212
Thesis for the degree of Doctor of Philosophy in Natural Science, specializing in Chemistry Quantum dynamical effects in complex chemical systems S. Karl-Mikael Svensson Department of Chemistry and Molecular Biology Gothenburg, Sweden, 2020 i

Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Thesis for the degree of Doctor of Philosophy inNatural Science, specializing in Chemistry

Quantum dynamical effects incomplex chemical systems

S. Karl-Mikael Svensson

Department of Chemistry and Molecular BiologyGothenburg, Sweden, 2020

i

Page 2: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

ISBN: 978-91-7833-922-8 (print)ISBN: 978-91-7833-923-5 (PDF)Available online at: http://hdl.handle.net/2077/64131

Thesis © S. Karl-Mikael Svensson, 2020Paper I © American Chemical Society, 2015Paper II Creative Commons Attribution license∗

Paper III © S. Karl-Mikael Svensson, Jens Aage Poulsen, GunnarNyman, 2020Paper IV © S. Karl-Mikael Svensson, Jens Aage Poulsen, GunnarNyman, 2020

Front cover: Artistic representation of an imaginary time pathintegral open polymer attached to a dividing surface on the top of apotential energy barrier.

Typeset with LATEX.

Printed by Stema Specialtryck ABBoras, Sweden, 2020

∗https://creativecommons.org/licenses/by/4.0/

ii

To my grandparents.

iii

SVANENMÄRKET

Trycksak3041 0234

Page 3: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

ISBN: 978-91-7833-922-8 (print)ISBN: 978-91-7833-923-5 (PDF)Available online at: http://hdl.handle.net/2077/64131

Thesis © S. Karl-Mikael Svensson, 2020Paper I © American Chemical Society, 2015Paper II Creative Commons Attribution license∗

Paper III © S. Karl-Mikael Svensson, Jens Aage Poulsen, GunnarNyman, 2020Paper IV © S. Karl-Mikael Svensson, Jens Aage Poulsen, GunnarNyman, 2020

Front cover: Artistic representation of an imaginary time pathintegral open polymer attached to a dividing surface on the top of apotential energy barrier.

Typeset with LATEX.

Printed by Stema Specialtryck ABBoras, Sweden, 2020

∗https://creativecommons.org/licenses/by/4.0/

ii

To my grandparents.

iii

Page 4: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Abstract

When using mathematical models to computationally investigate achemical system it is important that the methods used are accurateenough to account for the relevant properties of the system and atthe same time simple enough to be computationally affordable. Thisthesis presents research that so far has resulted in three publishedpapers and one unpublished manuscript. It concerns the applicationand development of computational methods for chemistry, with someextra emphasis on the calculation of reaction rate constants.

In astrochemistry radiative association is a relevant reactionmechanism for the formation of molecules. The rate constants forsuch reactions are often difficult to obtain though experiments. In thefirst published paper of the thesis a rate constant for the formationof the hydroxyl radical, through the radiative association of atomicoxygen and hydrogen, is presented. This rate constant was calculatedby a combination of different methods and should be an improvementover previously available rate constants.

In the the second published paper of this thesis two kinds ofbasis functions, for use with a variational principle for the dynamicsof quantum distributions in phase space, i.e. Wigner functions, ispresented. These are tested on model systems and found to havesome appealing properties.

The classical Wigner method is an approximate method of simu-lation, where an initial quantum distribution is propagated in timewith classical mechanics. In the third published paper of this thesisa new method of sampling the initial quantum distribution, withan imaginary time Feynman path integral, is derived and tested onmodel systems. In the unpublished manuscript, this new methodis applied to reaction rate constants and tested on two model sys-

v

Page 5: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Abstract

When using mathematical models to computationally investigate achemical system it is important that the methods used are accurateenough to account for the relevant properties of the system and atthe same time simple enough to be computationally affordable. Thisthesis presents research that so far has resulted in three publishedpapers and one unpublished manuscript. It concerns the applicationand development of computational methods for chemistry, with someextra emphasis on the calculation of reaction rate constants.

In astrochemistry radiative association is a relevant reactionmechanism for the formation of molecules. The rate constants forsuch reactions are often difficult to obtain though experiments. In thefirst published paper of the thesis a rate constant for the formationof the hydroxyl radical, through the radiative association of atomicoxygen and hydrogen, is presented. This rate constant was calculatedby a combination of different methods and should be an improvementover previously available rate constants.

In the the second published paper of this thesis two kinds ofbasis functions, for use with a variational principle for the dynamicsof quantum distributions in phase space, i.e. Wigner functions, ispresented. These are tested on model systems and found to havesome appealing properties.

The classical Wigner method is an approximate method of simu-lation, where an initial quantum distribution is propagated in timewith classical mechanics. In the third published paper of this thesisa new method of sampling the initial quantum distribution, withan imaginary time Feynman path integral, is derived and tested onmodel systems. In the unpublished manuscript, this new methodis applied to reaction rate constants and tested on two model sys-

v

Page 6: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Abstract

tems. The new sampling method shows some promise for futureapplications.

vi

Preface

You have guessed right; I have lately been sodeeply engaged in one occupation that I havenot allowed myself sufficient rest, as you see:but I hope, I sincerely hope, that all theseemployments are now at an end, and that I amat length free.

Victor Frankenstein (Mary Shelley)1

This thesis is a compilation thesis consisting of two main parts.To start from the back, the second part is a collection of papers thathave been coauthored by the author of the thesis and represent theresearch that the thesis is based upon.

The first part of this thesis is the frame, which is an introductionto and discussion of the papers in the second part. The frame isitself divided into three parts. First is an introduction where thegeneral subject of the thesis is presented and put on the scientificmap, with the aim of being accessible to a broader audience thanthe rest of this book. Second is a chapter with the theory on whichthe work is based. Third, and last in the frame, there is a chapterdescribing and discussing the new developments that has come outof the research. The contents of the third chapter overlaps with thecontent of the papers, but the aim is for it to be more pedagogicalthan the papers themselves are.

vii

Page 7: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Abstract

tems. The new sampling method shows some promise for futureapplications.

vi

Preface

You have guessed right; I have lately been sodeeply engaged in one occupation that I havenot allowed myself sufficient rest, as you see:but I hope, I sincerely hope, that all theseemployments are now at an end, and that I amat length free.

Victor Frankenstein (Mary Shelley)1

This thesis is a compilation thesis consisting of two main parts.To start from the back, the second part is a collection of papers thathave been coauthored by the author of the thesis and represent theresearch that the thesis is based upon.

The first part of this thesis is the frame, which is an introductionto and discussion of the papers in the second part. The frame isitself divided into three parts. First is an introduction where thegeneral subject of the thesis is presented and put on the scientificmap, with the aim of being accessible to a broader audience thanthe rest of this book. Second is a chapter with the theory on whichthe work is based. Third, and last in the frame, there is a chapterdescribing and discussing the new developments that has come outof the research. The contents of the third chapter overlaps with thecontent of the papers, but the aim is for it to be more pedagogicalthan the papers themselves are.

vii

Page 8: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Contents

Abstract v

Preface vii

Contents viii

List of Figures xv

List of Tables xvii

List of Papers xixContributions from the Author . . . . . . . . . . . . . . . xx

Acknowledgments xxi

Frame 1

1 Introduction 3

2 Background 72.1 Wigner transform and the Wigner phase space . . . . 92.2 Feynman path integral formulation of quantum me-

chanics . . . . . . . . . . . . . . . . . . . . . . . . . . 182.3 Chemical kinetics and thermal rate constants . . . . . 382.4 Common methods of approximate quantum dynamics 42

3 Developments 49

viii

Contents

3.1 Formation of the Hydroxyl Radical by Radiative As-sociation . . . . . . . . . . . . . . . . . . . . . . . . . 50

3.2 Dynamics of Gaussian basis functions . . . . . . . . . 653.3 The open polymer classical Wigner method . . . . . . 763.4 Future outlook . . . . . . . . . . . . . . . . . . . . . 100

Bibliography 103

Papers 117

I Formation of the Hydroxyl Radical by RadiativeAssociation 119

II Dynamics of Gaussian Wigner functions derivedfrom a time-dependent variational principle 127

IIIClassical Wigner Model Based on a Feynman PathIntegral Open Polymer 157

IVCalculation of Reaction Rate Constants From aClassical Wigner Model Based on a Feynman PathIntegral Open Polymer 179

ix

Page 9: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Contents

Abstract v

Preface vii

Contents viii

List of Figures xv

List of Tables xvii

List of Papers xixContributions from the Author . . . . . . . . . . . . . . . xx

Acknowledgments xxi

Frame 1

1 Introduction 3

2 Background 72.1 Wigner transform and the Wigner phase space . . . . 92.2 Feynman path integral formulation of quantum me-

chanics . . . . . . . . . . . . . . . . . . . . . . . . . . 182.3 Chemical kinetics and thermal rate constants . . . . . 382.4 Common methods of approximate quantum dynamics 42

3 Developments 49

viii

Contents

3.1 Formation of the Hydroxyl Radical by Radiative As-sociation . . . . . . . . . . . . . . . . . . . . . . . . . 50

3.2 Dynamics of Gaussian basis functions . . . . . . . . . 653.3 The open polymer classical Wigner method . . . . . . 763.4 Future outlook . . . . . . . . . . . . . . . . . . . . . 100

Bibliography 103

Papers 117

I Formation of the Hydroxyl Radical by RadiativeAssociation 119

II Dynamics of Gaussian Wigner functions derivedfrom a time-dependent variational principle 127

IIIClassical Wigner Model Based on a Feynman PathIntegral Open Polymer 157

IVCalculation of Reaction Rate Constants From aClassical Wigner Model Based on a Feynman PathIntegral Open Polymer 179

ix

Page 10: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Nomenclature

Come, let us go down and confuse theirlanguage so they will not understand eachother.

Genesis 11:7

Abbreviations

C3SE Chalmers Centre for Computational Science and Engineering

CMD Centroid molecular dynamics

DVR Discrete variable representation

FK-LPI Feynman-Kleinert linearized path integral

KIDA Kinetic Database for Astrochemistry

LPI Linearized path integral

LSC-IVR Linearized semi-classical initial value representation

OPCW Open polymer classical Wigner

PIMC Path integral Monte Carlo

PIMD Path integral molecular dynamics

RPMD Ring polymer molecular dynamics

SC-IVR Semi-classical initial value representation

SNIC Swedish National Infrastructure for Computing

xi

Page 11: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Nomenclature

Come, let us go down and confuse theirlanguage so they will not understand eachother.

Genesis 11:7

Abbreviations

C3SE Chalmers Centre for Computational Science and Engineering

CMD Centroid molecular dynamics

DVR Discrete variable representation

FK-LPI Feynman-Kleinert linearized path integral

KIDA Kinetic Database for Astrochemistry

LPI Linearized path integral

LSC-IVR Linearized semi-classical initial value representation

OPCW Open polymer classical Wigner

PIMC Path integral Monte Carlo

PIMD Path integral molecular dynamics

RPMD Ring polymer molecular dynamics

SC-IVR Semi-classical initial value representation

SNIC Swedish National Infrastructure for Computing

xi

Page 12: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Contents

UDfA UMIST Database for Astrochemistry

WKB Wentzel–Kramers–Brillouin

Physical and mathematical constants

Reduced Planck’s constant 1.054571800× 10−34Js

kB Boltzmann’s constant 1.38064852× 10−23JK−1

i Imaginary unit√−1

Variables, operators, functions, and transforms

β1

kBT

α Vector of parameters

η Position vector

λ Momentum vector

mod Modulus operator

Ω Arbitrary physical quantity

p Momentum vector

x Position vector

δ (ζ) Dirac delta function

[f(t)]F (ω) Fourier transform of function f(t) as a function of angularfrequency ω.

|p〉 Momentum eigenket

|x〉 Position eigenket

|Ψ〉 Ket describing the state of the system described by the wave-function Ψ

O Big O

xii

Contents

Hn nth order Hermite polynomial Hn (x) = (−1)n ex2 dn

dxne−x2

Ω Arbitrary quantum mechanical operator

p Momentum operator

x Position operator

F (x− s) Probability flux operator1

2m−1 (δ (x− s) p+ pδ (x− s))

H Hamiltonian operator T + V

L Liouvillian operator

T Kinetic energy operator

V Potential energy operator

ΦW Parametrized Wigner function

Ψ Wavefunction

ρ Classical probability distribution function

θ Heaviside step function

TrΩ

Trace of operator Ω.

[Ω]W(x,p) Wigner transform of operator Ω

∗ Complex conjugate

T Transpose of vector or matrix

D Number of degrees of freedom in the system

E Energy

H Classical Hamiltonian function

kr Reaction rate constant

QR Canonical reactant partition function

xiii

Page 13: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Contents

UDfA UMIST Database for Astrochemistry

WKB Wentzel–Kramers–Brillouin

Physical and mathematical constants

Reduced Planck’s constant 1.054571800× 10−34Js

kB Boltzmann’s constant 1.38064852× 10−23JK−1

i Imaginary unit√−1

Variables, operators, functions, and transforms

β1

kBT

α Vector of parameters

η Position vector

λ Momentum vector

mod Modulus operator

Ω Arbitrary physical quantity

p Momentum vector

x Position vector

δ (ζ) Dirac delta function

[f(t)]F (ω) Fourier transform of function f(t) as a function of angularfrequency ω.

|p〉 Momentum eigenket

|x〉 Position eigenket

|Ψ〉 Ket describing the state of the system described by the wave-function Ψ

O Big O

xii

Contents

Hn nth order Hermite polynomial Hn (x) = (−1)n ex2 dn

dxne−x2

Ω Arbitrary quantum mechanical operator

p Momentum operator

x Position operator

F (x− s) Probability flux operator1

2m−1 (δ (x− s) p+ pδ (x− s))

H Hamiltonian operator T + V

L Liouvillian operator

T Kinetic energy operator

V Potential energy operator

ΦW Parametrized Wigner function

Ψ Wavefunction

ρ Classical probability distribution function

θ Heaviside step function

TrΩ

Trace of operator Ω.

[Ω]W(x,p) Wigner transform of operator Ω

∗ Complex conjugate

T Transpose of vector or matrix

D Number of degrees of freedom in the system

E Energy

H Classical Hamiltonian function

kr Reaction rate constant

QR Canonical reactant partition function

xiii

Page 14: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Contents

s A dividing surface

T Absolute temperature

t Time

V Potential energy

xr Reaction coordinate

Z Canonical partition function

1

(2π)D|Ψ〉 〈Ψ | Probability density operator

xiv

List of Figures

2.1 Illustration of an imaginary time path integral ring poly-mer compared to a classical particle. . . . . . . . . . . . 26

2.2 Illustration of collision cross section. . . . . . . . . . . . 41

3.1 Potential energy surfaces for some electronic states of thehydroxyl radical. . . . . . . . . . . . . . . . . . . . . . . 51

3.2 Electric dipole moment and transition dipole moment forsome electronic states of the hydroxyl radical. . . . . . . 53

3.3 Reaction cross section for the reaction OH(X2Π)∗ →OH(X2Π) + γ. . . . . . . . . . . . . . . . . . . . . . . . 59

3.4 Reaction cross section for the reaction OH(12Σ−) →OH(X2Π) + γ. . . . . . . . . . . . . . . . . . . . . . . . 60

3.5 Reaction rate constant for the reactions OH(X2Π)∗ →OH(X2Π) + γ and OH(12Σ−) → OH(X2Π) + γ. . . . . 61

3.6 Reaction cross sections for the reactions OH(X2Π)∗ →OH(X2Π) + γ, OH(12Σ−) → OH(X2Π) + γ, andOH(12Σ−/a4Σ−/b4Π) → OH+(X3Σ−) + e−. . . . . . . 62

3.7 The total reaction rate constants for the radiative associ-ation reaction O(3P) + H(2S) → OH(X2Π) + γ. . . . . . 63

3.8 Average position of a particle in a double well potentialas a function of time, using a thawed Gaussian function. 70

3.9 Tunneling period of a double well potential with differingbarrier frequency, but constant well depth, using thawedGaussian basis functions. . . . . . . . . . . . . . . . . . . 71

3.10 Average position of a particle in a double well potentialas a function of time, using frozen Gaussian basis functions. 72

3.11 Average position of a particle in a quartic potential as afunction of time, using frozen Gaussian basis functions. . 73

xv

Page 15: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Contents

s A dividing surface

T Absolute temperature

t Time

V Potential energy

xr Reaction coordinate

Z Canonical partition function

1

(2π)D|Ψ〉 〈Ψ | Probability density operator

xiv

List of Figures

2.1 Illustration of an imaginary time path integral ring poly-mer compared to a classical particle. . . . . . . . . . . . 26

2.2 Illustration of collision cross section. . . . . . . . . . . . 41

3.1 Potential energy surfaces for some electronic states of thehydroxyl radical. . . . . . . . . . . . . . . . . . . . . . . 51

3.2 Electric dipole moment and transition dipole moment forsome electronic states of the hydroxyl radical. . . . . . . 53

3.3 Reaction cross section for the reaction OH(X2Π)∗ →OH(X2Π) + γ. . . . . . . . . . . . . . . . . . . . . . . . 59

3.4 Reaction cross section for the reaction OH(12Σ−) →OH(X2Π) + γ. . . . . . . . . . . . . . . . . . . . . . . . 60

3.5 Reaction rate constant for the reactions OH(X2Π)∗ →OH(X2Π) + γ and OH(12Σ−) → OH(X2Π) + γ. . . . . 61

3.6 Reaction cross sections for the reactions OH(X2Π)∗ →OH(X2Π) + γ, OH(12Σ−) → OH(X2Π) + γ, andOH(12Σ−/a4Σ−/b4Π) → OH+(X3Σ−) + e−. . . . . . . 62

3.7 The total reaction rate constants for the radiative associ-ation reaction O(3P) + H(2S) → OH(X2Π) + γ. . . . . . 63

3.8 Average position of a particle in a double well potentialas a function of time, using a thawed Gaussian function. 70

3.9 Tunneling period of a double well potential with differingbarrier frequency, but constant well depth, using thawedGaussian basis functions. . . . . . . . . . . . . . . . . . . 71

3.10 Average position of a particle in a double well potentialas a function of time, using frozen Gaussian basis functions. 72

3.11 Average position of a particle in a quartic potential as afunction of time, using frozen Gaussian basis functions. . 73

xv

Page 16: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

List of Figures

3.12 Kubo transformed position autocorrelation function fora quartic potential at βω = 8, calculated with thawedGaussian basis functions. . . . . . . . . . . . . . . . . . . 74

3.13 Kubo transformed position autocorrelation function fora quartic potential at βω = 1, calculated with thawedGaussian basis functions. . . . . . . . . . . . . . . . . . . 75

3.14 Illustration of an imaginary time path integral open poly-mer compared to a classical particle. . . . . . . . . . . . 86

3.15 The real part of the position autocorrelation function fora quartic oscillator calculated at βω = 8. . . . . . . . . 91

3.16 The real part of the position autocorrelation function fora quartic oscillator calculated at βω = 1. . . . . . . . . 92

3.17 The real part of the position autocorrelation function fora double well potential calculated at βω = 8. . . . . . . 93

3.18 The real part of the position-squared autocorrelationfunction for a double well potential calculated at βω = 8. 94

3.19 The real part of the position-squared autocorrelationfunction for a quartic potential bilinearly coupled toa bath of 3 harmonic oscillators, calculated at βω =8. Results for the y-version of OPCW, with differentnumbers of beads. . . . . . . . . . . . . . . . . . . . . . . 95

3.20 The real part of the position-squared autocorrelationfunction for a quartic potential bilinearly coupled toa bath of 3 harmonic oscillators, calculated at βω =8. Results for the x-version of OPCW, with differentnumbers of beads. . . . . . . . . . . . . . . . . . . . . . . 96

3.21 The real part of the position-squared autocorrelationfunction for a quartic potential bilinearly coupled to abath of 9 harmonic oscillators, calculated at βω = 8. . . 97

xvi

List of Tables

3.1 Rate constant for the Eckart potential at different inversetemperatures. . . . . . . . . . . . . . . . . . . . . . . . . 93

xvii

Page 17: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

List of Figures

3.12 Kubo transformed position autocorrelation function fora quartic potential at βω = 8, calculated with thawedGaussian basis functions. . . . . . . . . . . . . . . . . . . 74

3.13 Kubo transformed position autocorrelation function fora quartic potential at βω = 1, calculated with thawedGaussian basis functions. . . . . . . . . . . . . . . . . . . 75

3.14 Illustration of an imaginary time path integral open poly-mer compared to a classical particle. . . . . . . . . . . . 86

3.15 The real part of the position autocorrelation function fora quartic oscillator calculated at βω = 8. . . . . . . . . 91

3.16 The real part of the position autocorrelation function fora quartic oscillator calculated at βω = 1. . . . . . . . . 92

3.17 The real part of the position autocorrelation function fora double well potential calculated at βω = 8. . . . . . . 93

3.18 The real part of the position-squared autocorrelationfunction for a double well potential calculated at βω = 8. 94

3.19 The real part of the position-squared autocorrelationfunction for a quartic potential bilinearly coupled toa bath of 3 harmonic oscillators, calculated at βω =8. Results for the y-version of OPCW, with differentnumbers of beads. . . . . . . . . . . . . . . . . . . . . . . 95

3.20 The real part of the position-squared autocorrelationfunction for a quartic potential bilinearly coupled toa bath of 3 harmonic oscillators, calculated at βω =8. Results for the x-version of OPCW, with differentnumbers of beads. . . . . . . . . . . . . . . . . . . . . . . 96

3.21 The real part of the position-squared autocorrelationfunction for a quartic potential bilinearly coupled to abath of 9 harmonic oscillators, calculated at βω = 8. . . 97

xvi

List of Tables

3.1 Rate constant for the Eckart potential at different inversetemperatures. . . . . . . . . . . . . . . . . . . . . . . . . 93

xvii

Page 18: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

List of Papers

The included papers are reproduced with permission from the jour-nals when needed.

Paper I Formation of the Hydroxyl Radical by Radiative AssociationS. Karl-Mikael Svensson, Magnus Gustafsson, and GunnarNymanThe Journal of Physical Chemistry A, 2015, 119 (50), pp12263–12269, DOI: 10.1021/acs.jpca.5b06300

Paper II Dynamics of Gaussian Wigner functions derived from a time-dependent variational principleJens Aage Poulsen, S. Karl-Mikael Svensson, and GunnarNymanAIP Advances, 2017, 7 (11), pp 115018-1–115018-12,DOI: 10.1063/1.5004757

Paper III Classical Wigner Model Based on a Feynman Path IntegralOpen PolymerS. Karl-Mikael Svensson, Jens Aage Poulsen, and GunnarNymanThe Journal of Chemical Physics, 2020, 152 (9), pp 094111-1–094111-20, DOI: 10.1063/1.5126183

Paper IV Calculation of Reaction Rate Constants From a Classical WignerModel Based on a Feynman Path Integral Open PolymerS. Karl-Mikael Svensson, Jens Aage Poulsen, and Gunnar Ny-manManuscript

xix

Page 19: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

List of Papers

The included papers are reproduced with permission from the jour-nals when needed.

Paper I Formation of the Hydroxyl Radical by Radiative AssociationS. Karl-Mikael Svensson, Magnus Gustafsson, and GunnarNymanThe Journal of Physical Chemistry A, 2015, 119 (50), pp12263–12269, DOI: 10.1021/acs.jpca.5b06300

Paper II Dynamics of Gaussian Wigner functions derived from a time-dependent variational principleJens Aage Poulsen, S. Karl-Mikael Svensson, and GunnarNymanAIP Advances, 2017, 7 (11), pp 115018-1–115018-12,DOI: 10.1063/1.5004757

Paper III Classical Wigner Model Based on a Feynman Path IntegralOpen PolymerS. Karl-Mikael Svensson, Jens Aage Poulsen, and GunnarNymanThe Journal of Chemical Physics, 2020, 152 (9), pp 094111-1–094111-20, DOI: 10.1063/1.5126183

Paper IV Calculation of Reaction Rate Constants From a Classical WignerModel Based on a Feynman Path Integral Open PolymerS. Karl-Mikael Svensson, Jens Aage Poulsen, and Gunnar Ny-manManuscript

xix

Page 20: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

List of Papers

Contributions from the Author

Paper I The author performed the computations, did the analysis ofthe result, and wrote most of the paper.

Paper II The author ran some of the calculations and did a few minorderivations.

Paper III The author checked the derivations and did some on his own,rewrote and extended the code of the computer program, per-formed the computations using the new method, did the anal-ysis of the results, and wrote most of the paper.

Paper IV The author did all of the derivations, performed most of thecomputations, did the analysis of the results, and wrote mostof the manuscript.

xx

Acknowledgments

This thesis would not have come to fruition without the assistanceof multiple people, the most obvious of which are my supervisorand co-supervisor, Gunnar Nyman and Jens Poulsen. Also, MagnusGustafsson was supervising the work behind paper I. Furthermore, Iwould like to thank Johan Bergenholtz for being my examiner.

The supportive environment of the physical chemistry group isalso gratefully acknowledged.

A significant fraction of the computations that this thesis refers towere run on clusters, Glenn, Hebbe, and Vera, belonging to ChalmersCentre for Computational Science and Engineering (C3SE) that is apart of the Swedish National Infrastructure for Computing (SNIC).

Finally, many thanks to my family for their support during mystudies.

xxi

Page 21: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

List of Papers

Contributions from the Author

Paper I The author performed the computations, did the analysis ofthe result, and wrote most of the paper.

Paper II The author ran some of the calculations and did a few minorderivations.

Paper III The author checked the derivations and did some on his own,rewrote and extended the code of the computer program, per-formed the computations using the new method, did the anal-ysis of the results, and wrote most of the paper.

Paper IV The author did all of the derivations, performed most of thecomputations, did the analysis of the results, and wrote mostof the manuscript.

xx

Acknowledgments

This thesis would not have come to fruition without the assistanceof multiple people, the most obvious of which are my supervisorand co-supervisor, Gunnar Nyman and Jens Poulsen. Also, MagnusGustafsson was supervising the work behind paper I. Furthermore, Iwould like to thank Johan Bergenholtz for being my examiner.

The supportive environment of the physical chemistry group isalso gratefully acknowledged.

A significant fraction of the computations that this thesis refers towere run on clusters, Glenn, Hebbe, and Vera, belonging to ChalmersCentre for Computational Science and Engineering (C3SE) that is apart of the Swedish National Infrastructure for Computing (SNIC).

Finally, many thanks to my family for their support during mystudies.

xxi

Page 22: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Frame

1

Page 23: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Frame

1

Page 24: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Chapter 1

Introduction

The underlying physical laws necessary for themathematical theory of a large part of physicsand the whole of chemistry are thus completelyknown, and the difficulty is only that the exactapplication of these laws leads to equationsmuch too complicated to be soluble. Ittherefore becomes desirable that approximatepractical methods of applying quantummechanics should be developed, which can leadto an explanation of the main features ofcomplex atomic systems without too muchcomputation.

Paul A. M. Dirac2

Computational chemistry is the craft of calculating, preferablyon a computer, the answers to chemical questions from mathematicalmodels of how entities in chemistry such as atoms, molecules, fluids,or even more fundamental entities such as electrons and atomicnuclei behave. Running calculations on a computer instead of doingexperiments in a laboratory may have the advantage of being fasterand cheaper, and allowing many more things to be tried simulta-neously. However, the practical experiment in the laboratory hasdirect access to the physical reality of the universe and the chemistrywithin it, thus potentially giving the “truth”, while the mathematical

3

Page 25: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Chapter 1

Introduction

The underlying physical laws necessary for themathematical theory of a large part of physicsand the whole of chemistry are thus completelyknown, and the difficulty is only that the exactapplication of these laws leads to equationsmuch too complicated to be soluble. Ittherefore becomes desirable that approximatepractical methods of applying quantummechanics should be developed, which can leadto an explanation of the main features ofcomplex atomic systems without too muchcomputation.

Paul A. M. Dirac2

Computational chemistry is the craft of calculating, preferablyon a computer, the answers to chemical questions from mathematicalmodels of how entities in chemistry such as atoms, molecules, fluids,or even more fundamental entities such as electrons and atomicnuclei behave. Running calculations on a computer instead of doingexperiments in a laboratory may have the advantage of being fasterand cheaper, and allowing many more things to be tried simulta-neously. However, the practical experiment in the laboratory hasdirect access to the physical reality of the universe and the chemistrywithin it, thus potentially giving the “truth”, while the mathematical

3

Page 26: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

1. Introduction

models used in calculations are inevitably approximations of reality,thus giving potentially good but nevertheless approximate results.

There is another difference between computational and exper-imental chemistry, and that is the type of questions that can beanswered. In a simulation the movement of individual atoms, thatare experimentally untrackable, in a chemical reaction can be fol-lowed over time, scales of time and volume impossible in practicalexperimentation can be accessible, and environments, species andprocesses of exotic or even alchemical nature can be handled.

When choosing a computational method to answer a given ques-tion there are many choices to make. One of the common onesis the choice between quantum mechanics and classical mechanics.Quantum mechanics is correct but computationally expensive withits delocalization, tunnelling, zero point energy, and interferencewhile classical mechanics may be wrong but computationally cheapwith simple trajectories for the motion of a body, just as we ashumans experience things in our daily macroscopic lives. In manycases classical mechanics is good enough. However, when light atomssuch as hydrogen are involved, temperatures are low such as oftenin astrochemistry, or there is significant quantum interference, thenquantum mechanics may be essential to describe chemistry in ameaningful way. This leads back to Dirac’s quote2 in the beginningof this chapter:

It therefore becomes desirable that approximate practi-cal methods of applying quantum mechanics should bedeveloped, which can lead to an explanation of the mainfeatures of complex atomic systems without too muchcomputation.

This statement from 1929 is still as valid as it was back then, even asthe limits of what is considered too much computation have changed,and is a concise description of the mission of the work in this thesis.

Generally, chemistry deals with atomic nuclei, electrons, andaggregates of such particles. On the lowest level of chemistry, withelectrons and atomic nuclei as separate entities, it is almost alwaysclearly so that the electrons behave quantum mechanically, but thequestion is if the nuclei should be handled with classical or quantum

4

mechanics. Often the Born-Oppenheimer approximation3 is utilized,meaning that it is assumed that the movement of the electrons andthe movement of the nuclei can be handled separately. Entitiessignificantly heavier than atomic nuclei, such as colloidal particles,for all practical purposes move according to classical mechanics evenif the forces between them may be of a quantum mechanical nature.Some specific examples of when nuclear quantum effects can makean important difference in computations include:

• The volume of light atoms may become large due to thermalquantum fluctuations.4

• Resonances in quasi-bound states may make a significant con-tribution to a reaction rate constant.5

• The delocalization of hydrogen can have a significant impacton the acidity of an active site in an enzyme.6

The work presented in this thesis concerns methods used tohandle the movements of atomic nuclei, when some measure ofquantum mechanics is desirable. Of course the methods can beused for any type of particle, not only atomic nuclei, but atomicnuclei tend to be what chemists in this field of study focus on. Aparticular focus for parts of the thesis is the calculation of reactionrate constants.

5

Page 27: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

1. Introduction

models used in calculations are inevitably approximations of reality,thus giving potentially good but nevertheless approximate results.

There is another difference between computational and exper-imental chemistry, and that is the type of questions that can beanswered. In a simulation the movement of individual atoms, thatare experimentally untrackable, in a chemical reaction can be fol-lowed over time, scales of time and volume impossible in practicalexperimentation can be accessible, and environments, species andprocesses of exotic or even alchemical nature can be handled.

When choosing a computational method to answer a given ques-tion there are many choices to make. One of the common onesis the choice between quantum mechanics and classical mechanics.Quantum mechanics is correct but computationally expensive withits delocalization, tunnelling, zero point energy, and interferencewhile classical mechanics may be wrong but computationally cheapwith simple trajectories for the motion of a body, just as we ashumans experience things in our daily macroscopic lives. In manycases classical mechanics is good enough. However, when light atomssuch as hydrogen are involved, temperatures are low such as oftenin astrochemistry, or there is significant quantum interference, thenquantum mechanics may be essential to describe chemistry in ameaningful way. This leads back to Dirac’s quote2 in the beginningof this chapter:

It therefore becomes desirable that approximate practi-cal methods of applying quantum mechanics should bedeveloped, which can lead to an explanation of the mainfeatures of complex atomic systems without too muchcomputation.

This statement from 1929 is still as valid as it was back then, even asthe limits of what is considered too much computation have changed,and is a concise description of the mission of the work in this thesis.

Generally, chemistry deals with atomic nuclei, electrons, andaggregates of such particles. On the lowest level of chemistry, withelectrons and atomic nuclei as separate entities, it is almost alwaysclearly so that the electrons behave quantum mechanically, but thequestion is if the nuclei should be handled with classical or quantum

4

mechanics. Often the Born-Oppenheimer approximation3 is utilized,meaning that it is assumed that the movement of the electrons andthe movement of the nuclei can be handled separately. Entitiessignificantly heavier than atomic nuclei, such as colloidal particles,for all practical purposes move according to classical mechanics evenif the forces between them may be of a quantum mechanical nature.Some specific examples of when nuclear quantum effects can makean important difference in computations include:

• The volume of light atoms may become large due to thermalquantum fluctuations.4

• Resonances in quasi-bound states may make a significant con-tribution to a reaction rate constant.5

• The delocalization of hydrogen can have a significant impacton the acidity of an active site in an enzyme.6

The work presented in this thesis concerns methods used tohandle the movements of atomic nuclei, when some measure ofquantum mechanics is desirable. Of course the methods can beused for any type of particle, not only atomic nuclei, but atomicnuclei tend to be what chemists in this field of study focus on. Aparticular focus for parts of the thesis is the calculation of reactionrate constants.

5

Page 28: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Chapter 2

Background

A man would make but a very sorry chemist ifhe attended to that department of humanknowledge alone. If your wish is to becomereally a man of science, and not merely a pettyexperimentalist, I should advise you to applyto every branch of natural philosophy,including mathematics.

Fellow-professor M. Waldman (Mary Shelley)1

Quantum mechanics can be formulated in many ways. The onethat most people are familiar with is probably the wavefunctionformulation, which was published in 1926 by Schrodinger7–12∗. Thisformulation uses the Schrodinger equation

i∂

∂tΨ (x, t) = HΨ (x, t) (2.1)

or in bra-ket notation

i∂

∂t|Ψ (t)〉 = H |Ψ (t)〉 , (2.2)

where i is the imaginary unit (√−1 ), is the reduced Planck’s

constant, t is time, x is a position vector, Ψ (x, t) is the wavefunction

∗The original papers are in German. Schrodinger published a summary inEnglish in the same year.13

7

Page 29: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Chapter 2

Background

A man would make but a very sorry chemist ifhe attended to that department of humanknowledge alone. If your wish is to becomereally a man of science, and not merely a pettyexperimentalist, I should advise you to applyto every branch of natural philosophy,including mathematics.

Fellow-professor M. Waldman (Mary Shelley)1

Quantum mechanics can be formulated in many ways. The onethat most people are familiar with is probably the wavefunctionformulation, which was published in 1926 by Schrodinger7–12∗. Thisformulation uses the Schrodinger equation

i∂

∂tΨ (x, t) = HΨ (x, t) (2.1)

or in bra-ket notation

i∂

∂t|Ψ (t)〉 = H |Ψ (t)〉 , (2.2)

where i is the imaginary unit (√−1 ), is the reduced Planck’s

constant, t is time, x is a position vector, Ψ (x, t) is the wavefunction

∗The original papers are in German. Schrodinger published a summary inEnglish in the same year.13

7

Page 30: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

of the system at time t and position x, H is the Hamiltonian operator,and |Ψ (t)〉 is the ket representing the state of the system describedby the wavefunction Ψ at time t. This is, however, not always themost practical formulation to start with when trying to simplifyquantum mechanics.

In this chapter of this thesis two other formulations of quantummechanics are presented in sections 2.1 and 2.2, calculations ofreaction rate constants are introduced in section 2.3, and commonmethods to use for approximate quantum dynamics can be found insection 2.4.

8

2.1. Wigner transform and the Wigner phase space

2.1 Wigner transform and the Wigner

phase space

Of the many approaches to the semiclassicallimit from the quantum domain, the Wignermethod is one of the most immediatelyappealing.

Eric J. Heller14

A formulation of quantum mechanics ascribed to Wigner15 andMoyal16 is the phase space formulation. In this formulation oneworks with functions that depend on both position and momentumsimultaneously, something that may seem very strange from thewavefunction point of view, but that allows the equations to lookmore like classical mechanics.

The Wigner transform of an arbitrary operator Ω is

[Ω]W(x,p) =

∫dDη e−iη•p/

⟨x+

η

2

∣∣∣Ω∣∣∣x− η

2

=

∫dDλ eix•λ/

⟨p+

λ

2

∣∣∣∣Ω∣∣∣∣p− λ

2

⟩(2.3)

where p is a momentum vector, η is a vector where the elementshave the dimension of length, D is the number of degrees of freedomin the system, λ is a vector where the elements have the dimen-

sion of momentum, and the integrals are over all space.∣∣∣x± η

2

and

∣∣∣∣p± λ

2

⟩are eigenkets of position and momentum respectively,

meaning that 〈x|Ψ〉 = Ψ (x) and 〈p|Ψ〉 = Ψ (p).The Wigner transform of a product of two operators is

[Ω1Ω2

]W(x,p)

=[Ω1

]W(x,p) e

− i2

( ←∂∂p

•→∂∂x

−←∂∂x

•→∂∂p

) [Ω2

]W(x,p) (2.4)

where the arrows above the partial derivatives show in which directionthey act.

9

Page 31: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

of the system at time t and position x, H is the Hamiltonian operator,and |Ψ (t)〉 is the ket representing the state of the system describedby the wavefunction Ψ at time t. This is, however, not always themost practical formulation to start with when trying to simplifyquantum mechanics.

In this chapter of this thesis two other formulations of quantummechanics are presented in sections 2.1 and 2.2, calculations ofreaction rate constants are introduced in section 2.3, and commonmethods to use for approximate quantum dynamics can be found insection 2.4.

8

2.1. Wigner transform and the Wigner phase space

2.1 Wigner transform and the Wigner

phase space

Of the many approaches to the semiclassicallimit from the quantum domain, the Wignermethod is one of the most immediatelyappealing.

Eric J. Heller14

A formulation of quantum mechanics ascribed to Wigner15 andMoyal16 is the phase space formulation. In this formulation oneworks with functions that depend on both position and momentumsimultaneously, something that may seem very strange from thewavefunction point of view, but that allows the equations to lookmore like classical mechanics.

The Wigner transform of an arbitrary operator Ω is

[Ω]W(x,p) =

∫dDη e−iη•p/

⟨x+

η

2

∣∣∣Ω∣∣∣x− η

2

=

∫dDλ eix•λ/

⟨p+

λ

2

∣∣∣∣Ω∣∣∣∣p− λ

2

⟩(2.3)

where p is a momentum vector, η is a vector where the elementshave the dimension of length, D is the number of degrees of freedomin the system, λ is a vector where the elements have the dimen-

sion of momentum, and the integrals are over all space.∣∣∣x± η

2

and

∣∣∣∣p± λ

2

⟩are eigenkets of position and momentum respectively,

meaning that 〈x|Ψ〉 = Ψ (x) and 〈p|Ψ〉 = Ψ (p).The Wigner transform of a product of two operators is

[Ω1Ω2

]W(x,p)

=[Ω1

]W(x,p) e

− i2

( ←∂∂p

•→∂∂x

−←∂∂x

•→∂∂p

) [Ω2

]W(x,p) (2.4)

where the arrows above the partial derivatives show in which directionthey act.

9

Page 32: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

If taking the Wigner transform of the probability density opera-tor, 1

(2π)D|Ψ〉 〈Ψ |, the so called Wigner function is obtained. This

function is a quasi probability distribution in phase space that hasthe property that it can be used to obtain the expectation value ofa physical quantity Ω trough

⟨Ω⟩=

∫∫dDx dDp

[1

(2π)D|Ψ〉 〈Ψ |

]W(x,p)

[Ω]W(x,p)

(2.5)

which looks very similar to classical mechanics

〈Ω〉 =∫∫

dDx dDp ρ (x,p)Ω (x,p) (2.6)

where ρ (x,p) is the classical probability distribution function inphase space.

For the interested reader the following section shows how toderive equation 2.3 and 2.5 from the more common formulation ofquantum mechanics.

Derivation of the phase space formulation fromthe wavefunction formulation

To obtain equation 2.5, start with the standard equation⟨Ω⟩=

∫dDxΨ ∗ (x) ΩΨ (x) =

⟨Ψ∣∣∣Ω

∣∣∣Ψ⟩. (2.7)

Introduce two unity operators, 1 =

∫dDx |x〉 〈x|,

⟨Ω⟩=

∫∫dDx1 dDx2 〈Ψ |x1〉

⟨x1

∣∣∣Ω∣∣∣x2

⟩〈x2|Ψ〉

=

∫∫dDx1 dDx2 〈x2|Ψ〉 〈Ψ |x1〉

⟨x1

∣∣∣Ω∣∣∣x2

⟩(2.8)

and then introduce two more unity operators, 1 =

∫dDp |p〉 〈p|,

⟨Ω⟩=

∫∫∫∫dDx1 dDx2 dDp1 dDp2

× 〈x2|p1〉 〈p1|Ψ〉 〈Ψ |p2〉 〈p2|x1〉⟨x1

∣∣∣Ω∣∣∣x2

⟩. (2.9)

10

2.1. Wigner transform and the Wigner phase space

|x〉 and |p〉 are just a Fourier transform away from each other,

|p〉 = 1

(2π)D2

∫dDx eix•p/ |x〉 (2.10)

|x〉 = 1

(2π)D2

∫dDp e−ix•p/ |p〉 , (2.11)

which means that, since 〈x′|x〉 = δ (x′ − x) and 〈p′|p〉 = δ (p′ − p),where δ (x′ − x) is the Dirac delta function,

〈p|x〉 = 1

(2π)D2

∫dDx′ e−ix′•p/ 〈x′|x〉

=1

(2π)D2

∫dDx′ e−ix′•p/ δ (x′ − x)

=1

(2π)D2

e−ix•p/ . (2.12)

This leads to

⟨Ω⟩=

1

(2π)D

∫∫∫∫dDx1 dDx2 dDp1 dDp2

× eix2•p1/ e−ix1•p2/ 〈p1|Ψ〉 〈Ψ |p2〉⟨x1

∣∣∣Ω∣∣∣x2

⟩.

(2.13)

The variables of integration can be changed from x1, x2, p1, andp2 to x=

x1+x22

, η = x1 − x2, p=p1+p2

2, and λ = p1 − p2. For this

change the absolute value of the determinant of the Jacobian matrix,

11

Page 33: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

If taking the Wigner transform of the probability density opera-tor, 1

(2π)D|Ψ〉 〈Ψ |, the so called Wigner function is obtained. This

function is a quasi probability distribution in phase space that hasthe property that it can be used to obtain the expectation value ofa physical quantity Ω trough

⟨Ω⟩=

∫∫dDx dDp

[1

(2π)D|Ψ〉 〈Ψ |

]W(x,p)

[Ω]W(x,p)

(2.5)

which looks very similar to classical mechanics

〈Ω〉 =∫∫

dDx dDp ρ (x,p)Ω (x,p) (2.6)

where ρ (x,p) is the classical probability distribution function inphase space.

For the interested reader the following section shows how toderive equation 2.3 and 2.5 from the more common formulation ofquantum mechanics.

Derivation of the phase space formulation fromthe wavefunction formulation

To obtain equation 2.5, start with the standard equation⟨Ω⟩=

∫dDxΨ ∗ (x) ΩΨ (x) =

⟨Ψ∣∣∣Ω

∣∣∣Ψ⟩. (2.7)

Introduce two unity operators, 1 =

∫dDx |x〉 〈x|,

⟨Ω⟩=

∫∫dDx1 dDx2 〈Ψ |x1〉

⟨x1

∣∣∣Ω∣∣∣x2

⟩〈x2|Ψ〉

=

∫∫dDx1 dDx2 〈x2|Ψ〉 〈Ψ |x1〉

⟨x1

∣∣∣Ω∣∣∣x2

⟩(2.8)

and then introduce two more unity operators, 1 =

∫dDp |p〉 〈p|,

⟨Ω⟩=

∫∫∫∫dDx1 dDx2 dDp1 dDp2

× 〈x2|p1〉 〈p1|Ψ〉 〈Ψ |p2〉 〈p2|x1〉⟨x1

∣∣∣Ω∣∣∣x2

⟩. (2.9)

10

2.1. Wigner transform and the Wigner phase space

|x〉 and |p〉 are just a Fourier transform away from each other,

|p〉 = 1

(2π)D2

∫dDx eix•p/ |x〉 (2.10)

|x〉 = 1

(2π)D2

∫dDp e−ix•p/ |p〉 , (2.11)

which means that, since 〈x′|x〉 = δ (x′ − x) and 〈p′|p〉 = δ (p′ − p),where δ (x′ − x) is the Dirac delta function,

〈p|x〉 = 1

(2π)D2

∫dDx′ e−ix′•p/ 〈x′|x〉

=1

(2π)D2

∫dDx′ e−ix′•p/ δ (x′ − x)

=1

(2π)D2

e−ix•p/ . (2.12)

This leads to

⟨Ω⟩=

1

(2π)D

∫∫∫∫dDx1 dDx2 dDp1 dDp2

× eix2•p1/ e−ix1•p2/ 〈p1|Ψ〉 〈Ψ |p2〉⟨x1

∣∣∣Ω∣∣∣x2

⟩.

(2.13)

The variables of integration can be changed from x1, x2, p1, andp2 to x=

x1+x22

, η = x1 − x2, p=p1+p2

2, and λ = p1 − p2. For this

change the absolute value of the determinant of the Jacobian matrix,

11

Page 34: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

simplified because the matrix is block diagonal, becomes

∣∣∣∣∣∣∣∣∣∣∣∣∣∣∣

det

∂x

∂x1

∂η

∂x1

∂p

∂x1

∂λ

∂x1∂x

∂x2

∂η

∂x2

∂p

∂x2

∂λ

∂x2∂x

∂p1

∂η

∂p1

∂p

∂p1

∂λ

∂p1∂x

∂p2

∂η

∂p2

∂p

∂p2

∂λ

∂p2

∣∣∣∣∣∣∣∣∣∣∣∣∣∣∣

=

∣∣∣∣∣∣∣∣∣∣∣∣∣

det

1

21 0 0

1

2−1 0 0

0 01

21

0 01

2−1

∣∣∣∣∣∣∣∣∣∣∣∣∣

=

∣∣∣∣∣∣∣det

1

21

1

2−1

det

1

21

1

2−1

∣∣∣∣∣∣∣

=

∣∣∣∣∣∣∣

det

1

21

1

2−1

2∣∣∣∣∣∣∣=

∣∣∣∣∣

(−1

2− 1

2

)2∣∣∣∣∣ = 1. (2.14)

Thus, the equation becomes

⟨Ω⟩=

1

(2π)D

∫∫∫∫dDx dDη dDp dDλ

× ei(x−η2 )•(p+

λ2 )/ e−i(x+η

2 )•(p−λ2 )/

×⟨p+

λ

2

∣∣∣∣Ψ⟩⟨

Ψ

∣∣∣∣p− λ

2

⟩⟨x+

η

2

∣∣∣Ω∣∣∣x− η

2

=1

(2π)D

∫∫∫∫dDx dDη dDp dDλ

× ei(x•p+12x•λ− 1

2η•p− 1

4η•λ−x•p+ 1

2x•λ− 1

2η•p+ 1

4η•λ)/

×⟨p+

λ

2

∣∣∣∣Ψ⟩⟨

Ψ

∣∣∣∣p− λ

2

⟩⟨x+

η

2

∣∣∣Ω∣∣∣x− η

2

=1

(2π)D

∫∫∫∫dDx dDη dDp dDλ ei(x•λ−η•p)/

×⟨p+

λ

2

∣∣∣∣Ψ⟩⟨

Ψ

∣∣∣∣p− λ

2

⟩⟨x+

η

2

∣∣∣Ω∣∣∣x− η

2

⟩,

(2.15)

12

2.1. Wigner transform and the Wigner phase space

where Wigner transforms can be isolated, giving

⟨Ω⟩=

∫∫dDx dDp

×(∫

dDλ eix•λ/⟨p+

λ

2

∣∣∣∣ 1

(2π)D|Ψ〉 〈Ψ |

∣∣∣∣p− λ

2

⟩)

×(∫

dDη e−iη•p/⟨x+

η

2

∣∣∣Ω∣∣∣x− η

2

⟩)

=

∫∫dDx dDp

[1

(2π)D|Ψ〉 〈Ψ |

]W(x,p)

[Ω]W(x,p) .

(2.16)

This is equation 2.5.

To prove equation 2.3 two unity operators can be inserted

∫dDη e−iη•p/

⟨x+

η

2

∣∣∣Ω∣∣∣x− η

2

=

∫∫∫dDη dDp1 dDp2 e−iη•p/

×⟨x+

η

2

∣∣∣p1

⟩⟨p1

∣∣∣Ω∣∣∣p2

⟩⟨p2

∣∣∣x− η

2

=1

(2π)D

∫∫∫dDη dDp1 dDp2 e−iη•p/

× ei(x+η2 )•p1/

⟨p1

∣∣∣Ω∣∣∣p2

⟩e−i(x−η

2 )•p2/

=1

(2π)D

∫∫∫dDη dDp1 dDp2

× ei(−η•p+x•p1+12η•p1−x•p2+

12η•p2)/

⟨p1

∣∣∣Ω∣∣∣p2

=1

(2π)D

∫∫∫dDη dDp1 dDp2

× eiη•(−p+p1+p2

2 )/ eix•(p1−p2)/⟨p1

∣∣∣Ω∣∣∣p2

⟩. (2.17)

Changing the variables in the integration from p1, and p2 to p′=p1+p22

,and λ = p1 − p2, with the absolute value of the determinant of the

13

Page 35: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

simplified because the matrix is block diagonal, becomes

∣∣∣∣∣∣∣∣∣∣∣∣∣∣∣

det

∂x

∂x1

∂η

∂x1

∂p

∂x1

∂λ

∂x1∂x

∂x2

∂η

∂x2

∂p

∂x2

∂λ

∂x2∂x

∂p1

∂η

∂p1

∂p

∂p1

∂λ

∂p1∂x

∂p2

∂η

∂p2

∂p

∂p2

∂λ

∂p2

∣∣∣∣∣∣∣∣∣∣∣∣∣∣∣

=

∣∣∣∣∣∣∣∣∣∣∣∣∣

det

1

21 0 0

1

2−1 0 0

0 01

21

0 01

2−1

∣∣∣∣∣∣∣∣∣∣∣∣∣

=

∣∣∣∣∣∣∣det

1

21

1

2−1

det

1

21

1

2−1

∣∣∣∣∣∣∣

=

∣∣∣∣∣∣∣

det

1

21

1

2−1

2∣∣∣∣∣∣∣=

∣∣∣∣∣

(−1

2− 1

2

)2∣∣∣∣∣ = 1. (2.14)

Thus, the equation becomes

⟨Ω⟩=

1

(2π)D

∫∫∫∫dDx dDη dDp dDλ

× ei(x−η2 )•(p+

λ2 )/ e−i(x+η

2 )•(p−λ2 )/

×⟨p+

λ

2

∣∣∣∣Ψ⟩⟨

Ψ

∣∣∣∣p− λ

2

⟩⟨x+

η

2

∣∣∣Ω∣∣∣x− η

2

=1

(2π)D

∫∫∫∫dDx dDη dDp dDλ

× ei(x•p+12x•λ− 1

2η•p− 1

4η•λ−x•p+ 1

2x•λ− 1

2η•p+ 1

4η•λ)/

×⟨p+

λ

2

∣∣∣∣Ψ⟩⟨

Ψ

∣∣∣∣p− λ

2

⟩⟨x+

η

2

∣∣∣Ω∣∣∣x− η

2

=1

(2π)D

∫∫∫∫dDx dDη dDp dDλ ei(x•λ−η•p)/

×⟨p+

λ

2

∣∣∣∣Ψ⟩⟨

Ψ

∣∣∣∣p− λ

2

⟩⟨x+

η

2

∣∣∣Ω∣∣∣x− η

2

⟩,

(2.15)

12

2.1. Wigner transform and the Wigner phase space

where Wigner transforms can be isolated, giving

⟨Ω⟩=

∫∫dDx dDp

×(∫

dDλ eix•λ/⟨p+

λ

2

∣∣∣∣ 1

(2π)D|Ψ〉 〈Ψ |

∣∣∣∣p− λ

2

⟩)

×(∫

dDη e−iη•p/⟨x+

η

2

∣∣∣Ω∣∣∣x− η

2

⟩)

=

∫∫dDx dDp

[1

(2π)D|Ψ〉 〈Ψ |

]W(x,p)

[Ω]W(x,p) .

(2.16)

This is equation 2.5.

To prove equation 2.3 two unity operators can be inserted

∫dDη e−iη•p/

⟨x+

η

2

∣∣∣Ω∣∣∣x− η

2

=

∫∫∫dDη dDp1 dDp2 e−iη•p/

×⟨x+

η

2

∣∣∣p1

⟩⟨p1

∣∣∣Ω∣∣∣p2

⟩⟨p2

∣∣∣x− η

2

=1

(2π)D

∫∫∫dDη dDp1 dDp2 e−iη•p/

× ei(x+η2 )•p1/

⟨p1

∣∣∣Ω∣∣∣p2

⟩e−i(x−η

2 )•p2/

=1

(2π)D

∫∫∫dDη dDp1 dDp2

× ei(−η•p+x•p1+12η•p1−x•p2+

12η•p2)/

⟨p1

∣∣∣Ω∣∣∣p2

=1

(2π)D

∫∫∫dDη dDp1 dDp2

× eiη•(−p+p1+p2

2 )/ eix•(p1−p2)/⟨p1

∣∣∣Ω∣∣∣p2

⟩. (2.17)

Changing the variables in the integration from p1, and p2 to p′=p1+p22

,and λ = p1 − p2, with the absolute value of the determinant of the

13

Page 36: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

Jacobian matrix being

∣∣∣∣∣∣∣det

∂p′

∂p1

∂λ

∂p1∂p′

∂p2

∂λ

∂p2

∣∣∣∣∣∣∣=

∣∣∣∣∣∣∣det

1

21

1

2−1

∣∣∣∣∣∣∣=

∣∣∣∣−1

2− 1

2

∣∣∣∣ = 1

(2.18)

the equation becomes

∫dDη e−iη•p/

⟨x+

η

2

∣∣∣Ω∣∣∣x− η

2

=1

(2π)D

∫∫∫dDη dDp′ dDλ

× eiη•(p′−p)/ eix•λ/

⟨p′ +

λ

2

∣∣∣∣Ω∣∣∣∣p′ − λ

2

⟩. (2.19)

Since the Dirac delta function can be written

δ (ζ) =1

(2π)D

∫dDξ eiζ•ξ, (2.20)

where ζ and ξ are just dummy vector variables, and

δ (cζ) =1

|c|Dδ (ζ) , (2.21)

14

2.1. Wigner transform and the Wigner phase space

where c is a scalar constant, p′ can be integrated out,

∫dDη e−iη•p/

⟨x+

η

2

∣∣∣Ω∣∣∣x− η

2

=1

D

∫∫dDp′ dDλ

(1

(2π)D

∫dDη eiη•(p

′−p)/)

× eix•λ/⟨p′ +

λ

2

∣∣∣∣Ω∣∣∣∣p′ − λ

2

=1

D

∫∫dDp′ dDλ δ

(p′ − p

)

× eix•λ/⟨p′ +

λ

2

∣∣∣∣Ω∣∣∣∣p′ − λ

2

=

∫∫dDp′ dDλ δ (p′ − p)

× eix•λ/⟨p′ +

λ

2

∣∣∣∣Ω∣∣∣∣p′ − λ

2

=

∫dDλ eix•λ/

⟨p+

λ

2

∣∣∣∣Ω∣∣∣∣p− λ

2

⟩. (2.22)

As expected, this is equation 2.3.

The relation between the Wigner transform andFourier transform

An interesting, and sometime useful, property of the Wigner trans-form is that it is the Fourier transform of a matrix element. If inequation 2.3 〈x+η

2 |Ω|x−η2 〉 is just written as a function of η, f (η),

then it can be easily seen that

[Ω]W(x,p) =

∫dDη e−iη•p/ f (η) (2.23)

is a Fourier transform. Taking the inverse of this Fourier transformwould give back the original function.

1

(2π)D

∫dDp

eiη•p/

[Ω]W(x,p) = f (η) (2.24)

15

Page 37: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

Jacobian matrix being

∣∣∣∣∣∣∣det

∂p′

∂p1

∂λ

∂p1∂p′

∂p2

∂λ

∂p2

∣∣∣∣∣∣∣=

∣∣∣∣∣∣∣det

1

21

1

2−1

∣∣∣∣∣∣∣=

∣∣∣∣−1

2− 1

2

∣∣∣∣ = 1

(2.18)

the equation becomes

∫dDη e−iη•p/

⟨x+

η

2

∣∣∣Ω∣∣∣x− η

2

=1

(2π)D

∫∫∫dDη dDp′ dDλ

× eiη•(p′−p)/ eix•λ/

⟨p′ +

λ

2

∣∣∣∣Ω∣∣∣∣p′ − λ

2

⟩. (2.19)

Since the Dirac delta function can be written

δ (ζ) =1

(2π)D

∫dDξ eiζ•ξ, (2.20)

where ζ and ξ are just dummy vector variables, and

δ (cζ) =1

|c|Dδ (ζ) , (2.21)

14

2.1. Wigner transform and the Wigner phase space

where c is a scalar constant, p′ can be integrated out,

∫dDη e−iη•p/

⟨x+

η

2

∣∣∣Ω∣∣∣x− η

2

=1

D

∫∫dDp′ dDλ

(1

(2π)D

∫dDη eiη•(p

′−p)/)

× eix•λ/⟨p′ +

λ

2

∣∣∣∣Ω∣∣∣∣p′ − λ

2

=1

D

∫∫dDp′ dDλ δ

(p′ − p

)

× eix•λ/⟨p′ +

λ

2

∣∣∣∣Ω∣∣∣∣p′ − λ

2

=

∫∫dDp′ dDλ δ (p′ − p)

× eix•λ/⟨p′ +

λ

2

∣∣∣∣Ω∣∣∣∣p′ − λ

2

=

∫dDλ eix•λ/

⟨p+

λ

2

∣∣∣∣Ω∣∣∣∣p− λ

2

⟩. (2.22)

As expected, this is equation 2.3.

The relation between the Wigner transform andFourier transform

An interesting, and sometime useful, property of the Wigner trans-form is that it is the Fourier transform of a matrix element. If inequation 2.3 〈x+η

2 |Ω|x−η2 〉 is just written as a function of η, f (η),

then it can be easily seen that

[Ω]W(x,p) =

∫dDη e−iη•p/ f (η) (2.23)

is a Fourier transform. Taking the inverse of this Fourier transformwould give back the original function.

1

(2π)D

∫dDp

eiη•p/

[Ω]W(x,p) = f (η) (2.24)

15

Page 38: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

More nicely written as

1

(2π)D

∫dDp eiη•p/

[Ω]W(x,p) =

⟨x+

η

2

∣∣∣Ω∣∣∣x− η

2

⟩.

(2.25)

As any pair of positions x1 and x2 can be rewritten as x+η2and

x−η2through x=

x1+x22

and η = x1 −x2, this means that any positionmatrix element can be written as

⟨x1

∣∣∣Ω∣∣∣x2

⟩=

1

(2π)D

∫dDp ei(x1−x2)•p/

×[Ω]W

(x1 + x2

2,p

). (2.26)

The classical Wigner method

Although he did not recommend using the method, Heller14 in 1976introduced the first version of the classical Wigner method. It waslater, in 1998, introduced in a more general form by Wang, Sun,and Miller.17 The classical Wigner method approximates quantummechanics by propagating a Wigner transformed quantity forwardin time with classical mechanics, i.e.

[e

iHt Ω e−

iHt

]W(x,p)

[Ω]W(x(t),p(t)) . (2.27)

This gives exact quantum mechanics for a free particle, linear poten-tial, and harmonic potential. For other kinds of potentials it is anapproximation.14

Wang, Sun, and Miller17 developed the classical Wigner methodas a linearization approximation of the semi-classical initial valuerepresentation (SC-IVR) of Miller18 (well explained by the sameauthor in a later paper19). Because of the linearization approximationapplied to SC-IVR to obtain the classical Wigner method a commonname for the classical Wigner method is linearized semi-classicalinitial value representation (LSC-IVR). The thing that is linearizedis the differences between the positions and momenta in the pathsforward and backward in time. When going from SC-IVR to LSC-IVR the quantum coherence in real time, that SC-IVR has, is lost.

16

2.1. Wigner transform and the Wigner phase space

Condensed phase and large systems have many degrees of freedomcoupled together that typically can result in rapid decoherence, so forthese kinds of systems the loss of quantum real time coherence is notnecessarily a big problem.20 With the linearization approximationcomes the benefit of less oscillatory integrands which are easier toevaluate numerically than the ones in ordinary SC-IVR.20

Another, possibly more straightforward, way of deriving theclassical Wigner method is the linearized path integral (LPI) ap-proach.21,22 The derivation of LPI is shown in the last part of section2.2. There it is also proven that the classical Wigner method is exactfor constant, linear, and harmonic potentials and sums of these.

The classical Wigner method is exact at the initial time, withzero point energy and motion, static tunneling , interference and soon, but the method can not handle such things as dynamic tunnelingand dynamic quantum interference.

An example of successful usage of the classical Wigner methodis calculation of vibrational energy relaxation rate constants,23,24

where a few different systems were tested. Another example isthe calculation of kinetic energy and density fluctuation spectrumof liquid neon at 27 K.25 A third example is the calculation of aquantum correction factor for the far IR-spectrum of liquid water at296 K.26

An instance where the limitations of the classical Wigner methodhave a detrimental effect is the simulation of a graphite surface.27

In an anisotropic material, such as graphite, there will be morezero point energy in some directions than in others, and during theclassical propagation this energy can leak to the directions withless zero point energy. Another example where the leakage of zeropoint energy causes problems for the classical Wigner method is thecalculation of the self diffusion coefficient of liquid water.28

17

Page 39: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

More nicely written as

1

(2π)D

∫dDp eiη•p/

[Ω]W(x,p) =

⟨x+

η

2

∣∣∣Ω∣∣∣x− η

2

⟩.

(2.25)

As any pair of positions x1 and x2 can be rewritten as x+η2and

x−η2through x=

x1+x22

and η = x1 −x2, this means that any positionmatrix element can be written as

⟨x1

∣∣∣Ω∣∣∣x2

⟩=

1

(2π)D

∫dDp ei(x1−x2)•p/

×[Ω]W

(x1 + x2

2,p

). (2.26)

The classical Wigner method

Although he did not recommend using the method, Heller14 in 1976introduced the first version of the classical Wigner method. It waslater, in 1998, introduced in a more general form by Wang, Sun,and Miller.17 The classical Wigner method approximates quantummechanics by propagating a Wigner transformed quantity forwardin time with classical mechanics, i.e.

[e

iHt Ω e−

iHt

]W(x,p)

[Ω]W(x(t),p(t)) . (2.27)

This gives exact quantum mechanics for a free particle, linear poten-tial, and harmonic potential. For other kinds of potentials it is anapproximation.14

Wang, Sun, and Miller17 developed the classical Wigner methodas a linearization approximation of the semi-classical initial valuerepresentation (SC-IVR) of Miller18 (well explained by the sameauthor in a later paper19). Because of the linearization approximationapplied to SC-IVR to obtain the classical Wigner method a commonname for the classical Wigner method is linearized semi-classicalinitial value representation (LSC-IVR). The thing that is linearizedis the differences between the positions and momenta in the pathsforward and backward in time. When going from SC-IVR to LSC-IVR the quantum coherence in real time, that SC-IVR has, is lost.

16

2.1. Wigner transform and the Wigner phase space

Condensed phase and large systems have many degrees of freedomcoupled together that typically can result in rapid decoherence, so forthese kinds of systems the loss of quantum real time coherence is notnecessarily a big problem.20 With the linearization approximationcomes the benefit of less oscillatory integrands which are easier toevaluate numerically than the ones in ordinary SC-IVR.20

Another, possibly more straightforward, way of deriving theclassical Wigner method is the linearized path integral (LPI) ap-proach.21,22 The derivation of LPI is shown in the last part of section2.2. There it is also proven that the classical Wigner method is exactfor constant, linear, and harmonic potentials and sums of these.

The classical Wigner method is exact at the initial time, withzero point energy and motion, static tunneling , interference and soon, but the method can not handle such things as dynamic tunnelingand dynamic quantum interference.

An example of successful usage of the classical Wigner methodis calculation of vibrational energy relaxation rate constants,23,24

where a few different systems were tested. Another example isthe calculation of kinetic energy and density fluctuation spectrumof liquid neon at 27 K.25 A third example is the calculation of aquantum correction factor for the far IR-spectrum of liquid water at296 K.26

An instance where the limitations of the classical Wigner methodhave a detrimental effect is the simulation of a graphite surface.27

In an anisotropic material, such as graphite, there will be morezero point energy in some directions than in others, and during theclassical propagation this energy can leak to the directions withless zero point energy. Another example where the leakage of zeropoint energy causes problems for the classical Wigner method is thecalculation of the self diffusion coefficient of liquid water.28

17

Page 40: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

2.2 Feynman path integral formulation

of quantum mechanics

Yes, there is goal and meaning in our path -but it’s the way that is the labour’s worth.

Karin M. Boye (transl. David McDuff)29

The path integral formulation of quantum mechanics was intro-duced in 1948 by Feynman.30 It is also explained in the famous bookby Feynman and Hibbs.31 In this formulation of quantum mechanicsone looks at all possible paths from one position to another andmakes a weighted “average” over all the paths, with the weightsbeing complex numbers that all are equal in magnitude.

Derivation of the path integral formulation

We can start with the position matrix element of the time propagationoperator

⟨xFinal

∣∣∣e− iHt

∣∣∣xInitial

⟩.

Inserting the unit operator N−1 times and at the same time dividingthe time propagation operator into N parts gives

⟨xFinal

∣∣∣e− iHt

∣∣∣xInitial

⟩=

N−1∏

j=1

∫dDxj

⟨xFinal

∣∣∣e− iHtN

∣∣∣xN−1

. . .⟨x2

∣∣∣e− iHtN

∣∣∣x1

⟩⟨x1

∣∣∣e− iHtN

∣∣∣xInitial

⟩.

(2.28)

For simplicity xInitial will be called x0 and xFinal will be called xN

in the following.In the limit N → ∞ the Trotter product32 can be used, giving

limN→∞

e−iHtN = lim

N→∞e−

iT tN e−

iV tN , (2.29)

18

2.2. Feynman path integral formulation of quantum mechanics

where V is the potential energy operator and T is the kinetic energyoperator. If it is also assumed that the potential only depends onposition, then

⟨xN

∣∣∣e− iHt

∣∣∣x0

⟩= lim

N→∞

N−1∏

j=1

∫dDxj

×⟨xN

∣∣∣e− iT tN e−

iV tN

∣∣∣xN−1

. . .⟨x2

∣∣∣e− iT tN e−

iV tN

∣∣∣x1

×⟨x1

∣∣∣e− iT tN e−

iV tN

∣∣∣x0

= limN→∞

N−1∏

j=1

∫dDxj

×⟨xN

∣∣∣e− iT tN

∣∣∣xN−1

⟩e−

iV (xN−1)tN

. . .⟨x2

∣∣∣e− iT tN

∣∣∣x1

⟩e−

iV (x1)tN

×⟨x1

∣∣∣e− iT tN

∣∣∣x0

⟩e−

iV (x0)tN

= limN→∞

N−1∏

j=1

∫dDxj

⟨xN

∣∣∣e− iT tN

∣∣∣xN−1

. . .⟨x2

∣∣∣e− iT tN

∣∣∣x1

⟩⟨x1

∣∣∣e− iT tN

∣∣∣x0

× e−itN

∑N−1j=0 V (xj), (2.30)

where V (xj) is the potential energy

The kinetic energy operator is T =1

2

(m−1p

)• p, where p is the

momentum operator and m is a square diagonal matrix with themasses for the various degrees of freedom in the diagonal. For each

19

Page 41: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

2.2 Feynman path integral formulation

of quantum mechanics

Yes, there is goal and meaning in our path -but it’s the way that is the labour’s worth.

Karin M. Boye (transl. David McDuff)29

The path integral formulation of quantum mechanics was intro-duced in 1948 by Feynman.30 It is also explained in the famous bookby Feynman and Hibbs.31 In this formulation of quantum mechanicsone looks at all possible paths from one position to another andmakes a weighted “average” over all the paths, with the weightsbeing complex numbers that all are equal in magnitude.

Derivation of the path integral formulation

We can start with the position matrix element of the time propagationoperator

⟨xFinal

∣∣∣e− iHt

∣∣∣xInitial

⟩.

Inserting the unit operator N−1 times and at the same time dividingthe time propagation operator into N parts gives

⟨xFinal

∣∣∣e− iHt

∣∣∣xInitial

⟩=

N−1∏

j=1

∫dDxj

⟨xFinal

∣∣∣e− iHtN

∣∣∣xN−1

. . .⟨x2

∣∣∣e− iHtN

∣∣∣x1

⟩⟨x1

∣∣∣e− iHtN

∣∣∣xInitial

⟩.

