crystal structures, electron impact mass spectra and ...

251
Informatics and computational methods in physical chemistry: crystal structures, electron impact mass spectra and thermochemistry Peter Robert Spackman BA (Hons) & BSc, W. Aust. This thesis is presented for the degree of Doctor of Philosophy of The University of Western Australia School of Molecular Sciences December 2017 Supervisors Assoc. Prof. Dylan Jayatilaka* Asst. Prof. Amir Karton Assoc. Prof Mark Reynolds

Transcript of crystal structures, electron impact mass spectra and ...

Informatics and computational methods in physical

chemistry: crystal structures, electron impact mass

spectra and thermochemistry

Peter Robert Spackman

BA (Hons) & BSc, W. Aust.

This thesis is presented for the degree of

Doctor of Philosophy of The University of Western Australia

School of Molecular Sciences

December 2017

Supervisors

Assoc. Prof. Dylan Jayatilaka*

Asst. Prof. Amir Karton

Assoc. Prof Mark Reynolds

Thesis Declaration

I, Peter Robert Spackman, certify that:

This thesis has been substantially accomplished during enrolment in the degree.

This thesis does not contain material which has been accepted for the award of any other

degree or diploma in my name, in any university or other tertiary institution.

No part of this work will, in the future, be used in a submission in my name, for any other

degree or diploma in any university or other tertiary institution without the prior approval of

The University of Western Australia and where applicable, any partner institution responsible

for the joint-award of this degree.

This thesis does not contain any material previously published or written by another person,

except where due reference has been made in the text.

The work(s) are not in any way a violation or infringement of any copyright, trademark, patent,

or other rights whatsoever of any person.

The work described in this thesis was funded by the Danish National Research Foundation

(Center for Materials Crystallography, DNRF-93).

This thesis contains published work and/or work prepared for publication, some of which has

been co-authored.

Signature:

Date:

ii

20/03/2018

Abstract

This thesis comprises a body of work that investigates the development, assessment, and

application of computational methods and processes in chemistry. The work may be broadly

separated into two parts: the first exploring a novel and efficient description of molecular

shape and properties within molecular crystals, and the second focused on the assessment of

state-of-the-art quantum chemical predictions for small molecular geometries, energies and

electron impact mass spectra.

It is highly desirable to incorporate molecular shape into quantitative studies, especially

given the wealth of accurate experimental data available in the Cambridge Structural Database.

It is shown that molecular shape – defined as either the Hirshfeld surface or a promolecular

density surface – can be accurately described through spherical harmonic transform, and rotation

invariant shape descriptors may be used to measure similarity, and thus cluster or group like

with like. Further, by coupling surface properties like electrostatic potential with the shape

description, we show that molecules with similar interaction propensities can be described in

the same manner.

Accurate and timely prediction of an Electron Impact Mass Spectrum (EIMS) from a given

molecule would prove invaluable in metabolomics, natural products analysis and many other

fields where identification of unknown small molecules is labour intensive and time consuming.

To this end, the Quantum Chemical Electron Impact Mass Spectrum (QCEIMS) method is

examined, and its predicted spectra are compared with an entirely different prediction technique

– Competitive Fragmentation Modeling (CFM) – and experiment.

Finally, accurate energies and equilibrium geometries may be utilised for predicting a

wide variety of electronic and spectroscopic properties of small molecules. Knowledge and

understanding of trends associated with increasing basis-set size are of extreme importance if

chemical accuracy (i.e. an error bound within approximately 4 kJ mol−1) is sought. Therefore,

different basis-set extrapolation techniques for the ab initio method ‘CCSD’ are examined in order

to identify the most appropriate and cost-effective method to accelerate basis set convergence.

Similarly, the basis-set convergence of molecular geometries optimised with CCSD(T) and the

newer explicitly correlated CCSD(T)-F12 methods are examined in order to establish appropriate

standards for high accuracy theoretical calculations.

ii

Acknowledgments

I would never have been able to complete this thesis if not for the help of a large support base

of family, friends and colleagues.

First and foremost, my gratitude and thanks go to my supervisors, Dylan Jaytilaka, Amir

Karton and Mark Reynolds. Dylan, for taking me on as a student – a risk I’m not sure many

would take. When I began this research, I had no chemistry background at the higher degree

level. With Dylan’s knowledge, patience and assistance I have learned so much about chemistry,

physics, mathematics and much more. Simply put, I would not be in the position I am today

without his support. You’ve been a great teacher, friend and colleague, and have always made

me feel welcome, listened to my ideas or concerns and given me helpful feedback. Amir, who

likewise has imparted a great deal of scientific knowledge, but also provided great support and

advice. Thanks to you I’m a lot more savvy than I’d have been without you, and it has been a

privilege to work with you. I count myself very lucky to have had such supportive, patient and

enthusiastic researchers to collaborate with and learn from throughout my PhD.

Likewise, all of those I have collaborated closely with throughout my PhD. Sajesh Thomas,

whose ideas, discussions and perspectives had a huge impact on the direction of my research,

and my general scientific outlook. Mike Turner, whose guidance and discussion helped me

greatly in approaching software in science. Bjorn Bohman, whose perspectives and knowledge

of mass spectrometry and chemistry made Chapter 5 of this thesis possible.

I am incredibly thankful for the financial support of the Danish National Research Foundation

(Center for Materials Crystallography, DNRF-93), which made this PhD possible, and particularly

Bo Iversen (Aarhus) for having faith that I could do this. Additionally, this research was

supported by an Australian Government Research Training Program (RTP) Scholarship.

I would also like to thank those I have collaborated with in other publications, for their

valuable contributions: Simon Grabowsky (Bremen), Mingwen Shi, Li-Juan Yu, Campbell

Mackenzie, Scott Stewart, Alexandre Sobolev, Birger Dittrich (Gottingen), Tanja Schirmeister,

Peter Luger, Malte Hesse and Yu-Sheng Chen.

I owe many thanks to the people here at the School of Molecular Sciences at UWA, for

making my time here enjoyable, and for many interesting conversations. In particular Graham

Chandler, for his time, patience and knowledge of topics in theoretical chemistry, but also for

many great stories he’s shared with me – these have helped humanise many figures in theoretical

iii

chemistry. Also, the perennial UWA visitor Hans-Beat Burgi who I’ve known far longer than

I’ve known his influence in chemistry, your wisdom has been greatly appreciated.

Of course, my family and my friends for their support, patience and love throughout these

years. My mother, my sisters in particular have always been there for me, as has my girlfriend

Melinda – who has also no doubt helped me retain my sanity.

And finally, my father, Mark Spackman, without whom I would never have had the opportu-

nities I do in science. From you I have learned so much – some directly, a lot simply through

simply being around you most of my life. Through you I’ve had the opportunity to meet many

academics in chemistry for much of my life, and only through the research here did I really

come to grips with just who they were in this field. You’ve stayed out of direct involvement in

this work, but I’m sure anyone can directly see the influence you’ve had on me and my thought

processes, and I’m incredibly fortunate to have such a wonderful father.

iv

Authorship declaration: co-authored publications

This thesis contains the following work that has been published and/or prepared for

publication, in which the candidate is the first author and primary contributor:

Spackman, P. R., Thomas, S. P. & Jayatilaka, D. High Throughput Profiling of Molecular

Shapes in Crystals. Scientific Reports 6, 22204 (2016).

Location in thesis: Part II, Chapter 3

PRS and DJ designed the research. PRS performed all calculations, implemented all software

to describe molecular shape and analysed all results. PRS wrote the manuscript, which was

finalized with the help of DJ and SPT. (Overall contribution by PRS: 90%)

Spackman, P.R., Yu, L.-J., Morton, C., Bond, C.S., Spackman, M.A., Jayatilaka, D., &

Thomas, S.P, Fast Screening of Crystal Structures for ligands based on molecular shape and

intermolecular interactions. Planned for publication

Location in thesis: Part II, Chapter 4

PRS and SPT designed the research. PRS implemented all software to describe molecular

shapes and analyzed all results. PRS wrote the manuscript, with feedback from DJ, SPT.

(Overall contribution by PRS: 85%)

Spackman, P. R., Bohman, B., Karton, A. & Jayatilaka, D. Quantum chemical electron

impact mass spectrum prediction for de novo structure elucidation: Assessment against

experimental reference data and comparison to competitive fragmentation modeling.

International Journal of Quantum Chemistry e25460–N/A (2017)

Location in thesis: Part III, Chapter 5

PRS, AK and BB designed the research. PRS performed all calculations, and performed all

analysis of the results. PRS wrote the manuscript, which was finalized with AK, DJ and BB.

(Overall contribution by PRS: 90%)

Spackman, P. R., Jayatilaka, D. & Karton, A. Basis set convergence of CCSD(T) equilibrium

geometries using a large and diverse set of molecular structures. J. Chem. Phys. 145, 104101

(2016).

Location in thesis: Part III, Chapter 6

PRS and AK designed the research. PRS performed processing of calculation results, including

calculation of equilibrium bond lengths and angles. PRS and AK performed the analysis, and

wrote the manuscript. (Overall contribution by PRS: 75%)

v

Spackman, P. R. & Karton, A. Estimating the CCSD basis-set limit energy from small basis

sets: basis-set extrapolations vs additivity schemes. AIP Advances 5, 057148 (2015).

Location in thesis: Part III, Chapter 7

PRS and AK designed the research. PRS and AK performed the analysis of the results. PRS

and AK finalized the manuscript. (Overall contribution by PRS: 60%)

I, Peter Spackman, certify that the my statements regarding my contribution to each of the

works listed above are correct.

Student signature:

Date:

I, Dylan Jayatilaka, certify that the student statements regarding their contribution to each of

the works listed above are correct.

Coordinating supervisor signature:

Date:

vi

20/03/2018

23/03/2018

Other co-authored publications during the Ph.D candida-

ture

The first pages of other publications to which the candidate has contributed, but not as first

author, are included as appendices.

Shi, M. W. et al. Shi, M.W., Stewart, S.G., Sobolev, A.N., Dittrich, B., Schirmeister, T., Luger,

P., Hesse, M., Chen, Y.S., Spackman, P.R., Spackman, M.A. & Grabowsky, S. Approaching

an experimental electron density model of the biologically active trans-epoxysuccinyl amide

group—Substituent effects vs. crystal packing. Journal of Physical Organic Chemistry 36,

e3683 (2017).

Mackenzie, C. F., Spackman, P. R., Jayatilaka, D. & Spackman, M. A. CrystalExplorer

model energies and energy frameworks: extension to metal coordination compounds, organic

salts, solvates and open-shell systems. IUCrJ 4, (2017).

Edwards, A. J., Mackenzie, C., Spackman, P.R., Jayatilaka, D. & Spackman, M. A.

FDHALO17: Intermolecular interactions in molecular crystals: What’s in a name? Faraday

Discussions (2017).

vii

viii

Contents

Contents ix

List of Figures xv

List of Tables xix

List of Abbreviations and Symbols xxiii

I Introduction and Background 1

1 Background 3

1.1 Foreword . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

1.2 Science and computers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

1.2.1 History and philosophy . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

1.2.2 Data, information, knowledge and science . . . . . . . . . . . . . . . . . 5

1.2.3 Informatics and computer science . . . . . . . . . . . . . . . . . . . . . . 8

1.2.4 The role of computers in chemistry . . . . . . . . . . . . . . . . . . . . . 9

1.3 Structural and theoretical chemistry . . . . . . . . . . . . . . . . . . . . . . . . . 10

1.3.1 Chemical characterisation . . . . . . . . . . . . . . . . . . . . . . . . . . 10

1.3.2 The role of computers in chemical characterisation . . . . . . . . . . . . . 11

1.3.3 Crystal engineering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

1.3.4 The promolecule . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

1.3.5 Atoms in molecules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

1.3.6 Molecules in crystals: the Hirshfeld surface . . . . . . . . . . . . . . . . . 13

1.3.7 Hirshfeld surface fingerprints . . . . . . . . . . . . . . . . . . . . . . . . . 15

1.4 Quantum and computational chemistry . . . . . . . . . . . . . . . . . . . . . . . 15

1.4.1 The time-independent Schrodinger equation . . . . . . . . . . . . . . . . 16

1.4.2 The Born-Oppenheimer approximation and independent particle model . 17

1.4.3 Hartree-Fock theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

1.4.4 Basis functions and basis sets . . . . . . . . . . . . . . . . . . . . . . . . 19

ix

CONTENTS

1.4.5 Perturbation theory, configuration interaction and coupled cluster theory 23

1.4.6 Explicitly correlated methods . . . . . . . . . . . . . . . . . . . . . . . . 26

1.4.7 Density functional theory . . . . . . . . . . . . . . . . . . . . . . . . . . 26

1.4.8 Born-Oppenheimer molecular dynamics . . . . . . . . . . . . . . . . . . . 29

1.5 Chemoinformatics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

1.5.1 Molecules, their properties, and their representation . . . . . . . . . . . . 30

1.5.2 Data types in chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

1.5.3 Machine learning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

1.5.4 Chemical databases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

1.5.5 Virtual screening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

2 Introduction to a series of papers 47

2.1 Molecular shapes in crystals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

2.2 Assessment of the accuracy of quantum chemical methods . . . . . . . . . . . . 48

II Molecular shapes in crystals 51

3 High throughput profiling of molecules in crystals 53

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

3.2 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

3.2.1 Hirshfeld surfaces of metallic crystals . . . . . . . . . . . . . . . . . . . . 57

3.2.2 Hirshfeld surfaces of molecular crystals . . . . . . . . . . . . . . . . . . . 59

3.2.3 High throughput processing of structures in CSD . . . . . . . . . . . . . 61

3.3 Future research and prospects . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

3.4 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

3.4.1 Representing Hirshfeld surfaces with spherical harmonics . . . . . . . . . 64

3.4.2 Rotation invariant shape descriptors . . . . . . . . . . . . . . . . . . . . 65

3.4.3 Efficiency and computer implementation considerations . . . . . . . . . . 65

3.5 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

3.6 Author contributions statement . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

4 Fast screening of crystal structures for ligands based on molecular shape and

intermolecular interactions 69

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

4.2 Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

4.2.1 Selection of crystal structures . . . . . . . . . . . . . . . . . . . . . . . . 72

4.2.2 Selected ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

4.2.3 Calculation of the Hirshfeld and Promolecule electron density isosurfaces 73

4.2.4 Calculation of combined shape and surface property descriptors . . . . . 75

x

CONTENTS

4.2.5 Construction of the shape descriptor library . . . . . . . . . . . . . . . . 77

4.2.6 Wavefunction calculations . . . . . . . . . . . . . . . . . . . . . . . . . . 77

4.2.7 Electrostatic potential (ESP) calculations . . . . . . . . . . . . . . . . . . 78

4.2.8 Shape Matching . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

4.2.9 Docking . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

4.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

4.3.1 Docking scores for the best-matching ligands . . . . . . . . . . . . . . . . 79

4.3.2 Matches with known drugs in the CSD . . . . . . . . . . . . . . . . . . . 81

4.3.3 Small polymorphic drug molecules in the CSD . . . . . . . . . . . . . . . 83

4.4 Summary and future perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . 84

III Assessment of quantum chemical predictions in chemistry 91

5 Quantum Chemical Electron Impact Mass Spectrum prediction for de novo

structure elucidation 93

5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

5.2 Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96

5.2.1 Selection of compounds for analysis . . . . . . . . . . . . . . . . . . . . . 96

5.2.2 QCEIMS and CFM-EI calculations . . . . . . . . . . . . . . . . . . . . . 96

5.2.3 Spectrum matching . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98

5.2.4 Ranking comparison . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98

5.2.5 Spectrum normalization for statistical comparisons . . . . . . . . . . . . 98

5.2.6 Estimates of the errors in the calculated peak heights . . . . . . . . . . . 99

5.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99

5.3.1 The QCEIMS method . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100

5.3.2 The CFM-EI method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102

5.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

5.4.1 Detailed analysis of the predicted spectra . . . . . . . . . . . . . . . . . . 103

5.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107

5.6 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

6 Basis set convergence of CCSD(T) equilibrium geometries using a large and

diverse set of molecular structures 115

6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116

6.2 Computational Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117

6.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118

6.3.1 Overview of the molecules in the W4-11 database and reference geometries118

6.3.2 Basis set convergence of bond distances . . . . . . . . . . . . . . . . . . . 120

xi

CONTENTS

6.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132

6.5 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133

7 Estimating the CCSD basis-set limit energy from small basis sets 141

7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142

7.2 Computational Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143

7.3 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145

7.3.1 Performance of conventional basis set extrapolations for obtaining the

CCSD correlation component of the TAEs in the W4-11 dataset . . . . . 145

7.3.2 System-dependent basis set extrapolations of the CCSD correlation energy147

7.3.3 Performance of basis-set additivity schemes for obtaining the CCSD

correlation component of the TAEs in the W4-11 dataset . . . . . . . . . 151

7.3.4 Performance of basis-set extrapolations and additivity schemes for the

CCSD correlation energy for TAEs of larger molecules . . . . . . . . . . 153

7.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155

7.5 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156

IV Conclusion 163

8 Concluding remarks and future prospects 165

8.1 Molecular shape . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165

8.2 Assessment of quantum chemical predictions . . . . . . . . . . . . . . . . . . . . 167

Appendices 171

A Journal Article: Approaching an experimental electron density model of the

biologically active trans-epoxysuccinyl amide group—Substituent effects vs.

crystal packing 173

B Journal Article: Intermolecular interactions in molecular crystals: what’s in

a name? 177

C Journal Article: CrystalExplorer model energies and energy frameworks:

extension to metal coordination compounds, organic salts, solvates and open-

shell systems 181

D Description of CFM-EI and QCEIMS methods for prediction of mass spectra185

D.1 Competitive Fragmentation Modeling (CFM) . . . . . . . . . . . . . . . . . . . 185

D.2 Quantum Chemical prediction of Electronic Impact Mass-spectra (QCEIMS) . . 186

xii

CONTENTS

E Cartesian multipole expansion of electrostatic interaction energy 191

F Software: Simple Binary Format 195

F.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195

F.2 Performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196

xiii

CONTENTS

xiv

List of Figures

1.1 Relative abundance of words in the Google Books English corpus from years

1800-2000 (out of all words in this corpus). Note the rise of the terms ‘Science’

over the past 150 years, and ‘Computer’ or ‘Software’ since 1960, along with their

apparent correlation, indicating their fundamental connection. . . . . . . . . . . 4

1.2 A broad overview of the cyclical nature of the scientific method. The cyclical

nature is intended to demonstrate how our knowledge informs the way we design

experiments (or theoretical calculations), and from the resulting data we distill

information, which in turn may affect our knowledge of the world. . . . . . . . 6

1.3 2D Voronoi diagram vs. a Hirshfeld surface . . . . . . . . . . . . . . . . . . . . . 14

1.4 Visual representation of distance to the nearest internal atom (di) and nearest

external atom (de) for the Hirshfeld surface of a urea molecule in its crystal

structure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

1.5 Behaviour of a GTO (green) vs. a STO (blue), note the behaviour around the

cusp (0 A) and the long range behaviour (i.e. past 1 A). This is partially overcome

by using multiple GTOs to model each STO . . . . . . . . . . . . . . . . . . . . 21

1.6 Visualization of the occupied and virtual orbitals under HF and single and double

excitations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

1.7 Some examples of common representations and descriptions of chemical structure,

including skeleton/2D representations, 3D ball and stick models, IUPAC naming,

SMILES, InChi and excerpts from a text file representation of a Z-matrix and

cartesian geometries. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

3.1 Views of (a) crystalline environment in the co-crystal of the anti-inflammatory

drug indomethacin with nicotinamide, and (b) corresponding Hirshfeld dnorm

surface around indomethacin. Close contacts appear as red regions, while more

distant interactions will appear from white to blue. . . . . . . . . . . . . . . . . 55

xv

LIST OF FIGURES

3.2 From left to right, reconstructed (lmax = 9 and lmax = 20) and original Hirshfeld

surfaces for BENZEN07 (top) and INDMET (bottom). Surfaces have been

coloured based on the dnorm property at each vertex. While the reproductions at

lmax = 9 are not exact, descriptions at this level clearly capture the essential idea

of the shape of the HS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

3.3 2D PCA for the metallic crystals dataset, with squares indicating CCP structures,

and hexagons indicating HCP structures. . . . . . . . . . . . . . . . . . . . . . . 57

3.4 Hirshfeld surfaces for 3 CCP metals and 3 HCP metals with dnorm mapped on

the surface. Note the distinct patterns in both the shape of the surface and dnorm

correspond to the lattice structure in the crystal environment, along with the

heightened tendency toward asphericity as the packing becomes ’tighter’ (closer

interatomic distance with regard to electron density). . . . . . . . . . . . . . . . 58

3.5 Molecular structures of the three classes of substituted benzenes, naphthalenes

and phenylbenzamides and their pyridine analogues examined in this study. Note

that the pyridine ring N atom and R group may have varying positions. . . . . 59

3.6 (a) 2D PCA projection of selected benzene, naphthalene and phenylbenzamide

scaffolds , and (b) the same projection coloured by clusters assigned using

HDBSCAN. In both plots, circles are used to indicate phenylbenzamide type

scaffolds, squares to indicate naphthalene type scaffolds, and triangles used to

indicate benzene type scaffolds. . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

3.7 (a) A histogram of the interplanar angles between the two phenyl rings in

the phenylbenzamides. Note the distinct peaks around 0-20°and 60°, with a

more diffuse region between 60°and 90°, and (b) A 2D PCA plot of the set of

phenylbenzamides alone, again coloured by the clusters from HDBSCAN. . . . . 61

3.8 Hexagonally binned 2D histogram of the first two principal components from

invariants up to lmax = 9 along with the mean radius of all 14,772 structures.

Note that the the region highlighted as red square in the histogram represents

closely related structures in 10-dimensional principal component space, and not

necessarily the components in the 2D PCA. The similarity in corresponding

molecular shapes can be visualised in the representative structures contained

within this cluster. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

4.1 The ligand KNI-1689 (from PDB refcode 3A2O) in its binding pocket (left), and

Hirshfeld surface and surrounding region decorated with electrostatic potential

calculated using B3LYP/cc-pVDZ (right) . . . . . . . . . . . . . . . . . . . . . . 71

4.2 2D chemical structures of the four ligands presented in this study. . . . . . . . . 73

xvi

LIST OF FIGURES

4.3 Top: The Hirshfeld-ESP pharmacophore model generated from ligand (ganciclovir)

1 and the Hirshfeld- ESP surfaces of matching ligands discovered from the CSD;

bottom: the molecular structures of ligand 1 (ganciclovir) and the ligands showing

significant differences in their chemical structures. Docking scores included. . . . 79

4.4 Violin plots of the distributions of the docking scores of the top 100 shape-dnorm

matching ligands from the CSD database, for the four protein ligands and coloured

according to surface type. The thickness of the violins represents the relative

density of scores in that region. In the centre of each plot is a box-and-whisker

plot of docking scores. The red line represents the score of the reference ligand in

the protein-ligand complex. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80

4.5 Top 20 matches for small-molecule drugs with polymorphic structures in the CSD

(refcodes listed in Table 4.2) for Hirshfeld-dnorm, Promolecule-dnorm Hirshfeld-ESP,

Promolecule-ESP shape-property descriptors. Each square represents a match,

with colour indicating if it is either a polymorph of the molecule in question or

re-determinaton of the same crystal structure (green), or some other molecule.

The mean rankings are also given for each case. . . . . . . . . . . . . . . . . . . 85

5.1 Structures of the molecules chosen for analysis grouped into chemical classes . . 97

5.2 The number of molecules predicted correctly at a given rank for the QCEIMS

method. There were four molecules which were not matched to any molecule in

the NIST database. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100

5.3 Difference between predicted (QCEIMS/CFM-EI) and experimental (NIST) spec-

tra. Green bars indicate the presence of a correct peak, yellow indicates a missing

peak, and red indicates a peak which is incorrect i.e. should not have been

predicted. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104

5.4 Difference the between predicted (QCEIMS) and experimental (NIST) mass

spectrum for ethylbenzene. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107

6.1 Linear correlation between CCSD(T)/A′V6Z and CCSD(T)/VDZ bond distances

(in A) for the 181 bond lengths in the GEOM-AV6Z dataset. . . . . . . . . . . . 125

7.1 Normal distributions of system-dependent extrapolation exponents (αideal), ob-

tained from eq. 7.6 for the 140 molecules in the W4-11 database. The ideal

exponents reproduce the CCSD/CBS energy for each molecule from the (a)

A′V{D,T}Z and (b) A′V{T,Q}Z basis-set pairs. The Gaussians are centered

around 2.22 (A′V{D,T}Z) and 3.46 (A′V{T,Q}Z), and have standard deviations

of 0.50 and 0.25, respectively. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148

xvii

LIST OF FIGURES

xviii

List of Tables

1.1 Composition of primitive and contracted Gaussian basis functions for the correlation-

consistent (cc) basis sets (adapted from Jensen [165]) . . . . . . . . . . . . . . . 22

1.2 Terms for the different kinds of DFT functionals, including local density ap-

proximations (LDA), and generalized gradient approximations (GGA). The first

column indicates whether the functionals in this class utilise the density, ρ, its

derivatives and/or incorporate HF exchange (EHFx ). . . . . . . . . . . . . . . . . 28

4.1 Protein-ligand co-crystals examined within this study, with an indication of the

nature of their biological activity . . . . . . . . . . . . . . . . . . . . . . . . . . 72

4.2 Small drug molecules exhibiting polymorphism with their Cambridge Structural

Database (CSD) refcode(s). Those that appear in bold were used as the ‘query’

polymorphs for this investigation, and their references are provided. Refer to

text for details. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

4.3 Comparison of the root mean-squared deviation (RMSD), mean-signed deviation

(MSD) and mean absolute-deviation between the exact and approximate electro-

static potentials (ESPs) (in au) for the ligand in the ligand-protein complex using

electron densities for ligand wavefunctions at the B3LYP/6-31G(d,p) level. Also

shown are the number of atoms in the ligand natoms, the number of points in the

isosurfaces npoints, and representative times for the calculations of the exact ESP

texact and for the approximate ESP (tb=1) where the multipole approximation is

used for atoms separated by more than two bond connections. Refer to text for

further details. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

4.4 Pearson correlation coefficients R between shape-property descriptors used in this

paper and and docking scores from FRED/Chemgauss4 for the top 100 matches.

Refer to text for details. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82

4.5 Ligands found in the top 100 matches which are labelled as drugs in the Cambridge

Structural database. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

xix

LIST OF TABLES

5.1 Relative ranking position (RRP) of the matching mass spectra found by the

QCEIMS program produced by the NIST MS Search program (version 2.0g).

Ranking is relative to the total number isomers nisomers. NM indicates that no

match was found in the database. . . . . . . . . . . . . . . . . . . . . . . . . . . 99

5.2 Mean Relative ranking position (RRP) for the QCEIMS and CFM-EI methods

for different classes of compounds, with the number of compounds in each class.

Overall mean RRPs are given in the final line. Only molecules in which both

methods matched are compared. . . . . . . . . . . . . . . . . . . . . . . . . . . . 102

5.3 Mean root mean square deviation (RMSD) and mean absolute deviation (MAD)

for the QCEIMS and CFM-EI methods for different classes of compounds, .

Overall mean RMSDs and MADs are given in the final line. . . . . . . . . . . . 103

6.1 Overview of the 122 molecules in the W4-11-GEOM database for which we were

able to obtain CCSD(T)/A′V6Z or CCSD(T)/A′V5Z reference geometries. . . . 119

6.2 Overview of the bonds in the W4-11-GEOM and GEOM-AV6Z datasets (for

additional details see Table S1 of the Supporting Information). . . . . . . . . . . 120

6.3 Overview of the basis set convergence of the CCSD(T) and CCSD(T)-F12 methods

for the 181 unique bond lengths in the GEOM-AV6Z dataset (A). The reference

values are CCSD(T)/A′V6Z bond distances. . . . . . . . . . . . . . . . . . . . . 121

6.4 Computational resources used for the CCSD(T)/VnZ and CCSD(T)-F12/VnZ-

F12 single-point energy calculations for naphthalene and anthracene. . . . . . . 124

6.5 Overview of the performance of scaled and extrapolated CCSD(T) and CCSD(T)-

F12 bond distances for the 181 unique bonds in the GEOM-AV6Z dataset (A).

The reference values are CCSD(T)/A′V6Z bond distances. . . . . . . . . . . . . 126

6.6 Overview of the basis set convergence of the CCSD(T) and CCSD(T)-F12 methods

for the 75 H−X, 49 X−Y, 43 X−−Y, and 14 X−−−Y bonds in the GEOM-AV6Z

dataset (RMSDs, in A) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127

6.7 Overview of the performance of the CCSD(T) and CCSD(T)-F12 methods for

the 246 unique bonds in the W4-11-GEOM dataset (A). The reference values are

108 CCSD(T)/A′V6Z and 14 CCSD(T)/A′V5Z bond distances (See Table 6.1). . 129

6.8 Effect of CCSD(T) and CCSD(T)-F12 reference geometry on molecular energies

calculated at the CCSD(T)/CBS level of theory for the 108 molecules in the

GEOM-AV6Z database (kJ mol−1). . . . . . . . . . . . . . . . . . . . . . . . . . 131

7.1 Error statistics for global extrapolations of the valence CCSD correlation com-

ponent over the 140 total atomization energies in the W4-11 dataset (kJ mol−1).

The reference values are CCSD/A′V{5, 6}Z total atomization energies from W4

theory. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146

xx

LIST OF TABLES

7.2 Error statistics for the system-dependent extrapolations of the valence CCSD

correlation component over the 140 total atomization energies in the W4-11 test

set (kJ mol−1) . The reference values are CCSD/A′V{5, 6}Z total atomization

energies from W4 theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150

7.3 Error statistics for the valence CCSD correlation component over the 140 total

atomization energies in the W4-11 test set obtained via the CCSD/X(MP2/Y)

additivity scheme kJ mol−1. The reference values are CCSD/A′V{5, 6}Z total

atomization energies from W4 theory. . . . . . . . . . . . . . . . . . . . . . . . . 152

7.4 Error statistics for the global and system-dependent extrapolations of the valence

CCSD correlation energies over the 20 total atomization energies in the TAE20

dataset (kJ mol−1). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154

F.1 Relationship between SBF data types and their respective declaration in different

languages . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196

xxi

LIST OF TABLES

xxii

List of Abbreviations and Symbols

Abbreviations

ANO Atomic natural orbitals

BO Born-Oppenheimer

CBS Complete basis-set

CC Coupled cluster

cc correlation-consistent

CCSD Coupled cluster with single and double excitations

CCSD(T) Coupled cluster with singles, doubles and quasi-perturbative triples

CFM Competitive fragmentation modeling

CI Configuration interaction theory

CIF Crystallographic information file

CSD Cambridge Structural Database

DFT Density functional theory

EIMS Electronic impact mass spectrometry

GGA Generalised gradient approximation

GTO Gaussian-type orbital

HF Hartree-Fock

HS Hirshfeld surface

IAM Independent Atom Model

xxiii

LIST OF ABBREVIATIONS AND SYMBOLS

IR Infra-red (spectrometry)

IUPAC International Union of Pure and Applied Chemistry

LBVS Ligand-based virtual screening

LCAO Linear combination of atomic orbitals

LDA Local density approximation

MD Molecular Dynamics

ML Machine learning

MO Molecular orbital

MP Møller-Plesset perturbation theory

MS Mass spectrometry

NMR Nuclear magnetic resonance

PCA Principal component analysis

PDB Protein Data Bank

QCEIMS Quantum Chemical Electronic Impact Mass Spectrum

RS Range-separated

RSPT Rayleigh-Schrodinger perturbation theory

SBVS Structure-based virtual screening

SCF Self-consistent field

SMILES Simplified molecular-input line-entry system

STO Slater-type orbital

UV/VIS Ultra-violet/visible light (spectrometry)

xxiv

LIST OF ABBREVIATIONS AND SYMBOLS

Symbols

χ A basis function

de Distance from a point on the isosurface to the nucleus of the nearest external atom

di Distance from a point on the isosurface to the nucleus of the nearest internal atom

dnorm Combined distance from point on the isosurface to the nucleus of the nearest internal

and external atoms, normalized by van der Waals radii

E Energy

H Hamiltonian operator

Te Electronic kinetic energy operator

Tn Nuclear kinetic energy operator

∇ Laplacian operator

Z Nuclear charge

Vee Electron-electron electrostatic interaction operator

Ven Electron-nuclei electrostatic interaction operator

Vnn Nuclei-nuclei electrostatic interaction operator

r Particle coordinates

ρ Density (electronic), usually at a given position

Ψ Wavefunction of a system

Y ml Spherical harmonic function

xxv

LIST OF ABBREVIATIONS AND SYMBOLS

xxvi

Part I

Introduction and Background

1

Chapter 1

Background

1.1 Foreword

This thesis is primarily focussed on the development, evaluation and application of computational

techniques and their relation to physical chemistry problems. All of the work in this thesis

required genuine understanding of not only the science behind the work – be it crystal structures,

electron impact mass spectra or electronic structure theory – but also an in depth knowledge of

possibilities made available by contemporary computers, programming languages and software.

It is imperative that we understand the role that computation and theory play in contemporary

science: not only as tools to ‘get the job done’, but as bona fide pillars of science, capable

of changing our understanding and our scientific worldview. Computers are now ubiquitous

in science, and the impact of advances in computer science, informatics and mathematics on

chemical understanding is not to be underestimated. Only by understanding scientific problems

and their solution from the top-to-bottom, from the high-level abstract ontologies of science

down to the details of their representation on modern computer architectures is it possible

to solve some new scientific problems, or shift perspectives when addressing existing scientific

problems. To that end, the following chapter is intended to provide philsophical, historical and

theoretical context in order to help understand the work presented in the body of the thesis.

1.2 Science and computers

1.2.1 History and philosophy

The overwhelming majority of scientific research now relies on software of some kind. There is

little doubt that the ‘microcomputer revolution’ in the 1960s [1], coincided with an immense

shift in the importance of computer hardware and software in society. Likewise, the roles

performed by computers in science have changed in both nature and prevalence, moving beyond

mainstream and into ubiquity. This shift of computers into ubiquity is reflected in our language,

3

1.2. SCIENCE AND COMPUTERS

concepts of ”techne”, which refers to know-how or technology, and ”episteme”, which refers to

(particularly scientific) knowledge. This duality provides a concise framework in which we can

explore some of the impact of computers in science, without focussing too much on epistemology

from a philosophical standpoint.

For the purpose of providing a background to this thesis, it is primarily important to examine

the roles computers have played throughout their history in chemistry and chemical problems,

how those have changed, and where the current state of affairs lies on this landscape. To do

so, it is worth not only discussing chemistry in isolation, but bringing in broader scientific,

technological and philosophical context.

1.2.2 Data, information, knowledge and science

A computational perspective fundamentally alters our perceptions and attitudes about data, but

at first glance it may seem unlikely to some that it alters our sense of scientific reality. However,

the scientific method is built upon data, information and knowledge and their relation to one

another. From data we derive information, and from information we can derive knowledge.

Data, information and knowledge elude straightforward and precise definitions, and as the focus

of this thesis is not epistemology simple working definitions will suffice here.

Data can be understood as facts of the world – with or without context (often referred to as

metadata i.e. data about the data). Facts, in this case, can simply be true statements about

the world. Data may come in the form of scientific observations from a person, raw frames from

a digital camera or any other of the myriad data sources in contemporary science.

Information, then, is the encapsulation of data – be it summarised or annotated or whatever.

As an example, let us say I measured the mass of an electron through some experiment. I can

now report this mass verbally, write down the number in whatever units I wish – perhaps even

send it to a friend via a postcard or an email. None of this changes the actual mass of the

electron, I am sending information, but the mass itself is data.

Finally, knowledge is – to put it tautologically – what we know. It is a true, justified belief,

and may not be about the physical world at all. For example, I know my feelings, I know that

I exist (and that perhaps you do too!). If I know the mass of the electron, I can encode that

knowledge in information. Perhaps I can use the same knowledge to predict the mass of some

other particle or other based on a theoretical framework. Knowledge requires a subject (i.e.

me, or you etc.) and an object (the world, the electron etc.). The aforementioned concepts of

episteme and techne refer to knowledge, and attempt to subdivide that concept into knowledge

of the world (episteme) and practical knowledge or know-how (techne).

The relationship between data, information and knowledge is often visualized as a hierarchy

[5, 6] (also typically including wisdom), but in Figure 1.2 it is visualized as a cyclic graph.

This representation reinforces that not only does information provide us with knowledge, but it

5

1.2. SCIENCE AND COMPUTERS

as valid a concept as temperature. Indeed, some theories prove valuable before their reification.

For example, the existence of genes as a scientific concept predates their physical realisation in

DNA etc.

A theory can also be productive2 – even if it is not really correct. Perhaps the best historical

example of this is phlogiston, a substance which (in the 18th century) was said to exist in all

combustible materials and was released upon combustion. Of course, we now know that the

concept of phlogiston is clearly wrong (and indeed as a concept it was proven so by weighing

mercury before and after combustion – the mass increased rather than decreased). Nevertheless,

phlogiston theory helped scientists find the structure of N2, CO2, and even that water was not

an element but a compound.

Scientific reductionism, where some properties or entities are ‘nothing over and above’ more

fundamental entities and their interactions, is arguably the dominant scientific paradigm. As the

underlying causes for observable properties or previously held concepts are identified, concepts

may be be reduced to simply being useful but not ‘real’ (e.g. Lewis diagrams, bond order etc.)

or eliminated entirely (as was the case with phlogiston). These concepts are unreal only insofar

as they they do not reflect accurately the underlying physics, yet they still describe nature

sufficiently to prove useful in communication or teaching.

Indeed, scientific reductionism can be clearly seen in the following (all too common) quote

from Dirac [10]:

The underlying physical laws necessary for the mathematical theory of a large

part of physics and the whole of chemistry are thus completely known, and the

difficulty is only that the exact application of these laws leads to equations much too

complicated to be soluble. It therefore becomes desirable that approximate practical

methods of applying quantum mechanics should be developed, which can lead to

an explanation of the main features of complex atomic systems without too much

computation.

It is such approximations that form the majority of modern quantum chemistry and theoretical

chemistry, along with refining and developing the conceptual framework used in chemical

information. Computational chemistry, and by extension cheminformatics, then, provide us

the tools or techne that allow us to develop and test new approximations. Further, the

approximations and know-how (techne) we utilise to discuss scientific observations inform

our conception of reality (episteme), which elucidates the clear motivation in ensuring their

correctness – or at least adherence to reality.

2During his plenary talk at IUCr Florence 2005 [9] Roald Hoffmann said of Bader’s quantum theory of atomsin molecules that it was ”neither predictive nor productive.” Putting aside the accuracy of this claim, I believethe notion of predictive or productive encapsulates nicely two of the major values a theory can have in science:its capacity to provide accurate predictions, or how it might aid us in being productive through its concepts.

7

CHAPTER 1. BACKGROUND

1.2.3 Informatics and computer science

Informatics is the practice of information processing and engineering information systems.

Evidently, it exists at the intersection of scientific theory, concepts and their representation in

computers. Informatics is essential to science as it enables and expedites research that may

otherwise have been proven impossible, but also because the representation of information

influences our perception of reality. Consider, for example, the pervasiveness of the two-

dimensional structural diagram of molecules. From this simple abstraction, notions of chemical

bonding, Lewis structures, Natural Bond Orbital (NBO) theory, Valence Bond (VB) theory and

many more ideas find their grounding.

A significant change in the relationship between computers and science more recently has

been the shift from computational power as the ‘bottleneck’ in much scientific research. As

progress in computer throughput largely followed Moore’s law [11] (until recently [12]), a large

portion of scientific problems are no longer really limited by computation time but instead by the

time it takes to understand the data. Indeed, while scientific practice has undoubtedly changed

as a result of computers, the limits of human comprehension have not. An individual may be

able to process or understand hundreds or even thousands of chemical compounds, positions,

calculations. Perhaps some may even be able to comprehend tens of thousands. But hundreds

of thousands, or millions? It seems clear that there is an order of magnitude more chemical

information than a human can fathom, and past that we are entirely reliant on computers and

electronic storage of data. It is for the purpose of processing this vast amount of scientific data

that the field of informatics3 has burgeoned since the microcomputer revolution. Indeed, what

can be said to be the most transformative technological change for society over the last few

decades – the growth of the world wide web, and particularly search engines like Google – may

be readily understood as an informatics project.

Despite often being referred to as an ‘emerging’ discipline, informatics is not especially

new – or at least only insofar as computers are new or emerging. Informatics, or at least

the German Informatik was essentially synonymous with ‘computer science’ (and there is still

significant overlap). Further, much of the work performed in the early 20th century which

would be called informatics in contemporary science existed long before such a classification

was widespread – it was simply a part of the process of science or chemistry, i.e. information

processing. Such references to informatics as an emerging discipline, then, are more a result

of the rapidly expanding role of computers in science bringing these concepts to the forefront

within the broader scientific consciousness than the concepts themselves being new. This rapidly

expanding role is particularly exemplified by the burgeoning field of bioinformatics from its

roots in the late 20th century to the present, and even now may be seen through the changing

language in technology focused jobs. Increasing use of terms like ‘data science’, and ‘machine

3The term informatics is thought to have been coined by Karl Steinbuch in his 1957 book Informatik:Automatische Informationsverarbeitung i.e. ”Informatics: automatic information processing”

8

1.2. SCIENCE AND COMPUTERS

learning’ in job advertisements and consumer electronics make it plainly evident that the roles

of data and information analysis strategies are of increasing importance.

1.2.4 The role of computers in chemistry

Chemistry has as long and storied a history with computers as any other science. As such, this

section will focus only on areas touched on within this thesis: the intersection of computers and

cheminformatics developments with quantum chemistry, crystallography and electron impact

mass spectrometry.

Computational chemistry may broadly be described as chemistry explored using computers

rather than in the lab. It is something of a nebulous term, sometimes encompassing chemin-

formatics, statistical mechanics, molecular mechanics, semi-empirical methods and ab initio

quantum chemistry. Computational chemistry is not (and should not be) conceived as a re-

placement for experimental studies, but instead as in enabling chemists to both explain and

rationalise known chemistry or to explore new or unknown chemistry in ways not possible

through experiment.

Cheminformatics4 has been explained as “the design, creation, organization, management,

retrieval, analysis, dissemination, visualization and use of chemical information” [14], and

elsewhere as “the application of informatics methods to solve chemical problems” [13].

Prior to the advent of computers, quantum chemistry was, like any other mathematical

method, computed by hand. The limitations this enforced are evident in the early methods in

quantum chemistry which had to utilise significant approximations and simplifications to make

calculations achievable (e.g. using one-electron wave functions [15, 16]), and even then were

limited by the scale of the systems (generally limited to atoms or diatomic systems, or exploiting

symmetry [17]) Huckel theory [18], for example, allowed calculations of energy levels of the

π-electrons in conjugated hydrocarbons like ethene, butadiene and even benzene.5 Without the

simplification of these problems they were simply intractable, and indeed this may be readily

seen in the collections of Huckel theory molecular orbitals published by Heilbronner [20], and the

tables of molecular wavefunctions published (see McLean & Yoshimine [21] and Snyder & Basch

[22]), where largely diatomics or linear molecules (where confocal elliptic coordinate systems

could be used, greatly simplifying the calculations) are evaluated. Now, a simple Hartree-Fock

SCF calculation for these sorts of molecules takes seconds with computer programs. There is

little doubt that computers completely changed the scope and scale of quantum chemistry.

Likewise, it’s no exaggeration to state that computers revolutionised X-ray crystallography.

Prior to their use, it involved intensive calculations by hand, limiting the kinds of crystals which

4It has been noted that the terms cheminformatics and chemoinformatics occur with approximately equalfrequency in the literature [13]; throughout this thesis cheminformatics will be used

5It was later extended by Roald Hoffmann to include sigma orbitals [19], making it a bona-fide semi-empiricalquantum chemistry method.

9

CHAPTER 1. BACKGROUND

could be characterised.6 As computers have gotten faster, crystallographers have been more and

more ambitious in their investigations. The current state-of-the-art in crystallography might be

said to be the use of X-ray free electron lasers [24] in studying the structure of molecules in

nanocrystals. For this to be successful, in just one example [25] – solution of a crystal structure

of rhodopsin – 5 million detector frames were collected, including diffraction patterns from

18,874 crystals. It should be immediately apparent that without (fast) computers, such studies

would simply not be possible.

It was recognised early on that the amount of information embedded in mass spectra (MSp)

was too difficult for an individual to process, and so the DENDRAL [26] project was proposed.

The goal of the project was, broadly, to aid organic chemists in identifying unknown organic

molecules, by analyzing their MSp – using knowledge of chemistry to design a computer system

capable of performing the task. It is telling that the DENDRAL project is still cited as one of

the earliest examples of ‘expert systems’ on computing; it is one of many examples of scientific

goals pushing the boundaries of computing.

The dual motivation for cheminformatics, then, becomes clear: a) Many problems in chemistry

are too complex to be solved by methods based on first principles through theoretical calculations

and b) chemistry produces a huge amount of data.

Though they have different names and terminology, there is often no hard separation between

bioinformatics and cheminformatics, just as there is often no clear separation between chemistry

and biochemistry. The two fields are inextricably related. Likewise, it is common to see

terms like computational chemistry and molecular modelling used both interchangeably and

as separate monikers. Some would definitely claim there to be a sharp distinction between

cheminformatics and computational chemistry, with the latter primarily focused on theoretical

quantum mechanical calculations [13], but the boundaries between these fields are not always

clear, and certainly much of the work within this thesis blurs the boundaries.

1.3 Structural and theoretical chemistry

1.3.1 Chemical characterisation

Chemical characterisation, the determination of chemical structure and properties is such a

fundamental process in chemistry and materials science, that without it little to no scientific

understanding of chemicals and materials could be ascertained. Throughout the last few centuries,

a wide variety of characterisation methods including microscopy, spectroscopy, thermal analysis,

ultrasound and many more have been utilised to determine chemical structure or composition.

Arguably, the three current ‘workhorse’ methods7 for chemical characterisation are mass

6Dorothy Crowfoot Hodgkin was one of the first to use computers to solve crystal structures, the use ofHollerith machines aided in solving the structure of penicillin between 1937-1947 [23]

7Typically supplemented by other methods like infra-red spectra, ultra-violet/visible spectra, spectroscopic

10

1.3. STRUCTURAL AND THEORETICAL CHEMISTRY

spectral methods, including electron impact mass spectra (EIMS), crystallographic methods (i.e.

X-ray and neutron diffraction) and nuclear magnetic resonance (NMR). For the sake of brevity,

and because no work in this thesis focuses on NMR, only the first two methods will be detailed

in this section (crystallography and EIMS).

Crystallography, since pioneering work in X-ray diffraction (XRD) by William and Lawrence

Bragg [27] has been an immensely powerful characterisation tool – in many senses the ‘gold

standard’. Its use has led to the structure of many of the most important chemicals – biological

or otherwise. Indeed, this has been recognised through over 25 Nobel Prizes associated with

crystallography since 1914. The Braggs first used it in concert with NaCl crystals – demonstrating

the absence of molecules and salt’s structure as an ionic lattice. In 1928, Kathleen Lonsdale

reported the structure of benzene [28] – having six equal sized bonds and not alternating

double and single bonds. Dorothy Crowfoot Hodgkin won the Nobel Prize in 1964 for ”for

her determinations by X-ray techniques of the structures of important biochemical substances”

which included cholesterol, penicillin, vitamin B-12, insulin and many more [23]. To this day,

X-ray and neutron diffraction and crystallography are of the utmost importance in determining

chemical structure.

Mass spectrometry (MS), on the other hand, has its roots in J.J. Thomson and E. Everett’s

exploration of cathode rays, measuring e/m (i.e. the ratio of electronic charge e to electronic

mass m) in 1897. Two years later, they built an instrument that could simultaneously measure

e/m and e, i.e. indirectly measuring the mass of the electron. For this work in “discovering”

the electron, Thomson received the 1906 Nobel Prize in Physics. By stark contrast with the

contemporary state of affairs, until the 1940s MS was still dominated by physicists, predominantly

used to answer fundamental questions about the nature of atoms. Today different MS techniques

are used throughout chemistry and biochemistry in the identification of natural products [29],

metabolomics [30] and many other chemical problems.

1.3.2 The role of computers in chemical characterisation

Crystallographers were early adopters of computers, primarily to enable and speed up the

calculation of Fourier syntheses. Beginning mainly with home-brew analog computers, by the

late 1940s they gradually shifted to IBM punchcard tabulators programmed via plugboards [31].

The first crystallographic applications of programmable computers were done on EDSAC [32]

and the Manchester Mark II [33] in 1952–1953 on inorganic crystals. The first application of

computers to protein crystallography was also for the first high-resolution structure, that of

myoglobin, in 1958 [34].

By the 1960s, crystallographers were enthusiastic users of computers, not only for core primary

calculations but for many related routines as well; this even extended to visualization, including

methods etc.

11

CHAPTER 1. BACKGROUND

interactive molecular graphics first done by Cyrus Levinthal at the Massachusetts Institute

of Technology, who utilised an oscilloscope display connected to a time-sharing mainframe to

render a wireframe model [35].

Crystallography, in concert with computation, has been so prolific that there are now over

875 000 molecular crystal structures determined and available in the Cambridge structural

data base (CSD), and over 125 000 structures in the Protein Data Bank (PDB). These crystal

structures provide a wealth of experimental information regarding three dimensional chemical

structure, intermolecular interactions, biological activity and more.

With this background in mind, it is worth going into some more detail on the intersection of

crystallography and theoretical chemistry, which informs a huge part of the background and

motivation for Part II of this thesis.

1.3.3 Crystal engineering

The notion that macro-level properties of molecular crystals such as their shape, heat capacity

and plasticity are significantly determined by the micro-level properties of their constituents

can be said to be the primary basis of crystal engineering [36]. The major problem for crystal

engineering, then, is how to predict a crystal structure or associated physical properties of a

crystal from knowledge of its constituents. With the exception of specific cases such as the

aluminosilicates (zeolites) [37] and more recently the metal-organic frameworks (MOFS) [38]),

this problem remains unresolved. As a result, it is natural to pose the following question:

how best might we define a ‘constituent’ to predict and control molecular crystal structure?

The emphasis here is on organic molecular crystals, as these molecules can be constructed

more-or-less at will, unlike more general inorganic or metal-organic molecules.

One of the key notions utilised in crystal engineering is that of a ‘supramolecular synthon’

[39]. This term borrows from the term ‘synthon’ [40] as applied in organic chemistry – where

it was originally used in retrosynthesis to describe a unit within a molecule which is related

to a potential synthetic operation. A ‘supramolecular synthon’, is considered to be a building

block in supramolecular interactions i.e. in crystallisation. Canonical examples would include

carboxylic acid dimers, where the hydrogen bonding is strong enough that it makes it extremely

likely two molecules will bind through this interaction. It is difficult to say, however, how useful

the notion of supramolecular synthons becomes in structure determination when moving beyond

the canonical examples of hydrogen bonding.

1.3.4 The promolecule

The term promolecule was first used by Hirshfeld & Rzotkiewicz [41], and refers to a reference

electron density based on the Hartree product (see section 1.4.3) of spherical atomic wavefunctions

in a molecule. It is equivalent to the independent atom model (IAM) frequently used in electron

12

1.3. STRUCTURAL AND THEORETICAL CHEMISTRY

scattering and X-ray crystallography, and being well defined quantum mechanically, it constitutes

an approximate molecular wavefunction which can be used to make predictions based on electron

density.

Clearly, the promolecule wavefunction is not adequate for chemical accuracy in predictions

for some properties, but nevertheless the total electron density is not as far from ‘proper’

approximations to the wavefunction as one might conceive. Further, the deviation in electron

density from the the promolecule – typically referred to as the ‘deformation density’ – can be

used to provide insights into chemical bonding and other intramolecular properties [42].

1.3.5 Atoms in molecules

Atoms do not explicitly arise in quantum mechanics: all that appears is a Hamiltonian in terms

of interparticle interactions, where electrons and nuclei are the particles generally considered.

Nevertheless, the idea of an atom is so conceptually essential in chemistry that there have

repeated efforts to provide satisfactory definitions within the quantum mechanical framework.

The earliest (and possibly most complex) effort toward this end came from Moffitt [43],

who proposed that an ‘atom’ should be described by an n-electron wavefunction made up from

one-electron atomic orbitals. The rise in popularity of density functional theory subsequently

led researchers to look for the definition of an atom in the electron density. Bader [44] proposed

using a ‘zero-flux surface’, where flux is understood to mean the gradient of the electron density,

in order to define an atom. Such atoms have the property that their kinetic energy has a well

defined form, when evaluated using laplacian and gradient-squared forms. Bader’s atoms have

been used extensively for analysing experimental electron densities [45], however their derivation

from Schwinger’s principle has been questioned [46, 47]. Other, less popular, kinds of atoms

based on the electron density have been proposed by Gill [48] and Hunter [49]. Unlike density

based schemes, Roby [50] defined atoms in one-electron Hilbert space using projection operators,

and this was extended by Gould et al. [51] (a more complete discussion of these topics may be

found in Sukumar [52]). The most relevant method to this work is Hirshfeld’s atom partitioning

scheme [53] which is described in more detail in the following section.

1.3.6 Molecules in crystals: the Hirshfeld surface

The Hirshfeld surface (HS) [54–56] originally emerged from attempts to define the region of space

occupied by a molecule within a crystal structure. Construction of the HS involves partitioning

the electron density in a crystal structure into regions belonging to each constituent molecule

(akin to the Hirshfeld partitioning of a molecule into atoms [53]). In this way, the HS can be

considered as similar to a Wigner-Seitz [57] cell, or more generally a Voronoi diagram [58] which

partitions regions of space as belonging to a set of points, breaking it up into regions ‘belonging’

13

CHAPTER 1. BACKGROUND

idea is that chemistry is to be regarded as a collection of particles –protons and electrons –

interacting through fundamental physical forces, which are primarily electromagnetic. The

paradigm is reductive because chemistry, in effect, is reduced to the positions of these particles.

