. url: https://link.aps.org/doi/10.1103/PhysRevB.96.237102.

[98]

R. M. Hazen et al. “Single-crystal x-ray diffraction ofn-H2 at high pressure”. In: Phys. Rev. B 36.7 (1987), pp. 3944–3947. issn: 0163-1829. doi: 10.1103/PhysRevB.36.3944. url: https://link.aps.org/doi/10.1103/PhysRevB.36.3944.

[99]

H. K. MAao et al. “Synchrotron X-ray Diffraction Measurements of Single-Crystal Hydrogen to 26.5 Gigapascals”. In: Science 239.4844 (1988), pp. 1131–1134. issn: 0036-8075. doi: 10.1126/science.239.4844.1131. url: https://www.sciencemag.org/lookup/doi/10.1126/science.239.4844.1131.

[100]

P. Loubeyre et al. “X-ray diffraction and equation of state of hydrogen at megabar pressures”. In: Nature 383.6602 (1996), pp. 702–704. issn: 0028-0836. doi: 10.1038/383702a0. url: https://www.nature.com/articles/383702a0.pdfhttp://www.nature.com/articles/383702a0.

[101]

H. Kawamura et al. “X-ray powder diffraction from solid deuterium”. In: J. Phys. Condens. Matter 14.44 (2002), pp. 10407–10410. issn: 0953-8984. doi: 10.1088/0953-8984/14/44/301. url: http://stacks.iop.org/0953-8984/14/i=44/a=301?key=crossref.f7ecb62bd6245af471fba020c86831b0.

[102]

Igor Goncharenko and Paul Loubeyre. “Neutron and X-ray diffraction study of the broken symmetry phase transition in solid deuterium”. In: Nature 435.7046 (2005), pp. 1206–1209. issn: 0028-0836. doi: 10.1038/nature03699. url: http://www.nature.com/articles/nature03699.

[103]

M. I. Eremets and I. A. Trojan. “Evidence of maximum in the melting curve of hydrogen at megabar pressures”. In: JETP Lett. 89.4 (2009), pp. 174–179. issn: 0021-3640. doi: 10.1134/S0021364009040031. url: http://link.springer.com/10.1134/S0021364009040031.

[104]

Ross T. Howie et al. “Mixed Molecular and Atomic Phase of Dense Hydrogen”. In: Phys. Rev. Lett. 108.12 (2012), p. 125501. issn: 0031-9007. doi: 10.1103/PhysRevLett.108.125501. url: https://link.aps.org/doi/10.1103/PhysRevLett.108.125501.

[105]

Susanne Raynor. “Novel ab initio self‐consistent‐field approach to molecular solids under pressure. II. Solid H2 under high pressure”. In: J. Chem. Phys. 87.5 (1987), pp. 2795–2799. issn: 0021-9606. doi: 10.1063/1.453067. url: http://aip.scitation.org/doi/10.1063/1.453067.

[106]

T. W. Barbee, Marvin L. Cohen, and José Luís Martins. “Theory of high-pressure phases of hydrogen”. In: Phys. Rev. Lett. 62.10 (1989), pp. 1150–1153. issn: 0031-9007. doi: 10.1103/PhysRevLett.62.1150. url: https://link.aps.org/doi/10.1103/PhysRevLett.62.1150.

[107]

Efthimios Kaxiras, Jeremy Broughton, and R. J. Hemley. “Onset of metallization and related transitions in solid hydrogen”. In: Phys. Rev. Lett. 67.9 (1991), pp. 1138–1141. issn: 0031-9007. doi: 10.1103/PhysRevLett.67.1138. url: https://link.aps.org/doi/10.1103/PhysRevLett.67.1138.

[108]

E. Kaxiras and J. Broughton. “Energetics of Ordered Structures in Molecular Hydrogen”. In: Europhys. Lett. 17.2 (1992), pp. 151–155. issn: 0295-5075. doi: 10.1209/0295-5075/17/2/012. url: https://iopscience.iop.org/article/10.1209/0295-5075/17/2/012.

[109]

Hitose Nagara and Tuto Nakamura. “Stable phases of solid hydrogen at megabar pressures and at zero temperature”. In: Phys. Rev. Lett. 68.16 (1992), pp. 2468–2471. issn: 0031-9007. doi: 10.1103/PhysRevLett.68.2468. url: https://link.aps.org/doi/10.1103/PhysRevLett.68.2468.

[110]

B. Edwards, N. W. Ashcroft, and T. Lenosky. “Layering transitions and the structure of dense hydrogen”. In: Europhys. Lett. 34.7 (1996), pp. 519–524. issn: 0295-5075. doi: 10.1209/epl/i1996-00489-5. url: https://iopscience.iop.org/article/10.1209/epl/i1996-00489-5.

[111]

Kurt A. Johnson and N. W. Ashcroft. “Structure and bandgap closure in dense hydrogen”. In: Nature 403.6770 (2000), pp. 632–635. issn: 0028-0836. doi: 10.1038/35001024. url: http://www.ncbi.nlm.nih.gov/pubmed/10688193http://www.nature.com/articles/35001024.

[112]

T. Cui, Y. Ma, and G. Zou. “Crystal structure of solid molecular hydrogen under high pressures”. In: J. Phys. Condens. Matter 14.44 (2002), pp. 10901–10906. issn: 0953-8984. doi: 10.1088/0953-8984/14/44/398. url: https://iopscience.iop.org/article/10.1088/0953-8984/14/44/398.