(2.28)

For simplicity xInitial will be called x0 and xFinal will be called xN

in the following.In the limit N → ∞ the Trotter product32 can be used, giving

limN→∞

e−iHtN = lim

N→∞e−

iT tN e−

iV tN , (2.29)

18

2.2. Feynman path integral formulation of quantum mechanics

where V is the potential energy operator and T is the kinetic energyoperator. If it is also assumed that the potential only depends onposition, then

⟨xN

∣∣∣e− iHt

∣∣∣x0

⟩= lim

N→∞

N−1∏

j=1

∫dDxj

×⟨xN

∣∣∣e− iT tN e−

iV tN

∣∣∣xN−1

. . .⟨x2

∣∣∣e− iT tN e−

iV tN

∣∣∣x1

×⟨x1

∣∣∣e− iT tN e−

iV tN

∣∣∣x0

= limN→∞

N−1∏

j=1

∫dDxj

×⟨xN

∣∣∣e− iT tN

∣∣∣xN−1

⟩e−

iV (xN−1)tN

. . .⟨x2

∣∣∣e− iT tN

∣∣∣x1

⟩e−

iV (x1)tN

×⟨x1

∣∣∣e− iT tN

∣∣∣x0

⟩e−

iV (x0)tN

= limN→∞

N−1∏

j=1

∫dDxj

⟨xN

∣∣∣e− iT tN

∣∣∣xN−1

. . .⟨x2

∣∣∣e− iT tN

∣∣∣x1

⟩⟨x1

∣∣∣e− iT tN

∣∣∣x0

× e−itN

∑N−1j=0 V (xj), (2.30)

where V (xj) is the potential energy

The kinetic energy operator is T =1

2

(m−1p

)• p, where p is the

momentum operator and m is a square diagonal matrix with themasses for the various degrees of freedom in the diagonal. For each

19

Page 42: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

⟨xj

∣∣∣e− iT tN

∣∣∣xj′

⟩in the above equation

⟨xj

∣∣∣e− iT tN

∣∣∣xj′

⟩=

⟨xj

∣∣∣∣∣e−

i 12(m−1p)•ptN

∣∣∣∣∣xj′

=1

(2π)D

∫∫dDpj d

Dpj′ 〈xj|pj〉

×⟨pj

∣∣∣∣∣e−

i 12(m−1p)•ptN

∣∣∣∣∣pj′

⟩〈pj′ |xj′〉

=1

(2π)D

∫∫dDpj d

Dpj′ eixj•pj

e−ixj′ •pj′

×⟨pj

∣∣∣∣∣e−

i 12(m−1p)•ptN

∣∣∣∣∣pj′

=1

(2π)D

∫∫dDpj d

Dpj′ eixj•pj

e−ixj′ •pj′

× e−i 12(m−1pj)•pjt

N 〈pj|pj′〉

=1

(2π)D

∫∫dDpj d

Dpj′ eixj•pj

e−ixj′ •pj′

× e−i 12(m−1pj)•pjt

N δ (pj − pj′) . (2.31)

20

2.2. Feynman path integral formulation of quantum mechanics

Integrating over pj′ one acquires

⟨xj

∣∣∣e− iT tN

∣∣∣xj′

⟩=

1

(2π)D

∫dDpj e

i(xj−xj′)•pj e−

i 12(m−1pj)•pjtN

=1

(2π)D

∫dDpj e

− it2N((m−1pj)•pj− 2N

t (xj−xj′)•pj)

=1

(2π)D

∫dDpj e

− it2N

∑Dj′′=1

(p2j′′,jmj′′

− 2Nt (xj′′,j−xj′′,j′)pj′′,j

)

=1

(2π)D

∫dDpj

× e− it

2N∑D

j′′=1

(1

mj′′

(p2j′′,j−

2Nmj′′t (xj′′,j−xj′′,j′)pj′′,j

))

=1

(2π)D

∫dDpj

× e− it

2N∑D

j′′=1

(1

mj′′

(pj′′,j−

Nmj′′t (xj′′,j−xj′′,j′)

)2)

× eit

2N∑D

j′′=1

(N2m2

j′′mj′′ t2

(xj′′,j−xj′′,j′)2

)

=1

(2π)De

i∑D

j′′=1

(Nmj′′

2t (xj′′,j−xj′′,j′)2) ∫

dDpj

× e− it

2N∑D

j′′=1

(1

mj′′

(pj′′,j−

Nmj′′t (xj′′,j−xj′′,j′)

)2)

, (2.32)

where j′′ denotes a component of x or p. Changing variables ofintegration from pj′′,j to

21

Page 43: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

⟨xj

∣∣∣e− iT tN

∣∣∣xj′

⟩in the above equation

⟨xj

∣∣∣e− iT tN

∣∣∣xj′

⟩=

⟨xj

∣∣∣∣∣e−

i 12(m−1p)•ptN

∣∣∣∣∣xj′

=1

(2π)D

∫∫dDpj d

Dpj′ 〈xj|pj〉

×⟨pj

∣∣∣∣∣e−

i 12(m−1p)•ptN

∣∣∣∣∣pj′

⟩〈pj′ |xj′〉

=1

(2π)D

∫∫dDpj d

Dpj′ eixj•pj

e−ixj′ •pj′

×⟨pj

∣∣∣∣∣e−

i 12(m−1p)•ptN

∣∣∣∣∣pj′

=1

(2π)D

∫∫dDpj d

Dpj′ eixj•pj

e−ixj′ •pj′

× e−i 12(m−1pj)•pjt

N 〈pj|pj′〉

=1

(2π)D

∫∫dDpj d

Dpj′ eixj•pj

e−ixj′ •pj′

× e−i 12(m−1pj)•pjt

N δ (pj − pj′) . (2.31)

20

2.2. Feynman path integral formulation of quantum mechanics

Integrating over pj′ one acquires

⟨xj

∣∣∣e− iT tN

∣∣∣xj′

⟩=

1

(2π)D

∫dDpj e

i(xj−xj′)•pj e−

i 12(m−1pj)•pjtN

=1

(2π)D

∫dDpj e

− it2N((m−1pj)•pj− 2N

t (xj−xj′)•pj)

=1

(2π)D

∫dDpj e

− it2N

∑Dj′′=1

(p2j′′,jmj′′

− 2Nt (xj′′,j−xj′′,j′)pj′′,j

)

=1

(2π)D

∫dDpj

× e− it

2N∑D

j′′=1

(1

mj′′

(p2j′′,j−

2Nmj′′t (xj′′,j−xj′′,j′)pj′′,j

))

=1

(2π)D

∫dDpj

× e− it

2N∑D

j′′=1

(1

mj′′

(pj′′,j−

Nmj′′t (xj′′,j−xj′′,j′)

)2)

× eit

2N∑D

j′′=1

(N2m2

j′′mj′′ t2

(xj′′,j−xj′′,j′)2

)

=1

(2π)De

i∑D

j′′=1

(Nmj′′

2t (xj′′,j−xj′′,j′)2) ∫

dDpj

× e− it

2N∑D

j′′=1

(1

mj′′

(pj′′,j−

Nmj′′t (xj′′,j−xj′′,j′)

)2)

, (2.32)

where j′′ denotes a component of x or p. Changing variables ofintegration from pj′′,j to

21

Page 44: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

ζj′′ =

√t

2Nmj′′

(pj′′,j −

Nmj′′

t(xj′′,j − xj′′,j′)

)gives

⟨xj

∣∣∣e− iT tN

∣∣∣xj′

⟩=

1

(2π)De

i∑D

j′′=1

(Nmj′′

2t (xj′′,j−xj′′,j′)2)

×(2Nt

)D2

√√√√D∏

j′′=1

mj′′

∫dDζ e

−i∑D

j′′=1 ζ2j′′

=1

(2π)De

iN2t(m(xj−xj′))•(xj−xj′)

(2Nt

)D2

×√det (m)

∫dDζ

D∏

j′′=1

(cos

(ζ2j′′

)− i sin

(ζ2j′′

))

=1

(2π)De

iN2t(m(xj−xj′))•(xj−xj′)

(2Nt

)D2

×√det (m)

D∏

j′′=1

(√π

2− i

√π

2

)

=1

(2π)De

iN2t(m(xj−xj′))•(xj−xj′)

(2πNit

)D2

×√det (m)

=

(N

2πit

)D2 √

det (m) eiN2t(m(xj−xj′))•(xj−xj′) . (2.33)

22

2.2. Feynman path integral formulation of quantum mechanics

Putting this result into equation 2.30 gives

⟨xN

∣∣∣e− iHt

∣∣∣x0

⟩= lim

N→∞

N−1∏

j=1

∫dDxj

×((

N

2πit

)D2 √

det (m) eiN2t (m(xN−xN−1))•(xN−xN−1)

. . .

(N

2πit

)D2 √

det (m) eiN2t (m(x2−x1))•(x2−x1)

×(

N

2πit

)D2 √

det (m) eiN2t (m(x1−x0))•(x1−x0)

)

× e−itN

∑N−1j=0 V (xj)

= limN→∞

(N

2πit

)ND2

(det (m))N2

N−1∏

j=1

∫dDxj

× eiN2t

∑N−1j=0 (m(xj+1−xj))•(xj+1−xj)

× e−itN

∑N−1j=0 V (xj)

= limN→∞

(N

2πit

)ND2

(det (m))N2

N−1∏

j=1

∫dDxj

× eitN

∑N−1j=0

(N2

t212(m(xj+1−xj))•(xj+1−xj)−V (xj)

). (2.34)

Now, let’s introducet

N= ∆t and rewrite all xj as x (t′), where

t′ = j∆t.

⟨xN

∣∣∣e− iHt

∣∣∣x0

⟩= lim

N→∞

(det (m)

(2πi∆t)D

)N2

N−1∏

j=1

∫dDx (j∆t)

× ei∆t

∑N−1j=0 ( 1

2(mx((j+1)∆t)−x(j∆t)

∆t )•x((j+1)∆t)−x(j∆t)∆t )

× ei∆t

∑N−1j=0 (−V (x(j∆t))) (2.35)

where it can be recognized that as N → ∞ and ∆t → 0

lim∆t→0

x ((j + 1)∆t)− x (j∆t)

∆t=

dx (t′)

dt′= x (t′) . (2.36)

23

Page 45: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

ζj′′ =

√t

2Nmj′′

(pj′′,j −

Nmj′′

t(xj′′,j − xj′′,j′)

)gives

⟨xj

∣∣∣e− iT tN

∣∣∣xj′

⟩=

1

(2π)De

i∑D

j′′=1

(Nmj′′

2t (xj′′,j−xj′′,j′)2)

×(2Nt

)D2

√√√√D∏

j′′=1

mj′′

∫dDζ e

−i∑D

j′′=1 ζ2j′′

=1

(2π)De

iN2t(m(xj−xj′))•(xj−xj′)

(2Nt

)D2

×√

det (m)

∫dDζ

D∏

j′′=1

(cos

(ζ2j′′

)− i sin

(ζ2j′′

))

=1

(2π)De

iN2t(m(xj−xj′))•(xj−xj′)

(2Nt

)D2

×√

det (m)D∏

j′′=1

(√π

2− i

√π

2

)

=1

(2π)De

iN2t(m(xj−xj′))•(xj−xj′)

(2πNit

)D2

×√

det (m)

=

(N

2πit

)D2 √

det (m) eiN2t(m(xj−xj′))•(xj−xj′) . (2.33)

22

2.2. Feynman path integral formulation of quantum mechanics

Putting this result into equation 2.30 gives

⟨xN

∣∣∣e− iHt

∣∣∣x0

⟩= lim

N→∞

N−1∏

j=1

∫dDxj

×((

N

2πit

)D2 √

det (m) eiN2t (m(xN−xN−1))•(xN−xN−1)

. . .

(N

2πit

)D2 √

det (m) eiN2t (m(x2−x1))•(x2−x1)

×(

N

2πit

)D2 √

det (m) eiN2t (m(x1−x0))•(x1−x0)

)

× e−itN

∑N−1j=0 V (xj)

= limN→∞

(N

2πit

)ND2

(det (m))N2

N−1∏

j=1

∫dDxj

× eiN2t

∑N−1j=0 (m(xj+1−xj))•(xj+1−xj)

× e−itN

∑N−1j=0 V (xj)

= limN→∞

(N

2πit

)ND2

(det (m))N2

N−1∏

j=1

∫dDxj

× eitN

∑N−1j=0

(N2

t212(m(xj+1−xj))•(xj+1−xj)−V (xj)

). (2.34)

Now, let’s introducet

N= ∆t and rewrite all xj as x (t′), where

t′ = j∆t.

⟨xN

∣∣∣e− iHt

∣∣∣x0

⟩= lim

N→∞

(det (m)

(2πi∆t)D

)N2

N−1∏

j=1

∫dDx (j∆t)

× ei∆t

∑N−1j=0 ( 1

2(mx((j+1)∆t)−x(j∆t)

∆t )•x((j+1)∆t)−x(j∆t)∆t )

× ei∆t

∑N−1j=0 (−V (x(j∆t))) (2.35)

where it can be recognized that as N → ∞ and ∆t → 0

lim∆t→0

x ((j + 1)∆t)− x (j∆t)

∆t=

dx (t′)

dt′= x (t′) . (2.36)

23

Page 46: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

This is the very definition of a derivative and the sum in the exponentis the Riemann integral over dt′ , so

⟨xN

∣∣∣e− iHt

∣∣∣x0

⟩= lim

N→∞

(det (m)

(2πi∆t)D

)N2

t′

∫dDx (t′)

× ei∫ t0 dt′ ( 1

2(mx(t′))•x(t′)−V (x(t′))) . (2.37)

Here the integrand in the exponent is a Langrangian, and integratinga Lagrangian over time gives the classical action, Scl (x (t′) , x (t′) , t).∏

t′

∫dDx (t′)

is an integration over all paths from x (0) to x (t).

(det (m)

(2πi∆t)D

)N2

is a normalization factor, that can be called AN . A

shorter way to write then is

⟨xN

∣∣∣e− iHt

∣∣∣x0

⟩= lim

N→∞AN

t′

∫dDx (t′)

× eiScl(x(t

′),x(t′),t) . (2.38)

This is the path integral formulation. It uses a weighted averageover all paths, with the weights being complex exponentials of theclassical action of the path, all of which have equal magnitude.

The matrix element in equation 2.38 is called a kernel,

K (xN ,x0, t) =⟨xN

∣∣∣e− iHt

∣∣∣x0

⟩. (2.39)

The kernel can be used to propagate a wave function Ψ (x, 0) a timet forward in time to Ψ (x, t) through

Ψ (x, t) = 〈x|Ψ(t)〉 =∫

dDx′ K (x,x′, t) 〈x′|Ψ(0)〉

=

∫dDx′

⟨x∣∣∣e− iHt

∣∣∣x′⟩〈x′|Ψ(0)〉

=⟨x∣∣∣e− iHt

∣∣∣Ψ(0)⟩= e−

iHt 〈x|Ψ(0)〉

= e−iHt Ψ (x, 0) . (2.40)

24

2.2. Feynman path integral formulation of quantum mechanics

In general equations such as 2.38 and 2.40 are not analyticallysolvable and in a numerical approximation, with finite N , it is verydifficult to converge an integral of such a quickly oscillating complexfunction as is relevant here. Thus, this formulation may not be veryuseful for directly calculating real time propagation, but is rather,at least from a computational chemist’s point of view, of interest asa starting point from which approximate models and methods canbe developed.

Imaginary time path integrals

Looking at the Boltzmann operator e−βH , or e− H

kBT , where β =1

kBT, kB is Boltzmann’s constant, and T is absolute temperature,

it looks similar to the time propagation operator e−iHt . In fact, if

an imaginary time, given as t = −iβ, would be inserted into thetime propagation operator then the Boltzmann operator would beobtained.

Using this imaginary time in equation 2.34 gives

⟨xN

∣∣∣e−βH∣∣∣x0

⟩= lim

N→∞

(N

2πβ2

)ND2

(det (m))N2

×

N−1∏

j=1

∫dDxj

e

βN

∑N−1j=0

(N2

−β2212(m(xj+1−xj))•(xj+1−xj)−V (xj)

)

= limN→∞

(N

2πβ

)ND2

ND (det (m))N2

N−1∏

j=1

∫dDxj

× e− β

N

∑N−1j=0

(N2

2β22 (m(xj+1−xj))•(xj+1−xj)+V (xj)). (2.41)

−−iβN

is the new ∆t in imaginary time. The kinetic energy termsN2

2β22mj′′(xj′′,j+1−xj′′,j)2 are of the same form as the potential energy of

harmonic springs with force constant − N2

β22mj′′ . These are typicallycalled “spring terms”. Thus the integrand can be seen as a Boltzmannfactor, at the temperature NT , of the potential energy of a numberof copies of the system, xj, connected by harmonic springs.

Using this imaginary time kernel to calculate the canonical parti-tion function, Z, one simply has to set x0 = xN and integrate over

25

Page 47: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

This is the very definition of a derivative and the sum in the exponentis the Riemann integral over dt′ , so

⟨xN

∣∣∣e− iHt

∣∣∣x0

⟩= lim

N→∞

(det (m)

(2πi∆t)D

)N2

t′

∫dDx (t′)

× ei∫ t0 dt′ ( 1

2(mx(t′))•x(t′)−V (x(t′))) . (2.37)

Here the integrand in the exponent is a Langrangian, and integratinga Lagrangian over time gives the classical action, Scl (x (t′) , x (t′) , t).∏

t′

∫dDx (t′)

is an integration over all paths from x (0) to x (t).

(det (m)

(2πi∆t)D

)N2

is a normalization factor, that can be called AN . A

shorter way to write then is

⟨xN

∣∣∣e− iHt

∣∣∣x0

⟩= lim

N→∞AN

t′

∫dDx (t′)

× eiScl(x(t

′),x(t′),t) . (2.38)

This is the path integral formulation. It uses a weighted averageover all paths, with the weights being complex exponentials of theclassical action of the path, all of which have equal magnitude.

The matrix element in equation 2.38 is called a kernel,

K (xN ,x0, t) =⟨xN

∣∣∣e− iHt

∣∣∣x0

⟩. (2.39)

The kernel can be used to propagate a wave function Ψ (x, 0) a timet forward in time to Ψ (x, t) through

Ψ (x, t) = 〈x|Ψ(t)〉 =∫

dDx′ K (x,x′, t) 〈x′|Ψ(0)〉

=

∫dDx′

⟨x∣∣∣e− iHt

∣∣∣x′⟩〈x′|Ψ(0)〉

=⟨x∣∣∣e− iHt

∣∣∣Ψ(0)⟩= e−

iHt 〈x|Ψ(0)〉

= e−iHt Ψ (x, 0) . (2.40)

24

2.2. Feynman path integral formulation of quantum mechanics

In general equations such as 2.38 and 2.40 are not analyticallysolvable and in a numerical approximation, with finite N , it is verydifficult to converge an integral of such a quickly oscillating complexfunction as is relevant here. Thus, this formulation may not be veryuseful for directly calculating real time propagation, but is rather,at least from a computational chemist’s point of view, of interest asa starting point from which approximate models and methods canbe developed.

Imaginary time path integrals

Looking at the Boltzmann operator e−βH , or e− H

kBT , where β =1

kBT, kB is Boltzmann’s constant, and T is absolute temperature,

it looks similar to the time propagation operator e−iHt . In fact, if

an imaginary time, given as t = −iβ, would be inserted into thetime propagation operator then the Boltzmann operator would beobtained.

Using this imaginary time in equation 2.34 gives

⟨xN

∣∣∣e−βH∣∣∣x0

⟩= lim

N→∞

(N

2πβ2

)ND2

(det (m))N2

×

N−1∏

j=1

∫dDxj

e

βN

∑N−1j=0

(N2

−β2212(m(xj+1−xj))•(xj+1−xj)−V (xj)

)

= limN→∞

(N

2πβ

)ND2

ND (det (m))N2

N−1∏

j=1

∫dDxj

× e− β

N

∑N−1j=0

(N2

2β22 (m(xj+1−xj))•(xj+1−xj)+V (xj)). (2.41)

−−iβN

is the new ∆t in imaginary time. The kinetic energy termsN2

2β22mj′′(xj′′,j+1−xj′′,j)2 are of the same form as the potential energy of

harmonic springs with force constant − N2

β22mj′′ . These are typicallycalled “spring terms”. Thus the integrand can be seen as a Boltzmannfactor, at the temperature NT , of the potential energy of a numberof copies of the system, xj, connected by harmonic springs.

Using this imaginary time kernel to calculate the canonical parti-tion function, Z, one simply has to set x0 = xN and integrate over

25

Page 48: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

Figure 2.1: Illustration of a classical particle (1.) and its correspond-ing imaginary time path integral ring polymer (2.). The zigzag linesrepresent the spring terms.

xN

Z = Tre−βH

=

∫dDxN

⟨xN

∣∣∣e−βH∣∣∣xN

= limN→∞

(N

2πβ

)ND2

ND (det (m))N2

N∏

j=1

∫dDxj

× e− β

N

∑Nj=1

(N2

2β22 (m(x(j mod N)+1−xj))•(x(j mod N)+1−xj)+V (xj)),

(2.42)

where Tr denotes a trace and mod is the modulus operator.

The extended phase space of N copies of the system connectedby harmonic springs is often seen as turning each single particle intoa “polymer” with N monomers. The circular path integral found inequation 2.42 is then called a “ring polymer”. This is schematicallyillustrated in figure 2.1.

26

2.2. Feynman path integral formulation of quantum mechanics

Since the imaginary time kernel is real valued it is much easierto handle than the real time kernel, that is complex valued. Bysimply approximating N to be finite, equations 2.41 and 2.42 becomeaccessible to numerical integration. Indeed two molecular simulationmethods based immediately upon this is path integral Monte Carlo(PIMC)33 and path integral molecular dynamics (PIMD)34, whichrespectively use Monte Carlo and molecular dynamics to sample theintegrand.

There are also methods for quantum dynamics that use thisimaginary time path integral, such as centroid molecular dynamics(CMD)35 and ring polymer molecular dynamics (RPMD)36. Thesedynamical methods will be explained somewhat in section 2.4.

Linearized path integral

The classical Wigner method can be derived from the real time pathintegral. This was first published by Shi and Geva21 and somewhatlater by Poulsen, Nyman, and Rossky.22

Starting with a matrix element of an operator at time t,

⟨x0

∣∣∣Ω(t)∣∣∣x′

0

⟩=

⟨x0

∣∣∣e iHt Ω e−

iHt

∣∣∣x′0

⟩, (2.43)

and inserting unity operators,

⟨x0

∣∣∣Ω(t)∣∣∣x′

0

⟩=

∫∫dDxN dDx′

N

⟨x0

∣∣∣e iHt

∣∣∣xN

×⟨xN

∣∣∣Ω∣∣∣x′

N

⟩⟨x′N

∣∣∣e− iHt

∣∣∣x′0

⟩, (2.44)

a product of two kernels and a matrix element at time 0 is obtained.The interesting aspect here is the kernels. Drawing on equation 2.34,

27

Page 49: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

Figure 2.1: Illustration of a classical particle (1.) and its correspond-ing imaginary time path integral ring polymer (2.). The zigzag linesrepresent the spring terms.

xN

Z = Tre−βH

=

∫dDxN

⟨xN

∣∣∣e−βH∣∣∣xN

= limN→∞

(N

2πβ

)ND2

ND (det (m))N2

N∏

j=1

∫dDxj

× e− β

N

∑Nj=1

(N2

2β22 (m(x(j mod N)+1−xj))•(x(j mod N)+1−xj)+V (xj)),

(2.42)

where Tr denotes a trace and mod is the modulus operator.

The extended phase space of N copies of the system connectedby harmonic springs is often seen as turning each single particle intoa “polymer” with N monomers. The circular path integral found inequation 2.42 is then called a “ring polymer”. This is schematicallyillustrated in figure 2.1.

26

2.2. Feynman path integral formulation of quantum mechanics

Since the imaginary time kernel is real valued it is much easierto handle than the real time kernel, that is complex valued. Bysimply approximating N to be finite, equations 2.41 and 2.42 becomeaccessible to numerical integration. Indeed two molecular simulationmethods based immediately upon this is path integral Monte Carlo(PIMC)33 and path integral molecular dynamics (PIMD)34, whichrespectively use Monte Carlo and molecular dynamics to sample theintegrand.

There are also methods for quantum dynamics that use thisimaginary time path integral, such as centroid molecular dynamics(CMD)35 and ring polymer molecular dynamics (RPMD)36. Thesedynamical methods will be explained somewhat in section 2.4.

Linearized path integral

The classical Wigner method can be derived from the real time pathintegral. This was first published by Shi and Geva21 and somewhatlater by Poulsen, Nyman, and Rossky.22

Starting with a matrix element of an operator at time t,

⟨x0

∣∣∣Ω(t)∣∣∣x′

0

⟩=

⟨x0

∣∣∣e iHt Ω e−

iHt

∣∣∣x′0

⟩, (2.43)

and inserting unity operators,

⟨x0

∣∣∣Ω(t)∣∣∣x′

0

⟩=

∫∫dDxN dDx′

N

⟨x0

∣∣∣e iHt

∣∣∣xN

×⟨xN

∣∣∣Ω∣∣∣x′

N

⟩⟨x′N

∣∣∣e− iHt

∣∣∣x′0

⟩, (2.44)

a product of two kernels and a matrix element at time 0 is obtained.The interesting aspect here is the kernels. Drawing on equation 2.34,

27

Page 50: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

the product of the kernels can be written

⟨x0

∣∣∣e iHt

∣∣∣xN

⟩⟨x′N

∣∣∣e− iHt

∣∣∣x′0

= limN→∞

(N

2π(−i)t

)ND2

(det (m))N2

N−1∏

j=1

∫dDxj

× e− it

N∑N−1

j=0

(N2

t212(m(xj+1−xj))•(xj+1−xj)−V (xj)

)

×(

N

2πit

)ND2

(det (m))N2

N−1∏

j=1

∫dDx′

j

× eitN

∑N−1j=0

(N2

t212(m(x′

j+1−x′j))•(x′

j+1−x′j)−V (x′

j))

= limN→∞

(det (m)

(2π∆t)D

)N

N−1∏

j=1

∫∫dDxj d

Dx′j

× e− i

∆t∑N−1

j=0

(12

(m(xj+1−xj))•(xj+1−xj)∆t2

−V (xj)

)

× ei∆t

∑N−1j=0

(12

(m(x′j+1−x′

j))•(x′j+1−x′

j)∆t2

−V (x′j)

)

, (2.45)

where it is assumed that the same N is used for both kernels andt

N= ∆t. The coordinates can be changed to xj =

xj + x′j

2and

28

2.2. Feynman path integral formulation of quantum mechanics

∆x = x′j −xj (The Jacobian can be found in equation 2.14.), giving

⟨x0

∣∣∣e iHt

∣∣∣xN

⟩⟨x′N

∣∣∣e− iHt

∣∣∣x′0

= limN→∞

(det (m)

(2π∆t)D

)N

N−1∏

j=1

∫∫dDxj d

D∆xj

× e− i

∆t∑N−1

j=0

1

2

(m

(xj+1−

∆xj+12 −xj+

∆xj2

))•(xj+1−

∆xj+12 −xj+

∆xj2

)

∆t2

× e− i

∆t∑N−1

j=0

(−V

(xj−

∆xj2

))

× e

i∆t

∑N−1j=0

1

2

(m

(xj+1+

∆xj+12 −xj−

∆xj2

))•(xj+1+

∆xj+12 −xj−

∆xj2

)

∆t2

× ei∆t

∑N−1j=0

(−V

(xj+

∆xj2

))

= limN→∞

(det (m)

(2π∆t)D

)N

N−1∏

j=1

∫∫dDxj d

D∆xj

× e

− i∆t

∑N−1j=0

1

2

∑Dj′′=1

mj′′(xj′′,j+1−

∆xj′′,j+12 −xj′′,j+

∆xj′′,j2

)2

∆t2

× e− i

∆t∑N−1

j=0

(−V

(xj−

∆xj2

))

× e

i∆t

∑N−1j=0

1

2

∑Dj′′=1

mj′′(xj′′,j+1+

∆xj′′,j+12 −xj′′,j−

∆xj′′,j2

)2

∆t2

× ei∆t

∑N−1j=0

(−V

(xj+

∆xj2

)), (2.46)

where xj′′,j and ∆xj′′,j are the elements of the vectors xj and ∆xj.

29

Page 51: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

the product of the kernels can be written

⟨x0

∣∣∣e iHt

∣∣∣xN

⟩⟨x′N

∣∣∣e− iHt

∣∣∣x′0

= limN→∞

(N

2π(−i)t

)ND2

(det (m))N2

N−1∏

j=1

∫dDxj

× e− it

N∑N−1

j=0

(N2

t212(m(xj+1−xj))•(xj+1−xj)−V (xj)

)

×(

N

2πit

)ND2

(det (m))N2

N−1∏

j=1

∫dDx′

j

× eitN

∑N−1j=0

(N2

t212(m(x′

j+1−x′j))•(x′

j+1−x′j)−V (x′

j))

= limN→∞

(det (m)

(2π∆t)D

)N

N−1∏

j=1

∫∫dDxj d

Dx′j

× e− i

∆t∑N−1

j=0

(12

(m(xj+1−xj))•(xj+1−xj)∆t2

−V (xj)

)

× ei∆t

∑N−1j=0

(12

(m(x′j+1−x′

j))•(x′j+1−x′

j)∆t2

−V (x′j)

)

, (2.45)

where it is assumed that the same N is used for both kernels andt

N= ∆t. The coordinates can be changed to xj =

xj + x′j

2and

28

2.2. Feynman path integral formulation of quantum mechanics

∆x = x′j −xj (The Jacobian can be found in equation 2.14.), giving

⟨x0

∣∣∣e iHt

∣∣∣xN

⟩⟨x′N

∣∣∣e− iHt

∣∣∣x′0

= limN→∞

(det (m)

(2π∆t)D

)N

N−1∏

j=1

∫∫dDxj d

D∆xj

× e− i

∆t∑N−1

j=0

1

2

(m

(xj+1−

∆xj+12 −xj+

∆xj2

))•(xj+1−

∆xj+12 −xj+

∆xj2

)

∆t2

× e− i

∆t∑N−1

j=0

(−V

(xj−

∆xj2

))

× e

i∆t

∑N−1j=0

1

2

(m

(xj+1+

∆xj+12 −xj−

∆xj2

))•(xj+1+

∆xj+12 −xj−

∆xj2

)

∆t2

× ei∆t

∑N−1j=0

(−V

(xj+

∆xj2

))

= limN→∞

(det (m)

(2π∆t)D

)N

N−1∏

j=1

∫∫dDxj d

D∆xj

× e

− i∆t

∑N−1j=0

1

2

∑Dj′′=1

mj′′(xj′′,j+1−

∆xj′′,j+12 −xj′′,j+

∆xj′′,j2

)2

∆t2

× e− i

∆t∑N−1

j=0

(−V

(xj−

∆xj2

))

× e

i∆t

∑N−1j=0

1

2

∑Dj′′=1

mj′′(xj′′,j+1+

∆xj′′,j+12 −xj′′,j−

∆xj′′,j2

)2

∆t2

× ei∆t

∑N−1j=0

(−V

(xj+

∆xj2

)), (2.46)

where xj′′,j and ∆xj′′,j are the elements of the vectors xj and ∆xj.

29

Page 52: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

The kinetic energies can be further expanded and simplified,

⟨x0

∣∣∣e iHt

∣∣∣xN

⟩⟨x′N

∣∣∣e− iHt

∣∣∣x′0

= limN→∞

(det (m)

(2π∆t)D

)N

N−1∏

j=1

∫∫dDxj d

D∆xj

× exp

− i

∆t∑N−1

j=012

∑Dj′′=1

mj′′(x2j′′,j+1−xj′′,j+1∆xj′′,j+1−2xj′′,j+1xj′′,j+xj′′,j+1∆xj′′,j)

∆t2

+

mj′′

∆x2j′′,j+14 +xj′′,j∆xj′′,j+1−

∆xj′′,j+1∆xj′′,j2

∆t2

+

mj′′

x2

j′′,j−xj′′,j∆xj′′,j+∆x2

j′′,j4

∆t2

× exp

i

∆t∑N−1

j=012

∑Dj′′=1

mj′′(x2j′′,j+1+xj′′,j+1∆xj′′,j+1−2xj′′,j+1xj′′,j−xj′′,j+1∆xj′′,j)

∆t2

+

mj′′

∆x2j′′,j+14 −xj′′,j∆xj′′,j+1−

∆xj′′,j+1∆xj′′,j2

∆t2

+

mj′′

x2

j′′,j+xj′′,j∆xj′′,j+∆x2

j′′,j4

∆t2

× ei∆t

∑N−1j=0

(V(xj−

∆xj2

)−V

(xj+

∆xj2

)), (2.47)

30

2.2. Feynman path integral formulation of quantum mechanics

⟨x0

∣∣∣e iHt

∣∣∣xN

⟩⟨x′N

∣∣∣e− iHt

∣∣∣x′0

= limN→∞

(det (m)

(2π∆t)D

)N

N−1∏

j=1

∫∫dDxj d

D∆xj

× ei∆t

∑N−1j=0

(12

∑Dj′′=1

mj′′(2xj′′,j+1∆xj′′,j+1−2xj′′,j+1∆xj′′,j)∆t2

)

× ei∆t

∑N−1j=0

(12

∑Dj′′=1

mj′′(−2xj′′,j∆xj′′,j+1+2xj′′,j∆xj′′,j)∆t2

)

× ei∆t

∑N−1j=0

(V(xj−

∆xj2

)−V

(xj+

∆xj2

))

= limN→∞

(det (m)

(2π∆t)D

)N

N−1∏

j=1

∫∫dDxj d

D∆xj

× ei∆t

∑N−1j=0

((m(xj+1−xj))•∆xj+1

∆t2−(m(xj+1−xj))•∆xj

∆t2

)

× ei∆t

∑N−1j=0

(V(xj−

∆xj2

)−V

(xj+

∆xj2

))

= limN→∞

(det (m)

(2π∆t)D

)N

N−1∏

j=1

∫∫dDxj d

D∆xj

× ei∆t

∑N−1j=1

((m(xj−xj−1))•∆xj

∆t2−(m(xj+1−xj))•∆xj

∆t2

)

× ei∆t

(− (m(x1−x0))•∆x0

∆t2+(m(xN−xN−1))•∆xN

∆t2

)

× ei∆t

∑N−1j=0

(V(xj−

∆xj2

)−V

(xj+

∆xj2

)). (2.48)

The potential energy can then be Taylor-expanded around ∆xj = 0,

V

(xj −

∆xj

2

)− V

(xj +

∆xj

2

)

= −(

dV (x)

dx

)T

x=xj

•∆xj

− 1

24

D∑

j′′=1

(d3V (x)

dx3

)

x=xj′′,j

(∆xj′′,j)3 − . . . . (2.49)

31

Page 53: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

The kinetic energies can be further expanded and simplified,

⟨x0

∣∣∣e iHt

∣∣∣xN

⟩⟨x′N

∣∣∣e− iHt

∣∣∣x′0

= limN→∞

(det (m)

(2π∆t)D

)N

N−1∏

j=1

∫∫dDxj d

D∆xj

× exp

− i

∆t∑N−1

j=012

∑Dj′′=1

mj′′(x2j′′,j+1−xj′′,j+1∆xj′′,j+1−2xj′′,j+1xj′′,j+xj′′,j+1∆xj′′,j)

∆t2

+

mj′′

∆x2j′′,j+14 +xj′′,j∆xj′′,j+1−

∆xj′′,j+1∆xj′′,j2

∆t2

+

mj′′

x2

j′′,j−xj′′,j∆xj′′,j+∆x2

j′′,j4

∆t2

× exp

i

∆t∑N−1

j=012

∑Dj′′=1

mj′′(x2j′′,j+1+xj′′,j+1∆xj′′,j+1−2xj′′,j+1xj′′,j−xj′′,j+1∆xj′′,j)

∆t2

+

mj′′

∆x2j′′,j+14 −xj′′,j∆xj′′,j+1−

∆xj′′,j+1∆xj′′,j2

∆t2

+

mj′′

x2

j′′,j+xj′′,j∆xj′′,j+∆x2

j′′,j4

∆t2

× ei∆t

∑N−1j=0

(V(xj−

∆xj2

)−V

(xj+

∆xj2

)), (2.47)

30

2.2. Feynman path integral formulation of quantum mechanics

⟨x0

∣∣∣e iHt

∣∣∣xN

⟩⟨x′N

∣∣∣e− iHt

∣∣∣x′0

= limN→∞

(det (m)

(2π∆t)D

)N

N−1∏

j=1

∫∫dDxj d

D∆xj

× ei∆t

∑N−1j=0

(12

∑Dj′′=1

mj′′(2xj′′,j+1∆xj′′,j+1−2xj′′,j+1∆xj′′,j)∆t2

)

× ei∆t

∑N−1j=0

(12

∑Dj′′=1

mj′′(−2xj′′,j∆xj′′,j+1+2xj′′,j∆xj′′,j)∆t2

)

× ei∆t

∑N−1j=0

(V(xj−

∆xj2

)−V

(xj+

∆xj2

))

= limN→∞

(det (m)

(2π∆t)D

)N

N−1∏

j=1

∫∫dDxj d

D∆xj

× ei∆t

∑N−1j=0

((m(xj+1−xj))•∆xj+1

∆t2−(m(xj+1−xj))•∆xj

∆t2

)

× ei∆t

∑N−1j=0

(V(xj−

∆xj2

)−V

(xj+

∆xj2

))

= limN→∞

(det (m)

(2π∆t)D

)N

N−1∏

j=1

∫∫dDxj d

D∆xj

× ei∆t

∑N−1j=1

((m(xj−xj−1))•∆xj

∆t2−(m(xj+1−xj))•∆xj

∆t2

)

× ei∆t

(− (m(x1−x0))•∆x0

∆t2+(m(xN−xN−1))•∆xN

∆t2

)

× ei∆t

∑N−1j=0

(V(xj−

∆xj2

)−V

(xj+

∆xj2

)). (2.48)

The potential energy can then be Taylor-expanded around ∆xj = 0,

V

(xj −

∆xj

2

)− V

(xj +

∆xj

2

)

= −(

dV (x)

dx

)T

x=xj

•∆xj

− 1

24

D∑

j′′=1

(d3V (x)

dx3

)

x=xj′′,j

(∆xj′′,j)3 − . . . . (2.49)

31

Page 54: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

T denote the transpose of a vector. ∆xj is the separation betweenthe path the system will take going forward in time from 0 to t andwhen going backward in time from t to 0. In the limit of classicalmechanics the path forward in time and the path backward in timewill be the same, so we are going to approximate ∆xj as being smalland truncate the Taylor expansion after the linear term, i.e.

⟨x0

∣∣∣e iHt

∣∣∣xN

⟩⟨x′N

∣∣∣e− iHt

∣∣∣x′0

≈ limN→∞

(det (m)

(2π∆t)D

)N

N−1∏

j=1

∫∫dDxj d

D∆xj

× ei∆t

∑N−1j=1

((m(xj−xj−1))•∆xj

∆t2−(m(xj+1−xj))•∆xj

∆t2

)

× ei∆t

(− (m(x1−x0))•∆x0

∆t2+(m(xN−xN−1))•∆xN

∆t2

)

× ei∆t

∑N−1j=0

(−( dV (x)

dx )T

x=xj•∆xj

)

. (2.50)

The time-derivative ¨xj =xj+1−2xj+xj−1

(∆t)2can be used to simplify the

expression,⟨x0

∣∣∣e iHt

∣∣∣xN

⟩⟨x′N

∣∣∣e− iHt

∣∣∣x′0

≈ limN→∞

(det (m)

(2π∆t)D

)N

N−1∏

j=1

∫∫dDxj d

D∆xj

× ei∆t

∑N−1j=1 (−(m ¨xj)•∆xj)

× ei∆t

(− (m(x1−x0))•∆x0

(∆t)2+(m(xN−xN−1))•∆xN

(∆t)2

)

× ei∆t

∑N−1j=0

(−( dV (x)

dx )T

x=xj•∆xj

)

= limN→∞

(det (m)

(2π∆t)D

)N

N−1∏

j=1

∫∫dDxj d

D∆xj

× ei∆t

∑N−1j=1

(−m ¨xj−( dV (x)

dx )T

x=xj

)•∆xj

× ei∆t

(− (m(x1−x0))•∆x0

(∆t)2+(m(xN−xN−1))•∆xN

(∆t)2−( dV (x)

dx )T

x=x0•∆x0

)

.(2.51)

32

2.2. Feynman path integral formulation of quantum mechanics

Integrating over ∆xj then leads to delta functions,

⟨x0

∣∣∣e iHt

∣∣∣xN

⟩⟨x′N

∣∣∣e− iHt

∣∣∣x′0

≈ limN→∞

(det (m)

(∆t)D

)N

(2π)−D

N−1∏

j=1

∫dDxj

×(

N−1∏

j=1

δ

(∆t

(−m ¨xj −

(dV (x)

dx

)T

x=xj

)))

× ei∆t

(− (m(x1−x0))•∆x0

(∆t)2+(m(xN−xN−1))•∆xN

(∆t)2−( dV (x)

dx )T

x=x0•∆x0

)

= limN→∞

(det (m)

(∆t)2D

)N (∆t

)D

N−1∏

j=1

∫dDxj

×(

N−1∏

j=1

δ

(−m ¨xj −

(dV (x)

dx

)T

x=xj

))

× ei∆t

(− (m(x1−x0))•∆x0

(∆t)2+(m(xN−xN−1))•∆xN

(∆t)2−( dV (x)

dx )T

x=x0•∆x0

)

.(2.52)

Returning to equation 2.44 we have

⟨x0

∣∣∣Ω(t)∣∣∣x′

0

⟩=

⟨x0 −

∆x0

2

∣∣∣∣Ω(t)

∣∣∣∣x0 +∆x0

2

≈ limN→∞

∫∫dDxN dD∆xN

⟨xN − ∆xN

2

∣∣∣∣Ω∣∣∣∣xN +

∆xN

2

×(det (m)

(∆t)2D

)N (∆t

)D

N−1∏

j=1

∫dDxj

×(

N−1∏

j=1

δ

(−m ¨xj −

(dV (x)

dx

)T

x=xj

))

× ei∆t

(− (m(x1−x0))•∆x0

(∆t)2+(m(xN−xN−1))•∆xN

(∆t)2−( dV (x)

dx )T

x=x0•∆x0

)

.(2.53)

33

Page 55: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

T denote the transpose of a vector. ∆xj is the separation betweenthe path the system will take going forward in time from 0 to t andwhen going backward in time from t to 0. In the limit of classicalmechanics the path forward in time and the path backward in timewill be the same, so we are going to approximate ∆xj as being smalland truncate the Taylor expansion after the linear term, i.e.

⟨x0

∣∣∣e iHt

∣∣∣xN

⟩⟨x′N

∣∣∣e− iHt

∣∣∣x′0

≈ limN→∞

(det (m)

(2π∆t)D

)N

N−1∏

j=1

∫∫dDxj d

D∆xj

× ei∆t

∑N−1j=1

((m(xj−xj−1))•∆xj

∆t2−(m(xj+1−xj))•∆xj

∆t2

)

× ei∆t

(− (m(x1−x0))•∆x0

∆t2+(m(xN−xN−1))•∆xN

∆t2

)

× ei∆t

∑N−1j=0

(−( dV (x)

dx )T

x=xj•∆xj

)

. (2.50)

The time-derivative ¨xj =xj+1−2xj+xj−1

(∆t)2can be used to simplify the

expression,⟨x0

∣∣∣e iHt

∣∣∣xN

⟩⟨x′N

∣∣∣e− iHt

∣∣∣x′0

≈ limN→∞

(det (m)

(2π∆t)D

)N

N−1∏

j=1

∫∫dDxj d

D∆xj

× ei∆t

∑N−1j=1 (−(m ¨xj)•∆xj)

× ei∆t

(− (m(x1−x0))•∆x0

(∆t)2+(m(xN−xN−1))•∆xN

(∆t)2

)

× ei∆t

∑N−1j=0

(−( dV (x)

dx )T

x=xj•∆xj

)

= limN→∞

(det (m)

(2π∆t)D

)N

N−1∏

j=1

∫∫dDxj d

D∆xj

× ei∆t

∑N−1j=1

(−m ¨xj−( dV (x)

dx )T

x=xj

)•∆xj

× ei∆t

(− (m(x1−x0))•∆x0

(∆t)2+(m(xN−xN−1))•∆xN

(∆t)2−( dV (x)

dx )T

x=x0•∆x0

)

.(2.51)

32

2.2. Feynman path integral formulation of quantum mechanics

Integrating over ∆xj then leads to delta functions,

⟨x0

∣∣∣e iHt

∣∣∣xN

⟩⟨x′N

∣∣∣e− iHt

∣∣∣x′0

≈ limN→∞

(det (m)

(∆t)D

)N

(2π)−D

N−1∏

j=1

∫dDxj

×(

N−1∏

j=1

δ

(∆t

(−m ¨xj −

(dV (x)

dx

)T

x=xj

)))

× ei∆t

(− (m(x1−x0))•∆x0

(∆t)2+(m(xN−xN−1))•∆xN

(∆t)2−( dV (x)

dx )T

x=x0•∆x0

)

= limN→∞

(det (m)

(∆t)2D

)N (∆t

)D

N−1∏

j=1

∫dDxj

×(

N−1∏

j=1

δ

(−m ¨xj −

(dV (x)

dx

)T

x=xj

))

× ei∆t

(− (m(x1−x0))•∆x0

(∆t)2+(m(xN−xN−1))•∆xN

(∆t)2−( dV (x)

dx )T

x=x0•∆x0

)

.(2.52)

Returning to equation 2.44 we have

⟨x0

∣∣∣Ω(t)∣∣∣x′

0

⟩=

⟨x0 −

∆x0

2

∣∣∣∣Ω(t)

∣∣∣∣x0 +∆x0

2

≈ limN→∞

∫∫dDxN dD∆xN

⟨xN − ∆xN

2

∣∣∣∣Ω∣∣∣∣xN +

∆xN

2

×(det (m)

(∆t)2D

)N (∆t

)D

N−1∏

j=1

∫dDxj

×(

N−1∏

j=1

δ

(−m ¨xj −

(dV (x)

dx

)T

x=xj

))

× ei∆t

(− (m(x1−x0))•∆x0

(∆t)2+(m(xN−xN−1))•∆xN

(∆t)2−( dV (x)

dx )T

x=x0•∆x0

)

.(2.53)

33

Page 56: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

Since

[Ω]W

(xN ,

m (xN − xN−1)

∆t

)

=

∫dD∆xN e

i(m(xN−xN−1))•∆xN

∆t

×⟨xN − ∆xN

2

∣∣∣∣Ω∣∣∣∣xN +

∆xN

2

⟩(2.54)

equation 2.53 can be written as

⟨x0 −

∆x0

2

∣∣∣∣Ω(t)

∣∣∣∣x0 +∆x0

2

≈ limN→∞

∫dDxN

[Ω]W

(xN ,

m (xN − xN−1)

∆t

)

×(det (m)

(∆t)2D

)N (∆t

)D

N−1∏

j=1

∫dDxj

×(

N−1∏

j=1

δ

(−m ¨xj −

(dV (x)

dx

)T

x=xj

))

× ei∆t

(− (m(x1−x0))•∆x0

(∆t)2−( dV (x)

dx )T

x=x0•∆x0

)

. (2.55)

Defining the momentum vectors pN = m(xN−xN−1)∆t

and p0 =m(x0−x−1)

∆t,

and thus also the position x−1 at time −∆t, and multiplying both

34

2.2. Feynman path integral formulation of quantum mechanics

sides of the “≈” with ei p0•∆x0 gives

ei p0•∆x0

⟨x0 −

∆x0

2

∣∣∣∣Ω(t)

∣∣∣∣x0 +∆x0

2

≈ limN→∞

(det (m)

(∆t)2D

)N (∆t

)D

N∏

j=1

∫dDxj

×(

N−1∏

j=1

δ

(−m ¨xj −

(dV (x)

dx

)T

x=xj

))

×[Ω]W(xN , pN)

× ei∆t

(p0•∆x0

∆t− (m(x1−x0))•∆x0

(∆t)2−( dV (x)

dx )T

x=x0•∆x0

)

= limN→∞

(det (m)

(∆t)2D

)N (∆t

)D

N∏

j=1

∫dDxj

×(

N−1∏

j=1

δ

(−m ¨xj −

(dV (x)

dx

)T

x=xj

))

×[Ω]W(xN , pN)

× ei∆t

(− (m(x1−2x0+x−1))•∆x0

(∆t)2−( dV (x)

dx )T

x=x0•∆x0

)

= limN→∞

(det (m)

(∆t)2D

)N (∆t

)D

N∏

j=1

∫dDxj

×(

N−1∏

j=1

δ

(−m ¨xj −

(dV (x)

dx

)T

x=xj

))

×[Ω]W(xN , pN) e

i∆t

(−(m ¨x0)•∆x0−( dV (x)

dx )T

x=x0•∆x0

)

.

(2.56)

35

Page 57: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

Since

[Ω]W

(xN ,

m (xN − xN−1)

∆t

)

=

∫dD∆xN e

i(m(xN−xN−1))•∆xN

∆t

×⟨xN − ∆xN

2

∣∣∣∣Ω∣∣∣∣xN +

∆xN

2

⟩(2.54)

equation 2.53 can be written as

⟨x0 −

∆x0

2

∣∣∣∣Ω(t)

∣∣∣∣x0 +∆x0

2

≈ limN→∞

∫dDxN

[Ω]W

(xN ,

m (xN − xN−1)

∆t

)

×(det (m)

(∆t)2D

)N (∆t

)D

N−1∏

j=1

∫dDxj

×(

N−1∏

j=1

δ

(−m ¨xj −

(dV (x)

dx

)T

x=xj

))

× ei∆t

(− (m(x1−x0))•∆x0

(∆t)2−( dV (x)

dx )T

x=x0•∆x0

)

. (2.55)

Defining the momentum vectors pN = m(xN−xN−1)∆t

and p0 =m(x0−x−1)

∆t,

and thus also the position x−1 at time −∆t, and multiplying both

34

2.2. Feynman path integral formulation of quantum mechanics

sides of the “≈” with ei p0•∆x0 gives

ei p0•∆x0

⟨x0 −

∆x0

2

∣∣∣∣Ω(t)

∣∣∣∣x0 +∆x0

2

≈ limN→∞

(det (m)

(∆t)2D

)N (∆t

)D

N∏

j=1

∫dDxj

×(

N−1∏

j=1

δ

(−m ¨xj −

(dV (x)

dx

)T

x=xj

))

×[Ω]W(xN , pN)

× ei∆t

(p0•∆x0

∆t− (m(x1−x0))•∆x0

(∆t)2−( dV (x)

dx )T

x=x0•∆x0

)

= limN→∞

(det (m)

(∆t)2D

)N (∆t

)D

N∏

j=1

∫dDxj

×(

N−1∏

j=1

δ

(−m ¨xj −

(dV (x)

dx

)T

x=xj

))

×[Ω]W(xN , pN)

× ei∆t

(− (m(x1−2x0+x−1))•∆x0

(∆t)2−( dV (x)

dx )T

x=x0•∆x0

)

= limN→∞

(det (m)

(∆t)2D

)N (∆t

)D

N∏

j=1

∫dDxj

×(

N−1∏

j=1

δ

(−m ¨xj −

(dV (x)

dx

)T

x=xj

))

×[Ω]W(xN , pN) e

i∆t

(−(m ¨x0)•∆x0−( dV (x)

dx )T

x=x0•∆x0

)

.

(2.56)

35

Page 58: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

Integrating over ∆x0 on both sides of the “≈” leads to

∫dD∆x0 e

i p0•∆x0

⟨x0 −

∆x0

2

∣∣∣∣Ω(t)

∣∣∣∣x0 +∆x0

2

=[Ω(t)

]W(x0, p0)

≈ limN→∞

(det (m)

(∆t)2D

)N (∆t

)D

N∏

j=1

∫dDxj

×(

N−1∏

j=1

δ

(−m ¨xj −

(dV (x)

dx

)T

x=xj

))

×[Ω]W(xN , pN)

×∫

dD∆x0 ei∆t

(−(m ¨x0)•∆x0−( dV (x)

dx )T

x=x0•∆x0

)

= limN→∞

(det (m)

(∆t)2D

)N (∆t

)D

N∏

j=1

∫dDxj

×(

N−1∏

j=1

δ

(−m ¨xj −

(dV (x)

dx

)T

x=xj

))

×[Ω]W(xN , pN)

× δ

(∆t

(−m ¨x0 −

(dV (x)

dx

)T

x=x0

))

= limN→∞

(det (m)

(∆t)2D

)N

N∏

j=1

∫dDxj

×(

N−1∏

j=0

δ

(−m ¨xj −

(dV (x)

dx

)T

x=xj

))

×[Ω]W(xN , pN) . (2.57)

The delta functions restrict the path from x0 at time 0 to xN attime t to m ¨xj=−( dV (x)

dx )T

x=xj. This is Newton’s second law of motion.

The factor(

det(m)

(∆t)2D

)N

is a normalization to account for the fact that

36

2.2. Feynman path integral formulation of quantum mechanics

the dimensions of dDxj and −m ¨xj−( dV (x)dx )

T

x=xjdo not agree. Thus

[Ω(t)

]W(x0, p0)

≈ limN→∞

(det (m)

(∆t)2D

)N

N∏

j=1

∫dDxj

×(

N−1∏

j=0

δ

(−m ¨xj −

(dV (x)

dx

)T

x=xj

))

×[Ω]W(xN , pN) (2.58)

is the classical Wigner method. The only approximation introducedin this derivation is the truncation, after the linear term, of theTaylor expansion of the potential energy in equation 2.49. Thistruncation does of course not affect systems with all higher orderderivatives being zero. The classical Wigner method is thus provento be exact for e.g. bilinearly coupled harmonic oscillators.

This is a linearization of a path integral and is therefore calledlinearized path integral (LPI).

It can be noted that the LPI derivation presented here avoidsthe use of lim∆t→0 ∆t( dV (x)

dx )T

x=x0•∆x0→0 that both Shi and Geva21 and

Poulsen, Nyman, and Rossky22 use. As no other factor of ∆t, in thederivation, is assumed to be 0 this could be seen as an advantage.

37

Page 59: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

Integrating over ∆x0 on both sides of the “≈” leads to

∫dD∆x0 e

i p0•∆x0

⟨x0 −

∆x0

2

∣∣∣∣Ω(t)

∣∣∣∣x0 +∆x0

2

=[Ω(t)

]W(x0, p0)

≈ limN→∞

(det (m)

(∆t)2D

)N (∆t

)D

N∏

j=1

∫dDxj

×(

N−1∏

j=1

δ

(−m ¨xj −

(dV (x)

dx

)T

x=xj

))

×[Ω]W(xN , pN)

×∫

dD∆x0 ei∆t

(−(m ¨x0)•∆x0−( dV (x)

dx )T

x=x0•∆x0

)

= limN→∞

(det (m)

(∆t)2D

)N (∆t

)D

N∏

j=1

∫dDxj

×(

N−1∏

j=1

δ

(−m ¨xj −

(dV (x)

dx

)T

x=xj

))

×[Ω]W(xN , pN)

× δ

(∆t

(−m ¨x0 −

(dV (x)

dx

)T

x=x0

))

= limN→∞

(det (m)

(∆t)2D

)N

N∏

j=1

∫dDxj

×(

N−1∏

j=0

δ

(−m ¨xj −

(dV (x)

dx

)T

x=xj

))

×[Ω]W(xN , pN) . (2.57)

The delta functions restrict the path from x0 at time 0 to xN attime t to m ¨xj=−( dV (x)

dx )T

x=xj. This is Newton’s second law of motion.

The factor(

det(m)

(∆t)2D

)N

is a normalization to account for the fact that

36

2.2. Feynman path integral formulation of quantum mechanics

the dimensions of dDxj and −m ¨xj−( dV (x)dx )

T

x=xjdo not agree. Thus

[Ω(t)

]W(x0, p0)

≈ limN→∞

(det (m)

(∆t)2D

)N

N∏

j=1

∫dDxj

×(

N−1∏

j=0

δ

(−m ¨xj −

(dV (x)

dx

)T

x=xj

))

×[Ω]W(xN , pN) (2.58)

is the classical Wigner method. The only approximation introducedin this derivation is the truncation, after the linear term, of theTaylor expansion of the potential energy in equation 2.49. Thistruncation does of course not affect systems with all higher orderderivatives being zero. The classical Wigner method is thus provento be exact for e.g. bilinearly coupled harmonic oscillators.

This is a linearization of a path integral and is therefore calledlinearized path integral (LPI).

It can be noted that the LPI derivation presented here avoidsthe use of lim∆t→0 ∆t( dV (x)

dx )T

x=x0•∆x0→0 that both Shi and Geva21 and

Poulsen, Nyman, and Rossky22 use. As no other factor of ∆t, in thederivation, is assumed to be 0 this could be seen as an advantage.

37

Page 60: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

2.3 Chemical kinetics and thermal rate

constants

Kinetics is at the heart of physicalchemistry, [...]

G. Kristin Jonsson(Lab.-report in kinetics, 2011)

In chemistry a chemical reaction such as

A + B → C + D (2.a)

has a rate ofd[C]

dt, with the brackets denoting the concentration

of the enclosed species

(d[C]

dt=

d[D]

dt= − d[A]

dt= − d[B]

dt

). For

the simple kind of bimolecular (or more generally bispecies) reactionshown here the reaction rate can be expressed as

d[C]

dt= kr(T )[A][B], (2.59)

where kr(T ) is called the reaction rate constant. Miller and cowork-ers37,38 have shown how to calculate bimolecular reaction rate con-stants from three different traces:

kr(T )

=1

QR

limt→∞

d

dtTr

θ (s− xr) e

−βH eiHt θ (xr − s) e−

iHt

(2.60)

=1

QR

limt→∞

TrF (xr − s) e−βH e

iHt θ (xr − s) e−

iHt

(2.61)

=1

QR

∫ ∞

0

dt TrF (xr − s) e−βH e

iHt F (xr − s) e−

iHt

=1

QR

1

2

∫ ∞

−∞dt Tr

F (xr − s) e−βH e

iHt F (xr − s) e−

iHt

,

(2.62)

38

2.3. Chemical kinetics and thermal rate constants

where QR is the canonical partition function of the reactants, xr =xr (x) is the reaction coordinate operator, x is the position operator,s is the position of a dividing surface that separates reactants andproducts, F (x) is the probability flux operator, and θ(x) is theHeaviside step function.

The probability flux operator is

F (x− s) =1

2m−1 (δ (x− s) p+ pδ (x− s)) . (2.63)

To elucidate the physical meaning of the probability flux operator itcan be pointed out that the Wigner transform of the operator is

[F (x− s)

]W(x, p) =

p

mδ(x− s) (2.64)

which simply is the velocity with position fixed in the surface s.For reaction rate calculations the dividing surface s could be

placed e.g. in the transition state, but this is not a requirement.The dividing surface can formally be placed anywhere, as long as itseparates reactants and products from each other. Certain choices,such as placing s in the transition state, may be beneficial fornumerical convergence or even a requirement if approximations tothe exact expressions are used. E.g., in the transition state theorylimit, t → 0, where it is assumed that no recrossings of the dividingsurface occurs it makes a big difference if the dividing surface isplaced on the top of a potential barrier or below it.

The expressions 2.60-2.62 are rather intuitive. The trace inequation 2.60 is proportional to the probability of being on thereactant side of the dividing surface at time t = 0 and later on theproduct side of the dividing surface at t → ∞.

The trace in equation 2.61 is proportional to the flow of theprobability density thought the dividing surface at time t = 0multiplied with the probability of ending up on the product side ofthe dividing surface at t → ∞.

The trace in equation 2.62 is proportional to the product of theflow of the probability density thought the dividing surface at thetwo times t = 0, or t → −∞, and t → ∞.

A requirement for the validity of equations 2.60-2.62 is that iffollowing the reaction coordinate far enough away from the dividing

39

Page 61: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

2.3 Chemical kinetics and thermal rate

constants

Kinetics is at the heart of physicalchemistry, [...]

G. Kristin Jonsson(Lab.-report in kinetics, 2011)

In chemistry a chemical reaction such as

A + B → C + D (2.a)

has a rate ofd[C]

dt, with the brackets denoting the concentration

of the enclosed species

(d[C]

dt=

d[D]

dt= − d[A]

dt= − d[B]

dt

). For

the simple kind of bimolecular (or more generally bispecies) reactionshown here the reaction rate can be expressed as

d[C]

dt= kr(T )[A][B], (2.59)

where kr(T ) is called the reaction rate constant. Miller and cowork-ers37,38 have shown how to calculate bimolecular reaction rate con-stants from three different traces:

kr(T )

=1

QR

limt→∞

d

dtTr

θ (s− xr) e

−βH eiHt θ (xr − s) e−

iHt

(2.60)

=1

QR

limt→∞

TrF (xr − s) e−βH e

iHt θ (xr − s) e−

iHt

(2.61)

=1

QR

∫ ∞

0

dt TrF (xr − s) e−βH e

iHt F (xr − s) e−

iHt

=1

QR

1

2

∫ ∞

−∞dt Tr

F (xr − s) e−βH e

iHt F (xr − s) e−

iHt

,

(2.62)

38

2.3. Chemical kinetics and thermal rate constants

where QR is the canonical partition function of the reactants, xr =xr (x) is the reaction coordinate operator, x is the position operator,s is the position of a dividing surface that separates reactants andproducts, F (x) is the probability flux operator, and θ(x) is theHeaviside step function.

The probability flux operator is

F (x− s) =1

2m−1 (δ (x− s) p+ pδ (x− s)) . (2.63)

To elucidate the physical meaning of the probability flux operator itcan be pointed out that the Wigner transform of the operator is

[F (x− s)

]W(x, p) =

p

mδ(x− s) (2.64)

which simply is the velocity with position fixed in the surface s.For reaction rate calculations the dividing surface s could be

placed e.g. in the transition state, but this is not a requirement.The dividing surface can formally be placed anywhere, as long as itseparates reactants and products from each other. Certain choices,such as placing s in the transition state, may be beneficial fornumerical convergence or even a requirement if approximations tothe exact expressions are used. E.g., in the transition state theorylimit, t → 0, where it is assumed that no recrossings of the dividingsurface occurs it makes a big difference if the dividing surface isplaced on the top of a potential barrier or below it.

The expressions 2.60-2.62 are rather intuitive. The trace inequation 2.60 is proportional to the probability of being on thereactant side of the dividing surface at time t = 0 and later on theproduct side of the dividing surface at t → ∞.

The trace in equation 2.61 is proportional to the flow of theprobability density thought the dividing surface at time t = 0multiplied with the probability of ending up on the product side ofthe dividing surface at t → ∞.

The trace in equation 2.62 is proportional to the product of theflow of the probability density thought the dividing surface at thetwo times t = 0, or t → −∞, and t → ∞.

A requirement for the validity of equations 2.60-2.62 is that iffollowing the reaction coordinate far enough away from the dividing

39

Page 62: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

surface there should not be any possibility of reflecting back atrajectory or wavepacket.

These traces, or more specificly the flux-Heaviside one in equation2.61, are used for the calculation of reaction rate constants in paperIV, and section 3.3.

Another, older and more common, formulation for the reactionrate constant is collision theory. This theory is only valid for bispeciesreactions of ideal gases. It assumes classical mechanics for thetranslational movement of chemical species.

kr(T ) =

∫ ∞

0

dE σ(E)vrel(E)2β32

√E

πe−βE

=

√8

πµA,B

β32

∫ ∞

0

dE σ(E)E e−βE, (2.65)

where E is the collision energy, σ(E) is the reaction cross section,

vrel(E) =√

2E/µ is the relative velocity of the colliding species,

2β32

√E/π e−βE is the Maxwell-Boltzmann distribution function of

the energy, and µA,B = mAmBmA+mB

is the reduced mass of the collidingspecies with masses mA and mB.

In the situation where the two species can be described as classicalhard spheres, then if species A was trying to get past a stationary Bthe collision cross section would be the area that B covers so thatthe center of A can not pass through it without colliding with B,see figure 2.2. For hard spheres the collision cross section is simplythe area of a circle with a radius that is the sum of the radii ofthe species. For non-spherical species the area they cover would ofcourse depend on relative orientation, but the collision cross sectionwould be an average over all orientations. The reaction cross sectionis similar to the collision cross section, but with the addition that acollision only counts if it leads to a reaction. In reality, it is not welldefined where the borders of a chemical species are, so the speciescould be considered to come into proximity of each other, ratherthan to collide.

Thus the integral over collision energies in equation 2.65 hasthree parts: the Maxwell-Boltzmann distribution function, that takesaccount of the probability density of each particular collision energy,

40

2.3. Chemical kinetics and thermal rate constants

Figure 2.2: Illustration of the collision cross section of the twospherical particles A and B. The dashed line shows the area aroundparticle B that the center of particle A, when moving in from theleft of the figure, can not pass through without the two particlescolliding.

the reaction cross section, that takes account of the probability ofthe species colliding and reacting, and the relative velocity, thatdetermines at what rate collisions can occur.

Collision theory is used for the calculation of the radiative asso-ciation rate constant in paper I, and section 3.1.

41

Page 63: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

surface there should not be any possibility of reflecting back atrajectory or wavepacket.

These traces, or more specificly the flux-Heaviside one in equation2.61, are used for the calculation of reaction rate constants in paperIV, and section 3.3.

Another, older and more common, formulation for the reactionrate constant is collision theory. This theory is only valid for bispeciesreactions of ideal gases. It assumes classical mechanics for thetranslational movement of chemical species.

kr(T ) =

∫ ∞

0

dE σ(E)vrel(E)2β32

√E

πe−βE

=

√8

πµA,B

β32

∫ ∞

0

dE σ(E)E e−βE, (2.65)

where E is the collision energy, σ(E) is the reaction cross section,

vrel(E) =√

2E/µ is the relative velocity of the colliding species,

2β32

√E/π e−βE is the Maxwell-Boltzmann distribution function of

the energy, and µA,B = mAmBmA+mB

is the reduced mass of the collidingspecies with masses mA and mB.

In the situation where the two species can be described as classicalhard spheres, then if species A was trying to get past a stationary Bthe collision cross section would be the area that B covers so thatthe center of A can not pass through it without colliding with B,see figure 2.2. For hard spheres the collision cross section is simplythe area of a circle with a radius that is the sum of the radii ofthe species. For non-spherical species the area they cover would ofcourse depend on relative orientation, but the collision cross sectionwould be an average over all orientations. The reaction cross sectionis similar to the collision cross section, but with the addition that acollision only counts if it leads to a reaction. In reality, it is not welldefined where the borders of a chemical species are, so the speciescould be considered to come into proximity of each other, ratherthan to collide.

Thus the integral over collision energies in equation 2.65 hasthree parts: the Maxwell-Boltzmann distribution function, that takesaccount of the probability density of each particular collision energy,

40

2.3. Chemical kinetics and thermal rate constants

Figure 2.2: Illustration of the collision cross section of the twospherical particles A and B. The dashed line shows the area aroundparticle B that the center of particle A, when moving in from theleft of the figure, can not pass through without the two particlescolliding.

the reaction cross section, that takes account of the probability ofthe species colliding and reacting, and the relative velocity, thatdetermines at what rate collisions can occur.

Collision theory is used for the calculation of the radiative asso-ciation rate constant in paper I, and section 3.1.

41

Page 64: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

2.4 Common methods of approximate

quantum dynamics

It should be stressed that nuclei are heavyenough that quantum effects are almostnegligible, i.e. they behave to a goodapproximation as classical particles. Indeed, ifnuclei showed significant quantum aspects, theconcept of molecular structure (i.e. differentconfigurations and conformations) would nothave any meaning, since the nuclei wouldsimply tunnel through barriers and end up inthe global minimum.

Frank Jensen39

The quote that starts this section is from a standard textbookin computational chemistry. As already stated in chapter 1 of thisthesis, and indeed also acknowledged by Jensen elsewhere in hisbook,39 there are however situations in which quantum mechanics isnecessary to accurately describe the behavior of the atomic nuclei.

Apart from the classical Wigner method, that has been exten-sively described in sections 2.1 and 2.2, there are many methodsfor approximate quantum dynamics. A few notable methods with arelation to the work presented in this thesis will be presented in thissection, so that they can provide a context and serve as a comparisonfor some of the new developments that are presented in chapter 3.

Time correlation functions are important to describe the dynam-ics of physical systems. Examples of such functions can be found inequations 2.60-2.62. Those correlation functions are used to calculatereaction rate constants. The more general formulation of a thermaltime correlation function of operator A at time 0 and operator Bat time t is†

⟨AB(t)

⟩=

1

ZTr

A e−βH e

iHt B e−

iHt

. (2.66)

†An even more general formulation of a time correlation function would haveany number of operators at different times.

42

2.4. Common methods of approximate quantum dynamics

The Boltzmann operator can be placed on either side of A. Thiswill only change the sign of the imaginary part,

1

ZTr

A e−βH e

iHt B e−

iHt

=

1

ZTr

e−βH A e

iHt B e−

iHt

∗,

(2.67)

where ∗ denotes complex conjugate.Some quantum dynamical methods work with Kubo-transformed

time correlation functions,40

⟨AB(t)

⟩Kubo

=1

ZTr

1

β

∫ β

0

dζ e−ζH A e−(β−ζ)H eiHt B e−

iHt

,

(2.68)

which have the benefit that the quantum correlation functions aremore similar to their classical counterparts than is the case for thestandard “physical” correlation functions.

Another alternative is symmetrized correlation functions,41

⟨AB(t)

⟩Sym

=1

ZTr

e−

β2H A e−

β2H e

iHt B e−

iHt

, (2.69)

which have the benefit that they are real instead of complex quanti-ties. The correlation functions, or technically just traces, presentedin section 2.3 are of this type.

The standard, Kubo-transformed, and symmetrized versionsof the same correlation function are equivalent in the sense thatthey contain the same information. Using the Fourier transform,[f(t)]F (ω), the different correlation functions are related as:40,41

[⟨AB(t)

⟩]F(ω) = e−

βω2

[⟨AB(t)

⟩Sym

]

F

(ω)

=βω

eβω −1

[⟨AB(t)

⟩Kubo

]F(ω) . (2.70)

These relations are only valid for exact quantum mechanics. Ifapproximations to quantum mechanics are used, then equation 2.70

43

Page 65: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

2.4 Common methods of approximate

quantum dynamics

It should be stressed that nuclei are heavyenough that quantum effects are almostnegligible, i.e. they behave to a goodapproximation as classical particles. Indeed, ifnuclei showed significant quantum aspects, theconcept of molecular structure (i.e. differentconfigurations and conformations) would nothave any meaning, since the nuclei wouldsimply tunnel through barriers and end up inthe global minimum.

Frank Jensen39

The quote that starts this section is from a standard textbookin computational chemistry. As already stated in chapter 1 of thisthesis, and indeed also acknowledged by Jensen elsewhere in hisbook,39 there are however situations in which quantum mechanics isnecessary to accurately describe the behavior of the atomic nuclei.

Apart from the classical Wigner method, that has been exten-sively described in sections 2.1 and 2.2, there are many methodsfor approximate quantum dynamics. A few notable methods with arelation to the work presented in this thesis will be presented in thissection, so that they can provide a context and serve as a comparisonfor some of the new developments that are presented in chapter 3.

Time correlation functions are important to describe the dynam-ics of physical systems. Examples of such functions can be found inequations 2.60-2.62. Those correlation functions are used to calculatereaction rate constants. The more general formulation of a thermaltime correlation function of operator A at time 0 and operator Bat time t is†

⟨AB(t)

⟩=

1

ZTr

A e−βH e

iHt B e−

iHt

. (2.66)

†An even more general formulation of a time correlation function would haveany number of operators at different times.

42

2.4. Common methods of approximate quantum dynamics

The Boltzmann operator can be placed on either side of A. Thiswill only change the sign of the imaginary part,

1

ZTr

A e−βH e

iHt B e−

iHt

=

1

ZTr

e−βH A e

iHt B e−

iHt

∗,

(2.67)

where ∗ denotes complex conjugate.Some quantum dynamical methods work with Kubo-transformed

time correlation functions,40

⟨AB(t)

⟩Kubo

=1

ZTr

1

β

∫ β

0

dζ e−ζH A e−(β−ζ)H eiHt B e−

iHt

,

(2.68)

which have the benefit that the quantum correlation functions aremore similar to their classical counterparts than is the case for thestandard “physical” correlation functions.

Another alternative is symmetrized correlation functions,41

⟨AB(t)

⟩Sym

=1

ZTr

e−

β2H A e−

β2H e

iHt B e−

iHt

, (2.69)

which have the benefit that they are real instead of complex quanti-ties. The correlation functions, or technically just traces, presentedin section 2.3 are of this type.