However, unlike classical mechanics, on which quantum mechanics is based, the properties of

these particles, for example their velocity or acceleration in a classical and macroscopic picture

which we are accustomed to are not directly related to their positions. Rather the particles are

described by a complex function (in the sense of complex numbers) of the particle positions,

and operators on these complex functions are associated with measurements of properties (such

as velocity and acceleration). In this section the details of this paradigm relevant to this thesis

are described.

1.4.1 The time-independent Schrodinger equation

It is fair to say that modern theoretical chemistry really began with quantum mechanics, and

the Schrodinger equation. Erwin Schrodinger proposed his non-relativistic wave equation [63, 64]

incorporating the earlier formulation of Heisenberg’s matrix mechanics. Schrodinger postulated

how a system, described by a wavefunction Ψ, evolves in time analogous to how Newton

showed how the positions of particles evolve in response to the forces applied to those particles.

However, chemistry is often concerned with time-independent properties. For such systems, the

relationship between the total energy of a system E and the wave function representing the

system Ψ is through the following eigenvalue equation:11

HΨ = EΨ (1.2)

where Ψ is a function of the particle coordinates, and H is the Hamiltonian (energy) operator

of the system, given by

H = Te + Vee + Ven + Tn + Vnn. (1.3)

This Hamiltonian contains kinetic energy terms for the electrons and nuclei (Te, Tn), and

potential energy terms for the electron-electron, electron-nuclear and nuclear-nuclear electrostatic

11Note that in this and the sections that follow, Ψ and all other operators are functions of particle coordinatesR; r including nuclear and electronic location and spin

16

1.4. QUANTUM AND COMPUTATIONAL CHEMISTRY

interaction operators (Vee, Ven, Vnn), defined by

Te = −∇2i

2(1.4)

Tn = − ∇2A

2MA

(1.5)

Vee =∑i>j

1

rij(1.6)

Ven = −∑A,i

ZArAi

(1.7)

Vnn =∑A>B

ZAZBrAB

. (1.8)

Here ∇ is the Laplacian, i, j denote electrons, A,B denote nuclei with Z charge and M mass

and r is the distance between two particles. All operators are in atomic units.

Solution of the Schrodinger equation allows prediction of any chemical or physical property

of an atom or molecule from knowledge of its wavefunction, typically via derivatives of the

energy with respect to some external parameters. For example, knowledge of the first derivative

of E with respect to nuclear coordinates is related to the force on that nucleus, and thus can be

used to find equilibrium geometries and transition structures.

1.4.2 The Born-Oppenheimer approximation and independent par-

ticle model

The Born-Oppenheimer approximation [65] (BO) allows separate treatment of the motion of

nuclei and electrons. The basis of this stems from the huge discrepancy in the mass of electrons

and nuclei, in addition to the premise that electronic velocity tends to far exceed that of nuclei.

The nuclei are then treated in a fixed position, while the electrons are are not.

In this manner, the BO approximation greatly simplifies the Hamiltonian in eq. 1.3 as the

nuclear kinetic energy term (Tn) does not act on ψn, and the nuclear-nuclear electrostatic energy

is constant. Equation 1.3 then becomes:

H = Te + Vee + Ven + Vnn (1.9)

A BO wavefunction then would be written as the product of a nuclear and an electronic

wavefunction:

Ψ = ΨeΨn (1.10)

where Ψe is the electronic wavefunction in the fixed nuclei field and Ψn is the nuclear wavefunction.

The electronic and nuclear Schrodinger equations, then, are given by the following equations:

17

CHAPTER 1. BACKGROUND

Heψe = Eeψe (1.11)

[Hn + Ee]ψn = Enψn (1.12)

Hn = Tn (1.13)

The electronic energy eigenvalues Ee depend on the position of the nuclei, so varying these

positions in small steps and repeatedly solving the electronic Schrodinger equation yields Ee as

a function of the positions of the nuclei, known as the potential energy surface (PES), which

governs motion of the nuclei. In other words, once the solution to Ee is known, the nuclear

Schrodinger equation may be solved.

1.4.3 Hartree-Fock theory

The simplest ab initio method is Hartree-Fock (HF), so called because of its roots in work by

Douglas Hartree [66, 67] and Vladimir Fock [68] (see also Slater [69]). HF addresses the problem

of exact many-body solutions to he Schrodinger equation by reducing two electron integrals into

a series of one electron terms and an averaged field, with the restriction that the solution is

expressed as a single Slater determinant.

Underlying the HF approximation is the variational principle: that the energy determined

from any approximate wavefunction is greater than the energy for the exact wavefunction. This

provides a straightforward method to approach the exact wavefunction: minimize the total

energy of the system.12

In addition to the BO approximation, HF relies on the independent electron approximation,

and utilises the linear combination of atomic orbitals (LCAO) approximation.

Perhaps the best way of understanding the independent electron model, is to examine the

Hartree product. A Hartree wavefunction Ψ may be conceived of as the product of individual

one-electron wavefunctions.

Ψ = ψ1ψ2 . . . ψn (1.14)

The individual one-electron wavefunctions ψn are typically called molecular orbitals (MO). This

form of the wavefunction does not allow for instantaneous interactions between electrons, instead

they feel the averaged field of all other electrons in the system. This is the independent electron

model.

However, one of the postulates of quantum mechanics is that the total wavefunction must

be antisymmetric with respect to interchange of electron coordinates (the Pauli Principle). The

12For nonrelativistic Hamiltonians where there is a lowest energy eigenavlue. The variational theorem can alsobe applied to effective relativistic Hamiltonians, even though the true Hamiltonain does not have a minimumenergy eigenvalue.

18

1.4. QUANTUM AND COMPUTATIONAL CHEMISTRY

Hartree product is not antisymmetric, but may be made so (and thus into a Hartree-Fock

wavefunction) by adding all signed permutations:

Ψ =1√N !

∣∣∣∣∣∣∣∣∣∣∣∣

ψ1(r1) ψ2(r1) · · · ψN(r1)

ψ1(r2) ψ2(r2) · · · ψN(r2)...

.... . .

...

ψ1(rN) ψ2(rN) · · · ψN(rN)

∣∣∣∣∣∣∣∣∣∣∣∣(1.15)

where r are the electronic coordinates. This equation is in the form of a Slater determinant,

and eq. 1.15 would typically be written as |ψ1ψ2 . . . ψN〉. This is what is meant when HF is

called a single determinant method. By antisymmetrising the Hartree product we ensure that

all electrons are indistinguishable, and therefore associated with every orbital.

By conceiving of the MOs (or one electron wavefunctions) as LCAO, we may express an MO

ψi as follows:

ψi =∑µ

cµiχµ (1.16)

where cµi are the MO coefficients and χµ are the atomic orbitals or basis functions.

Computing the HF energy implies computing the MO coefficients. To compute the MO

coefficients we must minimise the HF energy according to the variational principle. In order to

find MO coefficients, HF (and indeed the majority of quantum chemical methods) rely on a

so-called self-consistent field (SCF) approach, where an initial guess for the MOs is provided via

an effective Hamiltonian method, then iterated until some convergence criteria are met.

Due to the independent electron model (the averaged-field electron-electron interactions) HF

theory neglects the so called ‘correlation’ of electrons. Methods that incorporate the remaining

correlation energy (Ecorr = E − EHF ) generally require multi-determinant wave functions13,

rendering them much more expensive to calculate.

In contemporary quantum chemistry, HF tends to constitute a starting point for further ab

initio calculations, be they so-called post-HF methods (coupled-cluster, configuration interaction)

or density functional theory. Before examining these methods, it is worth first exploring the

typical basis functions and basis-sets used in modern quantum chemistry.

1.4.4 Basis functions and basis sets

As outlined in the previous section, MOs in quantum chemistry are made up of basis functions.

These basis functions could take many (arbitrary) forms, but in quantum chemistry, because

we are typically interested in molecules – and atoms within them – work is predominantly

13However, DFT also handles dynamic correlation with a single determinant

19

CHAPTER 1. BACKGROUND

performed within the LCAO formalism, where so-called atomic basis functions are used.14

Generally atomic basis functions are broken into angular and radial components, (stemming

from the exact solution to the hydrogen atomic orbitals):

χ(r) = R(r)Y ml (r) (1.17)

where R(r) is the radial component and Y ml is the angular term from the corresponding spherical

harmonic function.

Slater-type orbitals (STOs) were the one of the earliest radial components for LCAO basis

sets [69]. STOs have the following radial component:

R(r) = Nrle−αr (1.18)

where N is some normalization constant typically chosen such that∞∫0

drr2|R(r)|2 = 1. Though

it may not be immediately apparent, the biggest issue with STOs is the difficulty in calculation

of the (many) two electron integrals which must be solved in order to calculate Vee.

By far the most common kinds of basis function used at present are Gaussian-type orbitals

(GTOs), which have the following radial component:

R(r) = Nrle−αr2

(1.19)

Despite having both incorrect short-range (or cusp) and long range behaviour, GTOs allow

rapid calculation of the two-electron integrals due to the Gaussian product theorem (reducing

four centre terms to two centre terms).

More GTOs must be used in order to achieve appropriate accuracy, and there is some history

behind their usage over STOs15 but it is clear that their use did not become favoured over STOs

prior to the involvement of computers in SCF calculations. Indeed, the biggest advantage of

GTOs is the ease of evaluating the integrals exactly in in computer SCF calculations.

Basis sets in the LCAO formalism can by classified by the number and type of basis functions

they include. In perhaps the majority of chemical situations(i.e. molecules, bonding situations),

it is the valence electrons that are principally responsible and of interest. As a result, it is

common to represent valence orbitals by multiple basis functions (particularly for GTOs) so there

are more degrees of freedom to describe the electronic behaviour. Such ‘split-valence‘ basis sets

with two or more basis functions describing valence orbitals are commonly classified as double-,

triple-, quadruple-zeta basis sets and so on [76]. Basis functions may be used to describe the

14In solid state or periodic systems, plane wave basis sets are commonly used. Further, in some multi-resolutionmethods e.g. Yanai et al. [70] wavelets are utilised.

15It is fair to credit S.F. Boys [71] with the introduction of GTOs into quantum chemistry, but theircontemporary dominance is largely due to work done by R.F Stewart (a pioneer of crystallographic chargedensity studies), W.J. Hehre and J.A. Pople [72–75]

20

CHAPTER 1. BACKGROUND

example, the first d-function provides a significant lowering in energy, but the contribution from

a second d-function is much more similar to that of the first f-function. The energy lowering

from a third d-function in turn is similar in magnitude to that from the second f-function and

the first g-function. Therefore, the inclusion of addition of polarization functions is performed

in the following order: 1d, 2d1f, 3d2f1g etc.

An additional aspect of the cc basis sets is that the energy error from the sp-basis should be

similar to (and not exceed) the correlation error arising from the incomplete polarization space,

and the sp-basis therefore also increases as the polarization space is extended. Typically, these

basis sets are denoted as cc-pVnZ, where p indicates incorporation of polarisation functions,

and nZ indicates the degree of split valency (see Table 1.1 for the composition of contracted

and primitive gaussians). When these basis sets are augmented with diffuse functions, they are

typically referred to as aug-cc-pVnZ.

Basis Hydrogen 1st row elements 2nd row elements

Contracted Primitive Contracted Primitive Contracted Primitive

cc-pVDZ 2s1p 4s 3s2p1d 9s4p 4s3p2d 12s8p

cc-pVTZ 3s2p1d 5s 4s3p2d1f 10s5p 5s4p3d1f 15s9p

cc-pVQZ 4s3p2d1f 6s 5s4p3d2f1g 12s6p 6s5p4d2f1g 16s11p

cc-pV5Z 5s4p3d2f1g 8s 6s5p4d3f2g1h 14s8p 7s6p5d3f2g1h 20s12p

cc-pV6Z 6s5p4d3f2g1h 10s 7s6p5d4f3g2h1i 16s10p 8s7p6d4f3g2h1i 21s14p

Table 1.1: Composition of primitive and contracted Gaussian basis functions for the correlation-consistent (cc) basis sets (adapted from Jensen [83])

The complete basis set limit

When introducing split-valence basis sets, the indexing scheme (double-, triple-, quadruple-

zeta etc.) implies a natural progression of larger basis sets, indeed the primary advantage

of these basis-sets is the capacity to generate a set of basis-sets which converge toward the

complete basis-set (CBS) limit systematically. Practical calculation of the CBS limit, then,

is only sensible when using groups of basis-sets designed specifically for the purpose e.g. the

correlation-consistent (cc-pvXZ) or polarization-consistent basis sets (pc-n) [84].

E(L) = E∞ + Ae−BL (1.20)

Given the systematic convergence of the cc basis sets several different schemes have been

proposed for extrapolation to the CBS limit, generally utilising the highest angular momentum

L included in the basis set as the extrapolation parameter. At the HF (or density functional

theory) level, the convergence is expected to be exponential, and functions of the form in

22

CHAPTER 1. BACKGROUND

H0 = T + Vne + V HF (1.21)

H = H0 + λV (1.22)

Here, H0 is termed known as the unperturbed Hamiltonian and V is termed the perturbation.

In RSPT, the perturbed wavefunction and perturbed energy are expressed as a power series in

λ:

Ψ = limm→∞

m∑i=0

λiΨ(i) (1.23)

E = limm→∞

m∑i=0

λiE(i) (1.24)

Møller-Plesset (MP) perturbation theory [87] takes H0 to be the sum of the one-electron Fock

operators, and treats electron correlation as the perturbation to the zeroth-order Hamiltonian.

MP perturbation theory is by far the most commonly used variant of perturbation theories

in the field of quantum chemistry. One of the most important features of MPn16 is its size-

extensivity: the predicted energy for every order of perturbation in MPn scales with the number

of non-interacting particles in a given system.

Second order MP perturbation theory, or MP2 is by far the most commonly used variant of

PT. Higher order perturbation expansions not only become significantly more computationally

expensive, but due to the convergence of MP-PT, do not necessarily improve the resulting

energy of the system. Further, the only new information required to obtain the MP2 energy is

the first order wavefunction.

The MP2 correction may be expressed as follows:

EMP2 =1

4

∑i,j,a,b

〈ψiψj| 1r12|ψaψb〉〈ψiψj| 1

r12|ψaψb〉

εi + εj − εa − εb(1.25)

where ψi and ψj are occupied orbitals and ψa and ψb are virtual orbitals, and ε are their

respective orbital energies.

Systematic studies of MP perturbation theory have shown that it is not necessarily a

convergent theory at high orders. Convergence can be highly erratic or even non-existent,

depending on the precise chemical system or basis set [88].

Where HF would normally scale as O(N4) (where N is the number of basis functions), full

16Orders of MP perturbation theory are generally indexed from two onward e.g. MP2, MP3 etc. indicating towhich order the power series is truncated.

24

1.4. QUANTUM AND COMPUTATIONAL CHEMISTRY

CI will scale as O(NN !), rendering it computationally intractable for any remotely large basis

set.

One of the rigorous, and arguably the most reliable and computationally affordable, methods

for estimating the electronic correlation energy is the coupled cluster (CC) method. Since its

introduction in the 1950s by Coester and Kummel, CC has been developed to correct a reference

wavefunction to a given order of perturbation.

Begin with an excitation operator of the following form:

T = T1 + T2 + T3 + · · ·+ Tn (1.26)

where n is the total number of electrons. The various operators Ti acting on a reference HF

wavefunction Φ0 generate all possible i-th excited Slater determinants. As an example:

T1Φ0 =∑i∈occ

∑a∈virt

caiΦai (1.27)

where c is the expansion coefficient of amplitude. A CI wavefunction may be obtained from this

excitation operator on a HF wavefunction:

ΨCI = (1 + T )Φ0 (1.28)

The corresponding CC wavefunction may be obtained by:

ΨCC = eTΦ0

eT =∞∑k=0

1

k!T k

(1.29)

This exponential operator may be rewritten as:

eT = 1 + T1 + (T2 +1

2T 2

1 ) + (T3 + T2T1 +1

6T 3

1 )

+(T4 + T3T1 +1

2T 2

2 +1

2T2T

21 +

1

24T 4

1 ) + · · ·(1.30)

The first term (1) generates the HF wavefunction and the second term (T1) generates all single

excitations. The first parenthesised group (T2 + 12T 2

1 ) generates all double excitations and so on.

With a CC wavefunction, the Schrodinger equation becomes:

HeTΦ0 = EeTΦ0 (1.31)

If all possible cluster operators Tn are included then all possible excited determinants will be

generated and the CC wavefunction is equivalent to full CI. In general, these will be truncated at

25

CHAPTER 1. BACKGROUND

some excitation level (e.g. CCSD truncates after the double excitation level). Clearly, this then

results in an approximate wavefunction. The truncated exponential form used in CC theory is,

however, generally preferable to a CI truncation. This is because, if an exponential form is used,

then when a system is broken into subsystems, say by pulling those parts apart, the energy is

additive, a rather fundamental property known as size extensivity. This additivity is due to

the property EA+B = eAeB of exponentials, so the wavefunction conveniently separates into a

product of subsystem wavefunctions. Remarkably, CI (and other multireference wavefunctions)

often do not have this property, which can lead to sizable errors for large systems, where the

different parts are effectively isolated from one another.

It is worth mentioning the approximate perturbative triples coupling term that is frequently

utilised as part of the CCSD(T) method. Here the T3 contribution is calculated using the MP4

formula with converged CCSD amplitudes, and one additional fifth-order term describing the

coupling between singles and triples. In fact CCSD(T) is generally a better approximation than

CCSDT due to error cancellation of opposite signs (ignored quadruple excitations, along with

overestimation of the triples correction) [89].

1.4.6 Explicitly correlated methods

Recently, explicitly correlated CC methods have achieved widespread usage – largely due to

their rapid basis-set convergence (relative to ‘vanilla’ CC). There are a wide variety of explicitly

correlated methods, but they can be broadly separated into two categories – those that use

explicitly correlated two-electron basis functions (also referred to as geminals), and those that

only utilise a one-electron basis. For detailed reviews of these methods, see Kong et al. [90],

Ten-no & Noga [91] and Hattig et al. [92]. Broadly, the CCSD-R12 formalism is given by the

following equations:

T = T1 + T2 + R (1.32)

R =1

2!3tyxyR

xyαβa

αβij (1.33)

The geminal operator, R, is the basis for the variety of approximations and formalisms in

CCSD-R12. One such family of approximations are the CCSD(T)-F12 [93] methods, which in

Chapter 6 are utilised via the CCSD(T)-F12b method [94], providing significant improvements

over CCSD(T) for smaller basis-set sizes.

1.4.7 Density functional theory

It would be remiss to provide an overview of modern quantum chemistry without at least touching

on density functional theory (DFT). To that end, it is worth examining some fundamentals

26

1.4. QUANTUM AND COMPUTATIONAL CHEMISTRY

of DFT, on the way providing some idea as to how and why there exists the current ‘zoo’ of

density functionals.

Just as a mathematical function f(x) = x2 might take as input a number and yield a resulting

value (e.g. f(3) = 9), a functional takes as its input a function and results in a value. DFT,

then, is focused on functionals of the electron density. For example, the following functional:

N[ρ]

=

∫ρ(r)dr (1.34)

takes the density as its argument and returns the number of electrons in a given system.

DFT in computational chemistry is primarily concerned with functionals which yield the energy

of a given system.

DFT is based primarily on the Hohenberg-Kohn theorems [95]. However, Levy’s constrained

search theorems [96] are simpler to derive and do not suffer from some technical problems in

Hohenberg-Kohn derivation. Levy says that the ground state wavefunction is a functional of

the electron density, specifically:

E [ρ] = Q [ρ] +

∫ven(r)ρ(r) (1.35)

Q [ρ] = minΨ→ρ〈Ψ|T + Vee|Ψ〉 (1.36)

Q [ρ] searches all antisymmetric wavefunctions which yield a fixed trial electron density ρ,

delivering the expectation value which is a minimum.

Almost all DFT calculations rely on the Kohn-Sham [97] (KS) approximation, which avoids

the need for a kinetic energy functional – as Te is very difficult to approximate, since it is a

direct functional of the Kohn-Sham orbitals rather than the electron density.

Under the KS formalism the electronic energy is obtained from a system of non-interacting

electrons, where the exact energy is given by:

Eee [ρ] =N∑i

εKSi = −1

2J [ρ] + EXC [ρ]−

∫vXC(r)ρ(r)dr (1.37)

EXC [ρ] = Vee [ρ]− J [ρ] + T [ρ]− Ts [ρ] (1.38)

where T is the exact kinetic energy functional, J is the classical coulomb energy, Vee is the

exact electron repulsion energy, Ts is the kinetic energy of a non-interacting system of electrons

described exactly by a single determinant, but having an electron density of ρ. Normally, EXC

and vXC (the potential) are approximated by the so-called density functionals we are discussing

(usually separated into Ecorrelation and Eexchange components). The difference between T and Ts

is ignored [98].

27

CHAPTER 1. BACKGROUND

Functional utilises Kind Examples

ρ LDA SVWN [95, 99]

ρ, ∇ρ GGA BLYP [100, 101], PBE [102], N12 [103]

ρ, ∇ρ, ∇2ρ meta-GGA TPSS [104], BB95 [105], HCTH [106]

ρ, ∇ρ, EHFx hybrid-GGA B3PW91 [107], B3LYP [108], PBE0 [109]

ρ, ∇ρ, ∇2ρ, EHFx hybrid-meta-GGA B1B95[105], TPSSh [110], M06 [111]

Table 1.2: Terms for the different kinds of DFT functionals, including local density approxi-mations (LDA), and generalized gradient approximations (GGA). The first column indicateswhether the functionals in this class utilise the density, ρ, its derivatives and/or incorporate HFexchange (EHF

x ).

Different DFT methods differ in the way they represent Eexchange and Ecorrelation. Typically,

these are functionals of the density ρ, and its derivatives (see Table 1.2 for a broad overview).

Despite the apparent simplicity of the differences between DFT and HF, actual computation

of DFT involves integrals which cannot be solved in closed form, requiring numerical quadratures.

It was largely due to Becke [112] and his work based on Lebedev grids [113, 114] that routine

DFT calculations with appropriate accuracy for molecules became possible.

Hybrid density functionals

DFT typically makes a local (LDA) or semilocal (GGA etc.) approximation for the exchange-

correlation energy functional Exc = Ecorrelation[ρ]

+ Eexchange[ρ]

of the electron density, despite

providing MOs from which the Fock or ‘exact’ exchange energy may be calculated. So called

‘hybrid’ density functionals are those that incorporate some amount of this exact exchange,

which has been demonstrated to provide improvement over non-empirical functionals when

calculating atomization energies, bond lengths, and vibrational frequencies for most molecules

[115, 116].

The original scheme developed by Becke [107] for the first hybrids e.g. B3PW91 is as follows:

Exc = (1− a)Elocalx + aEHF

x + b∆Enon-localx + cElocal

c + (1− c)Enon-localc (1.39)

Of all modern density functionals, the most widely used hybrid functional is almost certainly

B3LYP [108], which uses the same scheme in eq. 1.39 incorporating the Becke88 [100] exchange

functional, and the LYP [101] and VWN3 [99] correlation functionals in conjunction with the

HF exchange.

The scheme in Equation 1.39 is for a hybrid GGA functionals, but in a similar manner hybrid

meta-GGAs can be defined, typically with some fixed portion of the exact exchange. Additionally,

28

1.5. CHEMOINFORMATICS

there are popular range-separated (RS) functionals which vary this portion based on electron-

electron separation. Finally, in the last decade so-called ‘double-hybrids’, which incorporate

electron correlation from MP2, have been developed – producing even more accurate results

than hybrid density functionals or MP217 (approaching ‘chemical accuracy’ of approximately 4

kJ mol−1) over a wide variety of applications [117].

1.4.8 Born-Oppenheimer molecular dynamics

Chapter 5 of this thesis examines a method to predict electron impact mass spectra, based on

Born-Oppenheimer molecular dynamics. As such, a brief overview of molecular dynamics is

warranted here. Molecular dynamics (MD) is the procedure by which molecules are modelled by

empirical force fields (typically Hooke’s-law relations, Lennard-Jones relations, or charged-pair

Coulomb relations) between pairs, triples and quartets of atoms, and then these molecules

are subject to Newton’s Laws of motion constraining either volume or pressure. It is widely

employed in chemistry, biochemistry and materials science, and has been a common area of

study for chemists and biochemists. Popular visualisation and simulation packages like VMD

[118] provide sophisticated tools to investigate the formation of molecular assemblies, among

other things.

In ab initio molecular dynamics the interatomic forces FI = −∇RIψ are determined on-the-fly

using first principles electronic structure methods. That means that AIMD is not relying on

any adjustable parameter, but only on R, However, finding the corresponding many-body

Hamiltonian at each MD step comes at a significant computational cost, which has to be

carefully balanced against the size and sampling requirements of MD. In Born-Oppenheimer

molecular dynamics (BOMD) then, the potential energy E [ψi; R] is minimized at every MD

step with respect to ψi.

1.5 Chemoinformatics

The field of chemoinformatics arguably began with the problem of trying to search for chemical

substructures for use in literature and other databases, sixty years ago [119], and shortly

afterwards, with the attempt to correlate biological acitivity with partition coefficients and

other structural information [120]. Since then, the infiltration of computers into chemistry has

lead to an explosion of applications. Chemoinformatics is critical in the pharmaceutical and

agribusiness sectors for the design of drugs, herbicides and pesticides. In the following sections

the concept of chemical information is elaborated, focussing on the manner in which chemical

information is represented.

17It is worth noting that this success is not without sacrifice: these methods no longer adhere to the Hohenberg-Kohn theorem i.e. the energy is no longer computed as a unique functional of the density only. This means thatenergy is the only property recoverable by these methods.

29

CHAPTER 1. BACKGROUND

1.5.1 Molecules, their properties, and their representation

By far the most important abstraction in chemistry is that of the molecule (or the ion in charged

cases). But a molecule need not be described exclusively in terms of its quantum mechanical

representations: there are as many ways to describe molecules as there are molecular properties.

Which description is useful or valuable in a given context will obviously change depending on the

circumstances, and the questions of interest to the scientist. We might describe a molecule by the

coordinates of its nuclei in their equilibrium geometry, or by their relative positions and angles,

or by its connectivity according to chemical bonding. We might describe molecular properties

like shape (including symmetry, isotropy, size, promolecular density isosurface, or some other

isosurface), its electrostatics (charges, dipoles, quadrupoles etc.) or molecular connectivity

(bonding) or even magnetic or spin properties of the molecule. Although it hardly needs to be

said, the precise representation of this molecular description is referred to as the “descriptor”.

The chemical importance of the descriptor is clearly the first consideration when describing

molecules or their properties, and this will depend entirely on the chemical question we wish

to answer. It should be clear, when exploring chemical questions, that the foremost priority

is the representation of chemical structure (be it two- or three-dimensional) including atomic

composition and any possible bonding information. This is reflected in the typical requirements

for chemical publications, with standardised naming schemes to describe chemical structure

(IUPAC systematic naming), and standard two-dimensional drawing styles e.g. aromatic rings,

single-, double- and triple-bonds etc.

Representations of chemical structure and reactions

There is a plethora of representations of chemical structure used in modern chemistry (see Figure

1.7 for a visual representation of some of them). The most commonly used and understood

representation of chemical structure is still the 2D structure diagram or skeletal formula. Indeed,

this representation is so embedded in chemistry that it is one of the first things taught in

high-school classes, and the required representation for chemical structures in some journals.

This represenation embeds a huge amount of information (and assumptions) about chemical

structure, but many may not even consider its influence on our ideas of chemical bonding and

molecular structure. Consider how benzene rings are drawn, that carbon atoms are generally

not explicitly drawn or that single-, double- and triple-bonds are the canonical representations

of bonding. These diagrams represent a large class of organic chemistry, but in reality chemistry

occurs in three dimensions, and naturally some structures will be more difficult to represent in

2D than others (e.g. fullerenes).

Along with the solution of the 3D structure of molecules from crystallography and other

characterisation methods, 3D representations and depictions of molecules became more common

and more useful. 3D ball-and-stick representations – stemming from the physical models still

30

CHAPTER 1. BACKGROUND

1.5.2 Data types in chemistry

Chemistry deals with data and information in a wide variety of forms, many of them very complex.

The representation of numerical data, for example is generally straightforward (representing

numbers is, after all, widely standardised and accepted). How best to represent chemical

structure (2D or 3D) and spectral data (UV/VIS, NMR, IR, MS or otherwise) is far more

difficult. Yet, the representation of this kind of data is of the utmost importance in the nature

and scale of chemical problems we can solve when using this data. There is a vast number of

different chemical structure formats for electronic storage, predominantly based on connection

tables.

In the field of crystallography there is a standard file format for crystallographic data, the

crystallographic information file (CIF) [122–124], something that has been a great boon for

productivity in data dissemination and analysis. It is a shame, then, that there has been no such

central standardisation of file formats for wavefunctions, basis sets etc. in quantum chemistry

or indeed cheminformatics, as this would radically simplify a great many of the inconveniences

for users and developers in this area.

The proliferation of file formats, some proprietary and others open standards, is the raison

d’etre of software projects like Open Babel [125]. Indeed, that Open Babel can convert between

over 110 chemical file formats (the majority of which are chemical structure formats), speaks

volumes about how urgently standardisation is needed in this space – such redundancy is not

only a waste of productive time in science, every change in format increases the likelihood of

the introduction of errors into studies.

Numerical data

In addition to the storage of chemical structure information, representation of molecular

properties may involve numerical data of high dimensionality (i.e. greater than two). While

representation of numerical data is for the most part not particularly complicated, the transfer and

portability of this data between programs and programming languages can be cumbersome and

error prone. For this reason, primarily for wavefunction and HS data (along with accompanying

properties on the surface) the Simple Binary Format (SBF) described in Appendix F was

developed allowing the straightforward transfer of binary numerical data between the C, C++,

Fortran and Python programming languages.

1.5.3 Machine learning

Further to general informatics techniques, we might use machine learning (ML) in order to

find new molecular descriptors or properties (or refine old ones). Broadly speaking, ML is the

field of enabling computers to ‘learn’ patterns or relationships in data without being explicitly

programmed. ML has become increasingly popular in computer vision, natural language

32

1.5. CHEMOINFORMATICS

processing and indeed in scientific questions that do not lend themselves to straightforward

analytical descriptions, but where large amounts of (reliable) data are available. ML may be

conceived as a stepping stone or component of artificial intelligence, but it does not have as a

primary goal the creation of an intelligent system or being. The goal of ML in general is to

solve problems of a practical nature, usually to make testable predictions.

Because ML encompasses a huge number of disparate problems and approaches, it is well

and truly outside the scope of this introduction to describe them in any form that would do

them justice. Instead this section will focus on the concepts of classification and regression, and

further describe some of the ML techniques which are used in this thesis.

Classification and Regression

Classification problems result in class membership, they take a query object descriptor (or

descriptors) and return the resulting class to which the object is said to belong, based on this

descriptor. Multiple classification approaches are generally built up from binary classifiers,

where membership of one class is all that is interrogated. Classes may be known prior, for

example if we were training a model to determine whether an image contains a cat by providing

a series of images appropriately labelled. Classes may also be ‘discovered’, or unknown, for

example in cluster analysis.

Regression problems, on the other hand, result in a property: they take a query descriptor (or

descriptors) and return some resulting property. Straightforward regression might be something

like linear interpolation of functions, but generally ML concerns some more complicated function

or property. An example of a regression technique would be prediction of a mass spectrum from

a given chemical structure (see discussion of the CFM[126] techniques in Chapter 5).

Cluster analysis

Broadly speaking, cluster analysis is the task of classifying a set of observations in such a manner

that those in the same group or cluster are, in some sense, more similar to one-another than to

those outside the cluster. The clustering performed in this project will be hierarchical in nature.

That is to say that there will be levels or thresholds of proximity, and if we take a cut off at one

of these levels we are left with a group of flat clusters.

There are many different types of cluster analysis, and indeed clustering techniques themselves

(some of which, including methods for implementation are contained in [127]), and applications

of the techniques are as wide ranging as Google’s page-rank algorithms on the world-wide-web

to the field of bioinformatics where the similarity of genetic data is clustered to examine and

infer population structures or assign genotypes.

Clustering variants include (but are not limited to):

• Hierarchical clustering or connectivity based clustering, which is based on the idea that

33

CHAPTER 1. BACKGROUND

nearer objects are more related than those far away. This includes single linkage, complete

linkage, average linkage clustering etc.

• Centroid based clustering in which clusters are represented by a central vector. When the

number of clusters is fixed, k-means clustering provides a formalisation as an optimization

problem (where squared distance from the centre of each cluster is minimised.)

• Distribution based clustering and density based clustering.

Nearest neighbour search

A related problem to cluster analysis is that of nearest-neighbour search (NNS). NNS is a form

of proximity search, which is to say it is an optimization problem of finding the point in a given

set that is closest (most similar) to another given point. Typically, closeness is expressed in

terms of a distance function: less similar objects are ‘further’ from each other. Formally, the

NNS problem is defined as follows: given a set S of points in a space M and a query point

q ∈ M, find the closest point in S to q. An immediate generalization of this problem is to

extend it to the k nearest points, known as k-nearest neighbours (k-NN).

The use case for NNS and kNN is clear when we have a molecular descriptor represented as

a vector or numerical values. Here we can use Euclidean distance d(x) =√∑

x2i (or square

distance function d(x) =∑

x2i ) to find the closest matches to a given query molecule. NNS use

is also widespread in ML algorithms, particularly in the k-NN classification or regression.

Principal Component Analysis

Principal component analysis (PCA) is a procedure that utilises an orthogonal transformation to

convert a set of observations (vectors) in a (possibly) correlated vector space into a set of vectors

in a space of linearly uncorrelated variables termed principal components. More intuitively, it

can be thought of as fitting an n-dimensional ellipsoid to a set of data, such that the axes of the

ellipsoid maximally account for the variance within the set of observations.

PCA can be used for dimensionality reduction, as it identifies axes (which are linear

combinations of the original axes) which maximally account for the variance within the data.

This proves immensely useful for visualising and processing high-dimensional datasets (and it is

used for visualisation in Chapter 3), while still retaining as much of the variance in the dataset

as possible. It is frequently used with only the first two principal components to project a higher

dimensional data set onto the 2D plane, as the resulting projection will be optimally ‘spread

out’, enabling easier visualisation and representation of the differences in the data. PCA can

also be used to align objects or meshes in 3D, by determining the orientation of the primary

axes.

34

1.5. CHEMOINFORMATICS

1.5.4 Chemical databases

Perhaps the most visible achievement of cheminformatics in the last 50 years has been databases

of chemical information [128]. Access to high quality data is of crucial importance to modeling,

as any inaccuracies or errors will be propagated throughout the model. There is, at present, a

wide variety of chemical databases, many with enormous amounts of curated information. There

are general purpose chemical databases (e.g. Pubchem [129], ChemSpider [130]), which provide

free access to fundamental information about chemical structure, computed or experimentally

recorded properties like biological activity, and more. While these databases have found

widespread use in pharmaceutical research and drug discovery, their applicability is far more

widespread. Further, there are many more specific databases, some for mass spectrometry data

(e.g. NIST17 [131], The Wiley/NBS Registry [132]) others for crystallographic data, (e.g. the

aforementioned CSD [133] and PDB [134]).

This widespread availabilty of comprehensive chemical information provides a vital starting

point for chemical modeling: both for helping to parameterise models and for investigating their

success. Without a doubt, the success of cheminformatics (and indeed computational chemistry)

relies on the availabilty, accessibility and veracity of chemical data in databases like these.

Additionally, there are databases consisting entirely of virtual molecules i.e. those which

have been generated by exploration of the ‘chemical space’ but not synthesized. Such databases

find particular use in the field of drug discovery.

1.5.5 Virtual screening

Computers play many roles in drug discovery [135], but a particularly important role is so-called

Virtual screening (VS). The aim of a VS process is to identify molecules that may bind to a

macromolecular target of interest, usually binding to a drug target protein receptor or enzyme.

This is performed via searching or ‘screening’ a large library of potential chemical structures

(e.g. those mentioned in the prior section, ZINC [136] and many more), consisting or real or

virtual chemical structures. VS methods can be broadly separated into two distinct approaches:

methods focussed on matching ligands which are known to bind to the target – ligand-based

virtual screening (LBVS) – and methods focussed on matching potential ligands to the binding

site of the target – structure-based virtual screening (SBVS) [137, 138].

Ligand based virtual screening

Most LBVS methods attempt to design or discover new molecules that satisfy certain similarity

criteria with a known binding ligand [139, 140]. Such matching may be performed based on:

• Intermolecular interaction sites (such as hydrogen-bond donors and acceptors) [136],

35

CHAPTER 1. BACKGROUND

• So-called ‘pharmacophore‘ models built from knowledge of a set of ligands which bind to

a given receptor [141, 142],

• 2D chemical similarity analysis (i.e. skeletal or scaffold similarity) [143, 144],

• 3D molecular shape [145, 146] and/or electrostatics [147–150],

Some LBVS procedures make use of machine learning techniques to assess similarity [151], be it

using one or more of the above methods or otherwise. Because of its dependence only on a small

molecule ligand structure, typical LBVS processes are relatively cheap and efficient. Chapter 4

of this thesis demonstrates the potential use of the spherical harmonic shape descriptors in this

process, for searching small molecule crystal structure databases.

Structure based virtual screening

SBVS typically involves ‘docking’ procedures [152], which utilise a scoring function to estimate

binding or non-binding interactions between a candidate ligand and the desired protein-receptor

site. Docking procedures usually involve a step to find the most favourable orientation in the

receptor pocket, and a variety of molecular conformations (or conformers). Optimisation of the

conformers within the space is prohibitatively expensive (especially if the surrounding area of the

protein is not fixed) for large scale studies, and so generally some number of diverse low energy

molecular conformers are generated and treated as rigid. Further, SBVS approaches generally

require significantly more computation than LBVS procedures, due to their incorporation of a

larger chemical environment (the structure of the entire binding site). This limits the scale of

virtual screenings, or requires larger computational resources in order to perform these searches.

There are many approaches to SBVS (see Lionta et al. [153] or Cheng et al. [137] for recent

overviews), but as the topic is largely tangential to this thesis (with the exception of the docking

scores in Chapter 4) it will not be explored further here.

36

Bibliography

1. Calhoun, C. J. The Microcomputer Revolution? Sociol. Methods Res. 9, 397 (1981).

2. Kuhn, T. The Structure of Scientific Revolutions (University of Chicago Press, Chicago,

1962).

3. Hey, T. The Fourth Paradigm: Data-Intensive Scientific Discovery (eds Hey, T., Tolle,

K. M. & Tansley, S.) (Microsoft Research, Seattle, 2009).

4. Vardi, M. Y. Science has only two legs. Commun. ACM 53, 5 (2010).

5. Henry, N. L. Knowledge Management: A New Concern for Public Administration. Pub.

Admin. Rev. 34, 189 (1974).

6. Rowley, J. The wisdom hierarchy: representations of the DIKW hierarchy. J. Inf. Sci. 33,

163 (2007).

7. Bridgman, P. W. The Logic of Modern Physics (Macmillan, New York, 1927).

8. Popper, K. Conjectures and Refutations (Routlege & Kegan Paul, New York, 1962).

9. Hoffmann, R. A theoretician’s view of crystallography. Acta Crystallogr. A 61, c1 (2005).

10. Dirac, P. A. M. Quantum Mechanics of Many-Electron Systems. Proc. Royal Soc. A 123,

714 (1929).

11. Moore, G. E. Cramming more components onto integrated circuits. Electronics 38, 114

(1965).

12. Simonite, T. Moore’s Law is Dead. Now What? https://www.technologyreview.com/

s/601441/moores-law-is-dead-now-what/ (2017).

13. Engel, T. Basic Overview of Chemoinformatics. J. Cheminf. Model. 46, 2267 (2006).

14. Warr, W. A. Balancing the needs of the recruiters and the aims of the educators in Book

of Abstracts, 218th ACS National Meeting (New Orleans, 1999).

15. Mulliken, R. S. Electronic Structures of Polyatomic Molecules and Valence. Phys. Rev.

40, 55 (1932).

16. Mulliken, R. S. Electronic Structures of Polyatomic Molecules and Valence. II. Quantum

Theory of the Double Bond. Phys. Rev. 41, 751 (1932).

37

BIBLIOGRAPHY

17. Mulliken, R. S. Electronic Structures of Polyatomic Molecules and Valence. V. Molecules

RX n. J. Chem. Phys. 1, 492 (1933).

18. Huckel, E. Die freien Radikale der organischen Chemie. Z. Phys. 83, 632 (1933).

19. Hoffmann, R. An Extended Huckel Theory. I. Hydrocarbons. J. Chem. Phys. 39, 1397

(1963).

20. Heilbronner, E. & Straub. Huckel Molecular orbitals (Springer-Verlag, Berlin, 1966).

21. McLean, A. D. & Yoshimine, M. Computation of Molecular Properties and Structure.

IBM J. Res. Dev. 12, 206 (1968).

22. Snyder, L. C. & Basch, H. Molecular wave functions and properties: (ed Basch, H.)

(Wiley-Interscience, New York, 1972).

23. Ferry, G. Dorothy Hodgkin: A Life (Bloomsbury, London, 2014).

24. Huang, Z. & Kim, K.-J. Review of x-ray free-electron laser theory. Phys. Rev. Accel.

Beams 10 (2007).

25. Kang, Y. et al. Crystal structure of rhodopsin bound to arrestin by femtosecond X-ray

laser. Nature 523, 561 (2015).

26. Buchanan, B. G., Lederberg, J. & Feigenbaum, E. The Heuristic DENDRAL program for

explaining empirical data 1971.

27. Bragg, W. H. & Bragg, W. L. The Reflection of X-rays by Crystals. Proc. Royal Soc. A

88, 428 (1913).

28. Lonsdale, K. The structure of the benzene ring. Nature 122, 810 (1928).

29. Bouslimani, A., Sanchez, L. M., Garg, N. & Dorrestein, P. C. Mass spectrometry of

natural products: current, emerging and future technologies. Nat. Prod. Rep. 31, 718

(2014).

30. Dettmer, K., Aronov, P. A. & Hammock, B. D. Mass spectrometry-based metabolomics.

Mass Spec. Rev. 26, 51 (2007).

31. Cranswick, L. M. D. Busting out of crystallography’s Sisyphean prison: from pencil

and paper to structure solving at the press of a button: past, present and future of

crystallographic software development, maintenance and distribution. Acta Crystallogr. A

64, 65 (2008).

32. Bennett, J. M. & Kendrew, J. C. The computation of Fourier synthesis with a digital

electronic calculating machine. Acta Crystallogr. 5, 109 (1952).

33. Ahmed, F. R. & Cruickshank, D. W. J. A refinement of the crystal structure analyses of

oxalic acid dihydrate. Acta Crystallogr. 6, 385 (1953).

38

BIBLIOGRAPHY

34. Kendrew, J. C. et al. A Three-Dimensional Model of the Myoglobin Molecule Obtained

by X-Ray Analysis. Nature 181, 662 (1958).

35. Levinthal, C. Molecular model-bilding by computer. Sci. Amer. 214, 42 (1966).

36. Desiraju, G. R., Vittal, J. J. & Ramanan, A. Crystal engineering (Co-Published with

Indian Institute of Science (IISc), Bangalore, India, 2011).

37. Verheyen, E. et al. Design of zeolite by inverse sigma transformation. Nature Mat. 11,

1059 (2012).

38. Yaghi, O. M. et al. Reticular synthesis and the design of new materials. Nature 423, 705

(2003).

39. Desiraju, G. R. Supramolecular Synthons in Crystal Engineering—A New Organic Syn-

thesis. Angew. Chem. Int. Ed. 34, 2311 (1995).

40. Corey, E. General methods for the construction of complex molecules. Pure Appl. Chem.

14, 19 (1967).

41. Hirshfeld, F. L. & Rzotkiewicz, S. Electrostatic binding in the first-row AH and A2

diatomic molecules. Mol. Phys 27, 1319 (1974).

42. Coppens, P. & Hall, M. B. Electron Distributions and the Chemical Bond (Springer,

Boston, USA, 1982).

43. Moffitt, W. Atoms in Molecules and Crystals. Proc. Royal Soc. A 210, 245–268 (1951).

44. Bader, R. Atoms in Molecules: A Quantum Theory isbn: 9780198558651 (Clarendon

Press, Oxford, 1994).

45. Coppens, P. X-Ray Charge Densities and Chemical Bonding (Oxford University Press,

Oxford, 1997).

46. Cassam-Chenaı, P. & Jayatilaka, D. Some fundamental problems with zero flux partitioning

of electron densities. Theor. Chem. Acc. 104, 213–218 (3 2000).

47. Cassam-Chenaı, P. & Jayatilaka, D. A complement to “Some fundamental problems with

zero flux partitioning of electron densities”. Theor. Chem. Acc. 107, 383–384 (6 2002).

48. Gill, P. M. W. Extraction of Stewart Atoms from Electron Densities. J. Phys. Chem. 100,

15421–15427 (38 1996).

49. Hunter, G. Atoms in molecules from the exact one-electron wave function. Can. J. Chem.

74, 1008–1013 (6 1996).

50. Roby, K. R. Quantum theory of chemical valence concepts. Mol. Phys. 27, 81–104 (1

1974).

39

BIBLIOGRAPHY

51. Gould, M. D., Taylor, C., Wolff, S. K., Chandler, G. S. & Jayatilaka, D. A definition for

the covalent and ionic bond index in a molecule. Theor. Chem. Acc. 119, 275–290 (1–3

2008).

52. Sukumar, N. A Matter of Density (Wiley, New Jersey, 2012).

53. Hirshfeld, F. L. Bonded-atom fragments for describing molecular charge densities. Theor.

Chem. Acc. 44, 129 (1977).

54. Spackman, M. A. & Byrom, P. G. A novel definition of a molecule in a crystal. Chem.

Phys. Lett. 267, 215 (1997).

55. McKinnon, J. J., Spackman, M. A. & Mitchell, A. S. Novel tools for visualizing and

exploring intermolecular interactions in molecular crystals. Acta Crystallogr. B 60, 627

(2004).

56. Spackman, M. A. & Jayatilaka, D. Hirshfeld surface analysis. CrystEngComm 11, 19.

57. Wigner, E. & Seitz, F. On the Constitution of Metallic Sodium. Phys. Rev. 43, 804

(1933).

58. Voronoi, G. Nouvelles applications des parametres continus a la theorie des formes

quadratiques. Premier memoire. Sur quelques proprietes des formes quadratiques positives

parfaites. J. reine angew. Math. 1908, 97 (1908).

59. Blatov, V. A., Shevchenko, A. P. & Proserpio, D. M. Applied Topological Analysis of

Crystal Structures with the Program Package ToposPro. Cryst. Growth Des. 14, 3576

(2014).

60. Spackman, M. A. & McKinnon, J. J. Fingerprinting intermolecular interactions in molec-

ular crystals. Crystengcomm 4, 378 (2002).

61. McKinnon, J. J., Jayatilaka, D. & Spackman, M. A. Towards quantitative analysis of

intermolecular interactions with Hirshfeld surfaces. Chem. Commun. 267, 3814 (2007).

62. Parkin, A. et al. Comparing entire crystal structures: structural genetic fingerprinting.

Crystengcomm 9, 648 (2007).

63. Schrodinger, E. Quantisierung als Eigenwertproblem. Ann. Phys. (Berl). 384, 361 (1926).

64. Schrodinger, E. An Undulatory Theory of the Mechanics of Atoms and Molecules. Phys.

Rev. 28, 1049 (1926).

65. Born, M. & Oppenheimer, R. Zur Quantentheorie der Molekeln. Ann. Phys. (Berl.) 389,

457 (1927).

66. Hartree, D. R. The Wave Mechanics of an Atom with a Non-Coulomb Central Field. Part

I. Theory and Methods. Math. Proc. Camb. Philos. Soc. 24, 89 (1928).

40

BIBLIOGRAPHY

67. Hartree, D. R. The Wave Mechanics of an Atom with a Non-Coulomb Central Field. Part

II. Some Results and Discussion. Math. Proc. Camb. Philos. Soc. 24, 111 (1928).

68. Fock, V. Naherungsmethode zur Losung des quantenmechanischen Mehrkorperproblems.

Z. Phys. 61, 126 (1930).

69. Slater, J. C. Atomic Shielding Constants. Phys. Rev. 36, 57 (1930).

70. Yanai, T., Fann, G. I., Gan, Z., Harrison, R. J. & Beylkin, G. Multiresolution quantum

chemistry in multiwavelet bases: Analytic derivatives for Hartree–Fock and density

functional theory. J. Chem. Phys. 121, 2866 (2004).

71. Boys, S. F. Electronic wave functions - I. A general method of calculation for the stationary

states of any molecular system. Proc. Royal Soc. A 200, 542 (1950).

72. Stewart, R. F. Small Gaussian Expansions of Slater-Type Orbitals. J. Chem. Phys. 52,

431 (1970).

73. Hehre, W. J., Stewart, R. F. & Pople, J. A. Atomic electron populations by molecular

orbital theory. Symp. Faraday Soc. 2, 15 (1968).

74. Hehre, W. J., Stewart, R. F. & Pople, J. A. Self-Consistent Molecular-Orbital Methods. I.

Use of Gaussian Expansions of Slater-Type Atomic Orbitals. J. Chem. Phys. 51, 2657

(1969).

75. Hehre, W. J., Ditchfield, R. & Pople, J. A. Self—Consistent Molecular Orbital Methods.

XII. Further Extensions of Gaussian—Type Basis Sets for Use in Molecular Orbital

Studies of Organic Molecules. J. Chem. Phys. 56, 2257 (1972).

76. Davidson, E. R. & Feller, D. Basis set selection for molecular calculations. Chem. Rev.

86, 681 (1986).

77. Binkley, J. S., Pople, J. A. & Hehre, W. J. Self-consistent molecular orbital methods. 21.

Small split-valence basis sets for first-row elements. J. Am. Chem. Soc. 102, 939 (1980).

78. Krishnan, R., Binkley, J. S., Seeger, R. & Pople, J. A. Self-consistent molecular orbital

methods. XX. A basis set for correlated wave functions. J. Chem. Phys. 72, 650 (1980).

79. Frisch, M. J., Pople, J. A. & Binkley, J. S. Self-consistent molecular orbital methods 25.

Supplementary functions for Gaussian basis sets. J. Chem. Phys. 80, 3265 (1984).

80. Dunning, T. H. Gaussian-basis sets for use in correlated molecular calculations .1. the

atoms boron through neon and hydrogen. J. Chem. Phys. 90, 1007 (1989).

81. Kendall, R. A., Dunning, T. H. & Harrison, R. J. Electron-affinities of the 1st-row atoms

revisited - systematic basis-sets and wave-functions. J. Chem. Phys. 96, 6796 (1992).

82. Dunning, T. H., Peterson, K. A. & Wilson, A. K. Gaussian basis sets for use in correlated

molecular calculations. X. The atoms aluminum through argon revisited. J. Chem. Phys.

114, 9244 (2001).

41

BIBLIOGRAPHY

83. Jensen, F. Introduction to Computational Chemistry 2nd ed. (Wiley, Hoboken, USA,

2007).

84. Jensen, F. Polarization consistent basis sets: Principles. J. Chem. Phys. 115, 9113 (2001).

85. Karton, A. & Martin, J. M. L. Comment on: “Estimating the Hartree–Fock limit from

finite basis set calculations” [Jensen F (2005) Theor Chem Acc 113:267]. Theor. Chem.

Acc. 115, 330 (2006).

86. Szabo, A. & Ostlund, N. S. Modern Quantum Chemistry (Courier Corporation, London,

1989).

87. Møller, C. & Plesset, M. S. Note on an Approximation Treatment for Many-Electron

Systems. Phys. Rev. 46, 618 (1934).

88. Leininger, M. L., Allen, W. D., Schaefer III, H. F. & Sherrill, C. D. Is Møller–Plesset

perturbation theory a convergent ab initiomethod? J. Chem. Phys. 112, 9213 (2000).

89. Helgaker, T., Klopper, W., Koch, H. & Noga, J. Basis-set convergence of correlated

calculations on water. J. Chem. Phys. 106, 9639 (1997).

90. Kong, L., Bischoff, F. A. & Valeev, E. F. Explicitly Correlated R12/F12 Methods for

Electronic Structure. Chem. Rev. 112, 75 (2011).

91. Ten-no, S. & Noga, J. Explicitly correlated electronic structure theory from R12/F12

ansatze. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2, 114 (2012).

92. Hattig, C., Klopper, W., Kohn, A. & Tew, D. P. Explicitly Correlated Electrons in

Molecules. Chem. Rev. 112, 4 (2011).

93. Adler, T. B., Knizia, G. & Werner, H.-J. A simple and efficient CCSD(T)-F12 approxi-

mation. J. Chem. Phys. 127, 221106 (2007).

94. Knizia, G., Adler, T. B. & Werner, H.-J. Simplified CCSD(T)-F12 methods: Theory and

benchmarks. J. Chem. Phys. 130, 054104 (2009).

95. Hohenberg, P. & Kohn, W. Inhomogeneous Electron Gas. Phys. Rev. 136, B864B871

(1964).

96. Levy, M. Universal variational functionals of electron densities, first-order density matrices,

and natural spin-orbitals and solution of the v-representability problem. Proc. Natl. Acad.

Sci. U.S.A 76, 6062 (1979).