[113]

Sam Azadi and Thomas D. Kühne. “Unconventional phase III of high-pressure solid hydrogen”. In: Phys. Rev. B 100.15 (2019), p. 155103. issn: 2469-9950. doi: 10.1103/PhysRevB.100.155103. arXiv: 1906.10854. url: https://link.aps.org/doi/10.1103/PhysRevB.100.155103.

[114]

Ke Liao et al. “A comparative study using state-of-the-art electronic structure theories on solid hydrogen phases under high pressures”. In: npj Comput. Mater. 5.1 (2019), p. 110. issn: 2057-3960. doi: 10.1038/s41524-019-0243-7. url: http://dx.doi.org/10.1038/s41524-019-0243-7http://www.nature.com/articles/s41524-019-0243-7.

[115]

Jonathan H. Lloyd-Williams and Bartomeu Monserrat. “Lattice dynamics and electron-phonon coupling calculations using nondiagonal supercells”. In: Phys. Rev. B 92.18 (2015), p. 184301. issn: 1098-0121. doi: 10.1103/PhysRevB.92.184301. arXiv: 1510.04418. url: https://link.aps.org/doi/10.1103/PhysRevB.92.184301.

[116]

David M. Ceperley and B. J. Alder. “Ground state of solid hydrogen at high pressures”. In: Phys. Rev. B 36.4 (1987), pp. 2092–2106. issn: 0163-1829. doi: 10.1103/PhysRevB.36.2092. url: http://link.aps.org/doi/10.1103/PhysRevB.36.2092.

[117]

V Natoli, R M Martin, and D M Ceperley. “Crystal Structure of Atomic Hydrogen”. In: Phys. Rev. Lett. 70.13 (1993), p. 1952. issn: 0031-9007. doi: 10.1103/PhysRevLett.70.1952. url: file:///home/jarajar/Desktop/papers/General{\_}physics/MetallicHydrogen.pdf.

[118]

L. Reatto. “Structure of the ground-state wave function of quantum fluids and ”exact” numerical methods”. In: Phys. Rev. B 26.1 (1982), pp. 130–137. issn: 0163-1829. doi: 10.1103/PhysRevB.26.130. url: https://link.aps.org/doi/10.1103/PhysRevB.26.130.

[119]

N. D. Drummond et al. Data for ”Quantum Monte Carlo study of the phase diagram of solid molecular hydrogen at extreme pressures”. July 8, 2015. url: https://www.repository.cam.ac.uk/handle/1810/248864.

[120]

Hendra Kwee, Shiwei Zhang, and Henry Krakauer. “Finite-Size Correction in Many-Body Electronic Structure Calculations”. In: Phys. Rev. Lett. 100.12 (2008), p. 126404. issn: 0031-9007. doi: 10.1103/PhysRevLett.100.126404. url: https://link.aps.org/doi/10.1103/PhysRevLett.100.126404.

[121]

M. Born and R. Oppenheimer. “Zur Quantentheorie der Molekeln”. In: Ann. Phys. 389.20 (1927), pp. 457–484. issn: 00033804. doi: 10.1146/annurev.physchem.55.091602.094335. url: http://doi.wiley.com/10.1002/andp.19273892002.

[122]

Sharon Hammes-Schiffer and John C. Tully. “Vibrationally Enhanced Proton Transfer”. In: J. Phys. Chem. 99.16 (1995), pp. 5793–5797. issn: 0022-3654. doi: 10.1021/j100016a011. url: https://pubs.acs.org/doi/abs/10.1021/j100016a011.

[123]

David S. Sholl and John C. Tully. “A generalized surface hopping method”. In: J. Chem. Phys. 109.18 (1998), pp. 7702–7710. issn: 0021-9606. doi: 10.1063/1.477416. url: http://aip.scitation.org/doi/10.1063/1.477416.

[124]

Sharon Hammes-Schiffer. “Proton-Coupled Electron Transfer: Moving Together and Charging Forward”. In: J. Am. Chem. Soc. 137.28 (2015), pp. 8860–8871. issn: 0002-7863. doi: 10.1021/jacs.5b04087. url: https://pubs.acs.org/doi/10.1021/jacs.5b04087.

[125]

W. Kolos and L. Wolniewicz. “Improved Theoretical Ground‐State Energy of the Hydrogen Molecule”. In: J. Chem. Phys. 49.1 (1968), pp. 404–410. issn: 0021-9606. doi: 10.1063/1.1669836. url: http://aip.scitation.org/doi/10.1063/1.1669836.

[126]

L. Wolniewicz. “The X 1Σ+g state vibration-rotational energies of the H2, HD, and D2 molecules”. In: The Journal of Chemical Physics 78.10 (1983), p. 6173. issn: 00219606. doi: 10.1063/1.444580. url: http://link.aip.org/link/?JCP/78/6173/1{\&}Agg=doi.

[127]

Yubo Yang et al. “How large are nonadiabatic effects in atomic and diatomic systems?” In: Journal of Chemical Physics 143.12 (2015), pp. 0–9. issn: 00219606. doi: 10.1063/1.4931667. arXiv: 1507.05959. url: http://dx.doi.org/10.1063/1.4931667.

[128]

A. Szabo and N. S. Ostlund. Modern quantum chemistry : introduction to advanced electronic structure theory. Minelola, N. Y.: Dover Publications, 1996.

[129]

W. Heitler and F. London. “Wechselwirkung neutraler Atome und homopolare Bindung nach der Quantenmechanik”. In: Zeitschrift fur Phys. 44.6-7 (1927), pp. 455–472. issn: 1434-6001. doi: 10.1007/BF01397394. url: http://link.springer.com/10.1007/BF01397394.