The standard, Kubo-transformed, and symmetrized versionsof the same correlation function are equivalent in the sense thatthey contain the same information. Using the Fourier transform,[f(t)]F (ω), the different correlation functions are related as:40,41

[⟨AB(t)

⟩]F(ω) = e−

βω2

[⟨AB(t)

⟩Sym

]

F

(ω)

=βω

eβω −1

[⟨AB(t)

⟩Kubo

]F(ω) . (2.70)

These relations are only valid for exact quantum mechanics. Ifapproximations to quantum mechanics are used, then equation 2.70

43

Page 66: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

will also be approximate. As mentioned some methods work witha certain version of the correlation function. Other methods, e.g.the classical Wigner method, can be used for any of the versions ofthe correlation function. If one is not interested in the correlationfunction itself, but rather something that is calculated from it, thedifferent versions may in some cases give the same result. In thecase of equations 2.60-2.62, it is long time values and an integralthat is interesting rather than the correlation functions themselves.For these equations the ordinary version of the correlation function,or the Kubo-transformed correlation function, would give the sameresult as the symmetrized version. However, as soon as one startsto approximate the correlation functions this does not necessarilyapply any more.

Semi-classical initial value representation

The semi-classical initial value representation18 (SC-IVR) has alreadybeen mentioned in connection to the classical Wigner method. Ituses the IVR-representation of the time propagation operator:

e−iHt =

∫∫dDx0 dDp0

(2πi)D2

√det

(∂xt (x0,p0)

∂p0

)

× eiS(x0,p0,t) |xt〉 〈x0| , (2.71)

where the index 0 denotes initial values, xt (x0,p0) is the positionat time t as a function of the initial values, and Scl (x0,p0, t) is theclassical action of the trajectory that connects x0 and xt (x0,p0).There are some different implementations of SC-IVR. The interestedreader is directed to a paper by Miller.19

Matsubara dynamics

Matsubara dynamics was rather recently (2015) developed by Heleet al.42 and even more recently (2019) generalized from just singletime correlation functions to multitime correlation functions.43 Itis a method for calculating Kubo-transformed correlation functions.Matsubara dynamics is based on the discretized imaginary time

44

2.4. Common methods of approximate quantum dynamics

path integral. Every bead in the path integral polymer is given amomentum and kinetic energy in real time. A normal mode analysisis made for the path integral polymer, in the limit where the numberof beads goes to infinity, and the normal modes are divided intotwo groups. One group contains the so called Matsubara modes,i.e. the modes with the lowest frequencies, and the other groupcontains the non-Matsubara modes, i.e. the higher frequency modes.The propagation forward in time is made by using the quantum

mechanical propagation operator for phase space distributions, eLt,with the Liouvillian operator (here shown one-dimensional)

L =p

m

∂x− 2V (x)

sin

←∂

∂x

2

→∂

∂p

, (2.72)

but only including derivatives of the Matsubara mode coordinates.By using this method of propagation the quantum Boltzmann distri-bution is conserved over time. The intention of this method is to finda more correct way of combining a quantum Boltzmann distributionwith classical dynamics, compared to the classical Wigner method,which does not conserve the quantum Boltzmann distribution overtime.

Matsubara dynamics is a fairly accurate method of approximatequantum dynamics, but it is not practical to use for chemical prob-lems, as it is too computationally demanding. Approximations toMatsubara dynamics, to make it less demanding, have been stud-ied.44 Matsubara dynamics has also been used as a comparisonfor the classical Wigner method,42 and as an intermediate point toderive RPMD and CMD from first principles.45

Centroid molecular dynamics

Centroid molecular dynamics (CMD) was first introduced by Caoand Voth.35 CMD uses the discretized imaginary time path integraland gives Kubo-transformed correlation functions. In CMD theinteresting quantities are measured as centroid variables. For thecase of x and p the corresponding centroid variables are just theaverage position and momentum, respectively, of all the beads in the

45

Page 67: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

will also be approximate. As mentioned some methods work witha certain version of the correlation function. Other methods, e.g.the classical Wigner method, can be used for any of the versions ofthe correlation function. If one is not interested in the correlationfunction itself, but rather something that is calculated from it, thedifferent versions may in some cases give the same result. In thecase of equations 2.60-2.62, it is long time values and an integralthat is interesting rather than the correlation functions themselves.For these equations the ordinary version of the correlation function,or the Kubo-transformed correlation function, would give the sameresult as the symmetrized version. However, as soon as one startsto approximate the correlation functions this does not necessarilyapply any more.

Semi-classical initial value representation

The semi-classical initial value representation18 (SC-IVR) has alreadybeen mentioned in connection to the classical Wigner method. Ituses the IVR-representation of the time propagation operator:

e−iHt =

∫∫dDx0 dDp0

(2πi)D2

√det

(∂xt (x0,p0)

∂p0

)

× eiS(x0,p0,t) |xt〉 〈x0| , (2.71)

where the index 0 denotes initial values, xt (x0,p0) is the positionat time t as a function of the initial values, and Scl (x0,p0, t) is theclassical action of the trajectory that connects x0 and xt (x0,p0).There are some different implementations of SC-IVR. The interestedreader is directed to a paper by Miller.19

Matsubara dynamics

Matsubara dynamics was rather recently (2015) developed by Heleet al.42 and even more recently (2019) generalized from just singletime correlation functions to multitime correlation functions.43 Itis a method for calculating Kubo-transformed correlation functions.Matsubara dynamics is based on the discretized imaginary time

44

2.4. Common methods of approximate quantum dynamics

path integral. Every bead in the path integral polymer is given amomentum and kinetic energy in real time. A normal mode analysisis made for the path integral polymer, in the limit where the numberof beads goes to infinity, and the normal modes are divided intotwo groups. One group contains the so called Matsubara modes,i.e. the modes with the lowest frequencies, and the other groupcontains the non-Matsubara modes, i.e. the higher frequency modes.The propagation forward in time is made by using the quantum

mechanical propagation operator for phase space distributions, eLt,with the Liouvillian operator (here shown one-dimensional)

L =p

m

∂x− 2V (x)

sin

←∂

∂x

2

→∂

∂p

, (2.72)

but only including derivatives of the Matsubara mode coordinates.By using this method of propagation the quantum Boltzmann distri-bution is conserved over time. The intention of this method is to finda more correct way of combining a quantum Boltzmann distributionwith classical dynamics, compared to the classical Wigner method,which does not conserve the quantum Boltzmann distribution overtime.

Matsubara dynamics is a fairly accurate method of approximatequantum dynamics, but it is not practical to use for chemical prob-lems, as it is too computationally demanding. Approximations toMatsubara dynamics, to make it less demanding, have been stud-ied.44 Matsubara dynamics has also been used as a comparisonfor the classical Wigner method,42 and as an intermediate point toderive RPMD and CMD from first principles.45

Centroid molecular dynamics

Centroid molecular dynamics (CMD) was first introduced by Caoand Voth.35 CMD uses the discretized imaginary time path integraland gives Kubo-transformed correlation functions. In CMD theinteresting quantities are measured as centroid variables. For thecase of x and p the corresponding centroid variables are just theaverage position and momentum, respectively, of all the beads in the

45

Page 68: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

path integral. A phase space centroid distribution, that gives thethermal distribution, and a quasi density operator, that is used toacquire centroid variables and their propagation in time, are defined.

CMD was also derived as a centroid mean-field approximation ofMatsubara dynamics.45

For a harmonic oscillator, and an infinite number of beads in thepolymer, the original CMD method is exact, but only for correlationfunctions where at least one of the operators, A or B, is linear,46

Ω = Ω0 +Ω1x+Ω2p, (2.73)

where the Ω:s are constants. CMD works poorly for non-linearcorrelation functions for other potentials.45 The CMD method hasbeen extended to work better for correlation functions of non-linearoperators,46 but this is much more complicated than standard CMD.

One particular disadvantage of CMD is for vibrational spectra,where the peaks of a spectrum can be artificially broadened andshifted in frequency at low temperatures.47 This is an example ofwhat is called the curvature problem of CMD.

Ring polymer molecular dynamics

Ring polymer molecular dynamics (RPMD) was first introducedby Craig and Manolopoulos in 2004.36 Just as the previous twomethods, RPMD uses the discretized imaginary time path integralto calculate Kubo-transformed correlation functions. Contrary toCMD, the measurables of interest are evaluated at the position ofevery bead in the ring polymer and then taken as the mean of thesevalues. To get a propagation in time every bead in the ring polymeris propagated according to classical mechanics, with the addition ofthe interactions with neighboring beads.

RPMD is restricted to the calculation of position correlationfunctions, meaning that correlation functions of operators containingmomentum can not be directly obtained. Momentum correlationfunctions can, however, be calculated from time-derivatives of posi-tion correlation functions.48,49 Thus, momentum correlation func-tions can be obtained, but for certain operators, such as high powersof p, it may be impractical.

46

2.4. Common methods of approximate quantum dynamics

Hele et al.45 showed that, RPMD can be derived from Matsubaradynamics by removing part of the Liouvillian operator.

For a harmonic oscillator, and an infinite number of beads inthe polymer, the RPMD method is exact, but only for correlationfunctions where at least one of the operators, A or B, is linear. Ifthe correlation function happens to be 〈xx(t)〉Kubo, then the RPMDmethod using any number of beads gives the exact result.36 RPMDworks worse for correlation functions of non-linear operators thanfor correlation functions of linear ones, also for other potentials thanthe harmonic oscillator.45

An RPMD correlation function is exact a time 0 and the ensembleis conserved during propagation.

One particular disadvantage of RPMD, as for CMD, is for vi-brational spectra. In the case of RPMD, however, new spuriouspeaks can appear in a spectrum, and the peaks can be artificiallybroadened and split.47 By thermostating the vibrational modes ofthe path integral polymer the RPMD method has been found tobecome better at handling vibrational spectra.50

RPMD is used as a comparison, for the new method that isdeveloped, in papers III and IV, and section 3.3.

Dirac-Frenkel

The Dirac-Frenkel variational principle was first introduced byDirac51 and Frenkel,52 and later a somewhat different form wasderived by McLachlan.53

⟨δΨ(t)

∣∣∣∣i∂

∂t− H

∣∣∣∣Ψ(t)⟩

= 0, (2.74)

where δΨ(t) is a variation in the time dependent wavefunction. Forwavefunctions or densities that have been written in a parametrizedform this variational principle allows the derivation of equations ofmotion in these parameters, giving an approximate solution to thetime dependent Schrodinger equation.

Among the methods that have been derived using the Dirac-Frenkel variational principle can be found e.g. the multi-configurationaltime dependent Hartree (MCTDH) method,54 which is a very accu-rate method for time dependent quantum mechanics, but quickly

47

Page 69: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

path integral. A phase space centroid distribution, that gives thethermal distribution, and a quasi density operator, that is used toacquire centroid variables and their propagation in time, are defined.

CMD was also derived as a centroid mean-field approximation ofMatsubara dynamics.45

For a harmonic oscillator, and an infinite number of beads in thepolymer, the original CMD method is exact, but only for correlationfunctions where at least one of the operators, A or B, is linear,46

Ω = Ω0 +Ω1x+Ω2p, (2.73)

where the Ω:s are constants. CMD works poorly for non-linearcorrelation functions for other potentials.45 The CMD method hasbeen extended to work better for correlation functions of non-linearoperators,46 but this is much more complicated than standard CMD.

One particular disadvantage of CMD is for vibrational spectra,where the peaks of a spectrum can be artificially broadened andshifted in frequency at low temperatures.47 This is an example ofwhat is called the curvature problem of CMD.

Ring polymer molecular dynamics

Ring polymer molecular dynamics (RPMD) was first introducedby Craig and Manolopoulos in 2004.36 Just as the previous twomethods, RPMD uses the discretized imaginary time path integralto calculate Kubo-transformed correlation functions. Contrary toCMD, the measurables of interest are evaluated at the position ofevery bead in the ring polymer and then taken as the mean of thesevalues. To get a propagation in time every bead in the ring polymeris propagated according to classical mechanics, with the addition ofthe interactions with neighboring beads.

RPMD is restricted to the calculation of position correlationfunctions, meaning that correlation functions of operators containingmomentum can not be directly obtained. Momentum correlationfunctions can, however, be calculated from time-derivatives of posi-tion correlation functions.48,49 Thus, momentum correlation func-tions can be obtained, but for certain operators, such as high powersof p, it may be impractical.

46

2.4. Common methods of approximate quantum dynamics

Hele et al.45 showed that, RPMD can be derived from Matsubaradynamics by removing part of the Liouvillian operator.

For a harmonic oscillator, and an infinite number of beads inthe polymer, the RPMD method is exact, but only for correlationfunctions where at least one of the operators, A or B, is linear. Ifthe correlation function happens to be 〈xx(t)〉Kubo, then the RPMDmethod using any number of beads gives the exact result.36 RPMDworks worse for correlation functions of non-linear operators thanfor correlation functions of linear ones, also for other potentials thanthe harmonic oscillator.45

An RPMD correlation function is exact a time 0 and the ensembleis conserved during propagation.

One particular disadvantage of RPMD, as for CMD, is for vi-brational spectra. In the case of RPMD, however, new spuriouspeaks can appear in a spectrum, and the peaks can be artificiallybroadened and split.47 By thermostating the vibrational modes ofthe path integral polymer the RPMD method has been found tobecome better at handling vibrational spectra.50

RPMD is used as a comparison, for the new method that isdeveloped, in papers III and IV, and section 3.3.

Dirac-Frenkel

The Dirac-Frenkel variational principle was first introduced byDirac51 and Frenkel,52 and later a somewhat different form wasderived by McLachlan.53

⟨δΨ(t)

∣∣∣∣i∂

∂t− H

∣∣∣∣Ψ(t)⟩

= 0, (2.74)

where δΨ(t) is a variation in the time dependent wavefunction. Forwavefunctions or densities that have been written in a parametrizedform this variational principle allows the derivation of equations ofmotion in these parameters, giving an approximate solution to thetime dependent Schrodinger equation.

Among the methods that have been derived using the Dirac-Frenkel variational principle can be found e.g. the multi-configurationaltime dependent Hartree (MCTDH) method,54 which is a very accu-rate method for time dependent quantum mechanics, but quickly

47

Page 70: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

becomes computationally expensive as the size of the modeled systemis increased.

A variational principle corresponding to the Dirac-Frenkel one,but for Wigner distributions in phase space, is used in paper II, andsection 3.2.

48

Chapter 3

Developments

The general who wins a battle makes manycalculations in his temple ere the battle isfought. The general who loses a battle makesbut few calculations beforehand. Thus domany calculations lead to victory, and fewcalculations to defeat: How much more do nocalculation at all pave the way to defeat!

Sun Tzu (transl. L. Giles)55

The new contributions to science from the research presented inthis thesis are a reaction rate constant for a radiative associationreaction and two methods of approximate quantum dynamics. Thereaction rate constant is presented in section 3.1 (paper I). Thetwo new quantum dynamical methods are presented in sections 3.2(paper II) and 3.3 (papers III and IV) respectively. In section 3.4suggestions are made for the direction of future research.

49

Page 71: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

2. Background

becomes computationally expensive as the size of the modeled systemis increased.

A variational principle corresponding to the Dirac-Frenkel one,but for Wigner distributions in phase space, is used in paper II, andsection 3.2.

48

Chapter 3

Developments

The general who wins a battle makes manycalculations in his temple ere the battle isfought. The general who loses a battle makesbut few calculations beforehand. Thus domany calculations lead to victory, and fewcalculations to defeat: How much more do nocalculation at all pave the way to defeat!

Sun Tzu (transl. L. Giles)55

The new contributions to science from the research presented inthis thesis are a reaction rate constant for a radiative associationreaction and two methods of approximate quantum dynamics. Thereaction rate constant is presented in section 3.1 (paper I). Thetwo new quantum dynamical methods are presented in sections 3.2(paper II) and 3.3 (papers III and IV) respectively. In section 3.4suggestions are made for the direction of future research.

49

Page 72: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

3.1 Formation of the Hydroxyl Radical

by Radiative Association

In paper I5 new values for the reaction rate constant of the radiativeassociation reaction

O(3P) + H(2S) → OH(X2Π) + γ , (3.a)

where γ is a photon, is presented for the temperature range 10 K -30000 K.

The interest in this reaction comes from astrochemistry as thehydroxyl radical has been observed in the interstellar medium, i.e.the space between the stars, of the Milky Way56 and two othergalaxies.57Also the reactants of this reaction are the hydrogen atomand the oxygen atom. Hydrogen is the most abundant element andoxygen is the third most abundant element in our galaxy so thereshould not be a lack of reactants, from an interstellar point of view.

On earth it could be reasonable to believe that the formationof OH from O and H would happen through a three body collision,i.e. a third species colliding with the activated complex of O and Hand carrying away some energy. However, the Earths atmospherecan, through a back-of-the-envelope calculation, be estimated tohave a number density of ∼ 1019 cm−3 at the surface, while whatis considered a dense cloud in the interstellar medium has numberdensities of between 103 cm−3 and 106 cm−3.58 It is therefore possiblefor reactions that would be completely irrelevant at the surface ofthe Earth to be very important in the interstellar medium.

Three body reactions do noticeably occur in the interstellarmedium in the form of reactions between species adsorbed onto dustgrains, but these reactions are not the subject of paper I.

The radiative association reaction can take place through anumber of different pathways. The ones considered in paper I arethe ones starting from the electronic ground states of the atoms.

The ground state atoms O(3P) and H(2S) can form the moleculein the electronic states OH(X2Π), OH(12Σ−), OH(a4Σ−), andOH(b4Π). The potential energy for these electronic states canbe seen as a function of internuclear distance, xind, in figure 3.1.OH(X2Π) is the only one of these electronic states that has bound

50

3.1. Formation of the Hydroxyl Radical by Radiative Association

−5

0

5

10

15

0 1 2 3 4 5 6

V(x in

d) /

eV

xind / Å

OH(X2Π)OH(12Σ−)OH(A2Σ+)OH(a4Σ−)OH(b4Π)

OH+(X3Σ−) + e−

Figure 3.1: Potential energy surfaces for the electronic states X2Π,59

12Σ−,59 A2Σ+,60 a4Σ−,60 and b4Π60 of OH and a Morse potentialfor the electronic state X3Σ− of OH+61,62 shifted to account for thehydrogen ionization energy.63 The crosses show the original datapoints and the lines are inter- and extrapolations.

vibrational levels, and these bound levels are the product of thereaction. The free or quasi-bound vibrational states of OH(X2Π)will be denoted OH(X2Π)∗.

The two most direct pathways for the reaction are

O(3P) + H(2S) → OH(X2Π)∗ → OH(X2Π) + γ (3.b)

and

O(3P) + H(2S) → OH(12Σ−) → OH(X2Π) + γ . (3.c)

The atoms could also approach each other in any of the electronicstates corresponding to the ground states of the atoms and couple,through e.g. spin-orbit-interactions, to the state A2Σ+, that can beseen in figure 3.1. From the quasibound vibrational levels of thatstate a radiative transition to a bound vibrational level of X2Π canhappen. This, so called, inverse predissociation adds four possiblepathways for the reaction

51

Page 73: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

3.1 Formation of the Hydroxyl Radical

by Radiative Association

In paper I5 new values for the reaction rate constant of the radiativeassociation reaction

O(3P) + H(2S) → OH(X2Π) + γ , (3.a)

where γ is a photon, is presented for the temperature range 10 K -30000 K.

The interest in this reaction comes from astrochemistry as thehydroxyl radical has been observed in the interstellar medium, i.e.the space between the stars, of the Milky Way56 and two othergalaxies.57Also the reactants of this reaction are the hydrogen atomand the oxygen atom. Hydrogen is the most abundant element andoxygen is the third most abundant element in our galaxy so thereshould not be a lack of reactants, from an interstellar point of view.

On earth it could be reasonable to believe that the formationof OH from O and H would happen through a three body collision,i.e. a third species colliding with the activated complex of O and Hand carrying away some energy. However, the Earths atmospherecan, through a back-of-the-envelope calculation, be estimated tohave a number density of ∼ 1019 cm−3 at the surface, while whatis considered a dense cloud in the interstellar medium has numberdensities of between 103 cm−3 and 106 cm−3.58 It is therefore possiblefor reactions that would be completely irrelevant at the surface ofthe Earth to be very important in the interstellar medium.

Three body reactions do noticeably occur in the interstellarmedium in the form of reactions between species adsorbed onto dustgrains, but these reactions are not the subject of paper I.

The radiative association reaction can take place through anumber of different pathways. The ones considered in paper I arethe ones starting from the electronic ground states of the atoms.

The ground state atoms O(3P) and H(2S) can form the moleculein the electronic states OH(X2Π), OH(12Σ−), OH(a4Σ−), andOH(b4Π). The potential energy for these electronic states canbe seen as a function of internuclear distance, xind, in figure 3.1.OH(X2Π) is the only one of these electronic states that has bound

50

3.1. Formation of the Hydroxyl Radical by Radiative Association

−5

0

5

10

15

0 1 2 3 4 5 6

V(x in

d) /

eV

xind / Å

OH(X2Π)OH(12Σ−)OH(A2Σ+)OH(a4Σ−)OH(b4Π)

OH+(X3Σ−) + e−

Figure 3.1: Potential energy surfaces for the electronic states X2Π,59

12Σ−,59 A2Σ+,60 a4Σ−,60 and b4Π60 of OH and a Morse potentialfor the electronic state X3Σ− of OH+61,62 shifted to account for thehydrogen ionization energy.63 The crosses show the original datapoints and the lines are inter- and extrapolations.

vibrational levels, and these bound levels are the product of thereaction. The free or quasi-bound vibrational states of OH(X2Π)will be denoted OH(X2Π)∗.

The two most direct pathways for the reaction are

O(3P) + H(2S) → OH(X2Π)∗ → OH(X2Π) + γ (3.b)

and

O(3P) + H(2S) → OH(12Σ−) → OH(X2Π) + γ . (3.c)

The atoms could also approach each other in any of the electronicstates corresponding to the ground states of the atoms and couple,through e.g. spin-orbit-interactions, to the state A2Σ+, that can beseen in figure 3.1. From the quasibound vibrational levels of thatstate a radiative transition to a bound vibrational level of X2Π canhappen. This, so called, inverse predissociation adds four possiblepathways for the reaction

51

Page 74: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

O(3P) + H(2S) → OH(X2Π)∗ → OH(A2Σ+)→ OH(X2Π) + γ , (3.d)

O(3P) + H(2S) → OH(12Σ−) → OH(A2Σ+)→ OH(X2Π) + γ , (3.e)

O(3P) + H(2S) → OH(a4Σ−) → OH(A2Σ+)→ OH(X2Π) + γ , (3.f)

and

O(3P) + H(2S) → OH(b4Π) → OH(A2Σ+)→ OH(X2Π) + γ . (3.g)

Only single-photon processes are considered, as two-photon processesare highly unlikely to occur.

Possible competing reactions are the electron detachment reaction

O(3P) + H(2S) → OH(X2Π/12Σ−/a4Σ−/b4Π)→ OH+(X3Σ−) + e− (3.h)

and the charge transfer reactions

O(3P) + H(2S) → OH(X2Π/12Σ−/a4Σ−/b4Π)→ O−(2P) + H+ (3.i)

and

O(3P) + H(2S) → OH(X2Π/12Σ−/a4Σ−/b4Π)→ O+(4S) + H−(1S) . (3.j)

Theory and method

Several methods were used to calculate rate constants in this work.These used the potential energy surfaces shown in figure 3.1 togetherwith the dipole moment of the state X2Π and the transition dipolemoment of the transition 12Σ− → X2Π that are shown in figure 3.2.

52

3.1. Formation of the Hydroxyl Radical by Radiative Association

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 1 2 3 4 5 6

Dip

ole

mom

ent / (e

a0)

xind / Å

OH(X2Π)

OH(12Σ

−) −> OH(X

2Π)

Figure 3.2: Electric dipole moment of the electronic state X2Π64

and transition dipole moment for the transition between electronicstates 12Σ− and X2Π,59 of OH. ea0 is the elemental charge timesthe Bohr radius. The crosses show the original data points and thelines are inter- and extrapolations.

Quantum mechanical perturbation theory

To get quantum mechanical reaction cross sections a first orderperturbation theory was used. The perturbation, that is added tothe hamiltonian of the system, is in this case the interaction betweenthe electric field of the emitted photon and the dipole momentof the molecular system. This perturbation theory for transitionprobabilities is sometimes called “Fermi’s golden rule” or just the“golden rule”. It is more thoroughly described in e.g. the textbookby Bransden and Joachain.65 The application of this perturbationtheory to get radiative association cross sections has been describedin e.g. references 66 and 67. This method requires wavefunctionsto be computed for both the unbound and bound states so that thematrix elements of the dipole moment operator can be obtained.Here the unbound wavefunctions were calculated with Numerov’smethod68 and the bound wavefunctions were calculated with the

53

Page 75: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

O(3P) + H(2S) → OH(X2Π)∗ → OH(A2Σ+)→ OH(X2Π) + γ , (3.d)

O(3P) + H(2S) → OH(12Σ−) → OH(A2Σ+)→ OH(X2Π) + γ , (3.e)

O(3P) + H(2S) → OH(a4Σ−) → OH(A2Σ+)→ OH(X2Π) + γ , (3.f)

and

O(3P) + H(2S) → OH(b4Π) → OH(A2Σ+)→ OH(X2Π) + γ . (3.g)

Only single-photon processes are considered, as two-photon processesare highly unlikely to occur.

Possible competing reactions are the electron detachment reaction

O(3P) + H(2S) → OH(X2Π/12Σ−/a4Σ−/b4Π)→ OH+(X3Σ−) + e− (3.h)

and the charge transfer reactions

O(3P) + H(2S) → OH(X2Π/12Σ−/a4Σ−/b4Π)→ O−(2P) + H+ (3.i)

and

O(3P) + H(2S) → OH(X2Π/12Σ−/a4Σ−/b4Π)→ O+(4S) + H−(1S) . (3.j)

Theory and method

Several methods were used to calculate rate constants in this work.These used the potential energy surfaces shown in figure 3.1 togetherwith the dipole moment of the state X2Π and the transition dipolemoment of the transition 12Σ− → X2Π that are shown in figure 3.2.

52

3.1. Formation of the Hydroxyl Radical by Radiative Association

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 1 2 3 4 5 6

Dip

ole

mom

ent / (e

a0)

xind / Å

OH(X2Π)

OH(12Σ

−) −> OH(X

2Π)

Figure 3.2: Electric dipole moment of the electronic state X2Π64

and transition dipole moment for the transition between electronicstates 12Σ− and X2Π,59 of OH. ea0 is the elemental charge timesthe Bohr radius. The crosses show the original data points and thelines are inter- and extrapolations.

Quantum mechanical perturbation theory

To get quantum mechanical reaction cross sections a first orderperturbation theory was used. The perturbation, that is added tothe hamiltonian of the system, is in this case the interaction betweenthe electric field of the emitted photon and the dipole momentof the molecular system. This perturbation theory for transitionprobabilities is sometimes called “Fermi’s golden rule” or just the“golden rule”. It is more thoroughly described in e.g. the textbookby Bransden and Joachain.65 The application of this perturbationtheory to get radiative association cross sections has been describedin e.g. references 66 and 67. This method requires wavefunctionsto be computed for both the unbound and bound states so that thematrix elements of the dipole moment operator can be obtained.Here the unbound wavefunctions were calculated with Numerov’smethod68 and the bound wavefunctions were calculated with the

53

Page 76: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

version of the discrete variable representation (DVR) of Colbert andMiller.69

Since molecules can rotate these calculations need to account fordifferent rotational states. When rotation is added to the diatomicsystem a centrifugal potential, ∝ 1/x2

ind, has to be added to thepotential energy surface. When this, purely repulsive, centrifugalpotential is added to the X2Π potential a barrier is obtained for somerotational quantum numbers. This barrier can cause quasiboundvibrational states, i.e. states with an energy lower than the barrierbut above the dissociation limit, so tunneling is necessary to enter orescape. The reaction cross section calculated with this method willlook like a smooth function of energy, but when the collision energycoincides with the energy of a quasibound state there is a suddenpeak in the function. (These peaks can be seen in figure 3.3.) Thephenomenon is called resonance.

The quasibound states are long lived, compared with a wavepacketjust bouncing off the repulsive part of the potential. Somewhat sim-plified, the system could be considered as getting stuck inside thebarrier for a while. Thus, for a collision energy with a resonancethere will be much more time for the radiative transition from thequasibound state to occur, compared to the simple bounce-off situa-tion. These quasibound states have a finite lifetime, since there isthe possibility of escaping them through tunneling or the emissionof a photon. Due to this, there is a kind of uncertainty relation,

∆Eres =τres

, (3.1)

where τres is the lifetime of the resonance and ∆Eres is the so called“width” of the resonance. The width is actually related to the widthof the peaks in the reaction cross section.

A drawback of this perturbation theory is, as for all first orderperturbation theories, that the perturbation has to be small. If theperturbation, in this case the interaction between the photon and thesystem, is large enough to significantly affect the bound or unboundwavefunctions, then this method will not work that well.

Another drawback of this perturbation theory method is thatwhen putting it into collision theory and calculating the integralover energy, equation 2.65, that integral has to be done numerically

54

3.1. Formation of the Hydroxyl Radical by Radiative Association

and the cross section has to be calculated on a grid of energies. Ifthere are resonances to account for, the grid has to be fine enough toproperly represent these possibly very thin peaks in the cross sectionfunction. If the grid is too coarse the peaks may be distorted inshape or may not be found at all.

Semiclassical method

An alternative to perturbation theory when it comes to calculatingquantum mechanical cross sections is the use of an optical potential,

Vopt(xind) = − i2AEinstein(xind), (3.2)

where AEinstein(xind) is the Einstein A coefficient, that is the proba-bility of transition between two states per unit of time. This opticalpotential is added to the standard hamiltonian of the system. Whenthe system is propagated forward in time the imaginary opticalpotential will absorb part of the probability of the unbound states.This means that there is no need to calculate the wavefunctions forthe bound states, only for the unbound ones.

The method used in paper I is restricted by the Franck-Condonprinciple,70–72 which states that during a transition between twoelectronic states in a molecule the electrons will move to the ap-propriate new configuration, but the nuclei are so slow that theywill keep their position and momentum and only adapt after thetransition has already happened. This is a predecessor to the Born-Oppenheimer approximation.3 This optical potential method onlyworks for radiative transitions between different electronic states, inthis case reaction 3.c.

In this work the unbound wavefunctions were handled semiclassi-caly with the Wentzel-Kramers-Brillouin (WKB) approximation.73–75

With the semiclassical approximation, this method can not ac-count for resonances. But it can give the baseline of the reactioncross section.

55

Page 77: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

version of the discrete variable representation (DVR) of Colbert andMiller.69

Since molecules can rotate these calculations need to account fordifferent rotational states. When rotation is added to the diatomicsystem a centrifugal potential, ∝ 1/x2

ind, has to be added to thepotential energy surface. When this, purely repulsive, centrifugalpotential is added to the X2Π potential a barrier is obtained for somerotational quantum numbers. This barrier can cause quasiboundvibrational states, i.e. states with an energy lower than the barrierbut above the dissociation limit, so tunneling is necessary to enter orescape. The reaction cross section calculated with this method willlook like a smooth function of energy, but when the collision energycoincides with the energy of a quasibound state there is a suddenpeak in the function. (These peaks can be seen in figure 3.3.) Thephenomenon is called resonance.

The quasibound states are long lived, compared with a wavepacketjust bouncing off the repulsive part of the potential. Somewhat sim-plified, the system could be considered as getting stuck inside thebarrier for a while. Thus, for a collision energy with a resonancethere will be much more time for the radiative transition from thequasibound state to occur, compared to the simple bounce-off situa-tion. These quasibound states have a finite lifetime, since there isthe possibility of escaping them through tunneling or the emissionof a photon. Due to this, there is a kind of uncertainty relation,

∆Eres =τres

, (3.1)

where τres is the lifetime of the resonance and ∆Eres is the so called“width” of the resonance. The width is actually related to the widthof the peaks in the reaction cross section.

A drawback of this perturbation theory is, as for all first orderperturbation theories, that the perturbation has to be small. If theperturbation, in this case the interaction between the photon and thesystem, is large enough to significantly affect the bound or unboundwavefunctions, then this method will not work that well.

Another drawback of this perturbation theory method is thatwhen putting it into collision theory and calculating the integralover energy, equation 2.65, that integral has to be done numerically

54

3.1. Formation of the Hydroxyl Radical by Radiative Association

and the cross section has to be calculated on a grid of energies. Ifthere are resonances to account for, the grid has to be fine enough toproperly represent these possibly very thin peaks in the cross sectionfunction. If the grid is too coarse the peaks may be distorted inshape or may not be found at all.

Semiclassical method

An alternative to perturbation theory when it comes to calculatingquantum mechanical cross sections is the use of an optical potential,

Vopt(xind) = − i2AEinstein(xind), (3.2)

where AEinstein(xind) is the Einstein A coefficient, that is the proba-bility of transition between two states per unit of time. This opticalpotential is added to the standard hamiltonian of the system. Whenthe system is propagated forward in time the imaginary opticalpotential will absorb part of the probability of the unbound states.This means that there is no need to calculate the wavefunctions forthe bound states, only for the unbound ones.

The method used in paper I is restricted by the Franck-Condonprinciple,70–72 which states that during a transition between twoelectronic states in a molecule the electrons will move to the ap-propriate new configuration, but the nuclei are so slow that theywill keep their position and momentum and only adapt after thetransition has already happened. This is a predecessor to the Born-Oppenheimer approximation.3 This optical potential method onlyworks for radiative transitions between different electronic states, inthis case reaction 3.c.

In this work the unbound wavefunctions were handled semiclassi-caly with the Wentzel-Kramers-Brillouin (WKB) approximation.73–75

With the semiclassical approximation, this method can not ac-count for resonances. But it can give the baseline of the reactioncross section.

55

Page 78: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

Classical method

Since the semiclassical method described in the preceding sectiononly handles transitions between two different electronic states , itcan not be used for reaction 3.b. To get an alternative method tocalculate cross sections for this reaction a purely classical method wasused instead. This classical method was used in the form developedby Gustafsson.76

This classical method uses the classical equations of motion tocompute the electric dipole moment of the system as a function oftime. The frequency of the changes in dipole moment can be equatedwith the frequency of the emitted photon and the intensity of theemission can be calculated as the Larmor power,77 that depends onthe second time derivative of the dipole moment.

Since this method is based on the classical equations of motion, itcan not account for resonances, but just as the semiclassical method,it can give the baseline of the reaction cross section.

Breit-Wigner theory and inverse predissociation

Breit-Wigner theory78 allows the calculation of the contribution tothe reaction rate constant that comes from the resonances. A certainshape is assumed for the resonance-peaks, and the contributionto the reaction cross section or rate constant from each resonancecan be calculated from the width/lifetime of the resonance. Thiscontribution is

σres(E) ∝ ∆Eres,rad∆Eres,tun

(E − Eres)2 + (∆Eres,tot)

2

4

=

1τres,radτres,tun

(E−Eres)2

2 + 14τ2res,tot

(3.3)

kr,res(T ) ∝∆Eres,rad∆Eres,tun

∆Eres,tot

=τres,tot

τres,radτres,tun

=

τres,rad + τres,tun, (3.4)

where the relevant energy, E, is the collision energy compared tothe asymptotic potential energy of X2Π, and Eres is the energyof the quasibound state. The width/lifetime of the resonance has

56

3.1. Formation of the Hydroxyl Radical by Radiative Association

been divided into one part for the radiative process of escapingthe quasibound state and one part for the tunneling out of thequasibound state. The total process of escaping the quasibound statehas ∆Eres,tot = ∆Eres,rad +∆Eres,tun and 1/τres,tot = 1/τres,rad + 1/τres,tun.

As has already been mentioned, the quantum mechanical per-turbation theory gives the reaction cross section as a function ofenergy, and the integral over energy in equation 2.65 has to beperformed numerically. The ability of the perturbation theory todescribe the contribution from the resonance is thus dependent onthat the grid that the integration is conducted upon is fine enoughto capture the position and shape of the very narrow peaks in thecross section. Breit-Wigner theory instead assumes a peak shapethat is analytically integrable, and therefore does not have the sameproblems as the perturbation theory.

Of course computations are needed to find the energies of thequasibound states and calculate their widths/lifetimes. In this workthe program LEVEL79 was used. The methods to accomplish thesecomputations will not be further explored here. The Breit-Wignertheory was used to add a resonance contribution to the cross sectionof reaction 3.b when it was calculated with the classical method.

The use of equations 3.3 and 3.4 is not restricted to quasiboundvibrational states caused by tunneling. Reactions 3.d-3.g occur viaquasibound vibrational states in A2Σ+ that can be entered andexited, not by tunneling through a potential barrier, but rather bycoupling to the electronic states X2Π, 12Σ−, a4Σ−, and b4Π, e.g.through spin-orbit interactions.

This process is called inverse predissociation. Predissociation iswhen a bound molecule absorbs a photon and is excited to this kindof quasibound state and then changes electronic state and becomesunbound, dissociating. Inverse predissociation is thus the inverse ofthis.

As long as the energies of the quasibound states and thewidths/lifetimes for the radiative process and the coupling betweenthe electronic states can be obtained, a contribution to the reactionrate constant can be calculated through equations 3.3 and 3.4. Inpaper I these were taken from the literature.

57

Page 79: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

Classical method

Since the semiclassical method described in the preceding sectiononly handles transitions between two different electronic states , itcan not be used for reaction 3.b. To get an alternative method tocalculate cross sections for this reaction a purely classical method wasused instead. This classical method was used in the form developedby Gustafsson.76

This classical method uses the classical equations of motion tocompute the electric dipole moment of the system as a function oftime. The frequency of the changes in dipole moment can be equatedwith the frequency of the emitted photon and the intensity of theemission can be calculated as the Larmor power,77 that depends onthe second time derivative of the dipole moment.

Since this method is based on the classical equations of motion, itcan not account for resonances, but just as the semiclassical method,it can give the baseline of the reaction cross section.

Breit-Wigner theory and inverse predissociation

Breit-Wigner theory78 allows the calculation of the contribution tothe reaction rate constant that comes from the resonances. A certainshape is assumed for the resonance-peaks, and the contributionto the reaction cross section or rate constant from each resonancecan be calculated from the width/lifetime of the resonance. Thiscontribution is

σres(E) ∝ ∆Eres,rad∆Eres,tun

(E − Eres)2 + (∆Eres,tot)

2

4

=

1τres,radτres,tun

(E−Eres)2

2 + 14τ2res,tot

(3.3)

kr,res(T ) ∝∆Eres,rad∆Eres,tun

∆Eres,tot

=τres,tot

τres,radτres,tun

=

τres,rad + τres,tun, (3.4)

where the relevant energy, E, is the collision energy compared tothe asymptotic potential energy of X2Π, and Eres is the energyof the quasibound state. The width/lifetime of the resonance has

56

3.1. Formation of the Hydroxyl Radical by Radiative Association

been divided into one part for the radiative process of escapingthe quasibound state and one part for the tunneling out of thequasibound state. The total process of escaping the quasibound statehas ∆Eres,tot = ∆Eres,rad +∆Eres,tun and 1/τres,tot = 1/τres,rad + 1/τres,tun.

As has already been mentioned, the quantum mechanical per-turbation theory gives the reaction cross section as a function ofenergy, and the integral over energy in equation 2.65 has to beperformed numerically. The ability of the perturbation theory todescribe the contribution from the resonance is thus dependent onthat the grid that the integration is conducted upon is fine enoughto capture the position and shape of the very narrow peaks in thecross section. Breit-Wigner theory instead assumes a peak shapethat is analytically integrable, and therefore does not have the sameproblems as the perturbation theory.

Of course computations are needed to find the energies of thequasibound states and calculate their widths/lifetimes. In this workthe program LEVEL79 was used. The methods to accomplish thesecomputations will not be further explored here. The Breit-Wignertheory was used to add a resonance contribution to the cross sectionof reaction 3.b when it was calculated with the classical method.

The use of equations 3.3 and 3.4 is not restricted to quasiboundvibrational states caused by tunneling. Reactions 3.d-3.g occur viaquasibound vibrational states in A2Σ+ that can be entered andexited, not by tunneling through a potential barrier, but rather bycoupling to the electronic states X2Π, 12Σ−, a4Σ−, and b4Π, e.g.through spin-orbit interactions.

This process is called inverse predissociation. Predissociation iswhen a bound molecule absorbs a photon and is excited to this kindof quasibound state and then changes electronic state and becomesunbound, dissociating. Inverse predissociation is thus the inverse ofthis.

As long as the energies of the quasibound states and thewidths/lifetimes for the radiative process and the coupling betweenthe electronic states can be obtained, a contribution to the reactionrate constant can be calculated through equations 3.3 and 3.4. Inpaper I these were taken from the literature.

57

Page 80: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

Electron detachment

For reaction 3.h an upper bound to the reaction cross section wasestimated through a method that has been used for O− + H.80 Thisapproximate method assumes that, for a certain electronic state allclassical trajectories that reach the point where the potential energysurface of this electronic state crosses the potential energy surfaceof the electronic state with a detached electron, xcross, leads to areaction,

σElec.det.(E) ∝

0 if E < V (xcross)

πx2cross

(1− V (xcross)

E

)if E V (xcross)

.

(3.5)

Since the X2Π potential energy surface was not available at, norextrapolated to, the point where it would cross the potential energy ofthe X3Σ−-state of OH+ it was not counted in the total cross section.However, from looking at figure 3.1 it can be safely assumed thatthis V (xcross) will not be the crossing of potential energy surfacesthat is lowest in energy, and only the one that is lowest in energywill determine the energy at which the electron detachment crosssection starts being non-zero. As will be seen in the next section,only the energy at which the electron detachment may start to occur,rather than the estimated value of the cross section, is important.Not accounting for X2Π will thus not affect the end results of thiswork.

Results and discussion

In figure 3.3 the reaction cross section for reaction 3.b is shown. Thedifference between the cross section calculated with the perturbationtheory and the classical method with the addition of Breit-Wignertheory can be seen. The baselines of the two approaches are verysimilar except at the highest and lowest energies. Thus, the classicalmethod is a very good approximation for the non-resonant contri-bution to this cross section. It is visible that Breit-Wigner theoryfinds many more resonances than the perturbation theory does, par-ticularly at high energies, where the perturbation theory does not

58

3.1. Formation of the Hydroxyl Radical by Radiative Association

1E−14

1E−12

1E−10

1E−08

1E−06

1E−04

1E−02

1E−05 0.0001 0.001 0.01 0.1 1

σ(E

) / Å

2

E / eV

QMCL+BW

Figure 3.3: Reaction cross section for the reaction OH(X2Π)∗ →OH(X2Π) + γ (3.b) calculated with the quantum mechanical per-turbation theory (QM) and the classical method + Breit-Wignertheory (CL+BW).

find any resonances at all. The axes in the graph are logarithmic.To get a manageable number of grid points for the energy, the gridis coarser at higher energies than at lower ones. This can now beseen, as the perturbation theory is completely dependent on havinga fine enough energy grid for finding the resonances. For the highestenergies, where the grid is the coarsest, the perturbation theory doesnot find any resonances at all. Meanwhile the Breit-Wigner theoryhave found a lot of resonances, independent of the grid. Since theperturbation theory misses too many resonances, the cross sectionfrom the classical method with the addition of Breit-Wigner theorywas used for the rate constant calculations, see figure 3.5.

In figure 3.4 the reaction cross section for reaction 3.c is shown.The difference between the cross section calculated with the pertur-bation theory and the semiclassical method can be seen. Since theelectronic state 12Σ− is purely repulsive, there are no quasi-boundvibrational states and thus no resonances. At low energies the pertur-

59

Page 81: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

Electron detachment

For reaction 3.h an upper bound to the reaction cross section wasestimated through a method that has been used for O− + H.80 Thisapproximate method assumes that, for a certain electronic state allclassical trajectories that reach the point where the potential energysurface of this electronic state crosses the potential energy surfaceof the electronic state with a detached electron, xcross, leads to areaction,

σElec.det.(E) ∝

0 if E < V (xcross)

πx2cross

(1− V (xcross)

E

)if E V (xcross)

.

(3.5)

Since the X2Π potential energy surface was not available at, norextrapolated to, the point where it would cross the potential energy ofthe X3Σ−-state of OH+ it was not counted in the total cross section.However, from looking at figure 3.1 it can be safely assumed thatthis V (xcross) will not be the crossing of potential energy surfacesthat is lowest in energy, and only the one that is lowest in energywill determine the energy at which the electron detachment crosssection starts being non-zero. As will be seen in the next section,only the energy at which the electron detachment may start to occur,rather than the estimated value of the cross section, is important.Not accounting for X2Π will thus not affect the end results of thiswork.

Results and discussion

In figure 3.3 the reaction cross section for reaction 3.b is shown. Thedifference between the cross section calculated with the perturbationtheory and the classical method with the addition of Breit-Wignertheory can be seen. The baselines of the two approaches are verysimilar except at the highest and lowest energies. Thus, the classicalmethod is a very good approximation for the non-resonant contri-bution to this cross section. It is visible that Breit-Wigner theoryfinds many more resonances than the perturbation theory does, par-ticularly at high energies, where the perturbation theory does not

58

3.1. Formation of the Hydroxyl Radical by Radiative Association

1E−14

1E−12

1E−10

1E−08

1E−06

1E−04

1E−02

1E−05 0.0001 0.001 0.01 0.1 1

σ(E

) / Å

2

E / eV

QMCL+BW

Figure 3.3: Reaction cross section for the reaction OH(X2Π)∗ →OH(X2Π) + γ (3.b) calculated with the quantum mechanical per-turbation theory (QM) and the classical method + Breit-Wignertheory (CL+BW).

find any resonances at all. The axes in the graph are logarithmic.To get a manageable number of grid points for the energy, the gridis coarser at higher energies than at lower ones. This can now beseen, as the perturbation theory is completely dependent on havinga fine enough energy grid for finding the resonances. For the highestenergies, where the grid is the coarsest, the perturbation theory doesnot find any resonances at all. Meanwhile the Breit-Wigner theoryhave found a lot of resonances, independent of the grid. Since theperturbation theory misses too many resonances, the cross sectionfrom the classical method with the addition of Breit-Wigner theorywas used for the rate constant calculations, see figure 3.5.

In figure 3.4 the reaction cross section for reaction 3.c is shown.The difference between the cross section calculated with the pertur-bation theory and the semiclassical method can be seen. Since theelectronic state 12Σ− is purely repulsive, there are no quasi-boundvibrational states and thus no resonances. At low energies the pertur-

59

Page 82: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

1E−20

1E−18

1E−16

1E−14

1E−12

1E−10

1E−08

1E−06

1E−05 0.0001 0.001 0.01 0.1 1 10

σ(E

) / Å

2

E / eV

QMSC

Figure 3.4: Reaction cross section for the reaction OH(12Σ−) →OH(X2Π) + γ (3.c) calculated with the quantum mechanical per-turbation theory (QM) and the semiclassical method (SC).

bation theory cross sections are much higher than the semiclassicalones, due, at least in part, to tunneling into the classically forbiddenshort distances. At high energies the semiclassical cross section cannot be calculated because the Franck-Condon principle restricts theenergy of the emitted photon to the difference between the potentialenergy surfaces, and if this difference is lower than the collisionenergy the bound vibrational states are inaccessible. In figure 3.5the reaction rate constants calculated from these cross sections areshown. The noticeable difference between the perturbation theoryand semiclassical reaction rate constants is in the temperature range200 K - 1000 K, where reaction 3.b dominates. This difference isthus not a problem.

It is reassuring, for reaction 3.c, that different methods give thesame result for the reaction rate constant. This indicates that theresult should be reliable.

In figure 3.6 the cross sections for reactions 3.b, 3.c, and 3.h hasbeen gathered in the same graph. It can be seen that at low energies

60

3.1. Formation of the Hydroxyl Radical by Radiative Association

0.0001

0.001

0.01

0.1

1

10

100

1000

10 100 1000 10000

kr(

T)

/ (1

0−

20cm

3s

−1)

T / K

OH(X2Π) −> OH(X

2Π) CL+BW

OH(12Σ

−) −> OH(X

2Π) QM

OH(12Σ

−) −> OH(X

2Π) SCTotal

Figure 3.5: Reaction rate constant for the reactions OH(X2Π)∗ →OH(X2Π) + γ (3.b), calculated with the classical method + Breit-Wigner theory (CL+BW), and OH(12Σ−) → OH(X2Π) + γ (3.c),calculated with the quantum mechanical perturbation theory (QM)and the semiclassical method (SC). Additionally a total reactionrate constant is shown.

the cross section for reaction 3.b is much higher than the crosssection for reaction 3.c. At higher energies the processes have similarcross sections, and from around ∼1 eV reaction 3.c gets the highercross section. This is the same kind of trend as for the reaction rateconstants for the two reactions that become equal around ∼2000 K.

In figure 3.6 the estimate of the cross section for the electrondetachment process, reaction 3.h, can also be seen. This approxi-mate cross section is supposed to be an upper limit to the real crosssection. The energies at which this reaction could interfere with theradiative association are the ones for which there is no calculatedsemiclassical cross section for reaction 3.c. The perturbation theorycross section does, however, have finite values at these high collisionenergies. The highest collision energies are most important at thehighest temperatures. Since the rate constants calculated from the

61

Page 83: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

1E−20

1E−18

1E−16

1E−14

1E−12

1E−10

1E−08

1E−06

1E−05 0.0001 0.001 0.01 0.1 1 10

σ(E

) / Å

2

E / eV

QMSC

Figure 3.4: Reaction cross section for the reaction OH(12Σ−) →OH(X2Π) + γ (3.c) calculated with the quantum mechanical per-turbation theory (QM) and the semiclassical method (SC).

bation theory cross sections are much higher than the semiclassicalones, due, at least in part, to tunneling into the classically forbiddenshort distances. At high energies the semiclassical cross section cannot be calculated because the Franck-Condon principle restricts theenergy of the emitted photon to the difference between the potentialenergy surfaces, and if this difference is lower than the collisionenergy the bound vibrational states are inaccessible. In figure 3.5the reaction rate constants calculated from these cross sections areshown. The noticeable difference between the perturbation theoryand semiclassical reaction rate constants is in the temperature range200 K - 1000 K, where reaction 3.b dominates. This difference isthus not a problem.

It is reassuring, for reaction 3.c, that different methods give thesame result for the reaction rate constant. This indicates that theresult should be reliable.

In figure 3.6 the cross sections for reactions 3.b, 3.c, and 3.h hasbeen gathered in the same graph. It can be seen that at low energies

60

3.1. Formation of the Hydroxyl Radical by Radiative Association

0.0001

0.001

0.01

0.1

1

10

100

1000

10 100 1000 10000

kr(

T)

/ (1

0−

20cm

3s

−1)

T / K

OH(X2Π) −> OH(X

2Π) CL+BW

OH(12Σ

−) −> OH(X

2Π) QM

OH(12Σ

−) −> OH(X

2Π) SCTotal

Figure 3.5: Reaction rate constant for the reactions OH(X2Π)∗ →OH(X2Π) + γ (3.b), calculated with the classical method + Breit-Wigner theory (CL+BW), and OH(12Σ−) → OH(X2Π) + γ (3.c),calculated with the quantum mechanical perturbation theory (QM)and the semiclassical method (SC). Additionally a total reactionrate constant is shown.

the cross section for reaction 3.b is much higher than the crosssection for reaction 3.c. At higher energies the processes have similarcross sections, and from around ∼1 eV reaction 3.c gets the highercross section. This is the same kind of trend as for the reaction rateconstants for the two reactions that become equal around ∼2000 K.

In figure 3.6 the estimate of the cross section for the electrondetachment process, reaction 3.h, can also be seen. This approxi-mate cross section is supposed to be an upper limit to the real crosssection. The energies at which this reaction could interfere with theradiative association are the ones for which there is no calculatedsemiclassical cross section for reaction 3.c. The perturbation theorycross section does, however, have finite values at these high collisionenergies. The highest collision energies are most important at thehighest temperatures. Since the rate constants calculated from the

61

Page 84: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

1E−20

1E−15

1E−10

1E−05

1E+00

1E+05

1E−05 0.0001 0.001 0.01 0.1 1 10 100

σ(E

) / Å

2

E / eV

OH(X2Π) −> OH(X

2Π) QM

OH(X2Π) −> OH(X

2Π) CL+BW

OH(12Σ

−) −> OH(X

2Π) QM

OH(12Σ

−) −> OH(X

2Π) SC

OH(12Σ

−/a

−/b

4Π) −> OH

+(X

−) + e

Figure 3.6: Reaction cross sections for the reactions OH(X2Π)∗

→ OH(X2Π) + γ (3.b) and OH(12Σ−) → OH(X2Π) + γ (3.c),calculated with the quantum mechanical perturbation theory (QM),the classical method + Breit-Wigner theory (CL+BW), and thesemiclassical method (SC). Estimate of reaction cross section forOH(12Σ−/a4Σ−/b4Π) → OH+(X3Σ−) + e− (3.h).

semiclassical and perturbation theory cross sections are indistin-guishable at the high temperatures in figure 3.5, it can be concludedthat the high collision energies where the two different cross sectionsare very different are not important at the temperatures studiedin this work. If these collision energies are unimportant, then thecompeting electron detachment process will not affect the reactionrate constant studied here. Likewise, the charge transfer reactions,3.i and 3.j, have threshold energies of >12 eV∗, making also themnegligible.

In figure 3.7 the total reaction rate constants for the processes3.b and 3.c, and the total reaction rate constant for the inversepredissociation are shown. From around ∼20 K to ∼9000 K inverse

∗These energies are acquired from an addition of ionization energies63,81–83

and electron affinities.84,85

62

3.1. Formation of the Hydroxyl Radical by Radiative Association

0.1

1

10

100

1000

10 100 1000 10000

kr(

T)

/ (1

0−

20cm

3s

−1)

T / K

3.b + 3.cInv. predis.

TotalOld

Figure 3.7: The total reaction rate constants for the radiative asso-ciation reaction O(3P) + H(2S) → OH(X2Π) + γ (3.a) through theprocesses 3.b + 3.c, the inverse predissociation processes (reactions3.d-3.g), and the total sum of the processes. An older value for thesame reaction rate constant, taken from KIDA,86 is also shown.

predissociation dominates the reaction rate constant. At tempera-tures above ∼10000 K process 3.c is, however, dominating.

This dominance by the process 3.c at high temperature meansthat the temperatures where the inverse predissociation makes anoticeable contribution to the total reaction rate constant will betemperatures where the highest collision energies will not matter.Thus, the high energy processes of electron detachment and chargetransfer should not interfere with the inverse predissociation.

When it comes to possible interference between the processes 3.band 3.c and inverse predissociation reactions 3.d-3.g, 3.f and 3.g goesvia electronic states that are not involved with 3.b and 3.c and willnot noticeably affect or be affected by them. 3.f is assumed to bethe dominating process for inverse predissociation for most of thetemperature interval studied.60 Thus, it can also be assumed thatinverse predissociaton will not interfere with 3.b and 3.c for most

63

Page 85: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

1E−20

1E−15

1E−10

1E−05

1E+00

1E+05

1E−05 0.0001 0.001 0.01 0.1 1 10 100

σ(E

) / Å

2

E / eV

OH(X2Π) −> OH(X

2Π) QM

OH(X2Π) −> OH(X

2Π) CL+BW

OH(12Σ

−) −> OH(X

2Π) QM

OH(12Σ

−) −> OH(X

2Π) SC

OH(12Σ

−/a

−/b

4Π) −> OH

+(X

−) + e

Figure 3.6: Reaction cross sections for the reactions OH(X2Π)∗

→ OH(X2Π) + γ (3.b) and OH(12Σ−) → OH(X2Π) + γ (3.c),calculated with the quantum mechanical perturbation theory (QM),the classical method + Breit-Wigner theory (CL+BW), and thesemiclassical method (SC). Estimate of reaction cross section forOH(12Σ−/a4Σ−/b4Π) → OH+(X3Σ−) + e− (3.h).

semiclassical and perturbation theory cross sections are indistin-guishable at the high temperatures in figure 3.5, it can be concludedthat the high collision energies where the two different cross sectionsare very different are not important at the temperatures studiedin this work. If these collision energies are unimportant, then thecompeting electron detachment process will not affect the reactionrate constant studied here. Likewise, the charge transfer reactions,3.i and 3.j, have threshold energies of >12 eV∗, making also themnegligible.

In figure 3.7 the total reaction rate constants for the processes3.b and 3.c, and the total reaction rate constant for the inversepredissociation are shown. From around ∼20 K to ∼9000 K inverse

∗These energies are acquired from an addition of ionization energies63,81–83

and electron affinities.84,85

62

3.1. Formation of the Hydroxyl Radical by Radiative Association

0.1

1

10

100

1000

10 100 1000 10000

kr(

T)

/ (1

0−

20cm

3s

−1)

T / K

3.b + 3.cInv. predis.

TotalOld

Figure 3.7: The total reaction rate constants for the radiative asso-ciation reaction O(3P) + H(2S) → OH(X2Π) + γ (3.a) through theprocesses 3.b + 3.c, the inverse predissociation processes (reactions3.d-3.g), and the total sum of the processes. An older value for thesame reaction rate constant, taken from KIDA,86 is also shown.

predissociation dominates the reaction rate constant. At tempera-tures above ∼10000 K process 3.c is, however, dominating.

This dominance by the process 3.c at high temperature meansthat the temperatures where the inverse predissociation makes anoticeable contribution to the total reaction rate constant will betemperatures where the highest collision energies will not matter.Thus, the high energy processes of electron detachment and chargetransfer should not interfere with the inverse predissociation.

When it comes to possible interference between the processes 3.band 3.c and inverse predissociation reactions 3.d-3.g, 3.f and 3.g goesvia electronic states that are not involved with 3.b and 3.c and willnot noticeably affect or be affected by them. 3.f is assumed to bethe dominating process for inverse predissociation for most of thetemperature interval studied.60 Thus, it can also be assumed thatinverse predissociaton will not interfere with 3.b and 3.c for most

63

Page 86: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

temperatures. A possible problem is that both reactions 3.b and3.d become important at low temperatures. However, none of theresonances in the two pathways overlap, so it is unlikely that theywould affect each other.

If looking at collisions between O- and H-atoms that may haveexcited electronic states, discarding the excited states of hydrogenis unproblematic as they are very high in energy, ∼ 10 eV.87–89

The excited states of oxygen may however matter, at the highertemperatures studied here, as O(1D) is only ∼ 2 eV87,90 above theground state. This means that there are additional reaction pathwaysthat could be significant at high temperatures, possibly increasingthe rate constant.

There is an older value for the radiative association rate con-stant, shown in figure 3.7. This older value is the one found inthe databases Kinetic Database for Astrochemistry86 (KIDA) andUMIST database for astrochemistry91 (UDfA). This older value isvery different from the newly calculated one. The old value for thereaction rate constant is likely just a rough estimate†. With theabove discussion indicating that the newly calculated rate constantshould be relatively trustworthy, it is reasonable to assume that it isof higher quality than the old rate constant. Additionally, the newrate constant covers a much larger temperature interval.

†The author has been unsuccessful in locating the original source of the oldreaction rate constant. As of writing this (20 April 2020) KIDA denotes themethod for obtaining it as “estimation”, which, according to KIDA, means thatit is a guess, but possibly an educated one.

64

3.2. Dynamics of Gaussian basis functions

3.2 Dynamics of Gaussian basis

functions

In 2011 Poulsen published a work where the Dirac-Frenkel variationalprinciple, see section 2.4, was applied to Wigner functions.92 In paperII93 two new forms of basis functions to construct Wigner functions,for use with this variational principle are described and tested ona few model systems. The purpose is to evaluate if this could be auseful method for approximate quantum dynamics.

Theory and method

The Dirac-Frenkel variational principle for wavefunctions was brieflypresented in section 2.4. The variational principle introduced byPoulsen92 is the corresponding variational principle for Wigner dis-tributions in phase space. In analogy with the Dirac-Frenkel varia-tional principle the Wigner transformed operator can be written ina parametrized form. Here the parametrized Wigner functions willbe called ΦW (x,p,α(t)), where α(t) is the vector of parameters.The equation corresponding to equation 2.74 is

∫∫dDx dDp (δΦW (x,p,α(t)))∗

×(L− d

dt

)ΦW (x,p,α(t)) = 0, (3.6)

where δΦW (x,p,α(t)) is a variation in ΦW (x,p,α(t)) correspondingto the variation δα(t) in α(t). These variations are complex, eventhough both the Wigner function, and the parameters, may be real.L is the Liouvillian operator that was shown in equation 2.72. Sincethe time dependence of the Wigner function is contained in theparameters, the time derivative of the Wigner function is

d

dtΦW (x,p,α(t)) = α(t) • d

dα(t)ΦW (x,p,α(t)) , (3.7)

65

Page 87: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

temperatures. A possible problem is that both reactions 3.b and3.d become important at low temperatures. However, none of theresonances in the two pathways overlap, so it is unlikely that theywould affect each other.

If looking at collisions between O- and H-atoms that may haveexcited electronic states, discarding the excited states of hydrogenis unproblematic as they are very high in energy, ∼ 10 eV.87–89

The excited states of oxygen may however matter, at the highertemperatures studied here, as O(1D) is only ∼ 2 eV87,90 above theground state. This means that there are additional reaction pathwaysthat could be significant at high temperatures, possibly increasingthe rate constant.

There is an older value for the radiative association rate con-stant, shown in figure 3.7. This older value is the one found inthe databases Kinetic Database for Astrochemistry86 (KIDA) andUMIST database for astrochemistry91 (UDfA). This older value isvery different from the newly calculated one. The old value for thereaction rate constant is likely just a rough estimate†. With theabove discussion indicating that the newly calculated rate constantshould be relatively trustworthy, it is reasonable to assume that it isof higher quality than the old rate constant. Additionally, the newrate constant covers a much larger temperature interval.

†The author has been unsuccessful in locating the original source of the oldreaction rate constant. As of writing this (20 April 2020) KIDA denotes themethod for obtaining it as “estimation”, which, according to KIDA, means thatit is a guess, but possibly an educated one.

64

3.2. Dynamics of Gaussian basis functions

3.2 Dynamics of Gaussian basis

functions

In 2011 Poulsen published a work where the Dirac-Frenkel variationalprinciple, see section 2.4, was applied to Wigner functions.92 In paperII93 two new forms of basis functions to construct Wigner functions,for use with this variational principle are described and tested ona few model systems. The purpose is to evaluate if this could be auseful method for approximate quantum dynamics.

Theory and method

The Dirac-Frenkel variational principle for wavefunctions was brieflypresented in section 2.4. The variational principle introduced byPoulsen92 is the corresponding variational principle for Wigner dis-tributions in phase space. In analogy with the Dirac-Frenkel varia-tional principle the Wigner transformed operator can be written ina parametrized form. Here the parametrized Wigner functions willbe called ΦW (x,p,α(t)), where α(t) is the vector of parameters.The equation corresponding to equation 2.74 is

∫∫dDx dDp (δΦW (x,p,α(t)))∗

×(L− d

dt

)ΦW (x,p,α(t)) = 0, (3.6)

where δΦW (x,p,α(t)) is a variation in ΦW (x,p,α(t)) correspondingto the variation δα(t) in α(t). These variations are complex, eventhough both the Wigner function, and the parameters, may be real.L is the Liouvillian operator that was shown in equation 2.72. Sincethe time dependence of the Wigner function is contained in theparameters, the time derivative of the Wigner function is

d

dtΦW (x,p,α(t)) = α(t) • d

dα(t)ΦW (x,p,α(t)) , (3.7)

65

Page 88: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

where α(t) is the time derivative of α(t). This means that thevariational principle also can be written

∫∫dDx dDp (δΦW (x,p,α(t)))∗

×(L− α(t) • d

dα(t)

)ΦW (x,p,α(t)) = 0. (3.8)

From this, a formula for α(t) can be obtained,

α(t) =

∫∫dDx dDp

(d

dα(t)ΦW (x,p,α(t))

)∗LΦW (x,p,α(t))

∫∫dDx dDp

(d

dα(t)ΦW (x,p,α(t))

)∗• d

dα(t)ΦW (x,p,α(t))

,

(3.9)

which is an equation of motion for the parameters.This variational principle can be written as a principle of least

action,92 with an action functional, S (α(t)), that is

S (α(t)) =

∫dt

∫∫dDx dDp (ΦW (x,p,α(t)))∗

×(L− α(t) • d

dα(t)

)ΦW (x,p,α(t)) . (3.10)

Here the principle of least action states that you should find thepath in α(t) that makes S (α(t)) an extremum. This principle ofleast action is equivalent to equations 3.6 and 3.8.

In paper II93 two new types of basis functions, φ, are suggestedfor building up the parametrized Wigner functions. As in the paper,the one-dimensional case will be shown here.

ΦW (x, p,α(t)) =num. of basis func.∑

j=0

wjφj (x, p,αj(t)) , (3.11)

where wj is the weight coefficient of the basis function and αj(t) isthe vector of parameters for each basis function. The basis functionssuggested are a so called thawed real Gaussian function

φj (x, p,αj(t)) = N e−

x− x0 (t)p− p0 (t)

Ta (t) c (t)c (t) b (t)

x− x0 (t)p− p0 (t)

(3.12)

66

3.2. Dynamics of Gaussian basis functions

and a frozen complex Gaussian function

φj (x, p,αj(t)) = eiξ(t) e−

x− x0 (t)p− p0 (t)

Ta 00 b

x− x0 (t)p− p0 (t)

,

(3.13)

where N is a normalization constant, x0 and p0 are the averageposition and momentum of the basis function, and a, b, c, and ξ areother parameters of the basis function. The vectors αj(t) for thetwo kinds of basis functions are

αj,Thawed(t) =

x0(t)p0(t)a(t)b(t)c(t)

(3.14)

αj,Frozen(t) =

x0(t)p0(t)ab

ξ(t)

. (3.15)

While all other parameters are real, the time dependent parametersfor the frozen Gaussian, x0(t), p0(t), and ξ(t), are complex numbers.Even though the individual frozen Gaussian functions are complexvalued, the total Wigner function, that is the sum of them, shouldbe real valued. The frozen Gaussian is called “frozen” because aand b have fixed values that do not change with time, and eventhough ξ(t) change with time, it only affects the size and complexphase of the function. The frozen Gaussian thus preserves its shape,and only changes its average coordinate in phase space, its size,and its complex phase, when it evolves in time. In contrast, thethawed Gaussian is “thawed” (not frozen) because a(t), b(t), andc(t) change with time. So the thawed Gaussian changes shape andaverage coordinate in phase space, but it conserves its size as itevolves in time.

A Wigner function constructed from one of the types of basisfunctions presented in paper II can be propagated in time by prop-agating the individual basis functions separately. For the thawed

67

Page 89: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

where α(t) is the time derivative of α(t). This means that thevariational principle also can be written

∫∫dDx dDp (δΦW (x,p,α(t)))∗

×(L− α(t) • d

dα(t)

)ΦW (x,p,α(t)) = 0. (3.8)

From this, a formula for α(t) can be obtained,

α(t) =

∫∫dDx dDp

(d

dα(t)ΦW (x,p,α(t))

)∗LΦW (x,p,α(t))

∫∫dDx dDp

(d

dα(t)ΦW (x,p,α(t))

)∗• d

dα(t)ΦW (x,p,α(t))

,

(3.9)

which is an equation of motion for the parameters.This variational principle can be written as a principle of least

action,92 with an action functional, S (α(t)), that is

S (α(t)) =

∫dt

∫∫dDx dDp (ΦW (x,p,α(t)))∗

×(L− α(t) • d

dα(t)

)ΦW (x,p,α(t)) . (3.10)

Here the principle of least action states that you should find thepath in α(t) that makes S (α(t)) an extremum. This principle ofleast action is equivalent to equations 3.6 and 3.8.

In paper II93 two new types of basis functions, φ, are suggestedfor building up the parametrized Wigner functions. As in the paper,the one-dimensional case will be shown here.

ΦW (x, p,α(t)) =num. of basis func.∑

j=0

wjφj (x, p,αj(t)) , (3.11)

where wj is the weight coefficient of the basis function and αj(t) isthe vector of parameters for each basis function. The basis functionssuggested are a so called thawed real Gaussian function

φj (x, p,αj(t)) = N e−

x− x0 (t)p− p0 (t)

Ta (t) c (t)c (t) b (t)

x− x0 (t)p− p0 (t)

(3.12)

66

3.2. Dynamics of Gaussian basis functions

and a frozen complex Gaussian function

φj (x, p,αj(t)) = eiξ(t) e−

x− x0 (t)p− p0 (t)

Ta 00 b

x− x0 (t)p− p0 (t)

,

(3.13)

where N is a normalization constant, x0 and p0 are the averageposition and momentum of the basis function, and a, b, c, and ξ areother parameters of the basis function. The vectors αj(t) for thetwo kinds of basis functions are

αj,Thawed(t) =

x0(t)p0(t)a(t)b(t)c(t)

(3.14)

αj,Frozen(t) =

x0(t)p0(t)ab

ξ(t)

. (3.15)

While all other parameters are real, the time dependent parametersfor the frozen Gaussian, x0(t), p0(t), and ξ(t), are complex numbers.Even though the individual frozen Gaussian functions are complexvalued, the total Wigner function, that is the sum of them, shouldbe real valued. The frozen Gaussian is called “frozen” because aand b have fixed values that do not change with time, and eventhough ξ(t) change with time, it only affects the size and complexphase of the function. The frozen Gaussian thus preserves its shape,and only changes its average coordinate in phase space, its size,and its complex phase, when it evolves in time. In contrast, thethawed Gaussian is “thawed” (not frozen) because a(t), b(t), andc(t) change with time. So the thawed Gaussian changes shape andaverage coordinate in phase space, but it conserves its size as itevolves in time.

A Wigner function constructed from one of the types of basisfunctions presented in paper II can be propagated in time by prop-agating the individual basis functions separately. For the thawed

67

Page 90: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

Gaussian, the equations of motion, according to the variationalprinciple, are

x0(t) =p0(t)

m(3.16)

p0(t) = − dVSm (x0(t))

dx0(t)(3.17)

a(t) = 2c(t)d2VSm (x0(t))

d (x0(t))2 (3.18)

b(t) = −2c(t)

m(3.19)

c(t) = b(t)d2VSm (x0(t))

d (x0(t))2 − a(t)

m, (3.20)

where VSm (x0(t)) is a smeared potential defined by

VSm (x0(t)) =1

√πb′(t)

∫dη V (η) e

− (x0(t)−η)2

2b′(t) (3.21)

b′(t) =b(t)

2

(1

2 (a(t)b(t)− c(t)2)+ 1

). (3.22)

a(t)b(t)−c(t)2 is a constant of motion, and for a minimum uncertaintyWigner function, i.e. corresponding to a single state wavefunction,a(t)b(t) − c(t)2 = 1/2, resulting in b′(t) = b(t). In this particularsituation the proposed method is equivalent to methods that havebeen used before for Gaussian wavefunctions.94,95

For the frozen Gaussian, the equations of motion, according to

68

3.2. Dynamics of Gaussian basis functions

the variational principle, are

x0,R(t) =p0,R(t)

m(3.23)

x0,I(t) =1

22abd

dp0,I(t)

(VSm (x0,R(t)−bp0,I(t)) + VSm (x0,R(t)+bp0,I(t))) (3.24)

p0,R(t) = −1

2

d

dx0,R(t)

(VSm (x0,R(t)−bp0,I(t)) + VSm (x0,R(t)+bp0,I(t))) (3.25)

p0,I(t) = −ax0,I(t)

mb(3.26)

ξR(t) =1

(VSm (x0,R(t)+bp0,I(t))− VSm (x0,R(t)−bp0,I(t)))

− p0,I(t)bd

dx0,R(t)

(VSm (x0,R(t)+bp0,I(t)) + VSm (x0,R(t)−bp0,I(t))) (3.27)

ξI(t) = 2ax0,I(t)x0,I(t) + 2bp0,I(t)p0,I(t) (3.28)

a = b = 0, (3.29)

where x0(t), p0(t), and ξ(t) have been divided into their real andimaginary parts, respectively denoted by R and I. b′(t) for thesmeared potential is for the frozen Gaussian defined as before, butwith c = 0.