97. Kohn, W. & Sham, L. J. Self-Consistent Equations Including Exchange and Correlation

Effects. Phys. Rev. 140, A1133A1138 (1965).

98. Parr, R. G. & Yang, W. Density functional theory of atoms and molecules (Oxford

University Press, Oxford, 1989).

42

BIBLIOGRAPHY

99. Vosko, S. H., Wilk, L. & Nusair, M. Accurate spin-dependent electron liquid correlation

energies for local spin density calculations: a critical analysis. Can. J. Phys. 58, 1200

(1980).

100. Becke, A. D. Density-functional exchange-energy approximation with correct asymptotic

behavior. Phys. Rev. A 38, 3098 (1988).

101. Lee, C., Yang, W. & Parr, R. G. Development of the Colle-Salvetti correlation-energy

formula into a functional of the electron density. Phys. Rev. B 37, 785 (1988).

102. Perdew, J. P., Burke, K. & Ernzerhof, M. Generalized Gradient Approximation Made

Simple. Phys. Rev. Lett. 77, 3865 (1996).

103. Peverati, R. & Truhlar, D. G. Exchange–Correlation Functional with Good Accuracy for

Both Structural and Energetic Properties while Depending Only on the Density and Its

Gradient. J. Chem. Theory Comput. 8, 2310 (2012).

104. Tao, J., Perdew, J. P., Staroverov, V. N. & Scuseria, G. E. Climbing the Density Functional

Ladder: Nonempirical Meta21Generalized Gradient Approximation Designed for Molecules

and Solids. Phys. Rev. Lett. 91, 146401 (2003).

105. Becke, A. D. Density-functional thermochemistry. IV. A new dynamical correlation

functional and implications for exact-exchange mixing. J. Chem. Phys. 104, 1040 (1996).

106. Hamprecht, F. A., Cohen, A. J., Tozer, D. J. & Handy, N. C. Development and assessment

of new exchange-correlation functionals. J. Chem. Phys. 109, 6264 (1998).

107. Becke, A. D. Density-functional thermochemistry. III. The role of exact exchange. J.

Chem. Phys. 98, 5648 (1993).

108. Stephens, P. J., Devlin, F. J., Chabalowski, C. F. & Frisch, M. J. Ab Initio Calculation of

Vibrational Absorption and Circular Dichroism Spectra Using Density Functional Force

Fields. J. Phys. Chem. 98, 11623 (1994).

109. Adamo, C. & Barone, V. Toward reliable density functional methods without adjustable

parameters: The PBE0 model. J. Chem. Phys. 110, 6158 (1999).

110. Staroverov, V. N., Scuseria, G. E., Tao, J. & Perdew, J. P. Comparative assessment of

a new nonempirical density functional: Molecules and hydrogen-bonded complexes. J.

Chem. Phys. 119, 12129 (2003).

111. Zhao, Y. & Truhlar, D. G. The M06 suite of density functionals for main group thermo-

chemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition

elements: two new functionals and systematic testing of four M06-class functionals and

12 other functionals. Theor. Chem. Acc. 120, 215 (2008).

112. Becke, A. D. A multicenter numerical integration scheme for polyatomic molecules. J.

Chem. Phys. 88, 2547 (1988).

43

BIBLIOGRAPHY

113. Lebedev, V. I. Quadratures on the sphere. Zh. Vychisl. Mat. Mat. Fiz. 16, 293 (1976).

114. Lebedev, V. I. Spherical quadrature formulas exact to order-25-order-29. Sib. Math. J.

18, 99 (1977).

115. Perdew, J. P., Ernzerhof, M. & Burke, K. Rationale for mixing exact exchange with

density functional approximations. J. Chem. Phys. 105, 9982 (1996).

116. Goerigk, L. & Grimme, S. A thorough benchmark of density functional methods for

general main group thermochemistry, kinetics, and noncovalent interactions. Phys. Chem.

Chem. Phys. 13, 6670 (2011).

117. Goerigk, L. & Grimme, S. Double-hybrid density functionals. Wiley Interdiscip. Rev.

Comput. Mol. Sci. 4, 576 (2014).

118. Humphrey, W., Dalke, A. & Schulten, K. VMD Visual Molecular Dynamics. J. Mol.

Graph. 14, 33 (1996).

119. Ray, L. C. & Kirsch, R. A. Finding Chemical Records by Digital Computers. Science

126, 814 (1957).

120. Hansch, C., Maloney, P. P., Fujita, T. & Muir, R. M. Correlation of Biological Activity

of Phenoxyacetic Acids with Hammett Substituent Constants and Partition Coefficients.

Nature 194, 178 (1962).

121. Bonchev, D. & Rouvray, D. H. Chemical Topology: Introduction and Fundamentals (CRC

Press, Boca Raton, Florida).

122. Hall, S. R., Allen, F. H. & Brown, I. D. The crystallographic information file (CIF): a

new standard archive file for crystallography. Acta Crystallogr. A 47, 655 (1991).

123. Hall, S. R. et al. in International Tables for Crystallography Volume G: Definition

and exchange of crystallographic data (eds Hall, S. R. & McMahon, B.) 20 (Springer

Netherlands, Dordrecht, 2005).

124. Brown, I. D. & McMahon, B. CIF: the computer language of crystallography. Acta

Crystallogr. B 58, 317 (2002).

125. O’Boyle, N. M. et al. Open Babel: An open chemical toolbox. J. Cheminform 3, 33

(2011).

126. Allen, F., Pon, A., Greiner, R. & Wishart, D. Computational Prediction of Electron

Ionization Mass Spectra to Assist in GC/MS Compound Identification. Anal. Chem. 88,

7689 (2016).

127. Press, W. H., Teukolsky, S. A., Vetterling, W. T. & Flannery, B. P. Numerical Recipes

3rd Edition: The Art of Scientific Computing 3rd ed. (Cambridge University Press, New

York, NY, USA, 2007).

44

BIBLIOGRAPHY

128. Gasteiger, J. Chemoinformatics: Achievements and Challenges, a Personal View. Molecules

21 (2016).

129. Kim, S. et al. PubChem Substance and Compound databases. Nucleic Acids Res. 44,

D1202D1213 (2016).

130. ChemSpider Home Page https://www.chemspider.com (2017).

131. NIST Standard Reference Database 1A v17 https : / / www . nist . gov / srd / nist -

standard-reference-database-1a-v17 (2017).

132. Wiley Registry of Mass Spectral Data, 11th Edition http://wiley.com/WileyCDA/

WileyTitle/productCd-1119171016.html (2017).

133. Groom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. The Cambridge Structural

Database. Acta Crystallogr. B 72, 171 (2016).

134. Berman, H. M. The Protein Data Bank. Nucleic acids research 28, 235 (2000).

135. Jorgensen, W. L. The many roles of computation in drug discovery. Science 303, 1813

(2004).

136. Irwin, J. J. & Shoichet, B. K. ZINC − A Free Database of Commercially Available

Compounds for Virtual Screening. J. Cheminf. Model. 45, 177 (2005).

137. Cheng, T., Li, Q., Zhou, Z., Wang, Y. & Bryant, S. H. Structure-Based Virtual Screening

for Drug Discovery: a Problem-Centric Review. AAPS J. 14, 133 (2012).

138. Di Giovanni, A. L. & C. Virtual Screening Strategies in Drug Discovery: A Critical Review.

Curr. Med. Chem. 20, 2839 (2013).

139. Willett, P. Enhancing the Effectiveness of Ligand-Based Virtual Screening Using Data

Fusion. QSAR Comb. Sci. 25, 1143 (2006).

140. Stahura, F. L. & Bajorath, J. New methodologies for ligand-based virtual screening. Cur.

Pharm. Des. 11, 1189 (2005).

141. Cho, S. J. & Sun, Y. FLAME: A Program to Flexibly Align Molecules. J. Chem. Inf.

Model. 46, 298 (2006).

142. Khedkar, S. A., Malde, A. K., Coutinho, E. C. & Srivastava, S. Pharmacophore modeling

in drug discovery and development: an overview. Med. Chem. 3, 187 (2007).

143. Willett, P., Barnard, J. M. & Downs, G. M. Chemical Similarity Searching. J. Chem. Inf.

Model. 38, 983 (1998).

144. Sheridan, R. P. & Kearsley, S. K. Why do we need so many chemical similarity search

methods? Drug Discov. Today 7, 903 (2002).

145. Bemis, G. W. & Kuntz, I. D. A fast and efficient method for 2D and 3D molecular shape

description. J. Comput. Aided Mol. Des. 6, 607 (1992).

45

BIBLIOGRAPHY

146. Hahn, M. Three-Dimensional Shape-Based Searching of Conformationally Flexible Com-

pounds. J. Chem. Inf. Comput. Sci. 37, 80 (1997).

147. Hodgkin, E. E. & Richards, W. G. Molecular similarity based on electrostatic potential

and electric field. Int. J. Quantum Chem. 32, 105 (1987).

148. Buchbauer, G., Klinsky, A., Weiß-Greiler, P. & Wolschann, P. Ab initio Molecular

Electrostatic Potential Grid Maps for Quantitative Similarity Calculations of Organic

Compounds. Mol. Model. Ann. 6, 425 (2000).

149. Melani, F., Gratteri, P., Adamo, M. & Bonaccini, C. Field Interaction and Geometrical

Overlap: A New Simplex and Experimental Design Based Computational Procedure for

Superposing Small Ligand Molecules. J. Med. Chem. 46, 1359 (2003).

150. Vainio, M. J., Puranen, J. S. & Johnson, M. S. ShaEP: Molecular Overlay Based on

Shape and Electrostatic Potential. J. Chem. Inf. Model. 49, 492 (2009).

151. Hert, J. et al. New Methods for Ligand-Based Virtual Screening: Use of Data Fusion

and Machine Learning to Enhance the Effectiveness of Similarity Searching. J. Chem.

Inf. Model. 46, 462 (2006).

152. Kitchen, D. B., Decornez, H., Furr, J. R. & Bajorath, J. Docking and scoring in virtual

screening for drug discovery: methods and applications. Nat. Rev. Drug Discov. 3, 935

(2004).

153. Lionta, E., Spyrou, G., Vassilatis, D. K. & Cournia, Z. Structure-based virtual screening

for drug discovery: principles, applications and recent advances. Curr. Top. Med. Chem.

14, 1923 (2014).

46

Chapter 2

Introduction to a series of papers

It should be clear from the previous sections that computational chemistry encompasses a

wide variety of chemical, mathematical and software development problems. Throughout the

following series of papers, some these problems are explored in detail. With the exception of

Chapter 4, all of the following chapters have been previously published in peer-reviewed journals.

The content from these publications has been modified only to fit the formatting of this thesis,

no more. The papers comprising this thesis can be broadly separated into two parts: those

primarily focussed on molecular shape in crystals, and those focussed on assessing the accuracy

of quantum chemical calculations for their practical application.

2.1 Molecular shapes in crystals

The first part of this thesis is concerned with the development and application of a novel

description technique for molecular shapes. In these papers shape has been defined in terms of

electronic isosurfaces: the Hirshfeld Surface [292] and the promolecular density isosurface.

Chapter 3 consists of a publication laying out the description of molecular shapes with

spherical harmonic functions, and the process of constructing rotation-invariant shape descriptors

from the coefficients of the spherical harmonic transform. Further, this chapter contains

application to small examples: the classification of atomic Hirshfeld surfaces in metallic crystals

according to their lattice type, and the classification of different molecular scaffolds based on

these shape descriptors.

Chapter 4 represents the first significant application of these shape descriptors: to improve

the process of Ligand Based Virtual Screening (LBVS) in rational drug discovery, and enable

the use of large databases of small-molecule crystals as potential ligands. This paper consists of

the description of four protein-ligand systems in terms of the aforementioned spherical harmonic

shape descriptors, a description which is then utilised to find the top matching molecules

from a selection of over 110 thousand crystal structures from the CSD. Further, this paper

demonstrates the immense potential value of the usage of molecular conformers obtained from

47

CHAPTER 2. INTRODUCTION TO A SERIES OF PAPERS

their small-molecule crystal structures: an energetically favourable environment in which they

are interacting with neighbouring molecules. Finally, the impacts of different surface properties

(dnorm and electrostatic potential) and surface types (Hirshfeld surface and promolecule surface)

are explored with reference to their characteristic description of local molecular environment vs.

encapsulation of the crystal environment.

2.2 Assessment of the accuracy of quantum chemical

methods

The second part of this thesis is concerned with assessment and characterisation of existing

prediction techniques in quantum chemistry.

Chapter 5 primarily examines Grimme and Bauer’s method for the prediction of the electronic

impact mass spectrum of arbitrary molecules: Quantum Chemical Electronic Impact Mass

Spectrum (QCEIMS). This technique is evaluated for a set of 61 small molecules from different

common, and important (in terms of their fragmentation propensity), chemical compound

classes. The performance of this method is examined with reference to experimental spectra,

and compared with another (radically different) state-of-the-art approach in Competitive

Fragmentation Modelling (CFM). This chapter, then, consists of comparison between ab initio

predictions and machine learning predictions for EIMS, with the purpose of establishing a

baseline of correctness for use in identification of natural products and semiochemicals, or

metabolomics as complementary to experimental work.

Chapter 6 examines basis-set convergence of equilibrium geometries using CCSD(T), along

with newer explicitly correlated CCSD(T)-F12 methods. The basis-set convergence of equilibrium

geometries is especially important in high-accuracy thermochemical calculations, and this

publication investigates the relationship between correlation-consistent basis-set size and bond-

length convergence, along with the corresponding energies, providing practical recommendations

for their use. Further, the effect of the corresponding geometries on the total energies of

the systems, along with the relative scale of these effects in terms of other approximations

(relativistic effects, core-valence contributions etc.) present in typical CCSD(T) calculations is

established.

Finally, Chapter 7 explores two different categories of complete basis-set (CBS) extrapolations

for CCSD single point energy calculations from relatively small double-zeta and triple-zeta basis

sets. This is particularly important, as CC calculations with all single and double excitations

(CCSD) converge very slowly with the size of the one-particle basis set, and the scaling of the

computational cost with increasing basis-set size is severe. Further, the potential use of a system-

dependent basis-set extrapolation scheme, in which unique basis-set extrapolation exponents

for each system are obtained from lower-cost MP2 calculations is explored. This establishes a

48

2.2. ASSESSMENT OF THE ACCURACY OF QUANTUM CHEMICALMETHODS

protocol for high-accuracy CCSD calculations with accelerated basis-set convergence relative to

global extrapolation schemes.

49

CHAPTER 2. INTRODUCTION TO A SERIES OF PAPERS

50

Part II

Molecular shapes in crystals

51

Chapter 3

High throughput profiling of molecules

in crystals

ABSTRACT

Molecular shape is important in both crystallisation and supramolecular assembly, yet its role

is not completely understood. We present a computationally efficient scheme to describe and

classify the molecular shapes in crystals. The method involves rotation invariant description of

Hirshfeld surfaces in terms of spherical harmonic functions. Hirshfeld surfaces represent the

boundaries of a molecule in the crystalline environment, and are widely used to visualise and

interpret crystalline interactions. The spherical harmonic description of molecular shapes are

compared and classified by means of principal component analysis and cluster analysis. When

applied to a series of metals, the method results in a clear classification based on their lattice

type. When applied to around 300 crystal structures comprising of series of substituted benzenes,

naphthalenes and phenylbenzamide it shows the capacity to classify structures based on chemical

scaffolds, chemical isosterism, and conformational similarity. The computational efficiency of

the method is demonstrated with an application to over 14 thousand crystal structures. High

throughput screening of molecular shapes and interaction surfaces in the Cambridge Structural

Database (CSD) using this method has direct applications in drug discovery, supramolecular

chemistry and materials design.

3.1 Introduction

Molecular shape plays a fundamental role in our understanding of chemistry and biochemistry,

with the supramolecular assembly of molecular species generally being understood both in terms

of the intermolecular interactions and the complementarity of molecular shapes. In this area,

Lehn’s conception of supramolecular chemistry focused on specific molecular recognition leading

53

CHAPTER 3. HIGH THROUGHPUT PROFILING OF MOLECULES INCRYSTALS

to ‘supramolecules’.[1] Similarly, much of the known molecular recognition processes (such as

drug-receptor binding, enzymatic reactions etc.) can be understood with the ‘lock and key’

paradigm of steric fit proposed by Fischer[2] over a century ago.

Much of our present understanding of crystal structures, from Kitaigorodskii’s principles

of crystal packing [3] – the maximum utilisation of space and minimisation of free energy –

to Desiraju’s approach to crystal engineering[4] – the notion of ’supramolecular synthons’ as

recognition units based on chemical functionality – has emphasised the importance of molecular

shape.

Yet, beyond cartoon depictions of shape complementarity and qualitative arguments the

role of molecular shape is largely sidelined in quantitative analysis of supramolecular chemistry

and crystal engineering. It is our view that this is largely due to the lack of a straightforward

method allowing us to incorporate molecular shape into such studies. It would appear, then, that

computational approaches for describing molecular shapes in an accurate, systematic manner

are highly desirable.

With regard to molecular crystals, it is hard to overstate the scientific value that has been

derived from the magnitude of experimental crystal structure data available, curated within

the Cambridge Structural Database[5]. Such a database represents an opportune target for

the purposes of quantitative analysis. To adequately utilise this growing amount of data,

automation and algorithmic analysis (be it traditional statistical methods or machine learning)

are required. So, for the dual purposes of classification and understanding molecular shapes in

crystal structures, we present an efficient computational method based on spherical harmonics

with the capacity to accurately describe molecular shapes in their crystal environment.

In order to account for molecular shape, some degree of chemical information (i.e. the effects

of different elements), and some aspects of the crystal environment, we utilise the Hirshfeld

surface[6] (HS) as a summary object of molecules in crystals. The HS developed by Spackman

et al. is an isosurface surrounding a molecule in its crystal structure, defined as the boundary

where the contribution of electron density ‘belonging’ to the molecule is equal to that of its

crystalline surroundings. The densities in each case are approximated by a superposition of

quantum mechanically derived spherical atomic densities, a so-called promolecular[7] density.

Much like van der Waals or CPK[8] surfaces, the HS represents a realistic molecular surface,

albeit one generated from the experimental X-ray geometry. Further, it encompasses information

about the packing and intermolecular interactions within a crystal, and includes some molecular

size effects, both of which may be encompassed in its shape.

The HS is often decorated with properties such as di (the distance to the nearest internal

atom), de (the distance to the nearest external atom), and dnorm (the distance between internal

and external atoms normalised by van der Waals radii). Figure 3.1 demonstrates some of the

capacity for these properties to visually represent aspects of the crystal environment, such as

the red spots here indicating close contacts between molecules.

54

3.1. INTRODUCTION

Figure 3.1: Views of (a) crystalline environment in the co-crystal of the anti-inflammatorydrug indomethacin with nicotinamide, and (b) corresponding Hirshfeld dnorm surface aroundindomethacin. Close contacts appear as red regions, while more distant interactions will appearfrom white to blue.

HS properties, namely de and di, have been previously used by Gilmore and co-workers [9,

10] when they proposed so-called ‘genetic fingerprinting’ based on rasterisation of the Hirshfeld

fingerprints[11, 12] (2D histogram representations of de and di from the HS). This constituted a

similar use of the HS as a summary object of molecules in crystals, but the descriptors used

by these workers do not directly express the shape of the HS. As such they do not provide a

straightforward means of incorporating molecular shape in quantitative analysis.

Outside of the domain of molecules in crystals, there are a number of other methods routinely

applied to surfaces (e.g. Van der Waals surfaces) and domains (e.g. binding pockets in proteins)

that represent shapes, molecular or otherwise, using spherical harmonic functions. Rotation

invariant descriptors have been been applied in the field of drug design[13, 14] and more generally

in 3D shape recognition[15, 16], where they are used to perform shape matching without the high

computational cost of ‘docking’. Docking essentially consists of rotating the two shapes prior

to comparison so that they are in maximum coincidence, and it rapidly becomes an extremely

costly step in the comparison procedure as large numbers of structure comparisons are required.

An important aspect of utilising spherical harmonic descriptors is the natural parameter lmax

(i.e. the highest order of spherical harmonic functions used in the transform) which provides

a systematic parameter of the level of detail in the shape description. As a brief example of

of how accurate the description may be for higher values of lmax, Figure 3.2 shows two HS

reconstructions (i.e. meshes generated from the resulting coefficients of the spherical harmonic

transform), one for the benzene crystal and the other for an indomethacin crystal. Both

reconstructions include the dnorm property which colours the surface. In practice we have found

that typically lmax = 9 constitutes an acceptable compromise between precision and brevity in

55

CHAPTER 3. HIGH THROUGHPUT PROFILING OF MOLECULES INCRYSTALS

Figure 3.2: From left to right, reconstructed (lmax = 9 and lmax = 20) and original Hirshfeldsurfaces for BENZEN07 (top) and INDMET (bottom). Surfaces have been coloured based on thednorm property at each vertex. While the reproductions at lmax = 9 are not exact, descriptionsat this level clearly capture the essential idea of the shape of the HS

the description of the HS.

Since the method outlined here also involves rotationally invariant shape descriptors, it

enables computationally efficient shape matching in a very large number of structures (taken here

from the CSD). The full description of the details of the method, along with a brief description

of the HS, is provided in the methods section.

The potential value of an efficient, numerical, rotation invariant description of the HS in a

crystal will be demonstrated here through its application to selected sets of crystal structures.

The first dataset consists of 29 metallic crystal structures; a mix of hexagonal close-packed

(HCP) and cubic close-packed (CCP) crystal lattices. The second set consists of over 300

structures; comprising a series of substituted benzenes, naphthalenes and phenylbenzamides,

along with some pyridine analogues of each kind. The separation and grouping of both these

datasets in a principal component analysis (PCA) and cluster analysis based on the shape

descriptors is discussed with reference to crystal packing, chemical scaffolds, chemical isosterism

and molecular conformation. Finally, the computational efficiency of this technique will be

outlined by examining a dataset comprising over 14,000 organic crystal structures.

56

CHAPTER 3. HIGH THROUGHPUT PROFILING OF MOLECULES INCRYSTALS

Figure 3.4: Hirshfeld surfaces for 3 CCP metals and 3 HCP metals with dnorm mapped on thesurface. Note the distinct patterns in both the shape of the surface and dnorm correspond tothe lattice structure in the crystal environment, along with the heightened tendency towardasphericity as the packing becomes ’tighter’ (closer interatomic distance with regard to electrondensity).

and linearity of CCP metals in the descriptor space may be readily understood in terms of the

symmetry constraints brought by CCP: there is no degree of freedom in the shape of the unit

cell. Thus the only variation in this group must stem from a variation in interatomic distance

and electron density. Indeed, as we traverse the CCP metals along the approximate line from

Fe through to Pd (see Figure 3.4), we may see the increased similarity between the HS and the

space filling Voronoi or Wigner-Seitz cell[17].

The HCP group, on the other hand, exhibits more variation in its HS shape. This variation

may be accounted for by the varying degrees of of anisotropy in the different HCP metals, i.e.

the extra degree of freedom in the c axis of the unit cell. Examining the ratios of unit cell

lengths ca

in the HCP metals, it is evident that those with radically different ratios tend to be

separated, and the apparent outlier Cd can be explained by its notably large ratio (1.89) i.e. its

high degree of anisotropy. Indeed, Cd is the only element with a ratio higher than the ideal of

1.633[18]. The immediate separation of Cd indicates that its unusual degree of anisotropy in the

unit cell is directly associated with the HS shape, and that this shape is adequately described

by this technique.

Given the sharp differentiation seen for Cd, one might expect, then, that structures with

identical unit cell ratio would be co-located in the descriptor space. This is not the case, as while

they may have the same symmetry the different atomic lattices may, just as the CCP metals,

vary in their electron densities and interatomic distances. For example, while Co and Mg share

a close to ideal ca

ratio (both have a value 1.62 vs. the ideal value of 1.633) they differ in their

interatomic distance within the lattice, with Co being rather more tightly packed (interatomic

58

CHAPTER 3. HIGH THROUGHPUT PROFILING OF MOLECULES INCRYSTALS

the CSD with structure refinement R-factor < 0.03, only one molecular residue, the heaviest

permitted element was Xe, and for which the HS was ‘star shaped’ (17,646 were not ‘star

shaped’).

The meaning of the 2D PCA distribution is not itself the target of discussion (though the

lack of clear separation may indicate that two dimensions is not enough to visualise such a

diverse dataset). Rather, we emphasise that clustering this dataset takes less than five seconds

on a single-processor laptop. This demonstrates the potential for the technique in analysing

even larger scale datasets. More details of the speed and efficiency are outlined in the Efficiency

and implementation considerations section.

Even though it is not the main focus of this example, it would be remiss not to discuss

some aspects of the results obtained from this large cluster analysis. First, we observe very

many different small clusters, many of which are nearly identical. These nearly identical clusters

usually correspond to several determinations of the same crystal structure (for any reasonable

technique it should be expected that duplicate crystal structures will be co-located in the

descriptor space). Focusing on one cluster located in the selected region in in Figure 3.8, we

may see a visual similarity in the HS of the selected objects. This displays the potential of this

method to screen for a particular molecular shape or ‘interaction surface’ in the CSD, irrespective

of the similarity in chemical structure or elemental composition of the molecules. In other words,

this method provides a shape based structural comparison tool which is fundamentally distinct

from the conventional chemical connectivity-based CSD search tools[21].

3.3 Future research and prospects

In this paper we have presented the first application of rotationally invariant spherical harmonic

shape descriptors based on Hirshfeld surfaces for analysing the nature of molecular packing in

crystals. The advantage of using the technique is that once the descriptors are defined it can be

applied without bias, automatically and efficiently on potentially large datasets — too large for

direct examination by an individual.

The technique we have developed need not be applied to Hirshfeld surfaces: any surface

which is characteristic of the molecular packing in crystals may be used. As outlined, it may

also be applied to any properties mapped on the HS, such as dnorm (a property which generally

reflects the intermolecular interactions of the molecule).

Further, the capacity to process large-scale datasets provides promise in the fields of drug

discovery and crystal engineering. In addition to the conventional drug discovery techniques that

largely rely on functionalisation and systematic modification of selected chemical scaffolds, a

systematic and quantitative method based on shape affords new possibilities. For example, this

technique could be used to profile the shape of protein receptor pockets, along with a property

mapped onto the surface of the pocket (e.g. dnorm or electrostatic potential), subsequently

62

3.3. FUTURE RESEARCH AND PROSPECTS

Figure 3.8: Hexagonally binned 2D histogram of the first two principal components frominvariants up to lmax = 9 along with the mean radius of all 14,772 structures. Note that thethe region highlighted as red square in the histogram represents closely related structuresin 10-dimensional principal component space, and not necessarily the components in the 2DPCA. The similarity in corresponding molecular shapes can be visualised in the representativestructures contained within this cluster.

63

CHAPTER 3. HIGH THROUGHPUT PROFILING OF MOLECULES INCRYSTALS

searching through the large number of diverse structures in CSD extracting the best-matching

candidates, on the assumption that the HS of a molecule in a crystal or co-crystal constitutes

an acceptable proxy for the receptor pocket. Similarly, in the field of crystal engineering

and supramolecular chemistry, the molecular shape based approach could help in developing

systematic design strategies that utilise more of the chemical information inherent in shape and

interaction surfaces, information that may be difficult to incorporate for a human investigator.

3.4 Methods

3.4.1 Representing Hirshfeld surfaces with spherical harmonics

The Hirshfeld surface has been used primarily as a graphical object for visualisation. It is an

isosurface of a particular function of the type and positions of a subset of the atomic nuclei in

an infinite periodic crystal. It is represented as a set of vertices V = {vi, i = 1, . . . , nv} which

are connected in triangular facets. Typically, V comprises thousands of vertices, making it

prohibitively large for search and comparison algorithms on large datasets.

The first step in representing the HS with spherical harmonics is the determination a

suitable origin. Since the number of vertices is large and evenly distributed, the mean position

v = n−1v

∑Ni=1 vi represents an adequate centre.

Next, the surface must be normalised to have roughly unit radius via the transformation

ui = r−1(vi − v). The mean radius is r = n−1v

∑i |vi − v|. If u is one of the vertices ui, this

defines a function on the unit sphere f(θ, φ) = |u| where the polar angles are defined in the

usual way by

ux = |u| sin θ cosφ, uy = |u| sin θ sinφ, uz = |u| cos θ. (3.1)

f is the normalised HS; it can be defined at points other than the given vertices by interpolation.

Note that we consider only surfaces for which there is a unique normalised vertex for every polar

coordinate. This restricts the surfaces to so-called star-shaped domains, which comprise the

majority of small-molecule HS.

Any function of polar angles such as f may be represented to arbitrary accuracy (which

may be parameterised by lmax) using the spherical harmonic functions, Y ml (θ, φ), as a basis as

follows:

f =lmax∑l=0

m=l∑m=−l

cml Yml (θ, φ) (3.2)

64

3.4. METHODS

The coefficients cml are the spherical harmonic expansion coefficients.

cml = 〈Y ml |f〉 =

2π∫0

π∫0

Y m?l (θ, φ)f(θ, φ) sin θdθdφ (3.3)

These coefficients cml may be more readily computed through the use of a Lebedev quadrature[22,

23],

cml =

NLebedev∑i=0

wiYm?l (θi, φi)f(θi, φi). (3.4)

The quadrature weights and points (wi, θi, φi) are fixed for a given choice of l ≤ lmax. Such

grids are widely used in quantum chemistry, and provide an efficient means to exactly integrate

polynomials or spherical harmonics on the surface of a sphere. If the summation in (3.3) is

restricted to lmax there will be (lmax + 1)2 expansion coefficients.

The expansion procedure described above may also be used to encode other properties which

are recorded on the same set of vertices in V . The procedure is identical except that one uses

the set of property values P = {pi, i = 1, . . . , nv} instead of V to define the normalised HS i.e.

fP (θi, φi) = pi. One then obtains the spherical expansion coefficients for this colouring of the

surface.

3.4.2 Rotation invariant shape descriptors

Because we wish to obtain numerical descriptors independent of the orientation of the HS

it is desirable to process the coefficients of the spherical harmonic transform such that they

are rotation invariant. Weyl[24] has described the general procedure for constructing all such

invariants (see also Biedenharn and Louck[25]). Burel and Henocq[26] have proposed a more

limited set of invariants, and our experience has shown that using only the simplest “N” type

invariants yields acceptable results. These invariants may be evaluated as follows:

Nl =l∑

m=−l

cml [cml ]?. (3.5)

If it is desirable to factor size into the shape analysis, we need only include the mean radius

as an additional invariant by appending it to our feature vector for comparison. As previously

mentioned, These descriptors may also be applied to quantities decorating the HS or indeed any

other scalar function on the HS.

3.4.3 Efficiency and computer implementation considerations

On average, for the data set comprising 14,772 structures, the HS calculation and analysis took

between 1-3s per crystal structure on a typical Intel single processor laptop. The majority of

65

CHAPTER 3. HIGH THROUGHPUT PROFILING OF MOLECULES INCRYSTALS

this calculation time is spent in calculating the triangulated HS. There is potential for great

speed up by not triangulating the Hirshfeld surfaces at all, but by calculating the required

HS points and the quadrature points directly. For lmax = 9 which is more than sufficient for

descriptor purposes there are only 50 grid points; this is 2-3 orders of magnitude smaller than

the number of points needed for high quality graphical display. This has not been pursued as

the software is currently in a proof-of-concept state, and the meshes for the Hirshfeld surfaces

themselves are useful for comparison.

All associated surface data has been stored using Hierarchical Data Format HDF5 [27]. This

has minimal impact on the descriptors themselves (as we have only 10 per surface). Even for 1

million structures the storage of their entire feature vectors up to lmax = 9 would require less

than 100 MB of storage. Thus, the data retrieval aspect of the process requires only a trivial

amount of time.

By contrast, the distance calculations required for cluster analysis necessarily scale as O(N2)

where N is the number of structures considered, since we must calculate distances between each

possible pair of structures. Therefore, the computation of this distance matrix will become the

bottleneck as N gets large. Despite this, even for the large data set (14,772 objects) considered

here, the total computation for HDBSCAN clustering (once the shape descriptors have been

calculated) was less than five seconds on a consumer-grade laptop.

It is the efficiency of the representation of shape here that will allow shape to be incorporated

into further algorithmic analysis of the CSD (or any other crystal structure databases). Such

possibilities will be explored in future publications.

3.5 Acknowledgements

We gratefully acknowledge the financial support of the Danish National Research Foundation

(Center for Materials Crystallography, DNRF-93) to Peter R. Spackman.

3.6 Author contributions statement

P. R. S. implemented all software, analysed results and prepared the manuscript. S. P. T and D.

J. contributed to the design of this study and the preparation and revision of the manuscript.

66

Bibliography

1. Lehn, J.-M. Supramolecular Chemistry: Concepts and Perspectives isbn: 9783527607433.

doi:10.1002/3527607439. http://dx.doi.org/10.1002/3527607439 (Wiley, 2006).

2. Fischer, E. Einfluss der Configuration auf die Wirkung der Enzyme. Ber. Dtsch. Chem.

Ges. 27, 2985. issn: 1099-0682 (1894).

3. Kitaigorodskii, A. I. Principle of close packing and condition of thermodynamic stability

of organic crystals. Acta Crystallogr. 18, 585 (1965).

4. Desiraju, G. R. Supramolecular Synthons in Crystal Engineering—A New Organic Syn-

thesis. Angew. Chem. Int. Ed. 34, 2311 (1995).

5. Allen, F. H. The Cambridge Structural Database: a quarter of a million crystal structures

and rising. Acta Crystallogr. B 58, 380 (2002).

6. McKinnon, J. J., Spackman, M. A. & Mitchell, A. S. Novel tools for visualizing and

exploring intermolecular interactions in molecular crystals. Acta Crystallogr. B 60, 627

(2004).

7. Spackman, M. A. & Maslen, E. N. Chemical-properties from the promolecule. J. Phys.

Chem. 90, 2020 (1986).

8. Corey, R. B. & Pauling, L. Molecular models of amino acids, peptides, and proteins. Rev.

Sci. Instr. 24, 621 (1953).

9. Parkin, A. et al. Comparing entire crystal structures: structural genetic fingerprinting.

Crystengcomm 9, 648 (2007).

10. Collins, A., Wilson, C. C. & Gilmore, C. J. Comparing entire crystal structures using

cluster analysis and fingerprint plots. Crystengcomm 12, 801 (2010).

11. Spackman, M. A. & McKinnon, J. J. Fingerprinting intermolecular interactions in molec-

ular crystals. Crystengcomm 4, 378 (2002).

12. Spackman, M. A. & Jayatilaka, D. Hirshfeld surface analysis. Crystengcomm 11, 19

(2009).

67

BIBLIOGRAPHY

13. Morris, R. J., Najmanovich, R. J., Kahraman, A. & Thornton, J. M. Real spherical

harmonic expansion coefficients as 3D shape descriptors for protein binding pocket and

ligand comparisons. Bioinformatics 21, 2347 (2005).

14. Mak, L., Grandison, S. & Morris, R. J. An extension of spherical harmonics to region-based

rotationally invariant descriptors for molecular shape description and comparison. J. Mol.

Graph. Model. 26, 1035 (2008).

15. Tangelder, J. W. H. & Veltkamp, R. C. A survey of content based 3D shape retrieval

methods. Multimed. Tools Appl. 39, 441. issn: 1573-7721 (2007).

16. Xu, D. & Li, H. Geometric moment invariants. Pattern Recognit. 41, 240 (2008).

17. Spackman, M. A. Molecules in crystals. Phys. Scripta 87, 048103 (2013).

18. Hummel, R. E. Understanding Materials Science 32 (Springer, New York, 2004).

19. Campello, R. J. G. B., Moulavi, D., Zimek, A. & Sander, J. Hierarchical Density Estimates

for Data Clustering, Visualization, and Outlier Detection. ACM Trans. Knowl. Discov.

Data 10, 5:1. issn: 1556-4681 (2015).

20. Thornber, C. W. Isosterism and molecular modification in drug design. Chem. Soc. Rev.

8, 563 (1979).

21. Bruno, I. J. et al. New software for searching the Cambridge Structural Database and

visualizing crystal structures. Acta Crystallogr. B 58, 389 (2002).

22. Lebedev, V. I. Quadratures on the sphere. Zh. Vychisl. Mat. Mat. Fiz. 16, 293 (1976).

23. Lebedev, V. I. Spherical quadrature formulas exact to order-25-order-29. Sib. Math. J.

18, 99 (1977).

24. Weyl, H. The Classical Groups: Their Invariants and Representations 8th ed. (Princeton

University Press, 1973).

25. Biedenharn, L. C. & Louck, J. D. Angular momentum in quantum physics: theory and

application 307 (Addison-Wesley, 1981).

26. Burel, G. & Henocq, H. 3-dimensional invariants and their application to object recognition.

Sig. Proc. 45, 1 (1995).

27. The HDF Group. Hierarchical data format version 5 http://www.hdfgroup.org/HDF5.

68

Chapter 4

Fast screening of crystal structures for

ligands based on molecular shape and

intermolecular interactions

ABSTRACT

An efficient method to represent interaction surfaces of molecules in crystals has led to a new

approach to drug discovery. Interaction surfaces (Hirshfeld and promolecular density isosurfaces)

of molecules in crystal structures, and ligands in receptor-protein pockets have been described

using spherical harmonics and utilized to identify potential ligand molecules with similar shape

and intermolecular interaction propensity. The method has been demonstrated through a series

of small drug molecules and drug-protein crystal structures via a high-throughput profiling of

molecular shapes extracted from around 110,000 small molecule, high accuracy, crystal structures

in the Cambridge Structural Database (CSD). Molecular dynamics simulations of ligand-protein

structures in aqueous environment, followed by generation of Hirshfeld surfaces of ligands and

the electrostatic potentials mapped on these surfaces has resulted in the identification of suitable

molecules from the CSD as potential ligands to bind to the same location. The similarity of

these candidates to drug-like molecules obtained by our screening method has been validated

through docking scores, which were found to be comparable to, or in some cases higher than,

the original ligand.

4.1 Introduction

Identifying suitable ligand molecules that bind to the receptor protein sites is a significant

step in rational drug discovery. Drug-protein supramolecular chemistry may still be broadly

conceived of in terms of the simple lock and key model proposed by Fischer[1] more than a

69

CHAPTER 4. FAST SCREENING OF CRYSTAL STRUCTURES FORLIGANDS BASED ON MOLECULAR SHAPE AND INTERMOLECULARINTERACTIONS

century ago. Thus the ligand molecules that most effectively interact with the receptor – those

that best fit the receptor pockets – are ranked as the most suitable drug candidates in structure

based virtual screening (SBVS) procedures, [2, 3] be it via docking procedures, or ligand based

virtual screening (LBVS). Since the role of molecular shapes (and shape complementarity with

respect to the pockets in binding sites) is crucial in determining the most effective inhibitor, a

number of computer-aided drug design/discovery methods focus on profiling molecular, shapes[4,

5] and/or electrostatics [6–9], identification of so-called ‘pharmacophore’ points[10, 11], and

other mathematical descriptions or fingerprints associated with molecular shape.

While most LBVS methods attempt to design or discover new molecules that satisfy certain

structural criteria for intermolecular interactions (such as hydrogen bond donors and acceptors)

for a given query molecule, few (if any) LBVS approaches make use of the vast number of

experimental, small-molecule crystal structures present in the Cambridge Structural Database[11]

(CSD). The CSD contains an enormous amount of chemical information embedded in small-

molecule crystal structures. It comprises over 875 000 curated crystal structures, complete with

a wide array of metadata and structural information.1

The CCDC software distribution already contains tools for docking conformers to a given

binding site[13], and ligand based virtual screening (LBVS) tools involving pharmacophore

models. These packaged drug discovery tools, while informed by the wealth of connectivity

and geometry information embedded in the CSD, do not currently exploit the geometries (or

molecular conformations) from small-molecule structures themselves as potential ligands. In

light of the immense value[14] that small molecule crystal structures provide when used in

combination with protein-ligand structures from the PDB in drug discovery, it seems reasonable

to propose the CSD as a potential ligand database.

Further, it has previously been noted [15] that while known drugs binding to the same

target tend to exhibit a high level of two-dimensional similarity, it is more likely that this is

a consequence of inductive bias in the process of human reasoning in the drug development

process, than a result of some underlying physical basis.

Thus, virtual screening methods which account for molecular shapes in three dimensions

are desirable, and the intermolecular interactions that are intrinsic to small molecule crystal

structures – in an energetically favourable environment – may prove to be an underutilised

path to identifying potential ligands apart from structural analogues of a model molecule (and

potentially apart from patented chemistry).

In this context, we recently developed an efficient computational method[16] that enables high-

throughput profiling of molecular shapes in crystals using spherical harmonic shape descriptors

of Hirshfeld surfaces (HS). This method has been used to classify metals based on their lattice

types, to identify molecular structures in the CSD based on chemical scaffolds and conformational

similarity, and to identify chemically isosteric structures.

1For reference, the Protein Data Bank[12] (PDB) contains roughly 120 000 crystal structures.

70

CHAPTER 4. FAST SCREENING OF CRYSTAL STRUCTURES FORLIGANDS BASED ON MOLECULAR SHAPE AND INTERMOLECULARINTERACTIONS

Ligand PDB Recorded

# refcode Activity

1 1KI2 [20] DNA polymerase inhibitor

2 1UWH [21] Protein kinase inhibitor

3 1Z95 [22] Nonsteroidal antiandrogen

4 3A2O [23] HIV protease inhibitor

Table 4.1: Protein-ligand co-crystals examined within this study, with an indication of thenature of their biological activity

4.2.1 Selection of crystal structures

In order to produce a database of reliable single crystal structures for molecules with light nuclei

(suitable for quantum chemical calculations) the following protocol, was adopted. Structures

were selected from the CSD based on the following criteria: No disorder, no powder structures,

no ions, no polymers, 3D coordinates determined, and X-ray R-factor less than 0.05. Also,

the heaviest permitted element was Argon. In order to obtain reliable H atom position, the

geometries of these structures were adjusted so that B−H, C−H, N−H and O−H bond lengths

were equal to the standard neutron values.[19].

During the process of calculating surfaces for all molecules in the unit cell of each crys-

tal structure (see below for details) some calculations failed, and as such their respective

crystal structures were discarded. The complete list of 116 006 structures is available from

https://figshare.com/ under https://doi.org/10.6084/m9.figshare.5576350.v1.

4.2.2 Selected ligands

Protein-ligand systems

The four protein-ligand systems (Table 4.1 and Figure 4.2) studied in this paper were selected

based their reported drug activity, and to be representative of a variety of molecular sizes and

shapes, along with chemical diversity. More importantly, all four ligands have high quality

crystal structures (Table 4.1) with their respective proteins in their binding site of activity.

Small drug molecules exhibiting polymorphism

Further to the study of protein-ligand interactions, in order to further examine the impact

of ESP as the described surface property as opposed to dnorm, we have examined use of this

procedure for a set of eight drug molecules in the CSD exhibiting polymorphism (Table 4.2).

These structures all have at least one structural polymorph in the CSD, along with some

redeterminations of the same structures. The different crystal structures allow us to explore

72

4.2. METHODOLOGY

1 Ganciclovir

3 Bicalutamide

2 Sorafenib

4 KNI-1689

Figure 4.2: 2D chemical structures of the four ligands presented in this study.

not only the effect of the different conformations, but of different interaction environments

which may be exhibited through polymorphism (i.e. changes in close contacts). The ability

of the method to identify similar shaped molecules from the CSD for a given query ligand is

thus tested – based on the success of identifying its own polymorphs in the CSD i.e. the same

molecule exhibiting different crystal structures.

4.2.3 Calculation of the Hirshfeld and Promolecule electron density

isosurfaces

The promolecule[32] electron density of a molecule comprises the sum of the spherically-averaged

electron densities of the constituent atoms. The isosurfaces of the promolecule electron density

used in this work were calculated at an isovalue of 0.02 au, which has been found to produce

shapes which agree well with the CPK model[33] and which account well for the void spaces

observed from Helium pycnometry measurements [34].

The Hirshfeld Surfaces describes the shape of a molecule within some surroundings (for

example, the surrounding molecules in the infinite crystal lattice). It is defined as the bounding

surface such that, within the surface, more than half of the promolecule electron density comes

from the enclosed molecule[35].

In reality, however, the protein pockets and the binding ligands are generally surrounded by

water molecules in biological environment, constituting a crystal-like environment in the sense

73

CHAPTER 4. FAST SCREENING OF CRYSTAL STRUCTURES FORLIGANDS BASED ON MOLECULAR SHAPE AND INTERMOLECULARINTERACTIONS

Structure (Polymorph) Name Other Polymorphs

ACSALA21 (I) [24] Aspirin ACSALA20 (II)

ATDZSA03 (II) [25] Acetazolamide ATDZSA04 (I)

CBMZPN01 (III) [26] Carbamazepine CBMZPN11 (I)

CBMZPN03 (II)

CBMZPN12 (IV)

CBMZPN16 (V)

HCSBTZ03 (I) [27] Hydrochlorothiazide HCSBTZ04 (II)

HXACAN07 (I) [28] Paracetamol HXACAN08 (II)

NICOAM06 (I) [29] Nicotinamide NICOAM04 (II)

SLFNMA01 (Pbca) [30] Sulfamerazine SLFNMA03 (P21/c)

SLFNMA04 (Pna21)

SUTHAZ07 (I) [31] Sulfathiazole SUTHAZ03 (II)

SUTHAZ11 (III)

SUTHAZ04 (IV)

SUTHAZ05 (V)

Table 4.2: Small drug molecules exhibiting polymorphism with their Cambridge StructuralDatabase (CSD) refcode(s). Those that appear in bold were used as the ‘query’ polymorphs forthis investigation, and their references are provided. Refer to text for details.

74

4.2. METHODOLOGY

of intermolecular interactions which would contribute to a HS or surface properties like dnorm.

In order to account for this aqueous environment, and to ensure closed interaction surfaces

around the ligand molecules we performed molecular dynamics simulations with a large number

of water molecules solvating the protein-ligand systems.

Molecular Dynamics simulations

The initial coordinates used to build the protein-ligand models for the present study were based

on the X-ray crystal structures presented in Table 4.1. The bound ligand and the protein

were described by the AMBER GAFF force field[36] and the AMBER14SB force field[37],

respectively. The electrostatic potential (ESP) was calculated at the HF/6-31G(d) level of

theory using Gaussian 09 program suite.[38] The computed ESP of the bound ligandwas fitted

using restrained electrostatic potential (RESP) charge [39] fitting in antechamber.

The whole system was solvated into an 81× 85× 78 �A3rectangular box (taking 1KI2 for

example) of TIP3P water[40] with a 12 �A buffer distance between the box edge and the nearest

solute atoms. The system was neutralized by adding Na+/Cl– ions. The protons were added

automatically, and the topology parameters as well as initial coordinates were generated by the

tleap AmberTool. After the energy minimization (step (1)), the system was heated gradually

from 0 to 300 K for 100 ps (step (2)), and another 100 ps MD simulation was carried out to relax

the system density to 1.0 g cm−3 (step (3)). Finally, the standard 10 ns MD simulation under

isothermal-isobaric (NPT) ensemble was employed with an integration time step of 2 fs using the

periodic boundary condition (step (4)). The cutoff value was set to 10 �A for van der Waals and

electrostatic interaction calculations. Long-range electrostatic interactions were dealt with by

the Particle Mesh Ewald (PME) method.[41, 42] The Langevin thermostat (NTT = 3) was used

to maintain the temperature at 300 K. All the bonds involving hydrogen atoms were constrained

using the SHAKE scheme[43]. The whole protein (including the inhibitor) was frozen when

running steps (2)-(4), but no residues were frozen when running the energy minimisation step.

After all 4 steps finished, the final models were obtained by removing all the residues beyond 5�A of the inhibitor based on the stable protein structure. All MD simulations were accomplished

using the AMBER 14 software[44].

4.2.4 Calculation of combined shape and surface property descrip-

tors

The method used to describe Hirshfeld and promolecular surfaces by spherical harmonic transform

(SHT) is almost identical to that previously described[16], with one difference: the definition of

the function to be analysed is no longer three-dimensional shape alone. Because we wish to

examine both shape and properties mapped on the isosurface, we encode this information in a

75

CHAPTER 4. FAST SCREENING OF CRYSTAL STRUCTURES FORLIGANDS BASED ON MOLECULAR SHAPE AND INTERMOLECULARINTERACTIONS

single complex-valued function

f(θ, φ) = r′(θ, φ) + i prop′(θ, φ). (4.1)

Using a complex valued function, equation (4.1), with the real component as the shape and

the imaginary component reduces the number of invariants i.e. size of the matching vector (see

ESI for complete details) compared to performing separate transforms for each property. In the

above equation, (θ, φ) are the polar and azimuthal angles coordinates of a point on the surface

relative to the barycenter of the triangulated surface (see below), r′ is the normalized distance

from the barycentre of the surface, prop′ is the surface property (i.e. either the electrostatic

potential V or dnorm), scaled to fit the interval [0, 1] as follows:

r′(θ, φ) =r(θ, φ)

r(4.2)

prop′(θ, φ) =prop(θ, φ)−min(prop)

max(prop)−min(prop), (4.3)

where r(θ, φ) and prop(θ, φ) are the raw non-normalised surface radius and surface property,

respectively (i.e. either the electrostatic potential or dnorm) and r is the mean radius of the

shape. The shape description vector then has the following form:

{r,min(prop),max(prop)−min(prop), N0, N1 . . . Nlmax} (4.4)

encoded through a spherical harmonic transform, and then using N type rotation invariants[16],

clm =

2π∫0

π∫0

dθ sin(θ)f(θ, φ)Y ml (θ, φ)∗ (4.5)

Nl =l∑

m=−l

clmc∗lm (4.6)

The values of the function f on the triangulated surfaces (See below) were obtained by interpo-

lation, as described previously[16].

In this work lmax was taken to be 20 i.e. 20 N -type invariants, in combination with 3 other

invariants describing extent (mean radius), minimum surface property and range of the surface

property. Thus the shape descriptors were 23-dimensional vectors. All spherical harmonic

transforms (synthesis and analysis) in this work were performed using th SHTns library [45].

76

4.2. METHODOLOGY

4.2.5 Construction of the shape descriptor library

Triangulated Hirshfeld and promolecular surfaces with isovalue 0.02 were calculated as described

previously for each symmetry unique molecule in the unit cell, along with surface properties

like de, di, dnorm[46] etc. The surfaces, along with dnorm were used as the library for querying

matches for given molecules, consisting of 138 527 surfaces described up to lmax = 20 – yielding

24 descriptors for each molecular shape.

Once the top 100 matches for each candidate structure in Table 4.1 were found, their wave

functions were calculated and descriptors using the ESP on the surface were examined, in

order to assess the potential benefit of accurate potentials as opposed to simpler proxies for

interactions on the surface.

4.2.6 Wavefunction calculations

Molecular wavefunctions, used to calculate the ESP on the Hirshfeld or promolecular surfaces

for the top n matches (n = 100 in the case of the protein-ligand structures, n = 20 in the case

of polymorphic drug molecules), were calculated using B3LYP/6-31G(d,p) at the experimental

geometries, except that anyB−H, C−H, N−H and O−H bond lengths were adjusted to the

standard neutron values[19] (i.e. the same geometries were used for surfaces and wavefunctions).

Ligand natoms npoints RMSD MSD MAD texact tb=1

/s /s

1 31 6024 0.0071 -0.00069 0.0044 66 8

2 48 10114 0.0066 -0.00021 0.0052 286 24

3 43 8950 0.0057 -0.00130 0.0050 220 20

4 80 12786 0.0048 -0.00026 0.0032 871 47

Table 4.3: Comparison of the root mean-squared deviation (RMSD), mean-signed deviation(MSD) and mean absolute-deviation between the exact and approximate electrostatic potentials(ESPs) (in au) for the ligand in the ligand-protein complex using electron densities for ligandwavefunctions at the B3LYP/6-31G(d,p) level. Also shown are the number of atoms in theligand natoms, the number of points in the isosurfaces npoints, and representative times for thecalculations of the exact ESP texact and for the approximate ESP (tb=1) where the multipoleapproximation is used for atoms separated by more than two bond connections. Refer to textfor further details.

77

CHAPTER 4. FAST SCREENING OF CRYSTAL STRUCTURES FORLIGANDS BASED ON MOLECULAR SHAPE AND INTERMOLECULARINTERACTIONS

4.2.7 Electrostatic potential (ESP) calculations

Since ESP calculation on large grids can rapidly become prohibitively expensive for large

molecules, and since in this work we only use the ESP for the purposes of truncated SHT shape

and property descriptors, we have adopted a fast and approximate scheme for ESP evaluation.

First, the electron density is assigned to an atom using Tanaka’s partitioning technique

– itself a modification of Mulliken’s[47] method, applied at the primitive basis function level,

which makes use of the gaussian product theorem[48]. The contributions to the ESP involving

basis functions on atom-pairs which are within a certain number of bonds b are evaluated

exactly (whether atoms are ‘bonded’ is decided using a distance criteria also employed by the

Cambridge Structural Database[11]). The contributions from the remaining basis-function pairs

are evaluated using a multipole expansion up to octupole moments (details of the multipole

formulae are given in the appendix since they have not appeared in the literature in exactly the

form we use them). The required matrix elements are evaluated using the Rys method[49], and

the method was implemented in the open-source Tonto package[50].

In order to test and quantify the accuracy of the apprimation scheme we present statistics

comparing ESPs calculated exactly, and with our approximation with b = 1, in Table 4.3

evaluated on Hirshfeld Surfaces and promolecular density isosurfaces with isovalue equal to 0.01

au. Even though the mean signed deviation is systematically negative (the approximated ESP

is smaller than the exact ESP) the deviations are very small. We find the agreement to be very

satisfactory, and the timings for the ESP evaluation are much reduced.