Both of the types of basis functions presented are mathematicallycomplete, i.e. any function of x and p can be exactly expressed as asum of such Gaussians. An additional useful property of the frozenGaussians is that when summing them up to obtain the total Wignerfunction they can in principle account for quantum interference, dueto their complex phase. Another effect of this complex phase is thatthe total Wigner function will not keep its normalization. Everytime one wants to use the time evolved Wigner function it has tobe renormalized. The thawed Gaussians will keep their individualnormalization and they are completely positive real-valued, so thetotal Wigner function built from them will keep its nomalizationover time, but they can not account for quantum interference.

69

Page 91: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

Gaussian, the equations of motion, according to the variationalprinciple, are

x0(t) =p0(t)

m(3.16)

p0(t) = − dVSm (x0(t))

dx0(t)(3.17)

a(t) = 2c(t)d2VSm (x0(t))

d (x0(t))2 (3.18)

b(t) = −2c(t)

m(3.19)

c(t) = b(t)d2VSm (x0(t))

d (x0(t))2 − a(t)

m, (3.20)

where VSm (x0(t)) is a smeared potential defined by

VSm (x0(t)) =1

πb′(t)

∫dη V (η) e

− (x0(t)−η)2

2b′(t) (3.21)

b′(t) =b(t)

2

(1

2 (a(t)b(t)− c(t)2)+ 1

). (3.22)

a(t)b(t)−c(t)2 is a constant of motion, and for a minimum uncertaintyWigner function, i.e. corresponding to a single state wavefunction,a(t)b(t) − c(t)2 = 1/2, resulting in b′(t) = b(t). In this particularsituation the proposed method is equivalent to methods that havebeen used before for Gaussian wavefunctions.94,95

For the frozen Gaussian, the equations of motion, according to

68

3.2. Dynamics of Gaussian basis functions

the variational principle, are

x0,R(t) =p0,R(t)

m(3.23)

x0,I(t) =1

22abd

dp0,I(t)

(VSm (x0,R(t)−bp0,I(t)) + VSm (x0,R(t)+bp0,I(t))) (3.24)

p0,R(t) = −1

2

d

dx0,R(t)

(VSm (x0,R(t)−bp0,I(t)) + VSm (x0,R(t)+bp0,I(t))) (3.25)

p0,I(t) = −ax0,I(t)

mb(3.26)

ξR(t) =1

(VSm (x0,R(t)+bp0,I(t))− VSm (x0,R(t)−bp0,I(t)))

− p0,I(t)bd

dx0,R(t)

(VSm (x0,R(t)+bp0,I(t)) + VSm (x0,R(t)−bp0,I(t))) (3.27)

ξI(t) = 2ax0,I(t)x0,I(t) + 2bp0,I(t)p0,I(t) (3.28)

a = b = 0, (3.29)

where x0(t), p0(t), and ξ(t) have been divided into their real andimaginary parts, respectively denoted by R and I. b′(t) for thesmeared potential is for the frozen Gaussian defined as before, butwith c = 0.

Both of the types of basis functions presented are mathematicallycomplete, i.e. any function of x and p can be exactly expressed as asum of such Gaussians. An additional useful property of the frozenGaussians is that when summing them up to obtain the total Wignerfunction they can in principle account for quantum interference, dueto their complex phase. Another effect of this complex phase is thatthe total Wigner function will not keep its normalization. Everytime one wants to use the time evolved Wigner function it has tobe renormalized. The thawed Gaussians will keep their individualnormalization and they are completely positive real-valued, so thetotal Wigner function built from them will keep its nomalizationover time, but they can not account for quantum interference.

69

Page 92: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

−1.5

−1.0

−0.5

0.0

0.5

1.0

1.5

0 5 10 15 20 25 30 35

⟨ x∧ (

t) ⟩

/ (

− h1

/2m

−1

/2ω

−1

/2)

t / ω−1

CWThawed

QM

Figure 3.8: Average position of a particle in the double well potentialas a function of time. Results from the classical Wigner method(CW), a single thawed Gaussian function (Thawed), and accuratequantum mechanics (QM).

Both types of basis functions handle harmonic potentials exactly.For the frozen Gaussian basis functions the complex phase disappearsfor harmonic potentials, making the computations simpler. Forpotentials close to harmonic the complex phase factor, eiξR(t), willbe small.

Results and discussion

The new basis functions and equations of motion were tested for twodifferent one-dimensional systems, a quartic double well potential,

V (x) =m2ω3

10x4 − mω2

2x2, and a quartic potential V (x) =

m2ω3

4x4,

where ω is a unit of angular frequency. ω is also the absolute valueof the angular frequency on top of the barrier in the double well.

For a Wigner function corresponding to the ground state of anoscillation in one of the wells in the double well, the average positionas a function of time is shown in figure 3.8. Here a comparison can

70

3.2. Dynamics of Gaussian basis functions

0

10

20

30

40

50

0 1 2 3 4

tunnelin

g p

eriod / ω

−1

absolute value of barrier frequency / ω

ThawedQM

Figure 3.9: Tunneling periods of double well potentials with differentbarrier frequencies, but the same well depth. Results from thawedGaussian functions (Thawed) and from accurate quantum mechanics(QM).

be made between the use of a single thawed Gaussian basis function,the classical Wigner method, and accurate quantum mechanics. Theclassical Wigner method and the exact quantum mechanics are onlyused to calculate the dynamics, the initial distribution is still theGaussian basis function. It can be seen that the thawed Gaussians,quite well, account for the tunneling between the potential wells.The frequency of the oscillation is, however, a bit too slow. Thedynamics of the thawed Gaussian function is a clear improvementover the classical Wigner method. The classical Wigner function canbe seen to be stuck in the starting well, due to it not being able toaccount for dynamic tunneling.

The effect of adjusting the double well potential so that thefrequency of barrier changed, but the depth of the wells stayed thesame, was studied. In figure 3.9 the tunneling periods resulting fromthis are shown. The thawed Gaussian basis function gives goodresults for the tunneling period, except for the lowest frequency

71

Page 93: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

−1.5

−1.0

−0.5

0.0

0.5

1.0

1.5

0 5 10 15 20 25 30 35

⟨ x∧ (

t) ⟩

/ (

− h1

/2m

−1

/2ω

−1

/2)

t / ω−1

CWThawed

QM

Figure 3.8: Average position of a particle in the double well potentialas a function of time. Results from the classical Wigner method(CW), a single thawed Gaussian function (Thawed), and accuratequantum mechanics (QM).

Both types of basis functions handle harmonic potentials exactly.For the frozen Gaussian basis functions the complex phase disappearsfor harmonic potentials, making the computations simpler. Forpotentials close to harmonic the complex phase factor, eiξR(t), willbe small.

Results and discussion

The new basis functions and equations of motion were tested for twodifferent one-dimensional systems, a quartic double well potential,

V (x) =m2ω3

10x4 − mω2

2x2, and a quartic potential V (x) =

m2ω3

4x4,

where ω is a unit of angular frequency. ω is also the absolute valueof the angular frequency on top of the barrier in the double well.

For a Wigner function corresponding to the ground state of anoscillation in one of the wells in the double well, the average positionas a function of time is shown in figure 3.8. Here a comparison can

70

3.2. Dynamics of Gaussian basis functions

0

10

20

30

40

50

0 1 2 3 4

tunnelin

g p

eriod / ω

−1

absolute value of barrier frequency / ω

ThawedQM

Figure 3.9: Tunneling periods of double well potentials with differentbarrier frequencies, but the same well depth. Results from thawedGaussian functions (Thawed) and from accurate quantum mechanics(QM).

be made between the use of a single thawed Gaussian basis function,the classical Wigner method, and accurate quantum mechanics. Theclassical Wigner method and the exact quantum mechanics are onlyused to calculate the dynamics, the initial distribution is still theGaussian basis function. It can be seen that the thawed Gaussians,quite well, account for the tunneling between the potential wells.The frequency of the oscillation is, however, a bit too slow. Thedynamics of the thawed Gaussian function is a clear improvementover the classical Wigner method. The classical Wigner function canbe seen to be stuck in the starting well, due to it not being able toaccount for dynamic tunneling.

The effect of adjusting the double well potential so that thefrequency of barrier changed, but the depth of the wells stayed thesame, was studied. In figure 3.9 the tunneling periods resulting fromthis are shown. The thawed Gaussian basis function gives goodresults for the tunneling period, except for the lowest frequency

71

Page 94: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

−1.5

−1.0

−0.5

0.0

0.5

1.0

1.5

0 5 10 15 20 25 30

⟨ x∧ (

t) ⟩

/ (

− h1

/2m

−1

/2ω

−1

/2)

t / ω−1

CWFrozen

QM

Figure 3.10: Average position of a particle in the double well potentialas a function of time. Results from the classical Wigner method(CW), a Wigner function consisting of frozen Gaussian basis functions(Frozen), and accurate quantum mechanics (QM).

barriers, i.e. the widest ones.

Also, the frozen Gaussian basis functions were tested for the dou-ble well potential. For the standard double well the average positionas a function of time for a Wigner function starting in one of thewells is shown in figure 3.10. As for figure 3.8, the classical Wignermethod and the exact quantum mechanics are only used to calculatethe dynamics, the initial distribution is the one described by theGaussian basis functions. In contrast to the thawed Gaussian case,where only a single basis function was used, the Wigner functionis here described by thousands of frozen Gaussian basis functions,making the variational dynamics much more computationally de-manding. The frozen Gaussians, as the thawed ones, can be seento account quite well for tunneling. For the frozen Gaussians thefrequency of the oscillation is, however, significantly too fast. Thefrozen Gaussians definitely account for tunneling better than theclassical Wigner method does, but it does not accurately reproduce

72

3.2. Dynamics of Gaussian basis functions

−1.0

−0.8

−0.5

−0.2

0.0

0.2

0.5

0.8

1.0

0 5 10 15

⟨ x∧ (

t) ⟩

/ (

− h1

/2m

−1

/2ω

−1

/2)

t / ω−1

CWFrozen

QM

Figure 3.11: Average position of a particle in a quartic potential asa function of time. Results from the classical Wigner method (CW),a Wigner function consisting of frozen Gaussian basis functions(Frozen), and accurate quantum mechanics (QM).

quantum mechanical results for any time except t = 0.

The average position of a Wigner function, initialized as a Gaus-sian distribution on one side of the quartic potential, was calculatedwith the frozen Gaussian basis functions, and is displayed in figure3.11. For this situation dynamic tunneling is not nearly as importantas for the double well. For short times, the frozen Gaussian basisfunctions describe the accurate dynamics quite well, but worse thanthe classical Wigner method. The frequency of the dynamics of thefrozen Gaussian functions is too high, but the correlation functionretains a reasonable amplitude of the oscillation for much longer thanthe classical Wigner method does. In paper II it is shown that thefrozen Gaussian basis functions describe the quantum interferenceon this quartic potential qualitatively, but not quantitatively.

Using the Feynman-Kleinert approximation96 for the Boltzmanoperator, an analytic expression for the initial distribution for〈xx(t)〉Kubo, as described in section 2.4, for use with the thawed

73

Page 95: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

−1.5

−1.0

−0.5

0.0

0.5

1.0

1.5

0 5 10 15 20 25 30

⟨ x∧ (

t) ⟩

/ (

− h1

/2m

−1

/2ω

−1

/2)

t / ω−1

CWFrozen

QM

Figure 3.10: Average position of a particle in the double well potentialas a function of time. Results from the classical Wigner method(CW), a Wigner function consisting of frozen Gaussian basis functions(Frozen), and accurate quantum mechanics (QM).

barriers, i.e. the widest ones.

Also, the frozen Gaussian basis functions were tested for the dou-ble well potential. For the standard double well the average positionas a function of time for a Wigner function starting in one of thewells is shown in figure 3.10. As for figure 3.8, the classical Wignermethod and the exact quantum mechanics are only used to calculatethe dynamics, the initial distribution is the one described by theGaussian basis functions. In contrast to the thawed Gaussian case,where only a single basis function was used, the Wigner functionis here described by thousands of frozen Gaussian basis functions,making the variational dynamics much more computationally de-manding. The frozen Gaussians, as the thawed ones, can be seento account quite well for tunneling. For the frozen Gaussians thefrequency of the oscillation is, however, significantly too fast. Thefrozen Gaussians definitely account for tunneling better than theclassical Wigner method does, but it does not accurately reproduce

72

3.2. Dynamics of Gaussian basis functions

−1.0

−0.8

−0.5

−0.2

0.0

0.2

0.5

0.8

1.0

0 5 10 15

⟨ x∧ (

t) ⟩

/ (

− h1

/2m

−1

/2ω

−1

/2)

t / ω−1

CWFrozen

QM

Figure 3.11: Average position of a particle in a quartic potential asa function of time. Results from the classical Wigner method (CW),a Wigner function consisting of frozen Gaussian basis functions(Frozen), and accurate quantum mechanics (QM).

quantum mechanical results for any time except t = 0.

The average position of a Wigner function, initialized as a Gaus-sian distribution on one side of the quartic potential, was calculatedwith the frozen Gaussian basis functions, and is displayed in figure3.11. For this situation dynamic tunneling is not nearly as importantas for the double well. For short times, the frozen Gaussian basisfunctions describe the accurate dynamics quite well, but worse thanthe classical Wigner method. The frequency of the dynamics of thefrozen Gaussian functions is too high, but the correlation functionretains a reasonable amplitude of the oscillation for much longer thanthe classical Wigner method does. In paper II it is shown that thefrozen Gaussian basis functions describe the quantum interferenceon this quartic potential qualitatively, but not quantitatively.

Using the Feynman-Kleinert approximation96 for the Boltzmanoperator, an analytic expression for the initial distribution for〈xx(t)〉Kubo, as described in section 2.4, for use with the thawed

73

Page 96: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

−0.125

−0.100

−0.075

−0.050

−0.025

0.000

0.025

0.050

0.075

0.100

0.125

0 5 10 15 20 25 30 35

⟨ x∧ x∧

(t)

⟩K

ub

o / (

− hm

−1ω

−1)

t / ω−1

CWThawed

QM

Figure 3.12: Kubo transformed position autocorrelation functionfor the quartic potential at βω = 8. Results from the classicalWigner method (CW), thawed Gaussian basis functions (Thawed),and accurate quantum mechanics (QM).

Gaussian basis functions, could be obtained. This expression wasused to test the thawed Gaussian basis functions. In figures 3.12 and3.13, 〈xx(t)〉Kubo for the quartic potential at temperatures βω = 8and βω = 1 can be seen. At the lower temperature the thawedGaussian basis functions give a good approximation to the quantummechanical correlation function. At the higher temperature thethawed Gaussian basis functions, except for at very short times, failto capture either the amplitude or the frequency of the correlationfunction. A similar situation for the double well potential is shown inthe paper. That the thawed Gaussian functions from the Feynman-Kleinert approximation do not work well at high temperatures isdue to that the division into basis functions, at high temperature,may lead to the propagation of individual basis functions that areunphysical according to Heisenbergs uncertainty principle, i.e. beingtoo localized in both position and momentum simultaneously.

In summary the thawed Gaussian basis functions seem to be good

74

3.2. Dynamics of Gaussian basis functions

−0.75

−0.50

−0.25

0.00

0.25

0.50

0.75

0 5 10 15 20 25 30 35

⟨ x∧ x∧

(t)

⟩K

ub

o / (

− hm

−1ω

−1)

t / ω−1

CWThawed

QM

Figure 3.13: Kubo transformed position autocorrelation functionfor the quartic potential at βω = 1. Results from the classicalWigner method (CW), thawed Gaussian basis functions (Thawed),and accurate quantum mechanics (QM).

at describing tunneling, while the frozen Gaussian basis functions candescribe interference. However, the dynamics of the frozen Gaussianbasis functions are too fast.

75

Page 97: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

−0.125

−0.100

−0.075

−0.050

−0.025

0.000

0.025

0.050

0.075

0.100

0.125

0 5 10 15 20 25 30 35

⟨ x∧ x∧

(t)

⟩K

ub

o / (

− hm

−1ω

−1)

t / ω−1

CWThawed

QM

Figure 3.12: Kubo transformed position autocorrelation functionfor the quartic potential at βω = 8. Results from the classicalWigner method (CW), thawed Gaussian basis functions (Thawed),and accurate quantum mechanics (QM).

Gaussian basis functions, could be obtained. This expression wasused to test the thawed Gaussian basis functions. In figures 3.12 and3.13, 〈xx(t)〉Kubo for the quartic potential at temperatures βω = 8and βω = 1 can be seen. At the lower temperature the thawedGaussian basis functions give a good approximation to the quantummechanical correlation function. At the higher temperature thethawed Gaussian basis functions, except for at very short times, failto capture either the amplitude or the frequency of the correlationfunction. A similar situation for the double well potential is shown inthe paper. That the thawed Gaussian functions from the Feynman-Kleinert approximation do not work well at high temperatures isdue to that the division into basis functions, at high temperature,may lead to the propagation of individual basis functions that areunphysical according to Heisenbergs uncertainty principle, i.e. beingtoo localized in both position and momentum simultaneously.

In summary the thawed Gaussian basis functions seem to be good

74

3.2. Dynamics of Gaussian basis functions

−0.75

−0.50

−0.25

0.00

0.25

0.50

0.75

0 5 10 15 20 25 30 35

⟨ x∧ x∧

(t)

⟩K

ub

o / (

− hm

−1ω

−1)

t / ω−1

CWThawed

QM

Figure 3.13: Kubo transformed position autocorrelation functionfor the quartic potential at βω = 1. Results from the classicalWigner method (CW), thawed Gaussian basis functions (Thawed),and accurate quantum mechanics (QM).

at describing tunneling, while the frozen Gaussian basis functions candescribe interference. However, the dynamics of the frozen Gaussianbasis functions are too fast.

75

Page 98: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

3.3 The open polymer classical Wigner

method

A new method of computing initial phase space distributions forthe classical Wigner method, presented in sections 2.1 and 2.2, waspublished in paper III.97 In this method a discretized imaginary timepath integral, as presented in section 2.2, is used to sample a thermalquantum distribution. The new method was given the name “Openpolymer classical Wigner”. As can be surmised from the name, thepolymer corresponding to the imaginary time path integral is nota closed ring, but have an opening. Paper IV is a manuscript inwhich the new variant of the classical Wigner method is applied tothe calculation of reaction rate constants through the flux-Heavisidetrace presented in section 2.3.

Derivation of the open polymer expression

The canonical correlation function⟨AB (t)

⟩can be written

⟨AB (t)

⟩=

1

ZTr

A e−βH e

iHt B e−

iHt

=1

Z

∫dDx

⟨x∣∣∣A e−βH e

iHt B e−

iHt

∣∣∣x⟩

=1

Z

∫dDx

⟨x∣∣∣A e−βH B(t)

∣∣∣x⟩. (3.30)

By inserting N − 1 identity operators and dividing the Boltzmannoperator into many parts, as was done in section 2.2, this can bechanged into

⟨AB (t)

⟩=

1

Z

N∏

j=1

∫dDxj

⟨x1

∣∣∣A e−βNH∣∣∣x2

×⟨x2

∣∣∣e− βNH∣∣∣x3

⟩. . .

⟨xN−1

∣∣∣e− βNH∣∣∣xN

×⟨xN

∣∣∣e− βNH B(t)

∣∣∣x1

⟩. (3.31)

All the matrix elements can now be exchanged for inverse Fouriertransforms of Wigner transforms, with the help of equation 2.26,

76

3.3. The open polymer classical Wigner method

giving

⟨AB (t)

⟩=

1

Z

N∏

j=1

∫∫dDxj d

Dpj

(2π)Dei(xj−x(j mod N)+1)•pj/

×[A e−

βNH]W

(x1 + x2

2,p1

)

×[e−

βNH]W

(x2 + x3

2,p2

)

. . .[e−

βNH]W

(xN−1 + xN

2,pN−1

)

×[e−

βNH B(t)

]W

(xN + x1

2,pN

). (3.32)

In the limit where N → ∞, or equivalently the high temperaturelimit, the Trotter product,32

limN→∞

e−βNH = e−

β2N

V e−βNT e−

β2N

V , (3.33)

can be used, meaning that

limN→∞

[e−

βNH]W(x,p) = lim

N→∞

[e−

β2N

V e−βNT e−

β2N

V]W(x,p)

= limN→∞

[e−

β2N

V]W(x,p) e

− i2

( ←∂∂p

•→∂∂x

−←∂∂x

•→∂∂p

)

×[e−

βNT]W(x,p) e

− i2

( ←∂∂p

•→∂∂x

−←∂∂x

•→∂∂p

) [e−

β2N

V]W(x,p) ,

(3.34)

where equation 2.4 has been used. These Wigner transforms areeasy to evaluate, as long as the potential is restricted to only depend

77

Page 99: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

3.3 The open polymer classical Wigner

method

A new method of computing initial phase space distributions forthe classical Wigner method, presented in sections 2.1 and 2.2, waspublished in paper III.97 In this method a discretized imaginary timepath integral, as presented in section 2.2, is used to sample a thermalquantum distribution. The new method was given the name “Openpolymer classical Wigner”. As can be surmised from the name, thepolymer corresponding to the imaginary time path integral is nota closed ring, but have an opening. Paper IV is a manuscript inwhich the new variant of the classical Wigner method is applied tothe calculation of reaction rate constants through the flux-Heavisidetrace presented in section 2.3.

Derivation of the open polymer expression

The canonical correlation function⟨AB (t)

⟩can be written

⟨AB (t)

⟩=

1

ZTr

A e−βH e

iHt B e−

iHt

=1

Z

∫dDx

⟨x∣∣∣A e−βH e

iHt B e−

iHt

∣∣∣x⟩

=1

Z

∫dDx

⟨x∣∣∣A e−βH B(t)

∣∣∣x⟩. (3.30)

By inserting N − 1 identity operators and dividing the Boltzmannoperator into many parts, as was done in section 2.2, this can bechanged into

⟨AB (t)

⟩=

1

Z

N∏

j=1

∫dDxj

⟨x1

∣∣∣A e−βNH∣∣∣x2

×⟨x2

∣∣∣e− βNH∣∣∣x3

⟩. . .

⟨xN−1

∣∣∣e− βNH∣∣∣xN

×⟨xN

∣∣∣e− βNH B(t)

∣∣∣x1

⟩. (3.31)

All the matrix elements can now be exchanged for inverse Fouriertransforms of Wigner transforms, with the help of equation 2.26,

76

3.3. The open polymer classical Wigner method

giving

⟨AB (t)

⟩=

1

Z

N∏

j=1

∫∫dDxj d

Dpj

(2π)Dei(xj−x(j mod N)+1)•pj/

×[A e−

βNH]W

(x1 + x2

2,p1

)

×[e−

βNH]W

(x2 + x3

2,p2

)

. . .[e−

βNH]W

(xN−1 + xN

2,pN−1

)

×[e−

βNH B(t)

]W

(xN + x1

2,pN

). (3.32)

In the limit where N → ∞, or equivalently the high temperaturelimit, the Trotter product,32

limN→∞

e−βNH = e−

β2N

V e−βNT e−

β2N

V , (3.33)

can be used, meaning that

limN→∞

[e−

βNH]W(x,p) = lim

N→∞

[e−

β2N

V e−βNT e−

β2N

V]W(x,p)

= limN→∞

[e−

β2N

V]W(x,p) e

− i2

( ←∂∂p

•→∂∂x

−←∂∂x

•→∂∂p

)

×[e−

βNT]W(x,p) e

− i2

( ←∂∂p

•→∂∂x

−←∂∂x

•→∂∂p

) [e−

β2N

V]W(x,p) ,

(3.34)

where equation 2.4 has been used. These Wigner transforms areeasy to evaluate, as long as the potential is restricted to only depend

77

Page 100: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

on the position.

[e−

βNT]W(x,p)

=

∫dDλ eix•λ/

⟨p+

λ

2

∣∣∣∣e− β

N12(m−1p)•p

∣∣∣∣p− λ

2

=

∫dDλ eix•λ/ e−

βN

12(m−1(p+λ

2 ))•(p+λ2 ) δ (λ)

= e−βN

12(m−1p)•p (3.35)[

e−β2N

V]W(x,p)

=

∫dDη e−iη•p/

⟨x+

η

2

∣∣∣e− β2N

V (x)∣∣∣x− η

2

=

∫dDη e−iη•p/ e−

β2N

V (x+η2 ) δ (η)

= e−β2N

V (x) (3.36)

78

3.3. The open polymer classical Wigner method

Putting these results into equation 3.34 and Taylor expanding gives

limN→∞

[e−

βNH]W(x,p)

= limN→∞

e−β2N

V (x) e− i

2

( ←∂∂p

•→∂∂x

−←∂∂x

•→∂∂p

)

e−βN

12(m−1p)•p

× e− i

2

( ←∂∂p

•→∂∂x

−←∂∂x

•→∂∂p

)

e−β2N

V (x)

= limN→∞

e−β2N

V (x) e− i

2

( ←∂∂p

•→∂∂x

−←∂∂x

•→∂∂p

)

×(e−

βN

12(m−1p)•p

(1 +O

(β2

N2

))e−

β2N

V (x)

)

= limN→∞

e−β2N

V (x)

(1 +O

(β2

N2

))e−

βN

12(m−1p)•p e−

β2N

V (x)

+i2e−

β2N

V (x) O(β2

N2

)e−

βN

12(m−1p)•p e−

β2N

V (x)

+O(β2

N2e−

βNH(x,p)

)

= limN→∞

e−βNH(x,p) +O

(β2

N2e−

βNH(x,p)

)

= limN→∞

e−βNH(x,p), (3.37)

where H (x,p) is the classical Hamiltonian. O (β2/N2) shows theorder of the error, using the big O notation. Using this result, theoperator A can be separated from the Boltzmann operator in its

79

Page 101: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

on the position.

[e−

βNT]W(x,p)

=

∫dDλ eix•λ/

⟨p+

λ

2

∣∣∣∣e− β

N12(m−1p)•p

∣∣∣∣p− λ

2

=

∫dDλ eix•λ/ e−

βN

12(m−1(p+λ

2 ))•(p+λ2 ) δ (λ)

= e−βN

12(m−1p)•p (3.35)[

e−β2N

V]W(x,p)

=

∫dDη e−iη•p/

⟨x+

η

2

∣∣∣e− β2N

V (x)∣∣∣x− η

2

=

∫dDη e−iη•p/ e−

β2N

V (x+η2 ) δ (η)

= e−β2N

V (x) (3.36)

78

3.3. The open polymer classical Wigner method

Putting these results into equation 3.34 and Taylor expanding gives

limN→∞

[e−

βNH]W(x,p)

= limN→∞

e−β2N

V (x) e− i

2

( ←∂∂p

•→∂∂x

−←∂∂x

•→∂∂p

)

e−βN

12(m−1p)•p

× e− i

2

( ←∂∂p

•→∂∂x

−←∂∂x

•→∂∂p

)

e−β2N

V (x)

= limN→∞

e−β2N

V (x) e− i

2

( ←∂∂p

•→∂∂x

−←∂∂x

•→∂∂p

)

×(e−

βN

12(m−1p)•p

(1 +O

(β2

N2

))e−

β2N

V (x)

)

= limN→∞

e−β2N

V (x)

(1 +O

(β2

N2

))e−

βN

12(m−1p)•p e−

β2N

V (x)

+i2e−

β2N

V (x) O(β2

N2

)e−

βN

12(m−1p)•p e−

β2N

V (x)

+O(β2

N2e−

βNH(x,p)

)

= limN→∞

e−βNH(x,p) +O

(β2

N2e−

βNH(x,p)

)

= limN→∞

e−βNH(x,p), (3.37)

where H (x,p) is the classical Hamiltonian. O (β2/N2) shows theorder of the error, using the big O notation. Using this result, theoperator A can be separated from the Boltzmann operator in its

79

Page 102: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

Wigner transform,

limN→∞

[A e−

βNH]W(x,p)

= limN→∞

[A]W(x,p) e

− i2

( ←∂∂p

•→∂∂x

−←∂∂x

•→∂∂p

) [e−

βNH]W(x,p)

= limN→∞

[A]W(x,p) e

− i2

( ←∂∂p

•→∂∂x

−←∂∂x

•→∂∂p

)

e−βNH(x,p)

= limN→∞

[A]W(x,p)

(1 +O

N

))e−

βNH(x,p)

= limN→∞

[A]W(x,p) e−

βNH(x,p) +O

Ne−

βNH(x,p)

)

= limN→∞

[A]W(x,p) e−

βNH(x,p) . (3.38)

The same thing can be done for B(t),

limN→∞

[e−

βNH B(t)

]W(x,p)

= limN→∞

e−βNH(x,p)

[B(t)

]W(x,p) . (3.39)

Inserting the simplifications of the Wigner transforms, equations3.37-3.39, into equation 3.32 gives

⟨AB (t)

⟩= lim

N→∞

1

Z

N∏

j=1

∫∫dDxj dDpj

(2π)Dei(xj−x(j mod N)+1)•pj/

×[A]W

(x1 + x2

2,p1

)e−

βNH(x1+x2

2,p1)

× e−βNH(x2+x3

2,p2) . . . e

− βNH(

xN−1+xN2

,pN−1

)

× e−βNH(xN+x1

2,pN)

[B(t)

]W

(xN + x1

2,pN

).

(3.40)

80

3.3. The open polymer classical Wigner method

The potential energy has already been assumed to only depend onposition, not momentum, so

⟨AB (t)

⟩= lim

N→∞

1

Z

N∏

j=1

∫∫dDxj dDpj

(2π)Dei(xj−x(j mod N)+1)•pj/

×[A]W

(x1 + x2

2,p1

)e−

βN (

12(m−1p1)•p1+V (x1+x2

2 ))

× e−βN (

12(m−1p2)•p2+V (x2+x3

2 ))

. . . e− β

N

(12(m−1pN−1)•pN−1+V

(xN−1+xN

2

))

× e−βN (

12(m−1pN)•pN+V (xN+x1

2 ))

×[B(t)

]W

(xN + x1

2,pN

). (3.41)

By integrating over the momenta, as in section 2.2 we obtain

⟨AB (t)

⟩= lim

N→∞

1

Z

N∏

j=1

∫dDxj

(2π)D

×((

2πN

β

)D2 √

det (m)

)N−2

×∫∫

dDp1 dDpN

× e−βN (

12(m−1p1)•p1+V (x1+x2

2 )− iNβ (x1−x2)•p1)

× e−βN (

12(m−1pN)•pN+V (xN+x1

2 )− iNβ (xN−x1)•pN)

×[A]W

(x1 + x2

2,p1

)

× e− β

N

∑N−1j=2

(N2

2β22 (m(xj−xj+1))•(xj−xj+1)+V(

xj+xj+12

))

×[B(t)

]W

(xN + x1

2,pN

). (3.42)

Here p1 and pN have not been integrated out since they occur in[A]W(x1+x2

2,p1) and

[B(t)

]W(xN+x1

2,pN), which are unknown func-

tions. For the method derived here to work, we must restrict A to

81

Page 103: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

Wigner transform,

limN→∞

[A e−

βNH]W(x,p)

= limN→∞

[A]W(x,p) e

− i2

( ←∂∂p

•→∂∂x

−←∂∂x

•→∂∂p

) [e−

βNH]W(x,p)

= limN→∞

[A]W(x,p) e

− i2

( ←∂∂p

•→∂∂x

−←∂∂x

•→∂∂p

)

e−βNH(x,p)

= limN→∞

[A]W(x,p)

(1 +O

N

))e−

βNH(x,p)

= limN→∞

[A]W(x,p) e−

βNH(x,p) +O

Ne−

βNH(x,p)

)

= limN→∞

[A]W(x,p) e−

βNH(x,p) . (3.38)

The same thing can be done for B(t),

limN→∞

[e−

βNH B(t)

]W(x,p)

= limN→∞

e−βNH(x,p)

[B(t)

]W(x,p) . (3.39)

Inserting the simplifications of the Wigner transforms, equations3.37-3.39, into equation 3.32 gives

⟨AB (t)

⟩= lim

N→∞

1

Z

N∏

j=1

∫∫dDxj dDpj

(2π)Dei(xj−x(j mod N)+1)•pj/

×[A]W

(x1 + x2

2,p1

)e−

βNH(x1+x2

2,p1)

× e−βNH(x2+x3

2,p2) . . . e

− βNH(

xN−1+xN2

,pN−1

)

× e−βNH(xN+x1

2,pN)

[B(t)

]W

(xN + x1

2,pN

).

(3.40)

80

3.3. The open polymer classical Wigner method

The potential energy has already been assumed to only depend onposition, not momentum, so

⟨AB (t)

⟩= lim

N→∞

1

Z

N∏

j=1

∫∫dDxj dDpj

(2π)Dei(xj−x(j mod N)+1)•pj/

×[A]W

(x1 + x2

2,p1

)e−

βN (

12(m−1p1)•p1+V (x1+x2

2 ))

× e−βN (

12(m−1p2)•p2+V (x2+x3

2 ))

. . . e− β

N

(12(m−1pN−1)•pN−1+V

(xN−1+xN

2

))

× e−βN (

12(m−1pN)•pN+V (xN+x1

2 ))

×[B(t)

]W

(xN + x1

2,pN

). (3.41)

By integrating over the momenta, as in section 2.2 we obtain

⟨AB (t)

⟩= lim

N→∞

1

Z

N∏

j=1

∫dDxj

(2π)D

×((

2πN

β

)D2 √

det (m)

)N−2

×∫∫

dDp1 dDpN

× e−βN (

12(m−1p1)•p1+V (x1+x2

2 )− iNβ (x1−x2)•p1)

× e−βN (

12(m−1pN)•pN+V (xN+x1

2 )− iNβ (xN−x1)•pN)

×[A]W

(x1 + x2

2,p1

)

× e− β

N

∑N−1j=2

(N2

2β22 (m(xj−xj+1))•(xj−xj+1)+V(

xj+xj+12

))

×[B(t)

]W

(xN + x1

2,pN

). (3.42)

Here p1 and pN have not been integrated out since they occur in[A]W(x1+x2

2,p1) and

[B(t)

]W(xN+x1

2,pN), which are unknown func-

tions. For the method derived here to work, we must restrict A to

81

Page 104: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

operators whose Wigner transforms are such functions of p1 thatthe integral

∫dDp1 e−

βN (

12(m−1p1)•p1− iN

β (x1−x2)•p1)[A]W

(x1 + x2

2,p1

)

(3.43)

is analytically solvable, giving yet another spring term. This is

not such a severe restriction since at least all[A]W(x1+x2

2,p1) that

are polynomials with regard to p1 work fine. To simplify laterexpressions, a new function is defined:

A′(x1 + x2

2,x1 − x2

)

=

∫dDp1 e−

βN (

12(m−1p1)•p1− iN

β (x1−x2)•p1)[A]W

(x1+x2

2,p1

)

∫dDp1 e−

βN (

12(m−1p1)•p1− iN

β (x1−x2)•p1)

(3.44)

Proving that equation 3.44 can be solved for all polynomialsin p1 can be done by first calculating A′ for a Wigner transform[A]W(x1+x2

2,p1) = f (x1+x2

2)∑D

j′′=1pnj′′,1, where f (x1+x2

2) is any func-

82

3.3. The open polymer classical Wigner method

tion or constant and n is a positive integer.

A′p1,n

(x1 + x2

2,x1 − x2

)

=

∫dDp1 e−

βN (

12(m−1p1)•p1− iN

β (x1−x2)•p1) f(x1+x2

2

)∑Dj′′=1 p

nj′′,1∫

dDp1 e−βN (

12(m−1p1)•p1− iN

β (x1−x2)•p1)

=f(x1+x2

2

)(

2πNβ

)D2 √

det(m) e− β

NN2

2β22 (m(x1−x2))•(x1−x2)

×(−i

)n

D∑

j′′=1

dn

d(xj′′,1 − xj′′,2)n

×∫

dDp1 e−βN (

12(m−1p1)•p1− iN

β (x1−x2)•p1)

=f(x1+x2

2

)(

2πNβ

)D2 √

det(m) e− β

NN2

2β22 (m(x1−x2))•(x1−x2)

× (i)n

D∑

j′′=1

dn

d(xj′′,1 − xj′′,2)n

×(2πN

β

)D2 √

det(m) e− β

NN2

2β22 (m(x1−x2))•(x1−x2)

=f(x1+x2

2

)(i)n

e− β

NN2

2β22 (m(x1−x2))•(x1−x2)

×

D∑

j′′=1

dn

d(xj′′,1 − xj′′,2)n

e− β

NN2

2β22 (m(x1−x2))•(x1−x2)

(3.45)

Here the variable ζj′′ =

√mj′′N

1

(xj′′,1 − xj′′,2) can be introduced,

83

Page 105: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

operators whose Wigner transforms are such functions of p1 thatthe integral

∫dDp1 e−

βN (

12(m−1p1)•p1− iN

β (x1−x2)•p1)[A]W

(x1 + x2

2,p1

)

(3.43)

is analytically solvable, giving yet another spring term. This is

not such a severe restriction since at least all[A]W(x1+x2

2,p1) that

are polynomials with regard to p1 work fine. To simplify laterexpressions, a new function is defined:

A′(x1 + x2

2,x1 − x2

)

=

∫dDp1 e−

βN (

12(m−1p1)•p1− iN

β (x1−x2)•p1)[A]W

(x1+x2

2,p1

)

∫dDp1 e−

βN (

12(m−1p1)•p1− iN

β (x1−x2)•p1)

(3.44)

Proving that equation 3.44 can be solved for all polynomialsin p1 can be done by first calculating A′ for a Wigner transform[A]W(x1+x2

2,p1) = f (x1+x2

2)∑D

j′′=1pnj′′,1, where f (x1+x2

2) is any func-

82

3.3. The open polymer classical Wigner method

tion or constant and n is a positive integer.

A′p1,n

(x1 + x2

2,x1 − x2

)

=

∫dDp1 e−

βN (

12(m−1p1)•p1− iN

β (x1−x2)•p1) f(x1+x2

2

)∑Dj′′=1 p

nj′′,1∫

dDp1 e−βN (

12(m−1p1)•p1− iN

β (x1−x2)•p1)

=f(x1+x2

2

)(

2πNβ

)D2 √

det(m) e− β

NN2

2β22 (m(x1−x2))•(x1−x2)

×(−i

)n

D∑

j′′=1

dn

d(xj′′,1 − xj′′,2)n

×∫

dDp1 e−βN (

12(m−1p1)•p1− iN

β (x1−x2)•p1)

=f(x1+x2

2

)(

2πNβ

)D2 √

det(m) e− β

NN2

2β22 (m(x1−x2))•(x1−x2)

× (i)n

D∑

j′′=1

dn

d(xj′′,1 − xj′′,2)n

×(2πN

β

)D2 √

det(m) e− β

NN2

2β22 (m(x1−x2))•(x1−x2)

=f(x1+x2

2

)(i)n

e− β

NN2

2β22 (m(x1−x2))•(x1−x2)

×

D∑

j′′=1

dn

d(xj′′,1 − xj′′,2)n

e− β

NN2

2β22 (m(x1−x2))•(x1−x2)

(3.45)

Here the variable ζj′′ =

√mj′′N

1

(xj′′,1 − xj′′,2) can be introduced,

83

Page 106: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

giving

A′p1,n

(x1 + x2

2,x1 − x2

)

=f(x1+x2

2

)(i)n

e−

∑Dj′′=1 ζ

2j′′

×

D∑

j′′=1

(mj′′N

)n2 1

ndn

dζnj′′

e−

∑Dj′′=1 ζ

2j′′

=f(x1+x2

2

)in

e−

∑Dj′′=1 ζ

2j′′

×(

D∑

j′′=1

(mj′′N

)n2

(−1)nHn (ζj′′)

)e−

∑Dj′′=1 ζ

2j′′

= f

(x1 + x2

2

)(−i)n

×D∑

j′′=1

(mj′′N

)n2

Hn

(√mj′′N

(xj′′,1 − xj′′,2)

)(3.46)

where Hn is the nth order Hermite polynomial. This result can

easily be extended to nth order polynomials in p1,[A]W(x1+x2

2,p1) =

∑nj′=0fj′ (

x1+x22

)∑Dj′′=1p

nj′′,1, as a sum,

A′polynomial

(x1 + x2

2,x1 − x2

)

=n∑

j′=0

fj′

(x1 + x2

2

)(−i)j

×D∑

j′′=1

(mj′′N

) j′2

Hj′

(√mj′′N

(xj′′,1 − xj′′,2)

). (3.47)

Using the the definition of A′ in equation 3.44 equation 3.40 can

84

3.3. The open polymer classical Wigner method

be written as

⟨AB (t)

⟩= lim

N→∞

1

Z

N∏

j=1

∫dDxj

(2π)D

×((

2πN

β

)D2 √

det (m)

)N−1 ∫dDpN

× e−βN (

12(m−1pN)•pN+V (xN+x1

2 )− iNβ (xN−x1)•pN)

×A′(x1 + x2

2,x1 − x2

)

× e− β

N

∑N−1j=1

(N2

2β22 (m(xj−xj+1))•(xj−xj+1)+V(

xj+xj+12

))

×[B(t)

]W

(xN + x1

2,pN

). (3.48)

[B(t)

]W(xN+x1

2,pN) is most often much more tricky to calculate

compared to[A]W(x1+x2

2,p1), since it includes propagation in time.

For a general potential there will not be an analytic integral over

pN .[B(t)

]W(xN+x1

2,pN) will thus be kept as it is.

Some reordering of equation 3.48 can be done, to make it a bitmore readable,

⟨AB (t)

⟩= lim

N→∞

1

Z

(N

2πβ

) (N−1)D2

(det (m))N−1

2 −ND(2π)−D

×

N∏

j=1

∫dDxj

∫dDpN e

i (xN−x1)•pN

× e− β

N

(12(m−1pN)•pN+

∑Nj=1

(V

(xj+x(j mod N)+1

2

)))

× e− β

N

∑N−1j=1

(N2

2β22 (m(xj−xj+1))•(xj−xj+1))

×A′(x1 + x2

2,x1 − x2

)

×[B(t)

]W

(xN + x1

2,pN

). (3.49)

85

Page 107: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

giving

A′p1,n

(x1 + x2

2,x1 − x2

)

=f(x1+x2

2

)(i)n

e−

∑Dj′′=1 ζ

2j′′

×

D∑

j′′=1

(mj′′N

)n2 1

ndn

dζnj′′

e−

∑Dj′′=1 ζ

2j′′

=f(x1+x2

2

)in

e−

∑Dj′′=1 ζ

2j′′

×(

D∑

j′′=1

(mj′′N

)n2

(−1)nHn (ζj′′)

)e−

∑Dj′′=1 ζ

2j′′

= f

(x1 + x2

2

)(−i)n

×D∑

j′′=1

(mj′′N

)n2

Hn

(√mj′′N

(xj′′,1 − xj′′,2)

)(3.46)

where Hn is the nth order Hermite polynomial. This result can

easily be extended to nth order polynomials in p1,[A]W(x1+x2

2,p1) =

∑nj′=0fj′ (

x1+x22

)∑Dj′′=1p

nj′′,1, as a sum,

A′polynomial

(x1 + x2

2,x1 − x2

)

=n∑

j′=0

fj′

(x1 + x2

2

)(−i)j

×D∑

j′′=1

(mj′′N

) j′2

Hj′

(√mj′′N

(xj′′,1 − xj′′,2)

). (3.47)

Using the the definition of A′ in equation 3.44 equation 3.40 can

84

3.3. The open polymer classical Wigner method

be written as

⟨AB (t)

⟩= lim

N→∞

1

Z

N∏

j=1

∫dDxj

(2π)D

×((

2πN

β

)D2 √

det (m)

)N−1 ∫dDpN

× e−βN (

12(m−1pN)•pN+V (xN+x1

2 )− iNβ (xN−x1)•pN)

×A′(x1 + x2

2,x1 − x2

)

× e− β

N

∑N−1j=1

(N2

2β22 (m(xj−xj+1))•(xj−xj+1)+V(

xj+xj+12

))

×[B(t)

]W

(xN + x1

2,pN

). (3.48)

[B(t)

]W(xN+x1

2,pN) is most often much more tricky to calculate

compared to[A]W(x1+x2

2,p1), since it includes propagation in time.

For a general potential there will not be an analytic integral over

pN .[B(t)

]W(xN+x1

2,pN) will thus be kept as it is.

Some reordering of equation 3.48 can be done, to make it a bitmore readable,

⟨AB (t)

⟩= lim

N→∞

1

Z

(N

2πβ

) (N−1)D2

(det (m))N−1

2 −ND(2π)−D

×

N∏

j=1

∫dDxj

∫dDpN e

i (xN−x1)•pN

× e− β

N

(12(m−1pN)•pN+

∑Nj=1

(V

(xj+x(j mod N)+1

2

)))

× e− β

N

∑N−1j=1

(N2

2β22 (m(xj−xj+1))•(xj−xj+1))

×A′(x1 + x2

2,x1 − x2

)

×[B(t)

]W

(xN + x1

2,pN

). (3.49)

85

Page 108: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

Figure 3.14: Illustration of a classical particle (1.) and its correspond-ing imaginary time path integral open polymer (2.). The zigzag linesrepresent the spring terms. The dashed line represent the openingin the polymer, which has a corresponding complex phase factor.

This is an exact expression for a quantum time correlation function.It is written as a discretized imaginary time path integral, as insection 2.2, but the usual ring polymer is instead an open polymer,as the last spring term is missing. Instead of a spring term, the

opening in the polymer has a complex phase factor ei (xN−x1)•pN .

In figure 3.14 the open polymer is illustrated in comparison to aclassical particle. This figure can be compared with the ring polymerin figure 2.1. Another, minor, difference compared to the imaginarytime path integral in section 2.2 is that there the potential energyis evaluated at the beads, V (xj), while here it is evaluated at themidpoint between two beads, V

(xj+x(j mod N)+1

2

). This difference does

not matter when N → ∞, and it is caused by the different handlingof the Boltzmann operators in this section compared to section 2.2.

To make equation 3.49 useful in actual computations two approxi-

86

3.3. The open polymer classical Wigner method

mations are made. The first one is to approximate[B(t)

]W

(xN + x1

2,pN

)

with the classical Wigner model,

[B(t)

]W

(xN + x1

2,pN

)

≈[B]W

(x

(xN + x1

2,pN , t

),p

(xN + x1

2,pN , t

)).

(3.50)

The second approximation is to restrict N to finite numbers. Thesetwo approximations lead to

⟨AB (t)

⟩≈

⟨AB (t)

⟩OPCW,y

=1

Z

(N

2πβ

) (N−1)D2

(det (m))N−1

2 −ND(2π)−D

×

N∏

j=1

∫dDxj

∫dDpN e

i (xN−x1)•pN

× e− β

N

(12(m−1pN)•pN+

∑Nj=1

(V

(xj+x(j mod N)+1

2

)))

× e− β

N

∑N−1j=1

(N2

2β22 (m(xj−xj+1))•(xj−xj+1))

×A′(x1 + x2

2,x1 − x2

)

×[B]W

(x

(xN + x1

2,pN , t

),p

(xN + x1

2,pN , t

)),

(3.51)

where x (xN+x12

,pN , t) and p (xN+x12

,pN , t) are the positions and mo-menta of a classical trajectory starting with xN+x1

2and pN . This is

the open polymer classical Wigner (OPCW) method published inpaper III.97 When N → ∞ the initial distribution converges towardexact quantum mechanics, and the correlation function convergestoward the exact classical Wigner method.

In papers III and IV the methods used to numerically evaluateequation 3.51 are Metropolis Monte Carlo98 for the integrals and

87

Page 109: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

Figure 3.14: Illustration of a classical particle (1.) and its correspond-ing imaginary time path integral open polymer (2.). The zigzag linesrepresent the spring terms. The dashed line represent the openingin the polymer, which has a corresponding complex phase factor.

This is an exact expression for a quantum time correlation function.It is written as a discretized imaginary time path integral, as insection 2.2, but the usual ring polymer is instead an open polymer,as the last spring term is missing. Instead of a spring term, the

opening in the polymer has a complex phase factor ei (xN−x1)•pN .

In figure 3.14 the open polymer is illustrated in comparison to aclassical particle. This figure can be compared with the ring polymerin figure 2.1. Another, minor, difference compared to the imaginarytime path integral in section 2.2 is that there the potential energyis evaluated at the beads, V (xj), while here it is evaluated at themidpoint between two beads, V

(xj+x(j mod N)+1

2

). This difference does

not matter when N → ∞, and it is caused by the different handlingof the Boltzmann operators in this section compared to section 2.2.

To make equation 3.49 useful in actual computations two approxi-

86

3.3. The open polymer classical Wigner method

mations are made. The first one is to approximate[B(t)

]W

(xN + x1

2,pN

)

with the classical Wigner model,

[B(t)

]W

(xN + x1

2,pN

)

≈[B]W

(x

(xN + x1

2,pN , t

),p

(xN + x1

2,pN , t

)).

(3.50)

The second approximation is to restrict N to finite numbers. Thesetwo approximations lead to

⟨AB (t)

⟩≈

⟨AB (t)

⟩OPCW,y

=1

Z

(N

2πβ

) (N−1)D2

(det (m))N−1

2 −ND(2π)−D

×

N∏

j=1

∫dDxj

∫dDpN e

i (xN−x1)•pN

× e− β

N

(12(m−1pN)•pN+

∑Nj=1

(V

(xj+x(j mod N)+1

2

)))

× e− β

N

∑N−1j=1

(N2

2β22 (m(xj−xj+1))•(xj−xj+1))

×A′(x1 + x2

2,x1 − x2

)

×[B]W

(x

(xN + x1

2,pN , t

),p

(xN + x1

2,pN , t

)),

(3.51)

where x (xN+x12

,pN , t) and p (xN+x12

,pN , t) are the positions and mo-menta of a classical trajectory starting with xN+x1

2and pN . This is

the open polymer classical Wigner (OPCW) method published inpaper III.97 When N → ∞ the initial distribution converges towardexact quantum mechanics, and the correlation function convergestoward the exact classical Wigner method.

In papers III and IV the methods used to numerically evaluateequation 3.51 are Metropolis Monte Carlo98 for the integrals and

87

Page 110: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

molecular dynamics, using the velocity Verlet algorithm,99,100 forthe time evolution.

As is indicated with the subscript y, the⟨AB (t)

⟩OPCW,y

in

equation 3.51 is only one version of the OPCW method. Another,slightly different, version was also published in the same paper. Thisother version arises from the fact that if all dependence on x inA is to the left of all dependence on p, then the x-parts of A canoperate on 〈x1| in equation 3.31 and avoid being enclosed in theWigner transform in equation 3.32. After that step the derivationwould move on exactly as before, but A′ would of course not dependon x1+x2

2anymore. An A′

x (x1,x1 − x2) can be defined to be theproduct of the function of x1 and the A′ (x1 − x2) resulting from

the p-dependent parts of A. Then, this version of OPCW can bewritten as

⟨AB (t)

⟩≈

⟨AB (t)

⟩OPCW,x

=1

Z

(N

2πβ

) (N−1)D2

(det (m))N−1

2 −ND(2π)−D

×

N∏

j=1

∫dDxj

∫dDpN e

i (xN−x1)•pN

× e− β

N

(12(m−1pN)•pN+

∑Nj=1

(V

(xj+x(j mod N)+1

2

)))

× e− β

N

∑N−1j=1

(N2

2β22 (m(xj−xj+1))•(xj−xj+1))

×A′x (x1,x1 − x2)

×[B]W

(x

(xN + x1

2,pN , t

),p

(xN + x1

2,pN , t

)),

(3.52)

where it can now be understood that the subscript x in⟨AB (t)

⟩OPCW,x

denotes that the function corresponding to operator A is evaluated

at x1. The y in⟨AB (t)

⟩OPCW,y

denotes that the function corre-

sponding to operator A is evaluated at y1, which is the symbol usedfor x1+x2

2in paper III.

88

3.3. The open polymer classical Wigner method

For some simple cases it is easy to see what A′x (x1,x1 − x2) will

be.

• If A only depends on p and not on x, then A′x (x1,x1 − x2) =

A′ (x1+x22

,x1 − x2).

• If A is only a function of x, A = f (x), thenA′x (x1,x1 − x2) =

f (x1) and A′ (x1+x22

,x1 − x2) = f (x1+x22

).

• If A is the product of a function of x, f (x), and a function

of p, g (p), such that A = f (x) g (p) and A′g is the A′ when

A = g (p), then A′x (x1,x1 − x2) = f (x1)A

′g (

x1+x22

,x1 − x2)and A′ (x1+x2

2,x1 − x2) = f (x1+x2

2)A′

g (x1+x2

2,x1 − x2).

If A has p-dependence to the left of an x-dependence, then a rear-rangement of the expression for A may possible, such as for

px = xp+ [p, x] = xp+ [p, x] = xp− i. (3.53)

These kind of operators will lead to a more complicated derivation of⟨AB (t)

⟩OPCW,x

and a general formula for them will not be given

here. In paper IV the flux-Heaviside trace, equation 2.61, fromsection 2.3 is handled, and the flux operator, equation 2.63, needsthe kind of treatment illustrated. The interested reader is directedto the paper for an example of how a complicated A can be handled.

In the limit N → ∞ the y-version and the x-version of OPCWwill be equivalent, but in practical numerical calculations they canbehave differently. The OPCW method is interesting for runningthe classical Wigner method, because the main difficulty of anyclassical Wigner method is to obtain the initial quantum distribution.Sampling a discretized imaginary path integral should be relativelycheap, computationally, but the complex phase factor makes theintegrand oscillatory, which makes the sampling more demanding.For multidimensional systems the complex phase factor becomeseven worse, since each degree of freedom contributes to it. In paperIII it was suggested to describe only the most important, or mostquantum mechanical, degrees of freedom with the open polymer. Forthe less important, or less quantum mechanical, degrees of freedom

89

Page 111: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

molecular dynamics, using the velocity Verlet algorithm,99,100 forthe time evolution.

As is indicated with the subscript y, the⟨AB (t)

⟩OPCW,y

in

equation 3.51 is only one version of the OPCW method. Another,slightly different, version was also published in the same paper. Thisother version arises from the fact that if all dependence on x inA is to the left of all dependence on p, then the x-parts of A canoperate on 〈x1| in equation 3.31 and avoid being enclosed in theWigner transform in equation 3.32. After that step the derivationwould move on exactly as before, but A′ would of course not dependon x1+x2

2anymore. An A′

x (x1,x1 − x2) can be defined to be theproduct of the function of x1 and the A′ (x1 − x2) resulting from

the p-dependent parts of A. Then, this version of OPCW can bewritten as

⟨AB (t)

⟩≈

⟨AB (t)

⟩OPCW,x

=1

Z

(N

2πβ

) (N−1)D2

(det (m))N−1

2 −ND(2π)−D

×

N∏

j=1

∫dDxj

∫dDpN e

i (xN−x1)•pN

× e− β

N

(12(m−1pN)•pN+

∑Nj=1

(V

(xj+x(j mod N)+1

2

)))

× e− β

N

∑N−1j=1

(N2

2β22 (m(xj−xj+1))•(xj−xj+1))

×A′x (x1,x1 − x2)

×[B]W

(x

(xN + x1

2,pN , t

),p

(xN + x1

2,pN , t

)),

(3.52)

where it can now be understood that the subscript x in⟨AB (t)

⟩OPCW,x

denotes that the function corresponding to operator A is evaluated

at x1. The y in⟨AB (t)

⟩OPCW,y

denotes that the function corre-

sponding to operator A is evaluated at y1, which is the symbol usedfor x1+x2

2in paper III.

88

3.3. The open polymer classical Wigner method

For some simple cases it is easy to see what A′x (x1,x1 − x2) will

be.

• If A only depends on p and not on x, then A′x (x1,x1 − x2) =

A′ (x1+x22

,x1 − x2).

• If A is only a function of x, A = f (x), thenA′x (x1,x1 − x2) =

f (x1) and A′ (x1+x22

,x1 − x2) = f (x1+x22

).

• If A is the product of a function of x, f (x), and a function

of p, g (p), such that A = f (x) g (p) and A′g is the A′ when

A = g (p), then A′x (x1,x1 − x2) = f (x1)A

′g (

x1+x22

,x1 − x2)and A′ (x1+x2

2,x1 − x2) = f (x1+x2

2)A′

g (x1+x2

2,x1 − x2).

If A has p-dependence to the left of an x-dependence, then a rear-rangement of the expression for A may possible, such as for

px = xp+ [p, x] = xp+ [p, x] = xp− i. (3.53)

These kind of operators will lead to a more complicated derivation of⟨AB (t)

⟩OPCW,x

and a general formula for them will not be given

here. In paper IV the flux-Heaviside trace, equation 2.61, fromsection 2.3 is handled, and the flux operator, equation 2.63, needsthe kind of treatment illustrated. The interested reader is directedto the paper for an example of how a complicated A can be handled.

In the limit N → ∞ the y-version and the x-version of OPCWwill be equivalent, but in practical numerical calculations they canbehave differently. The OPCW method is interesting for runningthe classical Wigner method, because the main difficulty of anyclassical Wigner method is to obtain the initial quantum distribution.Sampling a discretized imaginary path integral should be relativelycheap, computationally, but the complex phase factor makes theintegrand oscillatory, which makes the sampling more demanding.For multidimensional systems the complex phase factor becomeseven worse, since each degree of freedom contributes to it. In paperIII it was suggested to describe only the most important, or mostquantum mechanical, degrees of freedom with the open polymer. Forthe less important, or less quantum mechanical, degrees of freedom

89

Page 112: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

the polymer would be collapsed into a single classical particle, butkeeping the couplings to all beads in degrees of freedom that still havea quantum description. This could reduce the computational costsignificantly, but at the cost of correlation functions not convergingtoward exact classical Wigner as the number of beads is increased.

Here, the OPCW method is only derived for the standard corre-lation function. In appendix B of paper III a version of the OPCWmethod for Kubo transformed correlation functions is presented.Paper IV contains derivations for symmetrized correlation functions,specifically the flux-Heaviside trace, equation 2.61.

It should be noted that the OPCW method is related to an ex-pression published, but not further explored, by Bose and Makri.101‡

OPCW is also related to a version of a method published by Bonellaet al.102§.

Results and discussion

In paper III OPCW was tested on the same two one-dimensional po-tentials as was used in paper II, see section 3.2, i.e. a quartic potential,V (x) = m2ω3

4 x4, and a double well potential, V (x) = m2ω3

10 x4 − mω2

2x2.

Additionally, OPCW was tested on the same quartic potential, butbilinearly coupled to a bath of harmonic oscillators.48,104 Thesetests were done with position, position-squared, momentum, andmomentum-squared autocorrelation functions.

In paper IV OPCW was tested for a parabolic barrier, V (x) =−mω2

2x2, and an Eckart potential, V (x) = 6ω

πsech2 (

√πmω12 x). These

tests where done with the symmetrized flux-Heaviside trace, equation2.61, and resulted in rate constants and tunneling factors for theEckart potential.

Some illustrative results from both papers III and IV are repro-duced here to illustrate the conclusions from the papers.

Common methods for approximating the initial distribution forthe classical Wigner model use a harmonic approximation.22,23,105

For comparison a representative from this group of methods is used

‡Equation 2.7 in reference 101.§The L = 1 version of the method published in reference 102. The relation

is more clear in a later paper by Bonella and Ciccotti, reference 103.

90

3.3. The open polymer classical Wigner method

−0.6

−0.4

−0.2

0.0

0.2

0.4

0.6

0 1 2 3 4 5 6 7 8 9 10

Re ⟨

x∧ x∧

(t)

⟩ / (

− hm

−1ω

−1)

t / ω−1

CMOPCW

CWFK−LPIRPMD

QM

Figure 3.15: The real part of the position autocorrelation functionfor a quartic oscillator calculated at βω = 8. Results for classicalmechanics (CM), OPCW, exact classical Wigner (CW), FK-LPI,RPMD, and exact quantum mechanics (QM).

as a comparison in the graphs. This method is the Feynman-Kleinertlinearized path integral (FK-LPI) method.22 It uses the Feynman-Kleinert approximation,96 which is a local harmonic approximation,for the initial quantum distribution.

In both papers III and IV it was found that, as the number ofbeads in the polymer is increased, the correlation functions (or trace)calculated with OPCW converge toward exact quantum mechanicsat time t = 0 and exact classical Wigner correlation functions for alltimes. This is as expected from the derivations.

In figures 3.15 and 3.16 Re 〈xx (t)〉 for the quartic oscillator atβω = 8 and βω = 1 can be seen, respectively. For both thesecases OPCW, as well as FK-LPI, has been converged to very nearthe exact classical Wigner correlation function. In this case the y-and x-versions of OPCW give very similar correlation functions, andonly the y-version is shown in the graph. These OPCW calculationsuse 160 beads at βω = 8 and 80 beads at βω = 1. It was found

91

Page 113: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

the polymer would be collapsed into a single classical particle, butkeeping the couplings to all beads in degrees of freedom that still havea quantum description. This could reduce the computational costsignificantly, but at the cost of correlation functions not convergingtoward exact classical Wigner as the number of beads is increased.

Here, the OPCW method is only derived for the standard corre-lation function. In appendix B of paper III a version of the OPCWmethod for Kubo transformed correlation functions is presented.Paper IV contains derivations for symmetrized correlation functions,specifically the flux-Heaviside trace, equation 2.61.

It should be noted that the OPCW method is related to an ex-pression published, but not further explored, by Bose and Makri.101‡

OPCW is also related to a version of a method published by Bonellaet al.102§.

Results and discussion

In paper III OPCW was tested on the same two one-dimensional po-tentials as was used in paper II, see section 3.2, i.e. a quartic potential,V (x) = m2ω3

4 x4, and a double well potential, V (x) = m2ω3

10 x4 − mω2

2x2.

Additionally, OPCW was tested on the same quartic potential, butbilinearly coupled to a bath of harmonic oscillators.48,104 Thesetests were done with position, position-squared, momentum, andmomentum-squared autocorrelation functions.

In paper IV OPCW was tested for a parabolic barrier, V (x) =−mω2

2x2, and an Eckart potential, V (x) = 6ω

πsech2 (

√πmω12 x). These

tests where done with the symmetrized flux-Heaviside trace, equation2.61, and resulted in rate constants and tunneling factors for theEckart potential.

Some illustrative results from both papers III and IV are repro-duced here to illustrate the conclusions from the papers.

Common methods for approximating the initial distribution forthe classical Wigner model use a harmonic approximation.22,23,105

For comparison a representative from this group of methods is used

‡Equation 2.7 in reference 101.§The L = 1 version of the method published in reference 102. The relation

is more clear in a later paper by Bonella and Ciccotti, reference 103.

90

3.3. The open polymer classical Wigner method

−0.6

−0.4

−0.2

0.0

0.2

0.4

0.6

0 1 2 3 4 5 6 7 8 9 10

Re ⟨

x∧ x∧

(t)

⟩ / (

− hm

−1ω

−1)

t / ω−1

CMOPCW

CWFK−LPIRPMD

QM

Figure 3.15: The real part of the position autocorrelation functionfor a quartic oscillator calculated at βω = 8. Results for classicalmechanics (CM), OPCW, exact classical Wigner (CW), FK-LPI,RPMD, and exact quantum mechanics (QM).

as a comparison in the graphs. This method is the Feynman-Kleinertlinearized path integral (FK-LPI) method.22 It uses the Feynman-Kleinert approximation,96 which is a local harmonic approximation,for the initial quantum distribution.

In both papers III and IV it was found that, as the number ofbeads in the polymer is increased, the correlation functions (or trace)calculated with OPCW converge toward exact quantum mechanicsat time t = 0 and exact classical Wigner correlation functions for alltimes. This is as expected from the derivations.

In figures 3.15 and 3.16 Re 〈xx (t)〉 for the quartic oscillator atβω = 8 and βω = 1 can be seen, respectively. For both thesecases OPCW, as well as FK-LPI, has been converged to very nearthe exact classical Wigner correlation function. In this case the y-and x-versions of OPCW give very similar correlation functions, andonly the y-version is shown in the graph. These OPCW calculationsuse 160 beads at βω = 8 and 80 beads at βω = 1. It was found

91

Page 114: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

−0.6

−0.4

−0.2

0.0

0.2

0.4

0.6

0.8

0 1 2 3 4 5 6 7 8 9 10

Re ⟨

x∧ x∧

(t)

⟩ / (

− hm

−1ω

−1)

t / ω−1

CMOPCW

CWFK−LPIRPMD

QM

Figure 3.16: The real part of the position autocorrelation functionfor a quartic oscillator calculated at βω = 1. Results for classicalmechanics (CM), OPCW, exact classical Wigner (CW), FK-LPI,RPMD, and exact quantum mechanics (QM).

in paper III that fewer beads are needed for convergence, of OPCW,at higher temperatures but that the Monte Carlo integration alsobecomes more demanding per bead, for the higher temperature.

In figures 3.17 and 3.18 Re 〈xx (t)〉 and Re⟨x2x2 (t)

⟩for the

double well potential at βω = 8 can be seen. This potential isinteresting to test since it contains some negative curvature. Asbefore only the y-version of OPCW is shown, in this case using 160beads. For Re 〈xx (t)〉 both OPCW and FK-LPI have more or lessconverged to the exact classical Wigner result. For Re

⟨x2x2 (t)

⟩,

however, OPCW has converged close to the exact classical Wignerresult, but FK-LPI has a significantly different behavior. Thisillustrates a possible benefit of OPCW compared to FK-LPI. OPCWworks equally well regardless of the potential energy surface it isapplied to, even if some will be more numerically demanding thanothers. Methods, such as FK-LPI, that use approximations for thepotential energy surface will work differently well, depending on how

92

3.3. The open polymer classical Wigner method

−1.5

−1.0

−0.5

0.0

0.5

1.0

1.5

2.0

2.5

0 1 2 3 4 5 6 7 8 9 10

Re ⟨

x∧ x∧

(t)

⟩ / (

− hm

−1ω

−1)

t / ω−1

CMOPCW

CWFK−LPIRPMD

QM

Figure 3.17: The real part of the position autocorrelation function fora double well potential calculated at βω = 8. Results for classicalmechanics (CM), OPCW, exact classical Wigner (CW), FK-LPI,RPMD, and exact quantum mechanics (QM).

kr/(12ω

12m− 1

2 )βω CM OPCW CW RPMD QM1 5.9084E-2 6.10E-2 6.0986E-2 6.17E-2 6.2856E-23 7.4821E-4 1.009E-3 1.0087E-3 1.07E-3 1.1409E-36 1.7186E-6 7.391E-6 7.5092E-6 7.55E-6 8.9347E-6

Table 3.1: Rate constant for the Eckart potential at different inversetemperatures. Values for OPCW, RPMD, classical mechanics (CM),exact classical Wigner (CW), and exact quantum mechanics (QM).

well the approximation describes the surface. It is however worthnoting that in figure 3.18 both OPCW and FK-LPI follow exactclassical Wigner well for as long as the classical Wigner results followexact quantum mechanics. It would thus be misleading to claim thatOPCW describes this double well better than FK-LPI does.

Another potential with a region of negative curvature is the

93

Page 115: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

−0.6

−0.4

−0.2

0.0

0.2

0.4

0.6

0.8

0 1 2 3 4 5 6 7 8 9 10

Re ⟨

x∧ x∧

(t)

⟩ / (

− hm

−1ω

−1)

t / ω−1

CMOPCW

CWFK−LPIRPMD

QM

Figure 3.16: The real part of the position autocorrelation functionfor a quartic oscillator calculated at βω = 1. Results for classicalmechanics (CM), OPCW, exact classical Wigner (CW), FK-LPI,RPMD, and exact quantum mechanics (QM).

in paper III that fewer beads are needed for convergence, of OPCW,at higher temperatures but that the Monte Carlo integration alsobecomes more demanding per bead, for the higher temperature.

In figures 3.17 and 3.18 Re 〈xx (t)〉 and Re⟨x2x2 (t)

⟩for the

double well potential at βω = 8 can be seen. This potential isinteresting to test since it contains some negative curvature. Asbefore only the y-version of OPCW is shown, in this case using 160beads. For Re 〈xx (t)〉 both OPCW and FK-LPI have more or lessconverged to the exact classical Wigner result. For Re

⟨x2x2 (t)

⟩,

however, OPCW has converged close to the exact classical Wignerresult, but FK-LPI has a significantly different behavior. Thisillustrates a possible benefit of OPCW compared to FK-LPI. OPCWworks equally well regardless of the potential energy surface it isapplied to, even if some will be more numerically demanding thanothers. Methods, such as FK-LPI, that use approximations for thepotential energy surface will work differently well, depending on how

92

3.3. The open polymer classical Wigner method

−1.5

−1.0

−0.5

0.0

0.5

1.0

1.5

2.0

2.5

0 1 2 3 4 5 6 7 8 9 10

Re ⟨

x∧ x∧

(t)

⟩ / (

− hm

−1ω

−1)

t / ω−1

CMOPCW

CWFK−LPIRPMD

QM

Figure 3.17: The real part of the position autocorrelation function fora double well potential calculated at βω = 8. Results for classicalmechanics (CM), OPCW, exact classical Wigner (CW), FK-LPI,RPMD, and exact quantum mechanics (QM).

kr/(12ω

12m− 1

2 )βω CM OPCW CW RPMD QM1 5.9084E-2 6.10E-2 6.0986E-2 6.17E-2 6.2856E-23 7.4821E-4 1.009E-3 1.0087E-3 1.07E-3 1.1409E-36 1.7186E-6 7.391E-6 7.5092E-6 7.55E-6 8.9347E-6

Table 3.1: Rate constant for the Eckart potential at different inversetemperatures. Values for OPCW, RPMD, classical mechanics (CM),exact classical Wigner (CW), and exact quantum mechanics (QM).

well the approximation describes the surface. It is however worthnoting that in figure 3.18 both OPCW and FK-LPI follow exactclassical Wigner well for as long as the classical Wigner results followexact quantum mechanics. It would thus be misleading to claim thatOPCW describes this double well better than FK-LPI does.

Another potential with a region of negative curvature is the

93

Page 116: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

0.0

1.0

2.0

3.0

4.0

5.0

6.0

7.0

0 1 2 3 4 5 6 7 8 9 10

Re ⟨

x∧2 x∧

2 (

t) ⟩

/ (

− h2m

−2ω

−2)

t / ω−1

CMOPCW

CWFK−LPIRPMD

QM

Figure 3.18: The real part of the position-squared autocorrelationfunction for a double well potential calculated at βω = 8. Resultsfor classical mechanics (CM), OPCW, exact classical Wigner (CW),FK-LPI, RPMD, and exact quantum mechanics (QM). The increasedthickness of the OPCW line shows the standard deviation.