4.2.8 Shape Matching

Once the vector of shape descriptors for each library compound is calculated, the closest matches

or most similar shapes for a given query shape reduces to the problem of nearest neighbours

in a coordinate space. In this analysis, this was performed through construction of a K-d-tree

with d = 12[51] and using Euclidean distances between the shape descriptor vectors. The

‘score’ associated with each shape was it’s distance from the reference shape. Matching the top

100 candidates using this scheme with the database described was very quick, less than one

millisecond on a normal computer.

4.2.9 Docking

Docking calculations (along with the previously described ESP calculations) was restricted only

to the top 100 matches for each compound with both surface types (Hirshfeld and promolecular

density). Thus the docking scores, and shape-ESP surfaces for approximately 800 compounds

were calculated in order to examine the value of this method in LBVS approaches, and to assess

the additional potential value offered by calculation of ESP in addition to the surface. A broad

78

4.3. RESULTS

a useful library to find potential drug candidates.

For both the Hirshfeld and promolecule surfaces, the distribution off the docking scores

is multimodal. Analysis of these main modes, perhaps by cluster analysis, may give useful

information as concerning the responsible shape-property patterns.

The Hirshfeld and promolecule surfaces show a somewhat similar distributions in docking

scores, with roughly the same number of modes. The Hirshfeld surface shows a slightly higher

peak in the main mode of the distribution. It is noteworthy that the top 100 matches for the

different surface types had only small overlaps: 3, 19, 9 and 10 respectively for ligands 1, 2,

3 and 4. This indicates that the HS and PS do indeed select for distinct molecules based on

shape and intermolecular interactions. Further, these overlaps between PS and HS matches

in concert with the distributions of docking scores in Figure 4.4, indicates that the available

similar molecules in the CSD are more abundant for smaller, simpler molecules like ligand 1

than for larger molecules like ligand 2.

It is interesting to test whether the shape-dnorm scores have any correlation with the docking

scores. Results for correlation coefficients of the top 100 matches are shwon in Table (4.4).

However, only a weak correlation was found (the best was 0.48). There are also some negative

correlations, indicating that further ‘distances’ of the shape descriptors from the reference are

associated with better docking scores. This indicates that the method presented here and the

docking scores are distinct, or in some sense ‘orthogonal’ or ‘independent’. Perhaps this is

unsurprising, given how different the methods are.

It is worth noting that the case of ligand 2 may represent a problematic case for single

centre expansions under the spherical harmonic transform as the barycentre of the molecule lies

outside its surface. This can be seen by small artifacts present in the reconstruction of the shape

vs. the original shape, with errors of 1.7%, 8.1%, 4.1% and 2.8% (RMSD, over all grid points)

respectively for ligands 1, 2, 3, and 4. Nevertheless, because these errors are small, largely

localised to the higher order spherical harmonic coefficients and similar errors associated with

this will be likely be present in molecules of the same shape in the CSD, it seems that the results

for the shape description are not significantly affected – indicating that despite limitations this

shape description is still good enough to prove useful.

4.3.2 Matches with known drugs in the CSD

The CSD contains roughly 2000 structures which are indicated to be drugs2, of which 500 or so

were within our restricted subset. It is interesting to examine if any of the top 100 matches are

drug molecules according to this definition. Table 4.5) shows that, in fact, 12 existing drugs

were found.

If the docking scores are any indication, then many of these drugs may share activity with the

2In this case, matching text search for ‘drug’ or possessing a DrugBank[55] identifier

81

CHAPTER 4. FAST SCREENING OF CRYSTAL STRUCTURES FORLIGANDS BASED ON MOLECULAR SHAPE AND INTERMOLECULARINTERACTIONS

Ligand Surface Property r

1KI2 Hirshfeld dnorm 0.17

ESP -0.01

Promolecule dnorm 0.24

ESP 0.33

1UWH Hirshfeld dnorm 0.48

ESP 0.15

Promolecule dnorm 0.27

ESP -0.25

1Z95 Hirshfeld dnorm 0.08

ESP 0.02

Promolecule dnorm -0.11

ESP 0.19

3A2O Hirshfeld dnorm -0.03

ESP -0.42

Promolecule dnorm 0.17

ESP 0.38

Table 4.4: Pearson correlation coefficients R between shape-property descriptors used in thispaper and and docking scores from FRED/Chemgauss4 for the top 100 matches. Refer to textfor details.

82

4.3. RESULTS

actual protein ligand. Indeed, in the case of 2, (QABCEE) is a known kinase inhibitor, and for 4,

both (WUWCUO) and (DIJWOJ) are drugs with known antiviral activity (specifically targeting

HIV). This gives us confidence that the methods developed here are producing potentially

valuable matches, through identifying molecules which have already been assessed for their

biocompatibility and utilising them for novel targets.

Ligand Surface CSD Molecular Docking

# type Refcode Formula Score

1 Hirshfeld BHHPHE C13H21NO3 -10.94

2 Hirshfeld KAJFEJ C23Cl2H27N3O2 -13.32

Promolecule QABCEE C21ClF4H15N4O3 -17.36

MELFIT C23Cl2H27N3O2 -13.32

3 Hirshfeld INDMET C19ClH16NO4 -11.46

CERRAS C22H25N3O2 -12.96

WUWCUO C20H32N5O8P 1.42

CIHFAA C16H14N2O6S -6.58

Promolecule VEHRAB C24Cl2H21NO6 -5.08

4 Hirshfeld WUWCUO C20H32N5O8P -9.24

Promolecule ACTDGN C29H44O9 -10.23

DIJWOJ C36H47N5O4 -10.00

Table 4.5: Ligands found in the top 100 matches which are labelled as drugs in the CambridgeStructural database.

4.3.3 Small polymorphic drug molecules in the CSD

In order to further examine the differences between using dnorm as opposed to the ESP and the

surface property in the descriptors, we applied both shape-property descriptors to find the top

20 matches for the given drug molecule polymorphs listed in Table 4.2. The intention is to

examine what effect molecular conformation crystal environment of the ligand might have on

potential drug candidates. Since polymorphs are different crystal structures associated to the

same molecule, they will have different Hirshfeld structures, and different conformations of these

molecules in these polymorphs will have different promolecular density isosurfaces. Results are

83

CHAPTER 4. FAST SCREENING OF CRYSTAL STRUCTURES FORLIGANDS BASED ON MOLECULAR SHAPE AND INTERMOLECULARINTERACTIONS

presented in Figure 4.5.

We observe that more polymorphs or structure re-determinations are found with the PS

rather than the HS shape. This is understandable, as the PS is independent of crystal structure

(but will depend on conformation) whereas the HS is dependent on both. Therefore we expect

that the PS (and the ESP on that surface) are largely conserved under polymorphism. These

results highlight the advantages and disadvantages in terms of conformational and crystalline

environment specificity between different surfaces and surface properties: if one wants matches

more related to the crystal environment then the HS or dnorm should be used, whereas if

one wants matches characteristic of only the molecule, then the PS should be used. This is

particularly evident in the case of sulfathiazole, where polymorphs and redeterminations made

up 3/8 (HS/PS) of the top 20 matches with dnorm, but 17/18 (HS/PS) when using ESP.

An intriguing result can be seen in the case of the top matches for sulfamerazine, where

sulfathiazole (SUTHAZ) and sulfamethoxazole (SLFNMB) can be found to in the top matches

when using PS, in many cases closer than other sulfamerazine structures. All three of these

ligands are sulfonamide molecules with antibiotic activity. They are also similar in terms of their

3D shapes, but because of their ‘U’ shape and the different angular conformations the shapes

may appear quite distinct under a spherical harmonic expansion. Nevertheless, the similarity of

these drugs in terms of structure, shape and even their effect (they are both antibiotic drugs) is

echoed in their matching under the proposed shape descriptors.

4.4 Summary and future perspectives

In summary, the search for potential ligand molecules in the CSD, based on the ‘molecular shape

plus surface property’ matching method shows promising outcomes. While a search based on the

HS generated from the known drug-receptor structures resulted in ligands with comparable or

higher docking scores, their chemical structures showed significant variety. The validity of this

method has been further verified through the examination of a series of drug molecules in the

CSD exhibiting polymorphism. The results demonstrate the potential of combining quantitative

approaches to profile intermolecular interaction surfaces in crystals with the wealth of chemical

information embedded in a small molecule crystal structure database (the CSD) for rational

drug discovery.

In this paper we have described a method to create rotation-invariant shape-property

descriptors for isolated molecules, and molecules in an environment. The method has been used

to search the Cambridge Structural Database. The method has proven to successfully describe

shape sufficiently to find related drug molecules, and efficiently enough to do perform real-time

querying of desired ligands for LBVS.

There remain some key areas for possible improvement. For example, it may be the case

that we are over-describing the shape, and more rudimentary (i.e. lower angular momenta l

84

CHAPTER 4. FAST SCREENING OF CRYSTAL STRUCTURES FORLIGANDS BASED ON MOLECULAR SHAPE AND INTERMOLECULARINTERACTIONS

levels) descriptions might prove more effective in LBVS approaches. Further, this method could

be extended by modification of the surface decoration: by only ‘colouring’ specific regions of

interest on the surface i.e. those which are associated with only specific interactions, rather

than the holistic approach undertaken here. This might prove more effective particularly for

larger compounds, where some regions of ESP might be largely irrelevant to the binding affinity.

While we used the model HS and ESP generated from known ligand-protein structures as a

pharmacophore descriptor in the present study, it is also possible to extend this method for drug

design. Instead of generating the HS around the ligand, it is possible to build a hypothetical,

ideal ligand molecule in the receptor pocket and use the volume occupied by this ideal ligand

for the search-and- match procedure. The ideal ligand can be designed utilizing the concept

of supramolecular synthons, as a molecule with adequate functional groups that can form the

most effective intermolecular interactions with the pocket. Such descriptors would not simply

represent the molecular structure, but the shape and intermolecular interactions in terms of the

Hirshfeld-ESP surfaces of the ideally designed ligands, and thus they may be used to find the

most suitable inhibitor molecules from the CSD or other small molecule databases.

86

Bibliography

1. Fischer, E. Einfluss der Configuration auf die Wirkung der Enzyme. Ber. Dtsch. Chem.

Ges. 27, 2985. issn: 1099-0682 (1894).

2. Cheng, T., Li, Q., Zhou, Z., Wang, Y. & Bryant, S. H. Structure-Based Virtual Screening

for Drug Discovery: a Problem-Centric Review. AAPS J. 14, 133 (2012).

3. Di Giovanni, A. L. & C. Virtual Screening Strategies in Drug Discovery: A Critical Review.

Curr. Med. Chem. 20, 2839 (2013).

4. Bemis, G. W. & Kuntz, I. D. A fast and efficient method for 2D and 3D molecular shape

description. J. Comput. Aided Mol. Des. 6, 607 (1992).

5. Hahn, M. Three-Dimensional Shape-Based Searching of Conformationally Flexible Com-

pounds. J. Chem. Inf. Comput. Sci. 37, 80 (1997).

6. Hodgkin, E. E. & Richards, W. G. Molecular similarity based on electrostatic potential

and electric field. Int. J. Quantum Chem. 32, 105 (1987).

7. Buchbauer, G., Klinsky, A., Weiß-Greiler, P. & Wolschann, P. Ab initio Molecular

Electrostatic Potential Grid Maps for Quantitative Similarity Calculations of Organic

Compounds. Mol. Model. Ann. 6, 425 (2000).

8. Melani, F., Gratteri, P., Adamo, M. & Bonaccini, C. Field Interaction and Geometrical

Overlap: A New Simplex and Experimental Design Based Computational Procedure for

Superposing Small Ligand Molecules. J. Med. Chem. 46, 1359 (2003).

9. Vainio, M. J., Puranen, J. S. & Johnson, M. S. ShaEP: Molecular Overlay Based on

Shape and Electrostatic Potential. J. Chem. Inf. Model. 49, 492 (2009).

10. Cho, S. J. & Sun, Y. FLAME: A Program to Flexibly Align Molecules. J. Chem. Inf.

Model. 46, 298 (2006).

11. Groom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. The Cambridge Structural

Database. Acta Crystallogr. B 72, 171 (2016).

12. Berman, H. M. The Protein Data Bank. Nucleic acids research 28, 235 (2000).

13. Jones, G., Willett, P., Glen, R. C., Leach, A. R. & Taylor, R. Development and validation

of a genetic algorithm for flexible docking. J. Mol. Biol. 267, 727.

87

BIBLIOGRAPHY

14. Groom, C. R. & Cole, J. C. The use of small-molecule structures to complement protein-

ligand crystal structures in drug discovery. Acta Crystallogr. D 73, 240 (2017).

15. Cleves, A. E. & Jain, A. N. Effects of inductive bias on computational evaluations of

ligand-based modeling and on drug discovery. J. Comput. Aided Mol. Des. 22, 147 (2008).

16. Spackman, P. R., Thomas, S. P. & Jayatilaka, D. High Throughput Profiling of Molecular

Shapes in Crystals. Sci. Rep. 6, 22204 (2016).

17. Huang, N., Shoichet, B. K. & Irwin, J. J. Benchmarking Sets for Molecular Docking. J.

Med. Chem. 49, 6789 (2006).

18. Mysinger, M. M., Carchia, M., Irwin, J. J. & Shoichet, B. K. Directory of Useful Decoys,

Enhanced (DUD-E): Better Ligands and Decoys for Better Benchmarking. J. Med. Chem.

55, 6582 (2012).

19. Allen, F. H. & Bruno, I. J. Bond lengths in organic and metal-organic compounds revisited:

X-H bond lengths from neutron diffraction data. Acta Crystallogr. B 66, 380 (2010).

20. Champness, J. N. et al. Exploring the active site of herpes simplex virus type-1 thymidine

kinase by X-ray crystallography of complexes with aciclovir and other ligands. Proteins

32, 350 (1998).

21. Wan, P. T. C. et al. Mechanism of activation of the RAF-ERK signaling pathway by

oncogenic mutations of B-RAF. Cell 116, 855 (2004).

22. Bohl, C. E., Gao, W., Miller, D. D., Bell, C. E. & Dalton, J. T. Structural basis for

antagonism and resistance of bicalutamide in prostate cancer. Proceedings of the National

Academy of Sciences of the United States of America 102, 6201 (2005).

23. Hidaka, K. et al. Small-sized human immunodeficiency virus type-1 protease inhibitors

containing allophenylnorstatine to explore the S2’ pocket. J Med Chem 52, 7604 (2009).

24. Arputharaj, D. S., Hathwar, V. R., Guru Row, T. N. & Kumaradhas, P. Topological

Electron Density Analysis and Electrostatic Properties of Aspirin: An Experimental and

Theoretical Study. Crystal Growth & Design 12, 4357 (2012).

25. Thomas, S. P., Jayatilaka, D. & Guru Row, T. N. S[three dots, centered]O chalcogen

bonding in sulfa drugs: insights from multipole charge density and X-ray wavefunction of

acetazolamide. Phys. Chem. Chem. Phys. 17, 25411 (2015).

26. Reboul, J. P., Cristau, B., Soyfer, J. C. & Astier, J. P. 5 H-Dibenz[ b, f]azepinecarboxamide-

5 (carbamazepine). Acta Crystallogr. B 37, 1844 (1981).

27. Leech, C. K., Fabbiani, F. P. A., Shankland, K., David, W. I. F. & Ibberson, R. M. Accurate

molecular structures of chlorothiazide and hydrochlorothiazide by joint refinement against

powder neutron and X-ray diffraction data. Acta Crystallogr. B 64, 101 (2008).

88

BIBLIOGRAPHY

28. Nichols, G. & Frampton, C. S. Physicochemical Characterization of the Orthorhombic

Polymorph of Paracetamol Crystallized from Solution. J. Pharma. Sci. 87, 684 (1998).

29. Jarzembska, K. N. et al. Combined Experimental and Computational Studies of Pyrazi-

namide and Nicotinamide in the Context of Crystal Engineering and Thermodynamics.

Cryst. Growth Des. 14, 3453 (2014).

30. Ravindra Acharya, K., Kuchela, K. N. & Kartha, G. Crystal structure of sulfamerazine.

J. Crystallogr. Spectrosc. Res. 12, 369 (1982).

31. Drebushchak, T. N., Boldyreva, E. V. & Mikhailenko, M. A. Crystal structures of

sulfathiazole polymorphs in the temperature range 100–295 K: A comparative analysis. J.

Struct. Chem. 49, 84 (2008).

32. Spackman, M. A. & Maslen, E. N. Chemical-properties from the promolecule. J. Phys.

Chem. 90, 2020 (1986).

33. Mitchell, A. S. & Spackman, M. A. Molecular surfaces from the promolecule: A comparison

with Hartree–Fock ab initio electron density surfaces. Journal of Computational Chemistry

21, 933 (2000).

34. Turner, M. J., McKinnon, J. J., Jayatilaka, D. & Spackman, M. A. Visualisation and

characterisation of voids in crystalline materials. CrystEngComm 13, 1804 (2011).

35. Spackman, M. A. & Byrom, P. G. A novel definition of a molecule in a crystal. Chem.

Phys. Lett. 267, 215 (1997).

36. Wang, J., Wolf, R. M., Caldwell, J. W., Kollman, P. A. & Case, D. A. Development

and testing of a general amber force field. Journal of Computational Chemistry 25, 1157

(2004).

37. Maier, J. A. et al. ff14SB: Improving the Accuracy of Protein Side Chain and Backbone

Parameters from ff99SB. Journal of Chemical Theory and Computation 11, 3696 (2015).

38. Frisch, M. J. et al. Gaussian 09 Revision D. 01

39. Bayly, C. I., Cieplak, P., Cornell, W. & Kollman, P. A. A well-behaved electrostatic

potential based method using charge restraints for deriving atomic charges: the RESP

model. J. Phys. Chem. 97, 10269 (1993).

40. Jorgensen, W. L., Chandrasekhar, J., Madura, J. D., Impey, R. W. & Klein, M. L.

Comparison of simple potential functions for simulating liquid water. J. Chem. Phys. 79,

926 (1983).

41. Darden, T., York, D. & Pedersen, L. Particle mesh Ewald: An Nlog( N) method for Ewald

sums in large systems. J. Chem. Phys. 98, 10089 (1993).

42. Essmann, U. et al. A smooth particle mesh Ewald method. J. Chem. Phys. 103, 8577

(1995).

89

BIBLIOGRAPHY

43. Ryckaert, J.-P., Ciccotti, G. & Berendsen, H. J. C. Numerical integration of the cartesian

equations of motion of a system with constraints: molecular dynamics of n-alkanes. Journal

of Computational Physics 23, 327 (1977).

44. Case, D. A. et al. Amber 14 (2014).

45. Schaeffer, N. Efficient spherical harmonic transforms aimed at pseudospectral numerical

simulations. Geochem. Geophys. 14, 751 (2013).

46. Spackman, M. A. & McKinnon, J. J. Fingerprinting intermolecular interactions in molec-

ular crystals. Crystengcomm 4, 378 (2002).

47. Mulliken, R. S. Electronic Population Analysis on LCAO–MO Molecular Wave Functions.

I. J. Chem. Phys. 23, 1833 (1955).

48. Tanaka, K. X-ray analysis of wavefunctions by the least-squares method incorporating

orthonormality. I. General formalism. Acta Crystallogr. A: Crystal Physics 44, 1002

(1988).

49. Rys, J., Dupuis, M. & King, H. F. Computation of electron repulsion integrals using the

rys quadrature method. J. Comput. Chem. 4, 154 (1983).

50. Jayatilaka, D. & Grimwood, D. J. in Computational Science — ICCS 2003: International

Conference, Melbourne, Australia and St. Petersburg, Russia, June 2–4, 2003 Proceedings,

Part IV (eds Sloot, P. M. A. et al.) 142 (Springer Berlin Heidelberg, Berlin, Heidelberg,

2003).

51. Jones, E., Oliphant, T., Peterson, P., et al. SciPy: Open source scientific tools for Python

[Online; accessed 01/11/2017]. 2001. http://www.scipy.org/.

52. Hawkins, P. C. D., Skillman, A. G., Warren, G. L., Ellingson, B. A. & Stahl, M. T.

Conformer Generation with OMEGA: Algorithm and Validation Using High Quality

Structures from the Protein Databank and Cambridge Structural Database. J. Chem.

Inf. Model. 50, 572 (2010).

53. McGann, M. FRED and HYBRID docking performance on standardized datasets. J.

Comput. Aided Mol. Des. 26, 897 (2012).

54. McGann, M. FRED Pose Prediction and Virtual Screening Accuracy. J. Chem. Inf. Model.

51, 578 (2011).

55. Wishart, D. S. et al. DrugBank: a comprehensive resource for in silico drug discovery and

exploration. Nucleic acids research 34, D668 (2006).

90

Part III

Assessment of quantum chemical

predictions in chemistry

91

Chapter 5

Quantum Chemical Electron Impact

Mass Spectrum prediction for de novo

structure elucidation

ABSTRACT

We investigate the success of the Quantum Chemical Electron Impact Mass Spectrum (QCEIMS)

method in predicting the electron impact mass spectra of a diverse test set of 61 small molecules

selected to be representative of common fragmentations and reactions in electron impact mass

spectra. Comparison with experimental spectra is performed using the standard matching

algorithms, and the relative ranking position of the actual molecule matching the spectra

within the NIST-11 library is examined. We find that the correct spectrum is ranked in

the top two matches from structural isomers in more than 50% of the cases. QCEIMS thus

reproduces the distribution of peaks sufficiently well to identify the compounds, with the

RMSD and mean absolute difference (MAD) between appropriately normalized predicted and

experimental spectra being at most 9% and 3% respectively, even though the most intense

peaks are often qualitatively poorly reproduced. We also compare the QCEIMS method to

competitive fragmentation modeling for electron ionization (CFM-EI), a training-based mass

spectrum prediction method, and remarkably we find the QCEIMS performs equivalently or

better. We conclude that QCEIMS will be very useful for those who wish to identify new

compounds which are not well represented in the mass spectral databases.

5.1 Introduction

Mass spectrometry (MS) is of major importance for identifying the presence of small quantities

of a compound in a mixture, particularly because it is several orders of magnitude more sensitive

93

CHAPTER 5. QUANTUM CHEMICAL ELECTRON IMPACT MASSSPECTRUM PREDICTION FOR DE NOVO STRUCTURE ELUCIDATION

than equivalent structure determination methods[1]. Indeed, MS is often the only feasible

identification method[2–4] in contexts like the identification of semiochemicals [5–7], where

abundances of critical compounds may be in the ng range.

Electron-impact mass spectrometry (EIMS) coupled to some chromatographic separation

techniques is, at present, likely to be the predominant form of MS practiced for small molecules.

The typical protocol for identifying unknown compounds using EIMS involves matching against

a known mass spectrum (MSp), using spectral libraries in conjunction with software to identify

matches[8–13]. For example, the National Institute of Standards and Technology (NIST)

mass spectral library (version 11) contains electron impact (EI) spectra of more than 2× 105

compounds, and may be readily searched using the NIST MS Search program. However large,

such a library may only contain a tiny fraction of the total number of small molecular species

found in the universe, which for molecules less than 500 Daltons in weight is estimated to be in

excess of 1060[14]. Simply put, the huge number of possible structural isomers for each molecular

formula renders it practically impossible for a library to be exhaustive. If the identification

of a compound whose spectra has yet to be characterized (i.e. its spectra is not in a library)

is desired, much of the value provided by MS matching/library methods is lost (though the

structure of the close matches may still provide insight).

In this paper, we are concerned with the problem of identifying the structures of unknown

compounds present in low abundances which have not been previously characterized by MS. We

will use the term de novo structure determination from the mass spectrum (MSp) to describe this.

In this case, the only way to confirm the structure is to synthesize several putative candidate

compounds in significant amounts, characterize their structure independently (say using nuclear

magnetic resonance (NMR), or X-ray crystallography), then compare the candidates MSp with

the original unknown spectrum to see which (if any of them) agree. It should come as no surprise

that this is a time- and labor-intensive process.

One way to ease the process of de novo compound identification is to predict the MSp for

each candidate using computational methods. If such computed MSp were accurate enough,

only one candidate compound need be synthesized and characterized in order to to confirm the

structure. Indeed, if the computed MSp were always reliably accurate it would not even be

necessary to confirm the structure by synthesis and alternative means; MS may then become

competitive with NMR or crystallography as a structure determination method.

Computational methods for the prediction of the MSp have been well reviewed by Bauer

and Grimme [15]. They may be broadly separated into two categories:

1. Expert systems. Expert systems produce a MSp using rules derived from experimental

spectra, for example fitted to experimental MSp using a model such as a neural network[16,

17]. The ‘system’ is trained or based on a library of MSps for which the structure

is known. Indeed, the prototypical example of an expert computer system based on

rules[18] concerned the identification of molecular structure from MSp: the DENDRAL[19]

94

5.1. INTRODUCTION

project. Regardless of the successes (or failures[1, 20]) of the original DENDRAL project,

recently learning-based methods such as competitive fragment modeling (CFM)[21, 22] for

predicting EIMS spectra (CFM-EI) have gained more traction, and proved more successful.

It should be noted, though, that within any such system there will always exist some bias,

both in the model it represents and the data used to ‘train’ it; such bias may lead to

unquantifiable inaccuracies.

2. First principles simulation techniques. First principles simulation techniques aim to

actively simulate the underlying physical processes involved in MS. Such methods are

not in principle dependent on an existing library of known spectra, but are limited by

the accuracy with which one is able to solve the time-dependent Schrodinger equation

(or more correctly, the time-dependent Liouville equations for the density matrix[23]).

Unfortunately such calculations were, in general, time consuming to the point of being

impractical. However, recently the QCEIMS program from Grimme et al.[24] has allowed

practical calculation of MSps using ab initio techniques. By directly modeling the time-

dependent fragmentation process via Born-Oppenheimer molecular dynamics (BOMD)

finite-temperature (fractional orbital occupied) quantum chemical (QC) methods[15, 24,

25], coupled with knowledge of isotopic distributions[26], QCEIMS has proven remarkably

accurate. Other quantum chemical methods based on bond indices have also been

proposed[27], and are examined by us in related work.

The purpose of this paper is to investigate the performance of, QCEIMS, which, although

having been applied to many molecules[15, 24, 25, 28, 29], has not yet been tested by the

conventional MSp matching algorithms used to identify compounds from the MSp; an essential

task if it is to be used for de novo compound identification. For this purpose, we make use of

the standard programs and databases from the National Institute of Standards (NIST). We

also provide an estimate of the errors associated with their predicted spectra, using counting

statistics. Our goal is to identify patterns or trends with any errors in the prediction, particularly

those seen to be associated with a chemical class of compounds; such information is invaluable

for practical application of QCEIMS.

Despite expert systems and first principles methods having radically different philosophical

approaches, for the purposes of de novo structure determination these differences mean little.

As such, we also assess the performance of QCEIMS method against the CFM-EI model-based

spectrum calculator, both as a contemporary reference point and to investigate the differences

in accuracy between the two techniques.

95

CHAPTER 5. QUANTUM CHEMICAL ELECTRON IMPACT MASSSPECTRUM PREDICTION FOR DE NOVO STRUCTURE ELUCIDATION

5.2 Methodology

All data analysis and processing of calculations for this paper was performed using the Python

programming language. Notably, the numpy[30], pandas[31] packages, and matplotlib[32] for

the generation of figures.

5.2.1 Selection of compounds for analysis

When selecting a set of molecules for investigation, the following criteria were considered: they

must have an experimental MSp and structural isomers with experimental MSp, they must

be representative of common/important fragmentations in EIMS i.e. made up of a variety of

functional groups, they must be relatively small molecules (natom < 25). Further, it is desirable

to have a mixture of easy and difficult cases for differentiating between structural isomers. As

such, a set of 61 small molecules were chosen, made up of alcohols (6), aldehydes (3), alkenes

(7), amines (6), alkylbenzenes (8), carboxylic acids (5), esters (16), ketones (2), alkylphenols

(8). The complete list is shown in Table 5.1, and the structures are available in Figure 5.1. The

primary motivation for the selection of small molecules is the identification of failed predictions

and their underlying cause. Both of these aspects are significantly more straightforward in

the case of small molecules. Further, the reduced computational time requirements associated

with quantum-chemical predictions with small molecules render them ideal for such a study.

Additionally, when establishing the viability for such methods in de novo structure elucidation,

having several structural isomers distinguishable via their MSp is vitally important.

5.2.2 QCEIMS and CFM-EI calculations

Initial geometries were drawn by hand in a standard molecular builder program Avogadro2 [33]

and then optimized using the generalized Amber force field [34]. MSp prediction calculations

were then performed using the QCEIMS program version 2.26, as described by Grimme and

coworkers[15, 24] using the default parameters, namely: using the OM2-D3[35] semi-empirical

quantum mechanical wavefunction method from MNDO99[36] version 7.0. The default QCEIMS

parameters were used, with an initial temperature of 500K, at 70 eV and the number of

trajectories set to 25× nheavyatoms. The resulting spectra were normalized in the typical manner:

such that the highest peak had an intensity of 100. The time taken for calculations varied

with molecular size (see ESI Table S2 for full details), with the shortest being methyl formate

(roughly 1.5 CPU hours) and the longest being 4-propan-2-ylphenol (roughly 80 CPU hours).

It should be noted that these calculations are readily parallelized across computing clusters,

drastically reducing the wall-clock time required.

CFM-EI Calculations were performed using the method described by Allen et al.[22] with

the provided trained parameters for EI-MS, using the Windows executable ‘cfm-predict’, version

96

5.2. METHODOLOGY

CH3

HOCH

3

CH3

HO

CH3

HO

CH3

H3C

HO

CH3

HO

H3C

CH3

CH3

OHH3C

CH3

O

CH3

O

H3C

CH3

O

CH2

H3C

CH3

H3C CH

2

H3C

CH3

CH3

H2C

CH3

CH3

H3C

CH3

CH3

CH3

H2N

CH3

CH3

H2N

CH3

H2N

CH3

H3C

H2N

CH3

H2N

H3C

CH3

CH3

NH2

H3C

CH3

H3C

CH3

H3C CH

3

H3C

H3C

CH3

CH3

CH3

H3C

OH

O

CH3

OH

OCH

3

HO

O

H3C

CH3

CH3

OH

O

CH3

OH

O

H3C

O

OH3C

O

O

H3C

H3C

O

O

CH3

H3C

CH3

O

OH3C

OO

CH3

CH3

O

O

H3C

CH3

H3C CH

3O

O CH3

H3C

H3C

O

O CH3

H3C

O

O

CH3

H3C

O

O

CH3

H3C

O

O CH3

CH3

H3C

O

O CH3

H3C

O

O

CH3

CH3

H3C

O

O

CH3

CH3

H3C

O

O

CH3

H3C

O

O CH3

CH3

CH3

CH3

O

CH3

H3C

O

OH

OH

CH3

OH

CH3

OH

CH3

OH

CH3

H3C

OH

CH3

OH

CH3

CH3

OH

CH3

H3C

Figure 5.1: Structures of the molecules chosen for analysis grouped into chemical classes

97

CHAPTER 5. QUANTUM CHEMICAL ELECTRON IMPACT MASSSPECTRUM PREDICTION FOR DE NOVO STRUCTURE ELUCIDATION

2.2.

5.2.3 Spectrum matching

Rankings between standard-normalized calculated and experimental MSp were made using the

NIST MS Search v2.0g software coupled with the NIST11 database. Search was performed with

the ‘identity’ matching method where the square root of intensity of the MSp was taken and

weighted by the mass-to-charge ratio squared. The difference between this and the available

‘similarity’ metric (which differs only in the the mass weighting) is negligible for our purposes

– see [10] for a comprehensive study of the impact of matching algorithms on the accuracy

of matches). Searches were carried out by limiting to molecules to have the same molecular

formula, using only the two most abundant isotopes.

5.2.4 Ranking comparison

It is not meaningful to directly compare ranks when the number of possible candidates differs.

To overcome this, the relative ranking position[3, 22] has been used:

RRP =r − 1

n− 1(5.1)

where r is the rank, and n is the total number of candidates with the same molecular weight

(this expression is equivalent to that presented in Kerber et al. [3]). Thus a rank of one maps to

zero, and a rank of n maps to 1. The RRP is not defined for a class with only one member.

This quantity constitutes the metric for direct comparison between compounds with different

molecular formulae, as they clearly have different possible numbers of structural isomers and

therefore differing numbers of entries in the database.

5.2.5 Spectrum normalization for statistical comparisons

It is standard to normalize the mass spectra such that the highest peak has an intensity with

value 100. However, such a normalization makes direct comparison relative between error

statistics such as root mean square deviation (RMSD) across different compounds meaningless.

As such, these metrics (see Table 5.3) have been calculated with spectral intensities normalized

such that the sum of the intensities is unity for both the spectra being compared. A byproduct

of this normalization scheme is that the mean signed error between two spectra is always zero.

Such a normalization scheme has been used previously, by Rasmussen and Isenhour [37] who

demonstrated that this kind of probability distribution normalization (which they call ‘total

ion current normalization’) does not lead to a significant effect when tested for the purposes of

identifying a known mass spectrum.

98

5.3. RESULTS

Chemical RRP

No. Name Formula /nisomers

Alcohols

1 propan-1-ol C3H8O 1/3

2 propan-2-ol C3H8O 1/3

3 butan-1-ol C4H10O 1/8

4 butan-2-ol C4H10O 1/8

5 2-methylpropan-1-ol C4H10O 1/8

6 2-methylpropan-2-ol C4H10O 5/8

Aldehydes

7 propanal C3H6O 1/7

8 2-methylpropanal C4H8O 3/21

9 butanal C4H8O 3/21

Alkenes

10 but-1-ene C4H8 4/6

11 but-2-ene C4H8 1/6

12 pent-1-ene C5H10 3/13

13 pent-2-ene C5H10 3/13

14 hex-1-ene C6H12 6/30

15 hex-2-ene C6H12 1/30

16 hex-3-ene C6H12 9/30

Amines

17 propan-1-amine C3H9N 1/4

18 propan-2-amine C3H9N 1/4

19 butan-1-amine C4H11N 1/8

20 butan-2-amine C4H11N 1/8

21 2-methylpropan-1-amine C4H11N 2/8

22 2-methylpropan-2-amine C4H11N 1/8

Benzenes

23 benzene C6H6 1/5

24 toluene C7H8 3/13

25 ethylbenzene C8H10 8/23

26 propylbenzene C9H12 2/42

27 cumene C9H12 16/42

28 butylbenzene C10H14 3/74

29 butan-2-ylbenzene C10H14 6/74

30 2-methylpropylbenzene C10H14 1/74

Chemical RRP

No. Name Formula /nisomers

Carboxylic acids

31 propanoic acid C3H6O2 1/9

32 butanoic acid C4H8O2 7/27

33 2-methylpropanoic acid C4H8O2 3/27

34 2-methylbutanoic acid C5H10O2 3/47

35 3-methylbutanoic acid C5H10O2 5/47

Esters

36 methyl formate C2H4O2 1/3

37 methyl acetate C3H6O2 2/9

38 ethyl acetate C4H8O2 1/27

39 methyl propanoate C4H8O2 1/ 27

40 propyl acetate C5H10O2 NM/47

41 propan-2-yl acetate C5H10O2 4/47

42 methyl 2-methylpropanoate C5H10O2 1/47

43 ethyl propanoate C5H10O2 1/47

44 methyl butanoate C5H10O2 5/47

45 propyl propanoate C6H12O2 1/80

46 propan-2-yl propanoate C6H12O2 3/80

47 ethyl butanoate C6H12O2 4/80

48 methyl 2-methylbutanoate C6H12O2 2/80

49 methyl 3-methylbutanoate C6H12O2 7/80

50 propyl butanoate C7H14O2 2/83

51 propan-2-yl butanoate C7H14O2 NM/83

Ketones

52 acetone C3H6O 1/7

53 butan-2-one C4H8O 1/21

Phenols

54 phenol C6H6O 2/4

55 4-methylphenol C7H8O 2/17

56 4-ethylphenol C8H10O 16/43

57 4-propylphenol C9H12O NM/112

58 4-propan-2-ylphenol C9H12O 16/112

59 4-butylphenol C10H14O NM/163

60 4-butan-2-ylphenol C10H14O 3/163

61 4-(2-methylpropyl)phenol C10H14O 2/163

Table 5.1: Relative ranking position (RRP) of the matching mass spectra found by theQCEIMS program produced by the NIST MS Search program (version 2.0g). Ranking is relativeto the total number isomers nisomers. NM indicates that no match was found in the database.

5.2.6 Estimates of the errors in the calculated peak heights

Since the spectra are built up by counting independent fragmentation processes we assume that

the number of counts in a particular peak of the mass spectrum has a Poisson distribution.

Then the absolute error in a particular peak with Ni counts is√Ni. If Ntotal =

∑iNi is

the total number of counts in the whole mass spectrum the absolute error in a particular

probability-normalized mass spectrum with peak heights Ni/Ntotal is√Ni/Ntotal. It is important

to keep in mind that the reproducibility error in a particular mass spectrum is about 5-10% in

the peak heights, and in some cases even up to 30% [24].

5.3 Results

Table 5.1 shows the predicted RRPs for the QCEIMS method for the 61 molecules considered

in this study; rankings for CFM-EI (Table S0), along with predicted and experimental MSp for

each compound (Figures S1-S122) are shown in the electronic supplementary material (ESI)

and selected spectra are presented and discussed later.

99

CHAPTER 5. QUANTUM CHEMICAL ELECTRON IMPACT MASSSPECTRUM PREDICTION FOR DE NOVO STRUCTURE ELUCIDATION

0.0 0.1 0.2 0.3 0.4 0.5 0.6QCEIMS RRP

0

5

10

15

20

25

30

Freq

uency

Figure 5.2: The number of molecules predicted correctly at a given rank for the QCEIMSmethod. There were four molecules which were not matched to any molecule in the NISTdatabase.

5.3.1 The QCEIMS method

Overall success rate. We see from the table that using the QCEIMS method we are able to

identify the correct structure from the NIST database (using the standard spectrum matching

algorithm for compounds with the same chemical formula) in 24 of the 61 cases. The correct

spectrum is ranked in the top two in 32 of the cases (greater than 50% success rate) and in

the top three in 42 of the cases (69% success rate). Thus if the QCEIMS method were to be

used to actually identify the structure of an unknown compound similar to those in the test

set one might expect to get the correct result in the top three matches two-thirds of the time.

However one must qualify this statement in at least two respects. First, the test set is rather

small; and even with this small set there were four cases providing unacceptable matches (shown

as NM in the table). Also, as we previously remarked, comparing actual ranks is problematic

because different chemical classes have different numbers of isomers. Therefore, in Figure 5.2

we present a histogram of the number of molecules given a particular rank. Our conclusions

based on actual rank are essentially the same as those based on RRP.

Dependence of the quality of the rankings on the chemical class of the compound. Some of

the classes have too few candidates to make any clear statements, but we may still make the

following observations:

• In the case of the alcohols, all the primary alcohols are correctly ranked, but the tertiary

alcohol 6 was not correctly predicted (rank 5!).

• Some of the alkenes are very poorly ranked, but there is no clear correlation of the results

with the position of the double bond.

• Rankings for the amines were remarkably good, with all with the exception of 21 being

100

5.3. RESULTS

matched correctly.

• For the substituted benzenes, poorer results are obtained for candidates with multiple

alkyl substituents. This trend is exacerbated for the phenols where two of the compounds

57 and 59 were not matched at all.

• Substituted esters are generally well ranked, but tend to be more poorly predicted when

(like the substituted benzene) one of the groups becomes more branched.

Dependence of the quality of the results on the mass of the molecules. There is a weak trend

that increasing size of molecule produces worse matches perhaps best exemplified by the esters

(Pearson’s correlation coefficients r = 0.40). It should be kept in mind that matches were only

considered in the database between molecules of the same molecular formula. It should also be

noted that the size of the molecules is very small, and one might expect a strong dependence

with molecular weight.

Mean RRPs for each chemical class. Table 5.2 summarizes some of the discussion above

in a more quantitative way, presenting the mean RRPs for each chemical class. Although the

results for ketones and aldehydes seem very good, there are too few systems to consider the

mean RRPs meaningful for these two classes of compounds. However, we see that the esters and

amines are very well predicted, with the means RRPs for carboxylic acids, substituted benzenes

and phenols all being quite similar and low. The alcohols seem not well predicted but are in

fact skewed by one bad result. The alkenes are ranked the worst. It also deserves to be noted

that in the cases where there are a large number of isomers the matching algorithm ranks the

QCEIMS spectra very near the top e.g. for 30 we have 1/74 and for 61 we have 2/163.

Comparison of spectra with experiment. Table 5.3 gives the mean root-mean-square deviation

(RMSD) and mean absolute error (MAD) in the probability normalized spectra. We observe

that the spectra which agree the worst on the RMSDs (which highlight outlier differences)

are in order: the alcohols (with one outlier already noted in the previous section), aldehydes,

carboxylic acids, esters and alkenes. With regard to MADs, the worst classes are again the

alchohols, alkenes, and ketones. In the ESI we present the mass spectra produced for all the

compounds. Most informative are the differences between the predicted and observed spectra:

peaks are coloured green if they were present in both spectra, yellow if present only in the

experiment, and red if present only in the QCEIMS model spectra. Positive differences indicate

that the calculated spectra overestimate the peak intensities, an d error bars on the QCEIMS

spectra are presented. We see that the errors due to counting statistics are very small, the

maximum always less than 8% and usually much smaller, and then only appearing on the peaks

which have the largest counts. Therefore the difference in intensity between the experimental

and QCEIMS peaks for the larger peaks is not to be attributed to counting statistics. Also,

the differences in the larger peaks far exceeds the sum of the counting statistics error and any

101

CHAPTER 5. QUANTUM CHEMICAL ELECTRON IMPACT MASSSPECTRUM PREDICTION FOR DE NOVO STRUCTURE ELUCIDATION

Mean RRP No. of

Class QCEIMS CFM-EI Compounds

Alcohols 0.11 0.21 5

Aldehydes 0.07 0.02 3

Alkenes 0.20 0.30 7

Amines 0.02 0.08 6

Benzenes 0.12 0.12 8

Carboxylic acids 0.10 0.14 3

Esters 0.03 0.04 14

Ketones 0.00 0.00 2

Phenols 0.18 0.14 5

Overall 0.09 0.12 53

Table 5.2: Mean Relative ranking position (RRP) for the QCEIMS and CFM-EI methodsfor different classes of compounds, with the number of compounds in each class. Overall meanRRPs are given in the final line. Only molecules in which both methods matched are compared.

reproducibility error associated with the EIMS experiment. A typical example showing error

bars in presented in Figure 5.4, ethylbenzene, discussed in more detail later.

5.3.2 The CFM-EI method

Overall success rate. The CFM-EI has a very similar first match rate compared to the QCEIM

method, with 24 matching in first position, 35 in the top two positions.

Dependence of the quality of results on the chemical class. The full set of rankings for the

CFM-EI method is shown in Table S1 of the ESI. The following observations can be made.

• The alcohols are well ranked except for one outlier (3) which is placed in the 5th position

(this is not the same compound ranked 5th in the QCEIMS method).

• The alkenes are rather poorly ranked: both 12 and 16 are ranked at position 10.

• The amines are remarkably well ranked: all are ranked in the first position except 18 and

22, which are ranked in second place.

• For the substituted benzenes and phenols there are some very poor rankings, which are a

bit worse than for the QCEIMS method. For example, 28 is in position 20, 59 in position

35, and 61 had no match.

Mean RRPs for each chemical class. Table 5.2 shows that, as for the QCEIMS method, the

alkenes are worst predicted, followed by the alcohols. The results for the alcohols and alkenes are

102

5.4. DISCUSSION

Mean RMSD Mean MAD

Class QCEIMS CFM-EI QCEIMS CFM-EI

Alcohols 0.09 0.05 0.03 0.03

Aldehydes 0.08 0.03 0.03 0.01

Alkenes 0.05 0.05 0.02 0.03

Amines 0.04 0.03 0.02 0.02

Benzenes 0.04 0.04 0.01 0.02

Carboxylic acids 0.06 0.04 0.02 0.02

Esters 0.06 0.04 0.02 0.02

Ketones 0.04 0.05 0.02 0.03

Phenols 0.04 0.04 0.01 0.02

Overall 0.05 0.04 0.02 0.02

Table 5.3: Mean root mean square deviation (RMSD) and mean absolute deviation (MAD) forthe QCEIMS and CFM-EI methods for different classes of compounds, . Overall mean RMSDsand MADs are given in the final line.

also significantly worse than in the QCEIMS method (RRPs respectively 0.21 and 0.30 versus

0.11 and 0.20). The next worst grouping was among the benzenes, phenols and carboxylic acids,

again generally worse ranked than compared to the QCEIMS method. Based on the individual

class results, it is therefore not surprising that the overall ranking from CFM-EI is worse than

from the QCEIMS method (mean RRPs respectively 0.12 versus 0.09).

Dependence of the quality of the results on the mass of the molecules. There is a weak trend

that results are worse predicted as the mass of the molecule increases (r = 0.47) and this trend

is a bit stronger than for the QCEIMS method.

Comparison of QCEIMS and CFM-EI. Table 5.3 shows rather uniformly low RMSDs and

MADs for the predicted and observed spectra: in both cases the MADs were less than 3%.

However, based on the RMSDs the CFM-EI seems to produce a slightly more accurate i.e. the

spectrum shows fewer outlying peaks relative to experiment. Nevertheless, a qualitatively better

MSp does not translate into a a superior ranking of the MSp (see discussion on the RRPs).

5.4 Discussion

5.4.1 Detailed analysis of the predicted spectra

Alcohols. QCEIMS underestimates the abundance of most small peaks it predicts. As the

peak intensities are relative, this indicates that QCEIMS is systematically overestimating the

abundance of the primary fragmentation relative to the others.

103

5.4. DISCUSSION

On the other hand, CFM-EI seems to be underestimating the primary fragmentation with

respect to others.

Further, there exists a tendency for CFM-EI to overestimate the abundance of the molecular

ion. This over-abundance is reflected in the poor results in predicting both propan-2-ol and

butan-1-ol (see Figure 5.3) by CFM-EI.

It is noteworthy that both methods generally underestimate the loss of water (M− 18) from

primary alcohols, which is reflected in the MS of butan-1-ol with the low m/z = 56 peak.

Both methods predict the existence of few peaks not in the experimental spectra, (shown in

red in Fig 5.3). CFM-EI particularly in the cases of 2-methylpropan-1-ol and, to a lesser extent,

butan-1-ol predicts significant peaks non-existent in the experimental MSp.

Aldehydes and ketones. Importantly, QCEIMS clearly demonstrates the capacity to ‘discover’

McLafferty rearrangements. An illustrative example,the case of butanal, shows in the m/z = 44

(M− 28) peak, from loss of ethene via a McLafferty rearrangement. While the relative abundance

of this peak was underestimated by QCEIMS, its existence is vitally important. In the case of

2-methylpropanal QCEIMS seems to be missing or underestimating characteristic m/z = 27

and m/z = 41 peaks.

The predictions by CFM in the case of 2-methylpropanal is closer to experiment, but still

systematically underestimated. Again, as was the case in the alcohols, CFM-EI seems to

overestimate the abundance of the molecular ion in some cases (particularly 2-propanone).

Neither MS prediction technique displays any egregious false peaks in the aldehydes or

ketones, they simply tend to underestimate the relative occurrence of certain rearrangements or

cleavages.

In the case of the aldehydes, both methods seem to consistently fail to match experimental

spectra, and this is reflected in the rankings (2s/3s for both butanal and 2-methylpropanal).

Contrast this with the case of the two ketones, where the predictions are almost perfectly on

the mark. This might be associated with the decreased importance of α-cleavages in the case of

the aldehydes as compared to the ketones, strong conclusions require more studies.

Esters. It is evident for both techniques that McLafferty type rearrangements are absent or

underestimated for all relevant esters – particularly in methyl butanoate, where m/z = 74 is

virtually absent in the predicted spectrum. This is not to say that these methods are not capable

of predicting rearrangement fragmentations in-principle, as may be seen upon the examination

of the the prominent m/z = 88 in ethyl butanoate.

Amines QCEIMS underestimates or fails to predict some portions of the MSp, notably in

the 25 < m/z < 30 region of propan-2-amine and 2-methylpropan-2-amine.

Intriguingly, despite the QCEIMS MS for propan-2-amine overall having many missing peaks

and generally incorrect intensities (i.e. looking qualitatively worse) than that produced by

CFM-EI, the spectra proves to be a better match. In this instance, CFM clearly overestimates

(once again) the abundance of the molecular ion, which may have resulted in this poor match

105

CHAPTER 5. QUANTUM CHEMICAL ELECTRON IMPACT MASSSPECTRUM PREDICTION FOR DE NOVO STRUCTURE ELUCIDATION

against experiment. The cases of 2-methylpropan-1- and 2-methylpropan-2-amine proved more

difficult for both prediciton techniques, though the spectra were very close to experiment. Both

cases yield significantly different spectra, something reflected in both predictive techniques.

A significant error in both methods appears to be the overestimated abundance of loss of

a proton (M− 1), highlighted particularly in the m/z = 72 peak of the CFM-EI MSp for

2-methylpropan-2-amine, but also in the m/z = 72 and m/z = 58 peaks for the QCEIMS

spectra for butan-1-amine and propan-2-amine, respectively.

Alkenes. The case of alkenes, in particular those chosen here, requires precise predictions in

order to achieve the closest match with experiment. The differences between structural isomers

in their MSp tend to be independent of the double bond position (unless highly substituted or

conjugated)[38]. This, coupled with the tendencies of isomerization through migration of the

double bond in these low molecular weight, straight chain alkenes makes discrimination between

candidates a much more difficult task.

In general, both techniques overestimate the abundance of the molecular ion, further

exaggerated in the hexene series where the relative abundance of the molecular ion is much

lower than in the pent- and butenes. The regions 25 < m/z < 30 and 40 < m/z < 45, and

the m/z = 55 peak for pent- and hexene are the discriminating regions of these spectra. It is

notable that in these regions, both QCEIMS and CFM come close to the experimental spectra

in their predictions. The poor rankings (as seen in Table 5.2) over the alkene set are a contrast

with the relatively good error statistics (see Table 5.3). This is consistent with the rather subtle

differences in MSps where the only difference in the molecule is the double bond position. It

is not the case, then that the two techniques are failing to accurately predict the spectra of

alkenes, merely that their predications are not accurate enough for the purposes of matching

within sets of structural isomers.

Carboxylic Acids. From the results presented in Table 5.2, it would appear that performance

for the majority of carboxylic acids was far from satisfactory. Both techniques showed failures

in their spectral predictions, but it would seem that in this case QCEIMS outperforms CFM in

producing an EIMS matching more closely to experiment.

For longer chain carboxylic acids, QCEIMS seems to severely underestimate crucial McLaf-

ferty rearrangements, giving rise to inadequate prediction of mass spectra of these compounds.

An illustrative example is the virtual absence of m/z = 74, the base peak of 2-methylbutanoic

acid. These issues seem to be even more significant for CFM-EI with several carboxylic acids,

giving rise to very inaccurate spectra.

Benzenes. The only significant discrepancy from experiment for QCEIMS is shown in the

case of (1-methylethyl)benzene, where this method seems to make a charged fragment of the

isopropyl moiety, which does not occur in the real spectrum.

As many alkylbenzenes show similarities in the mass spectra, errors in approximations of

abundance of fragments can mean that the rank is relatively low although the correct fragments

106

CHAPTER 5. QUANTUM CHEMICAL ELECTRON IMPACT MASSSPECTRUM PREDICTION FOR DE NOVO STRUCTURE ELUCIDATION

represented in the search library (a fact which unfortunately cannot be established before the

structure is known).

We have also established that the counting statistics errors on the peaks produced in QCEIMS

are much smaller than the experimental reproducibility error in the mass spectrum; and that

the spectra are often reproduced better than the corresponding rankings relative to isomers.

This demonstrates that the QCEIMS method achieves its successful matches not because of an

accurate match in the peak heights, but rather that the peak heights and the distribution of the

peak positions as a function of mass-to-charge ratio is sufficiently different for different isomers

to achieve a high ranking.

As regards the performance of QCEIMS for different chemical classes, we find that the

alcohols have the best matching spectra, with a mean MAD of about 0.03. The normalized

spectra are therefore reproduced rather well. On the other hand, the worst ranked spectra were

the alkenes and substituted phenols with mean relative ranking positions above 0.18. It should

be noted, though, that even in these cases the correct candidate is placed in the top fifth of

isomers.

The problem with QCEIMS, recognized by the authors, is that it is a time consuming

calculation to perform, often taking days or weeks for a single spectrum; by comparison the

CFM-EI method can be performed in the browser. We do not think this is necessarily a problem

for those working to identify semiochemicals, where in a worst case situation it takes years to

identify the structure of an unknown: in this case it is more useful for the information to be

correct rather than timely. Still, the length of time required for QCEIMS means that it is best

used sparingly after other methods have been used to narrow the field of candidates.

One of the great difficulties with EIMS prediction is that it is not only about simple

fragmentations, but frequently involves rearrangements within the molecule.

Both methods tended to suffer from the same problems: underestimation or failure to predict

specific rearrangements (e.g. McLafferty rearragements), and overestimation of the molecular

ion. These failures are especially salient given the radically different manner in which both

techniques function. Perhaps this indicates that reassessment of some of the fundamental

abstractions (namely the notion that it is simply a matter of predicting the fragmentation)

involved need to be re-evaluated for general EIMS prediction.

There remains work to be done – ideally the identification protocol would start from an

observed MS and predict the molecular geometry, not the current prediction of MS from a

geometry. While numerous existing tools function quite well at predicting the molecular formula

from the observed spectra (see Scheubert et al.[1]), the ‘brute force’ approach of generating all

possible isomers is simply infeasible for even slightly large molecules, due to the exponential

relationship between M and the number of possible isomers. Clearly, the goal of the DENDRAL

project remains yet to be realized.

108

5.6. ACKNOWLEDGEMENTS

5.6 Acknowledgements

We gratefully acknowledge financial support from the Australian Research Council (grant

DE160101313) to BB, and the financial support of the Danish National Research Foundation

(Center for Materials Crystallography, DNRF-93) to PRS.

109

CHAPTER 5. QUANTUM CHEMICAL ELECTRON IMPACT MASSSPECTRUM PREDICTION FOR DE NOVO STRUCTURE ELUCIDATION

110

Bibliography

1. Scheubert, K., Hufsky, F. & Bocker, S. Computational mass spectrometry for small

molecules. J. Cheminf. 5, 12 (2013).

2. Mun, I. K. & Mclafferty, F. W. in Supercomputers in Chemistry 117 ().

3. Kerber, A., Meringer, M. & Rucker, C. CASE via MS: Ranking Structure Candidates by

Mass Spectra. Croat. Chem. Acta 79, 449 (2006).

4. Kind, T. & Fiehn, O. Advances in structure elucidation of small molecules using mass

spectrometry. Bioanal. Rev. 2, 23 (2010).

5. Bohman, B. et al. Discovery of pyrazines as pollinator sex pheromones and orchid

semiochemicals: implications for the evolution of sexual deception. New Phytol. 203, 939

(2014).

6. Bohman, B., Flematti, G. R. & Barrow, R. A. Identification of hydroxymethylpyrazines

using mass spectrometry. J. Mass Spectrom. 50, 987 (2015).

7. Bohman, B., Phillips, R. D., Flematti, G. R., Barrow, R. A. & Peakall, R. The chemical

web of a spider orchid - Sulfurous semiochemicals seduce male wasp pollinator. Angew.

Chem. Int. Ed. (2017).

8. Hertz, H. S., Hites, R. A. & Biemann, K. Identification of mass spectra by computer-

searching a file of known spectra (Anal. Chem., 1971).

9. Stein, S. E. Estimating probabilities of correct identification from results of mass spectral

library searches. J. Am. Soc. Mass Spectrom. 5, 316 (1994).

10. Stein, S. E. & Scott, D. R. Optimization and testing of mass spectral library search

algorithms for compound identification. J. Am. Soc. Mass Spectrom. 5, 859 (1994).

11. Stein, S. E. An integrated method for spectrum extraction and compound identification

from gas chromatography/mass spectrometry data. J. Am. Soc. Mass Spectrom. 10, 770

(1999).

12. Wan, K. X., Vidavsky, I. & Gross, M. L. Comparing similar spectra: from similarity index

to spectral contrast angle. J. Am. Soc. Mass Spectrom. 13, 85 (2002).

111

BIBLIOGRAPHY

13. McLafferty, F. W., Stauffer, D. A., Loh, S. Y. & Wesdemiotis, C. Unknown identification

using reference mass spectra. Quality evaluation of databases. J. Am. Soc. Mass Spectrom.

10, 1229 (1999).

14. Dobson, C. M. Chemical space and biology. Nature 432, 824 (2004).

15. Bauer, C. A. & Grimme, S. How to Compute Electron Ionization Mass Spectra from First

Principles. J. Phys. Chem. A 120, 3755 (2016).

16. Zupan, J. & Gasteiger, J. Neural networks: A new method for solving chemical problems

or just a passing phase? Anal. Chim. Acta 248, 1 (1991).

17. Jalali-Heravi, M. & Fatemi, M. H. Simulation of mass spectra of noncyclic alkanes and

alkenes using artificial neural network. Anal. Chim. Acta 415, 95 (2000).

18. Russell, S. J. & Norvig, P. Artificial Intelligence (Prentice Hall, 2010).

19. Buchanan, B. G., Lederberg, J. & Feigenbaum, E. The Heuristic DENDRAL program for

explaining empirical data 1971.

20. Gasteiger, J., Hanebeck, W. & Schulz, K. P. Prediction of mass spectra from structural

information. J. Chem. Inf. Model. 32, 264 (1992).

21. Allen, F., Greiner, R. & Wishart, D. Competitive fragmentation modeling of ESI-MS/MS

spectra for putative metabolite identification. Metabolomics (2015).

22. Allen, F., Pon, A., Greiner, R. & Wishart, D. Computational Prediction of Electron

Ionization Mass Spectra to Assist in GC/MS Compound Identification. Anal. Chem. 88,

7689 (2016).

23. N N Bogolubov, J. Introduction to Quantum Statistical Mechanics (World Scientific,

2010).

24. Grimme, S. Towards First Principles Calculation of Electron Impact Mass Spectra of

Molecules. Angew. Chem. Int. Ed. 52, 6306 (2013).

25. Grimme, S. & Bauer, C. A. First principles calculation of electron ionization mass spectra

for selected organic drug molecules. Org. Biomol. Chem. 12, 8737 (2014).

26. Valkenborg, D., Mertens, I., Lemiere, F., Witters, E. & Burzykowski, T. The isotopic

distribution conundrum. Mass Spectrom. Rev. 31, 96 (2012).

27. Mayer, I., Knapp-Mohammady, M. & Suhai, S. Bond orders and energy components in

polymers. Chem. Phys. Lett. 389, 34 (2004).

28. Bauer, C. A. & Grimme, S. Elucidation of Electron Ionization Induced Fragmentations of

Adenine by Semiempirical and Density Functional Molecular Dynamics. J. Phys. Chem.

A 118, 11479 (2014).

112

BIBLIOGRAPHY

29. Asgeirsson, V., Bauer, C. A. & Grimme, S. Quantum chemical calculation of electron

ionization mass spectra for general organic and inorganic molecules. Chem. Sci. 52, 6306

(2017).

30. Van der Walt, S., Colbert, S. C. & Varoquaux, G. The NumPy Array: A Structure for

Efficient Numerical Computation. Comput. Sci. Eng. 13, 22 (2011).

31. McKinney, W. Data Structures for Statistical Computing in Python in Proceedings of the

9th Python in Science Conference (eds van der Walt, S. ’. & Millman, J.) (2010), 51.

32. Hunter, J. D. Matplotlib: A 2D Graphics Environment. Comput. Sci. Eng. 9, 90 (2007).

33. Hanwell, M. D. et al. Avogadro: an advanced semantic chemical editor, visualization, and

analysis platform. J. Cheminf. 4, 17 (2012).

34. Wang, J., Wolf, R. M., Caldwell, J. W., Kollman, P. A. & Case, D. A. Development and

testing of a general amber force field. J. Comput. Chem. 25, 1157 (2004).

35. Weber, W. & Thiel, W. Orthogonalization corrections for semiempirical methods. Theor.

Chem. Acc. 103, 495 (2000).

36. Thiel, W. MNDO99, version 7.0 2004.

37. Rasmussen, G. T. & Isenhour, T. L. The evaluation of mass spectral search algorithms. J.

Chem. Inf. (1979).

38. McLafferty, F. W. & Turecek, F. Interpretation Of Mass Spectra (University Science

Books, 1993).

113

BIBLIOGRAPHY

114

Chapter 6

Basis set convergence of CCSD(T)

equilibrium geometries using a large

and diverse set of molecular structures

ABSTRACT

We examine the basis set convergence of the CCSD(T) method for obtaining the structures

of the 108 neutral first- and second-row species in the W4-11 database (with up to five non-

hydrogen atoms). This set includes a total of 181 unique bonds: 75 H−X, 49 X−Y, 43 X−−Y,

and 14 X−−−Y bonds (where X and Y are first- and second-row atoms). As reference values,

geometries optimized at the CCSD(T)/aug′-cc-pV(6+d)Z level of theory are used. We consider

the basis set convergence of the CCSD(T) method with the correlation consistent basis sets

cc-pV(n+d)Z and aug′-cc-pV(n+d)Z (n = D,T,Q, 5) and the Weigend-Ahlrichs def2-n ZVPP

basis sets (n = T,Q). For each increase in the highest angular momentum present in the

basis set, the root-mean-square deviation (RMSD) over the bond distances is decreased by

a factor of ∼4 For example, the following RMSDs are obtained for the cc-pV(n+d)Z basis

sets 0.0196 (D), 0.0050 (T), 0.0015 (Q), and 0.0004 (5) A. Similar results are obtained for the

aug′-cc-pV(n+d)Z and def2-n ZVPP basis sets. The double-zeta and triple-zeta quality basis

sets systematically and significantly overestimate the bond distances. A simple and cost-effective

way to improve the performance of these basis sets is to scale the bond distances by an empirical

scaling factor of 0.9865 (cc-pV(D+d)Z) and 0.9969 (cc-pV(T+d)Z). This results in RMSDs of

0.0080 (scaled cc-pV(D+d)Z) and 0.0029 (scaled cc-pV(T+d)Z) A. The basis set convergence of

larger basis sets can be accelerated via standard basis-set extrapolations. In addition, the basis

set convergence of explicitly correlated CCSD(T)-F12 calculations is investigated in conjunction

with the cc-pVnZ-F12 basis sets (n = D,T ). Typically, one “gains” two angular momenta in the

explicitly correlated calculations. That is, the CCSD(T)-F12/cc-pVnZ-F12 level of theory shows

similar performance to the CCSD(T)/cc-pV(n+2)Z level of theory. In particular, the following

115

CHAPTER 6. BASIS SET CONVERGENCE OF CCSD(T) EQUILIBRIUMGEOMETRIES USING A LARGE AND DIVERSE SET OF MOLECULARSTRUCTURES

RMSDs are obtained for the cc-pVnZ-F12 basis sets 0.0019 (D) and 0.0006 (T) A. Overall,

the CCSD(T)-F12/cc-pVDZ-F12 level of theory offers a stellar price-performance ratio and we

recommend using it when highly accurate reference geometries are needed (e.g., in composite ab

initio theories such as W4 and HEAT).

6.1 Introduction

Coupled-cluster theory is one of the most cost-effective methods of approximating the exact

solution for the nonrelativistic electronic Schrodinger equation.[1, 2] Coupled-cluster theory

entails a hierarchy of approximations that can be systematically improved towards the exact

quantum mechanical solution, providing a roadmap for the determination of highly accurate and

reliable chemical properties.[3–7] The CCSD(T) method (coupled-cluster with single, double,

and quasiperturbative triple excitations) has been found to be a cost-effective approach for

the calculation of highly accurate thermochemical and kinetic data,[8–16] as well as molecular

properties based on energy derivatives (e.g., equilibrium structures, vibrational frequencies,

and electrical properties).[4, 17–21] The CCSD(T) model is therefore often referred to as the

“gold standard in quantum chemistry.”1 It should be stressed, however, that this expression can

be misleading since in some cases the CCSD(T) shows poor performance (most notably, but

not limited to,[6, 22] multireference systems).[3–7] The basis set convergence of the CCSD(T)

method has been extensively studied for energetic properties.[3, 4, 6–8, 13, 15, 17, 18, 23–27]

There has been substantial work[6, 18, 19, 21, 28–37] exploring the potential accuracy of

CCSD(T) molecular structures relative to experimental reference values, typically including

other energetic contributions (e.g., post-CCSD(T), core-valence, and relativistic effects). Fewer

studies have been dedicated to the basis set convergence of molecular geometries. Studies of the

basis set effects on the molecular structures have been predominantly limited to small species

(molecules with at most two non-hydrogen atoms) and pathologically multireference systems

such as halogen oxides.[35, 36] Heckert et al.[38] have explored the basis-set convergence of

CCSD(T) equilibrium structures for a set of 17 small first-row molecules, namely HF, H2O, CH2

(1A1), NH3, CH4, CO, N2, F2, HCN, HNC, C2H2, CO2, OH, CN, NH2, CH2 (3B1), and NO,

relative to CCSD(T)-R12 reference values at the CBS limit. The basis set convergence of these

geometries was found to be smooth. The mean-absolute deviations (MADs) relative to the CBS

reference values were found to be 0.0008 (cc-pVQZ), 0.00033 (cc-pV5Z), and 0.00021 (cc-pV6Z)

A. They further reported that CCSD(T)/cc-pV5,6Z extrapolations are required for target

accuracies of 0.0001 A, and that extrapolations using small basis sets are not recommended.

Knizia et al.[37] considered the basis set convergence of explicitly correlated CCSD(T)-F12b

calculations for a set of 13 first-row diatomics relative to CCSD(T)/aug-cc-pV5,6Z reference

1To the best of our knowledge, this expression was first coined by T.H. Dunning, Jr. in a lecture series in thelate 1990s

116

6.2. COMPUTATIONAL METHODS

values. They found that the CCSD(T)-F12b/aug-cc-pVnZ level of theory generates results

comparable in quality to the CCSD(T)/aug-cc-pV(n+2)Z level of theory. More recently, Feller

et al.[39] explored a set of somewhat larger hydrocarbons (of up to C6H12) and C2, again

finding a gain in accuracy of about two “zetas” in the basis set when using explicitly correlated

CCSD(T)-F12b calculations relative to the conventional CCSD(T) calculations. In the present

work, we investigate the basis set convergence of the CCSD(T) method for the molecular

structures in the W4-11 database.[5] Excluding pathologically multireference systems (e.g., O3,

C2, and BN) for which the CCSD(T) approximation breaks down, the W4-11 database includes

122 species (results for the multireference systems are provided in Figures S1 and S2 of the

supplementary material). These species cover a broad spectrum of bonding situations with a

range of single and multiple bonds that involve varying degrees of covalent and ionic characters.

As such the W4-11 database constitutes an excellent benchmark set for analysis of basis set

effects on the molecular structures. For most of the systems in the W4-11 database, we were

able to obtain reference structures at the CCSD(T)/aug′-cc-pV(6+d)Z level of theory, whilst for

larger molecules (with low spatial symmetries), we use CCSD(T)/aug′-cc-pV(5+d)Z reference

geometries. Using this large and diverse set of accurate reference geometries, we attempt to

answer questions such as the following:

1. What is the accuracy of bond distances calculated with the CCSD(T) method in conjunc-

tion with the cc-pV(n+d)Z, aug′-cc-pV(n+d)Z, and def2-n ZVPP basis sets?

2. What is the accuracy of bond distances calculated with the CCSD(T)-F12 method in

conjunction with the cc-pV(n+d)Z-F12 basis sets?

3. Do different bond types (e.g., single, double, and triple bonds) exhibit different rates of

basis set convergence?

4. Can we accelerate the basis set convergence of CCSD(T) bond distances using basis set

extrapolations or even simple scaling factors?

5. To what extent does the accuracy of the reference geometries affect the molecular energies

calculated at the CCSD(T)/CBS level of theory?

6.2 Computational Methods

All calculations were carried out on the Linux cluster of the Karton group at the University of

Western Australia. All calculations were carried out using the MOLPRO program suite.[40, 41]

Most of the CCSD(T) geometry optimizations and single-point energy calculations were carried

out with the correlation-consistent basis sets of Dunning and co-workers.[42–44] The notation

A′VnZ indicates the combination of the standard correlation-consistent cc-pVnZ basis sets on

117

CHAPTER 6. BASIS SET CONVERGENCE OF CCSD(T) EQUILIBRIUMGEOMETRIES USING A LARGE AND DIVERSE SET OF MOLECULARSTRUCTURES

hydrogen,[42] the aug-cc-pVnZ basis sets on first-row elements,[43] and the aug-cc-pV(n+d)Z

basis sets on second-row elements.[44] The notation VnZ indicates the combination of the cc-

pVnZ basis sets on hydrogen and first-row elements and the cc-pV(n+d)Z basis sets on second-row

elements. Geometry optimizations were also carried out with the Weigend–Ahlrichs def2-TZVPP

and def2-QZVPP basis sets.[45] The explicitly correlated CCSD(T)-F12b calculations[37, 46]

were carried out in conjunction with the Vn Z-F12 basis sets of Peterson et al.[47]

6.3 Results

6.3.1 Overview of the molecules in the W4-11 database and refer-

ence geometries

The W4-11 database contains 122 small first- and second-row molecules (excluding the 16

pathologically multireference and 3 beryllium-containing species which are not considered in

the present work).2 Table 6.1 lists the molecules in this set, which will be referred to as the

W4-11-GEOM dataset. The systems in the W4-11-GEOM dataset include 85 closed shell, 21

radical, 9 singlet carbene, and 7 triplet species. In terms of elemental composition, the dataset

includes 88 first-row species (containing H and B—F), 17 second-row species (containing H

and Al—Cl), and 17 mixed first- and second-row species (containing H, B—F, and Al—Cl

atoms). Overall, the W4-11-GEOM dataset includes hydrogen-containing (82), hydrogen-free

(40), organic (63), and inorganic (59) compounds.

Table 6.2 gives an overview of the types of bonds in the W4-11-GEOM dataset. Overall, it

includes 246 symmetry unique bonds. Of these, 182 are single bonds, 49 are double bonds, and

15 are triple bonds. The set of 182 single bonds includes 117 H−X and 65 X−Y bonds (where

X and Y are non-hydrogen atoms from the first and second rows of the periodic table). For

the complete list of bonds in the W4-11-GEOM dataset, see Table S1 of the supplementary

material.

We were able to optimize the geometries for a subset of 108 molecules at the CCSD(T)/A′V6Z

level of theory (hereinafter referred to as the GEOM-AV6Z dataset). These include molecules

with up to five non-hydrogen atoms such as BF3, CF4, C2N2, C2F2, F2CO, F2O2, AlF3, SiF4,

SO3, Cl2O2, P4, S4, and AlCl3 (Table 6.1 lists the molecules in the GEOM-AV6Z subset). The

entire W4-11-GEOM database includes 14 CCSD(T)/A′V5Z geometries of organic molecules

in addition to the 108 CCSD(T)/A′V6Z geometries in the GEOM-AV6Z subset. In Section

2

The 14 highly multireference systems, for which the %TAE[(T)] diagnostic is in excess of 10%, are: Be2, B2,C2(1Σ+), BN(1Σ+), OF, F2O, FOO, FOOF, Cl2O, ClOO, OClO, O3, S3, and S4. Note that cis−HO3 andtrans−HO3 for which %TAE[(T)] = 7.4 and 7.9, respectively, are also included in this subset, see refs. [3] and[5] for further details. We also exclude two two additional beryllium systems, for which not all the consideredbasis sets are available, and they are BeF2 and BeCl2.

118

6.3. RESULTS

CCSD(T)/A′V6Z reference geometriesb

AlCl CH2 (3B1) HCl NH3

AlCl3 CH2C HCN NO

AlF CH2F2 HCNH NO2

AlF3 CH2NH HCNO O2

AlH CH3 HCO OCS

AlH3 CH3F HCOF OH

allene CH4 HF oxirane

B2H6 Cl2 HNC oxirene

BF ClCN HNCO P2

BF3 ClF HN3 P4

BH ClO HNO PH3

BH3 CN HOCl S2

BHF2 CO HOCN S2O

BN (3Π) CO2 HOF Si2H6

cis−HCOH CS HONC SiF

cis−HONO CS2 HOO SiF4

cis−N2H2 dioxirane HOOH SiH

C2H2 F2 HS SiH3F

C2H4 F2CO ketene SiH4

C2H6 C2F2 N2 SiO

CCH glyoxal N2H SO

CCl2 H2 N2H4 SO2

CF H2CN N2O SO3

CF2 H2CO NCCN S2H

CF4 H2O NH trans−HCOH

CH H2S NH2 trans−HONO

CH2 (1A1) HCCF NH2Cl trans−N2H2

CCSD(T)/A′V5Z reference geometries

acetaldehyde CH2CH ethanol propene

acetic acid CH2NH2 formic acid propyne

C2H3F CH3NH methanol

C2H5F CH3NH2 propane

a The entire set of 122 molecules will be referred to as the W4-11-GEOM dataset.

b The subset of 108 molecules for which we were able to obtainCCSD(T)/A′V6Z reference geometries will be referred to as theGEOM-AV6Z dataset.

Table 6.1: Overview of the 122 molecules in the W4-11-GEOM database for which we wereable to obtain CCSD(T)/A′V6Z or CCSD(T)/A′V5Z reference geometries.a

119

CHAPTER 6. BASIS SET CONVERGENCE OF CCSD(T) EQUILIBRIUMGEOMETRIES USING A LARGE AND DIVERSE SET OF MOLECULARSTRUCTURES

Bond type Numberb Comment

W4-11-GEOM dataset (122 molecules, 246 unique bonds)

H−X 117 X = H, B-F, Al-Cl

X−Y 65 X, Y = B-F, Al-Cl

X−−Y 49 X, Y = C, N, O, Si, S, Cl

X−−−Y 15 X, Y = B, C, N, P

GEOM-AV6Z dataset (108 molecules, 181 unique bonds)

H−X 75 X = H, B-F, Al-Cl

X−Y 49 X, Y = B-F, Al-Cl

X−−Y 43 X, Y = C, N, O, Si, S, Cl

X−−−Y 14 X, Y = B, C, N, P

a For the complete list of bonds see Table S1 of the SupportingInformation

b Number of unique bonds (i.e., not equivalent by symmetry, forexample CH3Cl has two unique bonds).

Table 6.2: Overview of the bonds in the W4-11-GEOM and GEOM-AV6Z datasets (foradditional details see Table S1 of the Supporting Information).a

6.3.2.1-4, we use the GEOM-AV6Z dataset to evaluate the performance of the CCSD(T)/A′VnZ

(n = D − 5), CCSD(T)/VnZ (n = D − 5), CCSD(T)/def2-n ZVPP (n = T,Q), and CCSD(T)-

F12/VnZ-F12 (n = D,T ) levels of theory. In Section 6.3.2.5, we use the larger W4-11-GEOM

dataset to evaluate the performance of the CCSD(T) and CCSD(T)-F12 methods in conjunction

with basis sets of up to quadruple-zeta quality.

6.3.2 Basis set convergence of bond distances

Basis set convergence of conventional CCSD(T) calculations against CCSD(T)/A′V6Z

reference geometries

Table 6.3 gives the error statistics for the CCSD(T)/A′VnZ (n = D − 5), CCSD(T)/VnZ

(n = D−5), CCSD(T)/def2-nZVPP (n = T,Q), and CCSD(T)-F12/Vn Z-F12 (n = D,T ) levels

of theory relative to the CCSD(T)/A′V6Z bond distances for the GEOM-AV6Z dataset (which

includes 108 molecules and 181 unique bonds). We begin by making three general observations:

• All the levels of theory tend to systematically overestimate the bond lengths as evident

from MSD ≈ MAD. The extent of overestimation decreases with the size of the basis

set.

• The VnZ basis sets show similar (or even slightly better) performance than the A′VnZ

120

6.3. RESULTS

Basis set RMSD MAD MSD LND LPD

VDZ 0.0196 0.0174 0.0174 N/A 0.0611 (O−−Cl)

VTZ 0.0050 0.0037 0.0037 –0.0026 (C−F) 0.0172 (Cl−Cl)

VQZ 0.0015 0.0010 0.0009 –0.0016 (C−F) 0.0060 (O−−Cl)

V5Z 0.0004 0.0003 0.0001 –0.0005 (C−F) 0.0017 (Cl−Cl)

A′VDZ 0.0222 0.0200 0.0200 N/A 0.0525 (Cl−Cl)

A′VTZ 0.0059 0.0048 0.0048 N/A 0.0185 (Cl−Cl)

A′VQZ 0.0017 0.0013 0.0013 N/A 0.0060 (Cl−Cl)

A′V5Z 0.0005 0.0003 0.0003 –0.0001 (H−Cl) 0.0018 (Cl−Cl)

VDZ-F12 0.0019 0.0015 0.0014 –0.0030 (Al−Cl) 0.0081 (N−O)

VTZ-F12 0.0006 0.0005 0.0005 –0.0006 (Al−Cl) 0.0026 (F−F)

Def2-TZVPP 0.0047 0.0034 0.0034 –0.0023 (H−Cl) 0.0181 (Cl−Cl)

Def2-QZVPP 0.0015 0.0010 0.0009 –0.0006 (N−H) 0.0056 (Cl−Cl)

a RMSD = root-mean-squared deviation, MAD = mean absolute deviation, MSD = meansigned deviation, LND = largest negative deviation, and LPD = largest positive deviation(the molecules associated with the LND and LPD are given in parenthesis).

Table 6.3: Overview of the basis set convergence of the CCSD(T) and CCSD(T)-F12 methodsfor the 181 unique bond lengths in the GEOM-AV6Z dataset (A). The reference values areCCSD(T)/A′V6Z bond distances.a

basis sets.

• The def2-nZVPP basis sets show similar performance to the VnZ basis sets.

• The CCSD(T)-F12/VnZ-F12 level of theory shows similar performance to the CCSD(T)/A′V(n + 2)Z

level of theory at a significantly reduced computational cost. This is in agreement with

previous studies[37, 39], albeit the current study includes a more diverse set of molecules

consisting of both first- and second-row elements.

Let us first consider the basis set convergence with the orbital VnZ and A′VnZ basis

sets in conventional CCSD(T) calculations. The VDZ basis set has been found to significantly

overestimate bond lengths in MP2 and CCSD(T) geometry optimizations due to the lack of higher

angular momentum polarization functions.[31, 34, 48–51] Our results confirm these observations

over a very large and diverse set of chemical bonds. In particular, the CCSD(T)/VDZ level of

theory systematically overestimates the CCSD(T)/A′V6Z bond lengths by significant amounts

and results in a mean-signed deviation (MSD) of +0.0174 Aand a root-mean-square deviation

(RMSD) of 0.0196 A. The VDZ basis set shows particularly poor performance for bonds involving

second-row atoms. For example, overestimations ranging between 0.04 and 0.06 A are seen for

the bonds involving second-row atoms in SiF4, NH2Cl, P4, SiH3F, Cl2, SiF, HOCl, ClF, and

ClO. This poor performance for the second-row systems is also reflected in the percentage errors.

121

CHAPTER 6. BASIS SET CONVERGENCE OF CCSD(T) EQUILIBRIUMGEOMETRIES USING A LARGE AND DIVERSE SET OF MOLECULARSTRUCTURES

For instance, the largest percentage errors are obtained for SiF (2.8), Cl−O bond in HOCl

(2.8), ClF (3.4), F2 (3.4), and ClO (3.9%). We note, however, that removing the 35 second-row

systems from the training set only results in a minor improvement in performance.3 Namely,

the RMSD for the 73 first-row systems with 138 unique bonds is 0.0167 A (see Table S2 of

the supplementary material for additional error statistics for the subset of first-row systems).

Notably, the addition of diffuse functions does not bring succor and the A′VDZ basis set results

in similar performance to the VDZ basis set (Table 6.3).

Increasing the basis set size from a double-zeta to triple-zeta reduces the RMSD by a factor

of ∼4 The VTZ basis sets result in RMSDs of 0.0050 A (cf., an RMSD of 0.02 A for the VDZ

basis set). The VTZ basis set still tends to systematically overestimate the bond lengths, where

particularly large deviations of 0.015–0.017 A are obtained for the bonds involving second-row

atoms in S2O, ClO, S2H, and Cl2. We note that upon removing the second-row systems from

the training set, the RMSD is reduced to 0.0036 A (Table S2 of the supplementary material).

Similar to the performance of the VDZ and A′VDZ basis sets, the addition of diffuse functions

does not lead to an improvement in performance, namely, the CCSD(T)/A′VTZ level of theory

results in an RMSD and MAD of 0.0059 and 0.0048 A, respectively, for the entire GEOM-AV6Z

dataset (Table 6.3). The CCSD(T)/VQZ level of theory is used for optimizing geometries in

highly accurate composite theories such as the W4[27, 52, 53] and HEAT[54–56] thermochemical

protocols.[3] Here the performance of this level of theory is assessed against a large and diverse

set of geometries. Relative to our CCSD(T)/A′V6Z reference values, the CCSD(T)/VQZ level of

theory results in a respectable RMSD of 0.0015 A and a MAD of 0.0010 A. Again, the similarity

of the MSD of 0.0009 A to the MAD indicates that the bond lengths are still systematically

overestimated. The largest overestimations of 0.004–0.006 A are obtained for the bonds involving

second-row atoms in CCl2, P4, S2, NH2Cl, HOCl, S2O, S2H, Cl2, and ClO. Removing the 35

second-row systems from the training set has a noticeable effect on the performance, namely,

the error statistics are reduced to RMSD = 0.0009, MAD = 0.0007,and MSD = 0.0005 A

(Table S2 of the supplementary material). Further, the largest overestimations are reduced to

below 0.003 A, namely, 0.0029 (O−O bond in dioxirane) and 0.0027 (F2) A.

The V5Z and A′V5Z basis sets both yield bond distances that are very close to the

CCSD(T)/A′V6Z bond lengths. The V5Z basis set results in an RMSD of 0.0004 A and

a MAD of 0.0003 A. An MSD of 0.0001 indicates that there is a very small bias towards

overestimating the bond distances. The largest deviation, an overestimation of 0.0017 A, is

obtained for Cl2. Upon eliminating the second-row systems, the RMSD and MAD are reduced

to 0.0003 and 0.0002 A, respectively.

Lastly, we note that the def2-TZVPP and def2-QZVPP basis sets result in overall error

statistics that are very similar to those obtained for the VTZ and VQZ basis sets, respectively

3The 35 systems containing second-row elements are AlH, AlH3, AlF, AlF3, AlCl, AlCl3, SiH, SiH4, Si2H6,SiO, SiF, SiH3F, SiF4, PH3, P2, P4, HS, H2S, CS, CS2, SO, SO2, SO3, OCS, S2, S2H, S2O, HCl, CCl2, ClCN,NH2Cl, ClO, HOCl, ClF, and Cl2.

122

6.3. RESULTS

(Table 6.3).

Basis set convergence of explicitly correlated CCSD(T)-F12 calculations

It is well established that inclusion of ”geminal” terms that explicitly depend on the interelec-

tronic distance into the wavefunction drastically accelerates the basis set convergence.[57–59]

Experience with CCSD-F12 energy calculations has shown that typically the gain amounts to 1–2

angular momenta.[47, 57–60] Table 6.3 gives an overview of the performance of the CCSD(T)-

F12/Vn Z-F12 level of theory (n = D,T ) for the bond distances in the GEOM-AV6Z dataset.

The CCSD(T)-F12/VDZ-F12 level of theory results in RMSD = 0.0019, MAD = 0.0015, and

MSD = 0.0014A. This performance is significantly better than that of the CCSD(T)/VTZ

level of theory (RMSD = 0.0037) and is comparable to that of the CCSD(T)/VQZ level of

theory (RMSD = 0.0015 A, Table 6.3). The CCSD(T)-F12/VDZ-F12 level of theory tends to

systematically overestimate the bond distances. However, whilst the largest overestimations

for the CCSD(T)/VTZ and CCSD(T)/VQZ levels of theory are obtained for bond distances

involving second-row elements, the largest deviations for the CCSD(T)-F12/VDZ-F12 level of

theory are obtained for first-row systems such as F2 (0.007), the N−O bond in trans-HONO

(0.008), and the O−O bond in dioxirane (0.008 A). Consequently the error statistics for the

CCSD(T)-F12/VDZ-F12 level of theory for the 73 first-row systems are very similar to those

obtained for the entire GEOM-AV6Z dataset (Table S2 of the supplementary material).

We note that the CCSD(T)-F12/VDZ-F12 level of theory represents a significant saving

in computational resources compared to the CCSD(T)/VQZ level of theory. Table 6.4 gives

relative central processing unit (CPU) times and disk space used by these levels of theory for two

medium-sized systems (naphthalene and anthracene). For anthracene (C14H10), a single-point

energy CCSD(T)-F12/VDZ-F12 calculation which entails 510 basis functions in D2h symmetry

requires only 10% of the time required for the CCSD(T)/VQZ calculation which involves more

than twice the number of basis functions (1070 basis functions). In terms of the disk usage,

the CCSD(T)-F12/VDZ-F12 calculation uses 25 GB of scratch disk, whilst the CCSD(T)/VQZ

calculation uses a total of 335 GB of disk space. Considering these very significant savings in

terms of CPU time and disk usage for obtaining molecular geometries of similar quality, we

recommend using the CCSD(T)-F12/VDZ-F12 level of theory rather than CCSD(T)/VQZ when

high-quality geometries are needed (e.g., in thermochemical protocols such as W4lite, W4, and

W4-F12).[52, 53] Finally, we note that the CCSD(T)-F12/VDZ-F12 calculations have a similar

computational cost to that of CCSD(T)/VTZ calculations (Table 6.4); however, the former level

of theory leads to much more accurate bond distances (Table 6.3).

The CCSD(T)-F12/VTZ-F12 level of theory results in an RMSD of 0.0006 A, which is

comparable to that of the CCSD(T)/V5Z level of theory (RMSD = 0.0004 A, Table 6.3). Similar

to the VDZ-F12 basis set, the largest overestimations for the VTZ-F12 basis set are obtained

for bonds involving first-row atoms, such as F2 (0.003) and the O−O bond in dioxirane (0.002

123

CHAPTER 6. BASIS SET CONVERGENCE OF CCSD(T) EQUILIBRIUMGEOMETRIES USING A LARGE AND DIVERSE SET OF MOLECULARSTRUCTURES

Naphthalene Anthracene

Basis Set Timeb Diskc Timeb Diskc

VDZ 1 0.6 1 2

VTZ 25 10 22 31

VQZ 253 111 204 335

V5Z 1977 906 N/A N/A

VDZ-F12 31 7 20 25

VTZ-F12 191 68 138 218

a All the calculations ran on 8 cores of otherwiseidle dual Intel Xeon E5-2670v2 systems (3.1GHz, 256 GB RAM).

b The numbers given are the ratios between theCPU time for the different levels of theory andthat for the CCSD(T)/VDZ level of theory

c Scratch disk usage in GB.

Table 6.4: Computational resources used for the CCSD(T)/VnZ and CCSD(T)-F12/VnZ-F12single-point energy calculations for naphthalene and anthracene.a

A). Thus, it seems like the CCSD(T)-F12 method shows a more balanced performance for first-

and second-row systems than conventional CCSD(T) calculations in conjunction with double-

and triple-zeta basis sets. Table 6.4 illustrates the significant computational savings of the

CCSD(T)-F12/VTZ-F12 level of theory compared to the CCSD(T)/V5Z level of theory. For a

medium-sized system such as naphthalene, the CCSD(T)- F12/VTZ-F12 calculation uses about

10% of the CPU time and disk space as the CCSD(T)/V5Z calculation.

Accelerating the basis set convergence of CCSD(T) and CCSD(T)-F12 calculations

In Section 6.3.2.1, we have seen that the VDZ and A′VDZ basis sets systematically and severely

overestimate the bond lengths. For all the 181 bonds in the GEOM-AV6Z database, these

basis sets overestimate the CCSD(T)/ A′V6Z bond distances by amounts ranging from 0.0017

(C—F bond in CH3F) to 0.0611 (Cl=O) A. For both the VDZ and A′VDZ basis sets, we find

that there is a high statistical correlation with the CCSD(T)/A′V6Z bond distances. Namely,

the squared correlation coefficient (R2) is equal to 0.9991 (VDZ) and 0.9992 ( A′VDZ). For

the larger basis sets, we obtain R2 values > 0.99991, for additional details see Table S3 of the

supplementary material. The high statistical correlation between the bond distances obtained

at the CCSD(T)/A′V6Z and CCSD(T)/VDZ levels of theory is illustrated in Figure 6.1 and

suggests that simple linear scaling of the bond distances may improve the accuracy.

Table 6.5 gives an overview of the performance of the scaled CCSD(T)/VnZ and CCSD(T)/A′VnZ

bond distances in conjunction with basis sets of up to quadruple- zeta quality. Let us begin with

the performance of the double-zeta quality basis sets in conventional CCSD(T) calculations. The

124

6.3. RESULTS

Figure 6.1: Linear correlation between CCSD(T)/A′V6Z and CCSD(T)/VDZ bond distances(in A) for the 181 bond lengths in the GEOM-AV6Z dataset.

CCSD(T)/VDZ and CCSD(T)/A′VDZ levels of theory result in a large RMSD of ∼0.02 A(Table

6.3). Upon scaling of the bond distances by empirical factors of 0.9865 (VDZ) and 0.9842

(A′VDZ), the RMSDs are reduced by about 60% to 0.0080 (scaled VDZ) and 0.0076 (scaled

A′VDZ) A.

Moving to the triple-zeta quality basis sets, the RMSD for the unscaled bond lengths is

0.0050 (VTZ) and 0.0059 (A′VTZ) A (Table 6.3). These relatively large RMSDs are reduced

by about 50% upon scaling. Namely, they are 0.0029 (scaled VTZ) and 0.0027 (scaled A′VTZ)

A. Again a near-zero MSD of −0.0003 Aobtained for both basis sets indicates that scaling

eliminates the systematic bias towards overestimating the bond lengths. For comparison, the

unscaled VTZ and A′VTZ results lead to MSDs which are more than one order of magnitude

larger (namely, 0.004 and 0.005 A, respectively).

The unscaled CCSD(T)/VQZ and CCSD(T)/A′VQZ methods lead to respectable RMSDs

of 0.0015 and 0.0017 A, respectively. Scaling of the VQZ bond distances leads to a small

improvement in performance (RMSD = 0.0011 A); however, scaling the A′VQZ bond distances

reduces the RMSD by nearly 50% (RMSD = 0.0009 A). Finally, we note that scaling the

def2-TZVPP and def2-QZVPP bond distances results in overall error statistics that are very

similar to those obtained for the scaled VTZ and VQZ basis sets (Table 6.5).

What about extrapolating the bond lengths? This approach has been previously found to

accelerate the basis set convergence in conjunction with sufficiently large basis sets.[6, 28, 29, 32,

34, 38] Here we test this approach for a wider and more diverse set of bond lengths. We consider

a two-point extrapolation formula of the form D(L) = D inf +A/Lα where D is the bond

distance, L is the highest angular momentum present in the basis set, and α is an extrapolation

125

CHAPTER 6. BASIS SET CONVERGENCE OF CCSD(T) EQUILIBRIUMGEOMETRIES USING A LARGE AND DIVERSE SET OF MOLECULARSTRUCTURES

Basis sets αb RMSD MAD MSD LND LPD

Scaled VDZ 0.9865 0.0080 0.0056 –0.0001 –0.0197 (H−Al) 0.0390 (O−−Cl)

VTZ 0.9969 0.0029 0.0022 –0.0003 –0.0069 (C−F) 0.0109 (Cl−Cl)

VQZ 0.9992 0.0011 0.0008 –0.0001 –0.0027 (C−F) 0.0048 (O−−Cl)

A′VDZ 0.9842 0.0076 0.0055 –0.0004 –0.0228 (H−Al) 0.0256 (O−−Cl)

A′VTZ 0.9960 0.0027 0.0020 –0.0003 –0.0050 (H−S) 0.0105 (Cl−Cl)

A′VQZ 0.9989 0.0009 0.0007 –0.0001 –0.0016 (H−Al) 0.0038 (Cl−Cl)

VDZ-F12 0.9989 0.0013 0.0008 0.0000 –0.0053 (Al−Cl) 0.0065 (N−O)

Def2-TZVPP 0.9971 0.0028 0.0020 –0.0003 –0.0060 (H−Cl) 0.0123 (Cl−Cl)

Def2-QZVPP 0.9992 0.0010 0.0007 –0.0001 –0.0015 (H−N) 0.0040 (O−−Cl)

Extrap.c V{D,T}Z 4.0000 0.0033 0.0025 0.0003 –0.0060 (F−Si) 0.0120 (S−S)

V{T,Q}Z 4.5000 0.0007 0.0005 –0.0002 –0.0014 (H−F) 0.0024 (O−−Cl)

A′V{D,T}Z 3.5000 0.0026 0.0020 0.0001 –0.0046 (H−H) 0.0091 (S−S)

A′V{T,Q}Z 4.4000 0.0004 0.0003 –0.0001 –0.0008 (H−F) 0.0015 (O−−Cl)

Def2-{T,Q} 4.3000 0.0006 0.0004 –0.0001 –0.0015 (B−F) 0.0022 (H−C)

a Footnote a of Table 6.3 applies here.b Scaling factors and extrapolation exponents used in the two-point extrapolations, these are optimized to minimize

the RMSD over the set of 181 bond lengths in the GEOM-AV6Z dataset.c Extrapolated using the DL = D∞ + A/Lα extrapolation formula (where D is the bond distance and L is the

highest angular momentum represented in the basis set).

Table 6.5: Overview of the performance of scaled and extrapolated CCSD(T) and CCSD(T)-F12 bond distances for the 181 unique bonds in the GEOM-AV6Z dataset (A). The referencevalues are CCSD(T)/A′V6Z bond distances.a

126

6.3. RESULTS

Basis set H−X X−Y X−−Y X−−−Y

VDZ 0.0148 0.0245 0.0199 0.0215

VTZ 0.0014 0.0071 0.0058 0.0057

VQZ 0.0004 0.0022 0.0017 0.0013

V5Z 0.0002 0.0006 0.0005 0.0003

A′VDZ 0.0137 0.0304 0.0229 0.0228

A′VTZ 0.0022 0.0086 0.0064 0.0058

A′VQZ 0.0005 0.0026 0.0020 0.0016

A′V5Z 0.0001 0.0007 0.0005 0.0004

VDZ-F12 0.0007 0.0029 0.0020 0.0019

VTZ-F12 0.0004 0.0009 0.0006 0.0005

a For a comprehensive overview of the error statis-tics obtained for the H−X, X−Y, X−−Y, and X−−−Ysubsets see Table S4 of the Supporting Information

Table 6.6: Overview of the basis set convergence of the CCSD(T) and CCSD(T)-F12 methodsfor the 75 H−X, 49 X−Y, 43 X−−Y, and 14 X−−−Y bonds in the GEOM-AV6Z dataset (RMSDs,in A)a

exponent which is optimized to minimize the RMSD over the GEOM-AV6Z dataset. These

results are presented in Table 6.5. Extrapolating from the VD,TZ or A′V{D,T}Z basis set pairs

results in RMSDs of 0.0033 and 0.0026 A, respectively. This does not represent an improvement

over simple scaling (Table 6.5) and we do not recommend using basis set extrapolations that

involve double-zeta quality basis sets. Extrapolating from the VT,QZ or A′V{T,Q}Z basis

set pairs results in RMSDs of 0.0007 and 0.0004 A, respectively. This performance represents

an improvement over simple linear scaling of the VQZ (RMSD = 0.0011) and A′VQZ results

(RMSD = 0.0009 A). Thus, extrapolations from triple-zeta and quadruple- zeta quality basis sets

can be considered as a viable alternative to simple scaling. Extrapolations from the def2-TZVPP

and def2-QZVPP basis sets result in overall error statistics that are similar to those obtained

from the VT,QZ extrapolations (Table 6.5).

Basis set convergence for subsets of the GEOM-AV6Z dataset

In this subsection, we examine more closely the basis set convergence for the H−X, X−Y,

X−−Y, and X−−−Y bonds in the GEOM-AV6Z dataset (Table 6.2). Table 6.6 gathers the RMSDs

obtained for the H—X, X—Y, X−−Y, and X−−−Y subsets. For a comprehensive overview of the

error statistics, see Table S4 of the supplementary material.

A few interesting features emerge from Table 6.6. First, the performance of all the considered

basis sets is significantly better for the H—X bonds compared to the performance for the X−Y,

X−−Y, and X−−−Y bonds. This is not surprising since bonds involving hydrogen are expected

127

CHAPTER 6. BASIS SET CONVERGENCE OF CCSD(T) EQUILIBRIUMGEOMETRIES USING A LARGE AND DIVERSE SET OF MOLECULARSTRUCTURES

to converge faster to the basis set limit than the X−Y, X−−Y, and X−−−Y bonds. Second, all

the considered basis sets show similar performance for the X−−Y and X−−−Y bonds. This result

is somewhat counterintuitive, particularly for the double- and triple-zeta basis sets, and is

attributed to the fact that the X−−Y subset includes many bonds involving second- row elements,

whilst the X−−−Y subset includes mainly first-row elements. Table S5 of the supplementary

material gives the RMSDs over the bonds involving only first-row elements. After removing

the bonds involving second-row elements from both subsets, we obtain RMSDs of 0.0158 (X−−Y

bonds) and 0.0211 (X−−−Y) for the VDZ basis set, and 0.0045 (X−−Y) and 0.0050 (X−−−Y) for the

VTZ basis set. Thus, it appears that the triple bonds converge more slowly to the basis set

limit compared to the double bonds. It should be noted that for basis sets of quadruple-zeta

quality (and higher), the RMSDs for the X−−Y and X−−−Y bonds are practically the same (see

Table S5 of the supplementary material). Similar trends are observed for the A′VnZ basis sets.

The subset of 49 X—Y bonds seems to be particularly challenging for the conventional

CCSD(T) calculations involving double- and triple-zeta quality basis sets. The VDZ and A′VDZ

basis sets result in large RMSDs of 0.025 and 0.030 A, respectively. The VTZ and A′VTZ basis

sets result in RMSDs larger than 0.007 A. The VQZ, A′VQZ, and cost-effective VDZ-F12 basis

sets result in RMSDs of 0.002–0.003 A. The V5Z and A′V5Z basis sets result in small RMSDs

of about 0.0006 A (Table 6.6). These RMSDs are reduced after removing the bonds involving

second-row elements (see Table S5 of the supplementary material).

Basis set convergence for the entire W4-11-GEOM dataset

The GEOM-AV6Z dataset contains systems with up to 5 non-hydrogen atoms. The largest

systems in this database are highly symmetric, mostly inorganic systems, such as BF3, CF4,

C2N2, C2F2, F2CO, F2O2, AlF3, SiF4, SO3, Cl2O2, P4, S4, and AlCl3. The largest organic

systems that are present in the GEOM-AV6Z dataset are allene, ketene, glyoxal, oxirene, oxirane,

and dioxirane. It is therefore of interest to include larger organic systems in the dataset. For

this purpose, we generate an additional test set (W4-11-GEOM) with the remaining systems

from the W4-11 dataset (Table 6.1). The additional 14 systems are relatively large organic

systems involving first-row elements (namely, acetic acid, acetaldehyde, C2H3F, C2H5F, CH2CH,

CH2NH2, CH3NH, CH3NH2, ethanol, formic acid, methanol, propane, propene, and propyne).

We were able to optimize the geometries of these systems at the CCSD(T)/A′V5Z level of theory.

As we saw in Section 6.3.2.1, the performance of the A′V5Z basis set for bond involving only

first-row elements is very close to that of the A′V6Z basis set (RMSD = 0.0003 A, Table S2

of the supplementary material). Due to the use of 14 CCSD(T)/A′V5Z reference geometries,

we will only assess the performance of basis sets of up to quadruple-zeta quality against this

dataset. Overall, the W4-11-GEOM datasets contain 122 molecules involving 246 unique bond

distances. Table 6.7 gathers the error statistics for the W4-11-GEOM dataset.

Generally, the error statistics obtained for the W4-11-GEOM database (Table 6.7) are

128

6.3. RESULTS

Basis sets RMSD MAD MSD LND LPD

VDZ 0.0184 0.0165 0.0165 N/A 0.0611 (O−−Cl)

VTZ 0.0045 0.0033 0.0032 –0.0029 (C−F) 0.0172 (Cl−Cl)

VQZ 0.0013 0.0008 0.0007 –0.0018 (C−F) 0.006 (O−−Cl)

A′VDZ 0.0207 0.0188 0.0188 N/A 0.0525 (Cl−Cl)

A′VTZ 0.0053 0.0043 0.0043 N/A 0.0185 (Cl−Cl)

A′VQZ 0.0015 0.0011 0.0011 N/A 0.006 (Cl−Cl)

VDZ-F12 0.0017 0.0013 0.0013 –0.003 (Al−Cl) 0.0081 (N−O)

Scaledb VDZ 0.0075 0.0051 –0.0006 –0.0197 (H−Al) 0.039 (O−−Cl)

VTZ 0.0027 0.0022 –0.0007 –0.0073 (C−F) 0.0109 (Cl−Cl)

VQZ 0.0011 0.0008 –0.0003 –0.0029 (C−F) 0.0048 (O−−Cl)

A′VDZ 0.0070 0.0051 –0.0012 –0.0228 (H−Al) 0.0256 (O−−Cl)

A′VTZ 0.0026 0.0020 –0.0007 –0.005 (H−S) 0.0105 (Cl−Cl)

A′VQZ 0.0009 0.0007 –0.0003 –0.0016 (H−Al) 0.0038 (Cl−Cl)

VDZ-F12 0.0012 0.0007 –0.0001 –0.0053 (Al−Cl) 0.0065 (N−O)

a Footnote a to Table 6.3 applies hereb The scaling factors are given in Table 6.5

Table 6.7: Overview of the performance of the CCSD(T) and CCSD(T)-F12 methods for the 246unique bonds in the W4-11-GEOM dataset (A). The reference values are 108 CCSD(T)/A′V6Zand 14 CCSD(T)/A′V5Z bond distances (See Table 6.1).a

129

CHAPTER 6. BASIS SET CONVERGENCE OF CCSD(T) EQUILIBRIUMGEOMETRIES USING A LARGE AND DIVERSE SET OF MOLECULARSTRUCTURES

similar to those obtained for the GEOM-AV6Z dataset ( Tables 6.3 and 6.5), and thus, our

main conclusions in Sections 6.3.2.1-4 remain largely unchanged. We note that inclusion of

the 14 organic systems improves the performance for all the considered levels of theory. In

particular, the RMSDs for the entire W4-11-GEOM database are ∼0.02 (VDZ and A′VDZ),

∼0.005 (VTZ and A′VTZ), and ∼0.001 (VQZ and A′VQZ) A. For the scaled bond distances, we

obtain RMSDs of 0.008 (VDZ and A′VDZ), 0.003 (VTZ and A′VTZ), and ∼0.001 (VQZ and

A′VQZ) A. Thus, again we see that scaling (in particular of the VDZ and A′VDZ distances)

results in significant improvements in performance at no additional computational cost.

Energetic consequences of the level of theory used for optimizing the geometries

In Section 6.3.2, we have shown that the basis set incompleteness error can result in RMS

deviations from CCSD(T)/A′V6Z bond distances ranging from ∼0.02 (VDZ and A′VDZ) to

∼0.002 (VQZ and A′VQZ), to ∼0.0005 (V5Z and A′V5Z) A. We note that errors in the bond

distances (whether they are overestimations or underestimations) will always lead to overesti-

mation of the molecular energies (or underestimation of the atomization energies) relative to

those obtained at the CCSD(T)/CBS equilibrium geometries. However, a number of important

questions arise as follows: (i) by how much an RMSD of say 0.02, 0.002, or 0.0005 A in the

bond distances affects the CCSD(T)/CBS molecular energies? (ii) Is the geometry effect going

to increase with the molecular size (e.g., when going from diatomics, to triatomics, and to tetra-

atomics)? These questions are particularly relevant when deciding which reference geometry to

use in highly accurate composite ab initio methods such as Wn,[8, 27, 52, 53] HEAT,[54–56]

and Feller-Peterson-Dixon (FPD)[4, 6, 17, 61] theories.[3] Here we will address these questions

in the context of a large and diverse set of molecules.

We calculate the molecular energies at the CCSD(T)/CBS level of theory for the 108

molecules in the GEOM-AV6Z database using the following reference geometries: CCSD(T)/VnZ

(n = D − 5); CCSD(T)/A′VnZ (n = D − 5); and CCSD(T)-F12/VnZ-F12 (n = D,T ). The

single-point CCSD(T)/CBS energies are calculated using W2-F12 theory.[8] An overview of

these results is given in Table 6.8, whilst individual errors for all the molecules are given in

Table S6 of the supplementary material.

Calculating the W2-F12 energies using CCSD(T)/VDZ and CCSD(T)/A′VDZ reference

geometries leads to very large deviations from W2-F12 energies calculated using CCSD(T)/A′V6Z

geometries. In particular, the RMSDs are 2.4 (VDZ) and 3.0 (A′VDZ) kJ mol−1, and the largest

deviations exceed 10 kJ mol−1. Such large errors far exceed the intrinsic accuracy of highly

accurate composite ab initio methods,[3] and CCSD(T)/VDZ and CCSD(T)/A′VDZ geometries

should not be used for these purposes.

Moving to the triple-zeta quality basis sets, we obtain an RMSD of ∼0.2 kJ mol−1for the

CCSD(T)/VTZ and CCSD(T)/A′VTZ geometries. These geometries should not be used in

composite ab initio methods that attempt to approximate the full configuration interaction

130

6.3. RESULTS

Basis sets RMSD MAD MSD LPD

VDZ 2.380 1.900 1.900 10.39 (SiF4)

VTZ 0.170 0.130 0.130 0.62 (P4)

VQZ 0.020 0.010 0.010 0.07 (SO3)

V5Z 0.002 0.001 0.001 0.008 (SO3)

A′VDZ 3.010 2.420 2.420 10.42 (SO3)

A′VTZ 0.220 0.170 0.170 0.92 (SO3)

A′VQZ 0.030 0.020 0.020 0.13 (SO3)

A′V5Z 0.003 0.002 0.002 0.017 (SO3)

VDZ-F12 0.030 0.020 0.020 0.10 (CF4)

VTZ-F12 0.004 0.003 0.003 0.015 (SO3)

a Footnote a to Table 6.3 applies here.b The reference values are non-relativistic, valence

CCSD(T)/CBS values from W2-F12 theory calculatedwith CCSD(T)/A′V6Z reference geometries.