Eckart potential, studied in paper IV. In table 3.1 rate constants,taken from the long time values of the flux-Heaviside trace, equation2.61, for the Eckart barrier at different temperatures can be found.In this case results from the x-version of OPCW, with 40 beads,are shown. It can be seen that for these rate constants the OPCWresults are close to the exact classical Wigner value. For βω = 1,as little as 6 beads can be enough to get an accurate classicalWigner result. The classical Wigner rate constants, as could beexpected, place themselves between the classical and the quantummechanical ones. At βω = 1 the difference between the classicaland quantum rate constants is only ∼ 6 %, so there is not muchbenefit from using the classical Wigner method, which accounts forhalf of that difference. At βω = 3 and βω = 6 the differencesbetween classical and quantum rate constants are∼ 50 % and a factor∼ 5, respectively. At these temperatures the classical Wigner rate

94

3.3. The open polymer classical Wigner method

0.0

0.1

0.2

0.3

0.4

0.5

0 1 2 3 4 5 6 7 8 9 10

Re ⟨

x∧2 x∧

2 (

t) ⟩

/ (

− h2m

−2ω

−2)

t / ω−1

CMN=5

N=10N=20N=40

QM

Figure 3.19: The real part of the position-squared autocorrelationfunction for a quartic potential bilinearly coupled to a bath of 3harmonic oscillators, calculated at βω = 8. Results for classicalmechanics (CM) and the y-version of OPCW, with different numbersof beads (N). Exact quantum mechanics (QM) is shown at timet = 0. The top and bottom lines of each type show the standarddeviation.

constants are significant improvements over pure classical mechanics.The same thing can be seen for the position autocorrelation functionfor the quartic oscillator in figures 3.15 and 3.16, for the highertemperature the classical Wigner method is hardly an improvementover classical mechanics, but for the lower temperature it is a clearimprovement.

In figures 3.19 and 3.20 Re⟨x2x2 (t)

⟩for the quartic potential

coupled to a bath of 3 harmonic oscillators can be seen. These figuresillustrate the difference in convergence of Re

⟨x2x2 (t)

⟩OPCW,y

and

Re⟨x2x2 (t)

⟩OPCW,x

, as the number of beads used is increased. For

this multidimensional system no exact classical Wigner correlationfunction was readily calculable and the exact quantum mechanicalresult is only the time 0 value, not the entire correlation function. At

95

Page 117: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

0.0

1.0

2.0

3.0

4.0

5.0

6.0

7.0

0 1 2 3 4 5 6 7 8 9 10

Re ⟨

x∧2 x∧

2 (

t) ⟩

/ (

− h2m

−2ω

−2)

t / ω−1

CMOPCW

CWFK−LPIRPMD

QM

Figure 3.18: The real part of the position-squared autocorrelationfunction for a double well potential calculated at βω = 8. Resultsfor classical mechanics (CM), OPCW, exact classical Wigner (CW),FK-LPI, RPMD, and exact quantum mechanics (QM). The increasedthickness of the OPCW line shows the standard deviation.

Eckart potential, studied in paper IV. In table 3.1 rate constants,taken from the long time values of the flux-Heaviside trace, equation2.61, for the Eckart barrier at different temperatures can be found.In this case results from the x-version of OPCW, with 40 beads,are shown. It can be seen that for these rate constants the OPCWresults are close to the exact classical Wigner value. For βω = 1,as little as 6 beads can be enough to get an accurate classicalWigner result. The classical Wigner rate constants, as could beexpected, place themselves between the classical and the quantummechanical ones. At βω = 1 the difference between the classicaland quantum rate constants is only ∼ 6 %, so there is not muchbenefit from using the classical Wigner method, which accounts forhalf of that difference. At βω = 3 and βω = 6 the differencesbetween classical and quantum rate constants are∼ 50 % and a factor∼ 5, respectively. At these temperatures the classical Wigner rate

94

3.3. The open polymer classical Wigner method

0.0

0.1

0.2

0.3

0.4

0.5

0 1 2 3 4 5 6 7 8 9 10

Re ⟨

x∧2 x∧

2 (

t) ⟩

/ (

− h2m

−2ω

−2)

t / ω−1

CMN=5

N=10N=20N=40

QM

Figure 3.19: The real part of the position-squared autocorrelationfunction for a quartic potential bilinearly coupled to a bath of 3harmonic oscillators, calculated at βω = 8. Results for classicalmechanics (CM) and the y-version of OPCW, with different numbersof beads (N). Exact quantum mechanics (QM) is shown at timet = 0. The top and bottom lines of each type show the standarddeviation.

constants are significant improvements over pure classical mechanics.The same thing can be seen for the position autocorrelation functionfor the quartic oscillator in figures 3.15 and 3.16, for the highertemperature the classical Wigner method is hardly an improvementover classical mechanics, but for the lower temperature it is a clearimprovement.

In figures 3.19 and 3.20 Re⟨x2x2 (t)

⟩for the quartic potential

coupled to a bath of 3 harmonic oscillators can be seen. These figuresillustrate the difference in convergence of Re

⟨x2x2 (t)

⟩OPCW,y

and

Re⟨x2x2 (t)

⟩OPCW,x

, as the number of beads used is increased. For

this multidimensional system no exact classical Wigner correlationfunction was readily calculable and the exact quantum mechanicalresult is only the time 0 value, not the entire correlation function. At

95

Page 118: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

0.0

0.1

0.2

0.3

0.4

0.5

0 1 2 3 4 5 6 7 8 9 10

Re ⟨

x∧2 x∧

2 (

t) ⟩

/ (

− h2m

−2ω

−2)

t / ω−1

CMN=5

N=10N=20N=40

QM

Figure 3.20: The real part of the position-squared autocorrelationfunction for a quartic potential bilinearly coupled to a bath of 3harmonic oscillators, calculated at βω = 8. Results for classicalmechanics (CM) and the x-version of OPCW, with different numbersof beads (N). Exact quantum mechanics (QM) is shown at timet = 0. The top and bottom lines of each type show the standarddeviation.

time zero the classical Wigner method should however yield the exactquantum mechanical result, so this is still a useful comparison forOPCW. As can be seen, Re

⟨x2x2 (t)

⟩OPCW,x

converges significantly

faster toward the exact quantum result, when the number of beadsis increased, than Re

⟨x2x2 (t)

⟩OPCW,y

does. That the x-version of

OPCW converges faster toward exact classical Wigner than they-version was also observed for all other correlation functions, ortraces, where comparisons were made in papers III and IV. For thecalculations using the same number of beads in figures 3.19 and3.20 the same number of Monte Carlo steps were used. Thus, thestandard deviations illustrate another point from paper III. That isthat the x-version of OPCW tend to converge better with respectto the number of Monte Carlo steps used than the y-version does.

96

3.3. The open polymer classical Wigner method

−0.1

0.0

0.1

0.2

0.3

0.4

0.5

0 1 2 3 4 5 6 7 8 9 10

Re ⟨

x∧2 x∧

2 (

t) ⟩

/ (

− h2m

−2ω

−2)

t / ω−1

CMOPCW,yOPCW,x

OPCW,CBFK−LPIRPMD

QM

Figure 3.21: The real part of the position-squared autocorrelationfunction for a quartic potential bilinearly coupled to a bath of 9harmonic oscillators, calculated at βω = 8. Results for classicalmechanics (CM), OPCW, FK-LPI, and RPMD. The exact quantummechanical result (QM) is shown at time t = 0. The OPCW methodis shown for the y-version, the x-version, and the x-version with aclassical bath (CB). The top and bottom lines of each type show thestandard deviation.

Both these observations of course make the x-version seem morepromising as a useful computational method than the y-version, butthere is no guarantee that these convergence behaviors will be thesame for every system and correlation function.

In figure 3.21 Re⟨x2x2 (t)

⟩for the quartic oscillator coupled

to a bath of 9 harmonic oscillators can be seen. Due to the de-manding computations, only results using 5 beads are shown forRe

⟨x2x2 (t)

⟩OPCW,y

and Re⟨x2x2 (t)

⟩OPCW,x

. It can, as for figures

3.19 and 3.20 be seen that the x-version of OPCW is closer to theexact result than the y-version is. Both are however far off. Figure3.21 also shows how much of an improvement it can be to makethe harmonic bath classical and just let the quartic oscillator be

97

Page 119: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

0.0

0.1

0.2

0.3

0.4

0.5

0 1 2 3 4 5 6 7 8 9 10

Re ⟨

x∧2 x∧

2 (

t) ⟩

/ (

− h2m

−2ω

−2)

t / ω−1

CMN=5

N=10N=20N=40

QM

Figure 3.20: The real part of the position-squared autocorrelationfunction for a quartic potential bilinearly coupled to a bath of 3harmonic oscillators, calculated at βω = 8. Results for classicalmechanics (CM) and the x-version of OPCW, with different numbersof beads (N). Exact quantum mechanics (QM) is shown at timet = 0. The top and bottom lines of each type show the standarddeviation.

time zero the classical Wigner method should however yield the exactquantum mechanical result, so this is still a useful comparison forOPCW. As can be seen, Re

⟨x2x2 (t)

⟩OPCW,x

converges significantly

faster toward the exact quantum result, when the number of beadsis increased, than Re

⟨x2x2 (t)

⟩OPCW,y

does. That the x-version of

OPCW converges faster toward exact classical Wigner than they-version was also observed for all other correlation functions, ortraces, where comparisons were made in papers III and IV. For thecalculations using the same number of beads in figures 3.19 and3.20 the same number of Monte Carlo steps were used. Thus, thestandard deviations illustrate another point from paper III. That isthat the x-version of OPCW tend to converge better with respectto the number of Monte Carlo steps used than the y-version does.

96

3.3. The open polymer classical Wigner method

−0.1

0.0

0.1

0.2

0.3

0.4

0.5

0 1 2 3 4 5 6 7 8 9 10

Re ⟨

x∧2 x∧

2 (

t) ⟩

/ (

− h2m

−2ω

−2)

t / ω−1

CMOPCW,yOPCW,x

OPCW,CBFK−LPIRPMD

QM

Figure 3.21: The real part of the position-squared autocorrelationfunction for a quartic potential bilinearly coupled to a bath of 9harmonic oscillators, calculated at βω = 8. Results for classicalmechanics (CM), OPCW, FK-LPI, and RPMD. The exact quantummechanical result (QM) is shown at time t = 0. The OPCW methodis shown for the y-version, the x-version, and the x-version with aclassical bath (CB). The top and bottom lines of each type show thestandard deviation.

Both these observations of course make the x-version seem morepromising as a useful computational method than the y-version, butthere is no guarantee that these convergence behaviors will be thesame for every system and correlation function.

In figure 3.21 Re⟨x2x2 (t)

⟩for the quartic oscillator coupled

to a bath of 9 harmonic oscillators can be seen. Due to the de-manding computations, only results using 5 beads are shown forRe

⟨x2x2 (t)

⟩OPCW,y

and Re⟨x2x2 (t)

⟩OPCW,x

. It can, as for figures

3.19 and 3.20 be seen that the x-version of OPCW is closer to theexact result than the y-version is. Both are however far off. Figure3.21 also shows how much of an improvement it can be to makethe harmonic bath classical and just let the quartic oscillator be

97

Page 120: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

sampled by an open polymer. In this classical bath calculation itwas feasible to have 320 beads in the polymer in the quartic poten-tial. The OPCW, with a classical bath, correlation function followsFK-LPI quite well for the first 3 ω−1. After that the classical bathresult continues toward a different long time value. If assumingthat Re

⟨x2x2 (t)

⟩OPCW,x

is closer to the exact classical Wigner re-

sult than Re⟨x2x2 (t)

⟩OPCW,y

is, then the long time value of the

classical bath calculation should not be close to the exact classicalWigner value. Since the long time value of Re

⟨x2x2 (t)

⟩OPCW,x

is

in the neighborhood of well converged FK-LPI results, this is a fairassumption.

When a classical Wigner calcluation is run, all degrees of freedomstart off with some zero point energy. Since the evolution in time isclassical, this zero point energy leaks between the degrees of freedom,degrading the results over time. This phenomenon was mentioned insection 2.1. If only one degree of freedom has zero point energy andthe others do not, then the leakage of zero point energy from thatsingle degree of freedom should be even bigger and it would only goone way, making the effect even larger. This is a reason why the longtime values of classical bath OPCW correlation functions should notnecessarily be trusted to converge toward the exact classical Wignerresult. Another observation that can be made in figure 3.21 is thatthe OPCW, with classical bath, value at time zero is a bit below theexact quantum mechanical result. There is no guarantee that anynumber of beads in the quartic degree of freedom would make theOPCW value come any closer, due to the classical bath. In paperIII it it can be seen that the classical bath approximation of OPCWseems to give a good approximation to exact classical Wigner forshort times, for the quartic potential coupled to a harmonic bath,and that it is computationally much cheaper than standard OPCW.

In figures 3.15, 3.16, 3.17, 3.18, and 3.21, and table 3.1 theclassical Wigner results can be compared with RPMD. For most ofthese cases, and most of the cases in papers III and IV, RPMD givesresults that are better than, or equally good as, the classical Wignerresults. Notable exceptions to this are the correlation functionsfrom the double well potential, figures 3.17 and 3.18. For shorttimes the classical bath OPCW method also gives better results

98

3.3. The open polymer classical Wigner method

than RPMD for the position-squared autocorrelation function, forthe quartic potential with a harmonic bath. Since the classicalWigner method, and OPCW, give better results than RPMD forthe double well potential, and this is also the system where OPCWshows good potential compared to FK-LPI, the double well potential,and possibly other potentials with regions of negative curvature, areinteresting for further studies.

99

Page 121: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

sampled by an open polymer. In this classical bath calculation itwas feasible to have 320 beads in the polymer in the quartic poten-tial. The OPCW, with a classical bath, correlation function followsFK-LPI quite well for the first 3 ω−1. After that the classical bathresult continues toward a different long time value. If assumingthat Re

⟨x2x2 (t)

⟩OPCW,x

is closer to the exact classical Wigner re-

sult than Re⟨x2x2 (t)

⟩OPCW,y

is, then the long time value of the

classical bath calculation should not be close to the exact classicalWigner value. Since the long time value of Re

⟨x2x2 (t)

⟩OPCW,x

is

in the neighborhood of well converged FK-LPI results, this is a fairassumption.

When a classical Wigner calcluation is run, all degrees of freedomstart off with some zero point energy. Since the evolution in time isclassical, this zero point energy leaks between the degrees of freedom,degrading the results over time. This phenomenon was mentioned insection 2.1. If only one degree of freedom has zero point energy andthe others do not, then the leakage of zero point energy from thatsingle degree of freedom should be even bigger and it would only goone way, making the effect even larger. This is a reason why the longtime values of classical bath OPCW correlation functions should notnecessarily be trusted to converge toward the exact classical Wignerresult. Another observation that can be made in figure 3.21 is thatthe OPCW, with classical bath, value at time zero is a bit below theexact quantum mechanical result. There is no guarantee that anynumber of beads in the quartic degree of freedom would make theOPCW value come any closer, due to the classical bath. In paperIII it it can be seen that the classical bath approximation of OPCWseems to give a good approximation to exact classical Wigner forshort times, for the quartic potential coupled to a harmonic bath,and that it is computationally much cheaper than standard OPCW.

In figures 3.15, 3.16, 3.17, 3.18, and 3.21, and table 3.1 theclassical Wigner results can be compared with RPMD. For most ofthese cases, and most of the cases in papers III and IV, RPMD givesresults that are better than, or equally good as, the classical Wignerresults. Notable exceptions to this are the correlation functionsfrom the double well potential, figures 3.17 and 3.18. For shorttimes the classical bath OPCW method also gives better results

98

3.3. The open polymer classical Wigner method

than RPMD for the position-squared autocorrelation function, forthe quartic potential with a harmonic bath. Since the classicalWigner method, and OPCW, give better results than RPMD forthe double well potential, and this is also the system where OPCWshows good potential compared to FK-LPI, the double well potential,and possibly other potentials with regions of negative curvature, areinteresting for further studies.

99

Page 122: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

3.4 Future outlook

The research presented in this thesis contains a rate constant for theformation of the hydroxyl radical through radiative association, twonew types of basis functions for use with a variational principle forthe time evolution of Wigner functions, and a method for samplingthe initial distributions in the classical Wigner method. After thesummation of all this research, the question that remain is: Whereto go from here?

The new reaction rate constant was published in 2015, but hasstill not¶ been included in any astrochemical database and noneof its citations indicates that it has been used for astrochemicalmodeling.

The database KIDA86 has instructions for suggesting new ad-ditions to their database. It is an oversight by the author that nosuch suggestion has been sent. To make this rate constant actuallybenefit the astrochemical community, the first thing that has to bedone is to actually suggest that it is included in a database.

The calculation of reaction rate constants for other radiativeassociation reactions and the development of the methods to conductthese calculations has continued in the research group, and amongcollaborators, of the author since the publication of paper I.106–114

One of the main directions that further research is going toward is thecalculation of radiative association rate constants for systems largerthan diatoms. As radiative association generally becomes more likelythe larger the molecule, due to a more long lived activated complex,as explained e.g. in the textbook by Tielens,115 this seems to be theright way to go.

For the Gaussian basis functions for use with a variational princi-ple for the time propagation of Wigner functions, presented in paperII, more testing would be needed to know in what situations theycould be a beneficial route to go for calculations. Further testingcould also help to find ways of improving the current methods. Mul-tidimensional potentials and other correlation functions than theposition autocorrelation function, would be the obvious first stepsin such testing. As was suggested in the paper, a way of improving

¶When writing this, 30 April 2020.

100

3.4. Future outlook

the dynamics of the frozen Gaussian basis functions could be to givethem a semiclassical prefactor that approximately would account forcoupling between the separate basis functions.

When it comes to the OPCW method published in paper III andstudied for rate constants in paper IV, again, further testing is theway forward. Since OPCW should be able to converge toward theexact classical Wigner method, as the number of beads in the pathintegral is increased, for any physical potential energy surface, itwould be interesting to try it against other approximate classicalWigner methods, apart from FK-LPI, to see if the method is animprovement compared to what is already available. Since negativecurvature of the potential energy surface can be problematic for othermethods, this is of course one kind of system that is extra interestingto look at. The rate constant calculations should be extended tomultidimensional systems, such as a double well potential with a bathof harmonic oscillators, as a step of getting closer to real molecularsystems.

Improving the numerical performance of the OPCW method isalso an important aspect to consider, if the method is to be practicalfor large systems.

101

Page 123: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

3. Developments

3.4 Future outlook

The research presented in this thesis contains a rate constant for theformation of the hydroxyl radical through radiative association, twonew types of basis functions for use with a variational principle forthe time evolution of Wigner functions, and a method for samplingthe initial distributions in the classical Wigner method. After thesummation of all this research, the question that remain is: Whereto go from here?

The new reaction rate constant was published in 2015, but hasstill not¶ been included in any astrochemical database and noneof its citations indicates that it has been used for astrochemicalmodeling.

The database KIDA86 has instructions for suggesting new ad-ditions to their database. It is an oversight by the author that nosuch suggestion has been sent. To make this rate constant actuallybenefit the astrochemical community, the first thing that has to bedone is to actually suggest that it is included in a database.

The calculation of reaction rate constants for other radiativeassociation reactions and the development of the methods to conductthese calculations has continued in the research group, and amongcollaborators, of the author since the publication of paper I.106–114

One of the main directions that further research is going toward is thecalculation of radiative association rate constants for systems largerthan diatoms. As radiative association generally becomes more likelythe larger the molecule, due to a more long lived activated complex,as explained e.g. in the textbook by Tielens,115 this seems to be theright way to go.

For the Gaussian basis functions for use with a variational princi-ple for the time propagation of Wigner functions, presented in paperII, more testing would be needed to know in what situations theycould be a beneficial route to go for calculations. Further testingcould also help to find ways of improving the current methods. Mul-tidimensional potentials and other correlation functions than theposition autocorrelation function, would be the obvious first stepsin such testing. As was suggested in the paper, a way of improving

¶When writing this, 30 April 2020.

100

3.4. Future outlook

the dynamics of the frozen Gaussian basis functions could be to givethem a semiclassical prefactor that approximately would account forcoupling between the separate basis functions.

When it comes to the OPCW method published in paper III andstudied for rate constants in paper IV, again, further testing is theway forward. Since OPCW should be able to converge toward theexact classical Wigner method, as the number of beads in the pathintegral is increased, for any physical potential energy surface, itwould be interesting to try it against other approximate classicalWigner methods, apart from FK-LPI, to see if the method is animprovement compared to what is already available. Since negativecurvature of the potential energy surface can be problematic for othermethods, this is of course one kind of system that is extra interestingto look at. The rate constant calculations should be extended tomultidimensional systems, such as a double well potential with a bathof harmonic oscillators, as a step of getting closer to real molecularsystems.

Improving the numerical performance of the OPCW method isalso an important aspect to consider, if the method is to be practicalfor large systems.

101

Page 124: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Bibliography

1M. Shelley, Frankenstein, or The Modern Prometheus. 1831.

2 P. A. M. Dirac, “Quantum mechanics of many-electron systems,”Proceedings of the Royal Society of London. Series A, Contain-ing Papers of a Mathematical and Physical Character, vol. 123,pp. 714–733, Apr 1929.

3M. Born and R. Oppenheimer, “Zur Quantentheorie der Molekeln,”Annalen der Physik, vol. 389, no. 20, pp. 457–484, 1927.

4T. E. Markland, S. Habershon, and D. E. Manolopoulos, “Quan-tum diffusion of hydrogen and muonium atoms in liquid waterand hexagonal ice,” The Journal of Chemical Physics, vol. 128,p. 194506, May 2008.

5 S. K.-M. Svensson, M. Gustafsson, and G. Nyman, “Formationof the hydroxyl radical by radiative association,” The Journal ofPhysical Chemistry A, vol. 119, pp. 12263–9, Dec 2015.

6 L. Wang, S. D. Fried, S. G. Boxer, and T. E. Markland, “Quantumdelocalization of protons in the hydrogen-bond network of anenzyme active site,” Proceedings of the National Academy ofSciences, vol. 111, pp. 18454–18459, Dec 2014.

7 E. Schrodinger, “Quantisierung als Eigenwertproblem,” Annalender Physik, vol. 384, pp. 361–376, Jan 1926.

8 E. Schrodinger, “Quantisierung als Eigenwertproblem,” Annalender Physik, vol. 384, pp. 489–527, Feb 1926.

103

Page 125: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Bibliography

1M. Shelley, Frankenstein, or The Modern Prometheus. 1831.

2 P. A. M. Dirac, “Quantum mechanics of many-electron systems,”Proceedings of the Royal Society of London. Series A, Contain-ing Papers of a Mathematical and Physical Character, vol. 123,pp. 714–733, Apr 1929.

3M. Born and R. Oppenheimer, “Zur Quantentheorie der Molekeln,”Annalen der Physik, vol. 389, no. 20, pp. 457–484, 1927.

4T. E. Markland, S. Habershon, and D. E. Manolopoulos, “Quan-tum diffusion of hydrogen and muonium atoms in liquid waterand hexagonal ice,” The Journal of Chemical Physics, vol. 128,p. 194506, May 2008.

5 S. K.-M. Svensson, M. Gustafsson, and G. Nyman, “Formationof the hydroxyl radical by radiative association,” The Journal ofPhysical Chemistry A, vol. 119, pp. 12263–9, Dec 2015.

6 L. Wang, S. D. Fried, S. G. Boxer, and T. E. Markland, “Quantumdelocalization of protons in the hydrogen-bond network of anenzyme active site,” Proceedings of the National Academy ofSciences, vol. 111, pp. 18454–18459, Dec 2014.

7 E. Schrodinger, “Quantisierung als Eigenwertproblem,” Annalender Physik, vol. 384, pp. 361–376, Jan 1926.

8 E. Schrodinger, “Quantisierung als Eigenwertproblem,” Annalender Physik, vol. 384, pp. 489–527, Feb 1926.

103

Page 126: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Bibliography

9 E. Schrodinger, “Uber das Verhaltnis der Heisenberg-Born-Jordanschen Quantenmechanik zu der meinem,” Annalen derPhysik, vol. 384, pp. 734–756, Mar 1926.

10 E. Schrodinger, “Quantisierung als Eigenwertproblem,” Annalender Physik, vol. 385, pp. 437–490, May 1926.

11 E. Schrodinger, “Quantisierung als Eigenwertproblem,” Annalender Physik, vol. 386, pp. 109–139, Jun 1926.

12 E. Schrodinger, “Der stetige Ubergang von der Mikro- zurMakromechanik,” Die Naturwissenschaften, vol. 14, pp. 664–666,Jul 1926.

13 E. Schrodinger, “An undulatory theory of the mechanics of atomsand molecules,” Physical Review, vol. 28, pp. 1049–1070, Dec1926.

14 E. J. Heller, “Wigner phase space method: Analysis for semi-classical applications,” The Journal of Chemical Physics, vol. 65,pp. 1289–1298, Aug 1976.

15 E. Wigner, “On the quantum correction for thermodynamic equi-librium,” Physical Review, vol. 40, pp. 749–759, Jun 1932.

16 J. E. Moyal, “Quantum mechanics as a statistical theory,” Math-ematical Proceedings of the Cambridge Philosophical Society,vol. 45, pp. 99–124, Jan 1949.

17H.Wang, X. Sun, andW. H. Miller, “Semiclassical approximationsfor the calculation of thermal rate constants for chemical reactionsin complex molecular systems,” The Journal of Chemical Physics,vol. 108, pp. 9726 – 9736, Jun 1998.

18W. H. Miller, “Classical S matrix: Numerical application toinelastic collisions,” The Journal of Chemical Physics, vol. 53,pp. 3578–3587, Nov 1970.

19W. H. Miller, “The semiclassical initial value representation: apotentially practical way for adding quantum effects to classi-cal molecular dynamics simulations,” The Journal of PhysicalChemistry A, vol. 105, pp. 2942–2955, Mar 2001.

104

Bibliography

20X. Sun, H. Wang, and W. H. Miller, “On the semiclassical de-scription of quantum coherence in thermal rate constants,” TheJournal of Chemical Physics, vol. 109, pp. 4190–4200, Sep 1998.

21Q. Shi and E. Geva, “A relationship between semiclassical andcentroid correlation functions,” The Journal of Chemical Physics,vol. 118, pp. 8173–8184, May 2003.

22 J. A. Poulsen, G. Nyman, and P. J. Rossky, “Practical evaluationof condensed phase quantum correlation functions: A Feynman–Kleinert variational linearized path integral method,” The Journalof Chemical Physics, vol. 119, pp. 12179–12193, Dec 2003.

23Q. Shi and E. Geva, “Semiclassical theory of vibrational energyrelaxation in the condensed phase,” The Journal of PhysicalChemistry A, vol. 107, pp. 9059–9069, Oct 2003.

24Q. Shi and E. Geva, “Vibrational energy relaxation in liquidoxygen from a semiclassical molecular dynamics simulation,” TheJournal of Physical Chemistry A, vol. 107, pp. 9070–9078, Oct2003.

25 J. A. Poulsen, J. Scheers, G. Nyman, and P. J. Rossky, “Quantumdensity fluctuations in liquid neon from linearized path-integralcalculations,” Physical Review B, vol. 75, p. 224505, Jun 2007.

26 J. A. Poulsen, G. Nyman, and P. J. Rossky, “Static and dynamicquantum effects in molecular liquids: A linearized path integraldescription of water,” Proc Natl Acad Sci U S A, vol. 102, pp. 6709–6714, May 2005.

27N. Markovic and J. A. Poulsen, “A linearized path integral de-scription of the collision process between a water molecule and agraphite surface,” The Journal of Physical Chemistry A, vol. 112,pp. 1701–1711, Feb 2008.

28 S. Habershon and D. E. Manolopoulos, “Zero point energy leak-age in condensed phase dynamics: An assessment of quantumsimulation methods for liquid water,” The Journal of ChemicalPhysics, vol. 131, p. 244518, Dec 2009.

105

Page 127: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Bibliography

9 E. Schrodinger, “Uber das Verhaltnis der Heisenberg-Born-Jordanschen Quantenmechanik zu der meinem,” Annalen derPhysik, vol. 384, pp. 734–756, Mar 1926.

10 E. Schrodinger, “Quantisierung als Eigenwertproblem,” Annalender Physik, vol. 385, pp. 437–490, May 1926.

11 E. Schrodinger, “Quantisierung als Eigenwertproblem,” Annalender Physik, vol. 386, pp. 109–139, Jun 1926.

12 E. Schrodinger, “Der stetige Ubergang von der Mikro- zurMakromechanik,” Die Naturwissenschaften, vol. 14, pp. 664–666,Jul 1926.

13 E. Schrodinger, “An undulatory theory of the mechanics of atomsand molecules,” Physical Review, vol. 28, pp. 1049–1070, Dec1926.

14 E. J. Heller, “Wigner phase space method: Analysis for semi-classical applications,” The Journal of Chemical Physics, vol. 65,pp. 1289–1298, Aug 1976.

15 E. Wigner, “On the quantum correction for thermodynamic equi-librium,” Physical Review, vol. 40, pp. 749–759, Jun 1932.

16 J. E. Moyal, “Quantum mechanics as a statistical theory,” Math-ematical Proceedings of the Cambridge Philosophical Society,vol. 45, pp. 99–124, Jan 1949.

17H.Wang, X. Sun, andW. H. Miller, “Semiclassical approximationsfor the calculation of thermal rate constants for chemical reactionsin complex molecular systems,” The Journal of Chemical Physics,vol. 108, pp. 9726 – 9736, Jun 1998.

18W. H. Miller, “Classical S matrix: Numerical application toinelastic collisions,” The Journal of Chemical Physics, vol. 53,pp. 3578–3587, Nov 1970.

19W. H. Miller, “The semiclassical initial value representation: apotentially practical way for adding quantum effects to classi-cal molecular dynamics simulations,” The Journal of PhysicalChemistry A, vol. 105, pp. 2942–2955, Mar 2001.

104

Bibliography

20X. Sun, H. Wang, and W. H. Miller, “On the semiclassical de-scription of quantum coherence in thermal rate constants,” TheJournal of Chemical Physics, vol. 109, pp. 4190–4200, Sep 1998.

21Q. Shi and E. Geva, “A relationship between semiclassical andcentroid correlation functions,” The Journal of Chemical Physics,vol. 118, pp. 8173–8184, May 2003.

22 J. A. Poulsen, G. Nyman, and P. J. Rossky, “Practical evaluationof condensed phase quantum correlation functions: A Feynman–Kleinert variational linearized path integral method,” The Journalof Chemical Physics, vol. 119, pp. 12179–12193, Dec 2003.

23Q. Shi and E. Geva, “Semiclassical theory of vibrational energyrelaxation in the condensed phase,” The Journal of PhysicalChemistry A, vol. 107, pp. 9059–9069, Oct 2003.

24Q. Shi and E. Geva, “Vibrational energy relaxation in liquidoxygen from a semiclassical molecular dynamics simulation,” TheJournal of Physical Chemistry A, vol. 107, pp. 9070–9078, Oct2003.

25 J. A. Poulsen, J. Scheers, G. Nyman, and P. J. Rossky, “Quantumdensity fluctuations in liquid neon from linearized path-integralcalculations,” Physical Review B, vol. 75, p. 224505, Jun 2007.

26 J. A. Poulsen, G. Nyman, and P. J. Rossky, “Static and dynamicquantum effects in molecular liquids: A linearized path integraldescription of water,” Proc Natl Acad Sci U S A, vol. 102, pp. 6709–6714, May 2005.

27N. Markovic and J. A. Poulsen, “A linearized path integral de-scription of the collision process between a water molecule and agraphite surface,” The Journal of Physical Chemistry A, vol. 112,pp. 1701–1711, Feb 2008.

28 S. Habershon and D. E. Manolopoulos, “Zero point energy leak-age in condensed phase dynamics: An assessment of quantumsimulation methods for liquid water,” The Journal of ChemicalPhysics, vol. 131, p. 244518, Dec 2009.

105

Page 128: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Bibliography

29K. Boye, “In motion.” Karin Boye: Complete poems, Trans. D.McDuff, Bloodaxe Books, 1994.

30R. P. Feynman, “Space-time approach to non-relativistic quantummechanics,” Reviews of Modern Physics, vol. 20, pp. 367–387,Apr 1948.

31R. P. Feynman and A. R. Hibbs, Quantum Mechanics and PathIntegrals. McGraw-Hill Companies, Inc., New York, 1965.

32H. F. Trotter, “On the product of semi-groups of operators,” Pro-ceedings of the American Mathematical Society, vol. 10, pp. 545–551, Aug 1959.

33 J. A. Barker, “A quantum-statistical Monte Carlo method; pathintegrals with boundary conditions,” The Journal of ChemicalPhysics, vol. 70, pp. 2914–2918, Mar 1979.

34M. Parrinello and A. Rahman, “Study of an F center in moltenKCl,” The Journal of Chemical Physics, vol. 80, pp. 860–867, Jan1984.

35 J. Cao and G. A. Voth, “A new perspective on quantum timecorrelation functions,” The Journal of Chemical Physics, vol. 99,pp. 10070–10073, Dec 1993.

36 I. R. Craig and D. E. Manolopoulos, “Quantum statistics andclassical mechanics: real time correlation functions from ringpolymer molecular dynamics,” The Journal of Chemical Physics,vol. 121, pp. 3368–73, Aug 2004.

37W. H. Miller, “Quantum mechanical transition state theory and anew semiclassical model for reaction rate constants,” The Journalof Chemical Physics, vol. 61, pp. 1823–1834, Sep 1974.

38W. H. Miller, S. D. Schwartz, and J. W. Tromp, “Quantummechanical rate constants for bimolecular reactions,” The Journalof Chemical Physics, vol. 79, pp. 4889–4898, Nov 1983.

39 F. Jensen, Introduction to Computational Chemistry. England:John Wiley & Sons Ltd, 2nd ed., 2007.

106

Bibliography

40R. Kubo, “Statistical-mechanistical theory of irreversible pro-cesses. I. general theory and simple application to magnetic andconduction problems,” Journal of the Physical Society of Japan,vol. 12, pp. 570–586, Jun 1957.

41 P. Schofield, “Space-time correlation function formalism for slowneutron scattering,” Physical Review Letters, vol. 4, pp. 239–240,Mar 1960.

42T. J. H. Hele, M. J. Willatt, A. Muolo, and S. C. Althorpe,“Boltzmann-conserving classical dynamics in quantum time-correlation functions: “Matsubara dynamics”,” The Journal ofChemical Physics, vol. 142, p. 134103, Apr 2015.

43K. A. Jung, P. E. Videla, and V. S. Batista, “Multi-time formula-tion of Matsubara dynamics,” The Journal of Chemical Physics,vol. 151, p. 034108, Jul 2019.

44M. J. Willatt, M. Ceriotti, and S. C. Althorpe, “ApproximatingMatsubara dynamics using the planetary model: Tests on liq-uid water and ice,” The Journal of Chemical Physics, vol. 148,p. 102336, Jan 2018.

45T. J. H. Hele, M. J. Willatt, A. Muolo, and S. C. Althorpe,“Communication: Relation of centroid molecular dynamics andring-polymer molecular dynamics to exact quantum dynamics,”The Journal of Chemical Physics, vol. 142, p. 191101, May 2015.

46D. R. Reichman, P.-N. Roy, S. Jang, and G. A. Voth, “A Feynmanpath centroid dynamics approach for the computation of timecorrelation functions involving nonlinear operators,” The Journalof Chemical Physics, vol. 113, pp. 919–929, Jul 2000.

47A. Witt, S. D. Ivanov, M. Shiga, H. Forbert, and D. Marx, “On theapplicability of centroid and ring polymer path integral moleculardynamics for vibrational spectroscopy,” The Journal of ChemicalPhysics, vol. 130, p. 194510, May 2009.

48 I. R. Craig and D. E. Manolopoulos, “Chemical reaction ratesfrom ring polymer molecular dynamics,” The Journal of ChemicalPhysics, vol. 122, p. 084106, Feb 2005.

107

Page 129: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Bibliography

29K. Boye, “In motion.” Karin Boye: Complete poems, Trans. D.McDuff, Bloodaxe Books, 1994.

30R. P. Feynman, “Space-time approach to non-relativistic quantummechanics,” Reviews of Modern Physics, vol. 20, pp. 367–387,Apr 1948.

31R. P. Feynman and A. R. Hibbs, Quantum Mechanics and PathIntegrals. McGraw-Hill Companies, Inc., New York, 1965.

32H. F. Trotter, “On the product of semi-groups of operators,” Pro-ceedings of the American Mathematical Society, vol. 10, pp. 545–551, Aug 1959.

33 J. A. Barker, “A quantum-statistical Monte Carlo method; pathintegrals with boundary conditions,” The Journal of ChemicalPhysics, vol. 70, pp. 2914–2918, Mar 1979.

34M. Parrinello and A. Rahman, “Study of an F center in moltenKCl,” The Journal of Chemical Physics, vol. 80, pp. 860–867, Jan1984.

35 J. Cao and G. A. Voth, “A new perspective on quantum timecorrelation functions,” The Journal of Chemical Physics, vol. 99,pp. 10070–10073, Dec 1993.

36 I. R. Craig and D. E. Manolopoulos, “Quantum statistics andclassical mechanics: real time correlation functions from ringpolymer molecular dynamics,” The Journal of Chemical Physics,vol. 121, pp. 3368–73, Aug 2004.

37W. H. Miller, “Quantum mechanical transition state theory and anew semiclassical model for reaction rate constants,” The Journalof Chemical Physics, vol. 61, pp. 1823–1834, Sep 1974.

38W. H. Miller, S. D. Schwartz, and J. W. Tromp, “Quantummechanical rate constants for bimolecular reactions,” The Journalof Chemical Physics, vol. 79, pp. 4889–4898, Nov 1983.

39 F. Jensen, Introduction to Computational Chemistry. England:John Wiley & Sons Ltd, 2nd ed., 2007.

106

Bibliography

40R. Kubo, “Statistical-mechanistical theory of irreversible pro-cesses. I. general theory and simple application to magnetic andconduction problems,” Journal of the Physical Society of Japan,vol. 12, pp. 570–586, Jun 1957.

41 P. Schofield, “Space-time correlation function formalism for slowneutron scattering,” Physical Review Letters, vol. 4, pp. 239–240,Mar 1960.

42T. J. H. Hele, M. J. Willatt, A. Muolo, and S. C. Althorpe,“Boltzmann-conserving classical dynamics in quantum time-correlation functions: “Matsubara dynamics”,” The Journal ofChemical Physics, vol. 142, p. 134103, Apr 2015.

43K. A. Jung, P. E. Videla, and V. S. Batista, “Multi-time formula-tion of Matsubara dynamics,” The Journal of Chemical Physics,vol. 151, p. 034108, Jul 2019.

44M. J. Willatt, M. Ceriotti, and S. C. Althorpe, “ApproximatingMatsubara dynamics using the planetary model: Tests on liq-uid water and ice,” The Journal of Chemical Physics, vol. 148,p. 102336, Jan 2018.

45T. J. H. Hele, M. J. Willatt, A. Muolo, and S. C. Althorpe,“Communication: Relation of centroid molecular dynamics andring-polymer molecular dynamics to exact quantum dynamics,”The Journal of Chemical Physics, vol. 142, p. 191101, May 2015.

46D. R. Reichman, P.-N. Roy, S. Jang, and G. A. Voth, “A Feynmanpath centroid dynamics approach for the computation of timecorrelation functions involving nonlinear operators,” The Journalof Chemical Physics, vol. 113, pp. 919–929, Jul 2000.

47A. Witt, S. D. Ivanov, M. Shiga, H. Forbert, and D. Marx, “On theapplicability of centroid and ring polymer path integral moleculardynamics for vibrational spectroscopy,” The Journal of ChemicalPhysics, vol. 130, p. 194510, May 2009.

48 I. R. Craig and D. E. Manolopoulos, “Chemical reaction ratesfrom ring polymer molecular dynamics,” The Journal of ChemicalPhysics, vol. 122, p. 084106, Feb 2005.

107

Page 130: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Bibliography

49T. F. Miller and D. E. Manolopoulos, “Quantum diffusion inliquid para-hydrogen from ring-polymer molecular dynamics,”The Journal of Chemical Physics, vol. 122, p. 184503, May 2005.

50M. Rossi, M. Ceriotti, and D. E. Manolopoulos, “How to removethe spurious resonances from ring polymer molecular dynamics,”The Journal of Chemical Physics, vol. 140, p. 234116, Jun 2014.

51 P. A. M. Dirac, “Note on exchange phenomena in the Thomasatom,” Mathematical Proceedings of the Cambridge PhilosophicalSociety, vol. 26, pp. 376–385, Jul 1930.

52 J. I. Frenkel, Wave Mechanics; Advanced General Theory. Oxford:Clarendon Press, 1934.

53A. McLachlan, “A variational solution of the time-dependentSchrodinger equation,” Molecular Physics, vol. 8, no. 1, pp. 39–44, 1964.

54H.-D. Meyer, U. Manthe, and L. Cederbaum, “The multi-configurational time-dependent Hartree approach,” ChemicalPhysics Letters, vol. 165, pp. 73–78, Jan 1990.

55 S. Tzu, The Art of War, Trans. L. Giles. London: Luzac & Co.,1910.

56 S. Weinreb, A. H. Barrett, M. L. Meeks, and J. C. Henry, “Radioobservations of OH in the interstellar medium,” Nature, vol. 200,pp. 829–831, Nov 1963.

57 L. Weliachew, “Detection of interstellar OH in two external galax-ies,” The Astrophysical Journal, vol. 167, pp. L47–L52, Jul 1971.

58 E. Herbst, “Chemistry in the interstellar medium,” Annual Reviewof Physical Chemistry, vol. 46, pp. 27–54, Oct 1995.

59 E. F. van Dishoeck, S. R. Langhoff, and A. Dalgarno, “Thelow-lying 2Σ− states of OH,” The Journal of Chemical Physics,vol. 78, pp. 4552–4561, Apr 1983.

108

Bibliography

60D. R. Yarkony, “A theoretical treatment of the predissociation ofthe individual rovibronic levels of OH/OD(A 2Σ+),” The Journalof Chemical Physics, vol. 97, pp. 1838–1849, Aug 1992.

61H. van Lonkhuyzen and C. A. de Lange, “U.V. photoelectronspectroscopy of OH and OD radicals,” Molecular Physics, vol. 51,pp. 551–568, Sep 1984.

62K. P. Huber and G. Herzberg, Constants of diatomic molecules,vol. 4 of Molecular Spectra and Molecular Structure. Springer,Boston, MA, USA, 1979.

63A. Kramida, Y. Ralchenko, J. Reader, and NIST ASD Team,“Atomic spectra database, ver. 5.2.” National Institute of Stan-dards and Technology, Gaitersburg, MD, USA, Sep 2014.

64 S. R. Langhoff, C. W. Bauschlicher Jr., and P. R. Taylor, “Theo-retical study of the dipole moment function of OH(X 2Π),” TheJournal of Chemical Physics, vol. 91, pp. 5953–5959, Nov 1989.

65 B. H. Bransden and C. J. Joachain, Quantum Mechanics. Harlow;England: Pearson Education Limited, 2nd ed., 2000.

66 J. F. Babb and A. Dalgarno, “Radiative association and inversepredissociation of oxygen atoms,” Physical Review A, vol. 51,pp. 3021–3026, Apr 1995.

67 G. Barinovs and M. C. van Hemert, “CH+ radiative association,”The Astrophysical Journal, vol. 636, pp. 923–926, Jan 2006.

68B. V. Noumerov, “A method of extrapolation of perturbations,”Monthly Notices of the Royal Astronomical Society, vol. 84,pp. 592–602, Jun 1924.

69D. T. Colbert and W. H. Miller, “A novel discrete variable rep-resentation for quantum mechanical reactive scattering via theS-matrix Kohn method,” The Journal of Chemical Physics, vol. 96,pp. 1982–1991, Feb 1992.

70 J. Franck and E. G. Dymond, “Elementary processes of photo-chemical reactions,” Transactions of the Faraday Society, vol. 21,pp. 536–542, Feb 1926.

109

Page 131: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Bibliography

49T. F. Miller and D. E. Manolopoulos, “Quantum diffusion inliquid para-hydrogen from ring-polymer molecular dynamics,”The Journal of Chemical Physics, vol. 122, p. 184503, May 2005.

50M. Rossi, M. Ceriotti, and D. E. Manolopoulos, “How to removethe spurious resonances from ring polymer molecular dynamics,”The Journal of Chemical Physics, vol. 140, p. 234116, Jun 2014.

51 P. A. M. Dirac, “Note on exchange phenomena in the Thomasatom,” Mathematical Proceedings of the Cambridge PhilosophicalSociety, vol. 26, pp. 376–385, Jul 1930.

52 J. I. Frenkel, Wave Mechanics; Advanced General Theory. Oxford:Clarendon Press, 1934.

53A. McLachlan, “A variational solution of the time-dependentSchrodinger equation,” Molecular Physics, vol. 8, no. 1, pp. 39–44, 1964.

54H.-D. Meyer, U. Manthe, and L. Cederbaum, “The multi-configurational time-dependent Hartree approach,” ChemicalPhysics Letters, vol. 165, pp. 73–78, Jan 1990.

55 S. Tzu, The Art of War, Trans. L. Giles. London: Luzac & Co.,1910.

56 S. Weinreb, A. H. Barrett, M. L. Meeks, and J. C. Henry, “Radioobservations of OH in the interstellar medium,” Nature, vol. 200,pp. 829–831, Nov 1963.

57 L. Weliachew, “Detection of interstellar OH in two external galax-ies,” The Astrophysical Journal, vol. 167, pp. L47–L52, Jul 1971.

58 E. Herbst, “Chemistry in the interstellar medium,” Annual Reviewof Physical Chemistry, vol. 46, pp. 27–54, Oct 1995.

59 E. F. van Dishoeck, S. R. Langhoff, and A. Dalgarno, “Thelow-lying 2Σ− states of OH,” The Journal of Chemical Physics,vol. 78, pp. 4552–4561, Apr 1983.

108

Bibliography

60D. R. Yarkony, “A theoretical treatment of the predissociation ofthe individual rovibronic levels of OH/OD(A 2Σ+),” The Journalof Chemical Physics, vol. 97, pp. 1838–1849, Aug 1992.

61H. van Lonkhuyzen and C. A. de Lange, “U.V. photoelectronspectroscopy of OH and OD radicals,” Molecular Physics, vol. 51,pp. 551–568, Sep 1984.

62K. P. Huber and G. Herzberg, Constants of diatomic molecules,vol. 4 of Molecular Spectra and Molecular Structure. Springer,Boston, MA, USA, 1979.

63A. Kramida, Y. Ralchenko, J. Reader, and NIST ASD Team,“Atomic spectra database, ver. 5.2.” National Institute of Stan-dards and Technology, Gaitersburg, MD, USA, Sep 2014.

64 S. R. Langhoff, C. W. Bauschlicher Jr., and P. R. Taylor, “Theo-retical study of the dipole moment function of OH(X 2Π),” TheJournal of Chemical Physics, vol. 91, pp. 5953–5959, Nov 1989.

65 B. H. Bransden and C. J. Joachain, Quantum Mechanics. Harlow;England: Pearson Education Limited, 2nd ed., 2000.

66 J. F. Babb and A. Dalgarno, “Radiative association and inversepredissociation of oxygen atoms,” Physical Review A, vol. 51,pp. 3021–3026, Apr 1995.

67 G. Barinovs and M. C. van Hemert, “CH+ radiative association,”The Astrophysical Journal, vol. 636, pp. 923–926, Jan 2006.

68B. V. Noumerov, “A method of extrapolation of perturbations,”Monthly Notices of the Royal Astronomical Society, vol. 84,pp. 592–602, Jun 1924.

69D. T. Colbert and W. H. Miller, “A novel discrete variable rep-resentation for quantum mechanical reactive scattering via theS-matrix Kohn method,” The Journal of Chemical Physics, vol. 96,pp. 1982–1991, Feb 1992.

70 J. Franck and E. G. Dymond, “Elementary processes of photo-chemical reactions,” Transactions of the Faraday Society, vol. 21,pp. 536–542, Feb 1926.

109

Page 132: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Bibliography

71 E. U. Condon, “A theory of intensity distribution in band systems,”Physical Review, vol. 28, pp. 1182–1201, Dec 1926.

72 E. U. Condon, “Nuclear motions associated with electron transi-tions in diatomic molecules,” Physical Review, vol. 32, pp. 858–872,Dec 1928.

73G. Wentzel, “Eine Verallgemeinerung der Quantenbedingungenfur die Zwecke der Wellenmechanik,” Zeitschrift fur Physik, vol. 38,pp. 518–529, Jun 1926.

74H. A. Kramers, “Wellenmechanik und halbzahlige Quantisierung,”Zeitschrift fur Physik, vol. 39, pp. 828–840, Oct 1926.

75 L. Brillouin, “La mecanique ondulatoire de Schrodinger: unemethode generale de resolution par approximations successives,”Comptes Rendus de l’Academie des Sciences, vol. 183, pp. 24–26,1926.

76M. Gustafsson, “Classical calculations of radiative associationin absence of electronic transitions,” The Journal of ChemicalPhysics, vol. 138, pp. 074308–1–074308–7, Feb 2013.

77 J. Larmor, “On the theory of the magnetic influence on spectra;and on the radiation from moving ions,” The London, Edinburgh,and Dublin Philosophical Magazine and Journal of Science, Series5, vol. 44, no. 271, pp. 503–512, 1897.

78G. Breit and E. Wigner, “Capture of slow neutrons,” PhysicalReview, vol. 49, pp. 519–531, Apr 1936.

79R. J. Le Roy, “LEVEL 8.0 a computer program for solving theradial Schrodinger equation for bound and quasibound levels, in-cluding corrections as of December 2012.” University of WaterlooChemical Physics Research Report CP-663, Apr 2007.

80 J. A. Fedchak, M. A. Huels, L. D. Doverspike, and R. L. Champion,“Electron detachment and charge transfer for collisions of O- andS- with H,” Physical Review A, vol. 47, pp. 3796–3800, May 1993.

110

Bibliography

81U. Jentschura, S. Kotochigova, E. LeBigot, P. Mohr, and B. Taylor,“The energy levels of hydrogen and deuterium, ver. 2.1.” NationalInstitute of Standards and Technology, Gaitersburg, MD, USA,Jul 2005.

82K. Eriksson and H. Isberg, “New measurements in the spectrumof atomic oxygen, O I,” Arkiv for fysik, vol. 37, pp. 221–230, 1968.

83C. E. Moore, Selected Tables of Atomic Spectra, Atomic EnergyLevels and Multiplet Tables – O I, vol. 3 of National StandardReference Data System. National Bureau of Standards, U.S., Apr1976.

84K. R. Lykke, K. K. Murray, and W. C. Lineberger, “Thresholdphotodetachment of H-,” Physical Review A, vol. 43, pp. 6104–6107, Jun 1991.

85D. Neumark, K. R. Lykke, T. Andersen, and W. C. Lineberger,“Laser photodetachment measurement of the electron affinity ofatomic oxygen,” Physical Review A, vol. 32, pp. 1890–1892, Sep1985.

86V. Wakelam, E. Herbst, J.-C. Loison, I. Smith, V. Chan-drasekaran, B. Pavone, N. Adams, M.-C. Bacchus-Motabonel,A. Bergeat, K. Beroff, V. Bierbaum, M. Chabot, A. Dalgarno,E. F. van Dishoeck, A. Faure, W. Geppert, D. Gerlich, D. Galli,E. Hebrard, F. Hersant, K. Hickson, P. Honvault, S. Klippenstein,S. Le Picard, G. Nyman, P. Pernot, S. Schlemmer, F. Selsis,I. Sims, D. Talbi, J. Tennyson, J. Troe, R. Wester, and L. Wiesen-feld, “A KInetic Database for Astrochemistry (KIDA),” TheAstrophysical Journal Supplement Series, vol. 199, p. 21, Mar2012.

87A. Kramida, Y. Ralchenko, J. Reader, and NIST ASD Team,“NIST atomic spectra database, ver. 5.7.1.” National Institute ofStandards and Technology, Gaitersburg, MD, USA, Apr 2020.

88A. Kramida, “A critical compilation of experimental data onspectral lines and energy levels of hydrogen, deuterium, andtritium,” Atomic Data and Nuclear Data Tables, vol. 96, pp. 586–644, Nov 2010.

111

Page 133: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Bibliography

71 E. U. Condon, “A theory of intensity distribution in band systems,”Physical Review, vol. 28, pp. 1182–1201, Dec 1926.

72 E. U. Condon, “Nuclear motions associated with electron transi-tions in diatomic molecules,” Physical Review, vol. 32, pp. 858–872,Dec 1928.

73G. Wentzel, “Eine Verallgemeinerung der Quantenbedingungenfur die Zwecke der Wellenmechanik,” Zeitschrift fur Physik, vol. 38,pp. 518–529, Jun 1926.

74H. A. Kramers, “Wellenmechanik und halbzahlige Quantisierung,”Zeitschrift fur Physik, vol. 39, pp. 828–840, Oct 1926.

75 L. Brillouin, “La mecanique ondulatoire de Schrodinger: unemethode generale de resolution par approximations successives,”Comptes Rendus de l’Academie des Sciences, vol. 183, pp. 24–26,1926.

76M. Gustafsson, “Classical calculations of radiative associationin absence of electronic transitions,” The Journal of ChemicalPhysics, vol. 138, pp. 074308–1–074308–7, Feb 2013.

77 J. Larmor, “On the theory of the magnetic influence on spectra;and on the radiation from moving ions,” The London, Edinburgh,and Dublin Philosophical Magazine and Journal of Science, Series5, vol. 44, no. 271, pp. 503–512, 1897.

78G. Breit and E. Wigner, “Capture of slow neutrons,” PhysicalReview, vol. 49, pp. 519–531, Apr 1936.

79R. J. Le Roy, “LEVEL 8.0 a computer program for solving theradial Schrodinger equation for bound and quasibound levels, in-cluding corrections as of December 2012.” University of WaterlooChemical Physics Research Report CP-663, Apr 2007.

80 J. A. Fedchak, M. A. Huels, L. D. Doverspike, and R. L. Champion,“Electron detachment and charge transfer for collisions of O- andS- with H,” Physical Review A, vol. 47, pp. 3796–3800, May 1993.

110

Bibliography

81U. Jentschura, S. Kotochigova, E. LeBigot, P. Mohr, and B. Taylor,“The energy levels of hydrogen and deuterium, ver. 2.1.” NationalInstitute of Standards and Technology, Gaitersburg, MD, USA,Jul 2005.

82K. Eriksson and H. Isberg, “New measurements in the spectrumof atomic oxygen, O I,” Arkiv for fysik, vol. 37, pp. 221–230, 1968.

83C. E. Moore, Selected Tables of Atomic Spectra, Atomic EnergyLevels and Multiplet Tables – O I, vol. 3 of National StandardReference Data System. National Bureau of Standards, U.S., Apr1976.

84K. R. Lykke, K. K. Murray, and W. C. Lineberger, “Thresholdphotodetachment of H-,” Physical Review A, vol. 43, pp. 6104–6107, Jun 1991.

85D. Neumark, K. R. Lykke, T. Andersen, and W. C. Lineberger,“Laser photodetachment measurement of the electron affinity ofatomic oxygen,” Physical Review A, vol. 32, pp. 1890–1892, Sep1985.

86V. Wakelam, E. Herbst, J.-C. Loison, I. Smith, V. Chan-drasekaran, B. Pavone, N. Adams, M.-C. Bacchus-Motabonel,A. Bergeat, K. Beroff, V. Bierbaum, M. Chabot, A. Dalgarno,E. F. van Dishoeck, A. Faure, W. Geppert, D. Gerlich, D. Galli,E. Hebrard, F. Hersant, K. Hickson, P. Honvault, S. Klippenstein,S. Le Picard, G. Nyman, P. Pernot, S. Schlemmer, F. Selsis,I. Sims, D. Talbi, J. Tennyson, J. Troe, R. Wester, and L. Wiesen-feld, “A KInetic Database for Astrochemistry (KIDA),” TheAstrophysical Journal Supplement Series, vol. 199, p. 21, Mar2012.

87A. Kramida, Y. Ralchenko, J. Reader, and NIST ASD Team,“NIST atomic spectra database, ver. 5.7.1.” National Institute ofStandards and Technology, Gaitersburg, MD, USA, Apr 2020.

88A. Kramida, “A critical compilation of experimental data onspectral lines and energy levels of hydrogen, deuterium, andtritium,” Atomic Data and Nuclear Data Tables, vol. 96, pp. 586–644, Nov 2010.

111

Page 134: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Bibliography

89A. Kramida, “Erratum to “a critical compilation of experimentaldata on spectral lines and energy levels of hydrogen, deuterium,and tritium” [At. Data Nucl. Data Tables 96 (2010) 586–644],”Atomic Data and Nuclear Data Tables, vol. 126, pp. 295–298, Mar2019.

90C. E. Moore, Tables of Spectra of Hydrogen, Carbon, Nitrogen,and Oxygen Atoms and Ions. CRC Series in Evaluated Data inAtomic Physics, Boca Raton, Florida, U.S.A.: CRC Press, 1993.

91D. McElroy, C. Walsh, A. J. Markwick, M. A. Cordiner, K. W.Smith, and T. J. Millar, “The UMIST database for astrochemistry2012,” Astronomy and Astrophysics, vol. 550, p. A36, Feb 2013.

92 J. A. Poulsen, “A variational principle in Wigner phase-space withapplications to statistical mechanics,” The Journal of ChemicalPhysics, vol. 134, p. 034118, Jan 2011.

93 J. A. Poulsen, S. K.-M. Svensson, and G. Nyman, “Dynamicsof Gaussian Wigner functions derived from a time-dependentvariational principle,” AIP Advances, vol. 7, pp. 115018–1–115018–12, Nov 2017.

94R. D. Coalson and M. Karplus, “Multidimensional variationalGaussian wave packet dynamics with application to photodisso-ciation spectroscopy,” The Journal of Chemical Physics, vol. 93,pp. 3919–3930, Sep 1990.

95D. J. Coughtrie and D. P. Tew, “A Gaussian wave packet phase-space representation of quantum canonical statistics,” The Jour-nal of Chemical Physics, vol. 143, p. 044102, Jul 2015.

96R. P. Feynman and H. Kleinert, “Effective classical partitionfunctions,” Physical Review A, vol. 34, pp. 5080–5084, Dec 1986.

97 S. K.-M. Svensson, J. A. Poulsen, and G. Nyman, “ClassicalWigner model based on a Feynman path integral open polymer,”The Journal of Chemical Physics, vol. 152, p. 094111, Mar 2020.

98N. Metropolis, A. W. Rosenbluth, M. N. Rosenbluth, A. H. Teller,and E. Teller, “Equation of state calculations by fast computing

112

Bibliography

machines,” The Journal of Chemical Physics, vol. 21, pp. 1087–1092, Jun 1953.

99 L. Verlet, “Computer ”experiments´´ on classical fluids. I. ther-modynamical properties of Lennard-Jones molecules,” PhysicalReview, vol. 159, pp. 98–103, Jul 1967.

100W. C. Swope, H. C. Andersen, P. H. Berens, and K. R. Wilson,“A computer simulation method for the calculation of equilibriumconstants for the formation of physical clusters of molecules:Application to small water clusters,” The Journal of ChemicalPhysics, vol. 76, pp. 637–649, Jan 1982.

101A. Bose and N. Makri, “Coherent state-based path integralmethodology for computing the Wigner phase space distribution,”The Journal of Physical Chemistry A, vol. 123, pp. 4284–4294,Apr 2019.

102 S. Bonella, M. Monteferrante, C. Pierleoni, and G. Ciccotti, “Pathintegral based calculations of symmetrized time correlation func-tions. II,” The Journal of Chemical Physics, vol. 133, p. 164105,Oct 2010.

103 S. Bonella and G. Ciccotti, “Approximating time-dependent quan-tum statistical properties,” Entropy, vol. 16, pp. 86–109, Dec2013.

104A. O. Caldeira and A. J. Leggett, “Quantum tunnelling in adissipative system,” Annals of Physics, vol. 149, pp. 374–456,September 1983.

105 J. Liu and W. H. Miller, “Using the thermal Gaussian approxima-tion for the Boltzmann operator in semiclassical initial value timecorrelation functions,” The Journal of Chemical Physics, vol. 125,p. 224104, Dec 2006.

106 J. Ostrom, D. S. Bezrukov, G. Nyman, and M. Gustafsson, “Reac-tion rate constant for radiative association of CF+,” The Journalof Chemical Physics, vol. 144, p. 044302, Jan 2016.

113

Page 135: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Bibliography

89A. Kramida, “Erratum to “a critical compilation of experimentaldata on spectral lines and energy levels of hydrogen, deuterium,and tritium” [At. Data Nucl. Data Tables 96 (2010) 586–644],”Atomic Data and Nuclear Data Tables, vol. 126, pp. 295–298, Mar2019.

90C. E. Moore, Tables of Spectra of Hydrogen, Carbon, Nitrogen,and Oxygen Atoms and Ions. CRC Series in Evaluated Data inAtomic Physics, Boca Raton, Florida, U.S.A.: CRC Press, 1993.

91D. McElroy, C. Walsh, A. J. Markwick, M. A. Cordiner, K. W.Smith, and T. J. Millar, “The UMIST database for astrochemistry2012,” Astronomy and Astrophysics, vol. 550, p. A36, Feb 2013.

92 J. A. Poulsen, “A variational principle in Wigner phase-space withapplications to statistical mechanics,” The Journal of ChemicalPhysics, vol. 134, p. 034118, Jan 2011.

93 J. A. Poulsen, S. K.-M. Svensson, and G. Nyman, “Dynamicsof Gaussian Wigner functions derived from a time-dependentvariational principle,” AIP Advances, vol. 7, pp. 115018–1–115018–12, Nov 2017.

94R. D. Coalson and M. Karplus, “Multidimensional variationalGaussian wave packet dynamics with application to photodisso-ciation spectroscopy,” The Journal of Chemical Physics, vol. 93,pp. 3919–3930, Sep 1990.

95D. J. Coughtrie and D. P. Tew, “A Gaussian wave packet phase-space representation of quantum canonical statistics,” The Jour-nal of Chemical Physics, vol. 143, p. 044102, Jul 2015.

96R. P. Feynman and H. Kleinert, “Effective classical partitionfunctions,” Physical Review A, vol. 34, pp. 5080–5084, Dec 1986.

97 S. K.-M. Svensson, J. A. Poulsen, and G. Nyman, “ClassicalWigner model based on a Feynman path integral open polymer,”The Journal of Chemical Physics, vol. 152, p. 094111, Mar 2020.

98N. Metropolis, A. W. Rosenbluth, M. N. Rosenbluth, A. H. Teller,and E. Teller, “Equation of state calculations by fast computing

112

Bibliography

machines,” The Journal of Chemical Physics, vol. 21, pp. 1087–1092, Jun 1953.

99 L. Verlet, “Computer ”experiments´´ on classical fluids. I. ther-modynamical properties of Lennard-Jones molecules,” PhysicalReview, vol. 159, pp. 98–103, Jul 1967.

100W. C. Swope, H. C. Andersen, P. H. Berens, and K. R. Wilson,“A computer simulation method for the calculation of equilibriumconstants for the formation of physical clusters of molecules:Application to small water clusters,” The Journal of ChemicalPhysics, vol. 76, pp. 637–649, Jan 1982.

101A. Bose and N. Makri, “Coherent state-based path integralmethodology for computing the Wigner phase space distribution,”The Journal of Physical Chemistry A, vol. 123, pp. 4284–4294,Apr 2019.

102 S. Bonella, M. Monteferrante, C. Pierleoni, and G. Ciccotti, “Pathintegral based calculations of symmetrized time correlation func-tions. II,” The Journal of Chemical Physics, vol. 133, p. 164105,Oct 2010.

103 S. Bonella and G. Ciccotti, “Approximating time-dependent quan-tum statistical properties,” Entropy, vol. 16, pp. 86–109, Dec2013.

104A. O. Caldeira and A. J. Leggett, “Quantum tunnelling in adissipative system,” Annals of Physics, vol. 149, pp. 374–456,September 1983.

105 J. Liu and W. H. Miller, “Using the thermal Gaussian approxima-tion for the Boltzmann operator in semiclassical initial value timecorrelation functions,” The Journal of Chemical Physics, vol. 125,p. 224104, Dec 2006.

106 J. Ostrom, D. S. Bezrukov, G. Nyman, and M. Gustafsson, “Reac-tion rate constant for radiative association of CF+,” The Journalof Chemical Physics, vol. 144, p. 044302, Jan 2016.

113

Page 136: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Bibliography

107 J. Ostrom, D. S. Bezrukov, G. Nyman, and M. Gustafsson, “Er-ratum: “Reaction rate constant for radiative association of CF+”[J. Chem. Phys. 144, 044302 (2016)],” The Journal of ChemicalPhysics, vol. 150, p. 249901, Jun 2019.

108R. K. Kathir, G. Nyman, and M. Gustafsson, “The rate constantfor formation of HCl through radiative association,” MonthlyNotices of the Royal Astronomical Society, vol. 470, pp. 3068–3070, Jun 2017.

109 P. Szabo and M. Gustafsson, “A surface-hopping method forsemiclassical calculations of cross sections for radiative associationwith electronic transitions,” The Journal of Chemical Physics,vol. 147, p. 094308, Sep 2017.

110T. Stoecklin, P. Halvick, H.-G. Yu, G. Nyman, and Y. Ellinger,“On the gas-phase formation of the HCO radical: accurate quan-tum study of the H+CO radiative association,” Monthly Noticesof the Royal Astronomical Society, vol. 475, pp. 2545–2552, Jan2018.

111M. Zamecnıkova, P. Soldan, M. Gustafsson, and G. Nyman,“Formation of CO+ by radiative association,” Monthly Noticesof the Royal Astronomical Society, vol. 489, pp. 2954–2960, Aug2019.

112M. Gustafsson and R. C. Forrey, “Semiclassical methods forcalculating radiative association rate constants for different ther-modynamic conditions: Application to formation of CO, CN, andSiN,” The Journal of Chemical Physics, vol. 150, p. 224301, Jun2019.

113D. Burdakova, G. Nyman, and T. Stoecklin, “Formation of Na-containing complex molecules in the gas phase in dense molecularclouds: quantum study of the Na+ + H2 and Na+ + D2 radiativeassociation step,” Monthly Notices of the Royal AstronomicalSociety, vol. 485, pp. 5874–5879, Mar 2019.

114M. Zamecnıkova, M. Gustafsson, G. Nyman, and P. Soldan,“Formation of CO+ by radiative association II,” Monthly Notices

114

Bibliography

of the Royal Astronomical Society, vol. 492, pp. 3794–3802, Jan2020.

115A. G. G. M. Tielens, The Physics and Chemistry of the InterstellarMedium. Cambridge, UK: Cambridge University Press, 2005.

115

Page 137: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Bibliography

107 J. Ostrom, D. S. Bezrukov, G. Nyman, and M. Gustafsson, “Er-ratum: “Reaction rate constant for radiative association of CF+”[J. Chem. Phys. 144, 044302 (2016)],” The Journal of ChemicalPhysics, vol. 150, p. 249901, Jun 2019.

108R. K. Kathir, G. Nyman, and M. Gustafsson, “The rate constantfor formation of HCl through radiative association,” MonthlyNotices of the Royal Astronomical Society, vol. 470, pp. 3068–3070, Jun 2017.

109 P. Szabo and M. Gustafsson, “A surface-hopping method forsemiclassical calculations of cross sections for radiative associationwith electronic transitions,” The Journal of Chemical Physics,vol. 147, p. 094308, Sep 2017.

110T. Stoecklin, P. Halvick, H.-G. Yu, G. Nyman, and Y. Ellinger,“On the gas-phase formation of the HCO radical: accurate quan-tum study of the H+CO radiative association,” Monthly Noticesof the Royal Astronomical Society, vol. 475, pp. 2545–2552, Jan2018.

111M. Zamecnıkova, P. Soldan, M. Gustafsson, and G. Nyman,“Formation of CO+ by radiative association,” Monthly Noticesof the Royal Astronomical Society, vol. 489, pp. 2954–2960, Aug2019.

112M. Gustafsson and R. C. Forrey, “Semiclassical methods forcalculating radiative association rate constants for different ther-modynamic conditions: Application to formation of CO, CN, andSiN,” The Journal of Chemical Physics, vol. 150, p. 224301, Jun2019.

113D. Burdakova, G. Nyman, and T. Stoecklin, “Formation of Na-containing complex molecules in the gas phase in dense molecularclouds: quantum study of the Na+ + H2 and Na+ + D2 radiativeassociation step,” Monthly Notices of the Royal AstronomicalSociety, vol. 485, pp. 5874–5879, Mar 2019.

114M. Zamecnıkova, M. Gustafsson, G. Nyman, and P. Soldan,“Formation of CO+ by radiative association II,” Monthly Notices

114

Bibliography

of the Royal Astronomical Society, vol. 492, pp. 3794–3802, Jan2020.

115A. G. G. M. Tielens, The Physics and Chemistry of the InterstellarMedium. Cambridge, UK: Cambridge University Press, 2005.

115

Page 138: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Papers

117

Page 139: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

Papers

117

Page 140: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of
Page 141: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

I

Paper I

Formation of the HydroxylRadical by RadiativeAssociation

Reprinted with permission fromFormation of the Hydroxyl Radical by Radiative AssociationS. Karl-Mikael Svensson, Magnus Gustafsson, and Gunnar NymanThe Journal of Physical Chemistry A, 2015, 119 (50), pp 12263–12269,DOI: 10.1021/acs.jpca.5b06300.Copyright 2015 American Chemical Society.

119

I

Page 142: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

I

Formation of the Hydroxyl Radical by Radiative AssociationS. Karl-Mikael Svensson,† Magnus Gustafsson,‡ and Gunnar Nyman*,†

†Department of Chemistry and Molecular Biology, University of Gothenburg, SE-41296 Gothenburg, Sweden‡Applied Physics, Division of Material Science, Department of Engineering Science and Mathematics, Lulea University of Technology,SE-97187 Lulea, Sweden

ABSTRACT: The reaction rate constant for the radiative association of O(3P) and H(2S)has been calculated by combining a few different methods and taking account of bothdirect and resonance-mediated pathways. The latter includes both shape resonances andFeshbach type inverse predissociation. The reaction rate constant is expressed as afunction of temperature in the interval 10−30000 K. This reaction may be astrochemicallyrelevant and is expected to be of use in astrochemical networks.

INTRODUCTIONThe hydroxyl radical (OH) has been observed in the interstellarmedium both in the Milky Way1 and in other galaxies.2 OH isalso thought to have been an important species in the earlyuniverse, when heavy elements were more rare than in present-day molecular clouds.3

Due to the low density of the interstellar medium (for themost common element hydrogen it is usually <1 cm−3),4 three-body collisions essentially do not occur. This allows for two-body collisions with low reaction probabilities, such as radiativeassociation, to be important. The atoms needed to form thehydroxyl radical by this type of reaction are hydrogen (H) andoxygen (O), which respectively are the atoms of the most andthird most abundant elements in our galaxy and may haveatomic densities of approximately 0.354 and 0.0001 cm−3,respectively.5 Therefore, it is of interest to see if this type ofreaction could be a possible path for the formation of thehydroxyl radical.When the ground state atoms O(3P) and H(2S) collide, the

molecular states that are possible for them to form areOH(X2Π), OH(12Σ−), OH(a4Σ−), and OH(b4Π), whosepotential energies can be seen in Figure 1. The states X2Πand 12Σ− can radiate to give OH(X2Π) in a bound rovibrationalenergy level. Explicitly written, these reactions are (OH(X2Π)*denotes a free or quasibound state)

ν+ → Π * → Π + hO( P) H( S) OH(X ) OH(X )3 2 2 2(I)

and

ν+ → Σ → Π +− hO( P) H( S) OH(1 ) OH(X )3 2 2 2(II)

Alternatively, the two colliding atoms, in any of the formerlymentioned states, may undergo a radiationless transition to theexcited quasibound A2Σ+ state of OH from where a radiativetransition to X2Π can be made. This type of radiativeassociation reaction mechanism is called inverse predissociationand can occur via any of the pathways

ν

+ → Π * → Σ→ Π +

+

h

O( P) H( S) OH(X ) OH(A )

OH(X )

3 2 2 2

2 (III)

ν

+ → Σ → Σ→ Π +

− +

h

O( P) H( S) OH(1 ) OH(A )

OH(X )

3 2 2 2

2 (IV)

Special Issue: Dynamics of Molecular Collisions XXV: Fifty Years ofChemical Reaction Dynamics

Received: July 1, 2015Revised: September 3, 2015Published: September 9, 2015

Figure 1. Potential energies for the X2Π and 12Σ− states of OH fromvan Dishoeck et al.,6 the potential energies for the A2Σ+, a4Σ−, and b4Πfrom Yarkony,7 and a Morse potential for the potential energy of theX3Σ− state of OH+ (with parameters from van Lonkhuyzen and deLange8 and Huber and Herzberg9) placed one ionization energy ofhydrogen10 above OH. The points are the original data and the lines(except the Morse potential) are inter- and extrapolations.

Article

pubs.acs.org/JPCA

© 2015 American Chemical Society 12263 DOI: 10.1021/acs.jpca.5b06300J. Phys. Chem. A 2015, 119, 12263−12269

Dow

nloa

ded

via

CH

ALM

ERS

UN

IV O

F TE

CH

NO

LOG

Y o

n D

ecem

ber 9

, 201

9 at

15:

28:1

4 (U

TC).

See

http

s://p

ubs.a

cs.o

rg/s

harin

ggui

delin

es fo

r opt

ions

on

how

to le

gitim

atel

y sh

are

publ

ishe

d ar

ticle

s.

I

Page 143: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

I

ν

+ → Σ → Σ→ Π +

− +

h

O( P) H( S) OH(a ) OH(A )

OH(X )

3 2 4 2

2 (V)

or

ν

+ → Π → Σ→ Π +

+

h

O( P) H( S) OH(b ) OH(A )

OH(X )

3 2 4 2

2 (VI)

The rate constants for formation of OH through inversepredissociation III−VI have been estimated previously.11,12 Theradiative association reactions I and II have not been studiedprior to this work.The aim of this study is to compute a reliable reaction rate

constant for OH formation through radiative association thatcan be incorporated into chemical models of the interstellarmedium (through, e.g., KIDA13 and the UMIST database forastrochemistry14). This includes calculations of rate constantscorresponding to reactions I and II, and an improved estimateof the contribution from the inverse predissociation reactionsIII−VI. Attention is also paid to possible influence fromelectron detachment and charge transfer processes.

THEORETICAL METHODSFor a binary collisional process, such as radiative associationwhere the orbital electronic angular momentum quantumnumber changes from Λ to Λ′, an expression for the thermalrate constant is

∫πμσ= −

Λ→Λ′∞

Λ→Λ′⎛⎝⎜

⎞⎠⎟

⎛⎝⎜

⎞⎠⎟

k T

k TE E E k T E

( )

8 1( ) exp( / ) d

1/2

B

3/2

0B

(1)

where kΛ→Λ′(T) is the thermal rate constant, T is the absolutetemperature, μ is the reduced mass of the colliding particles, kBis Boltzmann’s constant, E is the kinetic energy of the collidingparticles, and σΛ→Λ′(E) is the cross-section for the process.There are a few different ways to calculate this cross-section.Four such approaches were used for reaction pathways I and IIin this work: a quantum mechanical perturbation theory, asemiclassical method, a classical method, and Breit−Wignertheory.Quantum Mechanical Perturbation Theory. The

perturbation theory employed here physically means that weapproximately account for the interaction between the electricfield of the photon produced in the reaction and the electricdipole moment of the molecule. When the quantummechanical perturbation theory is used, a Fermi Golden rulecross-section is obtained and can be written as (as described by,e.g., refs 15 and 16)

∑σ

ππ μ

ν= ℏϵ | |

Λ→Λ′

′ ′Λ Λ′ ′ ′ Λ →Λ′ ′ Λ Λ′ ′ ′

E

c EP S M

( )

323 (4 )

1

Jv JE v J J J EJ v J

5 2

03

3 2

(2)

where J and J′ are the total angular momentum quantumnumbers before and after the process, respectively, v′ is thevibrational quantum number for the bound state, ℏ is thereduced Planck constant, ϵ0 is the permittivity of vacuum, c isthe speed of light, PΛ is a factor for statistical weight, νEΛ′v′J′ isthe frequency of the emitted photon, SΛJ→Λ′J′ is the Honl−

London factor (rotational line intensity factor),17 and MΛEJΛ′v′J′is the matrix element of the transition dipole moment operator.The sum is over all J, v′, and J′. It is assumed in thesecalculations that the electronic spin quantum number (S) ismuch smaller than J, so that for the operators J − S ≈ J.MΛEJΛ′v′J′ is defined as

∫= ΨΛ Λ′ ′ ′∞

Λ ΛΛ′ Λ′ ′ ′M F R D R R R( ) ( ) ( ) dEJ v J EJ v J0 (3)

where R is the distance between the particles, FΛEJ(R) is thecollision-energy-normalized wave function of the unboundstate, ΨΛ′v′J′(R) is the normalized wave function of the boundstate, and DΛΛ′(R) is the electric dipole moment operator forthe transition between the electronic states of Λ and Λ′. PΛ isdefined as

δ= + −+ + + +Λ

ΛPS

L S L S

(2 1)(2 )

(2 1)(2 1)(2 1)(2 1)0,

A A B B (4)

where δ0,Λ is the Kronecker delta, LA and LB are the electronicorbital angular momentum quantum numbers of particles Aand B, respectively, and SA and SB are the electronic spinquantum numbers of particles A and B, respectively.

Semiclassical Method. The semiclassical method can beseen as the time integral of the transition rate between twoelectronic states, as it is given by the Einstein A coefficient.18

This can also be formulated as an integral over internuclearseparation, which is done here. This method only works fortransitions between different electronic states and has a cross-section according to the formula19,20

∫ ∫σ

π μ=− −

Λ→Λ′

Λ∞ ∞ Λ→Λ′

Λ

E

EP b

A RR b

( )

42

( )

1d d

R

Eb

V RE

bR

0 ( )c2

2 (5)

where b is the impact parameter, Rc is the outer classical turningpoint, VΛ(R) is the potential energy of the electronic statecharacterized by Λ, and AΛ→Λ′

Eb (R) is the probability oftransition between the states characterized by Λ and Λ′,respectively, per unit of time. AΛ→Λ′

Eb (R) is given by21

=< −

+ <Λ→Λ′

Λ→Λ′ Λ Λ′

Λ′

⎨⎪⎪

⎩⎪⎪

A R

A R E V R V R

V REbR

( )

( ) if ( ) ( ) and

( ) 0

0 otherwise

Eb 2

2

(6)

where VΛ′(R) is the potential energy of the electronic statecharacterized by Λ′, and AΛ→Λ′(R) is defined by

ππ

δδ

ν= ϵ−

− ′ ′

Λ→Λ′

Λ+Λ′

ΛΛ→Λ ΛΛ

⎛⎝⎜⎜

⎞⎠⎟⎟

A R

hcR D R

( )

643 (4 )

2

2( ) ( )

4

03

0,

0,

3 2

(7)

with

ν = −Λ→Λ′

Λ Λ′RV R V R

h( )

max(0, ( ) ( ))(8)

Classical Method. The classical method only applies totransitions within one electronic state and therefore comple-ments the semiclassical method. The method essentially takesthe frequency at which the electric dipole moment of the

The Journal of Physical Chemistry A Article

DOI: 10.1021/acs.jpca.5b06300J. Phys. Chem. A 2015, 119, 12263−12269

12264

I

Page 144: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

I

system changes, by using classical mechanics, and equates thatfrequency to that of the emitted photon. The intensity of theemission is given by the Larmor power22 and the radiativeassociation cross section is given by23

∫ ∫ ∫σ

πω ω= ℏ ϵ

ω

Λ→Λ

Λ∞ ∞

−∞

∞E

cP b b E t t bD

( )

43 (4 )

e ( , , ) d d dt3

0 0 0

3 i2

(9)

where ω is the angular frequency of the emitted photon, t istime, and D(b,E,t) is the time-dependent electric dipolemoment vector of the system. D(b,E,t) can be found fromDΛΛ(R) by solving the equations of motion,

μ= − −Λ

⎛⎝⎜

⎞⎠⎟

Rt

E V REbR

dd

2( )

2

2(10)

θμ

=t

bR

Edd

22

(11)

where θ is the polar angle of the system.Breit−Wigner Theory. Radiative association cross sections

will in many cases have peaks due to resonances and the shapeof these resonances will not always be correct when thequantum mechanical perturbation theory is used.24 Thesemiclassical and classical methods on the contrary cannottake resonances into account at all, but only give the directcontribution to the cross-section. Breit−Wigner theoryaccounts only for resonances and therefore can work as acomplement to the semiclassical or classical method and, whencombined with either of these, in some cases give better resultsthan the quantum mechanical perturbation theory.The Breit−Wigner theory assumes that the rate constant can

be divided into a direct contribution, kΛ→Λ′direct (T), and a

contribution from resonances, kΛ→Λ′res (T),

= +Λ→Λ′ Λ→Λ′ Λ→Λ′k T k T k T( ) ( ) ( )direct res(12)

The resonance contribution to the rate constant is given by25,26

πμ

= ℏ

× + Γ ΓΓ + Γ −

Λ→Λ′ Λ

Λ Λ →Λ′

Λ →Λ′ ΛΛ

⎛⎝⎜

⎞⎠⎟k T P

k T

J E k T

( )2

(2 1) exp( / )vJ

vJ vJ

vJ vJvJ

res 2

B

3/2

tun rad

rad tun B

(13)

where v is the vibrational quantum number of a quasiboundstate, EΛvJ is the energy of a quasibound state with quantumnumbers Λ, v, and J, ΓΛvJ→Λ′

rad is the total width of the radiativedecay to any of the bound levels, and ΓΛvJ

tun is the width oftunnelling out of the quasibound state. An alternate way toformulate eq 13 for calculating the speed of inversepredissociation is11

πμ

= ℏ

× + Γ ΓΓ −

Λ→Λ′ Λ

Λ →Λ″ Λ″ →Λ′

Λ″ →Λ′ ΛΛ″

⎛⎝⎜

⎞⎠⎟k T P

k T

J E k T

( )2

(2 1) exp( / )vJ

J vJ vJ

vJ JvJ

ip 2

B

3/2

pre rad

ortot B

(14)

where Λ″ is the orbital electronic angular momentum quantumnumber for the predissociatied (quasibound) state, ΓΛ″vJ→Λ′

rad is

the total width of the radiative transition from thepredissociated state with vibrational quantum number v androtational quantum number J to any of the bound states,ΓΛJ→Λ″vJpre is the width of escaping the quasibound state by

radiationless transitions to another electronic state, ΓΛ″vJ→Λ′orΛJtot

= ΓΛ″vJ→Λ′rad + ∑ΛΓΛJ→Λ″vJ

pre , and EΛ″vJ is the energy of thequasibound vibrational state with Λ″, v, and J.

Electron Detachment. A reaction that possibly couldcompete with the radiative association processes is electrondetachment,

+ → Σ → Σ +− + − −O( P) H( S) OH(1 ) OH (X ) e3 2 2 3

(VII)

+ → Σ → Σ +− + − −O( P) H( S) OH(a ) OH (X ) e3 2 4 3

(VIII)

and

+ → Π → Σ ++ − −O( P) H( S) OH(b ) OH (X ) e3 2 4 3

(IX)

A method to approximate an upper limit to the cross-section ofelectron detachment can be found in an experimental study onO− + H.27 In this approximation it is assumed that in all caseswhen the classical trajectory reaches a point in space where thepotential energy surface of the reactants cross the potentialenergy surface of the ground state of the products, the reactionoccurs. According to this approximation the cross-section,σdetachment(E), is

∑σ π= −Λ

Λ Λ× Λ Λ

×⎡⎣⎢

⎤⎦⎥E P R

V RE

( ) ( ) 1( )

detachment2

(15)

where RΛ× is the distance at which the potential energy surface

of the reactant state characterized by Λ and the product statecross. The sum is over all relevant Λ.

Charge Transfer. Apart from electron detachment, anotherreaction that could compete with the radiative association ischarge transfer:

+ → ++ −O H O H (X)

and

+ → +− +O H O H (XI)

The threshold energies for these reactions are 12.863880 and12.13736 eV, respectively (using electron affinities from Lykkeet al.28 and Neumark et al.29 and ionization energies fromKramida et al.10).

Computational Details. For the calculations of the crosssections of reactions I, II, and VII−IX the necessary data weretaken from the literature. Potential energy data points for theX2Π and 12Σ− states and the electrical transition dipolemoment between X2Π and 12Σ− were taken from van Dishoecket al.6 The electrical dipole moment of the X2Π state was takenfrom Langhoff et al.30 These data points were splined andextrapolated. At long distances the potentials were extrapolatedas ∝1/R6. At short distances the potential energy of the X2Πstate was extrapolated as a Morse potential with equilibriumdistance and dissociation energy from Huber and Herzberg.9

The potential energy of the 12Σ− state was at short distancesextrapolated as an exponential function.The dipole moments were extrapolated as exponential

functions going toward zero at long distances and as second-order polynomials going toward zero at short distances. Theforms for the extrapolations toward short internuclear distances

The Journal of Physical Chemistry A Article

DOI: 10.1021/acs.jpca.5b06300J. Phys. Chem. A 2015, 119, 12263−12269

12265

I

Page 145: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

I

were chosen to be simple forms that behave qualitativelycorrect. Different forms were tested, and it was found that theextrapolation was not important for the resulting reaction rateconstant.31

The potential energy of the X3Σ− state of OH+ wasapproximated by a Morse potential (V(R) = De1 − exp[a(R− Re)]

2) with equilibrium distance (Re) and dissociationenergy (De) from van Lonkhuyzen and de Lange8 and a fromHuber and Herzberg.9 The difference in asymptotic energybetween X2Π of OH and X3Σ− of OH+ was taken to be theionization energy of hydrogen.10 All the potential energies anddipole moments used can be seen in Figures 1 and 2.

The widths of the inverse predissociation (reactions III, IV,V, and VI) were gathered from many different sources.Vibrational quantum numbers up to 4 and rotational quantumnumbers up to a maximum of 30 were considered. Total,

radiative, and predissociation lifetimes were mostly taken fromYarkony.7 Total lifetimes not available in that article were takenfrom Brzozowski et al.33 and, where available, predissociationlifetimes were taken from Parlant and Yarkony34 instead. Theselifetimes do not take account of nonradiative coupling betweenthe continuum of X2Π and A2Σ+, so the predissociation widthsfrom Julienne and Krauss11 taking account of only that couplingwere added, and for v = 1 they were interpolated. Thecontributions from different rovibrational states to thepredissociation widths for vibrational quantum numbers 3and 4 were taken from the golden rule calculations forrotational quantum numbers 0 and 14 in Parlant and Yarkony34

and interpolated for the rotational quantum numbers inbetween.Radiative widths were interpolated and a few widths were

found by the use of ΓΛ″vJ→Λ′orΛJtot = ΓΛ″vJ→Λ′

rad + ∑ΛΓΛJ→Λ″vJpre .

Energies of the quasibound states were calculated with thespectroscopic constants from Luque and Crosley.35

The cross sections for reactions I and II were calculated withthe quantum mechanical perturbation theory. The bound wavefunctions were calculated with discrete variable representation36

and the unbound wave functions were calculated withNumerov’s method.37

The cross-section for reaction I was also calculated bycombining the classical method with Breit−Wigner theory. Theouter classical turning point, Rc, was found by the bisectionmethod, the range of time was found by Romberg integration38

of dt/dR from Rc to an outer distance limit. dR/dt and dθ/dtwere integrated with the fourth-order Runge−Kutta method,38

the Fourier transforms were done by fast Fourier transform,38

the integral over ω was made with Simpson’s 1/3 rule,38 andthe integral over b was evaluated with the trapezoidal rule. Thecross-section and reaction rate constant for the resonancecontribution were calculated by using quasibound stateparameters obtained with the program LEVEL, developed byLeRoy.39

The cross-section for reaction II was additionally calculatedby the semiclassical method. The bisection method was used to

Figure 2. Electric dipole moment of the X2Π state of OH fromLanghoff et al.30 and the electrical transition dipole moment betweenthe X2Π and 12Σ− states of OH from van Dishoeck et al.6 e is theelemental charge, and a0 is the Bohr radius. The points are the originaldata and the lines are inter- and extrapolations.

Table 1. Honl−London Factors SΛJΩϵ→Λ′J′Ω′ϵ′ Calculated by the Method of Watson17 for Hund’s Case aa

2S+1Λ Ω ϵ 2S+1Λ′ Ω′ ϵ′ J′ = J − 1 J′ = J J′ = J + 1

1Π 1 ±1 1Π 1 ±1 ≈+ − JJ JJ

( 1)( 1)0 ≈+

+ JJ JJ( 2)

1

1Π 1 ±1 1Π 1 ∓1 0 ≈++ 0J

J J2 1( 1) 0

1Σ− 0 −1 1Π 1 1 0 2J+1≈2J 01Σ− 0 −1 1Π 1 −1 J−1≈J 0 J+2≈J2Π 1/2 ±1 2Π 1/2 ±1 ≈+ − JJ J

J( 1 / 2)( 1 / 2)

0 ≈+ ++ JJ J

J( 3 / 2)( 1 / 2)

1

2Π 1/2 ±1 2Π 1/2 ∓1 0 ≈++ 0J

J J1 / 2

2 ( 1) 0

2Π 3/2 ±1 2Π 3/2 ±1 ≈+ − JJ JJ

( 3 / 2)( 3 / 2)0 ≈+ −

+ JJ JJ

( 5 / 2)( 1 / 2)1

2Π 3/2 ±1 2Π 3/2 ∓1 0 ≈++ 0J

J J9( 1 / 2)2 ( 1)

0

2Σ− 1/2 ±1 2Π 1/2 ±1 ≈+ −J JJ

J( 1 / 2)( 1 / 2)2 2 0 ≈+ +

+J J

JJ( 3 / 2)( 1 / 2)

2( 1) 2

2Σ− 1/2 ±1 2Π 1/2 ∓1 0 ≈++ JJ

J J( 1 / 2)

( 1)

30

2Σ− 1/2 ±1 2Π 3/2 ±1 ≈− −J JJ

J( 1 / 2)( 3 / 2)2 2 0 ≈+ +

+J J

JJ( 5 / 2)( 3 / 2)

2( 1) 2

2Σ− 1/2 ±1 2Π 3/2 ∓1 0 ≈+ + −+ JJ J J

J J( 3 / 2)( 1 / 2)( 1 / 2)

( 1)0

aThe approximations are valid for large J. ϵ is parity and Ω is the quantum number of the projection of the total electronic angular momentum onthe internuclear axis.

The Journal of Physical Chemistry A Article

DOI: 10.1021/acs.jpca.5b06300J. Phys. Chem. A 2015, 119, 12263−12269

12266

I

Page 146: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

I

find the classical turning point, the integral over R wasevaluated with the Romberg method,38 and the integral over bwas calculated with the trapezoidal rule.The direct (nonresonant) contribution to the rate constants

acquired by integration of the cross-section over the energy (eq1) was calculated with the trapezoidal rule.The programs used here are available with Honl−London

factors for singlets. Looking at Table 1, we can see that for thetransitions relevant here this is a good approximation for large J,except for Σ → Π transitions where an extra factor of 0.5appears. This factor is, however, accounted for by summationover Ω′ϵ′. The summation over Ωϵ is not done explicitly in thecalculations of the cross sections, but it is instead taken care ofby multiplication by the factor PΛ. In Table 2 the Honl-Londonfactors and PΛ used in the calculations can be seen.

When the cross sections were calculated, they wereconsidered converged when the corresponding reaction rateconstant at all considered temperatures had converged towithin 0.5% with respect to the individual computationalparameters.

RESULTS AND DISCUSSIONThe calculated cross sections can be seen in Figure 3. For theX2Π → X2Π pathway the quantum mechanical perturbation

theory and the combined classical/Breit−Wigner methods forcalculating the cross-section result in very similar baselines,apart from at high and low energies. With the Breit−Wignertheory we have, however, found many more resonances thanwith the quantum mechanical perturbation theory. This is likely

to be due to the energy resolution of the quantum calculation,but for the purpose of this work the comparison is satisfactory.For the 12Σ− → X2Π pathway the semiclassical and quantum

mechanical cross sections are almost identical for a large rangeof energies. The semiclassical program could not calculate crosssections at as high energies as the quantum mechanical programdue to limitations in the semiclassical computational procedure.The semiclassical cross-section will, however, approach zerovery quickly for the high energies where it is not plotted, due tothe Franck−Condon principle.40,41 It can be seen in Figure 4

that the semiclassical and quantum mechanical rate constantsare nearly indistinguishable, so the differences in cross-sectionat high energies do not affect the rate constant at the studiedtemperatures. Further the electron detachment reaction onlymatters at energies above 9 eV. Also because the charge transferreactions X and XI have threshold energies of >12 eV, which isnoticeably higher than the energy where the electrondetachment starts to dominate, they are not important forour study.At most temperatures we assume a4Σ− to be the most

important collision state for the inverse predissociationprocess,7 and this electronic state is not relevant for the otherradiative association pathways, meaning that even if reactions Iand III, or II and IV would interfere with each other, it wouldnot affect the total reaction rate noticeably for thosetemperatures. For low temperatures both reactions I and IIImake noticeable contributions to the total reaction rate.Because both these reactions occur via the state X2Π theycould possibly affect one another. However, the interferencebetween the processes will most likely not give a noticeabledifference, because none of the resonances in the differentpathways overlap.The total reaction rate constant is plotted versus temperature

in Figure 5, where the importance of the inverse predissociationis seen. The previous estimation for the rate constant forinverse predissociation12 was obtained up to 400 K and stayedbelow 3 × 10−20 cm3 s−1 at those temperatures. The parametersobtained by fitting the total rate constant to the Kooijformula,42 which is the form on which this kind of rateconstants are often presented in databases such as KIDA, canbe seen in Table 3. The fit is within 5% of the rate constant. Asshown in Figure 5, the rate constant already present in, for

Table 2. Honl−London Factors SΛJ→Λ′J′ and StatisticalWeight Factors PΛ Used for the Calculations of the RateConstant for Reactions I and II

SΛ J→Λ′J−1 SΛ J→Λ′J SΛ J→Λ′J+1 PΛ

X2Π → X2Π + −J JJ

( 1)( 1) ++

JJ J2 1( 1)

++

J JJ( 2)

14/18

12Σ− → X2Π J−1 2J+1 J+2 2/18

Figure 3. Reaction cross sections for reaction I calculated by thequantum mechanical method (QM) and the classical method withBreit−Wigner theory added (CL+BW), for reaction II calculated bythe quantum mechanical method (QM) and the semiclassical method(SC), and for reactions VII−IX by the use of eq 15.

Figure 4. Reaction rate constants for reaction I calculated with theclassical method and Breit−Wigner theory (CL+BW), for reaction IIcalculated with the golden rule (QM) and semiclassical method (SC),and for both reactions summed (Total).

The Journal of Physical Chemistry A Article

DOI: 10.1021/acs.jpca.5b06300J. Phys. Chem. A 2015, 119, 12263−12269

12267

I

Page 147: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

I

example, KIDA is larger than the one calculated in this study.43

The rate constant calculated in the present work should be animprovement over the old one, because this new rate constantincludes all the reactions I−VI and has values for a largertemperature interval.

CONCLUSIONS

The reaction rate constant for the radiative association ofO(3P) and H(2S) into OH(X2Π) has been calculated for thetemperature range 10−30000 K. The calculation includes directand shape-resonance-mediated radiative association, as well asinverse predissociation, and has been conducted by using a fewdifferent methods for the different pathways of the reaction.This rate constant is expected to be interesting for inclusion inastrochemical databases, because it has a much largertemperature interval than the rate constant already availablein for example KIDA13 and the UMIST database forastrochemistry.14

AUTHOR INFORMATION

Corresponding Author*G. Nyman. E-mail: [email protected]. Phone: +46 (0)317869035.

NotesThe authors declare no competing financial interest.

ACKNOWLEDGMENTS

Support from the Swedish research council and the COSTActions CM1401 “Our astrochemical history” and CM1405“Molecules in Motion (MOLIM)” is gratefully acknowledged.

REFERENCES(1) Weinreb, S.; Barret, A.; Meeks, M.; Henry, J. Radio Observationsof OH in the Interstellar Medium. Nature 1963, 200, 829−831.(2) Weliachew, L. Detection of Interstellar OH in Two ExternalGalaxies. Astrophys. J. 1971, 167, L47−L52.(3) Sternberg, A.; Dalgarno, A.; Pei, Y.; Herbst, E. In Conditions andImpact of Star Formation; Rollig, M., Simon, R., Ossenkopf, V., Stutzki,J., Eds.; EAS Publications Series, 52; European Astronomical Society:Geneva, 2011; pp 43−46.(4) Bohlin, R. C.; Savage, B. D.; Drake, J. F. A Survey of InterstellarH I from Lα Absorption Measurments. II. Astrophys. J. 1978, 224,132−142.(5) Meyer, D. M.; Jura, M.; Cardelli, J. A. The Definite Abundance ofInterstellar Oxygen. Astrophys. J. 1998, 493, 222−229.(6) van Dishoeck, E. F.; Langhoff, S. R.; Dalgarno, A. The Low-Lying

2Σ− States of OH. J. Chem. Phys. 1983, 78, 4552−4561.(7) Yarkony, D. A Theoretical Treatment of the Predissociation ofthe Individual Rovibronic Levels of OH/OD(A2Σ+). J. Chem. Phys.1992, 97, 1838−1849.(8) van Lonkhuyzen, H.; de Lange, C. A. U. V. PhotoelectronSpectroscopy of OH and OD Radicals. Mol. Phys. 1984, 51, 551−568.(9) Huber, K. P.; Herzberg, G. Constants of Diatomic Molecules;Molecular Spectra and Molecular Structure; Van Nostrand ReinholdCo.: New York, 1979; Vol. 4.(10) Kramida, A.; Ralchenko, Y.; Reader, J.; Team, N. A. NISTAtomic Spectra Database, ver. 5.2. http://physics.nist.gov/asd(accessed 150309), 2014; National Institute of Standards andTechnology, Gaithersburg, MD.(11) Julienne, P. S.; Krauss, M. In Molecules in the GalacticEnvironment; Gordon, M. A., Snyder, L. E., Eds.; John Wiley &Sons: New York, 1973; pp 353−373.(12) Julienne, P. S.; Krauss, M.; Donn, B. Formation of OH ThroughInverse Predissociation. Astrophys. J. 1971, 170, 65−70.(13) Wakelam, V.; Herbst, E.; Loison, J.-C.; Smith, I. W. M.;Chandrasekaran, V.; Pavone, B.; Adams, N. G.; Bacchus-Montabonel,M.-C.; Bergeat, A.; Beroff, K.; et al. A Kinetic Database forAstrochemistry (KIDA). Astrophys. J. Suppl. Ser. 2012, 199, 21-1−21-10 (the actual database can be found at the URL http://kida.obs.u-bordeaux1.fr).(14) McElroy, D.; Walsh, C.; Markwick, A. J.; Cordiner, M. A.;Smith, K.; Millar, T. J. The UMIST Database for Astrochemistry 2012.Astron. Astrophys. 2013, 550, A36-1−A36-13 (the actual database canbe found at the URL http://www.udfa.net/).(15) Babb, J. F.; Dalgarno, A. Radiative Association and InversePredissociation of Oxygen Atoms. Phys. Rev. A: At., Mol., Opt. Phys.1995, 51, 3021−3026.(16) Barinovs, G.; van Hemert, M. C. CH+ Radiative Association.Astrophys. J. 2006, 636, 923−926.(17) Watson, J. K. G. Honl-London Factors for Multiplet Transitionsin Hund’s Case a or b. J. Mol. Spectrosc. 2008, 252, 5−8.(18) Julienne, P. S. Theory of Rare Gas-Group VI 1 S-1 D Collision-Induced Transitions. J. Chem. Phys. 1978, 68, 32−41.(19) Bates, D. R. Rate of Formation of Molecules by RadiativeAssociation. Mon. Not. R. Astron. Soc. 1951, 111, 303−314.(20) Zygelman, B.; Dalgarno, A. Radiative Quenching of He(2 1S)Induced by Collisions with Ground-State Helium Atoms. Phys. Rev. A:At., Mol., Opt. Phys. 1988, 38, 1877−1884.(21) Gustafsson, M.; Antipov, S. V.; Franz, J.; Nyman, G. RefinedTheoretical Study of Radiative Association: Cross Sections and RateConstants for the Formation of SiN. J. Chem. Phys. 2012, 137, 104301.(22) Jackson, J. D. Classical Electrodynamics; John Wiley and SonsLtd.: New York, 1962.

Figure 5. Reaction rate constant for the combination of reactions I andII, for the combination of reactions III−VI, and for the total of allthese reactions. Also an older total reaction rate constant, taken fromKIDA,13 is shown.

Table 3. Reaction Rate Constants Fitted to the KooijFormula42 [Written as kΛ→Λ′(T) = α(T/300 K)β exp(−γ/T)]in a Few Different Temperature Intervalsa

temp interval (K) α (10−20 cm−3 s−1) β γ (K)

10−20 249551 4.90741 −44.348520−100 18.4379 −0.366037 50.6067100−500 22.7637 0.259761 4.71068500−2000 2.17451 1.72564 −776.0512000−5000 262.168 −0.0290599 2184.055000−15000 0.00646257 2.78182 −11183.415000−30000 85814.4 −0.593255 37216

aThe fit is within 5% of the original reaction rate constant.

The Journal of Physical Chemistry A Article

DOI: 10.1021/acs.jpca.5b06300J. Phys. Chem. A 2015, 119, 12263−12269

12268

I

Page 148: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

I

(23) Gustafsson, M. Classical Calculations of Radiative Association inAbsence of Electronic Transitions. J. Chem. Phys. 2013, 138, 074308.(24) Bennett, O.; Dickinson, A.; Leininger, T.; Gadea, F. RadiativeAssociation in Li+H Revisited: The Role of Quasi-Bound States. Mon.Not. R. Astron. Soc. 2003, 341, 361−368.(25) Bain, R. A.; Bardsley, J. N. Shape Resonances in Atom-AtomCollisions - I. Radiative Association. J. Phys. B: At. Mol. Phys. 1972, 5,277−285.(26) Breit, G.; Wigner, E. Capture of Slow Neutrons. Phys. Rev. 1936,49, 519−531.(27) Fedchak, J. A.; Huels, M. A.; Doverspike, L. D.; Champion, R. L.Electron Detachment and Charge Transfer for Collisions of O− and S−

with H. Phys. Rev. A: At., Mol., Opt. Phys. 1993, 47, 3796−3800.(28) Lykke, K. R.; Murray, K. K.; Lineberger, W. C. ThresholdPhotodetachment of H−. Phys. Rev. A: At., Mol., Opt. Phys. 1991, 43,6104−6107.(29) Neumark, D. M.; Lykke, K. R.; Andersen, T.; Lineberger, W. C.Laser Photodetachment Measurment of the Electron Affinity ofAtomic Oxygen. Phys. Rev. A: At., Mol., Opt. Phys. 1985, 32, 1890−1892.(30) Langhoff, S. R.; Bauschlicher, C. W.; Taylor, P. R. TheoreticalStudy of the Dipole Moment Function of OH(X2Π. J. Chem. Phys.1989, 91, 5953−5959.(31) It has come to our attention that there are more recentcalculations of the potential energies and transition dipole momentsfor OH than those used here.32 Most notably these have values forshorter internuclear distances than those used in the present study,meaning that we could have extrapolated less. However, because theforms of the extrapolations were not important for the reaction rateconstants, the use of the newer and better potentials and transitiondipole moments would most likely not have changed the results of thisstudy noticeably.(32) van der Loo, M. P. J.; Groenenboom, G. C. Ab InitioCalculation of (2 + 1) Resonance Enhanced Multiphoton IonizationSpectra and Lifetimes of the (D,3) 2Σ− States of OH and OD. J. Chem.Phys. 2005, 123, 074310.(33) Brzozowski, J.; Erman, P.; Lyyra, M. Precision Estimates of thePredissociation Rates of the OH A2Σ State (v ≤ 2). Phys. Scr. 1978, 17,507−511.(34) Parlant, G.; Yarkony, D. A Theoretical Analysis of the State-Specific Decomposition of OH(A2Σ+,v′,N′,F1/F2) Levels, Includingthe Effects of Spin-Orbit and Coriolis Interactions. J. Chem. Phys.1999, 110, 363−376.(35) Luque, J.; Crosley, D. Transition Probabilities in the A2Σ+ −X2Πi Electronic System of OH. J. Chem. Phys. 1998, 109, 439−448.(36) Colbert, D. T.; Miller, W. H. A Novel Discrete VariableRepresentation for Quantum Mechanical Reactive Scattering via theSmatrix Kohn Method. J. Chem. Phys. 1992, 96, 1982−1991.(37) Noumerov, B. A Method of Extrapolation of Perturbations.Mon. Not. R. Astron. Soc. 1924, 84, 592−602.(38) Press, W. H.; Teukolsky, S. A.; Vetterling, W. T.; Flannery, B. P.Numerical Recipes in FORTRAN - The Art of Scientific Computing, 2nded.; Cambridge University Press: Cambridge, U.K., 1992.(39) Roy, R. J. L. LEVEL 8.0 A Computer Program for Solving theRadial Schrodinger Equation for Bound and Quasibound Levels.University of Waterloo Chemical Physics Research Report CP-663,2007; Including corrections as of December 2012.(40) Franck, J.; Dymond, E. G. Elementary Processes of Photo-chemical Reactions. Trans. Faraday Soc. 1926, 21, 536−542.(41) Condon, E. U. Nuclear Motions Associated with ElectronTransitions in Diatomic Molecules. Phys. Rev. 1928, 32, 858−872.(42) Kooij, D. M. Uber die Zersetzung des GasformigenPhosphorwasserstoffs. Z. Phys. Chem. 1893, 12, 155−161.(43) We have unsuccessfully tried to trace the original source of therate constant available in KIDA.

The Journal of Physical Chemistry A Article

DOI: 10.1021/acs.jpca.5b06300J. Phys. Chem. A 2015, 119, 12263−12269

12269

I

Page 149: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II

Paper II

Dynamics of GaussianWigner functions derivedfrom a time-dependentvariational principle

Reprinted from Dynamics of Gaussian Wigner functions derivedfrom a time-dependent variational principleJens Aage Poulsen, S. Karl-Mikael Svensson, and Gunnar NymanAIP Advances, 2017, 7 (11), pp 115018-1–115018-12,DOI: 10.1063/1.5004757.

Creative Commons Attribution license∗

∗https://creativecommons.org/licenses/by/4.0/

127

II

Page 150: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II

AIP ADVANCES 7, 115018 (2017)

Dynamics of Gaussian Wigner functions derivedfrom a time-dependent variational principle

Jens Aage Poulsen,a S. Karl-Mikael Svensson,b and Gunnar Nymanc

Department of Chemistry and Molecular Biology, University of Gothenburg,41296 Gothenburg, Sweden

(Received 15 September 2017; accepted 21 October 2017; published online 16 November 2017)

By using a time-dependent variational principle formulated for Wigner phase-space functions, we obtain the optimal time-evolution for two classes of Gaus-sian Wigner functions, namely those of either thawed real-valued or frozen butcomplex Gaussians. It is shown that tunneling effects are approximately includedin both schemes. © 2017 Author(s). All article content, except where other-wise noted, is licensed under a Creative Commons Attribution (CC BY) license(http://creativecommons.org/licenses/by/4.0/). https://doi.org/10.1063/1.5004757

I. INTRODUCTION

Many times, the correct description of processes in molecular systems requires inclusion of quan-tum statistical effects. An obvious example is that of chemical reactions in small molecular systemsthat may be treated accurately (see e.g. Refs. 1 and 2). However, for processes in much larger sys-tems, quantum effects still may play an important role, but here one must resort to more approximatemethods. Molecular diffusion at low temperature3 is one such example. For such problems, manyapproximate schemes have been put forth. Among these we mention the family of path integral-basedapproches such as Ring Polymer Molecular Dynamics,4 Centroid Molecular Dynamics5 and PathIntegral Liouville dynamics.6 Also, various implementations of the so-called Classical Wigner modelor Linearized Semi-Classical Initial Value Representation7–10 have been quite successful. Commonfor all these methods is that they are exact in the high temperature limit, but also that they are unable toaccount for some genuine quantum effects, such as tunneling through barriers. The interested readermay consult e.g. Ref. 10 where some of these methods were tested on model problems. In this articlewe consider a computational method, which is also aimed at modeling dynamical quantum effectsin many-body systems. It adopts a variational principle for minimizing the error between exact andapproximate quantum dynamics. The variational principle is put to work in Wigners phase-spaceformulation of quantum mechanics.11,12

The Wigner phase-space formulation of quantum mechanics12 represents an alternative topath integral methods. This formulation has a mathematical structure which closely resembles thatfound in classical statistical mechanics. Indeed, operator averages may instead be computed usingclassical-like phase-space averages over so-called Wigner distribution functions, and the dynam-ics of Wigner functions follow equations of motion which are extensions of those of Newton.Given these desirable properties, a large body of litterature on Wigners formulation of quantummechanics has been published, and we refer to the famous articles collected in the textbook byZachos et al.12

In Wigners formulation, the Wigner function representation of the density operator plays a centralrole:

[ ρ]W (q, p; t)≡∫ ∞−∞

dη e−ipη/⟨q +η

2| ρ(t)| q − η

2

⟩. (1)

[email protected]@[email protected]

2158-3226/2017/7(11)/115018/12 7, 115018-1 © Author(s) 2017

II

Page 151: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II

115018-2 Poulsen, Svensson, and Nyman AIP Advances 7, 115018 (2017)

The density operator may be time-dependent and its Wigner distribution evolves according tothe Wigner-Moyal equation:12

ddt

[ ρ]W (q, p; t)=L[ ρ]W (q, p; t), (2)

where the Liouvillian is given by

L =− pm∂

∂q+ V (q)

∂p−

2

24V (q)

∂3

∂p3+ .. (3)

This evolution equation is general and applies to all time-dependent Wigner functions. Also,the important problem of computing quantum correlation functions can be formulated in Wignerphase-space. To see this, we use the identity,

⟨A(t)B(0)

⟩=

1Z

Tr(eiHt/e−H/kBT A e−iHt/ B

)

=1

Z 2π

∫ ∞−∞

dpdq [e−H/kBT A]W (q, p; t)[B]W (q, p). (4)

Thus the calculation of correlation functions is equivalent to the problem of evolving the moregeneral functions [e−H/kBT A]W (q, p; t) in Wigner phase-space. In Ref. 13 we derived a time-dependentvariational principle that enables one to propagate a general parameter-dependent Wigner-function inan optimal way. This principle can be considered as a generalization of the Dirac-Frenkel variationalprinciple for wavefunctions.14

In this paper we consider two different classes of Wigner functions that are parameter-dependent.Using the variational principle, we derive the equations of motion for their variational parameters andapply the dynamical theory for calculating various time-dependent expectation values for a particlein a quartic and a double well potential.

This paper is organized as follows: An introduction to the variational principle is presented inSection II. Sections III and IV apply the theory for calculating the time propagation of two differentforms of Gaussian Wigner functions. Various applications of the theory and comparison to accurateresults are also presented in these Sections. The conclusions are presented in Section V.

II. FORMULATION OF VARIATIONAL PRINCIPLE

Consider the propagation of a Wigner function through Eq.2. We will assume that the Wignerfunction is time-dependent through a vector of parameters θ. Hence, our problem is to find the optimumtime evolution of θ(t). This was the problem we solved in Ref. 13 by utilizing the time-dependentvariational principle. The principle can be explained by considering the functional

J[θ, θ]=∫ t

0ds

⟨ρ(q, p; θ(s)), L − ∂

∂sρ(q, p; θ(s))

=

∫ t

0ds⟨ρ(q, p; θ(s)), L − θ · ∇θ ρ(q, p; θ(s))

⟩. (5)

Here, L is the Liouvillian and the general scalar-product between two Wigner functions has beenintroduced:

AW (q, p; t), BW (q, p; t)=∫ ∫

dqdp2π

A∗W (q, p; t)BW (q, p; t). (6)

The time integration in Eq.5 goes from s = 0 to s = t, which is the time-window of interest. The

action principle states that J[θ, θ] is stationary with respect to complex variations δθ(s) (that vanishat end-points) if and only if θ(s) is the optimum “path” in a least squares sense, i.e. so that the error

Error =

⟨L − ∂

∂sρ(q, p; θ(s)), L − ∂

∂sρ(q, p; θ(s))

⟩(7)

is minimized along the path. To find the path that extremizes the functional, we use the Euler-Lagrangeequations:

∂s ∂∂θiJ[θ, θ]=

∂θiJ[θ, θ], (8)

II

Page 152: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II

115018-3 Poulsen, Svensson, and Nyman AIP Advances 7, 115018 (2017)

where θi is the time-derivative of the ith entry of the θ vector, θi etc. We notice that the Wigner function

should be an analytic - differentiable - function of its complex variational parameters, otherwise, theequivalence between Eq.5 and Eq.7 does not hold.

III. VARIATIONAL PRINCIPLE FOR THAWED REAL-VALUED GAUSSIANS

A. Equations of motion

In this Section we consider the following ansatz for the variational “Wigner packet”:

ρ(q, p; t)= ρ(q, p; q0(t), p0(t), α(t), γ(t), β(t))

=N(t) exp−(q − q0(t)p − p0(t)

)T(α(t) γ(t)γ(t) β(t)

)(q − q0(t)p − p0(t)

), (9)

where N(t) is a normalization constant. Thus, the variational parameters are α(t), β(t), γ(t), q0(t) andp0(t). We will be interested in real-valued parameter “trajectories” θ(s) and, as we will see, a real-valued initial set of parameters stay real for all times. By substituting Eq.9 into Eq.5 and evaluatingthe Euler-Lagrange equations, we derive the equation of motions for the five parameters. We referthe interested reader to the supplementary material. Here, we simply state the result:

q0(t)

p0(t)

α(t)

γ(t)

β(t)

=

p0(t)/M

− ddq0

Vsm(q0(t))

2γ(t) ×MΩ2(q0(t))

β(t) ×MΩ2(q0(t)) − α(t)M

−2γ(t)M

, (10)

where the effective frequency Ω(q0) has been defined by:

MΩ2(q0)=d2

dq20

Vsm(q0). (11)

In these equations, the time-dependent smeared potential

Vsm(q0)=1√

π2 β(t)

∫ +∞

−∞dyV (y) exp

(− (q0 − y)2

β(t)2

), (12)

where

β(t)=1

22

β(t)

α(t)β(t) − γ(t)2+

12β(t), (13)

has appeared. As is easily shown, the product α(t)β(t) γ(t)2 is a constant of motion. In the specialcase of a minimum uncertainty Wigner-function (which corresponds to a wave-function), we haveα(t)β(t) γ(t)2 = 1/2 and then β

(t) = β(t). Also, the above solution is exact for quadratic potentials.

We notice that the Dirac-Frenkel principle has been applied for complex Gaussian wavefunctionsin Ref. 16 in the context of photodissociation spectroscopy. Also, see Ref. 17. In the case, α(t)β(t) γ(t)2 = 1/2, these methods and our are equivalent.

B. Kubo position auto correlation function

Here we show that the thawed Gaussians via Eq.9 can be used for calculating so-called Kubotransformed thermal position auto correlation functions. Because the variational principle is formu-lated in Wigner phase-space we can solve quantum-statistical problems with no reference to wavefunctions. As shown in Appendix A, by adopting the so-called Feynman-Kleinert approximation to

II

Page 153: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II

115018-4 Poulsen, Svensson, and Nyman AIP Advances 7, 115018 (2017)

the Boltzmann-operator, one can derive an approximate expression for the Kubo transformed positionauto correlation function:

xx(t)Kubo =1

ZFK

∫dxcxc exp(−W1(xc)/kBT )

× 1

(2π)2

∫ ∫dqdpq exp(tL)2

√π2MkBT tanh( Ω(xc)

2kBT )

α(xc)

× exp(−MΩ(xc)α(xc)

(q − xc)2) exp(−tanh( Ω(xc)

2kBT )

MΩ(xc)p2). (14)

This expression is exact to the extent that the Feynman-Kleinert approximation is. The equation isevaluated by Monte Carlo sampling using the weight function derived from the centroid potential:exp(W1(xc)/kBT ). Clearly, the problem is to propagate the normalized Gaussian Wigner function:

ρxc (q, p)= 2

√tanh( Ω(xc)

2kBT )

α(xc)exp(−MΩ(xc)

α(xc)(q − xc)2) exp(−

tanh( Ω(xc)2kBT )

MΩ(xc)p2), (15)

which is similar to Eq.9 and so can be propagated using Eq.10.

C. Application to double well potential

We first test the method against a model problem where tunneling dynamics is important. Weconsider the double well potential V (x) = 0.5x2 + 0.1x4. We adopt natural units: = kB = m = 1.First, we will consider a pure state, i.e. the propagation of a Wigner function corresponding to awave-function. The initial Wigner function or wave function is chosen as follows. We approximatethe ground state wave function as a sum of two Gaussian wave functions:

ψ(x)=N[exp(−λ2

(x − x0)2) + exp(−λ2

(x + x0)2)]. (16)

By minimizing the energy of this wave function with respect to λ and x0 we obtain λ = 1.23 andx0 = 1.12. We next consider the intial Wigner function to be equivalent to one half of the abovewavefunction, i.e. the Gaussian in the right well. Its Wigner function thus becomes as in Eq.9 withα = 1.23, β = 1/1.23, γ = 0, q0 = 1.12 and p0 = 0. The time-dependent effective potential is

Vsm(x)=−0.5(x2 +12β(t)2) + 0.1x4 + 0.3β(t)2x2 + 0.1

34β(t)24. (17)

In Figure 1, the time-dependent average position of the Wigner function based on the variationalmethod is shown. Also shown are accurate results as well as the predictions of the so-called classical

FIG. 1. Average position for a particle in the double well starting from a single Gaussian using q0 = 1.12, p0 = γ(0) = 0,α(0) = 1.23 and β(0) = 1/1.23.

II

Page 154: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II

115018-5 Poulsen, Svensson, and Nyman AIP Advances 7, 115018 (2017)

FIG. 2. Time-development of variational parameters.

Wigner approximation, where the Wigner function is propagated classically assuming (falsely) thatLiouvilles theorem holds: ρ(q, p; t) = ρ(q

t , pt ; 0). While the latter method clearly fails dramatically,

it is seen that the variational method is capable of accounting for tunneling through the barrier. InFigure 2 is shown the time-development of the variational parameters. We see that β

′(t) = β(t) grows

in magnitude which in turn lowers the smeared potential barrier thereby making it possible to pass thebarrier. To further investigate the performance of the variational method, we repeated the above testwith different double well potentials on the form V (x) = ax2 + bx4. Thus a sequence of potentialswere generated, keeping the potential height as above, but using different widths. For each case, thetunneling period was extracted from accurate and variational dynamics. The results are presented inFigure 3. We see that the variational method does a good job for narrow potentials but eventuallyfails as the barrier widens.

We next consider the calculation of the Kubo-transformed position correlation function of thedouble well. We show results from variational theory, accurate calculations as well as the classicalWigner model. In Figures 4 and 5 are results for an inverse temperature of 8 and 1, respectively. For1/kBT = 8 it is seen that the variational method is in fairly good agreement with the accurate results.Again, the classical Wigner model performs badly. At 1/kBT = 1, the variational method actuallyperforms less well as compared to 1/kBT = 8: It correctly predicts tunneling but roughly twice as fastas it should be.

FIG. 3. Vibrational period of double well potential as a function of barrier width (imaginary frequency).

II

Page 155: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II

115018-6 Poulsen, Svensson, and Nyman AIP Advances 7, 115018 (2017)

FIG. 4. Kubo double well position correlation function at 1/kBT = 8.

FIG. 5. Kubo double well position correlation function at 1/kBT = 1.

FIG. 6. Kubo quartic position correlation function at 1/kBT = 8.

II

Page 156: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II

115018-7 Poulsen, Svensson, and Nyman AIP Advances 7, 115018 (2017)

FIG. 7. Kubo quartic position correlation function at 1/kBT = 1.

D. Application to quartic potential

We consider the potential V (x) = 0.25x4. We again adopt natural units: = kB = m = 1. InFigures 6 and 7 are shown the Kubo-transformed correlation functions obtained from variationaltheory, accurate calculations and the classical Wigner model. We observe a good agreement betweenaccurate and variational methods at 1/kBT = 8, while at 1/kBT = 1 the amplitude given by thevariational method is too large before it dephases.

IV. VARIATIONAL PRINCIPLE FOR COMPLEX-VALUED FROZEN GAUSSIANS

A. Equations of motion

In this Section we consider the following ansatz for the variational “Wigner packet”:

ρ(q, p; t)= ρ(q, p; ξ(t), q0(t), p0(t))= expiξ(t)× exp−(

q − q0(t)p − p0(t)

)T(α 00 β

)(q − q0(t)p − p0(t)

). (18)

The variational parameters are (ξ(t), q0(t), p0(t)). Let us write θ = (ξ, q0, p0), and introduce thewriting: ρ(q, p; t) = ρ(q, p; ξ(t), q0(t), p0(t)) = ρ(q, p; θ(t)). The parameters will be consideredcomplex and hence we write q0(t) = qr(t) + iqi(t) etc. As shown in the supplementary material, whensubstituting Eq.18 into Eq.5 the resulting equations of motion are

qr

pr

qi

pi

ξr

ξi

=

pr/M

− ddqr

Vsm(qr − pi β) + Vsm(qr + pi β)2

1

αβ2

ddpi

Vsm(qr − pi β) + Vsm(qr + pi β)2

− αβM

qi

Vsm(qr + pi β) − Vsm(qr − pi β)

− 2βV sm(qr − pi β) + V sm(qr + pi β)

2pi

ddtβp2

i + αq2i

. (19)

Here, V sm means ddqr

Vsm. The smeared potential is calculated as in Eq.12 but now with β = 122

1α + 1

2 β.It is not hard to demonstrate that if q0(t) and p0(t) are allowed to be complex, then the set of Wignerpackets in Eq.18 is mathematically complete, i.e. any function f (q, p) can be expanded in them, using

II

Page 157: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II

115018-8 Poulsen, Svensson, and Nyman AIP Advances 7, 115018 (2017)

the dot product in Eq.6:

f (q, p)=∫ ∫ ∫ ∫

dqrdqidprdpi

(π)2exp(−2αq2

i − 2βp2i )ρ(q, p; θ) ρ(q, p; θ), f (q, p) . (20)

In this expression, the ξ variable in θ is zero. This expansion is a desirable property since allWigner functions can then be time-evolved, at least approximately. The complex Wigner packet inEq.18 has several interesting properties. First, let us consider dynamics in a harmonic potential. Itturns out that an expansion-propagation scheme based on Eq.19 and Eq.20 is almost exact for aharmonic potential. More precisely, if we compute (again with ξ = 0):

f (q, p; t)=∫ ∫ ∫ ∫

dqrdqidprdpi

(π)2exp(−2αq2

i − 2βp2i )ρ(q, p; θ(t)) ρ(q, p; θ(0)), f (q, p; 0)

(21)

where the Wigner-packets ρ(q, p; θ(t)) are time-evolved using Eq.19, then the result is exact apartfrom an overall scaling factor of the final Wigner function. This is of course easy to fix. It is ageneral property of the above expansion-propagation scheme, that the phase-space integral overthe time-evolved Wigner function (“norm”) is not equal to unity, but instead is a function of time.Hence, the Wigner function always has to be renormalized. A further interesting property is thatsince the Wigner-packets include a phase-factor exp(iξr(t)/), the method can in principle accountfor interference effects, see below. From the equations of motion we also notice that this factorexp(iξr(t)/) is always unity for a harmonic potential. This means that the method does not haveoscillatory phases that causes problems for near-harmonic systems. This is clearly important if one isinterested in multi-dimensional problems. We finally observe that the variables qi and pi determinethe amount of off-diagonal character of the density operator of ρ(q, p; θ(t)) in the momentum andposition representation, respectively.

B. Application to double well potential

We consider again the double well potential from the previous section. We will consider thepropagation of a Wigner packet with non-complex initial conditions, i.e. qr = 1. and pr = pi = qi

= ξr = ξ i = 0. We further set α = β = 1. We compute its average position as a function of time.Our Wigner-packet is propagated through Eq.21 using some 45000 functions in the integral. Theresult is shown in Figure 8. As opposed to the classical Wigner model, we see that the variationalWigner-packet indeed tunnels through however a bit too fast. One also sees that the norm of theWigner-packet quickly decays to around 0.10. In Figure 9 is shown the propagated Wigner functionsafter a time of 2. It is seen that the variational result reproduces the two blue, negative bassins “north”and “south” of the central positive yellow peak. Indeed, closer inspection shows that both Wigner

FIG. 8. Average position for a particle in the double well starting from a single Gaussian. q0 = 1., p0 = 0 and α = β = 1.

II

Page 158: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II

115018-9 Poulsen, Svensson, and Nyman AIP Advances 7, 115018 (2017)

FIG. 9. Variational and accurate Wigner functions in the double well after a time of t = 2.

FIG. 10. Variational and accurate Wigner functions in the double well after a time of t = 4.

functions have negative values down to some -0.5. Similar results are shown in Figure 10 now for atime of 4. Here the agreement is less good and while the accurate function has a negative bassin witha “probability” of about -1., the corresponding variational “dip” only goes down to some -0.5.

C. Application to quartic potential

Here we revisit the quartic potential from the previous section. Again, we study the propagationof the Wigner packet with the initial conditions, qr = 1., pr = pi = qi = ξr = ξ i = 0. and α = β = 1. InFigure 11 is shown the average position of the quantum particle. We see a good agreement between

FIG. 11. Average position for a particle in the quartic potential starting from a single Gaussian. q0 = 1., p0 = 0 and α = β = 1.

II

Page 159: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II

115018-10 Poulsen, Svensson, and Nyman AIP Advances 7, 115018 (2017)

FIG. 12. Variational and accurate Wigner functions in the quartic potential after a time of t = 4.

the accurate and variational scheme for short times, but we also note a too fast oscillation predicted inthe variational result. The good thing is that the normalized Wigner function clearly is more coherentthan the classical Wigner result. Again the classical Wigner result erroneously decays on a fast time-scale. Finally, we show the time-evolved variational Wigner-packet for a time of 4. and compare to theaccurate result in Figure 12. We see that the interference pattern is clearly qualitatively reproduced bythe variational method. Indeed in both plots are clearly seen a black “dip” with probability of some -1.

V. DISCUSSION AND CONCLUSION

Let us first consider the variational method based on thawed Gaussians. It should be clear that thismethod is computationally simple and can be generalized to many dimensions. Also, as opposed tomany other practical methods, it does seem to account for tunneling effects which is indeed encourag-ing. It is interesting to compare this method with Hellers now classic method for propagating thawedGaussian wavefunctions.15 His method allows for complex widths of the wavepackets meaning thatthe corresponding Wigner function is identical to our thawed Wigner functions, see Eq.9. In his paper,however, one expands the classical potential to second order around the center of the wavepacket.Thus the method does not use a smeared potential as in this work. Here, we found it essential to doso in order to account for tunneling effects through the barrier of the double well. It was observedthat the thawed Gaussian method did not perform well at 1/kBT = 1 when used together with theFeynman-Kleinert model. Inspection of the equations of motion in Eq.10 and numerical experimen-tation reveals that the problem is the smeared potential. If β

′is put equal to 0 then we essentially get

the classical Wigner result, which is not a good result either. So β′is somewhat too big. The problem

is as follows: At high temperature, the factor MΩ(xc)α(xc) in Eq.15 is initially large, since α(xc) is small.

From Eq.10 it follows that after a transient time, γ turns into a big negative number which againmakes β and thereby β

′big. Thus after a transient time, the effective potential is lowered and the

particle is tunneling through. To conclude, in the high-temperature limit, we do not obtain dynamicsfor q0 and p0 “on” the classical potential if we use the Feynman-Kleinert sampling. At first this seemscounter intuitive but the variational principle is actually acting correctly! To see this, we notice thata large β value means a very localized Wigner distribution along the momentum direction in phase-space. It trivially follows that the corresponding density operator has large off-diagonal elementsin the position representation. It is well-known that in such a case, quantum effects are importantwhen considering the dynamics of the Wigner function. So it is correct that q0 and p0 evolve non-classically. The classical high-temperature limit is only obtained if we use the variational principlefor time-evolving a distribution function being broad in phase-space both along the q and p directions.Then both α and β are small all the time and the effective potential will be close to the classical one.The problem is then the form of the Feynman-Kleinert Gaussian functions that are too localized athigh temperatures. Indeed they are quantum mechanically “illegal” since they violate Heisenbergsuncertainty relation. We should mention that this discussion is not relevant for harmonic potentialswhere the method is always exact: Here the smearing of the potential does not have any effect on thedynamics.

II

Page 160: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II

115018-11 Poulsen, Svensson, and Nyman AIP Advances 7, 115018 (2017)

Let us next discuss the variational method based on complex frozen Gaussians. It was observedto keep coherence for a longer time than the classical Wigner model, but unfortunately, also pre-dicts a somewhat too fast time-dynamics. Also, the propagated Wigner function does not keep itsnorm, although this is easily fixed afterwards. We notice that the method we use is a fully uncoupledapproach: We expand a Wigner function in our set of complex Wigner packets which are afterwardsdriven forward in time individually. This method could be improved by including a semi-classicalprefactor multiplied onto the Wigner packets. Such a factor accounts for coupling between the evolv-ing Wigner-packets but it is evaluated in an approximate way. It could be evaluated in a similarmanner as in the derivation of the Herman-Kluk semi-classical initial value representation18–20 usedfor propagating wave functions. Here, the prefactor is evaluated by linearizing the dynamics aroundthe orbit of the coherent state of interest. Such an analysis will be the subject of future research.

SUPPLEMENTARY MATERIAL

See supplementary material for a derivation of the equations of motion of the variational Wignerfunctions.

ACKNOWLEDGMENTS

G.N. gratefully acknowledges the Swedish Research Council for support of this work.

APPENDIX A

In the following, the symbol β means 1/kBT. We will use the so-called Feynman-Kleinertapproximation to the Boltzmann operator:

exp(−βH)≈∫ ∫

dxcdpcρ(xc, pc)δFK (xc, pc), (A1)

where

ρ(xc, pc)=1

2πexp(−β p2

c

2m) exp(−βW1(xc)) (A2)

and

δFK (xc, pc)=∫ ∫

dxdx x⟩ x |√

MΩ(xc)πα(xc)

exp(i

pc(x − x))

× exp(−MΩ(xc)α(xc)

(x + x

2− xc)2 − MΩ(xc)α(xc)

4(x − x)2). (A3)

In reference 7 it is shown how to determine the centroid potential W1(xc), the effective frequencyΩ(xc) and α(xc). Using Eq.A1 enables us to establish the following approximate identity:21

1Z

∫ β

0ds exp(−(β − s)H)x exp(−sH)≈ 1

ZFK

∫ ∫dxcdpcxcρFK (xc, pc)δFK (xc, pc), (A4)

which would be exact if Eq.A1 were. By multiplying this equation with the Heisenberg operatorx(t) followed by taking the trace, we obtain an expression for the Kubo-transformed position autocorrelation function:

xx(t)Kubo = (A5)

1Z

∫ β

0dsTrexp(−(β − s)H)x exp(−sH)x(t) (A6)

1ZFK

∫ ∫dxcdpcxcρFK (xc, pc)

12π

×∫ ∫

dqdp(exp(−itH/)δFK (xc, pc) exp(itH/))W [q, p]q.

II

Page 161: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II

115018-12 Poulsen, Svensson, and Nyman AIP Advances 7, 115018 (2017)

We may explore the fact that δFK (xc, pc) is an operator which has a Wigner-transform that is aGaussian function of pc. Hence we may integrate out pc:

xx(t)Kubo = (A7)

1ZFK

∫dxcxcρFK (xc) × 1

∫ ∫dqdp ( exp(−itH/)

×∫

dpc exp(−β p2c

2m)δFK (xc, pc) exp(itH/) )W [q, p]q

=1

ZFK

∫dxcxcρFK (xc)

12π

∫ ∫dqdpq exp(tL)

×∫

dpc(exp(−β p2c

2m)δFK (xc, pc))W [q, p]

=1

ZFK

∫dxcxcρFK (xc) × 1

∫ ∫dqdpq × exp(tL)

2

√π2M tanh( Ω(xc)β

2 )

βα(xc)exp(−MΩ(xc)

α(xc)(q − xc)2) (A8)

× exp(− tanh( Ω(xc)β2 )

MΩ(xc)p2).

It is the last equation that can be evaluated using the varational method of Section II. It is calculatedby sampling xc by Monte Carlo and for each xc, we propagate the local Gaussian Wigner distribution

ρxc (q, p)= 2

√tanh( Ω(xc)β

2 )

α(xc)exp(−MΩ(xc)

α(xc)(q − xc)2) exp(− tanh( Ω(xc)β

2 )

MΩ(xc)p2). (A9)

1 Y. Zhao, T. Yamamoto, and W. H. Miller, J. Chem. Phys. 120, 3100 (2004).2 R. V. Harrevelt, G. Nyman, and U. Manthe, J. Chem. Phys. 126, 084303 (2007).3 T. D. Hone, J. A. Poulsen, P. J. Rossky, and D. E. Manolopoulos, J. Phys. Chem. B 112, 294–300 (2008).4 I. R. Craig and D. E. Manolopoulos, J. Chem. Phys. 121, 3368–3373 (2004).5 G. A. Voth, Adv. Chem. Phys. 93, 135 (1996).6 J. Liu, J. Chem. Phys. 140, 224107 (2014).7 J. A. Poulsen, G. Nyman, and P. J. Rossky, J. Chem. Phys. 119, 12179 (2003).8 Q. Shi and E. Geva, J. Phys. Chem. B 107, 9059 (2003).9 H. Wang, X. Sun, and W. H. Miller, J. Chem. Phys. 108, 9726 (1998).

10 K. K. G. Smith, J. A. Poulsen, G. Nyman, and P. J. Rossky, J. Chem. Phys. 142, 244113 (2015).11 E. Wigner, Phys. Rev. 40, 749 (1932).12 Quantum Mechanics in Phase Space, edited by C. K. Zachos, D. B. Fairlie, and T. L. Curtright, World Scientic Series in

20th Century Physics, Vol. 34, World Scientific (2005).13 J. A. Poulsen, J. Chem. Phys. 134, 034118 (2011).14 A. Raab, Chem. Phys. Lett. 319, 674–678 (2000).15 E. J. Heller, J. Chem. Phys. 64, 63 (1976).16 R. D. Coalson and M. Karplus, J. Chem. Phys. 93, 3919 (1990).17 D. J. Coughtrie and D. P. Tew, J. Chem. Phys. 143, 044102 (2015).18 M. F. Herman and E. Kluk, Chem. Phys. 91, 27 (1984).19 W. H. Miller, J. Phys. Chem. 106, 8132 (2002).20 D. V. Shalashilin and M. S. Child, Chem. Phys. 304, 103–120 (2004).21 J. A. Poulsen, G. Nyman, and P. J. Rossky, J. Phys. Chem. B 108, 19799–19808 (2004).

II

Page 162: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II

II

Page 163: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II

+ - - '-

- , / - -/ + , / ,

-/ . - -

ρ(q, p; t) = ρ(q, p; ξ(t), q0(t), p0(t))

= expiξ(t) exp−(q − q0(t)p− p0(t)

)T(

α 00 β

)(q − q0(t)p− p0(t)

)

ρ(q, p; t) = ρ(q, p; q0(t), p0(t), α(t), γ(t), β(t))

= exp−(q − q0(t)p− p0(t)

)T(

α(t) γ(t)γ(t) β(t)

)(q − q0(t)p− p0(t)

).

$ --- -- - / / $ -- --

/- N(t) . / % N(t) - --- --#

. - , -/ - /

- - "

ρ(q, p; t), ρ(q, p; t) =ˆ ˆ

dqdp

2πρ∗(q, p; t)ρ(q, p; t)) = 1.

$ - /-- / . , / . "

AW (q, p; t), BW (q, p; t) =ˆ ˆ

dqdp

2πA∗

W (q, p; t)BW (q, p; t)). !

) - q0 - p0 . / /- . - -/

% / - , / /- . - (/

-- -/ /- -- ,

/ *. , -/ / .

/- -"

• + - - ξ(t) - ξr(t) -- ,

-/ / - -/ -- (/ / - --

& - - -/ -

ξr(t) - (/ ξr(t) - - --/

.

• + - ξ(t) / -/ q0(t) - p0(t)

+ -/ expiξ(t) - / / - .-

, / + - " & - -

- , / F + -- %-/

, -/ - -/ -

II

Page 164: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II

* 22 1 24 3 + 1

ρ(q, p; q(t), p0(t), α(t), γ(t), β(t)) 3

24 4 ) 4 4

2 2'

θ(t) = q0(t), p0(t), α(t), γ(t), β(t) 0 222

4 42 3 2 3 4 42

J [θ, θ] =

ˆ t

0

ds

⟨ρ(q, p; θ(s)), L − ∂

∂sρ(q, p; θ(s))

⟩"

=

ˆ t

0

ds⟨ρ(q, p; θ(s)), L − θ · ∇θρ(q, p; θ(s))

⟩.

- L /2 0 2 s = 0 s = t

4 0 24 4 2 2 J [θ, θ]

22 4 4 22 δθ(s) 4 2

2

θ(s) 2 2 2 2 2

Error =

⟨L − ∂

∂sρ(q, p; θ(s)), L − ∂

∂sρ(q, p; θ(s))

⟩#

2 2 0 2 2 42

+/22 2'

∂s ∂

∂θiJ [θ, θ] = ∂

∂θiJ [θ, θ], $

θi 2 ith θ 4 4 1 4

2 24 3 2 224 23 4

4 222 22 24 3 + " 2

+ #

0 2 4 +/22 2 2

θ =⟨∇θρ(q, p;

θ),∇θρ(q, p;θ)⟩−1 ⟨

∇θρ(q, p;θ),Lρ(q, p; θ)

⟩. %

0 2 2 L. , 2 L = L1 + L2

4 L

L1 2 L2'θ = θ1 + θ2 - ( . 23

θ

2 2 2 2 2 0 2 2 ' 1 4

2 , 4'

V (x) =

ˆ +∞

−∞dka(k) exp(ikx). &

!

II

Page 165: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II+2 2 2 22 223 2 t 2

θ(t) 0

θ(t) 2 42 ),22 2 Lk

3

Lk = V k(q)

∂p− 2

24V k (q)

∂3

∂p3+ ..,

Vk(x) = exp(ikx) 2 t (2 42 θ(k, t) . 42

), Jk 2 3 3

∂s ∂

∂θ(k)iJk[θ, θ(k, t)] =

∂θiJk[θ, θ(k, t)].

/ 2 3

θ(t) = θ(free, t) +

ˆ +∞

−∞dka(k)θ(k, t), !

θ(free, t) L =− pM

∂∂q

* α(t) β(t) γ(t) q0(t) 2 p0(t) 3 222 22

2 4 ) ! ' 22 2 2 222

22 2 4 24 2 1 2

2 q0 = qr + iqi 4 4 3 ' 3

222 2 224 ξ(t) / 222 1

24

ρ(q, p; t) = N(t) exp−(q − q0(t)p− p0(t)

)T(

α(t) γ(t)γ(t) β(t)

)(q − q0(t)p− p0(t)

). "

/ 22 24 42 3 22 4&

N(t) = 2

√Λ(αrβr − γ2

r )1/4

#

ln Λ =1

2B∗

(q∗0p∗0

)+ c.c.TA−1B∗

(q∗0p∗0

)+ c.c.

−(

q∗0p∗0

)T

B∗(

q∗0p∗0

)−(

q0p0

)T

B

(q0p0

), $

A =

(αr γrγr βr

), B =

(α γγ β

). %

-4 2 Λ = 1 3 q0(t), p0(t) α(t), β(t), γ(t) 2 2

2 / 2 2 2 3 4 2 22 N 24

#

II

Page 166: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II ρ(q, p; t) --/ --- -- ' - $

$ / - --/- , /

/- %)-- -

- -- - - /- -- -/ - .

/- * /- . N -/ &

/ - / --- .

J [θ, θ] =

ˆ t

0

ds

⟨ρ(q, p; θ(s)), Lk − ∂

∂sρ(q, p; θ(s))

⟩, !

Lk % ( % % !

J [θ, θ] =

1

i

ˆ t

0

ds exp(−k2

8

βr

αrβr − γ2r

[2γ2i + 1]− 2k2

8

βi

αrβr − γ2r

[αrβi − 2γrγi])

exp(−2k2

8βr) exp(ik(λq + qr))

×exp(−k2

4

γiβr − βiγrαrβr − γ2

r

− ik(qiγr + piβr − γiλq − βiλp))

− exp(+k2

4

γiβr − βiγrαrβr − γ2

r

+ ik(qiγr + piβr − γiλq − βiλp)) "

+i

ˆ t

0

ds2αr qrqi + 2γr qrpi + 2βrprpi + 2γrprqi

+1

4

βiαr + αiβr − 2γiγrαrβr − γ2

r

− 2λqαr qi + αrqi + γrpi + γrpi

−2λpγr qi + βrpi + γrqi + βrpi−2qiαiqi − αiq

2i − 2piβipi − βip

2i − 2qiγipi − 2piγiqi − 2piγiqi.