Table 6.8: Effect of CCSD(T) and CCSD(T)-F12 reference geometry on molecular energiescalculated at the CCSD(T)/CBS level of theory for the 108 molecules in the GEOM-AV6Zdatabase (kJ mol−1).a,b

(FCI) CBS energy (e.g., HEAT, W4, and W4-F12), since these theories are capable of predicting

atomization energies with 95% (2σ) confidence intervals narrower (or even significantly narrower)

than 1 kJ mol−1(see Table 2 of ref. [3] for more details). On the other hand, for composite ab

initio methods that approximate the CCSD(T)/CBS energy (e.g., Wn and Wn-F12, n = 1, 2)

which are capable of 95% confidence intervals narrower than ∼1 kcal mol−1, one can consider

using VTZ and A′VTZ geometries. Nevertheless, it should be stressed that for a similar

computational cost, one can obtain CCSD(T)-F12/VDZ-F12 geometries which lead to much

better performance (vide infra).

The CCSD(T)/VQZ level of theory, which is used for optimizing geometries in the W4 and

HEAT thermochemical protocols, leads to a near-zero RMSD of 0.02 kJ mol−1. This confirms

that this level of theory is adequate for optimizing the geometries in these highly accurate

composite theories. The CCSD(T)/A′VQZ level of theory leads to a similar RMSD of 0.03

kJ mol−1. In both cases the largest deviations (of ∼0.1 kJ mol−1) are obtained for SO3.

An important finding is that the CCSD(T)-F12/VDZ-F12 geometries lead to a similar

performance to that of the CCSD(T)/VQZ and CCSD(T)/A′VQZ geometries, at a significantly

reduced computational cost (Table 6.4). In particular, the CCSD(T)-F12/VDZ-F12 geometries

result in an RMSD of 0.03 kJ mol−1and a largest deviation of 0.1 kJ mol−1(CF4). Thus, we

recommend using this economical level of theory for optimizing geometries in composite theories

such as HEAT and W4.

131

CHAPTER 6. BASIS SET CONVERGENCE OF CCSD(T) EQUILIBRIUMGEOMETRIES USING A LARGE AND DIVERSE SET OF MOLECULARSTRUCTURES

Finally, we note that quintuple-zeta quality basis sets lead to RMSD of below 0.003

kJ mol−1and largest deviations below 0.02 kJ mol−1. What about the dependence of the geome-

try effect on the size of the molecule? Inspection of Table S6 (supplementary material) reveals

that, in general, there is an increase in the geometry effect with the number of non-hydrogen

atoms in the system. For example, for the CCSD(T)/VDZ geometries, we obtain RMSDs of

0.9 (over the 21 systems with one non-hydrogen atom), 2.1 (over the 48 systems with two

non-hydrogen atoms), 2.5 (over the 27 systems with three non-hydrogen atoms), and 3.3 (over

the 9 systems with four non-hydrogen atoms). Similarly, for the CCSD(T)/VTZ geometries, we

obtain RMSDs of 0.03 (one non- hydrogen atom), 0.1 (two non-hydrogen atoms), 0.2 (three

non-hydrogen atoms), and 0.3 (four non-hydrogen atoms).

6.4 Conclusions

We have optimized reference geometries for a diverse set of 108 molecules at the CCSD(T)/A′V6Z

level of theory. This set includes inorganic species with up to five non-hydrogen atoms (e.g., BF3,

F2O2, AlF3, SiF4, SO3, Cl2O2, P4, S4, and AlCl3) as well as organic compounds of similar size

(e.g., allene, ketene, glyoxal, oxirene, oxirane, dioxirane, CF4, C2N2, C2F2, and F2CO). Overall,

the set includes a total of 181 unique bonds: 75 H−X, 49 X−Y, 43 X−−Y, and 14 X−−−Y bonds

(where X and Y are first- and second-row atoms). We use these CCSD(T)/A′V6Z reference

geometries to examine the basis set convergence of the CCSD(T) method with the VnZ and

A′VnZ basis sets (n = D,T,Q, 5), def2-nZVPP basis sets (n = T,Q), and the CCSD(T)-F12

method with the VnZ-F12 basis sets (n = D,T ). Our main conclusions can be summarized as

follows:

• In CCSD(T) calculations, the RMSD over the bond distances is reduced by a factor of 3–4

with each additional ‘zeta’ in the basis set. For example, the RMSD for the CCSD(T)/VnZ

levels of theory is 0.0196 (VDZ), 0.0050 (VTZ), 0.0015 (VQZ), and 0.0004 (V5Z) A. A

similar trend is obtained for the A′VnZ basis sets.

• An important finding is that the explicitly correlated CCSD(T)-F12/Vn Z-F12 levels of the-

ory show similar performance to the CCSD(T)/A′V(n + 2)Z levels of theory. For instance,

the CCSD(T)-F12/VDZ-F12 and CCSD(T)/VQZ levels of theory attain RMSDs of 0.0019

and 0.0015 A, respectively. Similarly, the CCSD(T)-F12/VTZ-F12 and CCSD(T)/V5Z

levels of theory attain RMSDs of 0.0006 and 0.0004 A, respectively. In both cases,

the CCSD(T)-F12 calculations use about 10% of the CPU time and disk space as the

conventional CCSD(T) calculations. Given these findings it seems pointless to use conven-

tional CCSD(T) calculations for geometry optimizations. We recommend replacing the

CCSD(T)/VQZ geometries used in highly accurate composite ab initio theories (e.g., W4

and HEAT) with computationally more economical CCSD(T)- F12/VDZ-F12 geometries.

132

6.5. ACKNOWLEDGEMENTS

• Due to systematic errors in conventional CCSD(T) calculations, we find that simple scaling

of the bond distances is an effective way to improve the performance at no additional

computational cost. This is particularly true for the smaller double-zeta and triple-zeta

quality basis sets. For example, the RMSD for the unscaled CCSD(T)/VDZ bond distances

(0.0196 A) is reduced by about 60% upon scaling (0.0080 A). Thus, in cases where a

CCSD(T)-F12/VDZ-F12 geometry optimization is computationally too expensive, we

recommend using scaled CCSD(T)/VDZ bond distances.

• Finally, we note that the use of CCSD(T)/VDZ geometries in composite ab initio theories

(such as W1-F12 and W2-F12) leads to an RMSD of over 2.4 kJ mol−1relative to the use of

CCSD(T)/A′V6Z geometries. This RMSD is reduced to 0.2 kJ mol−1(for CCSD(T)/VTZ

geometries) and merely 0.02 kJ mol−1(for CCSD(T)/VQZ geometries). However, the

computationally economical CCSD(T)-F12/VDZ-F12 geometries also result in a near-zero

RMSD of 0.02 kJ mol−1. We therefore recommend using the latter in highly accurate

composite ab initio methods such as W4lite, W4, and W4-F12.

• The performance of the def2-TZVPP and def2-QZVPP basis sets is very similar to that of

the VTZ and VQZ basis sets

6.5 Acknowledgements

We gratefully acknowledge the system administration support provided by the Faculty of Science

at UWA to the Linux cluster of the Karton group, the financial support of Danish National

Research Foundation (Center for Materials Crystallography, DNRF-93) to P.R.S., and an

Australian Research Council (ARC) Discovery Early Career Researcher Award to A.K. (Grant

No. DE140100311).

133

CHAPTER 6. BASIS SET CONVERGENCE OF CCSD(T) EQUILIBRIUMGEOMETRIES USING A LARGE AND DIVERSE SET OF MOLECULARSTRUCTURES

134

Bibliography

1. Shavitt, I. & Bartlett, R. J. Many-body methods in chemistry and physics: MBPT and

coupled-cluster theory, Cambridge molecular science series (Cambridge University Press,

Cambridge, 2009).

2. Bartlett, R. J. & Musia l, M. Coupled-cluster theory in quantum chemistry. Rev. Mod.

Phys. 79 (2007).

3. Karton, A. A computational chemist’s guide to accurate thermochemistry for organic

molecules. Wiley Interdiscip. Rev. Comput. Mol. Sci. 6, 292 (2016).

4. Peterson, K. A., Feller, D. & Dixon, D. A. Chemical accuracy in ab initio thermochemistry

and spectroscopy: current strategies and future challenges. Theor. Chem. Acc. 131, 1079

(2012).

5. Karton, A., Daon, S. & Martin, J. M. L. W4-11: A high-confidence benchmark dataset

for computational thermochemistry derived from first-principles W4 data. Chem. Phys.

Lett. 510, 165 (2011).

6. Feller, D., Peterson, K. A. & Dixon, D. A. A survey of factors contributing to accurate

theoretical predictions of atomization energies and molecular structures. J. Chem. Phys.

129, 204105 (2008).

7. Helgaker, T., Klopper, W. & Tew, D. P. Quantitative quantum chemistry. Mol. Phys.

106, 2107 (2008).

8. Raghavachari, K., Trucks, G. W., Pople, J. A. & Headgordon, M. A 5th-order perturbation

comparison of electron correlation theories. Chem. Phys. Lett. 157, 479 (1989).

9. Watts, J. D., Gauss, J. & Bartlett, R. J. Coupled-cluster methods with noniterative triple

excitations for restricted open-shell hartree-fock and other general single determinant

reference functions - energies and analytical gradients. J. Chem. Phys. 98, 8718 (1993).

10. Karton, A. & Martin, J. M. L. Explicitly correlated Wn theory: W1-F12 and W2-F12. J.

Chem. Phys. 136, 124114 (2012).

11. Curtiss, L. A., Redfern, P. C. & Raghavachari, K. Gn theory. Wiley Interdiscip. Rev.

Comput. Mol. Sci. 1, 810 (2011).

135

BIBLIOGRAPHY

12. N. DeYonker, T. R. C. & Wilson, A. K. Advances in the Theory of Atomic Molecular

Systems, Progress in Theoretical Chemistry Physics 197–224 (Springer, Netherlands,

Dordrecht, 2009).

13. Martin, J. M. L. Computational thermochemistry: A brief overview of quantum mechanical

approaches. Annu. Rep. Comput. Chem 1 (2005).

14. Helgaker, T. et al. in (ed Ciolowski, J.) 1 (Kluwer, Dordrecht, 2001).

15. Martin, J. M. L. & Parthiban, S. Quantum-Mechanical Prediction of Thermochemical

Data, Understanding Chemical Reactivity 31 (Kluwer, Dordrecht, 2001).

16. Raghavachari, K. Historical perspective on: A fifth-order perturbation comparison of

electron correlation theories [Volume 157, Issue 6, 26 May 1989, Pages 479-483]. Chem.

Phys. Lett. 589, 35 (2013).

17. D. A. Dixon, D. F. & Peterson, K. A. Chapter One - A Practical Guide to Reliable First

Principles Computational Thermochemistry Predictions Across the Periodic Table. Annu.

Rep. Comput. Chem. 8 (2012).

18. Feller, D., Peterson, K. A. & Dixon, D. A. Further benchmarks of a composite, con-

vergent, statistically calibrated coupled-cluster-based approach for thermochemical and

spectroscopic studies. Mol. Phys 110, 2381–2399 (2012).

19. Karton, A. & Martin, J. M. L. Performance of W4 theory for spectroscopic constants and

electrical properties of small molecules. J. Chem. Phys. 133, 144102 (2010).

20. Martin, J. M. L. & Kesharwani, M. K. Assessment of CCSD(T)-F12 Approximations and

Basis Sets for Harmonic Vibrational Frequencies. J. Chem. Theory Comput. 10, 2085

(2014).

21. Tentscher, P. R. & Arey, J. S. Geometries and Vibrational Frequencies of Small Radicals:

Performance of Coupled Cluster and More Approximate Methods. J. Chem. Theory

Comput. 8, 2165 (2012).

22. Karton, A. How large are post-CCSD(T) contributions to the total atomization energies

of medium-sized alkanes? Chem. Phys. Lett. 645, 118–122. issn: 0009-2614 (2016).

23. Schwenke, D. W. The extrapolation of one-electron basis sets in electronic structure

calculations: How it should work and how it can be made to work. J. Chem. Phys. 122,

014107 (2005).

24. Barnes, E. C., Petersson, G. A., Feller, D. & Peterson, K. A. The CCSD(T) complete

basis set limit for Ne revisited. J. Chem. Phys. 129, 194115 (2008).

25. Feller, D., Peterson, K. A. & Hill, J. G. On the effectiveness of CCSD(T) complete basis

set extrapolations for atomization energies. J. Chem. Phys. 135, 044102 (2011).

136

BIBLIOGRAPHY

26. Feller, D. Benchmarks of improved complete basis set extrapolation schemes designed for

standard CCSD(T) atomization energies. J. Chem. Phys. 138, 074103 (2013).

27. Karton, A., Taylor, P. R. & Martin, J. M. L. Basis set convergence of post-CCSD

contributions to molecular atomization energies. J. Chem. Phys. 127, 064104 (2007).

28. Feller, D., Peterson, K. A. & Crawford, T. D. Sources of error in electronic structure

calculations on small chemical systems. J. Chem. Phys. 124, 054107 (2006).

29. Heckert, M., Kallay, M. & Gauss, J. Molecular equilibrium geometries based on coupled-

cluster calculations including quadruple excitations. Mol. Phys 103, 2109–2115 (2005).

30. Feller, D., Peterson, K. A., de Jong, W. A. & Dixon, D. A. Performance of coupled cluster

theory in thermochemical calculations of small halogenated compounds. J. Chem. Phys.

118, 3510–3522 (2003).

31. Bak, K. L. et al. The accurate determination of molecular equilibrium structures. J. Chem.

Phys. 114, 6548–6556 (2001).

32. Feller, D. & Peterson, K. A. An examination of intrinsic errors in electronic structure

methods using the Environmental Molecular Sciences Laboratory computational results

database and the Gaussian-2 set. J. Chem. Phys. 108, 154 (1998).

33. Helgaker, T., Klopper, W., Koch, H. & Noga, J. Basis-set convergence of correlated

calculations on water. J. Chem. Phys. 106, 9639 (1997).

34. Feller, D. & Peterson, K. A. Probing the limits of accuracy in electronic structure

calculations: Is theory capable of results uniformly better than “chemical accuracy”? J.

Chem. Phys. 126, 114105 (2007).

35. Feller, D., Peterson, K. A. & Dixon, D. A. Refined Theoretical Estimates of the Atom-

ization Energies and Molecular Structures of Selected Small Oxygen Fluorides. J. Phys.

Chem. A 114. PMID: 19968310, 613–623 (2010).

36. Karton, A., Kaminker, I. & Martin, J. M. L. Economical Post-CCSD(T) Computational

Thermochemistry Protocol and Applications to Some Aromatic Compounds. J. Phys.

Chem. A 113, 7610 (2009).

37. Knizia, G., Adler, T. B. & Werner, H.-J. Simplified CCSD(T)-F12 methods: Theory and

benchmarks. J. Chem. Phys. 130, 054104 (2009).

38. Heckert, M., Kallay, M., Tew, D. P., Klopper, W. & Gauss, J. Basis-set extrapolation

techniques for the accurate calculation of molecular equilibrium geometries using coupled-

cluster theory. J. Chem. Phys. 125, 044108 (2006).

39. Feller, D., Peterson, K. A. & Hill, J. G. Calibration study of the CCSD(T)-F12a/b

methods for C2 and small hydrocarbons. J. Chem. Phys. 133, 184102 (2010).

137

BIBLIOGRAPHY

40. Werner, H.-J. et al. MOLPRO, version 2012.1, a package of ab initio programs see. 2012.

41. Werner, H.-J., Knowles, P. J., Knizia, G., Manby, F. R. & Schuetz, M. Molpro: a general-

purpose quantum chemistry program package. Wiley Interdiscip. Rev. Comput. Mol. Sci.

2, 242 (2012).

42. Dunning, T. H. Gaussian-basis sets for use in correlated molecular calculations .1. the

atoms boron through neon and hydrogen. J. Chem. Phys. 90, 1007 (1989).

43. Kendall, R. A., Dunning, T. H. & Harrison, R. J. Electron-affinities of the 1st-row atoms

revisited - systematic basis-sets and wave-functions. J. Chem. Phys. 96, 6796 (1992).

44. Dunning, T. H., Peterson, K. A. & Wilson, A. K. Gaussian basis sets for use in correlated

molecular calculations. X. The atoms aluminum through argon revisited. J. Chem. Phys.

114, 9244 (2001).

45. Weigend, F. & Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and

quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys.

Chem. Chem. Phys. 7, 3297 (2005).

46. Adler, T. B., Knizia, G. & Werner, H.-J. A simple and efficient CCSD(T)-F12 approxi-

mation. J. Chem. Phys. 127, 221106 (2007).

47. Peterson, K. A., Adler, T. B. & Werner, H.-J. Systematically convergent basis sets for

explicitly correlated wavefunctions: The atoms H, He, B-Ne, and Al-Ar. J. Chem. Phys.

128, 084102 (2008).

48. Wang, S. Y. & Schaefer, H. F. The small planarization barriers for the amino group in

the nucleic acid bases. J. Chem. Phys. 124, 044303 (2006).

49. Wiberg, K. B. Basis set effects on calculated geometries: 6-311++G** vs. aug-cc-pVDZ.

J. Comput. Chem. 25, 1342 (2004).

50. Martin, J. M. L. & Taylor, P. R. Structure and vibrations of small carbon clusters from

coupled-cluster calculations. J. Phys. Chem. 100, 6047 (1996).

51. Xie, Y. M., Scuseria, G. E., Yates, B. F., Yamaguchi, Y. & Schaefer, H. F. Methylnitrene

- theoretical predictions of its molecular-structure and comparison with the conventional

c-n single bond in methylamine. J. Am. Chem. Soc. 111, 5181 (1989).

52. Karton, A., Rabinovich, E., Martin, J. M. L. & Ruscic, B. W4 theory for computational

thermochemistry: In pursuit of confident sub-kJ/mol predictions. J. Chem. Phys. 125,

144108 (2006).

53. Sylvetsky, N., Peterson, K. A., Karton, A. & an M. L. Martin, J. Toward a W4-F12 ap-

proach: Can explicitly correlated and orbital-based ab initio CCSD(T) limits be reconciled?

J. Chem. Phys. 144, 214101 (2016).

138

BIBLIOGRAPHY

54. Tajti, A. et al. HEAT: High accuracy extrapolated ab initio thermochemistry. J. Chem.

Phys. 121, 11599 (2004).

55. Bomble, Y. J. et al. High-accuracy extrapolated ab initio thermochemistry. II. Minor

improvements to the protocol and a vital simplification. J. Chem. Phys. 125, 064108

(2006).

56. Harding, M. E. et al. High-accuracy extrapolated ab initio thermochemistry. III. Additional

improvements and overview. J. Chem. Phys. 128, 114111 (2008).

57. Hattig, C., Klopper, W., Kohn, A. & Tew, D. P. Explicitly Correlated Electrons in

Molecules. Chem. Rev. 112, 4 (2011).

58. Kong, L., Bischoff, F. A. & Valeev, E. F. Explicitly Correlated R12/F12 Methods for

Electronic Structure. Chem. Rev. 112, 75 (2011).

59. Ten-no, S. & Noga, J. Explicitly correlated electronic structure theory from R12/F12

ansatze. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2, 114 (2012).

60. Hill, J. G., Peterson, K. A., Knizia, G. & Werner, H.-J. Extrapolating MP2 and CCSD

explicitly correlated correlation energies to the complete basis set limit with first and

second row correlation consistent basis sets. J. Chem. Phys. 131, 194105 (2009).

61. Feller, D., Peterson, K. A. & Ruscic, B. Improved accuracy benchmarks of small molecules

using correlation consistent basis sets. Theo. Chem. Acc. 133, 1407. issn: 1432-2234

(2013).

139

BIBLIOGRAPHY

140

Chapter 7

Estimating the CCSD basis-set limit

energy from small basis sets

ABSTRACT

Coupled cluster calculations with all single and double excitations (CCSD) converge exceed-

ingly slowly with the size of the one-particle basis set. We assess the performance of a number

of approaches for obtaining CCSD correlation energies close to the complete basis-set limit

in conjunction with relatively small DZ and TZ basis sets. These include global and system-

dependent extrapolations based on the A + B/Lα two-point extrapolation formula, and the

well-known additivity approach that uses an MP2-based basis-set-correction term. We show that

the basis set convergence rate can change dramatically between different systems(e.g.it is slower

for molecules with polar bonds and/or second-row elements). The system-dependent basis-set

extrapolation scheme, in which unique basis-set extrapolation exponents for each system are

obtained from lower-cost MP2 calculations, significantly accelerates the basis-set convergence

relative to the global extrapolations. Nevertheless, we find that the simple MP2-based basis-set

additivity scheme outperforms the extrapolation approaches. For example, the following root-

mean-squared deviations are obtained for the 140 basis-set limit CCSD atomization energies in

the W4-11 database: 9.1 (global extrapolation), 3.7 (system-dependent extrapolation), and 2.4

(additivity scheme) kJ mol−1. The CCSD energy in these approximations is obtained from basis

sets of up to TZ quality and the latter two approaches require additional MP2 calculations with

basis sets of up to QZ quality. We also assess the performance of the basis-set extrapolations and

additivity schemes for a set of 20 basis-set limit CCSD atomization energies of larger molecules

including amino acids, DNA/RNA bases, aromatic compounds, and platonic hydrocarbon

cages. We obtain the following RMSDs for the above methods: 10.2 (global extrapolation), 5.7

(system-dependent extrapolation), and 2.9 (additivity scheme) kJ mol−1.

141

CHAPTER 7. ESTIMATING THE CCSD BASIS-SET LIMIT ENERGYFROM SMALL BASIS SETS

7.1 Introduction

Coupled-cluster theory is one of the most reliable, yet computationally affordable, methods for

solving the nonrelativistic electronic Schrodinger equation.[1] Coupled-cluster theory entails a

hierarchy of approximations that can be systematically improved towards the exact quantum

mechanical solution, providing a roadmap for highly accurate chemical properties.[2–8] In

particular, the CCSD(T) method (coupled-cluster with single, double, and quasiperturbative

triple excitations) has been found to be a cost-effective approach for the calculation of reliable

thermochemical and kinetic data(e.g. reaction energies and barrier heights) as well as molecular

properties based on energy derivatives (e.g. equilibrium/transition structures, vibrational

frequencies, and electrical properties).[9–14] However, one of the greatest weaknesses of the

CCSD(T) method is that it converges exceedingly slowly to the complete basis set (CBS) limit.

This is particularly true for the double excitations, as they reflect dynamical rather than static

electron correlation effects.[2, 3]

One way to accelerate the basis set convergence is to use basis set extrapolations that exploit

the systematic behavior of the correlation-consistent basis sets of Dunning and coworkers.[15–17]

There has been an extensive discussion in the literature on the form that this extrapolation

should take,[18–41], where the focus has been on extrapolations that use relatively large basis

sets of at least quadruple-zeta quality. References [32–41] provide a comprehensive review of

the previous works. These extrapolations use a global extrapolation formula for all systems

regardless of the unique electronic structure and bonding situations of each system. The use of

a global extrapolation formula is valid when large basis sets are used, however, it becomes less

valid for basis sets that are far from the asymptotic limit.

In the present work we consider two alternative approaches for obtaining CCSD atomization

energies close to the basis-set limit in conjunction with relatively small basis sets. The first is a

system-dependent extrapolation scheme that accounts for the different basis-set convergence

rate between different systems. We show that the basis-set convergence rate of the CCSD

correlation energy can differ substantially between different molecular systems, and we introduce

a system-dependent extrapolation scheme to account for these differences. In this scheme, unique

extrapolation exponents for each molecular system are obtained from lower-cost second-order

Møller-Plesset perturbation theory (MP2) calculations. In this context, it is worth mentioning

the work of Bakowies,[36] which highlights the limitations of conventional extrapolations and

proposes an empirically motivated extrapolation formula for MP2 and CCSD correlation energies.

The scaling factors in this extrapolation depend on the number of hydrogen atoms and the

number of correlated electrons in each system.The second approach is an additivity scheme

in which the CCSD energy is calculated with a relatively small basis set and an MP2-based

basis-set-correction term is added in order to approximate the CCSD/CBS energy.This cost-

effective approach, which is widely used in conjunction with the correlation-consistent basis

142

7.2. COMPUTATIONAL METHODS

sets for obtaining noncovalent interaction energies at the CCSD(T)/CBS limit[42–45], has been

recently shown to give good performance for reaction energies[46, 47] and barrier heights[48].

7.2 Computational Methods

All calculations were carried out on the Linux cluster of the Karton group at UWA. All MP2

and CCSD calculations were carried out with the correlation-consistent basis sets, using version

2012.1 of the MOLPRO program suite.[49, 50] The notation A′VnZ indicates the combination of

the standard correlation-consistent cc− pV nZ basis sets on hydrogen,[15] the aug− cc− pV nZbasis sets on first-row elements,[16] and the aug − cc− pV (n + d)Z basis sets on second-row

elements.[17] All correlated calculations were performed within the frozen-core approximation,

i.e. the 1s orbitals (for first-row atoms) and the 1s, 2s, and 2p orbitals (for second-row atoms) are

constrained to be doubly occupied in all configurations.We consider the two-point extrapolation

formula of Halkier et al.[26]

E(L) = E∞ + A/Lα (7.1)

where L is the highest angular momentum represented in the basis set, and α can be a

global exponent (Section III A) or a system-dependent exponent obtained individually for

each system from lower-cost MP2 calculations (Section III B). We note that the global two-

point extrapolations are extensively used in highly accurate composite methods (e.g. Wn[2, 4,

10, 51, 52], Wn-F12[53], and HEAT)[5–7] in conjunction with relatively large basis set of at

least quadruple-ζ quality. We also consider the empirical two-point extrapolation formula of

Schwenke[34].

E∞ = E(L) + F [E(L+ 1)− E(L)] (7.2)

which does not involve L (where F is an empirical scaling factor). We note that this expression

is equivalent to the above inverse power extrapolation with an empirical extrapolation exponent.

The notation A′V{n, n + 1}Z indicates extrapolation from the A′VnZ and A′V(n + 1)Z basis

sets using the aforementioned two-point formulas. Basis set extrapolations that use only a single

extrapolation parameter (α or F) for all systems will be referred to as global extrapolations,

whereas extrapolations that use a unique extrapolation exponent for each system will be referred

to as system-dependent extrapolations. We also assess the performance of an additivity-based

approach for approximating the CCSD/CBS energy. In this approach the CCSD base-energy

is calculated with a small basis set and an MP2-based basis-set-correction term (∆ MP2) is

calculated with larger basis sets

CCSD /CBS ≈ CCSD /X(MP2 /Y) = CCSD /A′VXZ + ∆ MP2 (7.3)

where ∆ MP2 is given by

143

CHAPTER 7. ESTIMATING THE CCSD BASIS-SET LIMIT ENERGYFROM SMALL BASIS SETS

∆ MP2 = MP2 /A′V{Y–1,Y}Z– MP2 /A′VXZ (7.4)

and the MP2/A′V{Y–1,Y}Z energy is extrapolated to the basis set limit with a global

two-point extrapolation formula (eq. 7.1). We note that the popular Gn(MP2) composite

thermochemical methods use a similar approach in conjunction with Pople-type basis sets

to approximate the CCSD(T)/TZ energy.[54–56] The performance of the various basis-set

extrapolations and additivity schemes are evaluated relative to the basis-set limit valence CCSD

atomization energies in the W4-11 database.[4] This database includes 140 total atomization

energies (TAEs) of species that cover a broad spectrum of bonding situations and as such it

is an excellent benchmark set for the validation of basis-set extrapolation techniques. The

W4-11 dataset includes the following species: closed shell (97), radicals (27), singlet carbenes

(9), and triplet systems (7). In terms of elemental composition the dataset includes 100 first-

row species, 19 second-row species, and 21 mixed first- and second-row species.[4] The CCSD

atomization energies in the W4-11 dataset were obtained by means of the W4 thermochemical

protocol.[2] In particular, they are extrapolated from the A′V5Z and A′V6Z basis sets. Following

the suggestion of Klopper[32], they are partitioned into singlet-coupled pair energies, triplet-

coupled pair energies, and T1 terms. The singlet-coupled and triplet-coupled pair energies

are extrapolated using eq. 7.1 with αs = 3 and αt = 5, respectively, and the T1 term (which

exhibits a weak basis-set dependence) is set equal to that in the largest basis set. In addition,we

evaluate the performance of the basis-set extrapolations and additivity schemes for a set of 20

TAEs of systems that are much larger than those in the W4-11 database. These include:

• Benzene and fulvene (C6H6), phenyl radical (C6H5), pyridine (C5H5N), pyrazine (C4H4N2),

furan (C4H4O), 1,3-butadiene (C4H6), and cyclobutene (cyc-C4H6). The CCSD correlation

component of the TAEs for these systems are obtained from W3.2 theory (i.e. extrapolated

from the A′V{Q, 5}Z basis-set pair, see Ref. 10 for further details).

• The DNA/RNA bases adenine (H5C5N5), cytosine (H5C4N3O), thymine (H6C5N2O2), and

uracil (H4C4N2O2). The amino acids alanine (H7C3NO2), cysteine (H7C3NO2S), glycine

(H5C2NO2), methionine (H11C5NO2S), and serine (H7C3NO3). The platonic hydrocarbon

cages tetrahedrane (C4H4), triprismane (C6H6), and cubane (C8H8). The CCSD/CBS

TAEs for these compounds are obtained from W2-F12 theory (i.e. extrapolated from the

V{T,Q}Z-F12 basis-set pair). These reference values are taken from Ref. [53] (DNA/RNA

bases), Ref. [46] (amino acids), and Ref. [48] (platonic hydrocarbons).

144

7.3. RESULTS AND DISCUSSION

7.3 Results and Discussion

7.3.1 Performance of conventional basis set extrapolations for ob-

taining the CCSD correlation component of the TAEs in the

W4-11 dataset

Table 7.1 gives the error statistics for the CCSD component of the TAEs extrapolated from the

A′V{D,T}Z, A′V{T,Q}Z, and A′V{Q, 5}Z basis-set pairs. The individual errors can be found

in Table S1 of the supplementary material. As expected, extrapolating the CCSD energy from

the A′V{D,T}Z basis sets with α = 3 results in a very large RMSD of 13.5kJ mol−1. This level

of theory tends to systematically and severely underestimate the basis-set-limit TAEs, as evident

from the mean signed deviation (MSD) and mean absolute deviation (MAD) being practically

the same (MAD ≈ MSD = –10.8kJ mol−1). Ad hoc optimization of α to minimize the RMSD

over the W4-11 training set leads to α = 2.357. Using this empirical extrapolation exponent

alleviates the bias toward underestimation of the basis-set limit TAEs (e.g. the MSD drops

to -2.2 kJ mol−1). However, an RMSD of 9.1 kJ mol−1is still unacceptably large (Table 7.1).

Remarkably, Truhlar obtained a very similar extrapolation exponent of α = 2.400 by minimizing

the RMSD over a much smaller set of correlation energies [27]. These two empirical exponents

(α = 2.357 and 2.400) result in similar overall error statistics for the W4-11 dataset. Using

Schwenke’s extrapolation in conjunction with the A′V{D,T}Z basis sets leads to a slightly higher

RMSD of 9.3 kJ mol−1[34]. We note that Schwenke’s extrapolation in this case is equivalent to

an inverse power extrapolation with α = 2.451.

Using the A′V{T,Q}Z basis sets in the extrapolations significantly improves the performance.

We note that these basis sets are used for extrapolating the CCSD correlation energy in W1w

theory [51]. An extrapolation exponent of α = 3 results in an RMSD of 2.8 kJ mol−1and a

largest positive deviation of 6.7 kJ mol−1(AlCl3). Using an empirical extrapolation exponent of

α = 3.403, which minimizes the RMSD over the TAEs in the W4-11 dataset, reduces the RMSD

to merely 1.2 kJ mol−1and the largest positive deviation to 3.08 kJ mol−1(Si2H6). We note that

Schwenke’s extrapolation (which is equivalent to an inverse power extrapolation with α = 3.803)

leads to an RMSD of 2.3 kJ mol−1and to a largest positive deviation of 5.8 kJ mol−1(Table 7.1).

Using the A′V{Q, 5}Z basis-set pair, which is used for the CCSD extrapolations in W2.2

theory,[2, 51],results in excellent performance. For example, with α = 3 we obtain an RMSD

of 0.8 kJ mol−1, and an empirical extrapolation exponent of 3.220 results in an RMSD of only

0.5 kJ mol−1. We note that in this case, Schwenke’s extrapolation leads to practically the same

RMSD (this extrapolation is equivalent to an inverse power extrapolation with α = 3.171).

145

CHAPTER 7. ESTIMATING THE CCSD BASIS-SET LIMIT ENERGYFROM SMALL BASIS SETS

Basis sets αd RMSD MAD MSD LND LPD

A′V{D,T}Zb 3 13.54 10.99 -10.83 -41.34 (SO3) 3.81 (B2H6)

A′V{D,T}Zb 2.357e 9.12 6.86 2.19 -32.57 (SO3) 20.67 (C3H8)

A′V{D,T}Zb 2.400f 9.16 6.89 -2.92 -32.32 (SO3) 18.55 (C3H8)

A′V{D,T}Zc g 9.30 6.99 -3.76 -34.16 (SO3) 16.51 (B2H6)

A′V{T,Q}Zb 3 2.80 2.38 2.37 -0.84 (ClO2) 6.66 (AlCl3)

A′V{T,Q}Zb 3.403e 1.15 0.88 0.07 -3.57 (S4) 3.08 (Si2H6)

A′V{T,Q}Zc g 2.27 1.87 1.83 -1.35 (ClO2) 5.78 (AlCl3)

A′V{Q, 5}Zb 3 0.81 0.54 0.51 -0.38 (CF4) 3.59 (AlCl3)

A′V{Q, 5}Zb 3.220e 0.54 0.38 -0.02 -1.01 (HCNO) 2.69 (AlCl3)

A′V{Q, 5}Zb g 0.55 0.42 -0.14 -1.17 (HCNO) 2.50 (AlCl3)

a RMSD = root-mean-squared deviation, MAD = mean absolute deviation, MSD = meansigned deviation, LND = largest negative deviation, and LPD = largest positive deviation(the molecules associated with the LND and LPD are given in parenthesis).

b Using eq. 7.1c Using Schwenke’s single parameter extrapolation formula eq. 7.2d Extrapolation exponent used in the two-point extrapolations.e Optimized to minimize RMSD over the 140 TAEs in the W4-11 dataset.f Extrapolation exponent taken from Ref. 27g The extrapolation fitting constants from Ref. 34 are F = 1.5877616 (A′V{D,T}Z),

1.7001115 (A′V{T,Q}Z), and 1.9303174 (A′V{Q, 5}Z)

Table 7.1: Error statistics for global extrapolations of the valence CCSD correlation componentover the 140 total atomization energies in the W4-11 dataset (kJ mol−1). The reference valuesare CCSD/A′V{5, 6}Z total atomization energies from W4 theory.a

146

7.3. RESULTS AND DISCUSSION

7.3.2 System-dependent basis set extrapolations of the CCSD cor-

relation energy

The global extrapolations considered in Section III A work well in conjunction with sufficiently

large basis sets (e.g. A′V{T,Q}Z and larger). With the A′V{T,Q}Z basis-set pair the 1/Lα

extrapolation results in RMSDs of 2.8 (α = 3) and 1.2 kJ mol−1(α = 3.403). However, using

this extrapolation in conjunction with the A′V{D,T}Z basis sets significantly worsens the

performance, such that even an empirical exponent results in an unacceptably large RMSD of

9.1 kJ mol–1. A natural question that arises is: can a single extrapolation exponent adequately

describe the basis set convergence of all systems in conjunction with small basis sets? For

systems for which we have results at the one-particle basis-set limit, this question can easily

be answeredby backtracking the ideal extrapolation exponents that will reproduce the CCSD

basis-set limit energies. A narrow distribution of the ideal extrapolation exponents around a

single value indicates a uniform basis set convergence, whereas a broad distribution of the ideal

extrapolation exponents indicates that a global extrapolation cannot possibly do a good job for

many systems. The two-point extrapolation formula from the A′VnZ and A′V(n + 1)Z basis

sets can be written in the form

E∞ = En+1 +En+1 − En(n+1

n)α − 1

(7.5)

Where En, En+1, and E∞ are the CCSD/A′VnZ, CCSD/A′V(n + 1)Z, and CCSD/CBS

energies, respectively. A simple rearrangement of eq. 7.1, gives the ideal extrapolation

exponent(αideal) that will exactly reproduce E∞ from En and En+1

αideal =ln( En+1−En

E∞−En+1+ 1)

ln(n+1n

)(7.6)

ideal extrapolation exponent can be obtained for any system for which En, En+1, and E∞

are known. Table S2 of the supplementary material1 gives the ideal exponents that will exactly

reproduce the CCSD/A′V{5, 6}Z energies from the A′V{D,T}Z, A′V{T,Q}Z, and A′V{Q, 5}Zbasis-set pairs, whilst Figure 1 shows the standard distributions of these ideal exponents. We

note three general observations:

• The Gaussian distribution for the ideal exponents for the A′V{D,T}Z, A′V{T,Q}Z and

A′V{Q, 5}Z basis-set pairs are centered at values that are progressively closer to 3.

1See supplementary material at http://dx.doi.org/10.1063/1.4921697 for the individual errors for the globalextrapolations (Table S1); ideal extrapolation exponents for system-dependent extrapolations that will reproducethe CCSD/CBS energies (Table S2 and Figure S1); individual errors for the system-dependent extrapolations(Table S3); extrapolation exponents for system-dependent extrapolations obtained from MP2 calculations (TableS4); individual errors for the CCSD/X(MP2/Y) additivity scheme (Table S5

147

CHAPTER 7. ESTIMATING THE CCSD BASIS-SET LIMIT ENERGYFROM SMALL BASIS SETS

Figure 7.1: Normal distributions of system-dependent extrapolation exponents (αideal), ob-tained from eq. 7.6 for the 140 molecules in the W4-11 database. The ideal exponents reproducethe CCSD/CBS energy for each molecule from the (a) A′V{D,T}Z and (b) A′V{T,Q}Z basis-setpairs. The Gaussians are centered around 2.22 (A′V{D,T}Z) and 3.46 (A′V{T,Q}Z), and havestandard deviations of 0.50 and 0.25, respectively.

• The ideal exponents for the A′V{D,T}Z basis-set pair are distributed over a much wider

interval than those for the A′V{T,Q}Z and A′V{Q, 5}Z basis-set pairs.

• Hydrogen-rich molecules tend to converge faster to the basis set limit whereas molecules

with polar bonds and/or second-row elements exhibit a slower basis set convergence.

The normal distribution functions for the ideal exponents are centered at 2.220 (A′V{D,T}Z),

3.462 (A′V{T,Q}Z), and 3.230 (A′V{Q, 5}Z). These mean values are in close agreement with

the empirical extrapolation exponents obtained for the global extrapolations by minimizing

the RMSD over the W4-11 dataset (Table 7.1). Specifically, the optimal global extrapolation

exponents are 2.357 (A′V{D,T}Z), 3.403 (A′V{T,Q}Z), and 3.220 (A′V{Q, 5}Z). The ideal

exponents for the A′V{D,T}Z basis-set pair span a wide range from 1.282 (SO2) to 3.504 (H2),

with a standard deviation of 0.502. This illustrates that a global extrapolation exponent cannot

possibly be a successful approach for estimating the CCSDbasis-set limit from the A′VDZ and

A′VTZ basis sets. The ideal exponents for the A′V{T,Q}Z basis-set pair span a narrower

interval, ranging from of 2.871 (ClO2) to 4.224 (AlH3), with a standard deviation of 0.251.

The narrower width of the normal distribution for the A′V{T,Q}Z basis sets indicates that a

global extrapolation scheme is more suitable to be used in conjunction with these larger basis

sets than with the A′V{D,T}Z basis-set pair. We note in passing that the ideal exponents

for the A′V{Q, 5}Z basis-set pair are asymmetrically distributed around the mean value and

therefore do not fit sufficiently well to a normal distribution. Nevertheless, they are narrowly

distributed around a value of 3.220 (Figure S1, supplementary material). The above results

illustrate the limitations of basis set extrapolations using a global exponent in conjunction with

148

7.3. RESULTS AND DISCUSSION

small basis sets and indicate that system-dependent extrapolations may result in performance

improvements. This raises the question: how can one obtain ideal (or effective) extrapolation

exponents without prior knowledge of the CCSD/CBS energy? In the present work,we consider

the possibility of estimating the exponents for each system from lower-cost MP2 calculations.

Such a system-dependent approach is not intended to be used as an actual extrapolation – as

it is not size consistent – but as a probe for the effective convergence rate. In particular, we

will use the following procedure to obtain effective (αeff) exponents for each system (atom or

molecule):

1. Extrapolate the MP2/CBS correlation energy from basis sets of at least A′V(n + 1)Z and

A′V(n + 2)Z quality using eq. 7.5.

2. Use eq. 7.6 to calculate an effective extrapolation exponent (αeff) that will reproduce the

MP2/CBS energy obtained in step (i) from the A′V{n, n + 1}Z basis-set pair.

3. Use the extrapolation exponent αeff to extrapolate the CCSD correlation energy from

the A′V{n, n + 1}Z basis-set pair. Table 7.2 gives error statistics for the CCSD TAEs

extrapolated from the A′V{D,T}Z basis-set pair in conjunction with system-dependent

exponents obtained via the above procedure. The supplementary material lists the errors

for each molecule (Table S3) and the system-dependent extrapolation exponents (Table

S4). Calculating the MP2/CBS energy from the A′V{T,Q}Z basis-set pair in step (i),

results in an RMSD of 4.4 kJ mol−1. This represents a significant improvement over the

global extrapolation from the same basis sets, which results in RMSDs of 13.5 (with α = 3)

and 9.1-9.3 kJ mol−1(with empirical exponents) (Table 7.1).

Note however, that our system-dependent approach is entirely devoid of empirical parameters.

The performance of the system-dependent extrapolation can be improved by introducing a global

scaling factor (λ) to account for systematic differences between the basis set convergence of the

MP2 and CCSDcorrelation energies [26, 27, 33, 36]. Introducing such an empirical scaling factor

optimized to minimize the RMSD over the W4-11 training set (i.e. multiplying all the system-

dependent extrapolation exponents by λ = 1.050) results in an RMSD of 3.7 kJ mol−1. We also

note that the largest negative and positive deviations for the system-dependent extrapolation

(–12.9 and 8.0 kJ mol−1, respectively) are significantly smaller than those obtained for the global

extrapolation (–32.6 and 20.7 kJ mol−1, respectively) (Table 7.1). Finally, it is worth pointing

out that the extra MP2/A′VQZ calculation that is needed for obtaining the system-dependent

extrapolation exponents is computationally more economical than the CCSD/A′VTZ calculation.

Therefore, the system-dependent approach does not significantly increase the computational cost

over the conventional approach. Calculating the MP2/CBS energy from the larger A′VQZ and

A′V5Z basis sets (eqs. 7.5 and 7.6) results in an RMSD of 3.2 kJ mol−1. It is also of interest to

assess the performance of the global and system-dependent CCSD/A′V{D,T}Z extrapolations

149

CHAPTER 7. ESTIMATING THE CCSD BASIS-SET LIMIT ENERGYFROM SMALL BASIS SETS

Basis setsb λf RMSD MAD MSD LND LPD

A′V{D,T}Zc N/A 4.42 3.47 1.80 -9.64 (SO3) 12.35 (C3H8)

A′V{D,T}Zc 1.050 3.65 2.77 -0.38 -12.88 (SO3) 8.02 (Si2H6)

A′V{D,T}Zd 1.050 3.18 2.52 -0.19 -10.68 (SO3) 7.83 (Si2H6)

A′V{T,Q}Zd 1.169 0.90 0.60 -0.02 -1.62 (ClO2) 5.36 (AlCl3)

A′V{T,Q}Ze 1.188 0.77 0.48 -0.01 -2.66 (ClO2) 3.40 (AlCl3)

A′V{Q, 5}Ze 1.114 0.24 0.17 0.00 -0.51 (C3H8) 1.06 (AlCl3)

a Footnore a from Table 7.1 applies here.b Basis sets used in the CCSD extrapolations (eq. 7.5)c The system-dependent extrapolation exponents are obtained from MP2 calculations in

conjunction with the A′VDZ, A′VTZ and A′VQZ basis sets (eq. 7.6)d The system-dependent extrapolation exponents are obtained from MP2 calculations in

conjunction with the A′VTZ, A′VQZ and A′V5Z basis sets (eq. 7.6)e The system-dependent extrapolation exponents are obtained from MP2 calculations in

conjunction with the A′VQZ, A′V5Z and A′V6Z basis sets (eq. 7.6)f Global scaling factor optimized to minimize the RMSD over the 140 TAEs in the W4-11

dataset.

Table 7.2: Error statistics for the system-dependent extrapolations of the valence CCSDcorrelation component over the 140 total atomization energies in the W4-11 test set (kJ mol−1) .The reference values are CCSD/A′V{5, 6}Z total atomization energies from W4 theorya

together with the Hartree–Fock (HF) and (T) contributions against CCSD(T)/CBS reference

values from the W4-11 database [4]. For this purpose we will take the HF and (T) components

from W1w theory [51],2 and add them to the CCSD/A′V{D,T}Z components calculated via

the global and system-dependent extrapolations. The global extrapolations result in RMSDs

that are close to those obtained for the CCSD component alone, namely they are 13.55 (α = 3)

and 9.50 (α = 2.357) kJ mol−1. The RMSDs for the system-dependent extrapolations increase

by about 10% relative to the RMSDs for the CCSDcomponent alone. In particular, calculating

the CCSD/A′V{D,T}Z energy using the non-empirical system-dependent extrapolation results

in an RMSD of 5.04 kJ mol−1. The empirical system-dependent extrapolations result in RMSDs

of 4.21 and 3.56 kJ mol−1(depending on the basis sets used for extrapolating the MP2/CBS

energy). So far we have shown that extrapolating the CCSD energy from the A′VDZ and A′VTZ

basis sets using system-dependent basis set extrapolations cuts the RMSD by over 50% relative

to the global basis set extrapolations.We now turn to CCSD extrapolations in conjunction with

the larger A′VTZ and A′VQZ basis sets. The system-dependent basis set extrapolation (with

an empirical scaling factor of λ = 1.169) results in an RMSD = 0.9 kJ mol−1.For comparison,

the global extrapolation with an optimized exponent of 3.403 results in an RMSD of 1.2

kJ mol−1(Table 7.1). Calculating the MP2/CBS energy from the A′V5Z and A′V6Z basis sets

results in a slightly lower RMSD of 0.8 kJ mol−1. Finally, extrapolating the CCSD energy from

2In W1w theory the HF energy is extrapolated from the A′V{T,Q}Z basis sets using eq. 7.1 with α = 5, andthe (T) contribution is extrapolated from the A′V{D,T}Z basis sets using the same extrapolation with α = 3.22.

150

7.3. RESULTS AND DISCUSSION

the A′V{Q, 5}Z basis-set pair using the system-dependent extrapolation results in a near-zero

RMSD of 0.2 kJ mol−1(in conjunction with a scaling factor of 1.114). Furthermore, the largest

deviations are very close to 1 kJ mol−1(Table 7.2). For comparison, the global extrapolation

(with an empirical exponent of α = 3.220) results in an RMSD that is roughly twice as large

(namely, 0.5 kJ mol−1), a largest positive deviation of 2.7 kJ mol−1, and a largest negative

deviation of –1.0 kJ mol−1(Table 7.1).

7.3.3 Performance of basis-set additivity schemes for obtaining the

CCSD correlation component of the TAEs in the W4-11 dataset

In the previous section we have shown that system-dependent extrapolations are superior to the

global basis-set extrapolations, most notably in conjunction with the A′V{D,T}Z basis set pair.

For example, the system-dependent approach results in RMSDs of 3.7 and 3.2 kJ mol−1(Table

7.2), whereas the global basis set extrapolation results in RMSDs of 9.1–13.5 kJ mol−1(Table

7.1). Another approach for estimating the CCSD/CBS energy through CCSD calculations with

smaller basis sets and MP2 calculations with larger basis sets is the additivity scheme (eqs. 7.3

and 7.4). Table 7.3 gives error statistics for the CCSD/X(MP2/Y) methods, the individual

deviations for the systems in the W4-11 database are given in Table S5 of the supplementary

material. Note that the CCSD/X(MP2/Y) notation indicates that the A′VXZ and A′VYZ basis

sets are the largest basis sets used in the CCSD and MP2 calculations, respectively.

Let us first consider the computationally most economical CCSD/D(MP2/Q) method. Using

an exponent of 3 for extrapolating the MP2 energy gives a relatively large RMSD of 11.8

kJ mol−1. Optimizing the exponent used in the MP2 extrapolation cuts the RMSD by more

than 50% and leads to an RMSD of 5.2 kJ mol−1. The performance of the CCSD/D(MP2/Q)

method with an empirical exponent is quite good considering its low computational cost. The

CCSD/T(MP2/Q) method results in an RMSD of 6.0 kJ mol−1with a non-empirical exponent

of α = 3. However, using an optimized exponent of α = 4.00 results in a remarkably low

RMSD of merely 2.4 kJ mol−1. For comparison, the system-dependent extrapolation from

the A′V{D,T}Z basis sets, which has the same computational cost as the CCSD/T(MP2/Q)

method, results in a larger RMSD of 3.7 kJ mol−1(Table 7.2).Extrapolating the MP2 energy

from the larger A′V{Q, 5}Z basis sets with an empirical extrapolation exponent further improves

the performance (RMSD = 1.9 kJ mol−1). Finally, calculating the CCSD energy in conjunction

with the A′VQZ basis set and extrapolating the MP2 energy from the A′V{Q, 5}Z basis

sets with an empirical exponent of α = 5.00 results in an RMSD of 0.78 kJ mol−1. The

system-dependent extrapolation with the same computational cost results in practically the

same RMSD of 0.77 kJ mol−1(Table 7.2). A useful way of looking at the performance of

the CCSD/X(MP2/Y) additivity scheme is to rewrite eq. 7.3 as ∆ CCSD ≈ ∆ MP2, where

∆ CCSD = CCSD /CBS– CCSD /A′VXZ and ∆ MP2 is given by eq. 7.4. Table S6 of the

151

CHAPTER 7. ESTIMATING THE CCSD BASIS-SET LIMIT ENERGYFROM SMALL BASIS SETS

CCSD MP2d αe RMSD MAD MSD LND LPD

A′VDZ A′V{T,Q}Z 3 11.76 10.23 10.20 -1.68 (BN) 29.54 (F2O2)

A′VDZ A′V{T,Q}Z 5.75f 5.18 3.62 0.54 -11.94 (AlCl3) 20.70 (F2O2)

A′VTZ A′V{T,Q}Z 3 6.01 5.23 5.18 -2.44 (AlCl3) 14.06 (C3H8)

A′VTZ A′V{T,Q}Z 4.00f 2.37 1.77 -0.05 -9.02 (AlCl3) 4.41 (F2O2)

A′VTZ A′V{Q, 5}Z 3 6.22 5.52 5.52 N/A 14.45 (P4)

A′VTZ A′V{Q, 5}Z 6.50f 1.90 1.44 0.22 -7.40 (AlCl3) 4.47 (F2O2)

A′VQZ A′V{Q, 5}Z 3 4.31 3.89 3.89 N/A 13.15 (P4)

A′VQZ A′V{Q, 5}Z 5.00f 0.78 0.62 -0.12 -3.12 (SO3) 1.70 (BN)

a Footnote a to Table 7.1 applies here.b The notation CCSD/X(MP2/Y) indicates that the CCSD correlation energy is calculated with the

A′VXZ basis set and the MP2correlation energy is extrapolated from the A′V(Y-1)Z A′VYZ basissets.

c Basis set used in the CCSD calculations (eq. (7.3)).d Basis sets used in the two-point MP2 extrapolations (eq. (7.4)).e Extrapolation exponent used in the MP2 extrapolations (eq. (7.1)).f Optimized to minimize the RMSD over the 140 TAEs in the W4-11 dataset.

Table 7.3: Error statistics for the valence CCSD correlation component over the 140 totalatomization energies in the W4-11 test set obtained via the CCSD/X(MP2/Y) additivity schemekJ mol−1. The reference values are CCSD/A′V{5, 6}Z total atomization energies from W4theory.a,b

supplementary material lists the ∆ MP2 and ∆ CCSD terms for the molecules in the W4-11

database. The magnitude of the ∆ CCSD term calculated in conjunction with the A′VDZ basis

set spans a wide range from 11.5 (F2) to 193.2 (P4) kJ mol−1. Inspection of Table S6 reveals

that for many systems the ∆ MP2 = MP2 /A′V{T,Q}Z– MP2 /A′VDZ basis-set correction

term does not provide a very good approximation to this ∆ CCSD term. For example, for

nearly 10% of the systems the difference between the ∆ MP2 and ∆ CCSD terms is greater

than 10 kJ mol−1(most notably, for oxides such as Cl2O, ClO2, F2O, FO2, HO3, O3, and

F2O2). On the other hand, the ∆ MP2 = MP2 /A′V{T,Q}Z– MP2 /A′VTZ term provides

an excellent approximation to the ∆ CCSD term calculated in conjunction with the A′VTZ

basis set. The difference between the ∆ MP2 and ∆ CCSD terms is smaller than 2 kJ mol−1for

67% of the systems, and it exceeds 5 kJ mol−1for only five molecules (S4, SiF4, SO3, P4, and

AlCl3). Upon removing these five second-row molecules from the evaluation set the RMSD for

the CCSD/T(MP2/Q) method drops from 2.4 to 1.9 kJ mol−1.

152

7.3. RESULTS AND DISCUSSION

7.3.4 Performance of basis-set extrapolations and additivity schemes

for the CCSD correlation energy for TAEs of larger molecules

The W4-11 dataset contains 140 first- and second-row species containing up to five non-hydrogen

atoms. For example, the largest systems in the W4-11 dataset are molecules such as acetic

acid (H4C2O2), cyanogen (N2C2), dioxygen difluoride (F2O2), aluminium trichloride (AlCl3),

sulfur trioxide (SO3), tetraphosphorus (P4), tetrafluoromethane (CF4), and tetrafluorosilane

(SiF4). It is also of interest to evaluate the performance of the basis-set extrapolations and

additivity schemes for larger molecules. This is important since the CCSD component to the

TAE increases with the size of the system. For example, the CCSD correlation component

for the molecules in the W4-11 database ranges between 77.5 (AlH) and 874.7 (propane)

kJ mol−1, whereas for the larger molecules considered in the present section (vide infra) it

ranges between 906.2 (1,3-butadiene) and 2130.1 (adenine) kJ mol−1. In addition, it is of

interest to evaluate the performance of the empirical extrapolations in which a single adjustable

parameter has been parameterized over the W4-11 dataset for molecules that are not included

in the training set. In particular, we will consider a set of 20 TAEs of larger molecules for

which we have CCSD/CBS reference values from W3.2lite (i.e. CCSD/A′V{Q, 5}Z) or W2-

F12 (i.e. CCSD-F12/V{T,Q}Z-F12) theory, this database will be referred to as the TAE20

dataset. These CCSD/CBS values allow us to evaluate the performance of conventional and

system-dependent basis-set extrapolations in conjunction with the A′V{D,T}Z basis-set pair.

The TAE20 dataset includes amino acids (alanine, cysteine, glycine, methionine, and serine);

DNA/RNA bases (adenine, cytosine, thymine, and uracil); aromatic compounds (benzene, phenyl

radical, pyridine, pyrazine, and furan); hydrocarbons (fulvene, 1,3-butadiene, and cyclobutene);

and hydrocarbon cages (tetrahedrane, triprismane, and cubane). Table 7.4 gives error statistics

for the CCSD/A′V{D,T}Z correlation energies extrapolated using the conventional basis-set

extrapolations. Using the two-point extrapolation with an exponent of α = 3 results in very

poor performance, which is much worse than the performance for the systems in the W4-11

database. To be specific, the RMSD for the TAE20 set is 27.0 kJ mol−1, compared to an RMSD

of 13.5 kJ mol−1for the W4-11 dataset. Using empirical exponents cuts the RMSD by about

60–70%, resulting in RMSDs of 15.1 (α = 2.357, this work), 12.6 (α = 2.357, Truhlar) [27], and

10.2 (α = 2.451, Schwenke) kJ mol−1. In all cases the largest deviations of 24.5–33.7 kJ mol−1are

obtained for cubane.

Using the system-dependent extrapolations in conjunction with the A′V{D,T}Z basis-set

pair results in RMSDs of 5.5–5.7 kJ mol−1(depending on the basis sets used for extrapolating the

MP2/CBS energy). These RMSDs are higher by about 2 kJ mol−1relative to the RMSDs obtained

for the W4-11 database (Table 7.2), nevertheless, they represent a substantial improvement over

the performance of the global extrapolations. The additivity approach significantly outperforms

the system-dependent extrapolations at no increase in the computational cost. In particular,

153

CHAPTER 7. ESTIMATING THE CCSD BASIS-SET LIMIT ENERGYFROM SMALL BASIS SETS

Basis setsc j RMSD MAD MSD LND LPD

Global extrapolations

A′V{D,T}Zd 3k 26.99 24.36 -24.36 -49.38 (adenine) N/A

A′V{D,T}Zd 2.357k 15.13 12.83 12.68 -1.52 (uracil) 33.71 (cubane)

A′V{D,T}Zd 2.400k 12.60 10.03 9.52 -4.88 (uracil) 29.39 (cubane)

A′V{D,T}Ze l 10.22 8.23 5.94 -8.69 (uracil) 24.48 (cubane)

System-dependent extrapolations

A′V{D,T}Zf 1.050m 5.70 5.27 5.27 N/A 9.70 (cubane)

A′V{D,T}Zg 1.050m 5.45 4.81 5.06 N/A 9.01 (cubane)

Additivity schemes

A′VDZh 5.75n 14.61 12.69 -12.69 -29.60 (adenine) N/A

A′VTZh 4.00n 2.88 2.18 -1.28 -7.35 (adenine) 2.15 (methionine)

A′VTZi 6.50n 2.53 1.93 -1.58 -5.32 (adenine) 1.20 (alanine)

a Footnote a to Table 7.1 applies here.b The reference values in the TAE20 database are calculated at the CCSD/A′V{Q, 5}Z (for benzene,

fulvene, phenyl radical, pyridine, pyrazine, furan, 1,3-butadiene, and cyclobutene) or CCSD-F12/V{T,Q}Z-F12 (for adenine, cytosine, thymine, uracil, alanine, cysteine, glycine, methionine,serine, tetrahedrane, triprismane, and cubane) levels of theory

c Basis sets used in the CCSD calculations.d Extrapolated using eq. (7.1).e Extrapolated using eq. (7.2).f The system-dependent extrapolation exponents are obtained from MP2calculations in conjunction

with the A′VDZ, A′VTZ, and A′VQZ basis sets (eq. (7.6)g The system-dependent extrapolation exponents are obtained from MP2calculations in conjunction

with the A′VTZ, A′VQZ, and A′V5Z basis sets (eq. (7.6)h The ∆ MP2 basis-set correction is calculated with the A′V{T,Q}Z basis sets (eq. (7.4)).i The ∆ MP2 basis-set correction is calculated with the A′V{Q, 5}Z basis sets (eq. (7.4)).j Single parameter used in the global and system-dependent extrapolations and in the additivity

schemes (see text).k Extrapolation exponent used in the two-point extrapolations (eq. (7.1)).l The extrapolation fitting constant from Ref. 34 is F = 1.5877616 (eq. (7.2)).m Scaling factor used in the system-dependent extrapolations.n Extrapolation exponent used in the MP2extrapolations (eq. (7.4)).

Table 7.4: Error statistics for the global and system-dependent extrapolations of the va-lence CCSD correlation energies over the 20 total atomization energies in the TAE20 dataset(kJ mol−1).a,b

154

7.4. CONCLUSIONS

the CCSD/T(MP2/Q)method results in an RMSD of 2.9 kJ mol−1, and the CCSD/T(MP2/5)

additivity scheme results in an RMSD of 2.5 kJ mol−1. We also note that the latter approach

results in a largest deviation (in absolute value) of only 5.3 kJ mol−1.

7.4 Conclusions

We assess the performance of a number of approaches for obtaining CCSDcorrelation energies

close to the complete basis set limit in conjunction with relatively small basis sets.These

approaches are evaluated against two datasets of highly accurate CCSD/CBS limit total

atomization energies. The first is the W4-11 database of 140 total atomization energies of

small molecules. The species in the W4-11 database cover a broad spectrum of bonding

situations and multireference character, and as such it is an excellent benchmark set for the

validation of basis-set extrapolation techniques. The reference values in the W4-11 database

are CCSD/A′V{5, 6}Z TAEs obtained from W4 theory. The second database is the TAE20 set

of 20 TAEs of larger molecules. This set includes five amino acids, four DNA/RNA bases, six

(hetero)aromatic compounds, three platonic hydrocarbon cages, and three simple hydrocarbons.

The reference values in the TAE20 database are either CCSD/A′V{Q, 5}Z TAEs (from W3.2

theory) or CCSD-F12/V{T,Q}Z-F12 TAEs (from W2-F12 theory). We consider the following

approaches for obtaining CCSD/CBS energies in conjunction with relatively small basis sets: (i)

global and system-dependent extrapolations based on the A+ B/Lα two-point extrapolation

formula, and (ii) the well-known additivity approach that uses an MP2-based basis-set-correction

term. In the system-dependent extrapolations the CCSD energy is extrapolated using the two-

point extrapolation formula A+B/Lαeff , where αeff is an effective extrapolation exponent that

is uniquely determined for each molecule (or atom) from lower-cost MP2 calculations. We

show that this system-dependent extrapolation scheme is superior to conventional basis-set

extrapolations, which use a global extrapolation exponent for all systems.For example, for the

W4-11 database we obtain the following RMSDs for the CCSD energies extrapolated from the

aug’-cc-pV(D+d)Z and aug’-cc-pV(T+d)Z basis sets: 9.1 (best global extrapolation), 3.7 (best

system-dependent extrapolation) kJ mol−1. For the TAE20 database we obtain somewhat higher

RMSDs, namely: 10.2 (best global extrapolation), 5.7 (best system-dependent extrapolation)

kJ mol−1. However, the additivity approach, which uses an MP2-based basis-set-correction term

and has the same computational cost as the above system-dependent extrapolation, significantly

outperforms the system-dependent extrapolation. In particular, it results in RMSDs of 2.4

(W4-11 dataset) and 2.9 (TAE20 dataset) kJ mol−1. Both the system-dependent extrapolation

and additivity approaches use a single adjustable parameter that minimizes the RMSD over the

W4-11 database. The CCSDcalculations in these two approaches are carried out in conjunction

with basis sets of up to aug’-cc-pV(T+d)Z quality and the MP2 calculations with basis sets of

up to aug’-cc-pV(Q+d)Z quality.

155

CHAPTER 7. ESTIMATING THE CCSD BASIS-SET LIMIT ENERGYFROM SMALL BASIS SETS

7.5 Acknowledgements

We gratefully acknowledge the system administration support provided by the Faculty of

Science at UWA to the Linux cluster of the Karton group, the financial support of Danish

National Research Foundation (Center for Materials Crystallography, DNRF-93) to PRS, and an

Australian Research Council (ARC) Discovery Early Career Researcher Award (DE140100311) to

AK.The authors would like to thank the referees of this manuscript for their valuable comments.

156

Bibliography

1. Shavitt, I. & Bartlett, R. J. Many-body methods in chemistry and physics: MBPT and

coupled-cluster theory, Cambridge molecular science series (Cambridge University Press,

Cambridge, 2009).

2. Karton, A., Rabinovich, E., Martin, J. M. L. & Ruscic, B. W4 theory for computational

thermochemistry: In pursuit of confident sub-kJ/mol predictions. J. Chem. Phys. 125,

144108 (2006).

3. Karton, A., Taylor, P. R. & Martin, J. M. L. Basis set convergence of post-CCSD

contributions to molecular atomization energies. J. Chem. Phys. 127, 064104 (2007).

4. Karton, A., Daon, S. & Martin, J. M. L. W4-11: A high-confidence benchmark dataset

for computational thermochemistry derived from first-principles W4 data. Chem. Phys.

Lett. 510, 165 (2011).

5. Tajti, A. et al. HEAT: High accuracy extrapolated ab initio thermochemistry. J. Chem.

Phys. 121, 11599 (2004).

6. Bomble, Y. J. et al. High-accuracy extrapolated ab initio thermochemistry. II. Minor

improvements to the protocol and a vital simplification. J. Chem. Phys. 125, 064108

(2006).

7. Harding, M. E. et al. High-accuracy extrapolated ab initio thermochemistry. III. Additional

improvements and overview. J. Chem. Phys. 128, 114111 (2008).

8. Klopper, W., Ruscic, B., Tew, D. P., Bischoff, F. A. & Wolfsegger, S. Atomization energies

from coupled-cluster calculations augmented with explicitly-correlated perturbation theory.

Chem. Phys. 356, 14 (2009).

9. Peterson, K. A., Feller, D. & Dixon, D. A. Chemical accuracy in ab initio thermochemistry

and spectroscopy: current strategies and future challenges. Theor. Chem. Acc. 131, 1079

(2012).

10. Karton, A., Kaminker, I. & Martin, J. M. L. Economical Post-CCSD(T) Computational

Thermochemistry Protocol and Applications to Some Aromatic Compounds. J. Phys.

Chem. A 113, 7610 (2009).

157

BIBLIOGRAPHY

11. Helgaker, T., Klopper, W. & Tew, D. P. Quantitative quantum chemistry. Mol. Phys.

106, 2107 (2008).

12. Martin, J. M. L. Computational thermochemistry: A brief overview of quantum mechanical

approaches. Annu. Rep. Comput. Chem 1 (2005).

13. Karton, A. & Martin, J. M. L. Performance of W4 theory for spectroscopic constants and

electrical properties of small molecules. J. Chem. Phys. 133, 144102 (2010).

14. Martin, J. M. L. & Kesharwani, M. K. Assessment of CCSD(T)-F12 Approximations and

Basis Sets for Harmonic Vibrational Frequencies. J. Chem. Theory Comput. 10, 2085

(2014).

15. Dunning, T. H. Gaussian-basis sets for use in correlated molecular calculations .1. the

atoms boron through neon and hydrogen. J. Chem. Phys. 90, 1007 (1989).

16. Kendall, R. A., Dunning, T. H. & Harrison, R. J. Electron-affinities of the 1st-row atoms

revisited - systematic basis-sets and wave-functions. J. Chem. Phys. 96, 6796 (1992).

17. Dunning, T. H., Peterson, K. A. & Wilson, A. K. Gaussian basis sets for use in correlated

molecular calculations. X. The atoms aluminum through argon revisited. J. Chem. Phys.

114, 9244 (2001).

18. Feller, D. The use of systematic sequences of wave-functions for estimating the complete

basis set, full configuration-interaction limit in water. J. Chem. Phys. 98, 7059 (1993).

19. Peterson, K. A., Woon, D. E. & Dunning, T. H. Benchmark calculations with correlated

molecular wave-functions .4. the classical barrier height of the h+h-2-]h-2+h reaction. J.

Chem. Phys. 100, 7410 (1994).

20. Martin, J. M. L. On the performance of correlation consistent basis sets for the calculation

of total atomization energies, geometries, and harmonic frequencies. J. Chem. Phys. 100,

8186–8193 (1994).

21. Martin, J. M. L. Ab initio total atomization energies of small molecules - Towards the

basis set limit. Chem. Phys. Lett. 259, 669 (1996).

22. Martin, J. M. L. & Lee, T. J. The atomization energy and proton affinity of NH3. An ab

initio calibration study. Chem. Phys. Lett. 258, 136 (1996).

23. Martin, J. M. L. & Taylor, P. R. Benchmark quality total atomization energies of small

polyatomic molecules. J. Chem. Phys. 106, 8620 (1997).

24. Wilson, A. K. & Dunning, T. H. Benchmark calculations with correlated molecular wave

functions .10. Comparison with ”exact” MP2 calculations on Ne, HF, H2O, and N-2. J.

Chem. Phys. 106, 8718 (1997).

25. Helgaker, T., Klopper, W., Koch, H. & Noga, J. Basis-set convergence of correlated

calculations on water. J. Chem. Phys. 106, 9639 (1997).

158

BIBLIOGRAPHY

26. Halkier, A. et al. Basis-set convergence in correlated calculations on Ne, N-2, and H2O.

Chem. Phys. Lett. 286, 243 (1998).

27. Truhlar, D. G. Basis-set extrapolation. Chem. Phys. Lett. 294, 45 (1998).

28. Feller, D. & Peterson, K. A. An examination of intrinsic errors in electronic structure

methods using the Environmental Molecular Sciences Laboratory computational results

database and the Gaussian-2 set. J. Chem. Phys. 108, 154 (1998).

29. Fast, P. L., Sanchez, M. L. & Truhlar, D. G. Infinite basis limits in electronic structure

theory. J. Chem. Phys. 111, 2921 (1999).

30. Feller, D. & Peterson, K. A. Probing the limits of accuracy in electronic structure

calculations: Is theory capable of results uniformly better than “chemical accuracy”? J.

Chem. Phys. 126, 114105 (2007).

31. Varandas, A. J. C. Basis-set extrapolation of the correlation energy. J. Chem. Phys. 113,

8880 (2000).

32. Klopper, W. Highly accurate coupled-cluster singlet and triplet pair energies from explicitly

correlated calculations in comparison with extrapolation techniques. Mol. Phys. 99, 481

(2001).

33. Valeev, E. F., Allen, W. D., Hernandez, R., Sherrill, C. D. & Schaefer, H. F. On the

accuracy limits of orbital expansion methods: Explicit effects of k-functions on atomic

and molecular energies. J. Chem. Phys. 118, 8594 (2003).

34. Schwenke, D. W. The extrapolation of one-electron basis sets in electronic structure

calculations: How it should work and how it can be made to work. J. Chem. Phys. 122,

014107 (2005).

35. Feller, D., Peterson, K. A. & Crawford, T. D. Sources of error in electronic structure

calculations on small chemical systems. J. Chem. Phys. 124, 054107 (2006).

36. Irk Bakowies, D. Accurate extrapolation of electron correlation energies from small basis

sets. J. Chem. Phys. 127, 164109 (2007).

37. Barnes, E. C., Petersson, G. A., Feller, D. & Peterson, K. A. The CCSD(T) complete

basis set limit for Ne revisited. J. Chem. Phys. 129, 194115 (2008).

38. Feller, D., Peterson, K. A. & Hill, J. G. On the effectiveness of CCSD(T) complete basis

set extrapolations for atomization energies. J. Chem. Phys. 135, 044102 (2011).

39. Feller, D. Benchmarks of improved complete basis set extrapolation schemes designed for

standard CCSD(T) atomization energies. J. Chem. Phys. 138, 074103 (2013).

40. Ranasinghe, D. S. & Petersson, G. A. CCSD(T)/CBS atomic and molecular benchmarks

for H through Ar. J. Chem. Phys. 138, 144104 (2013).

159

BIBLIOGRAPHY

41. Pansini, F. & Varandas, A. Toward a unified single-parameter extrapolation scheme for

the correlation energy: Systems formed by atoms of hydrogen through neon. Chem. Phys.

Lett. 631, 70. issn: 0009-2614 (2015).

42. Jurecka, P., Sponer, J., Cerny, J. & Hobza, P. Benchmark database of accurate (MP2

and CCSD(T) complete basis set limit) interaction energies of small model complexes,

DNA base pairs, and amino acid pairs. Phys. Chem. Chem. Phys. 8, 1985 (2006).

43. Goerigk, L., Karton, A., Martin, J. M. L. & Radom, L. Accurate quantum chemical

energies for tetrapeptide conformations: why MP2 data with an insufficient basis set

should be handled with caution. Phys. Chem. Chem. Phys. 15, 7028 (2013).

44. Liakos, D. G. & Neese, F. Improved Correlation Energy Extrapolation Schemes Based on

Local Pair Natural Orbital Methods. J. Phys. Chem. A 116, 4801 (2012).

45. Rina Brauer, B., anoj K. . Kesharwani, M. & an M. . L. . Martin, J. Some Observa-

tions on Counterpoise Corrections for Explicitly Correlated Calculations on Noncovalent

Interactions. J. Chem. Theory Comput. 10, 3791 (2014).

46. Karton, A., Yu, L.-J., Kesharwani, M. K. & Martin, J. M. L. Heats of formation of the

amino acids re-examined by means of W1-F12 and W2-F12 theories. Theoretical Chem.

Accounts 133, 1483 (2014).

47. He, X., Fusti-Molnar, L. & Merz, K. M. J. Accurate Benchmark Calculations on the

Gas-Phase Basicities of Small Molecules. J. Phys. Chem. A 113, 10096 (2009).

48. Mir Karton, A. & ars Goerigk, L. Accurate Reaction Barrier Heights of Pericyclic

Reactions: Surprisingly Large Deviations for the CBS-QB3 Composite Method and Their

Consequences in DFT Benchmark Studies. J. Computational Chem. 36, 622 (2015).

49. Werner, H.-J. et al. MOLPRO, version 2012.1, a package of ab initio programs see. 2012.

50. Werner, H.-J., Knowles, P. J., Knizia, G., Manby, F. R. & Schuetz, M. Molpro: a general-

purpose quantum chemistry program package. Wiley Interdiscip. Rev. Comput. Mol. Sci.

2, 242 (2012).

51. Martin, J. M. L. & de Oliveira, G. Towards standard methods for benchmark quality ab

initio thermochemistry - W1 and W2 theory. J. Chem. Phys. 111, 1843 (1999).

52. Boese, A. D. et al. W3 theory: Robust computational thermochemistry in the kJ/mol

accuracy range. J. Chem. Phys. 120, 4129 (2004).

53. Karton, A. & Martin, J. M. L. Explicitly correlated Wn theory: W1-F12 and W2-F12. J.

Chem. Phys. 136, 124114 (2012).

54. Curtiss, L. A., Redfern, P. C., Raghavachari, K., Rassolov, V. & Pople, J. A. Gaussian-3

theory using reduced Moller-Plesset order. J. Chem. Phys. 110, 4703 (1999).

160

BIBLIOGRAPHY

55. Curtiss, L. A., Redfern, P. C. & Raghavachari, K. Gaussian-4 theory using reduced order

perturbation theory. J. Chem. Phys. 127, 124105 (2007).

56. Curtiss, L. A., Redfern, P. C. & Raghavachari, K. Gn theory. Wiley Interdiscip. Rev.

Comput. Mol. Sci. 1, 810 (2011).

161

BIBLIOGRAPHY

162

Part IV

Conclusion

163

Chapter 8

Concluding remarks and future

prospects

This thesis focussed on the development, assessment, and application of cheminformatics and

quantum chemical methods for a range of important problems in physical chemistry. The

research presented here has covered a range of topics: molecular shape and volume in crystal

structures along with their description using spherical harmonics, approximating electrostatic

interactions, assessing the accuracy of a technique for quantum chemical prediction of electron

impact mass spectra, the basis-set convergence on CCSD(T) geometry optimisations, and

different methods for basis-set extrapolation for single-point energies when using CCSD and

MP2. This chapter will summarise the findings, successes and limitations, and summarise future

plans and possibilities involving this research.

8.1 Molecular shape

Part II of this thesis detailed the development of novel molecular shape descriptors in the crystal

environment, though the broader ideas of the technique need not be limited to crystal structures.

It was demonstrated that molecular surfaces (in this case, the Hirshfeld and promolecule surfaces)

could be accurately described via a spherical harmonic transform, with roughly 2% error for

typical surfaces at the level lmax = 20. This degree of accuracy is sufficient to be confident that

we are genuinely describing the shape and/or properties to a degree that can be utilised in

large-scale quantitative studies. By utilising rotation invariant descriptors based on the spherical

harmonic coefficients that result from the transform, we avoid the need to orient shapes in order

to determine similarity. It was demonstrated in Chapter 3 that these shape descriptors were

sufficient to classify metallic crystal structures according to their (an)isotropy, and to cluster

different molecular crystals based on their chemical scaffolds (i.e. structural similarity). Further,

in Chapter 4 these descriptors were modified to incorporate properties calculated on the surfaces

e.g. dnorm, electrostatic potential. The combined shape-property descriptors were utilised to

165

CHAPTER 8. CONCLUDING REMARKS AND FUTURE PROSPECTS

search through a subset of the Cambridge Structural Database (CSD) in order to find suitable

ligands as part of ligand-based virtual screening processes. The efficiency of these descriptors

was proven here, where finding the top n matches out of a set of approx. 140 000 molecules was

effectively instant (< 1 ms). The results of this study demonstrated the potential application of

these shape descriptors in ligand-based virtual screening (LBVS) processes, but also showed

the potential value of the CSD as a small-molecule database of potential ligands – something

largely overlooked in the drug discovery literature.

This straightforward vector of real numbers, each representing the magnitudes of different

angular momenta (l) spherical harmonic functions produced results exceeding our expectations.

It was particularly surprising how well these descriptors performed when examining large drug

ligands, such as the HIV-protease complex (PDB code 3A2O), where multiple existing antiviral

drugs (some targeting HIV) were found in the top 100 matches.

When initially developing this technique, the only goal was to find a method to incorporate

some description of a crystal structure involving molecular shape and interactions into quantita-

tive studies, but it seems the technique could be used as a bona fide step in LBVS procedures.

This being said, there is much that could be explored or optimised when using these descriptors.

Firstly, it remains to be explored to what maximum angular momenta (lmax) is required

for different purposes. Reducing this degree not only reduces computational time for the

initial description calculation, it further reduces the dimensionality of the dataset. Second, the

same shape-property descriptors could be applied to difference surfaces (e.g. CPK or other

surfaces which exist outside crystal environments) which may not only significantly speedup

the description process, but allow the description of molecules in the gas phase. The surface

property being described has much room for improvement; while dnorm is cheap to calculate,

incorporating some scalar function describing pharmacophores, hydrogen-bond donors/acceptors

or other typical notions used in drug discovery may prove incredibly useful, and as yet is

entirely unexplored. Finally, incorporation of these techniques with existing software (e.g.

CrystalExplorer, or the software suite accompanying the CSD) could prove incredibly useful to

researchers.

In the process of this research, software to calculate these shape descriptors, and find matches

for given molecules was written, which will be publicly available upon publication of the work

in Chapter 4. Further, a novel linear scaling method to calculate approximate electrostatic

potentials from a wavefunction on large grids (such as Hirshfeld surfaces) was developed using

a combination of the exact potential and atomic multipoles. This method should be explored

further and characterised completely, as the core of the idea (describing most of the electron

density exactly, and the rest with atomic multipoles) has more applications than just ESP and

could provide accurate, if approximate, results for a wide variety of electronic properties.

In summary, the technique to describe molecular shape presented in Chapters 3 and 4 proved

effective beyond expectations, and represents a fast, efficient and accurate way to incorporate

166

8.2. ASSESSMENT OF QUANTUM CHEMICAL PREDICTIONS

molecular shape into large scale studies of crystal structures. Further, this work is likely the first

use of the complex spherical harmonic transform to jointly describe molecular shape coupled

with a surface property, which provides a concise and efficient framework for a wide range of

possible chemical descriptors.

8.2 Assessment of quantum chemical predictions

Part III explored the accuracy of state-of-the-art quantum chemical predictions for thermochem-

istry, equilibrium geometries and electron impact mass spectra.

The capacity to accurately predict EIMS would be extremely valuable in some chemical

characterisation problems, particularly areas like de novo structure elucidation of natural

products. To that end, Chapter 5 assessed the performance of Grimme and Bauer’s QCEIMS

method against experimental MSp and compared it with the machine learning method CFM-EI.

It was found that QCEIMS overall outperformed CFM-EI, which was ‘trained’ on a wide variety

of metabolites and their experimental MS. It is immensely promising that such an ab initio

technique is capable of predicting, with a high degree of accuracy, the EIMS of a variety of

alcohols, ketones, phenols, aldehydes and other common classes of compounds encountered in de

novo structure elucidation. QCEIMS succeeded in matching the correct spectra in the first or

second rank from a field of isomers in more than half of the cases for our test set of 61 molecules.

Further, it was established that the counting statistics errors on the peaks produced by

QCEIMS were much smaller than the experimental reproducibility error in the mass spectrum;

and that these predicted spectra are often reproduced better than the corresponding rankings

relative to isomers. Importantly, this demonstrates that the QCEIMS method achieves its

success not simply due to accurate matching in the distribution of peak heights, but rather that

the peak heights and the distribution of the peak positions as a function of mass-to-charge ratio

(m/z) is sufficiently different for different isomers to achieve a high ranking.

One of the difficult barriers to successful EIMS prediction is that it is not simply a case

of predicting fragmentations, but frequently involves rearrangements or reactions within the

molecule in question. We found that both QCEIMS and CFM-EI tended to suffer from the

same problems: underestimation or failure to predict specific rearrangements (e.g. McLafferty

rearragements), and overestimation of the relative abundance of the molecular ion. These

failures are especially salient given the radically different approaches in both techniques. It may

be that this indicates that reassessment of some of the fundamental abstractions involved in

EIMS predictions must be re-evaluated for success. In the case of QCEIMS, it might be that

different Hamiltonians could improve the prediction (i.e. using quantum chemical methods of

higher accuracy to determine the Hamiltonian)– something which was not explored in the work

in Chapter 5.

Future work in this area would ideally involve working in concert with a genuine structure

167

CHAPTER 8. CONCLUDING REMARKS AND FUTURE PROSPECTS

elucidation problem: i.e. attempting to predict which structural isomer was observed through

its MSp without synthesizing the compounds, and confirming the accuracy of the prediction.

Such a study would not only prove the utility of these techniques, but showcase the exciting

intersection of quantum chemical predictions with experimental questions. QCEIMS (and CFM-

EI) may indeed prove valuable in the lab, particularly if accuracy of the predictions is known

and understood, so that it may save significant labour and synthesis time over synthesising all

possible structural isomers and recording their MSp;

Chapter 6 primarily focussed on the basis set convergence of equilibrium bond-lengths

optimized at the CCSD(T) level of theory. High accuracy thermochemistry requires highly

accurate geometries, and it was found that at least a quadruple-zeta basis is required for

convergence within 0.002A. Every increase in the highest angular momenta present in the basis

set (i.e. double-, triple- quadruple-zeta etc.) reduced the RMSD over bond distances by a factor

of ∼4 Similar results were found for both the A′VXZ basis sets and the def2-nZVPP basis sets.

Assessment of the explicitly correlated CCSD(T)-F12 calculations coupled with cc-pVnZ-F12

optimised basis sets demonstrated accelerated basis set convergence, with calculations at the

nZ level providing roughly equivalent accuracy to those at the n + 2Z level using CCSD(T).

CCSD(T)-F12/cc-pVDZ, then, provides an impressive perfomance for its computational cost,

and is recommended for use when highly accurate reference geometries are required. Due to the

systematic errors in conventional CCSD(T) calculations, it was found that simple scaling of

bond lengths was an effective method to improve performance at no additional cost. Double-

and triple-zeta basis sets significantly overestimate bond lengths, with RMSDs of 0.0196A and

0.0050A respectively, and these were reduced by roughly 60% upon scaling. As such, in cases

where explicitly correlated CCSD(T)-F12/VDZ-F12 are too expensive, it is recommended to

use scaled CCSD(T)/VDZ bond distances. The energetic implications of these errors in bond

distances were characterised, and the difference between CCSD(T)/VDZ and CCSD(T)/A′V6Z

geometries led to significant errors with an RMSD of over 2.4 kJ mol−1 for composite ab initio

theories like W1-F12 and W2-F12. This error is reduced to only 0.02 kJ mol−1 when using either

a QZ basis or CCSD(T)-F12/VDZ-F12, and since the latter is roughly 4 times computationally

cheaper it should be utilised when available.

For the future, similar studies should be performed on DFT and MP2, in order characterise

both the basis-set convergence of their predicted equilbrium geometries and the corresponding

energetic errors. This is especially important, as both DFT and MP2 are applicable to much

larger molecules, and as such are far more common in practical geometry optimizations in

computational studies. Further, it would be interesting to examine similar bond scaling

techniques for these methods as compared to CCSD(T) geometries. Clearly not all functionals

will perform in the same way, but if some DFT or MP2 techniques can accurately model the

potential energy surface, then they should have the capacity to identify roughly the correct

conformations and bond scaling methods may be readily applicable.

168

8.2. ASSESSMENT OF QUANTUM CHEMICAL PREDICTIONS

Finally, Chapter 7 evaluated the performance of a number of different approaches for

obtaining CCSD correlation energies close to the CBS limit while only using comparatively small

double- and triple-zeta basis sets. Both global (system-independent) and system-dependent

extrapolations based on the A+B/Lα two-point extrapolation formula were examined, along

with the well known additivity approach using an MP2-based basis set correction term. It was

found that, perhaps unsurprisingly, basis set convergence varied significantly between different

systems. In particular, systems with polar bonds and elements from the second row of the

periodic table proved to have much slower convergence when using CCSD or MP2 at the double-

or triple-zeta level of theory.

As such, the system dependent basis set extrapolation technique, where we obtain unique

extraploation exponents for each system using lower cost MP2 calculations significantly improves

convergence to the CBS limit over global extrapolation coefficients. This was demonstrated

through RMSDs against reference CBS calculations over the W4-11 database: 9.1 kJ mol−1

for global extrapolation, 3.7 kJ mol−1 using system dependent extrapolations and 2.4 kJ mol−1

when using the additivity scheme. Similar results were found over a set of 20 basis-set limit

CCSD calculations on larger molecules, including amino acids, DNA/RNA bases, aromatics and

platonic hydrocarbon cages. Overall this demonstrates that utilising lower cost MP2 calculations

can be used effectively to facilitate rapid predictions of CBS energies at the CCSD level of

theory, which generally converge very slowly with an increase in the size of the one-particle

basis sets.

Importantly, the additivity scheme, which uses an MP2-based basis set correction term was

found to significanly outperform system dependent extrapolation, resulting in RMSDs of 2.4

(over the W4-11 dataset) and 2.9 (over the TAE20 dataset) kJ mol−1.

In the future, it would be prudent to explore these schemes for explicitly correlated F12

methods, as these generally provide significant improvements in performance over conventional

CCSD(T) methods for the same cost.

In short, when it comes to CC calculations use of the CCSD(T)-F12 methods with a double-

zeta basis set provide highly accurate reference energies and geometries, but where these are

not available CCSD(T) with double-zeta may be used in concert with extrapolation (energies)

or scaling (bond lengths) to achieve similar results.

This thesis has explored both the development and application of a variety of techniques

applied to physical chemistry problems (crystal structures, electron impact mass spectra). In

doing so, the focus has been on the tools to advance scientific enquiry – something which

increasingly constitutes a large portion of the time (and money) spent on scientific research.

There is little doubt that the increasing complexity of the problems we face, and the tools we

use to solve them, is likely to necessitate more and more computer literacy among scientists. For

scientific research to be successful, then, those who write software (i.e. create the tools) must

bring people along for the ride (i.e. make their techniques more accessible and understandable),

169

CHAPTER 8. CONCLUDING REMARKS AND FUTURE PROSPECTS

and those who use these tools must continue to endeavour to understand enough about what

they’re using to not only justify their reasoning, but to find new ways to use existing tools –

both of which are essential to the topics of this body of work.

170

Appendices

171

Appendix A

Journal Article: Approaching an

experimental electron density model of

the biologically active

trans-epoxysuccinyl amide

group—Substituent effects vs. crystal

packing

173

Title Approaching an experimental electron density model of the

biologically active trans-epoxysuccinyl amide group –

Substituent effects vs. crystal packing

Authors Ming W. Shi, Scott G. Stewart, Alexandre N. Sobolev,

Birger Dittrich, Tanja Schirmeister, Peter Luger,

Malte Hesse, Yu-Sheng Chen, Peter R. Spackman,

Mark A. Spackman, Simon Grabowsky

Journal Journal of Physical Organic Chemistry

DOI 10.1002/poc.3683

Date of publication 24 January 2017

175

APPENDIX A. JOURNAL ARTICLE: APPROACHING AN EXPERIMENTALELECTRON DENSITY MODEL OF THE BIOLOGICALLY ACTIVETRANS-EPOXYSUCCINYL AMIDE GROUP—SUBSTITUENT EFFECTS VS.CRYSTAL PACKING

176

Appendix B

Journal Article: Intermolecular

interactions in molecular crystals:

what’s in a name?

177

Title Intermolecular interactions in molecular crystals: what’s in a name?

Authors Alison J. Edwards, Campbell F. Mackenzie, Peter R. Spackman,

Dylan Jayatilaka and Mark A. Spackman

Journal Faraday Discuss.

DOI 10.1039/C7FD00072C

Date of publication 08 March 2017

179

APPENDIX B. JOURNAL ARTICLE: INTERMOLECULAR INTERACTIONSIN MOLECULAR CRYSTALS: WHAT’S IN A NAME?

180

Appendix C

Journal Article: CrystalExplorer model

energies and energy frameworks:

extension to metal coordination

compounds, organic salts, solvates and

open-shell systems

181

research papers

IUCrJ (2017). 4 https://doi.org/10.1107/S205225251700848X 1 of 13

IUCrJISSN 2052-2525

CHEMISTRYjCRYSTENG

Received 12 April 2017

Accepted 7 June 2017

Edited by C. Lecomte, Universite de Lorraine,

France

Keywords: CrystalExplorer; model energies;

energy frameworks; coordination compounds;

open-shell systems; intermolecular interactions.

Supporting information: this article has

supporting information at www.iucrj.org

CrystalExplorer model energies and energy frame-works: extension to metal coordination compounds,organic salts, solvates and open-shell systems

Campbell F. Mackenzie, Peter R. Spackman, Dylan Jayatilaka and Mark A.Spackman*

School of Molecular Sciences, University of Western Australia, Perth, 6009, Australia. *Correspondence e-mail:

[email protected]

The application domain of accurate and efficient CE-B3LYP and CE-HF modelenergies for intermolecular interactions in molecular crystals is extended bycalibration against density functional results for 1794 molecule/ion pairsextracted from 171 crystal structures. The mean absolute deviation of CE-B3LYP model energies from DFT values is a modest 2.4 kJ mol!1 for pairwiseenergies that span a range of 3.75 MJ mol!1. The new sets of scale factorsdetermined by fitting to counterpoise-corrected DFT calculations result inminimal changes from previous energy values. Coupled with the use of separatepolarizabilities for interactions involving monatomic ions, these model energiescan now be applied with confidence to a vast number of molecular crystals.Energy frameworks have been enhanced to represent the destabilizinginteractions that are important for molecules with large dipole moments andorganic salts. Applications to a variety of molecular crystals are presented indetail to highlight the utility and promise of these tools.

1. Introduction

The detailed analysis of the interactions between moleculesand ions in crystals plays an increasingly important role inmodern solid-state chemistry and, in particular, crystal engi-neering, where the derivation of predictive structure propertyrelationships is key to a genuine ‘engineering’ of crystals. This,of course, was articulated some time ago by Desiraju (1989),who described crystal engineering as the ‘understanding ofintermolecular interactions in the context of crystal packingand the utilization of such understanding in the design of newsolids with desired physical and chemical properties’. Utili-zation and design require understanding as an essentialprecursor, and the context of crystal packing in this statementis especially important. Recent years have witnessed a rapidgrowth of publications that focus on the relative importance ofnoncanonical interactions (e.g. halogen, chalcogen, pnicogenand tetrel bonds), but it is not obvious that they haveenhanced our understanding of the relationship between thestructure of molecules (geometric and electronic), the crystalstructures they form and their consequent chemical andphysical properties. It seems pertinent to ask whether we areconverging on the requisite intimate, and ultimately useful,understanding of why molecules and ions are arranged incrystals as observed, or merely cataloguing an increasingnumber of examples of relatively weak intermolecular ‘inter-actions’ while ignoring their common origins?

The recent series of essays published in this journal (Dunitz,2015; Lecomte et al., 2015; Thakur et al., 2015) has highlighted

Title CrystalExplorer model energies and energy frameworks: extension

to metal coordination compounds, organic salts, solvates and

open-shell systems

Authors Campbell F. Mackenzie, Peter R. Spackman, Dylan Jayatilaka

and Mark A. Spackman*

Journal IUCrJ

DOI 10.1107/S205225251700848X

Date of publication 01 September 2017

183

APPENDIX C. JOURNAL ARTICLE: CRYSTALEXPLORER MODELENERGIES AND ENERGY FRAMEWORKS: EXTENSION TO METALCOORDINATION COMPOUNDS, ORGANIC SALTS, SOLVATES ANDOPEN-SHELL SYSTEMS

184

Appendix D

Description of CFM-EI and QCEIMS

methods for prediction of mass spectra

D.1 Competitive Fragmentation Modeling (CFM)

CFM models the fragmentation processes within a mass spectrometer in a generative, probablistic

manner. Their model consists of a stochastic Markov process made up of transitions between

discrete fragment states F0, F1, . . . , Fd that each take values from the set of all possible

fragments.

The set of all possible fragments is enumerated by a combinatorial approach, based on

systematic discconection of bonds, [1] which involves breaking every bond in the molecule (and

every pair in each ring) in turn, while considering all possible hydrogen rearrangements within

each pair of resulting fragments.

The transition between fragments fi and fj is assigned a break tendency value θij = wTΦij ∈R, where Φij is a vector of ‘chemical features’ which describe the possible transition (e.g. atoms

either side of the broken bond) and w is a vector of parameters corresponding to those features.

The approximate probability, then, that fi transitions to fj at a single time step is defined as:

ρ(fi, fj) =

exp θij

1 +∑k

exp θikwhere fi 6= fj

1

1 +∑k

exp θikwhere fi = fj

(D.1)

This softmax function models the competition between different possible fragmentation

events which originate from the same parent fragment fi, and it ensures all probablities sum to

unity. The second part of eq. D.1 shows that self-transitions are permitted, but are essentially

185

APPENDIX D. DESCRIPTION OF CFM-EI AND QCEIMS METHODS FORPREDICTION OF MASS SPECTRA

assigned a tendency θii of zero.

Isotope distribution has an important impact on EIMS – helping decide between alternative

explanations for the same peak. Within the CFM model, isotopic peaks are incorporated using

the weighted sum of Gaussians corresponding to the peaks in the fragment’s expected distribution

of isotopes. The expected isotopic spectrum for a fragment fi is denoted as E(Sfi) : {(m′, h′)}where {(m′, h′)} are a set of normalized mass and intensity pairs. CFM uses the program

emass[emass] to compute the expected isotope spectrum for a given molecular formula, with

the result thresholded to only include peaks with a normalized intensity above 0.01.

The probability, considering isotopic distrubutions, that a peak of mass m is produced can

be modeled by a random variable P , where P = m|Fd in this case is expressed as a weighted

sum of Gaussians with variance σ determined by the mass tolerance of the instrument and a

mean given by the mass Fd:

Pr(P = m|Fd) ≈ g(m,Fd;σ) =∑

(m′,h′)∈E(SFd)

h′

σ√

2πexp

{− 1

2

(m−m′)

σ

)2}(D.2)

The predicted MSp is then computed as the marginal of P given the input molecule F0:

Pr(P = m|F0;w) ≈∑F1

· · ·∑Fd

ρ(F0, F1) · · · ρ(Fd−1, Fd)g(m,Fd;σ) (D.3)

The machine learning aspect of CFM, then, primarily lies in training the model parameters

governing the calculation of the break tendency θij from the chemical feature vector describing

the break Φij. In the case of CFM-EI, θij is taken to be the output of a multilayer perceptron

(i.e. this relationship is modeled using a neural network) whose inputs are Φij and model

parameters are denoted by w. These parameters are estimated using expectation minimisation

(i.e. maximising the expected log-likelihood Q), and have been trained on a subset of the

NIST/EPA/NIH Mass Spectral Library (see Allen et al. [2] and Allen et al. [3] for complete

details).

D.2 Quantum Chemical prediction of Electronic Impact

Mass-spectra (QCEIMS)

QCEIMS is a procedure combining so-called ‘Ab-Initio Molecular Dynamics‘ (AIMD) with

stochastic and statistical elements in order to predict accurate EIMS without regard to pre-

conceived decomposition pathways. The program itself is coupled to several external pieces of

electronic structure software (e.g. MOPAC, ORCA, TURBOMOLE) allowing the internal MD

186

D.2. QUANTUM CHEMICAL PREDICTION OF ELECTRONIC IMPACTMASS-SPECTRA (QCEIMS)

procedure to be calculated with a variety of semi-empirical methods (e.g. DFTB3, OM2, PM3

and PM6) along with DFT functionals. For a detailed description of the methods involved, see

Grimme [4], Grimme & Bauer [5], Bauer & Grimme [6, 7], and Asgeirsson et al. [8].

Broadly, the QCEIMS procedure is made up of three steps: (1) equilibration and sampling of

potential neutral conformers, (2) calculation of the MO spectrum and (3) so-called ‘production

runs’. The first and last steps involve MD, where the neutral molecule (1) or cation (3) are

propagated in time by numerically integrating Newton’s equations of motion using the leap-frog

algorithm. Each time step (0.5 fs) the atomic forces are calculated on the fly using the requested

ab-initio method (i.e. semi-empirical or DFT).

(1) Equilibration and sampling The neutral input molecule is equlibrated over a short period

of time (e.g. 12.5 ps) in the canonical ensemble (NVT), with a constant temperature (default

500 K). The equilibration is followed by a conformer sampling in the micro-canonical ensemble

(NVE), where samples (1000s) of molecular geometry and nuclear velocities are randomly taken

and saved along a longer (e.g. 25.0 ps) trajectory.

(2) Molecular orbital spectrum A single point energy calculation is then performed (using

the specified semi-empirical or DFT method) in order to determine the MO spectrum, followed

by an MO resolved Mulliken population analysis. This calculation allows estimation of internal

excess energy (IEE), internal conversion (IC) time and MO-population derived nuclear velocity

scaling factors.

The IEE represents energy given to the molecule from the collision with the electron, and is

distributed among the vibrational modes of the molecular ion, by scaling the nuclear velocities

i.e. heating. The value of IEE for each run is chosen stochastically, and is assumed to be a

Poisson-type random variable:

P (E) =exp (cE(1 + ln(b/cE))− b)√

aE + 1(D.4)

where P(E) is the probability to have an IEE equal to E, a = 0.2 eV, b = 1.0 and c =1

aNel

. The

maximum value of IEE is Eimpact − εHOMO, where the former is an input parameter representing

the kinetic energy of the free electron (e.g. 70 eV in most EI experiments), and the IEE

distribution is set to have its mode at 0.6 eV per atom.

IC represents the time interval in which the ion is heated, and after the IC process all of the

IEE is converted into nuclear kinetic energy. The IC is calculated as follows from the energy-gap

law, depending on relative energies of the MOs:

tIC =M∑j>i

kh

Nel

exp (α(εi − εj)) (D.5)

where α = 0.5 eV−1 and kh = 2 ps, M is the ordinal number of the HOMO and εi is the orbital

187

APPENDIX D. DESCRIPTION OF CFM-EI AND QCEIMS METHODS FORPREDICTION OF MASS SPECTRA

energy of the i-th MO. For molecules with less than 35 atoms, MO-based velocity scaling factors

are used – proportional to the Mulliken population of that nucleus in the ionized MO.

(3) Production runs Here the sampled conformers are ionized (valence) and the coordinates

and nuclear velocities are used as initial conditions for the molecular ion in the production runs.

As each run is independent, they can be performed entirely in parallel, and a large number of

runs are performed for each molecule in order to ensure statistical convergence of the resulting

EIMS with regard to the observed fragments.

Within each run, the molecular ion is heated by scaling the nuclear velocities (as described in

the previous paragraph) in the first few picoseconds (IC time) of the simulation. Conceptually,

this corresponds to the notion that after an electron impact and electronically excited ion

will form, then relax to an excited level of the electronic ground state through IC followed

by intramolecular vibrational redistribution (IVR). Further propagation can then result in

decomposition of the molecular ion to favourable (possibly radical) neutral and charged moieties.

If fragmentation occurs, then the program evaluates the vertical IP of each fragmentation

product via SCF calculation. The statistical weight of each product is then given by:

Ci =exp

(−∆ESCFi

kBTAv

)M∑j

exp(−∆ESCFj

kBTAv

) (D.6)

where M is the number of fragments, TAv is the average temperature of the fragments and kB

is the Boltzmann constant. The product of this statistical weight Ci and the total molecular

charge yields the statistical charge of fragment i.

The fragment with the highest statistical charge is then propagated further in a so-called

‘cascade’, while other fragments are counted and stored. This allows the highest charged

secondary fragments to decompose further, introducing tertiary fragments where this is repeated

until fragmentation finishes. The sum of the statistical charges of all order fragments within

one run will be equal to the total molecular charge (i.e. +1), thus the sum of the statistcal

charges for a specific fragment over the ensemble of production runs will yield the total relative

intensity of that particular fragment, which along with the introduction of isotope distributions

after the simulation yields a purely theoretical MSp.

188

Bibliography

1. Hill, A. W. & Mortishire-Smith, R. J. Automated assignment of high-resolution collisionally

activated dissociation mass spectra using a systematic bond disconnection approach. Rapid

Commun. Mass Spectrom. 19, 3111–3118. issn: 1097-0231 (2005).

2. Allen, F., Greiner, R. & Wishart, D. Competitive fragmentation modeling of ESI-MS/MS

spectra for putative metabolite identification. Metabolomics (2015).

3. Allen, F., Pon, A., Greiner, R. & Wishart, D. Computational Prediction of Electron

Ionization Mass Spectra to Assist in GC/MS Compound Identification. Anal. Chem. 88,

7689 (2016).

4. Grimme, S. Towards First Principles Calculation of Electron Impact Mass Spectra of

Molecules. Angew. Chem. Int. Ed. 52, 6306 (2013).

5. Grimme, S. & Bauer, C. A. First principles calculation of electron ionization mass spectra

for selected organic drug molecules. Org. Biomol. Chem. 12, 8737 (2014).

6. Bauer, C. A. & Grimme, S. Elucidation of Electron Ionization Induced Fragmentations of

Adenine by Semiempirical and Density Functional Molecular Dynamics. J. Phys. Chem.

A 118, 11479 (2014).

7. Bauer, C. A. & Grimme, S. How to Compute Electron Ionization Mass Spectra from First

Principles. J. Phys. Chem. A 120, 3755 (2016).

8. Asgeirsson, V., Bauer, C. A. & Grimme, S. Quantum chemical calculation of electron

ionization mass spectra for general organic and inorganic molecules. Chem. Sci. 52, 6306

(2017).

189

BIBLIOGRAPHY

190

Appendix E

Cartesian multipole expansion of

electrostatic interaction energy

The electrostatic interaction energy between two densities ρA and ρB is

U =

∫∫drAdrB ρA(rA)ρB(rB) T (R), (E.1)

T (R) = R−1, (E.2)

R = |R|, (E.3)

R = rA − rB. (E.4)

We are interested in the Taylor expansion of this energy around a fixed point A in system A,

and fixed point B in system B (this will be the multipole expansion) so it is better to make

these fixed points new origins. Then by a slight abuse of notation we get

U =

∫∫drAdrB ρA(rA)ρB(rB) T (R), (E.5)

R = |R|, (E.6)

R = a− b, (E.7)

a = rA −A, (E.8)

b = rB −B. (E.9)

191

APPENDIX E. CARTESIAN MULTIPOLE EXPANSION OFELECTROSTATIC INTERACTION ENERGY

We will require derivatives of T = T (rA, rB) with respect to rA and rB so we have

TAα ≡∂T

∂rAα=

∂T

∂Rα

∂Rα

∂rAα=

∂T

∂Rα

= Tα, (E.10)

TBα ≡∂T

∂rBα=

∂T

∂Rα

∂Rα

∂rBα= − ∂T

∂Rα

,= −Tα, where (E.11)

Tα =∂T

∂Rα

. (E.12)

Therefore all derivatives of T with respect to a or b may be replaced with signed derivatives

of T with respect to R. The result is easy to generalise to higher derivatives using an obvious

notation e.g.

TABBBαβγδ ≡ ∂4T

∂rAα∂rBβ∂rBγ∂rBδ= TBABBβαγδ = TBBABβγαδ = TBBBAβγδα = −Tαβγδ. (E.13)

i.e. there is a negative sign for terms with odd number of derivatives with respect to components

of rB. Using this observation we can write down the Taylor expansion of around U(rA, rB) as:

U =

∫∫drAdrB ρA(rA)ρB(rB) T (R), (E.14)

=

∫∫drAdrB ρA(rA)ρB(rB)

×[T + aαT

Aα + bαT

+1

2!(aαaβT

AAαβ + 2aαbβT

ABαβ + bαbβT

BBαβ )

+1

3!(aαaβaγT

AAAαβγ + 3aαaβbγT

AABαβγ + 3aαbβbγT

ABBαβγ + bαbβbγT

BBBαβγ )

+1

4!(aαaβaγaδT

AAAAαβγδ + 4aαaβaγbδT

AAABαβγδ + 6aαaβbγbδT

AABBαβγδ

+ 4aαbβbγbδTABBBαβγδ + bαbβbγbδT

BBBBαβγδ ) + . . .

](E.15)

In the above lines we have made use of the symmetry of the mixed derivatives of the T ’s, e.g.

for the mixed second derivative terms in the third line of the equation above we should have

written

aαbβTABαβ + bαaβT

BAαβ = aαbβT

ABαβ + bβaαT

BAβα

= aαbβTABαβ + aαbβT

BAβα

= aαbβTABαβ + aαbβT

ABαβ (E.16)

= 2aαbβTABαβ . (E.17)

192

In the first line above we changed the dummy indices α and β around, in the second line

we swapped the two terms around, and in the third line we used the symmetry of the mixed

derivatives, TBAβα = TABαβ . From this it is clear that only terms with different numbers T

derivatives with respect to A and B will appear, as we have shown. The coefficients in front of

the terms are binomial coefficients. To get the final multipole expansion we reduce all terms to

derivatives with respect to R,

U =

∫∫drAdrB ρA(rA)ρB(rB)

×[T + aαTα − bαTα

+1

2!(aαaβTαβ − 2aαbβTαβ + bαbβTαβ)

+1

3!(aαaβaγTαβγ − 3aαaβbγTαβγ + 3aαbβbγTαβγ − bαbβbγTαβγ)

+1

4!(aαaβaγaδTαβγδ − 4aαaβaγbδTαβγδ + 6aαaβbγbδTαβγδ

− 4aαbβbγbδTαβγδ + bαbβbγbδTαβγδ) + . . .]

= M0AM

0BT + (M1

AαM0B −M0

AM1Bα)Tα +

1

2!(M2

AαβM0B − 2M1

AαM1Bβ +M0

AM2Bαβ)Tαβ

+1

3!(M3

AαβγM0B − 3M2

AαβM1Bγ + 3M1

AαM2Bβγ −M0

AM3Bαβγ)Tαβγ

+1

4!(M4

AαβγδM0B − 4M3

AαβγM1Bδ + 6M2AαβM2

Bγδ − 4M1AαM

3Bβγδ +M0

BM4Bαβγδ)Tαβγδ + . . .

(E.18)

The multipole moments are easily defined by inspection.

The electrostatic potential φB(rA) of density ρB evaluated at a point rA = A + a is easily

read off from the above expansion: it is simply all the terms involving the monopole term M0A

with that term removed. Likewise the potential gradient (that is the electric field) at the same

point from the density B is given by all those terms with M1Aα appearing, with that moment

term removed.