+ -- - - -- % "

(λq

λp

)=

−1

αrβr − γ2r

(βr −γr−γr αr

)(αi γiγi βi

)(qipi

). #

% " //- &- / &

- - % " / - -

- - /. -- -

- /- . / %) -

, - - -- -

θ = q0(t), p0(t), α(t), β(t), γ(t) %)-- %) - L

θ

- - t = 0 + - - --

θ - /- %) -

. -- % - - - - , -

- / /- /- %

II

Page 167: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II ( 0 00 . 02

0 00 20 (+ 0 0 00

002 * 02 2 (+ 0 001

0 0 2 ( $ 0 0 0 0

0 0 2 00 001 20 1 20 )

0 0 λq 001

00 00 0 00

θ - 0 ( $ 0

2 %

J [θ, θ] =1

i

ˆ t

0

ds exp(−k2

8

βr

αrβr − γ2r

− 2k2

8βr) exp(ikqr)

×exp(−k2

4

γiβr − βiγrαrβr − γ2

r

− ik(qiγr + piβr))

− exp(+k2

4

γiβr − βiγrαrβr − γ2

r

+ ik(qiγr + piβr))

+i

ˆ t

0

ds2αr qrqi + 2γr qrpi + 2βrprpi + 2γrprqi

+1

4

βiαr + αiβr − 2γiγrαrβr − γ2

r

.

' (+00 0 ) 2 001

qr / 1 00 0 0 0 0 2

0 %

∂qrJ =

d

dt

∂qrJ ⇔ 0 =

d

dt[αrqi + γrpi].

. qi 0 pi 0 0 0 0

, 2 pr /

∂prJ =

d

dt

∂prJ ⇔ 0 =

d

dt[2βrpi + 2γrqi].

&0 0 qi 0 pi 0 0 0 /

0 qi%∂

∂qiJ =

d

dt

∂qiJ ⇔

−2kγr exp(−k2

8

βr

αrβr − γ2r

− 2k2

8βr) exp(ikqr) + 2iαr qr + 2iγrpr = 0. !

, 2 pi%∂

∂piJ =

d

dt

∂piJ ⇔

−2kβr exp(−k2

8

βr

αrβr − γ2r

− 2k2

8βr) exp(ikqr) + i2γr qr + i2βrpr = 0. "

#

II

Page 168: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II/ ) $ qr = 0 1 pr = −ik exp(−k2

8βr

αrβr−γ2r−

2k2

8 βr) exp(ikqr) = − ddqr

exp(−k2

8βr

αrβr−γ2r− 2k2

8 βr) exp(ikqr) 1 3

1 1 3 + 1 2$ (1

1 pr 11 112 1 3 1

3 11 313 1 pi = 0 ' 2$ 01 1

1 1 21 pr 1 1 1 *

1 13 ,1 & 1 pr1 - 3 αi 0 1

∂αiJ =

d

dt

∂αiJ ⇔

0 =d

dt

βr

αrβr − γ2r

⇔ βr = 0, !

1 1 1 αrβr − γ2r % 1 13 1

2 11 1 * βi

∂βiJ =

d

dt

∂βi

J ⇔

1

iexp(−k2

8

βr

αrβr − γ2r

− 2k2

8βr) exp(ikqr)

k2

2

γrαrβr − γ2

r

=i

4

d

dt

αr

αrβr − γ2r

αr = 2γr ×d2

dq2rexp(−k2

8

βr

αrβr − γ2r

− 2k2

8βr) exp(ikqr) "

* γi

∂γiJ =

d

dt

∂γiJ ⇔

−1

iexp(−k2

8

βr

αrβr − γ2r

− 2k2

8βr) exp(ikqr)

k2

2

βr

αrβr − γ2r

= − i

2

d

dt

γrαrβr − γ2

r

γr = βr ×d2

dq2rexp(−k2

8

βr

αrβr − γ2r

− 2k2

8βr) exp(ikqr). #

- 1 αr / ), 1

∂αrJ =

d

dt

∂αrJ

3 1 1 1 1 111 11

1 1 1 / 1 3 1 βr 1 γr1 1 ), 1 2 . 1

1 1 11 111 11 1

11 1 t = 0 0 1 1 1 1 1

αrβr − γ2r $

d

dtαrβr − γ2

r

#

II

Page 169: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II= αrβr + αrβr − 2γrγr

= 2γr ×d2

dq2rexp(−k2

8

βr

αrβr − γ2r

− 2k2

8βr) exp(ikqr)βr !

+αr · 0− 2γrβr ×d2

dq2rexp(−k2

8

βr

αrβr − γ2r

− 2k2

8βr) exp(ikqr) = 0.

+ ,. - * ,,, -

J [θ, θ] =

ˆ t

0

ds

⟨ρ(q, p; θ(s)), − p

M

∂q− ∂

∂sρ(q, p; θ(s))

⟩"

= i

ˆ t

0

ds2αr(qr −prM

)qi + 2γr(qr −prM

)pi + 2βrprpi + 2γrprqi

+1

2M

γiαr − αiγrαrβr − γ2

r

+1

4

βiαr + αiβr − 2γiγrαrβr − γ2

r

.

' , ,, , , , ,, , ,

. ,, ., - ., , % * %(,,

%( , , ,, ,, & qr#

∂qrJ =

d

dt

∂qrJ ⇔ 0 =

d

dt[αrqi + γrpi].

* qi , pi , , , ,

) . pr#

∂prJ =

d

dt

∂prJ ⇔ 0 =

d

dt[2βrpi + 2γrqi].

$, , qi , pi , , , +

, qi#∂

∂qiJ =

d

dt

∂qiJ ⇔

αr(qr −prM

) + γrpr = 0.

* %( , pi

∂piJ =

d

dt

∂piJ ⇔

γr(qr −prM

) + βrpr = 0.

* , , qr − pr

M , pr = 0 ) . αi

* %( , ,

∂αiJ =

d

dt

∂αiJ ⇔

!

II

Page 170: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II− 1

2M

γrαrβr − γ2

r

=1

4

d

dt

βr

αrβr − γ2r

⇔ βr = −2γrM

,

,, , , , ( βi -,

∂βiJ =

d

dt

∂βi

J ⇔ αr = 0.

(, . γi$∂

∂γiJ =

d

dt

∂γiJ ⇔

αr

2M= −1

2γr ⇔ γr = −αr

M.

&, , , ,

ddtαrβr − γ2

r

) . αr * ') , -.

∂αrJ =

d

dt

∂αrJ

2(qr −prM

)qi +1

2M

∂αr

γiαr − αiγrαrβr − γ2

r

= 0,

. , ,, , * ') , βr , γr ,

, % , ,, ,

+ ., ' , , , ,

, + ,

d

dt

⎛⎜⎜⎜⎜⎝

qrprαr

γrβr

⎞⎟⎟⎟⎟⎠

!

=

⎛⎜⎜⎜⎜⎝

pr/M00

−αr

M

− 2γr

M

⎞⎟⎟⎟⎟⎠+

ˆ +∞

−∞dka(k)

⎛⎜⎜⎜⎜⎜⎝

0

− ddqr

exp(−k2

8βr

αrβr−γ2r− 2k2

8 βr) exp(ikqr)

2γr × d2

dq2rexp(−k2

8βr

αrβr−γ2r− 2k2

8 βr) exp(ikqr)

βr × d2

dq2rexp(−k2

8βr

αrβr−γ2r− 2k2

8 βr) exp(ikqr)

0

⎞⎟⎟⎟⎟⎟⎠

.

) βr(t) =

βr

2 1 + 12(αrβr−γ2

r ) * '! -.

d

dt

⎛⎜⎜⎜⎜⎝

qrprαr

γrβr

⎞⎟⎟⎟⎟⎠

"

#

II

Page 171: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II=

⎛⎜⎜⎜⎜⎝

pr/M00

−αr

M

− 2γr

M

⎞⎟⎟⎟⎟⎠

+

ˆ +∞

−∞dka(k) exp(−2k2

4βr)

⎛⎜⎜⎜⎜⎜⎝

0− d

dqrexp(ikqr)

2γr × d2

dq2rexp(ikqr)

βr × d2

dq2rexp(ikqr)

0

⎞⎟⎟⎟⎟⎟⎠

.

' + ) ) * )) * ) a(k) = 12π

´ +∞−∞ dyV (y) exp(−iky)#

ˆ +∞

−∞dka(k) exp(−2k2

4βr) exp(ikqr)

=

ˆ +∞

−∞dk exp(−2k2

4βr) exp(ikqr)

1

ˆ +∞

−∞dyV (y) exp(−iky)

=1

ˆ +∞

−∞dyV (y)

ˆ +∞

−∞dk exp(−2k2

4βr + ik(qr − y))

=1

ˆ +∞

−∞dyV (y)

ˆ +∞

−∞dk exp(−β

r

2

4(k − 2i

(qr − y)

βr2

)2 − (qr − y)2

βr2

)

=1

π2βr

ˆ +∞

−∞dyV (y) exp(− (qr − y)2

βr2

)

≡ Vsm(qr). "

%! ) ) ) )) #

d

dt

⎛⎜⎜⎜⎜⎝

qrprαr

γrβr

⎞⎟⎟⎟⎟⎠

=

⎛⎜⎜⎜⎜⎝

pr/M00

−αr

M−2 γr

M

⎞⎟⎟⎟⎟⎠

+

⎛⎜⎜⎜⎜⎜⎝

0− d

dqrVsm(qr)

2γr × d2

dq2rVsm(qr)

βr × d2

dq2rVsm(qr)

0

⎞⎟⎟⎟⎟⎟⎠

,

d

dt

⎛⎜⎜⎜⎜⎝

qrprαr

γrβr

⎞⎟⎟⎟⎟⎠

=

⎛⎜⎜⎜⎜⎝

pr/M− d

dqrVsm(qr)

2γr ×MΩ2(qr)βr ×MΩ2(qr)− αr

M−2 γr

M

⎞⎟⎟⎟⎟⎠

,

+ + Ω(qr) ) * *#

MΩ2(qr) =d2

dq2rVsm(qr).

$) ) ) )) ) & )

) ) % +) )+ ) ) + )

' ) # ( ) ) ) )) +

II

Page 172: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II1 1 .13% + 1 1

1 1 1 31 1 * 1

1

V (x) = −0.5x2 + 0.1x4.

' 1 31 . 13 1 1 & $

3131

Vsm(x) = −0.5(x2 +1

2βr2) + 0.1x4 + 0.1 ∗ 3βr2x2 + 0.1 ∗ 3

4(βr2)2.

- 3 23

F (x) = −V sm(x) = x− 4 ∗ 0.1x3 − 0.1 ∗ 6βr2x.

) βr2 = 1 3 1 1

,/0 3 1 13 21 (

3 & # (3 1 1 111

1 3

' & $ 1 1 3131 1 31

1 βr2 = 2/2⟨p2⟩ - 1 13

3131 21

( 3 11 & %

ρ(q, p; t) = 2 expiξ(t)− (q − q0(t)p− p0(t)

)T(

α 00 β

)(q − q0(t)p− p0(t)

). !

- 2 3 111 11 1 % q0(t) p0(t) 1 ξ(t). ξ(t) = ξr(t)+ iξi(t) 3 - . 13 113 112

2 1 3 & %

ρ(q, p; t), ρ(q, p; t) =ˆ ˆ

dqdp

2πρ∗(q, p; t)ρ(q, p; t)) = exp(2[αq2i + βp2i − ξi]).

"

) & ! & "

J [θ, θ] = #

1

i

ˆ t

0

ds exp(−2k2

4β) exp(ikqr)

×exp(−ikpiβ)− exp(ikpiβ) × exp(2[αq2i + βp2i − ξi])

+i

ˆ t

0

ds[2αqrqi+2βprpi+2iαqiqi+2βipipi− iξi− ξr]× exp(2[αq2i +βp2i − ξi]).

II

Page 173: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II& β′ = 1

2α2 + β2 # + - + +

+

J [θ, θ] = !

1

i

ˆ t

0

ds exp(−2k2

4β′) exp(ikqr)

×exp(−ikpiβ)− exp(ikpiβ) × exp(2[αq2i + βp2i − ξi])

+i

ˆ t

0

ds[2αqrqi + 2βprpi − ξr]× exp(2[αq2i + βp2i − ξi]).

) $' + + % qi"

∂qiJ =

d

dt

∂qiJ ⇔

i2αqr × exp(2[αq2i +p2iα2

− ξi])

+4αqi1

iexp(−2k2

4β′) exp(ikqr)exp(−ikpiβ)− exp(ikpiβ)

+i[2αqrqi + 2βprpi − ξr] × exp(2[αq2i + βp2i − ξi]) = 0.

* , qr = 0 +

ξr =

−1

exp(−2k2

4β′) exp(ikqr)exp(−ikpiβ)− exp(ikpiβ)+ 2βprpi.

( - pi"

− exp(−2k2

4β′) exp(ikqr)kβexp(−ikpiβ) + exp(ikpiβ)+ 2iβpr

× exp(2[αq2i + βp2i − ξi])

+2βpi1

iexp(−2k2

4β′) exp(ikqr)exp(−ikpiβ)− exp(ikpiβ)

+i[2αqrqi + 2βprpi − ξr] × exp(2[αq2i +p2iα2

− ξi]) = 0.

* + , + + - + +-

pr = −ik exp(ikqr) exp(−2k2

4β′)

exp(−ikpiβ) + exp(ikpiβ)2

= − d

dqr

exp(ik(qr − piβ)) + exp(ik(qr + piβ))2

exp(−2k2

4β′),

II

Page 174: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II*

ξr =

−1

exp(−2k2

4β′) exp(ikqr)exp(−ikpiβ)− exp(ikpiβ)+ 2βprpi

= −1

exp(−2k2

4β′)exp(ik(qr − piβ))− exp(ik(qr + piβ)) + 2βprpi,

+ ** $ ' * $& + ξi

∂ξiJ =

d

dt

∂ξiJ

( +* ( * * *

++ ξr ' + qr ( $& * *

∂qrJ =

d

dt

∂qrJ ⇔

k

exp(−2k2

4β′) exp(ikqr)

×exp(−ikpiβ)− exp(ikpiβ) × exp(2[αq2i + βp2i − ξi])

= i2αqi × exp(2[αq2i + βp2i − ξi]) + i2αqid

dtexp(2[αq2i + βp2i − ξi]).

)

ξi = αq2i + βp2i

*#

qi = − ik

2αexp(−2k2

4β′) exp(ikqr)× exp(−ikpiβ)− exp(ikpiβ) !

= − 1

2αexp(−2k2

4β′)

d

dqrexp(ik(qr − piβ)− exp(ik(qr + piβ))

=1

αβ2exp(−2k2

4β′)

d

dpi

exp(ik(qr − piβ) + exp(ik(qr + piβ))2

.

' + $& * + pr#

∂prJ =

d

dt

∂prJ ⇔

0 = 2βpi ⇔ pi = 0. "

%* + ξr ( $&

∂ξrJ =

d

dt

∂ξrJ ⇔

II

Page 175: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II0 = − d

dtexp(2[αq2i + βp2i − ξi]),

* ) "

' ( * ) & ( (( )

J [θ, θ] =

ˆ t

0

ds

⟨ρ(q, p; θ(s)), − p

M

∂q− ∂

∂sρ(q, p; θ(s))

= i

ˆ t

0

ds2[− α

Mprqi + αqrqi + βprpi − ξr]× exp(2[αq2i + βp2i − ξi]).

# "$ * ξi!

∂ξiJ =

d

dt

∂ξiJ ⇔

ξr = − α

Mprqi + αqrqi + βprpi.

"$ * qi!∂

∂qiJ =

d

dt

∂qiJ ⇔

qr −prM

= 0,

( " "$ * pi!

∂piJ =

d

dt

∂piJ ⇔ pr = 0.

% "$ * ξr!

∂ξrJ =

d

dt

∂ξrJ ⇔

0 = − d

dtexp(2[αq2i + βp2i − ξi]),

* (( ( " "$ * qr!

∂qrJ =

d

dt

∂qrJ ⇔

qi = 0.

"$ * pr!∂

∂prJ =

d

dt

∂prJ ⇔

− α

Mqi = βpi ⇔ pi = − α

βMqi.

II

Page 176: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II( ) ) + * ) %

+ ) )+ "

⎛⎜⎜⎜⎜⎜⎜⎝

qrprqipiξrξi

⎞⎟⎟⎟⎟⎟⎟⎠

=

⎛⎜⎜⎜⎜⎜⎜⎝

pr/M00

− αβM qi0

2βpipi

⎞⎟⎟⎟⎟⎟⎟⎠

+

ˆ +∞

−∞dka(k) exp(−2k2

4β)×

⎛⎜⎜⎜⎜⎜⎜⎜⎝

0

− ddqr

exp(ik(qr−piβ))+exp(ik(qr+piβ))2

1αβ2

ddpi

exp(ik(qr−piβ)+exp(ik(qr+piβ))2

0

− 1 exp(ikqr)exp(−ikpiβ)− exp(ikpiβ) − 2β d

dqr

exp(ik(qr−piβ))+exp(ik(qr+piβ))2 pi

2αqiqi

⎞⎟⎟⎟⎟⎟⎟⎟⎠

.

#) + +) * $ ! '

⎛⎜⎜⎜⎜⎜⎜⎝

qrprqipiξrξi

⎞⎟⎟⎟⎟⎟⎟⎠

= !

⎛⎜⎜⎜⎜⎜⎜⎜⎝

pr/M

− ddqr

Vsm(qr−piβ)+Vsm(qr+piβ)2

1αβ2

ddpi

Vsm(qr−piβ)+Vsm(qr+piβ)2

− αβM qi

Vsm(qr+piβ)−Vsm(qr−piβ) − 2β d

dqr

Vsm(qr−piβ)+Vsm(qr+piβ)2 pi

ddtβp2i + αq2i

⎞⎟⎟⎟⎟⎟⎟⎟⎠

.

& ( )+ )"

ρ(q, p; t) = 2 expiξr(t)− α(q − qr(t))2 + 2iαqi(t)(q − qr(t))

× exp−β(p− pr(t))2 + 2iβpi(t)(p− pr(t))

× expαq2i (t) + βp2i (t)− ξi(t).' ) ξi ) ))

))) ( )+ )

ρ(q, p; t) = 2 expiξr(t)− α(q − qr(t))2 + 2iαqi(t)(q − qr(t))

× exp−β(p− pr(t))2 + 2iβpi(t)(p− pr(t))

II

Page 177: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II× expαq2i (0) + βp2i (0)− ξi(0).

'/ - / ξi(0) = αq2i (0) + βp2i (0) , -/

-- - #

ρ(q, p; t) = 2 expiξr(t)− α(q − qr(t))2 + 2iαqi(t)(q − qr(t)) "

× exp−β(p− pr(t))2 + 2iβpi(t)(p− pr(t)).

+ - / - ξi --. , -/

/ - - ( - / -

-// & & "

% - -/ /- V (x) = M2 Ω2x2

+ - -

- - - - /- + -

⎛⎜⎜⎜⎜⎜⎜⎝

qrprqipiξrξi

⎞⎟⎟⎟⎟⎟⎟⎠

=

⎛⎜⎜⎜⎜⎜⎜⎝

pr/M−MΩ2qrβ/αMΩ2pi− α

βM qi0

2qipi(− αM + βMΩ2)

⎞⎟⎟⎟⎟⎟⎟⎠

. "

+ - - . -/ - /

-/ /-

% -/ ρ(q, p; θ) / - --/ -

-- -- $ - &) - - --

-#

∂s ∂

∂θiρ(q, p; θ(s)), Lk − ∂

∂sρ(q, p; θ(s))

"

=∂

∂θi

ρ(q, p; θ(s)), Lk − ∂

∂sρ(q, p; θ(s))

.

* / -/ N(θ) × ρ(q, p; θ) N(θ) -

. --/- -- /- ( - &) -

-- - N(θ) .

+ /

N(θ)× ρ(q, p; θ(s)), L − ∂

∂sN(θ)× ρ(q, p; θ(s))

= N∗(θ)N(θ)

ρ(q, p; θ(s)), L − ∂

∂sρ(q, p; θ(s))

!

II

Page 178: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

II−N∗(θ)

⟨ρ(q, p; θ(s)), ρ(q, p; θ(s))

∂sN(θ)

⟩"

=∣∣∣N(θ)

∣∣∣2⟨ρ(q, p; θ(s)), L − ∂

∂sρ(q, p; θ(s))

− d

dtlnN(θ).

- 0 00 2 2 0 0 0 ,

0 0 &) 0

J =∣∣∣N(θ)

∣∣∣2⟨ρ(q, p; θ(s)), Lk − ∂

∂sρ(q, p; θ(s))

⟩. "!

- &) 0 J

∂s ∂

∂θi∣∣∣N(θ)

∣∣∣2⟨ρ(q, p; θ(s)), L − ∂

∂sρ(q, p; θ(s))

=∂

∂θi

∣∣∣N(θ)∣∣∣2⟨ρ(q, p; θ(s)), L − ∂

∂sρ(q, p; θ(s))

⟩""

∂s

∣∣∣N(θ)∣∣∣2

∂θi⟨ρ(q, p; θ(s)), L − ∂

∂sρ(q, p; θ(s))

=∂

∂θi

∣∣∣N(θ)∣∣∣2⟨ρ(q, p; θ(s)), L − ∂

∂sρ(q, p; θ(s))

⟩. "#

' N(θ) 0 0 0

∣∣∣N(θ)∣∣∣2 ∂

∂s ∂

∂θi⟨ρ(q, p; θ(s)), L − ∂

∂sρ(q, p; θ(s))

= ∂

∂θi

∣∣∣N(θ)∣∣∣2

⟨ρ(q, p; θ(s)), L − ∂

∂sρ(q, p; θ(s))

⟩+ "$

∣∣∣N(θ)∣∣∣2 ∂

∂θi

⟨ρ(q, p; θ(s)), L − ∂

∂sρ(q, p; θ(s))

⟩.

- 00 0 0 2

0 102 - 0 0 & "

* 0 0 0 N(θ)×ρ(q, p; θ) 0 0 00 00

0 20 00 02 - & "!"$ 0 0 2

0 0 & "$ 0 0 0 0 & "

./ + ( %0 #

"

II

Page 179: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

III

Paper III

Classical Wigner ModelBased on a Feynman PathIntegral Open Polymer

Reproduced fromClassical Wigner Model Based on a Feynman Path Integral OpenPolymerS. Karl-Mikael Svensson, Jens Aage Poulsen, and Gunnar NymanThe Journal of Chemical Physics, 2020, 152 (9), pp 094111-1–094111-20, DOI: 10.1063/1.5126183,with the permission of AIP Publishing

157

III

Page 180: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

III

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

Classical Wigner model based on a Feynmanpath integral open polymer

Cite as: J. Chem. Phys. 1 52, 0941 1 1 (2020); doi: 1 0.1 063/1 .51 261 83Submitted: 6 September 201 9 • Accepted: 1 6 February 2020 •Published Online: 5 March 2020

S. Karl-Mikael Svensson, Jens Aage Poulsen,a) and Gunnar Nymanb)

AFFILIATIONSDepartment of Chemistry and Molecular Biology, University of Gothenburg, SE 405 30Gothenburg, Sweden

a)Electronic mail: [email protected])Author to whom correspondence should be addressed: [email protected]

ABSTRACTThe classical Wigner model is one way to approximate the quantum dynamics of atomic nuclei. Here, a newmethod is presented for samplingthe initial quantummechanical distribution that is required in the classicalWignermodel. The newmethod is tested for the position, position-squared, momentum, and momentum-squared autocorrelation functions for a one-dimensional quartic oscillator and double well potentialas well as a quartic oscillator coupled to harmonic baths of different sizes. Two versions of the new method are tested and shown to possiblybe useful. Both versions always converge toward the classical Wigner limit. For the one-dimensional cases, some results that are essentiallyconverged to the classical Wigner limit are acquired and others are not far off. For the multi-dimensional systems, the convergence is slower,but approximating the sampling of the harmonic bath with classical mechanics was found to greatly improve the numerical performance.For the double well, the new method is noticeably better than the Feynman–Kleinert linearized path integral method at reproducing the exactclassicalWigner results, but they are equally good at reproducing exact quantummechanics. The newmethod is suggested as being interestingfor future tests on other correlation functions and systems.Published under license by AIP Publishing. https://doi.org/10.1063/1.5126183., s

I. INTRODUCTION

When studying molecular systems computationally, a prob-lem that can arise is how to account for the quantum mechani-cal behavior of the atomic nuclei without having to solve the timedependent Schrödinger equation using wavefunctions and, instead,work with, e.g., trajectories, which has the potential to be signif-icantly cheaper computationally for many chemically interestingproblems. Several different methods exist, which can be used toaddress this problem, such as centroid molecular dynamics (CMD)1that has been well explained by Jang and Voth,2 ring polymer molec-ular dynamics (RPMD),3 semi-classical initial value representation(SC-IVR)4 that has been well explained by Miller,5 Matsubaradynamics,6 and the classical Wigner (CW) method7,8 [also calledlinearized semi-classical initial value representation (LSC-IVR), andlinearized path integral (LPI)]. A new implementation of the classi-cal Wigner method is the topic of the present article.

The classical Wigner method starts with a quantum mechan-ical phase space distribution, which is propagated forward in timeclassically. Finding the initial distribution is typically problematic. In

this work, a new way to sample the initial phase space distribution ispresented and tested for some simple problems.

The classical Wigner method, CMD, and RPMD can all be seenas approximations to Matsubara dynamics and give worse resultsthan Matsubara dynamics.6,9 However, for large systems, Mat-subara dynamics would be too computationally demanding to bepractical.6

Comparing CMD and RPMD to the classical Wigner method,one major advantage of the former two is that the quantummechan-ical ensemble is conserved during the dynamics, while it is not inthe classical Wigner method. On the other hand, for a harmonicoscillator, the classical Wigner method is exact for any correlationfunction,7 while CMD can formally be done for non-linear oper-ators,10 but it is a much more complicated process than for linearoperators, and RPMD is only exact for correlation functions whereat least one operator is linear.3

The classical Wigner method can be seen as an approximationof the semi-classical initial value representation. In this approxima-tion, quantum real time coherence is, however, lost, which may notbe very important in large or condensed phase systems since these

J. Chem. Phys. 152, 094111 (2020); doi: 10.1063/1.5126183 152, 094111-1

Published under license by AIP Publishing

III

Page 181: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

III

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

typically decohere rapidly.11 An advantage with the approximationis less oscillatory integrands to handle numerically.11

From the above comparisons, it can be seen that one of thereasons for using the classical Wigner method is that it may betterhandle non-linear correlation functions at a reasonable expense ofcomputational resources, compared to the other available methods.

In the new implementation of the classical Wigner method thatis presented in this paper, the initial phase space distribution is sam-pled with an imaginary time path integral polymer similar to thosethat can be found in CMD and RPMD, but in those cases, thesepolymers are closed rings and in the present implementation of theclassical Wigner method, the polymer has an opening. The openpolymer presented here is more closely related to those that can befound in the work of Bose and Makri12 and Bonella et al.13,14

In what follows, first, the classical Wigner method is explained(Sec. II), and a new path integral open polymer implementation of itis presented (Sec. III). After that, the computational details (Sec. IV)and the results (Sec. V) are presented, and conclusions (Sec. VI) aredrawn.

II. CLASSICAL WIGNER METHODThe classical Wigner method was in its first form introduced,

but not necessarily recommended by Heller.7 In its current moregeneral form, which is applicable to correlation functions, it wasintroduced byWang, Sun, andMiller.8 The classical Wigner methodhas as its basis the Wigner phase space distribution.15,16 In onedimension, easily generalized to any number of dimensions, theWigner transform is

(Ω)W[x, p] = ∫∞

−∞dη e−iηp/h⟨x + η

2∣Ω∣x − η

2⟩, (1)

where Ω is an arbitrary operator, x is the position, p is the momen-tum, η is a variable with the dimension of the length, h is the reducedPlanck constant, and i is the imaginary unit. The Wigner function isthe Wigner transform of the probability density operator 1

2πh ∣Ψ⟩⟨Ψ∣,where ∣Ψ⟩ is the ket of the state of the system. This distribution is anexact quantummechanical quasi-probability distribution and can beused to calculate expectation values,

⟨Ω⟩ = ∫∞

−∞dx ∫

−∞dp ( 1

2πh ∣Ψ⟩⟨Ψ∣)W[x, p](Ω)W[x, p]. (2)

The classical Wigner method consists of taking the quantummechanical transformed quantity and propagating it forward in timewith classical mechanics (CM), i.e.,

( e iHth Ω e−

iHth )

W[x, p] ≃ (Ω)W[x(t), p(t)], (3)

where H is theHamiltonian operator and t is time. Thus, the classicalWigner method works by initiating with a quantummechanical dis-tribution and then propagating it classically, thereby taking accountof all equilibrium quantum effects e.g., zero-point energy, but over-looking e.g., dynamic tunneling and quantum interference withinthe dynamics. Themethod is exact for potentials up to and includingharmonic terms.7 For a detailed derivation of the classical Wignermethod, the review by Liu17 can be recommended.

The classical Wigner method has been applied successfully to,e.g., calculating the kinetic energy and density fluctuation spectrumin liquid neon18 and vibrational energy relaxation rate constants,19,20but the classical Wigner method has also been found to have limita-tions for calculating the self-diffusion coefficient of liquid water21 orhandling anisotropic materials.22

III. FEYNMAN PATH INTEGRAL OPEN POLYMERLet us assume that the quantity of interest for a system is the

canonical time correlation function,

⟨AB(t)⟩ = 1ZTrA e−βH e

iHth B e−

iHth , (4)

where  and B are arbitrary operators, β is the inverse of Boltz-mann’s constant times the absolute temperature, Z is the partitionfunction, and Tr denotes a trace. The choice of placing the Boltz-mann operator after Â, instead of before, is arbitrary and only deter-mines the sign of the imaginary part of the correlation function,since we have,

1ZTr e−βHA e

iHth B e−

iHth = 1

Z( TrA e−βH e

iHth B e−

iHth )

∗. (5)

The trace can be written as an integral over the position, x1,eigenkets,

⟨AB(t)⟩ = 1Z ∫

−∞dx1 ⟨x1∣A e−βH e

iHth B e−

iHth ∣x1⟩. (6)

By dividing the Boltzmann operator e−βH into N factors e−βN H

and inserting N − 1 identity operators, 1 = ∫∞−∞dxj ∣xj⟩⟨xj∣, thecorrelation function can be written as

⟨AB(t)⟩ = 1Z

⎧⎪⎪⎨⎪⎪⎩

N∏j=1∫∞

−∞dxj⎫⎪⎪⎬⎪⎪⎭⟨x1∣A e−

βN H∣x2⟩

×⟨x2∣ e−βN H∣x3⟩ . . . ⟨xN−1∣ e−

βN H∣xN⟩

×⟨xN ∣ e−βN H e

iHth B e−

iHth ∣x1⟩, (7)

which can be recognized as a Feynman path integral23 in imag-inary time (−ihβ), and this in turn can be rewritten as Wignertransforms,

⟨AB(t)⟩ = 1Z

⎧⎪⎪⎨⎪⎪⎩

N∏j=1∫∞

−∞ ∫∞

−∞dxj dpj2πh

⎫⎪⎪⎬⎪⎪⎭e−

ih ∑N

j=1 pj(xj+1−xj)

×(A e−βN H)

W[x1 + x2

2, p1]( e−

βN H)

W[x2 + x3

2, p2]

. . .( e− βN H)

W[xN−1 + xN

2, pN−1]

×( e− βN H e

iHth B e−

iHth )

W[xN + x1

2, pN], (8)

where xN+1 = x1 so that the coordinates make a loop.

J. Chem. Phys. 152, 094111 (2020); doi: 10.1063/1.5126183 152, 094111-2

Published under license by AIP Publishing

III

Page 182: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

III

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

If βN is approaching 0 and only finite temperatures are of inter-

est, i.e., N is approaching infinity, then the Wigner transform ofthe Boltzmann operator is simply the classical Boltzmann factor. Inthis same limit, the Boltzmann operators can be separated from thetransforms involving  and B and thus make up their own Wignertransforms, without consequence,

⟨AB(t)⟩ = limN→∞

1Z

⎧⎪⎪⎨⎪⎪⎩

N∏j=1∫∞

−∞ ∫∞

−∞dxj dpj2πh

⎫⎪⎪⎬⎪⎪⎭× e−

ih ∑N

j=1 pj(xj+1−xj)(A)W[y1, p1]× e−

βN ∑N

j=1 H(yj ,pj)( e iHth B e−

iHth )

W[yN , pN], (9)

where yj = xj+xj+12 , yN = xN+x1

2 , and H(yj, pj) is the classicalHamiltonian. Now, assuming that the potential energy V(yj) isindependent of momentum, whereby most of the momenta onlyoccur in the kinetic energy term of the classical Hamiltonians,p2j2m , and the symplectic area, ∑N

j=1 pj(xj+1 − xj), it is easy to inte-grate those momenta out analytically. The momenta that occur inother places in Eq. (9) are p1 and pN . p1 can occur within theWigner transform of Â. Depending on how (A)W[y1, p1] dependson p1, the integration over p1 will turn out differently. As long as∫∞−∞dp1 (A)W[y1, p1] e−

βN H(y1,p1) e−

ih p1(x2−x1) can be evaluated ana-

lytically, e.g., for (A)W[y1, p1] being any polynomial of p1 (seeAppendix A), all momenta except pN can be integrated out ana-lytically. pN can usually not be integrated out analytically since if Bincludes either position or momentum, then ( e iHt

h B e−iHth )

W[yN , pN]

will depend on both, unless the potential is constant, and typicallythis dependence will be such that the integral is not easily evaluatedanalytically. Hence, these integrations lead to

⟨AB(t)⟩ = limN→∞

1Z(mN2πβ)

N2√

β2πmN

h−N⎧⎪⎪⎨⎪⎪⎩

N∏j=1∫∞

−∞dxj⎫⎪⎪⎬⎪⎪⎭

× ∫∞

−∞dpN e−

ih pN(x1−xN)A′(y1, x2 − x1)

× e−mN2h2β ∑

N−1j=1 (xj+1−xj)2 e−

βN ∑N

j=1 V(yj)

× e−βN

p2N2m ( e iHt

h B e−iHth )

W[yN , pN], (10)

where A′(y1, x2 − x1) is

A′(y1, x2 − x1) = ∫∞−∞ dp1 (A)W[y1, p1] e−

βN

p212m e−

ih p1(x2−x1)

∫ ∞−∞ dp1 e−βN

p212m e−

ih p1(x2−x1)

= ∫∞−∞ dp1 (A)W[y1, p1] e−

βN

p212m e−

ih p1(x2−x1)

√2πmN

β e−βN

mN2(x2−x1)22h2β2

(11)

and may depend on y1 and/or x2 − x1.

Finally, the Wigner transform of the time-evolved oper-ator B can be very difficult to derive. This is where theclassical Wigner method comes in through the approximation( e iHt

h B e−iHth )

W[yN , pN] ≈ (B)W[x(yN , pN , t), p(yN , pN , t)] [as in

Eq. (3)] giving the final expression,

⟨AB(t)⟩ ≈ 1Z(mN2πβ)

N2√

β2πmN

h−N⎧⎪⎪⎨⎪⎪⎩

N∏j=1∫∞

−∞dxj⎫⎪⎪⎬⎪⎪⎭

× ∫∞

−∞dpN e−

ih pN(x1−xN)

× e−βN (

p2N2m +∑N

j=1 V(yj)+ mN22h2β2 ∑

N−1j=1 (xj+1−xj)2)

×A′(y1, x2 − x1)(B)W[x(yN , pN , t), p(yN , pN , t)], (12)

where x(yN , pN , t) and p(yN , pN , t) are the position and momen-tum coordinates describing the classical trajectory starting with yNand pN , and where also for practical reasons the limit of N → ∞has been removed, thereby also making the value of the correlationfunction at t = 0 approximate. When referring to this expression, itwill be called ⟨AB(t)⟩y1 . It is noted that Eq. (12) is closely relatedto an expression recently published by Bose and Makri12 [Eq. (2.7)in the reference], but not further explored by them. The main dif-ference is that they use a Boltzmann operator that is symmetrizedaround Â, i.e., e−

β2 HA e−

β2 H . Equation (12) is also related to the L = 1

version of the method published by Bonella et al.,13 more clearlyseen in the paper by Bonella and Ciccotti.14 This method also han-dles symmetrized Boltzmann operators but uses sum and differencevariables for the polymer.

Apart from the asymmetric placement of the Boltzmann opera-tor explored in this paper and the symmetric placement explored inthe cited papers, another common handling of the Boltzmann oper-ator is the Kubo transform.24 This will not be further explored here;however, the Kubo transform of the new method can be found inAppendix B.

It can be noted that in Eq. (12), the effective equilibriumHamil-tonian has terms mN2

2h2β2 (xj+1 − xj)2. These look exactly like the poten-tial energy of harmonic springs and are, therefore, called “springterms.” This kind of model for a system is usually described as beadson a necklace, with the beads placed at the positions of the xj:s andconnected by springs. The ends of this polymer are connected via theimaginary exponential − i

h pN(x1 − xN), thus making it “open” in thesense that there is no force keeping the ends together, in contrast toother path integral methods such as path integral molecular dynam-ics,25 ring polymer molecular dynamics,3 or path integral MonteCarlo.26 Even such a method as the open chain imaginary time pathintegral of Cendagorta et al.27 would be considered a closed polymerfrom this point of view. A graphical representation of the open poly-mer in this work is shown in Fig. 1. This is the polymer advertised inthe title of this paper.

Looking back at Eq. (7), it can be seen that if  only depends onthe position, then can operate to the left and transform into a func-tion of x1, A(x1), thereby leaving a matrix element of the Boltzmannoperator. This operation removes the approximation of separatingthe Wigner transforms of  and e−

βN H , making this derivation less

approximate than the previous one. This means that a new version

J. Chem. Phys. 152, 094111 (2020); doi: 10.1063/1.5126183 152, 094111-3

Published under license by AIP Publishing

III

Page 183: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

III

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

FIG. 1. A drawing to exemplify a 7-bead open polymer, where the positions ofall the x:s and y:s have been marked. The springs represent the separationsthat give a spring-term, and the line is the separation that gives the imagi-nary exponential. y1 and y7 (in the circles) are where A′(y1, x2 − x1) and(B)

W[x(y7, p7, t), p(y7, p7, t)] are evaluated, respectively.

of Eq. (12) is

⟨AB(t)⟩ ≈ 1Z(mN2πβ)

N2√

β2πmN

h−N⎧⎪⎪⎨⎪⎪⎩

N∏j=1∫∞

−∞dxj⎫⎪⎪⎬⎪⎪⎭

× ∫∞

−∞dpN e−

ih pN(x1−xN)

× e−βN (

p2N2m +∑N

j=1 V(yj)+ mN22h2β2 ∑

N−1j=1 (xj+1−xj)2)

×A(x1)(B)W[x(yN , pN , t), p(yN , pN , t)], (13)

where the only change compared to before is that A′(y1, x2 − x1)has been replaced by A(x1). This expression will be referred to as⟨AB(t)⟩x1 . If  depends on both position and momentum, it is inmany cases possible to reorder operators so that all position depen-dence is to the left of all momentum dependence, and then let theposition dependent parts operate to the left in Eq. (7), keeping themomentum dependent parts in the matrix element. Of course thiswould lead to a more complicated expression than A(x1). Opera-tors of both position and momentum will, however, not be furtherdiscussed in this article

If  depends only on momentum, then Eq. (12) will be equiva-lent to Eq. (13) so that ⟨AB(t)⟩x1 = ⟨AB(t)⟩y1 . For this situation, thenotation ⟨AB(t)⟩y1,x1 will be used.

For simplicity of notation, the path integral open polymermethod will usually be called Open Polymer Classical Wigner, orOPCW for short, in this paper.

For multidimensional systems, the path integral open polymermethod for sampling initial conditions can potentially get a severesign problem due to the factor e−

ih pN(x1−xN) appearing in each degree

of freedom. The way this problem will be tackled in this article is tosample the initial distribution of the most quantum mechanical ormost important degrees of freedom by the path integral open poly-mer method and the other degrees of freedom by classical mechan-ics, coupling each bead in the quantum mechanical part(s) with thesingle bead in the classical parts.

IV. COMPUTATIONAL DETAILSEquations (12) and (13) were evaluated by Monte Carlo for the

integrals over xj and pN and with molecular dynamics for the timepropagation of (B)W[x(yN , pN , t), p(yN , pN , t)].

The correlation functions studied in this paper are autocor-relation functions of position, position-squared, momentum, andmomentum-squared for a one-dimensional quartic potential. Inaddition, a one-dimensional double well potential and a quarticpotential with many-dimensional harmonic baths were studied forthe position and position-squared autocorrelation functions.

A. Potentials and system parametersThe potentials studied in this work are a quartic potential, a

double well potential, and a quartic potential with various harmonicbaths.

The quartic potential was taken as

Vquartic(x) = m2ω3

4hx4, (14)

where ω is a unit of angular frequency.The double well potential was taken as

Vdouble well(x) = m2ω3

10hx4 − 1

2mω2x2. (15)

The multidimensional systems all use the quartic potentialtogether with the types of baths of Caldeira and Leggett.28 Thesesystems, thus, have one mostly quartic degree of freedom and sev-eral harmonic degrees of freedom that form a bath. The harmonicdegrees of freedom in the bath are bilinearly coupled to the quar-tic degree of freedom. The form of the bath is the one suggestedby Craig and Manolopoulos.29 The bath is a discretized version ofa bath with a linear spectral density with an exponential cutoff, andthe parameters chosen for the current work are a cutoff frequencyof ω and a system–bath coupling strength of mω. The completepotential is

VFD(x1, x2, . . . , xF−1, xF) = m2ω3

4hx41

+F∑l=2

12mω2(xl ln(

l − 32

F − 1) + x1√

2(F − 1)π)

2

, (16)

where F is the number of degrees of freedom.

B. Monte CarloThe maximum stepsize, Δxmax, l, was individually set for all

degrees of freedom l, with an initial value of

Δxmax,l =

2 ln 2

Fβ(∣∂2V∂x2l∣xl=0

+ ml(N−1)Nh2β2 )

. (17)

This choice for stepsize is based on a harmonic approximation ofthe potential energy, assuming the same average stepsize in xj andyj, and aiming for 50% acceptance rate. The change in energy thatwould give an acceptance likelihood of 50% for a Monte Carlo step

J. Chem. Phys. 152, 094111 (2020); doi: 10.1063/1.5126183 152, 094111-4

Published under license by AIP Publishing

III

Page 184: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

III

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

is N ln 2/β. 1/F of this energy could then be assigned to each degreeof freedom. For each degree of freedom, the maximum stepsize is setso that the total energy change resulting from a change in position ofΔxmax, l away from the minimum in the harmonically approximatedpotential, weighted by N, and the path integral spring potential,weighted by N − 1, would give this energy. This results in Eq. (17).The above does not work for a situation where ∣∂2V

∂x2l∣xl=0= 0 and at

the same time N = 1, and in such cases, the maximum stepsize wasset to the thermal de Broglie wavelength.

The maximal stepsizes were updated as a group every 50 000thstep according to the algorithm of Allen and Tildesley30 in order tokeep the acceptance rate close to 50%.

The momentum, pN , was sampled from a Maxwell–Boltzmanndistribution at the temperature N

kBβ , where kB is Boltzmann’s con-stant. The pseudo-random number generator used was ran2 of Presset al.31 Data were collected, i.e., a molecular dynamics trajectory wasrun, each 100th Monte Carlo-step.

The Monte Carlo chain begun in a part of phase space that hasa low probability of occurring, i.e., far from the equilibrium distribu-tion. It was, however, found that the number of Monte Carlo stepsit takes to come close to equilibrium for the various simulated sys-tems is negligible compared to the total length of the Monte Carlosimulation runs and was, therefore, not explicitly accounted for.

C. Molecular dynamicsThe molecular dynamics was conducted using the velocity Ver-

let algorithm.32,33 The time step was 0.050 ω−1 for the quartic anddouble well one-dimensional systems, and 0.035ω−1, 0.025ω−1, and0.020ω−1 for the quartic potential with harmonic baths containing3, 6, and 9 degrees of freedom, respectively. The total time length ofeach molecular dynamics run was 10ω−1.

D. Statistical evaluation of dataThe block average method, explained by, e.g., Friedberg and

Cameron34 and Flegal, Haran, and Jones,35 was used to calculatethe standard deviations of the correlation functions. The minimumblock size used was 106 Monte Carlo-steps. These standard devi-ations were used as a measure of uncertainty and to determineconvergence.

E. Exact correlation functionsIn order to have exact classical, quantum mechanical, and clas-

sical Wigner results to compare against, numerically exact resultswere produced for the systems and correlation functions where thiswas deemed doable.

For the quartic potential and double well potential, the classicalmechanics comparison was obtained by setting N = 1, in the sameprogram as was used to find the classical Wigner results, using 109Monte Carlo steps.

The quantum mechanical autocorrelation functions for thequartic potential were calculated with a numerically exact programthat uses the lowest 2000 particle in-the-box energy eigenfunc-tions, with a box length of 40

√hmω , as a basis set to approximate

the 40 lowest energy eigenfunctions of the quartic oscillator. These

eigenfunctions were then used to evaluate the necessary matrix ele-ments and could be propagated in time analytically. The quantummechanical autocorrelation functions for the double well potentialwere calculated in a similar way, but the basis set was the 12 lowestenergy eigenfunctions of the harmonic oscillator, V(x) = 1

2mω2x2.For the quartic potential with harmonic bath, just as without

the bath, a classical comparison was calculated with one bead, N = 1,in the classical Wigner routine, using 109 Monte Carlo steps. Thequantum mechanical comparison at t = 0 for this case was a pathintegral Monte Carlo simulation with 80 beads, N = 80, run withthe same Monte Carlo parameters as the classical Wigner runs.64 × 109 Monte Carlo steps were used for 3 degrees of freedom in thebath and 16 × 109 Monte Carlo steps for 6 and 9 degrees of freedomin the bath.

The exact classicalWigner data were generated as follows: First,the matrix elements of the Boltzmann operator were calculatedusing the numerical matrix multiplication scheme.36 In this scheme,N = 50 and N = 14 matrix multiplications of e−

βN H were used for

βhω = 8 and βhω = 1, respectively. Afterward, a numerical Fouriertransform of these data was used for computing the BoltzmannWigner transform. Finally, this Boltzmann Wigner function wasrepresented on a grid for doing classical dynamics.

F. Feynman–Kleinert classical Wigner methodMany of the methods for acquiring an approximate initial dis-

tribution for a classical Wigner calculation use a harmonic approxi-mation for the potential.19,37,38 However, for potentials with negativecurvature, these methods all encounter problems when the tem-perature is too low. For these situations, a modified local Gaussianapproximation39 can be used instead. This is, however, not necessaryfor the systems and temperatures in this paper.

The Feynman–Kleinert approximation40 is one of the harmonicapproximation methods, and its application to the classical Wignermethod is usually called Feynman–Kleinert linearized path integral(FK-LPI).37 FK-LPI will be used as a comparison for OPCW.

G. Ring polymer molecular dynamicsRing polymer molecular dynamics (RPMD)3 is a popu-

lar method for calculating approximate quantum dynamics. Thismethod was used as a comparison for the classical Wigner results.The RPMD results were generated by using 32 and 5 beads forβhω = 8 and βhω = 1, respectively, using a time step of 0.009ω−1 forthe quartic oscillator and a time step of 0.0125ω−1 for the double welland quartic oscillator with harmonic baths. The Kubo-transformedresults acquired from the calculations were transformed into theasymmetric placement of the Boltzmann operator through themethod by Braams, Miller, and Manolopoulos.41 The way to calcu-late momentum-correlation functions with RPMD is through timederivatives of position correlation functions.29,42 This was consid-ered too complicated for ⟨p2p2(t)⟩ so that this function has not beenincluded.

V. RESULTS AND DISCUSSIONIn this section, the results from the calculations of a few differ-

ent autocorrelation functions for a few model systems are presented.Each correlation function is presented by three lines. The middle

J. Chem. Phys. 152, 094111 (2020); doi: 10.1063/1.5126183 152, 094111-5

Published under license by AIP Publishing

III

Page 185: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

III

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

FIG. 2. The position autocorrelation func-tion for a quartic potential (βhω = 8).Comparison between different numbersof beads for the two versions of OPCWand numerically exact solutions for clas-sical mechanics (CM), classical Wigner(CW), and quantum mechanics (QM).The number of Monte Carlo steps usedfor each number of beads, N, is 1 × 109

for N = 10 and N = 40, and 16× 109 for N= 160. The outer lines of each type showthe standard deviations for the results. Ifthe standard deviation is small enough,the outer lines are not visible. (a) Realpart of ⟨xx(t)⟩y1 , (b) imaginary part of⟨xx(t)⟩y1 , (c) real part of ⟨xx(t)⟩x1 ,and (d) imaginary part of ⟨xx(t)⟩x1 .

line is the correlation function itself, and the upper and lower onesshow the standard deviation of the result. In most cases presentedhere, the standard deviations are within the width of the middleline.

A. Quartic potential βhω = 8When looking at the correlation functions for the quartic

potential calculated with the new method (Figs. 2–4) and numer-ically exact classical Wigner (Figs. 2 and 3), it can be seen thatthey flatten out and become constant at long times. This is due

to the fact that the classical Wigner method relies on classicalmechanics for propagation forward in time. For all systems withpotentials of higher order than harmonic, the classical trajecto-ries do not give the correct coherence, meaning that the individ-ual classical trajectories dephase against each other. Results for aone-dimensional harmonic oscillator, where classical propagationis exact, can be found in Appendix C. Analytic expressions for theharmonic oscillator autocorrelation functions have been placed inAppendix D.

In Fig. 2, it can be seen that at time t = 0, the correlationfunctions calculated with the new method converge from classical

FIG. 3. The position and position-squared autocorrelation functions for aquartic potential (βhω = 8). Comparisonbetween the two versions of OPCW andnumerically exact solutions for classicalmechanics (CM), classical Wigner (CW),FK-LPI, RPMD, and quantum mechan-ics (QM). The number of beads used inthe y1- and x1-calculations is N = 160,and the number of Monte Carlo steps is16 × 109. The outer lines of each typeshow the standard deviations for theresults. If the standard deviation is smallenough, the outer lines are not visi-ble. (a) Real part of ⟨xx(t)⟩, (b) imag-inary part of ⟨xx(t)⟩, (c) real part of⟨x2x2(t)⟩, and (d) imaginary part of⟨x2x2(t)⟩.

J. Chem. Phys. 152, 094111 (2020); doi: 10.1063/1.5126183 152, 094111-6

Published under license by AIP Publishing

III

Page 186: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

III

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

FIG. 4. The momentum and momentum-squared autocorrelation functions for a quartic potential (βhω = 8). Comparison between OPCW [with ⟨pp(t)⟩y1 and ⟨p2p2(t)⟩y1being identical to ⟨pp(t)⟩x1 and ⟨p2p2(t)⟩x1 , respectively] and numerically exact solutions for classical mechanics (CM), FK-LPI, RPMD, and quantum mechanics (QM).

The number of beads used in the calculations of ⟨pp(t)⟩y1 ,x1 is N = 160, and the number of Monte Carlo steps is 64 × 109. The number of beads used in the calculations of

⟨p2p2(t)⟩y1 ,x1 is N = 80, and the number of Monte Carlo steps is 128× 109. The outer lines of each type show the standard deviations for the results. If the standard deviation

is small enough, the outer lines are not visible. (a) Real part of ⟨pp(t)⟩, (b) imaginary part of ⟨pp(t)⟩, (c) real part of ⟨p2p2(t)⟩, and (d) imaginary part of ⟨p2p2(t)⟩.

FIG. 5. The position and position-squared autocorrelation functions for aquartic potential (βhω = 1). Comparisonbetween the two versions of OPCW andnumerically exact solutions for classicalmechanics (CM), classical Wigner (CW),FK-LPI, RPMD, and quantum mechan-ics (QM). The number of beads used inthe y1- and x1-calculations of ⟨xx(t)⟩ isN = 80, and the number of Monte Carlosteps is 16 × 109. The number of beadsused in the y1- and x1-calculations of⟨x2x2(t)⟩ is N = 80, and the number ofMonte Carlo steps is 64 × 109. The outerlines of each type show the standarddeviations for the results. If the standarddeviation is small enough, the outer linesare not visible. (a) Real part of ⟨xx(t)⟩,(b) imaginary part of ⟨xx(t)⟩, (c) realpart of ⟨x2x2(t)⟩, and (d) imaginary partof ⟨x2x2(t)⟩.

J. Chem. Phys. 152, 094111 (2020); doi: 10.1063/1.5126183 152, 094111-7

Published under license by AIP Publishing

III

Page 187: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

III

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

FIG. 6. The momentum and momentum-squared autocorrelation functions for a quartic potential (βhω = 1). Comparison between OPCW [with ⟨pp(t)⟩y1 and ⟨p2p2(t)⟩y1being identical to ⟨pp(t)⟩x1 and ⟨p2p2(t)⟩x1 , respectively] and numerically exact solutions for classical mechanics (CM), FK-LPI, RPMD, and quantum mechanics (QM).

The number of beads used in the calculations of ⟨pp(t)⟩y1 ,x1 is N = 80, and the number of Monte Carlo steps is 128 × 109. The number of beads used in the calculations

of ⟨p2p2(t)⟩y1 ,x1 is N = 20, and the number of Monte Carlo steps is 256 × 109. The outer lines of each type show the standard deviations for the results. If the standard

deviation is small enough, the outer lines are not visible. (a) Real part of ⟨pp(t)⟩y1 ,x1 , (b) imaginary part of ⟨pp(t)⟩y1 ,x1 , (c) real part of ⟨p2p2(t)⟩y1 ,x1 , and (d) imaginary

part of ⟨p2p2(t)⟩y1 ,x1 .

FIG. 7. The position and position-squared autocorrelation functions for adouble well potential (βhω = 8). Com-parison between the two versions ofOPCW and numerically exact solutionsfor classical mechanics (CM), classicalWigner (CW), FK-LPI, RPMD, and quan-tum mechanics (QM). The number ofbeads used in the y1- and x1-calculationsis N = 160, and the number of MonteCarlo steps is 16 × 109. The outer linesof each type show the standard devia-tions for the results. If the standard devi-ation is small enough, the outer lines arenot visible. (a) Real part of ⟨xx(t)⟩, (b)imaginary part of ⟨xx(t)⟩, (c) real partof ⟨x2x2(t)⟩, and (d) imaginary part of⟨x2x2(t)⟩.

J. Chem. Phys. 152, 094111 (2020); doi: 10.1063/1.5126183 152, 094111-8

Published under license by AIP Publishing

III

Page 188: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

III

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

mechanics toward quantum mechanics as the number of beads isincreased. For all times, the correlation functions calculated with thenew method converge toward the exact classical Wigner method. Itcan also be seen that the x1-method converges faster than the y1-method with respect to the number of beads. This is also the case forthe harmonic oscillator, see Appendix E, where an explanation forthis is also given.

Looking at the individual versions of the new method in Figs. 2and 3, it can be seen that at least the real parts of ⟨xx(t)⟩x1 and⟨x2x2(t)⟩x1 have converged essentially to within the thickness of theline of the exact classical Wigner result for N = 160. ⟨xx(t)⟩y1 and⟨x2x2(t)⟩y1 converge quite slowly in comparison, and the results forN = 160 are not entirely converged to the exact classical Wignerresult, even if they are close. ⟨pp(t)⟩y1,x1 in Figs. 4(a) and 4(b) hasnot converged all the way to exact quantum mechanics at t = 0 forN = 160 but is close.

In Figs. 4(c) and 4(d), it stands out that ⟨p2p2(t)⟩y1,x1 is far fromconverged to exact quantummechanics at t = 0, but this is a compli-cated correlation function. Even though the result for N = 80 is notvery close to exact quantum mechanics at time t = 0, the shapes ofthe curves have some qualitative agreement.

Correlation functions with pn in the first operator can beexpected to be more difficult to converge than correlation functionswith xn in the first operator, since (pn)W will be integrated into annth-order polynomial of a difference between positions, while (xn)Wwill just be a position, or an average of positions, to the power of n(see Appendix A).

In Fig. 3, it can be seen that FK-LPI gives results very closeto exact classical Wigner for ⟨xx(t)⟩ and the imaginary part of⟨x2x2(t)⟩, while the real part of ⟨x2x2(t)⟩ is a little bit further off.OPCW gives as good, or slightly better, results as FK-LPI, except y1for Im⟨x2x2(t)⟩.

In Figs. 4(a) and 4(b), it can be seen that FK-LPI gives a bet-ter starting value for ⟨pp(t)⟩ than the calculations with the newmethod. The oscillations are, however, qualitatively similar. For⟨p2p2(t)⟩ in Figs. 4(c) and 4(d), the results acquired with the newmethod have a better value at t = 0 and has more qualitative agree-ment with the exact quantum mechanical result than the FK-LPIresult.

It is shown in Figs. 3, 4(a), and 4(b) that the classical Wignermethod gives worse amplitudes than RPMD for the linear auto-correlation functions, ⟨xx(t)⟩ and ⟨pp(t)⟩. It can, however, alsobe seen that the classical Wigner method gives better amplitudeat short times and better phase overall than RPMD does for thenon-linear autocorrelation function ⟨x2x2(t)⟩. This is not surpris-ing as it is known that for RPMD to be exact for the harmonicoscillator at least one of the operators in a correlation function hasto be linear.3

B. Quartic potential βhω = 1When, for the quartic potential, the temperature is increased so

that βhω = 1, instead of βhω = 8, it is shown in Fig. 5 that ⟨xx(t)⟩and ⟨x2x2(t)⟩ for both versions of the new method are almost per-fectly converged to the exact classical Wigner result for N = 80. Thisis fewer beads than what seems necessary to achieve a similar con-vergence at βhω = 8. This is hardly surprising as the classical andquantum mechanical correlation functions are much more similarat βhω = 1 than at βhω = 8.

In Fig. 6, it can be seen that ⟨pp(t)⟩y1,x1 is rather well con-verged toward exact quantummechanics at t = 0. ⟨p2p2(t)⟩y1,x1 doesnot reach the limit of the exact quantum mechanics for the N = 20calculation presented here but shows a qualitative agreement withquantum mechanics up to t = 4ω−1

FIG. 8. The position autocorrelation func-tion for a quartic potential with a har-monic bath with 3 degrees of freedom(βhω = 8). Comparison between differ-ent numbers of beads for the two ver-sions of OPCW and numerically exactsolutions for classical mechanics (CM),at all times and quantum mechanics(QM), at time t = 0. The number of MonteCarlo steps used for each number ofbeads, N, is 1× 109 for N = 5 and N = 10,16 × 109 for N = 20, and 64 × 109 for N= 40. The outer lines of each type showthe standard deviations for the results. Ifthe standard deviation is small enough,the outer lines are not visible. (a) Realpart of ⟨xx(t)⟩y1 , (b) imaginary part of⟨xx(t)⟩y1 , (c) real part of ⟨xx(t)⟩x1 ,and (d) imaginary part of ⟨xx(t)⟩x1 .

J. Chem. Phys. 152, 094111 (2020); doi: 10.1063/1.5126183 152, 094111-9

Published under license by AIP Publishing

III

Page 189: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

III

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

Generally, even for a lower number of beads the convergencewith respect to the number ofMonte Carlo steps is worse for βhω = 1compared to βhω = 8. This is, however, not a problem in practice asfewer beads are required to converge the result to the exact classicalWigner result at the higher temperature.

For ⟨xx(t)⟩ and ⟨x2x2(t)⟩, FK-LPI results are essentially thesame as the exact classical Wigner results. For ⟨pp(t)⟩, the resultsfrom the new method and the FK-LPI results are very similar.However, for ⟨p2p2(t)⟩, the FK-LPI results are somewhat closerto the quantum mechanical result than the results produced withOPCW.

For ⟨xx(t)⟩, the classical Wigner method gives worse ampli-tudes than RPMD. For ⟨x2x2(t)⟩, the classical Wigner method pos-sibly gets a somewhat worse amplitude than RPMD. For ⟨pp(t)⟩,it is not obvious if one of the methods performs better than theother.

C. Double well potential βhω = 8In Fig. 7, it can be seen that all the approximate methods quite

early become very different from exact quantum mechanics. This isbecause dynamical tunneling is important for describing the dynam-ics of a system like this, which is not taken into account at all in theclassical Wigner method and not properly in RPMD. For ⟨xx(t)⟩,the y1- and x1-versions of the open polymer method have convergedto the exact classicalWigner result by using 160 beads. For ⟨x2x2(t)⟩,

the open polymermethod has almost converged to the exact classicalWigner result using 160 beads.

The FK-LPI results can be seen to essentially agree with theexact classical Wigner result for the real part of ⟨xx(t)⟩ and tobe close to it for the imaginary part of ⟨x2x2(t)⟩. For the othercases, FK-LPI is further off. For the real part of ⟨x2x2(t)⟩, FK-LPIis substantially off compared to exact classical Wigner. The FK-LPIresults do, however, stay almost equal to the exact quantummechanical results for as long as the exact classical Wignerresults do.

The classical Wigner method gives significantly better resultsthan RPMD for the real part of ⟨xx(t)⟩ and the imaginary part of⟨x2x2(t)⟩, and slightly better results for the real part of ⟨x2x2(t)⟩.For the imaginary part of ⟨xx(t)⟩, the classical Wigner method andRPMD perform equally well.

D. Quartic potential in harmonic bath βhω = 8In Figs. 8 and 9, the position and position-squared autocor-

relation functions are shown for the quartic oscillator with a har-monic bath of 3 degrees of freedom. The numerically exact quan-tum mechanical comparison is only available for time t = 0 dueto the method used for its calculation (see Sec. IV E). It can,in these figures, be seen that the new method converges fromclassical mechanics toward quantum mechanics at t = 0 as thenumber of beads is increased. The possible exception is for low

FIG. 9. The position-squared autocorrelation function for a quartic potential with a harmonic bath with 3 degrees of freedom (βhω = 8). Comparison between different numbersof beads for the two versions of OPCW and numerically exact solutions for classical mechanics (CM), at all times and quantum mechanics (QM), at time t = 0. The number ofMonte Carlo steps used for each number of beads, N, is 1 × 109 for N = 5 and N = 10, 16 × 109 for N = 20, and 64 × 109 for N = 40. The outer lines of each type show thestandard deviations for the results. If the standard deviation is small enough, the outer lines are not visible. The value in parentheses gives the number of Monte Carlo steps.(a) Real part of ⟨x2x2(t)⟩y1 , (b) imaginary part of ⟨x2x2(t)⟩y1 , (c) real part of ⟨x2x2(t)⟩x1 , and (d) imaginary part of ⟨x2x2(t)⟩x1 .

J. Chem. Phys. 152, 094111 (2020); doi: 10.1063/1.5126183 152, 094111-10

Published under license by AIP Publishing

III

Page 190: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

III

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

numbers of beads for y1, as forN = 5 Re⟨xx(t)⟩y1 is further from thequantum mechanical result than classical mechanics is.

In Figs. 10–12, the position and position-squared autocorrela-tion functions are shown for the quartic oscillator with a harmonicbath of 3, 6, and 9 degrees of freedom, respectively.

In Figs. 8, 9, 11, and 12, it can be seen that the x1-results aresignificantly closer to quantummechanics at t = 0 than the y1-resultsare.

For the case with 3 degrees of freedom in the bath, results areshown that are not entirely converged with respect to the number ofMonte Carlo steps used, and it is visible that the x1-version of themethod converges better with respect to the number of Monte Carlosteps than the y1-version.

Overall, the calculations for the larger numbers of degrees offreedom are fairly computationally intensive and have, therefore, notbeen numerically converged with respect to the number of beads.However, it is shown in Figs. 10–12 that when the harmonic bathis sampled from a classical distribution, much larger numbers ofbeads can be used in the quartic oscillator degree of freedom andthe number of Monte Carlo steps used is still the same or smallerthan used for a one-dimensional quartic potential with a lower num-ber of beads. Making the bath classical, thus, improves the overallconvergence drastically. The classical bath calculations are so wellconverged with regard to both numbers of beads and Monte Carlosteps that the difference between y1 and x1 is almost unnoticeablefor the correlation functions shown here, and therefore, only thex1-version is shown. The difference in numerical performance

between the full OPCW calculations and the calculations with clas-sical bath can be seen to increase when the size of the bath isincreased.

At time t = 0, the calculations employing a classical bath givebetter values than those using the open polymer for all degreesof freedom. However, even if the quartic oscillator part of a clas-sical bath calculation were to be sampled with an infinite num-ber of beads in the polymer, the classical mechanics of the bathwould still mean that the initial value of the correlation func-tions would not necessarily be the exact quantum mechanical value.The correlation functions for t > 0 for the classical bath calcula-tions are qualitatively similar to but have higher amplitudes thanthe full OPCW calculations. This is the behavior that would beexpected from a full OPCW calculation with a larger number ofbeads.

For the real part of ⟨x2x2(t)⟩, the long time value given by theclassical bath calculations is lower than the corresponding result forthe full OPCW calculations, except y1 for 9 degrees of freedom inthe bath. From the full OPCW results for the quartic oscillator with aharmonic bath with 3 degrees of freedom (Figs. 8 and 9) and the one-dimensional quartic oscillator (Fig. 2), it can be expected that for thesame number of beads in the open polymer, x1 will be more con-verged toward exact classical Wigner than y1 will be. For the quarticoscillator with harmonic baths with 6 and 9 degrees of freedom,Figs. 11 and 12, the results from calculations with classical bathsare closer to the y1-results than to the x1-results. This indicates thatthe long time values of Re⟨x2x2(t)⟩ may not be very well described

FIG. 10. The position and position-squared autocorrelation functions for a quartic potential with a harmonic bath with 3 degrees of freedom (βhω = 8). Comparison betweenthe x1-version of OPCW and numerically exact solutions for classical mechanics (CM), FK-LPI, and RPMD, at all times and quantum mechanics (QM), at time t = 0. Resultsfor the x1-version of OPCW with 320 beads in the quartic oscillator and a classical bath, CB, are also shown. The number of beads used in the y1- and x1-calculations is N =40, and the number of Monte Carlo steps is 64 × 109. For the calculations with classical bath, the number of Monte Carlo steps used is 16 × 109. The outer lines of each typeshow the standard deviations for the results. If the standard deviation is small enough, the outer lines are not visible. (a) Real part of ⟨xx(t)⟩, (b) imaginary part of ⟨xx(t)⟩,(c) real part of ⟨x2x2(t)⟩, and (d) imaginary part of ⟨x2x2(t)⟩.

J. Chem. Phys. 152, 094111 (2020); doi: 10.1063/1.5126183 152, 094111-11

Published under license by AIP Publishing

III

Page 191: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

III

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

FIG. 11. The position and position-squared autocorrelation functions for a quartic potential with a harmonic bath with 6 degrees of freedom (βhω = 8). Comparison betweenthe two versions of OPCW and numerically exact solutions for classical mechanics (CM), FK-LPI, and RPMD, at all times and quantum mechanics (QM), at time t = 0. Resultsfor the x1-version of OPCW with 320 beads in the quartic oscillator and a classical bath, CB, are also shown. The number of beads used in the y1- and x1-calculations is N =10, and the number of Monte Carlo steps is 64 × 109. For the calculations with classical bath, the number of Monte Carlo steps used is 16 × 109. The outer lines of each typeshow the standard deviations for the results. If the standard deviation is small enough, the outer lines are not visible. (a) Real part of ⟨xx(t)⟩, (b) imaginary part of ⟨xx(t)⟩,(c) real part of ⟨x2x2(t)⟩, and (d) imaginary part of ⟨x2x2(t)⟩.

with the classical bath. This was to be expected as the zero pointenergy leakage from the system into the bath should be consider-able in this type of calculation, and the correlation function, even atlong times, is highly dependent on the magnitude of the oscillationin the system degree of freedom.

It can be seen in Figs. 10–12 that the results from the classicalbath calculations and the FK-LPI results follow each other closely forthe first 2–4 ω−1. At t = 0, the classical bath results are as good as orslightly worse than the FK-LPI results.

In Figs. 10(a), 10(b), 11(a), 11(b), 12(a), and 12(b), it can be seenthat the classical Wigner method with a classical bath gives slightlyworse than or equally good results as RPMD does for ⟨xx(t)⟩ att = 0. Looking at Figs. 10(c), 10(d), 11(c), 11(d), 12(c), and 12(d),it can be seen that the classical Wigner method with a classical bathgives slightly better results than RPMD does for ⟨x2x2(t)⟩ at t = 0.RPMD, however, goes to higher values at long times for the real partof ⟨x2x2(t)⟩, and this may be a better value as it follows ⟨x2x2(t)⟩x1 ,which should be better converged than ⟨x2x2(t)⟩y1 .

E. Summary of resultsFor all the cases studied here, the results of the new method

converge toward exact quantummechanics at time t = 0 as the num-ber of beads increases. Additionally, for all cases where the exact clas-sical Wigner result is available, the new method converges towardthis result as the number of beads increases. These convergences arewhat should be observed according to the derivation of the method.

Some of the results for the one-dimensional quartic oscillator anddouble well have converged essentially to within the thickness of theline of the exact classical Wigner result.

For the correlation functions where a comparison has beenmade, the x1-version of the new method converges faster than they1-version with respect to the number of beads used. The x1-versionalso converges better than the y1-version with respect to the num-ber of Monte Carlo steps for these cases. Note also that for almostevery graph shown, it can be seen that for a larger number ofbeads, more Monte Carlo steps have been used to converge theresults.

The results for the multidimensional systems using many beadsin the polymer for sampling the initial distribution of the quarticoscillator degree of freedom and classical mechanics to sample theinitial distribution of the bath show a significant improvement inconvergence toward the exact result at t = 0 compared to the resultsof using fewer beads for the full OPCW quartic oscillator with har-monic bath. The long time values of Re⟨x2x2(t)⟩ may, however, besignificantly different from the exact classical Wigner result. This islikely to be a result of increased zero point energy leakage when aclassical bath is used.

If looking at the one-dimensional andmultidimensional poten-tials, using a classical bath for the multidimensional cases, the resultsfrom the open polymer sampled classical Wigner method is aboutas good as the results from FK-LPI. For the double well potential,the newmethod clearly reproduces the exact classical Wigner resultsbetter than FK-LPI, but the new method does not come closer to

J. Chem. Phys. 152, 094111 (2020); doi: 10.1063/1.5126183 152, 094111-12

Published under license by AIP Publishing

III

Page 192: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

III

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

FIG. 12. The position and position-squared autocorrelation functions for a quartic potential with a harmonic bath with 9 degrees of freedom (βhω = 8). Comparison betweenthe two versions of OPCW and numerically exact solutions for classical mechanics (CM), FK-LPI, and RPMD, at all times and quantum mechanics (QM), at time t = 0. Resultsfor the x1-version of OPCW with 320 beads in the quartic oscillator and a classical bath, CB, are also shown. The number of beads used in the y1- and x1-calculations is N =5, and the number of Monte Carlo steps is 16 × 109. For the calculations with classical bath, the number of Monte Carlo steps used is 16 × 109. The outer lines of each typeshow the standard deviations for the results. If the standard deviation is small enough, the outer lines are not visible. (a) Real part of ⟨xx(t)⟩, (b) imaginary part of ⟨xx(t)⟩,(c) real part of ⟨x2x2(t)⟩, and (d) imaginary part of ⟨x2x2(t)⟩.

exact quantummechanics than FK-LPI. Comparing Figs. 3, 5, and 7,it can be seen that the OPCWmethod works essentially equally wellindependent of the potential involved, while in the case of FK-LPI, itworks worse for the double well, which contains a region of negativecurvature.

In comparison to RPMD, it can be seen that for the one-dimensional quartic oscillator, the classical Wigner method andthereby also the OPCW method perform worse than, or in a singlecase equally well as, RPMD for autocorrelation functions of linearoperators. For the autocorrelation function ⟨x2x2(t)⟩, the classicalWigner method gives better results than RPMD at the lower temper-ature employed here and possibly worse results at the higher temper-ature. For the one-dimensional double well potential, the classicalWigner method is seen to give better results than RPMD for both⟨xx(t)⟩ and ⟨x2x2(t)⟩. Comparing classical Wigner with a classicalbath to RPMD for the multidimensional systems, the classical bathcalculations tend to be as good as or better than RPMD, apart fromthe long time values of ⟨x2x2(t)⟩.

VI. CONCLUSIONIn this article, two versions of a new way of sampling the ini-

tial quantum distribution used in the classical Wigner method forthe calculation of correlation functions have been presented andtested for the one-dimensional quartic oscillator and double well

and a quartic oscillator with linearly coupled harmonic baths. Thename used for the new method is Open Polymer Classical Wigner(OPCW).

The newmethod will always converge toward the exact classicalWigner result as the number of beads in the open polymer neck-lace goes to infinity. For the y1-version of the new method and thecorrelation functions and potentials tested here, this convergenceis mostly slow. For some cases, the x1-version of the new methodconverges considerably faster.

Compared to FK-LPI, the open polymer sampling for the clas-sical Wigner method can give better, worse, or equal results. Thedouble well potential is a case where the two methods give notice-ably different results, with the ones from OPCW being closer toexact classical Wigner. Both methods, however, follow exact quan-tum mechanics equally well for the double well potential. For a wellbehaved molecular potential, OPCW will always converge towardthe exact classical Wigner result as the calculation gets more refined.A harmonic approximation method such as FK-LPI will not nec-essarily converge toward the exact classical Wigner result for allpotential energy surfaces.

The way forward from this study would be to test the methoddeveloped here on other potentials and correlation functions. Oneset of correlation functions that are of chemical interest and thatpossibly could be calculated by the presented method are the ones ofMiller, Schwartz, and Tromp43,44 that can be used to acquire reactionrate constants. Potential energy surfaces in reaction rate calculations

J. Chem. Phys. 152, 094111 (2020); doi: 10.1063/1.5126183 152, 094111-13

Published under license by AIP Publishing

III

Page 193: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

III

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

tend to have barriers, such as the one in the double well tested here,and this is where the new method may be an improvement overmethods such as FK-LPI, since OPCW for the double well approxi-mates the exact classicalWigner result noticeably better than FK-LPIdoes.

Systems with many degrees of freedom are seen to be compu-tationally demanding. Describing the harmonic baths studied withclassical mechanics improves the situation considerably. It wouldthus be of interest to try this out on other multidimensional sys-tems. This should be particularly useful when the coupling betweenthe system and the bath is weak. It would also be of interest totry a less approximate simplification for the less quantum mechan-ical degrees of freedom in a system, such as an open polymerequivalent to the ring polymer contraction of Markland andManolopoulos.45 Another approach of interest for handling themore computationally demanding systems would be to try toenhance convergence using the techniques recently introduced byBose and Makri.12

ACKNOWLEDGMENTSFinancial support from the Swedish Research Council

(Vetenskapsrådet), Diary No. 2016-03275, is acknowledged.The computations in this work were performed at the Chalmers

Centre for Computational Science and Engineering (C3SE), a part-ner center of the Swedish National Infrastructure for Computing(SNIC).

APPENDIX A: ANALYTIC FORMS OF A ′(y 1 , x 2 − x 1 )FOR ( A)W[y1 p1 ] BEING A POLYNOMIAL IN p 1

If (A)W[y1, p1] is a polynomial with respect to p1, i.e.,

(A)W[y1, p1] = knpn1 + kn−1pn−11 . . . k2p21 + k1p1 + k0, (A1)

where kn, . . ., k0 are constants, then the solution to Eq. (11) is

A′(y1, x2 − x1) = ∫∞−∞ dp1 (A)W[y1, p1] e−

βN

p212m e−

ih p1(x2−x1)

√2πmN

β e−βN

mN2(x2−x1)22h2β2

= kn(−i)n(mN2β)

n2

Hn(√

mN2β

x2 − x1h) + kn−1(−i)n−1(mN

2β)

n−12

Hn−1(√

mN2β

x2 − x1h) . . .

+ k2(−i)2(mN2β)

22

H2(√

mN2β

x2 − x1h) + k1(−i)(mN

2β)

12

H1(√

mN2β

x2 − x1h) + k0

=n∑j=0

kj(−i)j(mN2β)

j2

Hj(√

mN2β

x2 − x1h), (A2)

where Hj(χ) is the Hermite polynomial defined by

Hj(χ) = (−1)je χ2 d j

dχje−χ

2, (A3)

where χ is a dummy variable.

APPENDIX B: KUBO TRANSFORMOne possible form of the Kubo transform24 of the open polymer expression presented in this article is

⟨AB(t)⟩Kubo =1ZTr 1

β ∫β

0dλ e−λHA e−(β−λ)H B(t)

≈ 1Z(mN2πβ)

N2√

β2πmN

h−N⎧⎪⎪⎨⎪⎪⎩

N∏j=1∫∞

−∞dxj⎫⎪⎪⎬⎪⎪⎭∫∞

−∞dpN e−

ih pN(x1−xN)

× e− β

N (p2N2m +∑N

j=1 V(yj)+ mN22h2β2 ∑

N−1j=1 (xj+1−xj)2)(B)W[x(yN , pN , t), p(yN , pN , t)]

× 1N(12A′(y1, x2 − x1) +

N−2∑k=1

A′(yk+1, xk+2 − xk+1) + 32A′(yN−1, xN − xN−1)). (B1)

J. Chem. Phys. 152, 094111 (2020); doi: 10.1063/1.5126183 152, 094111-14

Published under license by AIP Publishing

III

Page 194: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

III

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

It can be noted that the double counting of A′(yN−1, xN − xN−1) is due to the approximation (Ω e−βN H)

W[x, p] ≈ (Ω)W[x, p]( e−

βN H)

W[x, p]

≈ ( e− βN HΩ)

W[x, p]. This double counting leads to an asymmetry that means that the resulting correlation function may have an imaginary

part. If using the Kubo transform, this may be an unwanted property, so the mean of the above expression and its complex conjugate may beused instead,

⟨AB(t)⟩Kubo ≈1Z(mN2πβ)

N2√

β2πmN

h−N⎧⎪⎪⎨⎪⎪⎩

N∏j=1∫∞

−∞dxj⎫⎪⎪⎬⎪⎪⎭∫∞

−∞dpN e

− βN (

p2N2m +∑N

j=1 V(yj)+ mN22h2β2 ∑

N−1j=1 (xj+1−xj)2)(B)W[x(yN , pN , t), p(yN , pN , t)]

× 12N( e− i

h pN(x1−xN)(12A′(y1, x2 − x1) +

N−1∑k=2

A′(yk, xk+1 − xk) + 32A′(yN−1, xN − xN−1))

+ eih pN(x1−xN)(1

2A′(yN−1, xN−1 − xN) +

N−2∑k=1

A′(yk, xk − xk+1) + 32A′(y1, x1 − x2))), (B2)

which should not give an imaginary part. If A′(yj, xj+1 − xj) is either an even or odd function with regard to xj+1 − xj, the expression can besimplified to

⟨AB(t)⟩Kubo, even ≈1Z(mN2πβ)

N2√

β2πmN

h−N⎧⎪⎪⎨⎪⎪⎩

N∏j=1∫∞

−∞dxj⎫⎪⎪⎬⎪⎪⎭∫∞

−∞dpN e

− βN (

p2N2m +∑N

j=1 V(yj)+ mN22h2β2 ∑

N−1j=1 (xj+1−xj)2)(B)W[x(yN , pN , t), p(yN , pN , t)]

× 1N(cos(pN(x1 − xN)

h)(1

2A′(y1, x2 − x1) +

N−2∑k=2

A′(yk, xk+1 − xk) + 12A′(yN−1, xN − xN−1))

+ eih pN(x1−xN)A′(y1, x2 − x1) + e−

ih pN(x1−xN)A′(yN−1, xN − xN−1)) (B3)

for even A′ and

⟨AB(t)⟩Kubo, odd ≈1Z(mN2πβ)

N2√

β2πmN

h−N⎧⎪⎪⎨⎪⎪⎩

N∏j=1∫∞

−∞dxj⎫⎪⎪⎬⎪⎪⎭∫∞

−∞dpN e

− βN (

p2N2m +∑N

j=1 V(yj)+ mN22h2β2 ∑

N−1j=1 (xj+1−xj)2)

×(B)W[x(yN , pN , t), p(yN , pN , t)]

× 1N(−i sin(pN(x1 − xN)

h)(1

2A′(y1, x2 − x1) +

N−2∑k=2

A′(yk, xk+1 − xk) + 12A′(yN−1, xN − xN−1))

− eih pN(x1−xN)A′(y1, x2 − x1) + e−

ih pN(x1−xN)A′(yN−1, xN − xN−1)) (B4)

for odd A′.

APPENDIX C: RESULTS FOR HARMONICPOTENTIAL βhω = 8

Apart from the calculations shown in themain body of this arti-cle, the four autocorrelation functions position, position-squared,momentum, and momentum-squared were also calculated for aone-dimensional harmonic oscillator.

The harmonic potential was taken as

Vharmonic(x) = 12mω2x2, (C1)

where ω is the angular frequency of the harmonic oscillation.

TheMonte Carlo procedure andmolecular dynamics were con-ducted as for the other systems. The time step used in the moleculardynamics was 0.050ω−1.

In order to have exact values to compare our calculatedresults against, anaytical classical and quantum mechanical cor-relation functions were used. These functions are collected inAppendix D.

As shown in Fig. 13 for both versions of the new method,the real and imaginary parts of the position autocorrelation func-tion converge from exact classical mechanics toward exact quan-tum mechanics as the number of beads increases. ⟨xx(t)⟩x1 con-verges toward quantum mechanics, with respect to the number of

J. Chem. Phys. 152, 094111 (2020); doi: 10.1063/1.5126183 152, 094111-15

Published under license by AIP Publishing

III

Page 195: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

III

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

FIG. 13. The position autocorrelationfunction for a harmonic potential (βhω= 8). Comparison between different num-bers of beads for the two versions ofOPCW and exact solutions for classicalmechanics (CM) and quantum mechan-ics (QM). The number of Monte Carlosteps used for each number of beads,N, is 1 × 109 for N = 5, N = 10, andN = 40, and 64 × 109 for N = 160.The standard deviations are in all casessmall enough not to be visible. (a) Realpart of ⟨xx(t)⟩y1 , (b) imaginary part of⟨xx(t)⟩y1 , (c) real part of ⟨xx(t)⟩x1 ,and (d) imaginary part of ⟨xx(t)⟩x1 .

beads, noticeably faster than ⟨xx(t)⟩y1 . This ordering of speed ofconvergence with regard to the number of beads,N, is what could beexpected since ⟨xx(t)⟩x1 requires the positions of neighboring beadsto converge to the same value, while ⟨xx(t)⟩y1 also requires the posi-tions of next neighboring beads to converge to the same value (seeAppendix E 1).

In Fig. 14, the position-squared autocorrelation function canbe seen for both versions of the method studied. Both versionsof the method, just as for the previous correlation function, con-verge from classical toward quantum mechanics as the number

of beads increases. Similar to the previous correlation function,⟨x2x2(t)⟩x1 converges faster with respect to the number of beadsthan ⟨x2x2(t)⟩y1 , as could be expected (see Appendix E 2).

In Fig. 15, the position and position-squared autocorrelationfunctions from the two versions of the open polymer method canbe compared to each other, exact classical mechanics, exact quan-tum mechanics, and RPMD. FK-LPI is always exact for a harmonicpotential, so it is equivalent to exact quantum mechanics. As RPMDis exact for correlation functions with at least one linear operatorin a harmonic potential,3 the exact RPMD result is also equivalent

FIG. 14. The position-squared autocor-relation function for a harmonic potential(βhω = 8). Comparison between differ-ent numbers of beads for the two ver-sions of OPCW and exact solutions forclassical mechanics (CM) and quantummechanics (QM). The number of MonteCarlo steps used for each number ofbeads, N, is 1× 109 for N = 5 and N = 10,4 × 109 for N = 40, and 64 × 109

for N = 160. The outer lines of eachtype show the standard deviations for theresults. If the standard deviation is smallenough, the outer lines are not visible.(a) Real part of ⟨x2x2(t)⟩y1 , (b) imagi-

nary part of ⟨x2x2(t)⟩y1 , (c) real part of

⟨x2x2(t)⟩x1 , and (d) imaginary part of

⟨x2x2(t)⟩x1 .

J. Chem. Phys. 152, 094111 (2020); doi: 10.1063/1.5126183 152, 094111-16

Published under license by AIP Publishing

III

Page 196: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

III

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

FIG. 15. The position and position-squared autocorrelation functions for aharmonic potential (βhω = 8). Compar-ison between the two versions of OPCWand numerically exact solutions for clas-sical mechanics (CM), classical Wigner(CW), RPMD, and quantum mechanics(QM). The number of beads used inthe y1- and x1-calculations is N = 160,and the number of Monte Carlo steps is64 × 109. The outer lines of each typeshow the standard deviations for theresults. If the standard deviation is smallenough, the outer lines are not visi-ble. (a) Real part of ⟨xx(t)⟩, (b) imag-inary part of ⟨xx(t)⟩, (c) real part of⟨x2x2(t)⟩, and (d) imaginary part of⟨x2x2(t)⟩.

to exact quantum mechanics for ⟨xx(t)⟩. For ⟨x2x2(t)⟩, it can beseen that the classical Wigner method gives better results thanRPMD. This is to be expected as the classical Wigner method isexact for any correlation function for a harmonic potential, while

RPMD is not exact when both operators in the correlation functionare non-linear.

In Fig. 16, the momentum and momentum-squared autocor-relation functions are shown. For these correlation functions, the

FIG. 16. The momentum and momentum-squared autocorrelation functions for a harmonic potential (βhω = 8). Comparison between different numbers of beads for OPCW[with ⟨pp(t)⟩y1 and ⟨p2p2(t)⟩y1 being identical to ⟨pp(t)⟩x1 and ⟨p2p2(t)⟩x1 , respectively] and exact solutions for classical mechanics (CM) and quantum mechanics

(QM). The number of Monte Carlo steps used for each number of beads, N, for the calculation of ⟨pp(t)⟩y1 ,x1 is 1 × 109 for N = 5 and N = 10, 4 × 109 for N = 40, and

64 × 109 for N = 160. The number of Monte Carlo steps used for each number of beads for the calculation of ⟨p2p2(t)⟩y1 ,x1 is 1 × 109 for N = 5, 4 × 109 for N = 20, and

128 × 109 for N = 80. The outer lines of each type show the standard deviations for the results. If the standard deviation is small enough, the outer lines are not visible. (a)Real part of ⟨pp(t)⟩y1 ,x1 , (b) imaginary part of ⟨pp(t)⟩y1 ,x1 , (c) real part of ⟨p2p2(t)⟩y1 ,x1 , and (d) imaginary part of ⟨p2p2(t)⟩y1 ,x1 .

J. Chem. Phys. 152, 094111 (2020); doi: 10.1063/1.5126183 152, 094111-17

Published under license by AIP Publishing

III

Page 197: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

III

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

two versions of the method are identical. ⟨pp(t)⟩y1,x1 converges in asimilar way as ⟨xx(t)⟩y1 does with respect to the number of beads.⟨p2p2(t)⟩y1,x1 , as all the other correlation functions, converges withrespect to the number of beads from classical toward quantummechanics. In this work, ⟨p2p2(t)⟩y1,x1 is the most difficult corre-lation function to converge with respect to the number of MonteCarlo steps. That is why no results from calculations with N = 160are shown, and why the standard deviations are visible in the resultsfor N = 80.

APPENDIX D: ANALYTIC CORRELATION FUNCTIONSFOR THE HARMONIC OSCILLATOR

For the harmonic oscillator described in Appendix C, correla-tion functions can be calculated analytically. ⟨xx(t)⟩ and ⟨pp(t)⟩ arestraightforward to derive. ⟨x2x2(t)⟩ and ⟨p2p2(t)⟩ can be acquiredfrom the simpler correlation functions by using the cumulant expan-sion of Cao and Voth.46 The correlation functions are, for classicalmechanics,

⟨xx(t)⟩ = 1βmω2 cos(ωt), (D1)

⟨x2x2(t)⟩ = 1β2m2ω4 (1 + 2 cos2(ωt)), (D2)

⟨pp(t)⟩ = mβcos(ωt), (D3)

⟨p2p2(t)⟩ = m2

β2(1 + 2 cos2(ωt)), (D4)

and, for quantum mechanics,

⟨xx(t)⟩ = h2mω

( eβhω

eβhω − 1 e−iωt + 1eβhω − 1 eiωt), (D5)

⟨x2x2(t)⟩ = h2

4m2ω2⎛⎝(

eβhω + 1eβhω − 1)

2

+ 2( eβhω

eβhω − 1 e−iωt + 1eβhω − 1 eiωt)

2⎞⎠, (D6)

⟨pp(t)⟩ = mhω2( eβhω

eβhω − 1 e−iωt + 1eβhω − 1 eiωt), (D7)

⟨p2p2(t)⟩ = m2h2ω2

4⎛⎝(

eβhω + 1eβhω − 1)

2

+ 2( eβhω

eβhω − 1 e−iωt + 1eβhω − 1 eiωt)

2⎞⎠. (D8)

APPENDIX E: COMPARISON OF THE REALAND IMAGINARY PARTS OF CORRELATIONFUNCTIONS FOR THE HARMONIC POTENTIAL

For the harmonic oscillator, the analytical equations ofmotion can be put into (B)W[x(yN , pN , t), p(yN , pN , t)] and

(B)W[x(xN , pN , t), p(xN , pN , t)] in Eqs. (12) and (13). This appendixshows how some autocorrelation functions behave for the twoversions of the method presented in this paper.

1 . Position autocorrelation function, ⟨ xx( t) ⟩For the case of ⟨xx(t)⟩, entering the analytical equations of

motion into Eqs. (12) and (13) leads to

⟨xx(t)⟩y1 =1Z(mN2πβ)

N2√

β2πmN

h−N⎧⎪⎪⎨⎪⎪⎩

N∏j=1∫∞

−∞dxj⎫⎪⎪⎬⎪⎪⎭

× ∫∞

−∞dpN e−

ih pN(x1−xN)

× e− β

N (p2N2m +∑N

j=1 V(yj)+ mN22h2β2 ∑

N−1j=1 (xj+1−xj)2)

× y1(yN cos(ωt) + pNmω

sin(ωt)), (E1)

⟨xx(t)⟩x1 =1Z(mN2πβ)

N2√

β2πmN

h−N⎧⎪⎪⎨⎪⎪⎩

N∏j=1∫∞

−∞dxj⎫⎪⎪⎬⎪⎪⎭

× ∫∞

−∞dpN e−

ih pN(x1−xN)

× e−βN (

p2N2m +∑N

j=1 V(yj)+ mN22h2β2 ∑

N−1j=1 (xj+1−xj)2)

× x1(yN cos(ωt) + pNmω

sin(ωt)). (E2)

To simplify, pN can be integrated out and all constants that areidentical in both cases can be collected into a single constant, C,

⟨xx(t)⟩y1 = C⎧⎪⎪⎨⎪⎪⎩

N∏j=1∫∞

−∞dxj⎫⎪⎪⎬⎪⎪⎭e−

βN (∑N

j=1 V(yj)+ mN22h2β2 ∑

Nj=1(xj+1−xj)2)

× y1(yN cos(ωt) − iNβhω(x1 − xN) sin(ωt)), (E3)

⟨xx(t)⟩x1 = C⎧⎪⎪⎨⎪⎪⎩

N∏j=1∫∞

−∞dxj⎫⎪⎪⎬⎪⎪⎭e−

βN (∑N

j=1 V(yj)+ mN22h2β2 ∑

Nj=1(xj+1−xj)2)

× x1(yN cos(ωt) − iNβhω(x1 − xN) sin(ωt)). (E4)

These correlation functions are Boltzmann-weighted averages,which can be simplified to

⟨xx(t)⟩y1 = ⟨y1(yN cos(ωt) − iNβhω(x1 − xN) sin(ωt))⟩, (E5)

⟨xx(t)⟩x1 = ⟨x1(yN cos(ωt) − iNβhω(x1 − xN) sin(ωt))⟩. (E6)

Now, the correlation functions can be separated into the realparts,

Re⟨xx(t)⟩y1 = cos(ωt)⟨y1yN⟩

= cos(ωt)4(⟨x21⟩ + ⟨x1xN⟩ + ⟨x1x2⟩ + ⟨x2xN⟩)

= cos(ωt)4(⟨x21⟩ + 2⟨x1xN⟩ + ⟨x2xN⟩), (E7)

J. Chem. Phys. 152, 094111 (2020); doi: 10.1063/1.5126183 152, 094111-18

Published under license by AIP Publishing

III

Page 198: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

III

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

Re⟨xx(t)⟩x1 = cos(ωt)⟨x1yN⟩

= cos(ωt)2(⟨x21⟩ + ⟨x1xN⟩), (E8)

and the imaginary parts,

Im⟨xx(t)⟩y1 = −Nβhω

sin(ωt)⟨y1(x1 − xN)⟩

= − N2βhω

sin(ωt)(⟨x21⟩ + ⟨x1x2⟩ − ⟨x1xN⟩ − ⟨x2xN⟩)

= − N2βhω

sin(ωt)(⟨x21⟩ − ⟨x2xN⟩), (E9)

Im⟨xx(t)⟩x1 = −Nβhω

sin(ωt)⟨x1(x1 − xN)⟩

= − Nβhω

sin(ωt)(⟨x21⟩ − ⟨x1xN⟩), (E10)

where it has been used that all beads are equivalent after thelast momentum was integrated out, so, e.g., ⟨x2N⟩ = ⟨x21⟩ and⟨x1xN⟩ = ⟨x1x2⟩.

For the real part, ⟨x21⟩ cos(ωt) is the exact quantum mechan-ics, apart from that the Boltzmann weighting factor in the averageis approximate as long as N is finite. Re⟨xx(t)⟩x1 consists to a largerdegree of ⟨x21⟩ cos(ωt) than Re⟨xx(t)⟩y1 does. When N → ∞ andthe distance between beads becomes smaller both ⟨x1xN⟩ and ⟨x2xN⟩will converge toward ⟨x21⟩. ⟨x2xN⟩ will most likely converge moreslowly than ⟨x1xN⟩ as it depends on next neighboring beads insteadof immediately neighboring beads. This means that Re⟨xx(t)⟩x1could be expected to convergence toward exact quantum mechanicsfaster than Re⟨xx(t)⟩y1 with respect to the number of beads. Withthe same kind of reasoning, Im⟨xx(t)⟩x1 could be expected to con-verge faster with respect to the number of beads than Im⟨xx(t)⟩y1 ,since the former depends on ⟨x1xN⟩ and the latter dependson ⟨x2xN⟩.

2. Comparison of real and imaginary partsof ⟨ x2x2( t) ⟩ for the harmonic potential

For ⟨x2x2(t)⟩, the equivalent of Eqs. (E5) and (E6) is

⟨x2x2(t)⟩y1 = ⟨y21(y2N cos2(ωt) − iN

βhω(x1 − xN)yN sin(ωt) cos(ωt) − N2

β2h2ω2 (x1 − xN)2 sin2(ωt) + N

βω2msin2(ωt))⟩, (E11)

⟨x2x2(t)⟩x1 = ⟨x21(y2N cos2(ωt) − iN

βhω(x1 − xN)yN sin(ωt) cos(ωt) − N2

β2h2ω2 (x1 − xN)2 sin2(ωt) + N

βω2msin2(ωt))⟩. (E12)

Separating into real parts

Re⟨x2x2(t)⟩y1 = ⟨y21y2N cos2(ωt) − y21 N2

β2h2ω2 (x1 − xN)2 sin2(ωt) + y21

Nβω2m

sin2(ωt)⟩

= cos2(ωt)16

⟨x41 + 2x31x2 + 2x31xN + x21x22 + 4x21x2xN + x21x2N + 2x1x22xN + 2x1x2x2N + x22x2N⟩

− N2 sin2(ωt)4β2h2ω2 ⟨x

41 + 2x31x2 − 2x31xN + x21x22 − 4x21x2xN + x21x2N − 2x1x22xN + 2x1x2x2N + x22x2N⟩

+ N sin2(ωt)4βω2m

⟨x21 + 2x1x2 + x22⟩

= cos2(ωt)16

(⟨x41⟩ + 4⟨x31xN⟩ + 2⟨x21x2N⟩ + 4⟨x21x2xN⟩ + 4⟨x1x2x2N⟩ + ⟨x22x2N⟩)

− N2 sin2(ωt)4β2h2ω2 (⟨x

41⟩ + 2⟨x21x2N⟩ − 4⟨x21x2xN⟩ + ⟨x22x2N⟩) + N sin2(ωt)

2βω2m(⟨x21⟩ + ⟨x1xN⟩), (E13)

Re⟨x2x2(t)⟩x1 = ⟨x21y2N cos2(ωt) − x21 N2

β2h2ω2 (x1 − xN)2 sin2(ωt) + x21

Nβω2m

sin2(ωt)⟩

= cos2(ωt)4

(⟨x41⟩ + 2⟨x31xN⟩ + ⟨x21x2N⟩) − N2 sin2(ωt)β2h2ω2 (⟨x41⟩ − 2⟨x31xN⟩ + ⟨x21x2N⟩) + N sin2(ωt)

βω2m⟨x21⟩ (E14)

and imaginary parts

J. Chem. Phys. 152, 094111 (2020); doi: 10.1063/1.5126183 152, 094111-19

Published under license by AIP Publishing

III

Page 199: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

III

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

Im⟨x2x2(t)⟩y1 = −Nβhω

sin(ωt) cos(ωt)⟨y21(x1 − xN)yN⟩

= − N8βhω

sin(ωt) cos(ωt)(⟨x41⟩ + 2⟨x31x2⟩

+ ⟨x21x22⟩ − ⟨x21x2N⟩ − 2⟨x1x2x2N⟩ − ⟨x22x2N⟩)= − N

8βhωsin(ωt) cos(ωt)(⟨x41⟩ + 2⟨x31x2⟩

− 2⟨x1x2x2N⟩ − ⟨x22x2N⟩), (E15)

Im⟨x2x2(t)⟩x1 = −Nβhω

sin(ωt) cos(ωt)⟨x21(x1 − xN)yN⟩

= − N2βhω

sin(ωt) cos(ωt)(⟨x41⟩ − ⟨x21x2N⟩). (E16)

From these expressions, it can be seen that ⟨x2x2(t)⟩x1 for boththe real and imaginary parts is a combination of fewer and lesscomplex averages than ⟨x2x2(t)⟩y1 . Less complex in this case meansaverages of fewer different positions and of positions closer to eachother. Thus, it can be expected that ⟨x2x2(t)⟩x1 converges faster withregard to the number of beads than ⟨x2x2(t)⟩y1 does.

REFERENCES1 J. Cao and G. A. Voth, J. Chem. Phys. 99, 10070 (1993).2S. Jang and G. A. Voth, J. Chem. Phys. 111, 2357 (1999).3I. R. Craig and D. E. Manolopoulos, J. Chem. Phys. 121, 3368 (2004).4W. H. Miller, J. Chem. Phys. 53, 3578 (1970).5W. H. Miller, J. Phys. Chem. A 105, 2942 (2001).6T. J. H. Hele, M. J. Willatt, A. Muolo, and S. C. Althorpe, J. Chem. Phys. 142,134103 (2015).7E. J. Heller, J. Chem. Phys. 65, 1289 (1976).8H. Wang, X. Sun, and W. H. Miller, J. Chem. Phys. 108, 9726 (1998).9T. J. H. Hele, M. J. Willatt, A. Muolo, and S. C. Althorpe, J. Chem. Phys. 142,191101 (2015).1 0D. R. Reichman, P.-N. Roy, S. Jang, and G. A. Voth, J. Chem. Phys. 113, 919(2000).1 1 X. Sun, H. Wang, and W. H. Miller, J. Chem. Phys. 109, 4190 (1998).1 2A. Bose and N. Makri, J. Phys. Chem. A 123, 4284 (2019).

1 3S. Bonella, M. Monteferrante, C. Pierleoni, and G. Ciccotti, J. Chem. Phys. 133,164105 (2010).1 4S. Bonella and G. Ciccotti, Entropy 16, 86 (2013).1 5E. Wigner, Phys. Rev. 40, 749 (1932).1 6J. E. Moyal, Math. Proc. Cambridge Philos. Soc. 45, 99 (1949).1 7J. Liu, Int. J. Quant. Chem. 115, 657 (2015).1 8J. A. Poulsen, J. Scheers, G. Nyman, and P. J. Rossky, Phys. Rev. B 75, 224505(2007).1 9Q. Shi and E. Geva, J. Phys. Chem. A 107, 9059 (2003).20Q. Shi and E. Geva, J. Phys. Chem. A 107, 9070 (2003).21 S. Habershon and D. E. Manolopoulos, J. Chem. Phys. 131, 244518 (2009).22N. Markovic and J. A. Poulsen, J. Phys. Chem. A 112, 1701 (2008).23R. P. Feynman and A. R. Hibbs, Quantum Mechanics and Path Integrals(McGraw-Hill Companies, Inc., New York, 1965).24R. Kubo, J. Phys. Soc. Jpn. 12, 570 (1957).25M. Parrinello and A. Rahman, J. Chem. Phys. 80, 860 (1984).26J. A. Barker, J. Chem. Phys. 70, 2914 (1979).27J. R. Cendagorta, Z. Bacic, and M. E. Tuckerman, J. Chem. Phys. 148, 102340(2018).28A. O. Caldeira and A. J. Leggett, Ann. Phys. 149, 374 (1983).29I. R. Craig and D. E. Manolopoulos, J. Chem. Phys. 122, 084106 (2005).30M. P. Allen and D. J. Tildesley, Computer Simulation of Liquids (OxfordUniversity Press, Oxford, UK, 1987).31 W. H. Press, S. A. Teukolsky, W. T. Vetterling, and B. P. Flannery, Numeri-cal Recipes in FORTRAN: The Art of Scientific Computing, 2nd ed. (CambridgeUniversity Press, Cambridge, UK, 1992).32L. Verlet, Phys. Rev. 159, 98 (1967).33W. C. Swope, H. C. Andersen, P. H. Berens, and K. R. Wilson, J. Chem. Phys.76, 637 (1982).34R. Friedberg and J. E. Cameron, J. Chem. Phys. 52, 6049 (1970).35J. M. Flegal, M. Haran, and G. L. Jones, Stat. Sci. 23, 250 (2008).36D. Thirumalai, E. J. Bruskin, and B. J. Berne, J. Chem. Phys. 79, 5063 (1983).37J. A. Poulsen, G. Nyman, and P. J. Rossky, J. Chem. Phys. 119, 12179 (2003).38J. Liu and W. H. Miller, J. Chem. Phys. 125, 224104 (2006).39J. Liu and W. H. Miller, J. Chem. Phys. 131, 074113 (2009).40R. P. Feynman and H. Kleinert, Phys. Rev. A 34, 5080 (1986).41 B. J. Braams, T. F. Miller, and D. E. Manolopoulos, Chem. Phys. Lett. 418, 179(2006).42T. F. Miller and D. E. Manolopoulos, J. Chem. Phys. 122, 184503 (2005).43W. H. Miller, S. D. Schwartz, and J. W. Tromp, J. Chem. Phys. 79, 4889 (1983).44W. H. Miller, J. Chem. Phys. 61, 1823 (1974).45T. E. Markland and D. E. Manolopoulos, J. Chem. Phys. 129, 024105 (2008).46J. Cao and G. A. Voth, J. Chem. Phys. 100, 5106 (1994).

J. Chem. Phys. 152, 094111 (2020); doi: 10.1063/1.5126183 152, 094111-20

Published under license by AIP Publishing

III

Page 200: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

III

III

Page 201: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

IV

Paper IV

Calculation of ReactionRate Constants From aClassical Wigner ModelBased on a Feynman PathIntegral Open Polymer

179

IV

Page 202: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

IV

Calculation of Reaction Rate Constants From a Classical Wigner ModelBased on a Feynman Path Integral Open Polymer

S. Karl-Mikael Svensson,1 Jens Aage Poulsen,1, a) and Gunnar Nyman1, b)

Department of Chemistry and Molecular Biology, University of Gothenburg, SE 405 30 Gothenburg,Sweden

(Dated: 3 May 2020)

The Open Polymer Classical Wigner (OPCW) method [J. Chem. Phys. 152, 094111 (2020)] is applied to aflux-Heaviside correlation function used to calculate reaction rate constants. The obtained expression is testedon a parabolic barrier and a symmetric Eckart potential. The OPCW method is shown to converge towardthe exact classical Wigner result. The OPCW method shows some promise for rate constant calculations andsuggestions for future tests are made.

I. INTRODUCTION

Reaction rates are of central importance in chemistryand methods to calculate reaction rate constants aretherefore of interest. In many cases classical mechanics isenough to calculate a rate constant, but for light atomssuch as hydrogen or cold low density situations quantummechanics can be essential.Methods to acquire reaction rate constants from

a Wigner distribution include, to some extent, thetransition state theory with quantum partition func-tions of Wigner1 and Eyring2, the quantum transi-tion state theory of Pollak and Liao3, and the classicalWigner method4,5, with alternatives such as FK-LPI6

(Feynman-Kleinert Linearized Path Integral) and LGA-LSC-IVR7 (Local Gaussian Approximation LinearizedSemi-Classical Initial Value Representation). Other no-table approximate quantum mechanical methods to cal-culate reaction rate constants are e.g. instanton reactionrate theory8–10 and Ring Polymer Molecular Dynamics(RPMD)11.This article presents one way to calculate reaction rate

constants with a recent implementation of the classicalWigner method by the same authors12. Section II is anintroduction to a general formulation of chemical reac-tion rates, then section III combines this theory withthe OPCW method. Section IV notes some importantcorrections to factor in when running the calculations.Sections V, VI, and VII respectively declare the detailsof the computations that have been run, show the resultsof these computations, and contain the conclusions thathave been drawn.

II. REACTION RATE CONSTANTS

Miller and coworkers published13,14 three quantum me-chanical traces that can be used to acquire bimolecular

a)Electronic mail: [email protected])Electronic mail: [email protected]

reaction rate constants. These traces are

Css(t) = Tre−

β2 H h (−x) e−

β2 H e

iHt h (x) e−

iHt

(1)

Cfs(t) = Tre−

β2 H F e−

β2 H e

iHt h (x) e−

iHt

(2)

Cff(t) = Tre−

β2 H F e−

β2 H e

iHt F e−

iHt

(3)

where h(x) is the heaviside function, x is the position op-erator, β = (kBT )

−1, where kB is Boltzmann’s constant

and T is the absolute temperature, H is the Hamiltonianoperator of the system, i is the imaginary unit, t is time, is the reduced Planck constant, and F is the probabil-ity density flux operator. These functions can be used toobtain reaction rate constants, kr, through the relations

krQR = limt→∞

d

dtCss(t) (4)

= limt→∞Cfs(t) (5)

=

∫ ∞

0

dt Cff(t) =1

2

∫ ∞

−∞dt Cff(t) (6)

where QR is the canonical partition function per unitvolume of the reactants.It can be noted that the traces presented here have a

symmetrized Boltzmann operator15, i.e. e−β2 H on both

sides of the first operator in each trace. The relations4-6 are just as valid for the real part of the trace with anasymmetrically placed Boltzmann operator13. Importantto notice is that these relations are defined for potentialswhere the reactants start infinitely far from each otherand the products end up infinitely far from each other.This requires the potential to be unbound. This restric-tion applies to the potentials employed in this article.In the computations presented in this article the trace

used for the calculation of rate constants is Cfs(t).

III. REACTION RATE CALCULATION WITH THEOPEN POLYMER CLASSICAL WIGNER METHOD

The quantity of interest in this work is the sym-metrized trace Cfs(t) in equation 2.

IV

Page 203: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

IV

2

In this section expressions for calculating this quantitywith OPCWwill be derived. These derivations will followthe ones in the original OPCW article12 quite closely, butwill be using the symmetrized Boltzmann operator andwill be specialized for the flux-side trace.Initially Cfs(t) is written as the integral over position,

x1, of a matrix element.

Cfs(t) =

−∞dx1

x1

e−β2 H F e−

β2 H e

iHt h (x) e−

iHtx1

(7)

The Boltzmann operators are each divided into N2 fac-

tors of e−βN H , where N must be an even number. N − 1

identity operators are inserted between these operators.

Cfs(t) =

⎧⎨⎩

N

j=1

−∞dxj

⎫⎬⎭

x1

e−βN H

x2

x2

e−βN H

x3

. . .

xN

2

e−βN H F

xN2 +1

. . .xN−1

e−βN H

xN

xN

e−βN H e

iHt h(x) e−

iHtx1

(8)

This is an imaginary time, −iβ, Feynman path integral16, which can be expressed with Wigner transforms insteadof matrix elements,

Cfs(t) =

⎧⎨⎩

N

j=1

−∞

dxj dpj2π

⎫⎬⎭ e−

i∑N

j=1 pj(x(j mod N)+1−xj)e−

βN H

W

x1 + x2

2, p1

e−

βN H

W

x2 + x3

2, p2

. . .e−

βN H F

W

xN

2+ xN

2 +1

2, pN

2

. . .

e−

βN H

W

xN−1 + xN

2, pN−1

×e−

βN H e

iHt h(x) e−

iHtW

xN + x1

2, pN

. (9)

In the limit of N → ∞ the Wigner transform of the Boltzmann operator simplifies to the classical Boltzmannfactor. Also the Boltzmann operators can be separated from the flux and Heaviside operators in this limit.

Cfs(t) = limN→∞

⎧⎨⎩

N

j=1

−∞

dxj dpj2π

⎫⎬⎭ e−

i∑N

j=1 pj(x(j mod N)+1−xj) e−βN

∑Nj=1 H(yj ,pj)

FW

yN

2, pN

2

×e

iHt h(x) e−

iHtW

[yN , pN ] (10)

where yj =xj+x(j mod N)+1

2 and H (yj, pj) is the classical Hamiltonian. For a V (yj) that is independent of momentum

all momenta except pN2and pN can be integrated out analytically as done in the original derivation of OPCW12.

Cfs(t) = limN→∞

mN

2πβ

N2 β

2πmN−2N

⎧⎨⎩

N

j=1

−∞dxj

⎫⎬⎭

−∞dpN

2dpN e−

βN

∑Nj=1 V (yj)

× e− mN

22β

(∑N2

−1

j=1 (xj+1−xj)2+

∑N−1

j=N2

+1(xj+1−xj)

2

)

e− i

pN2

(xN

2+1

−xN2

)

e−βN

p2N2

2m

FW

yN

2, pN

2

× e−ipN (x1−xN ) e−

βN

p2N2m

e

iHt B e−

iHtW

[yN , pN ] (11)

For the flux operator F = 12m (δ(x)p+ pδ(x)) the Wigner transform is

FW

[x, p] =p

mδ(x). (12)

IV

Page 204: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

IV

3

Using the momentum integrals of appendix A in our pre-

vious article12 a quantity F yN

2, xN

2 +1 − xN2

can be

defined as

F yN

2, xN

2 +1 − xN2

=

∞−∞ dpN

2

FW

yN

2, pN

2

e−

βN

p2N2

2m e− i

pN2

(xN

2+1

−xN2

)

∞−∞ dpN

2e−

βN

p2N2

2m e− i

pN2

(xN

2+1

−xN2

)

=

∞−∞ dpN

2

pN2

m δyN

2

e−

βN

p2N2

2m e− i

pN2

(xN

2+1

−xN2

)

2πmN

β e− β

N

mN2

(xN

2+1

−xN2

)2

22β2

= − iN

β

xN

2 +1 − xN2

δyN

2

. (13)

This can be put into equation 11, giving

Cfs(t) = limN→∞

mN

2πβ

N2

β

2πmN−N

⎧⎨⎩

N

j=1

−∞dxj

⎫⎬⎭

−∞dpN e−

βN

∑Nj=1 V (yj)

× e− mN

22β

∑N−1j=1 (xj+1−xj)

2 iN

β

xN

2− xN

2 +1

δyN

2

× e−ipN (x1−xN ) e−

βN

p2N2m

e

iHt h(x) e−

iHtW

[yN , pN ] . (14)

Applying the classical Wigner approximation,

e

iHt h(x) e−

iHtW

[yN , pN ] ≈ (h(x))W [x (yN , pN , t) , p (yN , pN , t)] = h (x (yN , pN , t)) , (15)

and assuming finite N results in

Cfs,y(t) ≈mN

2πβ

N2

β

2πmN−N

⎧⎨⎩

N

j=1

−∞dxj

⎫⎬⎭

−∞dpN e−

ipN (x1−xN )

× e− β

N

(p2N2m +

∑Nj=1 V (yj)+

mN2

22β2

∑N−1j=1 (xj+1−xj)

2

)iN

β

xN

2− xN

2 +1

δyN

2

h (x (yN , pN , t)) , (16)

which is the expression corresponding to the y1-versionof OPCW for the flux-Heaviside trace. Due to the sym-metrization of the Boltzmann operator, yN

2rather than

y1 is used. This trace will be called Cfs,y(t).

To get an expression corresponding to the x1-versionof OPCW we start over from equation 8. There the x-dependent parts of F has to operate to the right in thematrix element before rewriting as Wigner transforms.

To be able to operate to the right, the flux operatorhas to be rearranged into a form where all x-dependenceis to the right of all p-dependence. Using that p = −i d

dxthis can be achieved as

F =1

2m(δ(x)p+ pδ(x)) =

1

2m([δ(x), p] + pδ(x) + pδ(x))

=1

2m

−i

δ(x)

d

dx− d

dxδ(x)

+ 2pδ(x)

=1

2m

−i

δ(x)

d

dx− dδ(x)

dx− δ(x)

d

dx

+ 2pδ(x)

=1

2m

i

dδ(x)

dx+ 2pδ(x)

, (17)

where the square brackets denote a commutator. Usingthis new form of the flux operator and operating x to theright we acquire

IV

Page 205: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

IV

4

Cfs(t) =

⎧⎨⎩

N

j=1

−∞dxj

⎫⎬⎭

x1

e−βN H

x2

x2

e−βN H

x3

. . .

⎛⎝xN

2

e−βN H p

xN2 +1

1

mδxN

2 +1

+xN

2

e−βN H

xN2 +1

i2m

dδxN

2 +1

dxN2 +1

⎞⎠

. . .xN−1

e−βN H

xN

xN

e−βN H e

iHt h(x) e−

iHtx1

. (18)

Rewriting as Wigner transforms, taking the limit where the Wigner transform of the Boltzmann operator is theclassical Boltzmann factor, and assuming that V (yj) is independent of momentum, a new way to write equation 11can be obtained:

Cfs(t) = limN→∞

mN

2πβ

N2 β

2πmN−2N

⎧⎨⎩

N

j=1

−∞dxj

⎫⎬⎭

−∞dpN

2dpN e−

βN

∑Nj=1 V (yj)

× e− mN

22β

(∑N2

−1

j=1 (xj+1−xj)2+

∑N−1

j= N2

+1(xj+1−xj)

2

)

e− i

pN2

(xN

2+1

−xN2

)

e−βN

p2N2

2m

×

⎛⎝(p)W

yN

2, pN

2

1

mδxN

2 +1

+

i2m

dδxN

2 +1

dxN2 +1

⎞⎠

× e−ipN (x1−xN ) e−

βN

p2N2m

e

iHt h(x) e−

iHtW

[yN , pN ] . (19)

Knowing that (p)W [x, p] = p, pN2can be integrated out.

∞−∞ dpN

2(p)W

yN

2, pN

2

e−

βN

p2N2

2m e− i

pN2

(xN

2+1

−xN2

)

∞−∞ dpN

2e−

βN

p2N2

2m e− i

pN2

(xN

2+1

−xN2

)

=

∞−∞ dpN

2pN

2e−

βN

p2N2

2m e− i

pN2

(xN

2+1

−xN2

)

2πmN

β e− β

N

mN2

(xN

2+1

−xN2

)2

22β2

= − iNm

β

xN

2 +1 − xN2

(20)

Cfs(t) = limN→∞

mN

2πβ

N2

β

2πmN−N

⎧⎨⎩

N

j=1

−∞dxj

⎫⎬⎭

−∞dpN e−

βN

∑Nj=1 V (yj)

× e− mN

22β

∑N−1j=1 (xj+1−xj)

2

⎛⎝− iN

β

xN

2 +1 − xN2

δxN

2 +1

+

i2m

dδxN

2 +1

dxN2 +1

⎞⎠

× e−ipN (x1−xN ) e−

βN

p2N2m

e

iHt h(x) e−

iHtW

[yN , pN ] (21)

If f(x) is a function of x then∞−∞ dx f(x)dδ(x)dx =

∞−∞ dx

−df(x)

dx

δ(x). This can be used to eliminate the

IV

Page 206: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

IV

5

derivative of the delta function through

−∞dxN

2e− β

N

(V

(yN

2

)+V

(yN

2+1

))

e− mN

22β

((xN

2+1

−xN2

)2

+

(xN

2+2

−xN2

+1

)2)i2m

dδxN

2 +1

dxN2 +1

=

−∞dxN

2(−1)

⎛⎝− β

N

⎛⎝1

2

dVyN

2

dyN2

+1

2

dVyN

2 +1

dyN2 +1

⎞⎠− mN

22β

2xN

2 +1 − xN2

− 2

xN

2 +2 − xN2 +1

⎞⎠

× e− β

N

(V

(yN

2

)+V

(yN

2+1

))

e− mN

22β

((xN

2+1

−xN2

)2

+

(xN

2+2

−xN2

+1

)2)i2m

δxN

2 +1

=

−∞dxN

2e− β

N

(V

(yN

2

)+V

(yN

2+1

))

e− mN

22β

((xN

2+1

−xN2

)2

+

(xN

2+2

−xN2

+1

)2)

×

⎛⎝ iβ4Nm

⎛⎝dV

yN

2

dyN2

+dV

yN

2 +1

dyN2 +1

⎞⎠+

iN

2xN

2 +1 − xN2− xN

2 +2

⎞⎠ δ

xN

2 +1

(22)

leading to

Cfs(t) = limN→∞

mN

2πβ

N2

β

2πmN−N

⎧⎨⎩

N

j=1

−∞dxj

⎫⎬⎭

−∞dpN e−

βN

∑Nj=1 V (yj) e

− mN22β

∑N−1j=1 (xj+1−xj)

2

×

⎛⎝− iN

β

xN

2 +1 − xN2

δxN

2 +1

+

⎛⎝ iβ4Nm

⎛⎝dV

yN

2

dyN2

+dV

yN

2 +1

dyN2 +1

⎞⎠+

iN

2xN

2 +1 − xN2− xN

2 +2

⎞⎠ δ

xN

2 +1

⎞⎠

× e−ipN (x1−xN ) e−

βN

p2N2m

e

iHt h(x) e−

iHtW

[yN , pN ]

= limN→∞

mN

2πβ

N2

β

2πmN−N

⎧⎨⎩

N

j=1

−∞dxj

⎫⎬⎭

−∞dpN e−

βN

∑Nj=1 V (yj) e

− mN22β

∑N−1j=1 (xj+1−xj)

2

×

⎛⎝ iβ4Nm

⎛⎝dV

yN

2

dyN2

+dV

yN

2 +1

dyN2 +1

⎞⎠+

iN

2xN

2 +1 − xN2− xN

2 +2 − 2xN2 +1 + 2xN

2

⎞⎠ δ

xN

2 +1

× e−ipN (x1−xN ) e−

βN

p2N2m

e

iHt h(x) e−

iHtW

[yN , pN ]

= limN→∞

mN

2πβ

N2

β

2πmN−N

⎧⎨⎩

N

j=1

−∞dxj

⎫⎬⎭

−∞dpN e−

βN

∑Nj=1 V (yj) e

− mN22β

∑N−1j=1 (xj+1−xj)

2

×

⎛⎝ iβ4Nm

⎛⎝dV

yN

2

dyN2

+dV

yN

2 +1

dyN2 +1

⎞⎠+

iN

xN

2− xN

2 +2

⎞⎠ δ

xN

2 +1

× e−ipN (x1−xN ) e−

βN

p2N2m

e

iHt h(x) e−

iHtW

[yN , pN ] . (23)

Up to this point the new derivation is equivalent to what was previously done up to equation 14. However, once

IV

Page 207: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

IV

6

the expression is approximated through classical Wigner and the assumption of finite N , giving

Cfs,x(t) ≈mN

2πβ

N2

β

2πmN−N

⎧⎨⎩

N

j=1

−∞dxj

⎫⎬⎭

−∞dpN e−

ipN (x1−xN )

× e− β

N

(p2N2m +

∑Nj=1 V (yj)+

mN2

22β2

∑N−1j=1 (xj+1−xj)

2

)

×

⎛⎝ iβ4Nm

⎛⎝dV

yN

2

dyN2

+dV

yN

2 +1

dyN2 +1

⎞⎠+

iN

xN

2− xN

2 +2

⎞⎠ δ

xN

2 +1

× h (x (yN , pN , t)) , (24)

this is not the case anymore. Equation 24 is the expres-sion corresponding to the x1-version of OPCW for theflux-Heaviside trace. Due to the symmetrization of thecorrelation function xN

2 +1 rather than x1 is used. This

trace will be called Cfs,x(t).

IV. CORRECTION FACTORS FOR EVALUATION OFTRACES WITH MONTE CARLO

In this work the position integrals in equations 16 and24 were evaluated with Metropolis Monte Carlo17.

Compared to calculating a correlation function, cal-culating a trace with Monte Carlo can be a bit morecomplicated. The flux-Heaviside correlation function is

F e

iHt h(x) e−

iHt=

1

ZTr

F e−βH e

iHt h(x) e−

iHt,

(25)

where the angular brackets denote a thermal average andZ is the canonical partition function

Z = Tre−βH

. (26)

In Monte Carlo sampling a weight function is needed.The delta functions in equation 16 and 24 complicatethings somewhat, since it is impossible to integrate themnumerically. The weight function used is

mN

2πβ

N2

β

2πmN−N

× e− β

N

(p2N2m +

∑Nj=1 V (yj)+

mN2

22β2

∑N−1j=1 (xj+1−xj)

2

)

δ (yy,x) ,(27)

where yy,x is either yN2or xN

2 +1. The integrand is thus

e−ipN (x1−xN ) iN

β

xN

2− xN

2 +1

h (x (yN , pN , t)) (28)

for the y-version of OPCW and

e−ipN (x1−xN )

×

⎛⎝ iβ4Nm

⎛⎝dV

yN

2

dyN2

+dV

yN

2 +1

dyN2 +1

⎞⎠+

iN

xN

2− xN

2 +2

⎞⎠

× h (x (yN , pN , t)) (29)

for the x-version of OPCW. This means that the quantity

actually calculated is Cfs(t)Z

δ, where Z

δ is

Z δ =

mN

2πβ

N2

β

2πmN−N

⎧⎨⎩

N

j=1

−∞dxj

⎫⎬⎭

−∞dpN

× e− β

N

(p2N2m +

∑Nj=1 V (yj)+

mN2

22β2

∑N−1j=1 (xj+1−xj)

2

)

× δ (yy,x) , (30)

and is something similar to a partition function, but theintegrand includes a delta function and lacks a factor of

e−ipN (x1−xN ). By sampling the integrand e−

ipN (x1−xN )

in the same Monte Carlo run as Cfs(t)Z

δa correction factor

Zδcan be acquired, where

Zδ =

mN

2πβ

N2

β

2πmN−N

⎧⎨⎩

N

j=1

−∞dxj

⎫⎬⎭

−∞dpN

× e− β

N

(p2N2m +

∑Nj=1 V (yj)+

mN2

22β2

∑N−1j=1 (xj+1−xj)

2

)

× e−ipN (x1−xN) δ (yy,x) . (31)

Zδ is very similar to Z δ, but it has e

− ipN (x1−xN ) in the

integrand, and is thus a partition function with a deltafunction in it. For a parabolic barrier Zδ can be calcu-lated analytically for the exact quantum mechanical case,N → ∞. It is

Zδ =

2π sin(βω). (32)

For the classical case it is

Zδ =

m

2πβ−1. (33)

IV

Page 208: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

IV

7

For the Eckart potential and other potentials withoutan analytic solution, the correction factor has to be cal-culated numerically.The quantity obtained from the Monte Carlo integra-

tion is thus Cfs(t)Z′

δ. A correction factor of Zδ

Z′δcan likewise

be obtained from the Monte Carlo run, but an additionalcorrection factor Zδ has to the computed.

Cfs(t) =Cfs(t)

Z δ

× Z δ

Zδ× Zδ (34)

V. COMPUTATIONAL DETAILS

In this work, the integrals in Eqs. 16 and 24 wereevaluated by Monte Carlo and the time propagation wasconducted with molecular dynamics.

A. Potentials and system parameters

Two different potentials are studied in this work: aparabolic barrier and a symmetric Eckart barrier.The parabolic barrier has the potential

Vparabolic(x) = −1

2mω2x2, (35)

where ω is the absolute value of the angular frequency ofthe system.The symmetric Eckart barrier used is of the form

VEckart(x) =6ωπ

sech2(√

πmω

12x

), (36)

where ω is the absolute value of the angular frequency atthe top of the barrier, i.e. at the transition state.

B. Traces

The Monte Carlo integration was conducted throughthe Metropolis Monte Carlo method17. The maximumstepsize for the Monte Carlo was choosen and updatedaccording to the procedure in the original OPCWwork12.A Maxwell–Boltzmann distribution of the inverse tem-perature β/N was used to sample the momentum, pN .For the sampling, the ran2 pseudo-random number gen-erator of Press et al.18 was used.A molecular dynamics trajectory was run for each 10th

Monte Carlo step. The dynamics were run with the ve-locity Verlet algorithm19,20, with a timestep of 0.05 ω−1

and a total time of 10 ω−1.The standard deviations of the results where calculated

with the block average method, as explained in e.g.21,22,using a minimum block size of 106 Monte Carlo steps.To get the correction factor Zδ for the Eckart potential

the numerical matrix multiplication scheme23 was used tocalculate the matrix elements of the Boltzmann operator.

C. Exact comparisons

The exact classical (CM) and quantum mechanical13

(QM) flux-Heaviside traces for the parabolic barrier canbe acquired as analytic functions,

Cfs,CM(t) =1

2πβ(37)

Cfs,QM(t) =1

2πβ

βω2

sin(

βω2

) sinh(ωt)√sinh2(ωt) + sin2

(βω2

) .

(38)

For the Eckart barrier the classical result is the sameanalytic function as for the parabolic barrier, apart froma Boltzmann factor that accounts for the height of thebarrier compared to the reactant energy,

kr,CMQR =1

2πβe−βV (0) . (39)

The corresponding quantum mechanical result can becalculated from an analytic transmission coefficient24,

κEckart(E) =cosh

(24

√πE6ω

)− 1

cosh(24

√πE6ω

)+ cosh

(√576− π2

) ,

(40)

but has to be numerically integrated,

kr,QMQR =1

∫ ∞

0

dE e−βE κEckart(E) (41)

. (42)

In this work the integral is performed by Rombergintegration25, using an integration interval from 0 to 100ω and a convergence criterion of 10−10.

In the case of the Eckart potential, the reactants arerepresented as a free particle, which has the partitionfunction per unit of length

QR =

√m

2πβ−1. (43)

This partition function is the same in the classical andquantum mechanical cases. The parabolic barrier is in-finitely high and the partition function for the reactantsis therefore infinite, and can not be calculated. The quan-tity krQR is, however, a finite calculable quantity, as therate constant is infinitely small and compensates for theinfinite reactant partition function.To obtain exact classical Wigner flux-Heaviside traces

for the Eckart potential the numerical matrix multiplica-tion scheme23 was used to calculate the matrix elementsof the Boltzmann operator.An RPMD comparison was calculated for the

Eckart barrier, using the formulation of Craig andManolopoulos26.

IV

Page 209: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

IV

8

0.10

0.12

0.14

0.16

0.18

0 1 2 3 4 5 6 7 8 9 10

C fs(t

) / (− h1/2

ω1/2

m-1/2 )

t / ω-1

CMN= 6 (64E9)N=10 (64E9)N=20 (64E9)N=40 (64E9)

QM

FIG. 1. Cfs(t) for a parabolic barrier with βω = 1. Thetraces are calculated with the y-version of OPCW for differ-ent N , exact Classical Mechanics (CM), and exact QuantumMechanics (QM). The numbers in parenthesis are the numberof Monte Carlo steps used. The top and bottom lines of eachtype show the standard deviation of the results.

0.10

0.12

0.14

0.16

0.18

0 1 2 3 4 5 6 7 8 9 10

C fs(t

) / (− h1/2

ω1/2

m-1/2 )

t / ω-1

CMN= 6 (64E9)N=10 (64E9)N=20 (64E9)N=40 (64E9)

QM

FIG. 2. Cfs(t) for a parabolic barrier with βω = 1. Thetraces are calculated with the x-version of OPCW for differ-ent N , exact Classical Mechanics (CM), and exact QuantumMechanics (QM). The numbers in parenthesis are the numberof Monte Carlo steps used. The top and bottom lines of eachtype show the standard deviation of the results.

VI. RESULTS AND DISCUSSION

For the parabolic barrier the results of the case βω =1 can be seen in Figs. 1 and 2. The classical Wignermethod is exact for harmonic potentials4, so the OPCWresults converge toward the exact quantum mechanicalresult as N is increased. It can be seen that the y-versionof OPCW needs more beads than the x-version to con-verge to the exact quantum mechanical result. The x-version of OPCW has excellent convergence with respectto the number of beads. It almost gives the exact resultfor N = 6.For the parabolic barrier the results of the case βω =

0.5 can be seen in Figs. 3 and 4. The standard deviationsare visible around the OPCW-results for this temper-

0.30

0.31

0.32

0.33

0 1 2 3 4 5 6 7 8 9 10

C fs(t

) / (− h1/2

ω1/2

m-1/2 )

t / ω-1

CMN= 6 ( 64E9)N=10 (128E9)N=20 (128E9)

QM

FIG. 3. Cfs(t) for a parabolic barrier with βω = 0.5. Thetraces are calculated with the y-version of OPCW for differ-ent N , exact Classical Mechanics (CM), and exact QuantumMechanics (QM). The numbers in parenthesis are the numberof Monte Carlo steps used. The top and bottom lines of eachtype show the standard deviation of the results.

0.30

0.31

0.32

0.33

0 1 2 3 4 5 6 7 8 9 10

C fs(t

) / (− h1/2

ω1/2

m-1/2 )

t / ω-1

CMN= 6 ( 64E9)N=10 (128E9)N=20 (128E9)

QM

FIG. 4. Cfs(t) for a parabolic barrier with βω = 0.5. Thetraces are calculated with the x-version of OPCW for differ-ent N , exact Classical Mechanics (CM), and exact QuantumMechanics (QM). The numbers in parenthesis are the numberof Monte Carlo steps used. The top and bottom lines of eachtype show the standard deviation of the results.

ature. Thus these higher temperature calculations arevisibly more difficult than those with βω = 1. Thesampling of the initial conditions in OPCW is known tobecome more difficult when raising the temperature butkeeping the same N12. Generally, however, fewer beadsare needed for a calculation at higher temperature. Thex-version of OPCW converges toward the exact quantummechanical result significantly faster than the y-versionalso at βω = 0.5, when the number of beads is increased.Particularly for βω = 0.5 it can be seen that Cfs,y(t)

converges toward exact quantum mechanics from aboveand Cfs,x(t) converges toward exact quantum mechanicsfrom below.In Figs. 5-10 Cfs(t) for the Eckart barrier at βω = 1,

3, and 6 are shown. The rate constants and tunneling

IV

Page 210: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

IV

9

2.0E-022.1E-022.2E-022.3E-022.4E-022.5E-022.6E-02

0 1 2 3 4 5 6 7 8 9 10

C fs(t

) / (− h1/2

ω1/2

m-1/2 )

t / ω-1

CMN= 6 ( 1E9)N=10 (16E9)N=20 (16E9)N=40 (16E9)

CWQM

FIG. 5. Cfs(t) for an Eckart potential with βω = 1. Thetraces are calculated with the y-version of OPCW for differentN , exact Classical Mechanics (CM), exact Classical Wigner(CW), and exact Quantum Mechanics (QM). The QM-lineshows the long time result, not the time dependent function.The numbers in parenthesis are the number of Monte Carlosteps used. The top and bottom lines of each type show thestandard deviation of the results.

2.0E-022.1E-022.2E-022.3E-022.4E-022.5E-022.6E-02

0 1 2 3 4 5 6 7 8 9 10

C fs(t

) / (− h1/2

ω1/2

m-1/2 )

t / ω-1

CMN= 6 ( 1E9)N=10 (16E9)N=20 (16E9)N=40 (16E9)

CWQM

FIG. 6. Cfs(t) for an Eckart potential with βω = 1. Thetraces are calculated with the x-version of OPCW for differentN , exact Classical Mechanics (CM), exact Classical Wigner(CW), and exact Quantum Mechanics (QM). The QM-lineshows the long time result, not the time dependent function.The numbers in parenthesis are the number of Monte Carlosteps used. The top and bottom lines of each type show thestandard deviation of the results.

factors, Γ = kr,QM/kr,CM, are presented in Tabs. I andII. For the Eckart potential at βω = 1 there is only∼ 6% difference between the quantum mechanical andclassical results. For the lower temperatures the differ-ence is significantly bigger, i.e. ∼ 50% for βω = 3 anda factor ∼ 5 for βω = 6.

The classical Wigner method, as could be expected,gives results that are intermediate between classical me-chanics and quantum mechanics.

The OPCW results, for the Eckart potential, convergeto the exact classical Wigner result as the number of

0.0E+005.0E-051.0E-041.5E-042.0E-042.5E-043.0E-04

0 1 2 3 4 5 6 7 8 9 10

C fs(t

) / (− h1/2

ω1/2

m-1/2 )

t / ω-1

CMN= 6 ( 1E9)N=10 ( 1E9)N=20 (16E9)N=40 (16E9)

CWQM

FIG. 7. Cfs(t) for an Eckart potential with βω = 3. Thetraces are calculated with the y-version of OPCW for differentN , exact Classical Mechanics (CM), exact Classical Wigner(CW), and exact Quantum Mechanics (QM). The QM-lineshows the long time result, not the time dependent function.The numbers in parenthesis are the number of Monte Carlosteps used. The top and bottom lines of each type show thestandard deviation of the results.

0.0E+005.0E-051.0E-041.5E-042.0E-042.5E-043.0E-04

0 1 2 3 4 5 6 7 8 9 10

C fs(t

) / (− h1/2

ω1/2

m-1/2 )

t / ω-1

CMN= 6 ( 1E9)N=10 ( 1E9)N=20 (16E9)N=40 (16E9)

CWQM

FIG. 8. Cfs(t) for an Eckart potential with βω = 3. Thetraces are calculated with the x-version of OPCW for differentN , exact Classical Mechanics (CM), exact Classical Wigner(CW), and exact Quantum Mechanics (QM). The QM-lineshows the long time result, not the time dependent function.The numbers in parenthesis are the number of Monte Carlosteps used. The top and bottom lines of each type show thestandard deviation of the results.

TABLE I. Rate constants (kr) for different inverse tempera-tures (β) for the Eckart barrier described in Eq. 36. Compari-sion of OPCWa, Ring Polymer Molecular Dynamics (RPMD),classical mechanics (CM), exact Classical Wigner (CW), andquantum mechanics (QM).

kr/(12ω

12m− 1

2 )βω OPCW RPMD CM CW QM1 6.10E-2 6.17E-2 5.9084E-2 6.0986E-2 6.2856E-23 1.009E-3 1.07E-3 7.4821E-4 1.0087E-3 1.1409E-36 7.391E-6 7.55E-6 1.7186E-6 7.5092E-6 8.9347E-6

a x-version, N = 40

IV

Page 211: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

IV

10

0.0E+005.0E-071.0E-061.5E-062.0E-062.5E-063.0E-063.5E-064.0E-06

0 1 2 3 4 5 6 7 8 9 10

C fs(t

) / (− h1/2

ω1/2

m-1/2 )

t / ω-1

CMN= 6 (16E9)N=10 (16E9)N=20 (32E9)N=40 (32E9)

CWQM

FIG. 9. Cfs(t) for an Eckart potential with βω = 6. Thetraces are calculated with the y-version of OPCW for differentN , exact Classical Mechanics (CM), exact Classical Wigner(CW), and exact Quantum Mechanics (QM). The QM-lineshows the long time result, not the time dependent function.The numbers in parenthesis are the number of Monte Carlosteps used. The top and bottom lines of each type show thestandard deviation of the results.

0.0E+005.0E-071.0E-061.5E-062.0E-062.5E-063.0E-063.5E-064.0E-06

0 1 2 3 4 5 6 7 8 9 10

C fs(t

) / (− h1/2

ω1/2

m-1/2 )

t / ω-1

CMN= 6 (16E9)N=10 (16E9)N=20 (16E9)N=40 (16E9)

CWQM

FIG. 10. Cfs(t) for an Eckart potential with βω = 6. Thetraces are calculated with the x-version of OPCW for differentN , exact Classical Mechanics (CM), exact Classical Wigner(CW), and exact Quantum Mechanics (QM). The QM-lineshows the long time result, not the time dependent function.The numbers in parenthesis are the number of Monte Carlosteps used. The top and bottom lines of each type show thestandard deviation of the results.

TABLE II. Tunneling factors (Γ) for different inverse tem-peratures (β) for the Eckart barrier described in Eq. 36.Comparision of OPCWa, Ring Polymer Molecular Dynamics(RPMD), exact Classical Wigner (CW), and exact results.

βω ΓOPCW ΓRPMD ΓCW ΓExact

1 1.03 1.04 1.0322 1.06383 1.349 1.43 1.3481 1.52496 4.300 4.39 4.3693 5.1987

a x-version, N = 40

beads is increased. The x-version converges faster thanthe y-version. As for the parabolic barrier, Cfs,y(t) con-verge from above and Cfs,x(t) converge from below. Thisis an interesting occurrence, but too few potentials havebeen studied here to draw conclusions from it.For both potentials and all temperatures the x-version

of OPCW gives results that are very close to the exactclassical Wigner result. Particularly for the Eckart po-tential at βω = 3 and 6 the classical Wigner methodcan be seen to be a significant improvement over classi-cal mechanics.Compared to RPMD the classical Wigner method gives

slightly worse results. However, in the perspective oftrying to approximate the quantum mechanical reactionrates, the differences between the rate constants calcu-lated by the classical Wigner method and RPMD, at thelowest temperatures, is insignificant compared to the dif-ference between those rate constants from the the exactquantum mechanical result. Results from a single sym-metric one-dimensional potential is, however, not enoughto draw general conclusions.

VII. CONCLUSION

In this work the recently developed Open PolymerClassical Wigner (OPCW) method12 is extended to theflux-Heaviside correlation function of Miller14 and ap-plied to a parabolic barrier and a symmetric Eckart po-tential, for a few different temperatures. The OPCWmethod is shown to converge to the exact classicalWigner result as the number of beads in the path in-tegral polymer is increased. Particularly the x-version ofOPCW shows promise for further tests.In the future, it would be of interest to apply the

OPCW reaction rate constant calculations to multidime-sional systems, such as the double well potential ofTopaler and Makri27 coupled to the harmonic bath ofCaldeira and Leggett28. For such a system modifyingOPCW by making the bath classical, as introduced in thepreceding article12, would be of interest. It would also beof interest to try the method on asymmetric potentials,as both potentials tried thus far have been symmetric.For systems where the reactants are bound it could

be of interest to use the Heaviside-Heaviside correlationfunction of Miller et al.13 instead of the flux-Heavisidecorrelation function.

ACKNOWLEDGMENTS

Financial support from the Swedish Research Council(Vetenskapsradet), diary number 2016-03275, is acknowl-edged.The computations in this work were performed at

Chalmers Centre for Computational Science and Engi-neering (C3SE), a partner center of the Swedish NationalInfrastructure for Computing (SNIC).

IV

Page 212: Quantum dynamical effects in complex chemical systems · extra emphasis on the calculation of reaction rate constants. ... for use with a variational principle for the dynamics of

IV

11

1E. Wigner, Zeitschrift fur physikalische Chemie B (Leipzig) 19,203 (1932).

2H. Eyring, The Journal of Chemical Physics 3, 107 (1935).3E. Pollak and J.-L. Liao, The Journal of Chemical Physics 108,2733 (1998).

4E. J. Heller, The Journal of Chemical Physics 65, 1289 (1976).5H. Wang, X. Sun, and W. H. Miller, The Journal of ChemicalPhysics 108, 9726 (1998).

6J. A. Poulsen, G. Nyman, and P. J. Rossky, The Journal ofChemical Physics 119, 12179 (2003).

7J. Liu and W. H. Miller, The Journal of Chemical Physics 131,074113 (2009).

8J. Langer, Annals of Physics 41, 108 (1966).9J. Langer, Annals of Physics 54, 258 (1969).

10W. H. Miller, The Journal of Chemical Physics 62, 1899 (1975).11I. R. Craig and D. E. Manolopoulos, The Journal of ChemicalPhysics 122, 084106 (2005).

12S. K.-M. Svensson, J. A. Poulsen, and G. Nyman, The Journalof Chemical Physics 152, 094111 (2020).

13W. H. Miller, S. D. Schwartz, and J. W. Tromp, The Journal ofChemical Physics 79, 4889 (1983).

14W. H. Miller, The Journal of Chemical Physics 61, 1823 (1974).15P. Schofield, Physical Review Letters 4, 239 (1960).16R. P. Feynman, Reviews of Modern Physics 20, 367 (1948).17N. Metropolis, A. W. Rosenbluth, M. N. Rosenbluth, A. H.Teller, and E. Teller, The Journal of Chemical Physics 21, 1087

(1953).18W. H. Press, S. A. Teukolsky, W. T. Vetterling, and B. P.Flannery, Numerical Recipes in FORTRAN: The Art of Sci-entific Computing, 2nd ed. (Cambridge University Press, ThePitt Building, Trumpington Street, Cambridge CB2 1RP, UnitedKingdom, 1992).

19L. Verlet, Physical Review 159, 98 (1967).20W. C. Swope, H. C. Andersen, P. H. Berens, and K. R. Wilson,The Journal of Chemical Physics 76, 637 (1982).

21R. Friedberg and J. E. Cameron, The Journal of ChemicalPhysics 52, 6049 (1970).

22J. M. Flegal, M. Haran, and G. L. Jones, Statistical Science 23,250 (2008).

23D. Thirumalai, E. J. Bruskin, and B. J. Berne, The Journal ofChemical Physics 79, 5063 (1983).

24H. S. Johnston and D. Rapp, Journal of the American ChemicalSociety 83, 1 (1961).

25W. Romberg, Det Kongelige Norske Videnskabers SelskabForhandlinger 28, 30 (1955).

26I. R. Craig and D. E. Manolopoulos, The Journal of ChemicalPhysics 123, 034102 (2005).

27M. Topaler and N. Makri, The Journal of Chemical Physics 101,7500 (1994).

28A. O. Caldeira and A. J. Leggett, Annals of Physics 149, 374(1983).

IV