This cartesian multipole expansion has been incorporated into the Tonto [163] software

package, and will be utilised in CrystalExplorer to calculate lattice energies, in addition to use

in electrostatic potential calculations.

193

APPENDIX E. CARTESIAN MULTIPOLE EXPANSION OFELECTROSTATIC INTERACTION ENERGY

194

Appendix F

Software: Simple Binary Format

F.1 Motivation

In scientific software, we frequently deal with numerical data; arrays, matrices, tensors etc. are

used for a wide variety of purposes.

All too often, scientific programmers rely on text (ASCII or UTF-8) representations of

numerical data on disk. Primarily this is an issue of convenience. While there are robust, feature

filled libraries for all kinds of data [1] for many tasks these may be overkill, or their installation

and incorporation into existing code might be too difficult a task when the focus is simply to

get the job done.

ASCII data is particularly bad. Data corruption

In order to address some of these concerns, I designed the Simple Binary Formt (SBF) file.

The goals were twofold: be straightforward to incorporate into existing software, and be as low

overhead as possible in reading and writing data. The SBF library supports the Fortran (2008),

C++, C and python programming languages.

A SBF file consists of a header, identifying that it is an SBF file along with the version

number it is compatible with along with how many datasets are stored within the file, and some

number of datasets. Each dataset can be considered ‘self-describing’: information about the

data type (see Table F.1), the shape of the data, whether it is little- or big-endian, along with

an identifier for each dataset is stored.

Because the C language Application Binary Interface is somewhat well defined across

platforms (and indeed most programming languages have some way of interoperating with C

code), C data types were chosen as the basis in order to define SBF data types.

195

APPENDIX F. SOFTWARE: SIMPLE BINARY FORMAT

SBF Type C/C++ declaration Fortran declaration Numpy declaration

sbf byte uint8 t integer(c uint8 t) "uint8"

sbf character char character(c char) "S1"

sbf integer int32 t integer(c int32 t) "int32"

sbf long int64 t integer(c int64 t) "int64"

sbf float float real(c float) "float32"

sbf double double real(c double) "float64"

sbf complex float complex float complex(c float) "complex64"

sbf complex double complex doubles complex(c float) "complex128"

Table F.1: Relationship between SBF data types and their respective declaration in differentlanguages

F.2 Performance

As the goal of this library was to provide a straightforward, easy to use interface between

multiple programming languages, not a great deal of effort has been expended on fine-grained

performance. However, using self-describing binary representations is still orders of magnitude

faster than writing textual numerical data. As an example, in the CrystalExplorer software,

typically roughly 50-60% of the calculation time for an isosurface was spent writing and Tonto

output files and reading them into the GUI. This time has been reduced to a matter of 100ms or

so (depending on disk speed). Further, the file size of these intermediates has been reduced by

an order of magnitude. From a user perspective, this has reduced the calculation time required

by a huge margin. Relatedly, the work performed in Chapter 4 of this thesis would have been

significantly more difficult with the old output formats – it required approximately 1 TB of

Hirshfeld and promolecular surface data in SBF files, which would have required approx. 10 TB

in the old file format.

The SBF library is freely licensed, and available at github.com/peterspackman/sbf.

196

Bibliography

1. The HDF Group. Hierarchical data format version 5 http://www.hdfgroup.org/HDF5.

197

BIBLIOGRAPHY

198

Complete Bibliography

1. Adamo, C. & Barone, V. Toward reliable density functional methods without adjustable

parameters: The PBE0 model. J. Chem. Phys. 110, 6158 (1999).

2. Adler, T. B., Knizia, G. & Werner, H.-J. A simple and efficient CCSD(T)-F12 approxi-

mation. J. Chem. Phys. 127, 221106 (2007).

3. Ahmed, F. R. & Cruickshank, D. W. J. A refinement of the crystal structure analyses of

oxalic acid dihydrate. Acta Crystallogr. 6, 385 (1953).

4. Allen, F. H. The Cambridge Structural Database: a quarter of a million crystal structures

and rising. Acta Crystallogr. B 58, 380 (2002).

5. Allen, F., Greiner, R. & Wishart, D. Competitive fragmentation modeling of ESI-MS/MS

spectra for putative metabolite identification. Metabolomics (2015).

6. Allen, F., Pon, A., Greiner, R. & Wishart, D. Computational Prediction of Electron

Ionization Mass Spectra to Assist in GC/MS Compound Identification. Anal. Chem. 88,

7689 (2016).

7. Allen, F. H. & Bruno, I. J. Bond lengths in organic and metal-organic compounds revisited:

X-H bond lengths from neutron diffraction data. Acta Crystallogr. B 66, 380 (2010).

8. Alsberg, B. K., Goodacre, R., Rowland, J. J. & Kell, D. B. Classification of pyrolysis

mass spectra by fuzzy multivariate rule induction-comparison with regression, K-nearest

neighbour, neural and decision-tree methods. Anal. Chim. Acta 348, 389 (1997).

9. Arputharaj, D. S., Hathwar, V. R., Guru Row, T. N. & Kumaradhas, P. Topological

Electron Density Analysis and Electrostatic Properties of Aspirin: An Experimental and

Theoretical Study. Crystal Growth & Design 12, 4357 (2012).

10. Asgeirsson, V., Bauer, C. A. & Grimme, S. Quantum chemical calculation of electron

ionization mass spectra for general organic and inorganic molecules. Chem. Sci. 52, 6306

(2017).

11. Bader, R. Atoms in Molecules: A Quantum Theory isbn: 9780198558651 (Clarendon

Press, Oxford, 1994).

199

COMPLETE BIBLIOGRAPHY

12. Bak, K. L. et al. The accurate determination of molecular equilibrium structures. J. Chem.

Phys. 114, 6548–6556 (2001).

13. Barnes, E. C., Petersson, G. A., Feller, D. & Peterson, K. A. The CCSD(T) complete

basis set limit for Ne revisited. J. Chem. Phys. 129, 194115 (2008).

14. Bartlett, R. J. & Musia l, M. Coupled-cluster theory in quantum chemistry. Rev. Mod.

Phys. 79 (2007).

15. Bauer, C. A. & Grimme, S. Elucidation of Electron Ionization Induced Fragmentations of

Adenine by Semiempirical and Density Functional Molecular Dynamics. J. Phys. Chem.

A 118, 11479 (2014).

16. Bauer, C. A. & Grimme, S. How to Compute Electron Ionization Mass Spectra from First

Principles. J. Phys. Chem. A 120, 3755 (2016).

17. Bayly, C. I., Cieplak, P., Cornell, W. & Kollman, P. A. A well-behaved electrostatic

potential based method using charge restraints for deriving atomic charges: the RESP

model. J. Phys. Chem. 97, 10269 (1993).

18. Becke, A. D. A multicenter numerical integration scheme for polyatomic molecules. J.

Chem. Phys. 88, 2547 (1988).

19. Becke, A. D. Density-functional exchange-energy approximation with correct asymptotic

behavior. Phys. Rev. A 38, 3098 (1988).

20. Becke, A. D. Density-functional thermochemistry. III. The role of exact exchange. J.

Chem. Phys. 98, 5648 (1993).

21. Becke, A. D. Density-functional thermochemistry. IV. A new dynamical correlation

functional and implications for exact-exchange mixing. J. Chem. Phys. 104, 1040 (1996).

22. Bemis, G. W. & Kuntz, I. D. A fast and efficient method for 2D and 3D molecular shape

description. J. Comput. Aided Mol. Des. 6, 607 (1992).

23. Bennett, J. M. & Kendrew, J. C. The computation of Fourier synthesis with a digital

electronic calculating machine. Acta Crystallogr. 5, 109 (1952).

24. Berman, H. M. The Protein Data Bank. Nucleic acids research 28, 235 (2000).

25. Biedenharn, L. C. & Louck, J. D. Angular momentum in quantum physics: theory and

application 307 (Addison-Wesley, 1981).

26. Binkley, J. S., Pople, J. A. & Hehre, W. J. Self-consistent molecular orbital methods. 21.

Small split-valence basis sets for first-row elements. J. Am. Chem. Soc. 102, 939 (1980).

27. Blatov, V. A., Shevchenko, A. P. & Proserpio, D. M. Applied Topological Analysis of

Crystal Structures with the Program Package ToposPro. Cryst. Growth Des. 14, 3576

(2014).

200

COMPLETE BIBLIOGRAPHY

28. Boese, A. D. et al. W3 theory: Robust computational thermochemistry in the kJ/mol

accuracy range. J. Chem. Phys. 120, 4129 (2004).

29. Bohl, C. E., Gao, W., Miller, D. D., Bell, C. E. & Dalton, J. T. Structural basis for

antagonism and resistance of bicalutamide in prostate cancer. Proceedings of the National

Academy of Sciences of the United States of America 102, 6201 (2005).

30. Bohman, B., Flematti, G. R. & Barrow, R. A. Identification of hydroxymethylpyrazines

using mass spectrometry. J. Mass Spectrom. 50, 987 (2015).

31. Bohman, B., Phillips, R. D., Flematti, G. R., Barrow, R. A. & Peakall, R. The chemical

web of a spider orchid - Sulfurous semiochemicals seduce male wasp pollinator. Angew.

Chem. Int. Ed. (2017).

32. Bohman, B. et al. Discovery of pyrazines as pollinator sex pheromones and orchid

semiochemicals: implications for the evolution of sexual deception. New Phytol. 203, 939

(2014).

33. Bomble, Y. J. et al. High-accuracy extrapolated ab initio thermochemistry. II. Minor

improvements to the protocol and a vital simplification. J. Chem. Phys. 125, 064108

(2006).

34. Bonchev, D. & Rouvray, D. H. Chemical Topology: Introduction and Fundamentals (CRC

Press, Boca Raton, Florida).

35. Born, M. & Oppenheimer, R. Zur Quantentheorie der Molekeln. Ann. Phys. (Berl.) 389,

457 (1927).

36. Bouslimani, A., Sanchez, L. M., Garg, N. & Dorrestein, P. C. Mass spectrometry of

natural products: current, emerging and future technologies. Nat. Prod. Rep. 31, 718

(2014).

37. Boys, S. F. Electronic wave functions - I. A general method of calculation for the stationary

states of any molecular system. Proc. Royal Soc. A 200, 542 (1950).

38. Bragg, W. H. & Bragg, W. L. The Reflection of X-rays by Crystals. Proc. Royal Soc. A

88, 428 (1913).

39. Breitling, R., Ritchie, S., Goodenowe, D., Stewart, M. L. & Barrett, M. P. Ab initio predic-

tion of metabolic networks using Fourier transform mass spectrometry data. Metabolomics

2, 155 (2006).

40. Bridgman, P. W. The Logic of Modern Physics (Macmillan, New York, 1927).

41. Brown, I. D. & McMahon, B. CIF: the computer language of crystallography. Acta

Crystallogr. B 58, 317 (2002).

42. Bruno, I. J. et al. New software for searching the Cambridge Structural Database and

visualizing crystal structures. Acta Crystallogr. B 58, 389 (2002).

201

COMPLETE BIBLIOGRAPHY

43. Buchanan, B. G., Lederberg, J. & Feigenbaum, E. The Heuristic DENDRAL program for

explaining empirical data 1971.

44. Buchbauer, G., Klinsky, A., Weiß-Greiler, P. & Wolschann, P. Ab initio Molecular

Electrostatic Potential Grid Maps for Quantitative Similarity Calculations of Organic

Compounds. Mol. Model. Ann. 6, 425 (2000).

45. Burel, G. & Henocq, H. 3-dimensional invariants and their application to object recognition.

Sig. Proc. 45, 1 (1995).

46. Calhoun, C. J. The Microcomputer Revolution? Sociol. Methods Res. 9, 397 (1981).

47. Campello, R. J. G. B., Moulavi, D., Zimek, A. & Sander, J. Hierarchical Density Estimates

for Data Clustering, Visualization, and Outlier Detection. ACM Trans. Knowl. Discov.

Data 10, 5:1. issn: 1556-4681 (2015).

48. Case, D. A. et al. Amber 14 (2014).

49. Cassam-Chenaı, P. & Jayatilaka, D. Some fundamental problems with zero flux partitioning

of electron densities. Theor. Chem. Acc. 104, 213–218 (3 2000).

50. Cassam-Chenaı, P. & Jayatilaka, D. A complement to “Some fundamental problems with

zero flux partitioning of electron densities”. Theor. Chem. Acc. 107, 383–384 (6 2002).

51. Champness, J. N. et al. Exploring the active site of herpes simplex virus type-1 thymidine

kinase by X-ray crystallography of complexes with aciclovir and other ligands. Proteins

32, 350 (1998).

52. ChemSpider Home Page https://www.chemspider.com (2017).

53. Cheng, T., Li, Q., Zhou, Z., Wang, Y. & Bryant, S. H. Structure-Based Virtual Screening

for Drug Discovery: a Problem-Centric Review. AAPS J. 14, 133 (2012).

54. Cho, S. J. & Sun, Y. FLAME: A Program to Flexibly Align Molecules. J. Chem. Inf.

Model. 46, 298 (2006).

55. Cleves, A. E. & Jain, A. N. Effects of inductive bias on computational evaluations of

ligand-based modeling and on drug discovery. J. Comput. Aided Mol. Des. 22, 147 (2008).

56. Collins, A., Wilson, C. C. & Gilmore, C. J. Comparing entire crystal structures using

cluster analysis and fingerprint plots. Crystengcomm 12, 801 (2010).

57. Coppens, P. X-Ray Charge Densities and Chemical Bonding (Oxford University Press,

Oxford, 1997).

58. Coppens, P. & Hall, M. B. Electron Distributions and the Chemical Bond (Springer,

Boston, USA, 1982).

59. Corey, E. General methods for the construction of complex molecules. Pure Appl. Chem.

14, 19 (1967).

202

COMPLETE BIBLIOGRAPHY

60. Corey, R. B. & Pauling, L. Molecular models of amino acids, peptides, and proteins. Rev.

Sci. Instr. 24, 621 (1953).

61. Cranswick, L. M. D. Busting out of crystallography’s Sisyphean prison: from pencil

and paper to structure solving at the press of a button: past, present and future of

crystallographic software development, maintenance and distribution. Acta Crystallogr. A

64, 65 (2008).

62. Curry, B. & Rumelhart, D. E. MSnet: A Neural Network which Classifies Mass Spectra.

Tetrahedron Comput. Methodol. 3, 213 (1990).

63. Curtiss, L. A., Redfern, P. C., Raghavachari, K., Rassolov, V. & Pople, J. A. Gaussian-3

theory using reduced Moller-Plesset order. J. Chem. Phys. 110, 4703 (1999).

64. Curtiss, L. A., Redfern, P. C. & Raghavachari, K. Gaussian-4 theory using reduced order

perturbation theory. J. Chem. Phys. 127, 124105 (2007).

65. Curtiss, L. A., Redfern, P. C. & Raghavachari, K. Gn theory. Wiley Interdiscip. Rev.

Comput. Mol. Sci. 1, 810 (2011).

66. D. A. Dixon, D. F. & Peterson, K. A. Chapter One - A Practical Guide to Reliable First

Principles Computational Thermochemistry Predictions Across the Periodic Table. Annu.

Rep. Comput. Chem. 8 (2012).

67. Darden, T., York, D. & Pedersen, L. Particle mesh Ewald: An Nlog( N) method for Ewald

sums in large systems. J. Chem. Phys. 98, 10089 (1993).

68. Davidson, E. R. & Feller, D. Basis set selection for molecular calculations. Chem. Rev.

86, 681 (1986).

69. Desiraju, G. R. Supramolecular Synthons in Crystal Engineering—A New Organic Syn-

thesis. Angew. Chem. Int. Ed. 34, 2311 (1995).

70. Desiraju, G. R., Vittal, J. J. & Ramanan, A. Crystal engineering (Co-Published with

Indian Institute of Science (IISc), Bangalore, India, 2011).

71. Dettmer, K., Aronov, P. A. & Hammock, B. D. Mass spectrometry-based metabolomics.

Mass Spec. Rev. 26, 51 (2007).

72. Di Giovanni, A. L. & C. Virtual Screening Strategies in Drug Discovery: A Critical Review.

Curr. Med. Chem. 20, 2839 (2013).

73. Dirac, P. A. M. Quantum Mechanics of Many-Electron Systems. Proc. Royal Soc. A 123,

714 (1929).

74. Dobson, C. M. Chemical space and biology. Nature 432, 824 (2004).

75. Drebushchak, T. N., Boldyreva, E. V. & Mikhailenko, M. A. Crystal structures of

sulfathiazole polymorphs in the temperature range 100–295 K: A comparative analysis. J.

Struct. Chem. 49, 84 (2008).

203

COMPLETE BIBLIOGRAPHY

76. Dunning, T. H. Gaussian-basis sets for use in correlated molecular calculations .1. the

atoms boron through neon and hydrogen. J. Chem. Phys. 90, 1007 (1989).

77. Dunning, T. H., Peterson, K. A. & Wilson, A. K. Gaussian basis sets for use in correlated

molecular calculations. X. The atoms aluminum through argon revisited. J. Chem. Phys.

114, 9244 (2001).

78. Engel, T. Basic Overview of Chemoinformatics. J. Cheminf. Model. 46, 2267 (2006).

79. Essmann, U. et al. A smooth particle mesh Ewald method. J. Chem. Phys. 103, 8577

(1995).

80. Fahrenkamp-Uppenbrink, J., Szuromi, P., Yeston, J. & Coontz, R. Theoretical Possibilities.

Science 321, 783 (2008).

81. Fast, P. L., Sanchez, M. L. & Truhlar, D. G. Infinite basis limits in electronic structure

theory. J. Chem. Phys. 111, 2921 (1999).

82. Feller, D. The use of systematic sequences of wave-functions for estimating the complete

basis set, full configuration-interaction limit in water. J. Chem. Phys. 98, 7059 (1993).

83. Feller, D. & Peterson, K. A. An examination of intrinsic errors in electronic structure

methods using the Environmental Molecular Sciences Laboratory computational results

database and the Gaussian-2 set. J. Chem. Phys. 108, 154 (1998).

84. Feller, D., Peterson, K. A. & Crawford, T. D. Sources of error in electronic structure

calculations on small chemical systems. J. Chem. Phys. 124, 054107 (2006).

85. Feller, D. Benchmarks of improved complete basis set extrapolation schemes designed for

standard CCSD(T) atomization energies. J. Chem. Phys. 138, 074103 (2013).

86. Feller, D. & Peterson, K. A. Probing the limits of accuracy in electronic structure

calculations: Is theory capable of results uniformly better than “chemical accuracy”? J.

Chem. Phys. 126, 114105 (2007).

87. Feller, D., Peterson, K. A., de Jong, W. A. & Dixon, D. A. Performance of coupled cluster

theory in thermochemical calculations of small halogenated compounds. J. Chem. Phys.

118, 3510–3522 (2003).

88. Feller, D., Peterson, K. A. & Dixon, D. A. A survey of factors contributing to accurate

theoretical predictions of atomization energies and molecular structures. J. Chem. Phys.

129, 204105 (2008).

89. Feller, D., Peterson, K. A. & Dixon, D. A. Refined Theoretical Estimates of the Atom-

ization Energies and Molecular Structures of Selected Small Oxygen Fluorides. J. Phys.

Chem. A 114. PMID: 19968310, 613–623 (2010).

204

COMPLETE BIBLIOGRAPHY

90. Feller, D., Peterson, K. A. & Dixon, D. A. Further benchmarks of a composite, con-

vergent, statistically calibrated coupled-cluster-based approach for thermochemical and

spectroscopic studies. Mol. Phys 110, 2381–2399 (2012).

91. Feller, D., Peterson, K. A. & Hill, J. G. Calibration study of the CCSD(T)-F12a/b

methods for C2 and small hydrocarbons. J. Chem. Phys. 133, 184102 (2010).

92. Feller, D., Peterson, K. A. & Hill, J. G. On the effectiveness of CCSD(T) complete basis

set extrapolations for atomization energies. J. Chem. Phys. 135, 044102 (2011).

93. Feller, D., Peterson, K. A. & Ruscic, B. Improved accuracy benchmarks of small molecules

using correlation consistent basis sets. Theo. Chem. Acc. 133, 1407. issn: 1432-2234

(2013).

94. Ferry, G. Dorothy Hodgkin: A Life (Bloomsbury, London, 2014).

95. Fischer, E. Einfluss der Configuration auf die Wirkung der Enzyme. Ber. Dtsch. Chem.

Ges. 27, 2985. issn: 1099-0682 (1894).

96. Fock, V. Naherungsmethode zur Losung des quantenmechanischen Mehrkorperproblems.

Z. Phys. 61, 126 (1930).

97. Frisch, M. J. et al. Gaussian 09 Revision D. 01

98. Frisch, M. J., Pople, J. A. & Binkley, J. S. Self-consistent molecular orbital methods 25.

Supplementary functions for Gaussian basis sets. J. Chem. Phys. 80, 3265 (1984).

99. Gasteiger, J., Hanebeck, W. & Schulz, K. P. Prediction of mass spectra from structural

information. J. Chem. Inf. Model. 32, 264 (1992).

100. Gasteiger, J. Chemoinformatics: Achievements and Challenges, a Personal View. Molecules

21 (2016).

101. Gasteiger, J. & Engel, T. Chemoinformatics (eds Engel, T. & Gasteiger, J.) (Wiley,

Weinheim, 2003).

102. Geurts, P. et al. Proteomic mass spectra classification using decision tree based ensemble

methods. Bioinformatics 21, 3138 (2005).

103. Gill, P. M. W. Extraction of Stewart Atoms from Electron Densities. J. Phys. Chem. 100,

15421–15427 (38 1996).

104. Goerigk, L. & Grimme, S. A thorough benchmark of density functional methods for

general main group thermochemistry, kinetics, and noncovalent interactions. Phys. Chem.

Chem. Phys. 13, 6670 (2011).

105. Goerigk, L. & Grimme, S. Double-hybrid density functionals. Wiley Interdiscip. Rev.

Comput. Mol. Sci. 4, 576 (2014).

205

COMPLETE BIBLIOGRAPHY

106. Goerigk, L., Karton, A., Martin, J. M. L. & Radom, L. Accurate quantum chemical

energies for tetrapeptide conformations: why MP2 data with an insufficient basis set

should be handled with caution. Phys. Chem. Chem. Phys. 15, 7028 (2013).

107. Gould, M. D., Taylor, C., Wolff, S. K., Chandler, G. S. & Jayatilaka, D. A definition for

the covalent and ionic bond index in a molecule. Theor. Chem. Acc. 119, 275–290 (1–3

2008).

108. Grimme, S. Towards First Principles Calculation of Electron Impact Mass Spectra of

Molecules. Angew. Chem. Int. Ed. 52, 6306 (2013).

109. Grimme, S. & Bauer, C. A. First principles calculation of electron ionization mass spectra

for selected organic drug molecules. Org. Biomol. Chem. 12, 8737 (2014).

110. Groom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. The Cambridge Structural

Database. Acta Crystallogr. B 72, 171 (2016).

111. Groom, C. R. & Cole, J. C. The use of small-molecule structures to complement protein-

ligand crystal structures in drug discovery. Acta Crystallogr. D 73, 240 (2017).

112. Hahn, M. Three-Dimensional Shape-Based Searching of Conformationally Flexible Com-

pounds. J. Chem. Inf. Comput. Sci. 37, 80 (1997).

113. Halkier, A. et al. Basis-set convergence in correlated calculations on Ne, N-2, and H2O.

Chem. Phys. Lett. 286, 243 (1998).

114. Hall, S. R., Allen, F. H. & Brown, I. D. The crystallographic information file (CIF): a

new standard archive file for crystallography. Acta Crystallogr. A 47, 655 (1991).

115. Hall, S. R. et al. in International Tables for Crystallography Volume G: Definition

and exchange of crystallographic data (eds Hall, S. R. & McMahon, B.) 20 (Springer

Netherlands, Dordrecht, 2005).

116. Hamprecht, F. A., Cohen, A. J., Tozer, D. J. & Handy, N. C. Development and assessment

of new exchange-correlation functionals. J. Chem. Phys. 109, 6264 (1998).

117. Hansch, C., Maloney, P. P., Fujita, T. & Muir, R. M. Correlation of Biological Activity

of Phenoxyacetic Acids with Hammett Substituent Constants and Partition Coefficients.

Nature 194, 178 (1962).

118. Hansen, M. E. & Smedsgaard, J. A new matching algorithm for high resolution mass

spectra. J. Am. Soc. Mass Spectrom. 15, 1173 (2004).

119. Hanwell, M. D. et al. Avogadro: an advanced semantic chemical editor, visualization, and

analysis platform. J. Cheminf. 4, 17 (2012).

120. Harding, M. E. et al. High-accuracy extrapolated ab initio thermochemistry. III. Additional

improvements and overview. J. Chem. Phys. 128, 114111 (2008).

206

COMPLETE BIBLIOGRAPHY

121. Hartree, D. R. The Wave Mechanics of an Atom with a Non-Coulomb Central Field. Part

I. Theory and Methods. Math. Proc. Camb. Philos. Soc. 24, 89 (1928).

122. Hartree, D. R. The Wave Mechanics of an Atom with a Non-Coulomb Central Field. Part

II. Some Results and Discussion. Math. Proc. Camb. Philos. Soc. 24, 111 (1928).

123. Hattig, C., Klopper, W., Kohn, A. & Tew, D. P. Explicitly Correlated Electrons in

Molecules. Chem. Rev. 112, 4 (2011).

124. Hawkins, P. C. D., Skillman, A. G., Warren, G. L., Ellingson, B. A. & Stahl, M. T.

Conformer Generation with OMEGA: Algorithm and Validation Using High Quality

Structures from the Protein Databank and Cambridge Structural Database. J. Chem.

Inf. Model. 50, 572 (2010).

125. He, X., Fusti-Molnar, L. & Merz, K. M. J. Accurate Benchmark Calculations on the

Gas-Phase Basicities of Small Molecules. J. Phys. Chem. A 113, 10096 (2009).

126. Heckert, M., Kallay, M. & Gauss, J. Molecular equilibrium geometries based on coupled-

cluster calculations including quadruple excitations. Mol. Phys 103, 2109–2115 (2005).

127. Heckert, M., Kallay, M., Tew, D. P., Klopper, W. & Gauss, J. Basis-set extrapolation

techniques for the accurate calculation of molecular equilibrium geometries using coupled-

cluster theory. J. Chem. Phys. 125, 044108 (2006).

128. Hehre, W. J., Ditchfield, R. & Pople, J. A. Self—Consistent Molecular Orbital Methods.

XII. Further Extensions of Gaussian—Type Basis Sets for Use in Molecular Orbital

Studies of Organic Molecules. J. Chem. Phys. 56, 2257 (1972).

129. Hehre, W. J., Stewart, R. F. & Pople, J. A. Atomic electron populations by molecular

orbital theory. Symp. Faraday Soc. 2, 15 (1968).

130. Hehre, W. J., Stewart, R. F. & Pople, J. A. Self-Consistent Molecular-Orbital Methods. I.

Use of Gaussian Expansions of Slater-Type Atomic Orbitals. J. Chem. Phys. 51, 2657

(1969).

131. Heilbronner, E. & Straub. Huckel Molecular orbitals (Springer-Verlag, Berlin, 1966).

132. Helgaker, T., Klopper, W., Koch, H. & Noga, J. Basis-set convergence of correlated

calculations on water. J. Chem. Phys. 106, 9639 (1997).

133. Helgaker, T. et al. in (ed Ciolowski, J.) 1 (Kluwer, Dordrecht, 2001).

134. Helgaker, T., Klopper, W. & Tew, D. P. Quantitative quantum chemistry. Mol. Phys.

106, 2107 (2008).

135. Henry, N. L. Knowledge Management: A New Concern for Public Administration. Pub.

Admin. Rev. 34, 189 (1974).

207

COMPLETE BIBLIOGRAPHY

136. Hert, J. et al. New Methods for Ligand-Based Virtual Screening: Use of Data Fusion

and Machine Learning to Enhance the Effectiveness of Similarity Searching. J. Chem.

Inf. Model. 46, 462 (2006).

137. Hertz, H. S., Hites, R. A. & Biemann, K. Identification of mass spectra by computer-

searching a file of known spectra (Anal. Chem., 1971).

138. Hey, T. The Fourth Paradigm: Data-Intensive Scientific Discovery (eds Hey, T., Tolle,

K. M. & Tansley, S.) (Microsoft Research, Seattle, 2009).

139. Hidaka, K. et al. Small-sized human immunodeficiency virus type-1 protease inhibitors

containing allophenylnorstatine to explore the S2’ pocket. J Med Chem 52, 7604 (2009).

140. Hill, A. W. & Mortishire-Smith, R. J. Automated assignment of high-resolution collisionally

activated dissociation mass spectra using a systematic bond disconnection approach. Rapid

Commun. Mass Spectrom. 19, 3111–3118. issn: 1097-0231 (2005).

141. Hill, J. G., Peterson, K. A., Knizia, G. & Werner, H.-J. Extrapolating MP2 and CCSD

explicitly correlated correlation energies to the complete basis set limit with first and

second row correlation consistent basis sets. J. Chem. Phys. 131, 194105 (2009).

142. Hirshfeld, F. L. Bonded-atom fragments for describing molecular charge densities. Theor.

Chem. Acc. 44, 129 (1977).

143. Hirshfeld, F. L. & Rzotkiewicz, S. Electrostatic binding in the first-row AH and A2

diatomic molecules. Mol. Phys 27, 1319 (1974).

144. Hodgkin, E. E. & Richards, W. G. Molecular similarity based on electrostatic potential

and electric field. Int. J. Quantum Chem. 32, 105 (1987).

145. Hoffmann, R. A theoretician’s view of crystallography. Acta Crystallogr. A 61, c1 (2005).

146. Hoffmann, R. An Extended Huckel Theory. I. Hydrocarbons. J. Chem. Phys. 39, 1397

(1963).

147. Hohenberg, P. & Kohn, W. Inhomogeneous Electron Gas. Phys. Rev. 136, B864B871

(1964).

148. Huang, N., Shoichet, B. K. & Irwin, J. J. Benchmarking Sets for Molecular Docking. J.

Med. Chem. 49, 6789 (2006).

149. Huang, Z. & Kim, K.-J. Review of x-ray free-electron laser theory. Phys. Rev. Accel.

Beams 10 (2007).

150. Huckel, E. Die freien Radikale der organischen Chemie. Z. Phys. 83, 632 (1933).

151. Hummel, R. E. Understanding Materials Science 32 (Springer, New York, 2004).

152. Humphrey, W., Dalke, A. & Schulten, K. VMD Visual Molecular Dynamics. J. Mol.

Graph. 14, 33 (1996).

208

COMPLETE BIBLIOGRAPHY

153. Hunter, G. Atoms in molecules from the exact one-electron wave function. Can. J. Chem.

74, 1008–1013 (6 1996).

154. Hunter, J. D. Matplotlib: A 2D Graphics Environment. Comput. Sci. Eng. 9, 90 (2007).

155. I-J. uan Yu, L. & mir Karton, A. Assessment of theoretical procedures for a diverse set

of isomerization reactions involving double-bond migration in conjugated dienes. Chem.

Phys. 441, 166 (2014).

156. Interpretation of Mass Spectra, 4th ed. (McLafferty, Fred W.; Turecek, Frantisek). J.

Chem. Ed. 71, A54 (1994).

157. Irk Bakowies, D. Accurate extrapolation of electron correlation energies from small basis

sets. J. Chem. Phys. 127, 164109 (2007).

158. Irwin, J. J. & Shoichet, B. K. ZINC − A Free Database of Commercially Available

Compounds for Virtual Screening. J. Cheminf. Model. 45, 177 (2005).

159. Jacob, C. R. How Open Is Commercial Scientific Software? J. Phys. Chem. Lett. 7, 351

(2016).

160. Jalali-Heravi, M. & Fatemi, M. H. Simulation of mass spectra of noncyclic alkanes and

alkenes using artificial neural network. Anal. Chim. Acta 415, 95 (2000).

161. Jan H Schuur, Selzer, P. & Gasteiger, J. The Coding of the Three-Dimensional Structure of

Molecules by Molecular Transforms and Its Application to Structure-Spectra Correlations

and Studies of Biological Activity. J. Chem. Inf. Comput. Sci. 36, 334 (1996).

162. Jarzembska, K. N. et al. Combined Experimental and Computational Studies of Pyrazi-

namide and Nicotinamide in the Context of Crystal Engineering and Thermodynamics.

Cryst. Growth Des. 14, 3453 (2014).

163. Jayatilaka, D. & Grimwood, D. J. in Computational Science — ICCS 2003: International

Conference, Melbourne, Australia and St. Petersburg, Russia, June 2–4, 2003 Proceedings,

Part IV (eds Sloot, P. M. A. et al.) 142 (Springer Berlin Heidelberg, Berlin, Heidelberg,

2003).

164. Jensen, F. Polarization consistent basis sets: Principles. J. Chem. Phys. 115, 9113 (2001).

165. Jensen, F. Introduction to Computational Chemistry 2nd ed. (Wiley, Hoboken, USA,

2007).

166. Jones, E., Oliphant, T., Peterson, P., et al. SciPy: Open source scientific tools for Python

[Online; accessed 01/11/2017]. 2001. http://www.scipy.org/.

167. Jones, G., Willett, P., Glen, R. C., Leach, A. R. & Taylor, R. Development and validation

of a genetic algorithm for flexible docking. J. Mol. Biol. 267, 727.

168. Jorgensen, W. L. The many roles of computation in drug discovery. Science 303, 1813

(2004).

209

COMPLETE BIBLIOGRAPHY

169. Jorgensen, W. L., Chandrasekhar, J., Madura, J. D., Impey, R. W. & Klein, M. L.

Comparison of simple potential functions for simulating liquid water. J. Chem. Phys. 79,

926 (1983).

170. Jurecka, P., Sponer, J., Cerny, J. & Hobza, P. Benchmark database of accurate (MP2

and CCSD(T) complete basis set limit) interaction energies of small model complexes,

DNA base pairs, and amino acid pairs. Phys. Chem. Chem. Phys. 8, 1985 (2006).

171. Kamat, P. V. et al. Reaching Out with Physical Chemistry. J. Phys. Chem. Lett. 7, 103

(2016).

172. Kang, Y. et al. Crystal structure of rhodopsin bound to arrestin by femtosecond X-ray

laser. Nature 523, 561 (2015).

173. Karton, A. A computational chemist’s guide to accurate thermochemistry for organic

molecules. Wiley Interdiscip. Rev. Comput. Mol. Sci. 6, 292 (2016).

174. Karton, A. How large are post-CCSD(T) contributions to the total atomization energies

of medium-sized alkanes? Chem. Phys. Lett. 645, 118–122. issn: 0009-2614 (2016).

175. Karton, A., Daon, S. & Martin, J. M. L. W4-11: A high-confidence benchmark dataset

for computational thermochemistry derived from first-principles W4 data. Chem. Phys.

Lett. 510, 165 (2011).

176. Karton, A., Kaminker, I. & Martin, J. M. L. Economical Post-CCSD(T) Computational

Thermochemistry Protocol and Applications to Some Aromatic Compounds. J. Phys.

Chem. A 113, 7610 (2009).

177. Karton, A. & Martin, J. M. L. Comment on: “Estimating the Hartree–Fock limit from

finite basis set calculations” [Jensen F (2005) Theor Chem Acc 113:267]. Theor. Chem.

Acc. 115, 330 (2006).

178. Karton, A. & Martin, J. M. L. Performance of W4 theory for spectroscopic constants and

electrical properties of small molecules. J. Chem. Phys. 133, 144102 (2010).

179. Karton, A. & Martin, J. M. L. Explicitly correlated Wn theory: W1-F12 and W2-F12. J.

Chem. Phys. 136, 124114 (2012).

180. Karton, A., Rabinovich, E., Martin, J. M. L. & Ruscic, B. W4 theory for computational

thermochemistry: In pursuit of confident sub-kJ/mol predictions. J. Chem. Phys. 125,

144108 (2006).

181. Karton, A., Taylor, P. R. & Martin, J. M. L. Basis set convergence of post-CCSD

contributions to molecular atomization energies. J. Chem. Phys. 127, 064104 (2007).

182. Karton, A., Yu, L.-J., Kesharwani, M. K. & Martin, J. M. L. Heats of formation of the

amino acids re-examined by means of W1-F12 and W2-F12 theories. Theoretical Chem.

Accounts 133, 1483 (2014).

210

COMPLETE BIBLIOGRAPHY

183. Kendall, R. A., Dunning, T. H. & Harrison, R. J. Electron-affinities of the 1st-row atoms

revisited - systematic basis-sets and wave-functions. J. Chem. Phys. 96, 6796 (1992).

184. Kendrew, J. C. et al. A Three-Dimensional Model of the Myoglobin Molecule Obtained

by X-Ray Analysis. Nature 181, 662 (1958).

185. Kerber, A., Meringer, M. & Rucker, C. CASE via MS: Ranking Structure Candidates by

Mass Spectra. Croat. Chem. Acta 79, 449 (2006).

186. Khedkar, S. A., Malde, A. K., Coutinho, E. C. & Srivastava, S. Pharmacophore modeling

in drug discovery and development: an overview. Med. Chem. 3, 187 (2007).

187. Kim, S. et al. PubChem Substance and Compound databases. Nucleic Acids Res. 44,

D1202D1213 (2016).

188. Kind, T. & Fiehn, O. Advances in structure elucidation of small molecules using mass

spectrometry. Bioanal. Rev. 2, 23 (2010).

189. Kitaigorodskii, A. I. Principle of close packing and condition of thermodynamic stability

of organic crystals. Acta Crystallogr. 18, 585 (1965).

190. Kitchen, D. B., Decornez, H., Furr, J. R. & Bajorath, J. Docking and scoring in virtual

screening for drug discovery: methods and applications. Nat. Rev. Drug Discov. 3, 935

(2004).

191. Klopper, W. Highly accurate coupled-cluster singlet and triplet pair energies from explicitly

correlated calculations in comparison with extrapolation techniques. Mol. Phys. 99, 481

(2001).

192. Klopper, W., Ruscic, B., Tew, D. P., Bischoff, F. A. & Wolfsegger, S. Atomization energies

from coupled-cluster calculations augmented with explicitly-correlated perturbation theory.

Chem. Phys. 356, 14 (2009).

193. Knizia, G., Adler, T. B. & Werner, H.-J. Simplified CCSD(T)-F12 methods: Theory and

benchmarks. J. Chem. Phys. 130, 054104 (2009).

194. Koga, T., Kanayama, K., Watanabe, S. & Thakkar, A. J. Analytical Hartree–Fock wave

functions subject to cusp and asymptotic constraints: He to Xe, Li+ to Cs+, H– to I– .

Int. J. Quantum Chem. 71, 491 (1999).

195. Kohn, W. & Sham, L. J. Self-Consistent Equations Including Exchange and Correlation

Effects. Phys. Rev. 140, A1133A1138 (1965).

196. Kong, L., Bischoff, F. A. & Valeev, E. F. Explicitly Correlated R12/F12 Methods for

Electronic Structure. Chem. Rev. 112, 75 (2011).

197. Krishnan, R., Binkley, J. S., Seeger, R. & Pople, J. A. Self-consistent molecular orbital

methods. XX. A basis set for correlated wave functions. J. Chem. Phys. 72, 650 (1980).

211

COMPLETE BIBLIOGRAPHY

198. Krylov, A. I. et al. What Is the Price of Open-Source Software? J. Phys. Chem. Lett. 6,

2751 (2015).

199. Kuhn, T. The Structure of Scientific Revolutions (University of Chicago Press, Chicago,

1962).

200. Lebedev, V. I. Quadratures on the sphere. Zh. Vychisl. Mat. Mat. Fiz. 16, 293 (1976).

201. Lebedev, V. I. Spherical quadrature formulas exact to order-25-order-29. Sib. Math. J.

18, 99 (1977).

202. Lee, C., Yang, W. & Parr, R. G. Development of the Colle-Salvetti correlation-energy

formula into a functional of the electron density. Phys. Rev. B 37, 785 (1988).

203. Leech, C. K., Fabbiani, F. P. A., Shankland, K., David, W. I. F. & Ibberson, R. M. Accurate

molecular structures of chlorothiazide and hydrochlorothiazide by joint refinement against

powder neutron and X-ray diffraction data. Acta Crystallogr. B 64, 101 (2008).

204. Lehn, J.-M. Supramolecular Chemistry: Concepts and Perspectives isbn: 9783527607433.

doi:10.1002/3527607439. http://dx.doi.org/10.1002/3527607439 (Wiley, 2006).

205. Leininger, M. L., Allen, W. D., Schaefer III, H. F. & Sherrill, C. D. Is Møller–Plesset

perturbation theory a convergent ab initiomethod? J. Chem. Phys. 112, 9213 (2000).

206. Levinthal, C. Molecular model-bilding by computer. Sci. Amer. 214, 42 (1966).

207. Levy, M. Universal variational functionals of electron densities, first-order density matrices,

and natural spin-orbitals and solution of the v-representability problem. Proc. Natl. Acad.

Sci. U.S.A 76, 6062 (1979).

208. Liakos, D. G. & Neese, F. Improved Correlation Energy Extrapolation Schemes Based on

Local Pair Natural Orbital Methods. J. Phys. Chem. A 116, 4801 (2012).

209. Lionta, E., Spyrou, G., Vassilatis, D. K. & Cournia, Z. Structure-based virtual screening

for drug discovery: principles, applications and recent advances. Curr. Top. Med. Chem.

14, 1923 (2014).

210. Lonsdale, K. The structure of the benzene ring. Nature 122, 810 (1928).

211. Maier, J. A. et al. ff14SB: Improving the Accuracy of Protein Side Chain and Backbone

Parameters from ff99SB. Journal of Chemical Theory and Computation 11, 3696 (2015).

212. Mak, L., Grandison, S. & Morris, R. J. An extension of spherical harmonics to region-based

rotationally invariant descriptors for molecular shape description and comparison. J. Mol.

Graph. Model. 26, 1035 (2008).

213. Martin, J. M. L. Ab initio total atomization energies of small molecules - Towards the

basis set limit. Chem. Phys. Lett. 259, 669 (1996).

212

COMPLETE BIBLIOGRAPHY

214. Martin, J. M. L. Computational thermochemistry: A brief overview of quantum mechanical

approaches. Annu. Rep. Comput. Chem 1 (2005).

215. Martin, J. M. L. & de Oliveira, G. Towards standard methods for benchmark quality ab

initio thermochemistry - W1 and W2 theory. J. Chem. Phys. 111, 1843 (1999).

216. Martin, J. M. L. & Lee, T. J. The atomization energy and proton affinity of NH3. An ab

initio calibration study. Chem. Phys. Lett. 258, 136 (1996).

217. Martin, J. M. L. & Parthiban, S. Quantum-Mechanical Prediction of Thermochemical

Data, Understanding Chemical Reactivity 31 (Kluwer, Dordrecht, 2001).

218. Martin, J. M. L. & Taylor, P. R. Structure and vibrations of small carbon clusters from

coupled-cluster calculations. J. Phys. Chem. 100, 6047 (1996).

219. Martin, J. M. L. & Taylor, P. R. Benchmark quality total atomization energies of small

polyatomic molecules. J. Chem. Phys. 106, 8620 (1997).

220. Martin, J. M. L. On the performance of correlation consistent basis sets for the calculation

of total atomization energies, geometries, and harmonic frequencies. J. Chem. Phys. 100,

8186–8193 (1994).

221. Martin, J. M. L. & Kesharwani, M. K. Assessment of CCSD(T)-F12 Approximations and

Basis Sets for Harmonic Vibrational Frequencies. J. Chem. Theory Comput. 10, 2085

(2014).

222. Mayer, I., Knapp-Mohammady, M. & Suhai, S. Bond orders and energy components in

polymers. Chem. Phys. Lett. 389, 34 (2004).

223. McGann, M. FRED Pose Prediction and Virtual Screening Accuracy. J. Chem. Inf. Model.

51, 578 (2011).

224. McGann, M. FRED and HYBRID docking performance on standardized datasets. J.

Comput. Aided Mol. Des. 26, 897 (2012).

225. McKinney, W. Data Structures for Statistical Computing in Python in Proceedings of the

9th Python in Science Conference (eds van der Walt, S. ’. & Millman, J.) (2010), 51.

226. McKinnon, J. J., Jayatilaka, D. & Spackman, M. A. Towards quantitative analysis of

intermolecular interactions with Hirshfeld surfaces. Chem. Commun. 267, 3814 (2007).

227. McKinnon, J. J., Spackman, M. A. & Mitchell, A. S. Novel tools for visualizing and

exploring intermolecular interactions in molecular crystals. Acta Crystallogr. B 60, 627

(2004).

228. McLafferty, F. W., Hertel, R. H. & Villwock, R. D. Probability based matching of mass

spectra. Rapid identification of specific compounds in mixtures. J. Mass Spectrom. 9, 690

(1974).

213

COMPLETE BIBLIOGRAPHY

229. McLafferty, F. W., Stauffer, D. A., Loh, S. Y. & Wesdemiotis, C. Unknown identification

using reference mass spectra. Quality evaluation of databases. J. Am. Soc. Mass Spectrom.

10, 1229 (1999).

230. McLafferty, F. W. & Turecek, F. Interpretation Of Mass Spectra (University Science

Books, 1993).

231. McLafferty, F. W., Zhang, M. Y., Stauffer, D. B. & Loh, S. Y. Comparison of algorithms

and databases for matching unknown mass spectra. J. Am. Soc. Mass Spectrom. 9, 92

(1998).

232. McLean, A. D. & Yoshimine, M. Computation of Molecular Properties and Structure.

IBM J. Res. Dev. 12, 206 (1968).

233. McWeeny, R. X-ray scattering by aggregates of bonded atoms. III. The bond scattering

factor: simple methods of approximation in the general case. Acta Crystallogr. 6, 631

(1953).

234. Melani, F., Gratteri, P., Adamo, M. & Bonaccini, C. Field Interaction and Geometrical

Overlap: A New Simplex and Experimental Design Based Computational Procedure for

Superposing Small Ligand Molecules. J. Med. Chem. 46, 1359 (2003).

235. Mir Karton, A. & ars Goerigk, L. Accurate Reaction Barrier Heights of Pericyclic

Reactions: Surprisingly Large Deviations for the CBS-QB3 Composite Method and Their

Consequences in DFT Benchmark Studies. J. Computational Chem. 36, 622 (2015).

236. Mitchell, A. S. & Spackman, M. A. Molecular surfaces from the promolecule: A comparison

with Hartree–Fock ab initio electron density surfaces. Journal of Computational Chemistry

21, 933 (2000).

237. Moffitt, W. Atoms in Molecules and Crystals. Proc. Royal Soc. A 210, 245–268 (1951).

238. Møller, C. & Plesset, M. S. Note on an Approximation Treatment for Many-Electron

Systems. Phys. Rev. 46, 618 (1934).

239. Moore, G. E. Cramming more components onto integrated circuits. Electronics 38, 114

(1965).

240. Morozik, Y. I., Galyaev, G. V. & Smirnov, A. O. Prediction of electron-ionization mass

spectra of O-Alkyl-N,N-dialkylphosphoramido cyanidates. J. Anal. Chem. 66, 1285 (2011).

241. Morris, R. J., Najmanovich, R. J., Kahraman, A. & Thornton, J. M. Real spherical

harmonic expansion coefficients as 3D shape descriptors for protein binding pocket and

ligand comparisons. Bioinformatics 21, 2347 (2005).

242. Mulliken, R. S. Electronic Structures of Polyatomic Molecules and Valence. Phys. Rev.

40, 55 (1932).

214

COMPLETE BIBLIOGRAPHY

243. Mulliken, R. S. Electronic Population Analysis on LCAO–MO Molecular Wave Functions.

I. J. Chem. Phys. 23, 1833 (1955).

244. Mulliken, R. S. Electronic Structures of Polyatomic Molecules and Valence. II. Quantum

Theory of the Double Bond. Phys. Rev. 41, 751 (1932).

245. Mulliken, R. S. Electronic Structures of Polyatomic Molecules and Valence. V. Molecules

RX n. J. Chem. Phys. 1, 492 (1933).

246. Mun, I. K. & Mclafferty, F. W. in Supercomputers in Chemistry 117 ().

247. Mysinger, M. M., Carchia, M., Irwin, J. J. & Shoichet, B. K. Directory of Useful Decoys,

Enhanced (DUD-E): Better Ligands and Decoys for Better Benchmarking. J. Med. Chem.

55, 6582 (2012).

248. N N Bogolubov, J. Introduction to Quantum Statistical Mechanics (World Scientific,

2010).

249. N. DeYonker, T. R. C. & Wilson, A. K. Advances in the Theory of Atomic Molecular

Systems, Progress in Theoretical Chemistry Physics 197–224 (Springer, Netherlands,

Dordrecht, 2009).

250. Nichols, G. & Frampton, C. S. Physicochemical Characterization of the Orthorhombic

Polymorph of Paracetamol Crystallized from Solution. J. Pharma. Sci. 87, 684 (1998).

251. NIST Standard Reference Database 1A v17 https : / / www . nist . gov / srd / nist -

standard-reference-database-1a-v17 (2017).

252. O-ohata, K., Taketa, H. & Huzinaga, S. Gaussian Expansions of Atomic Orbitals. J. Phys.

Soc. Jpn. 21, 2306 (1966).

253. O’Boyle, N. M. et al. Open Babel: An open chemical toolbox. J. Cheminform 3, 33

(2011).

254. Pansini, F. & Varandas, A. Toward a unified single-parameter extrapolation scheme for

the correlation energy: Systems formed by atoms of hydrogen through neon. Chem. Phys.

Lett. 631, 70. issn: 0009-2614 (2015).

255. Parkin, A. et al. Comparing entire crystal structures: structural genetic fingerprinting.

Crystengcomm 9, 648 (2007).

256. Parr, R. G. & Yang, W. Density functional theory of atoms and molecules (Oxford

University Press, Oxford, 1989).

257. Perdew, J. P., Burke, K. & Ernzerhof, M. Generalized Gradient Approximation Made

Simple. Phys. Rev. Lett. 77, 3865 (1996).

258. Perdew, J. P., Ernzerhof, M. & Burke, K. Rationale for mixing exact exchange with

density functional approximations. J. Chem. Phys. 105, 9982 (1996).

215

COMPLETE BIBLIOGRAPHY

259. Peterson, K. A., Woon, D. E. & Dunning, T. H. Benchmark calculations with correlated

molecular wave-functions .4. the classical barrier height of the h+h-2-]h-2+h reaction. J.

Chem. Phys. 100, 7410 (1994).

260. Peterson, K. A., Adler, T. B. & Werner, H.-J. Systematically convergent basis sets for

explicitly correlated wavefunctions: The atoms H, He, B-Ne, and Al-Ar. J. Chem. Phys.

128, 084102 (2008).

261. Peterson, K. A., Feller, D. & Dixon, D. A. Chemical accuracy in ab initio thermochemistry

and spectroscopy: current strategies and future challenges. Theor. Chem. Acc. 131, 1079

(2012).

262. Peverati, R. & Truhlar, D. G. Exchange–Correlation Functional with Good Accuracy for

Both Structural and Energetic Properties while Depending Only on the Density and Its

Gradient. J. Chem. Theory Comput. 8, 2310 (2012).

263. Popper, K. Conjectures and Refutations (Routlege & Kegan Paul, New York, 1962).

264. Press, W. H., Teukolsky, S. A., Vetterling, W. T. & Flannery, B. P. Numerical Recipes

3rd Edition: The Art of Scientific Computing 3rd ed. (Cambridge University Press, New

York, NY, USA, 2007).

265. Raghavachari, K., Trucks, G. W., Pople, J. A. & Headgordon, M. A 5th-order perturbation

comparison of electron correlation theories. Chem. Phys. Lett. 157, 479 (1989).

266. Raghavachari, K. Historical perspective on: A fifth-order perturbation comparison of

electron correlation theories [Volume 157, Issue 6, 26 May 1989, Pages 479-483]. Chem.

Phys. Lett. 589, 35 (2013).

267. Ranasinghe, D. S. & Petersson, G. A. CCSD(T)/CBS atomic and molecular benchmarks

for H through Ar. J. Chem. Phys. 138, 144104 (2013).

268. Rasmussen, G. T. & Isenhour, T. L. The evaluation of mass spectral search algorithms. J.

Chem. Inf. (1979).

269. Ravindra Acharya, K., Kuchela, K. N. & Kartha, G. Crystal structure of sulfamerazine.

J. Crystallogr. Spectrosc. Res. 12, 369 (1982).

270. Ray, L. C. & Kirsch, R. A. Finding Chemical Records by Digital Computers. Science

126, 814 (1957).

271. Reboul, J. P., Cristau, B., Soyfer, J. C. & Astier, J. P. 5 H-Dibenz[ b, f]azepinecarboxamide-

5 (carbamazepine). Acta Crystallogr. B 37, 1844 (1981).

272. Rina Brauer, B., anoj K. . Kesharwani, M. & an M. . L. . Martin, J. Some Observa-

tions on Counterpoise Corrections for Explicitly Correlated Calculations on Noncovalent

Interactions. J. Chem. Theory Comput. 10, 3791 (2014).

216

COMPLETE BIBLIOGRAPHY

273. Roby, K. R. Quantum theory of chemical valence concepts. Mol. Phys. 27, 81–104 (1

1974).

274. Rowley, J. The wisdom hierarchy: representations of the DIKW hierarchy. J. Inf. Sci. 33,

163 (2007).

275. Russell, S. J. & Norvig, P. Artificial Intelligence (Prentice Hall, 2010).

276. Ryckaert, J.-P., Ciccotti, G. & Berendsen, H. J. C. Numerical integration of the cartesian

equations of motion of a system with constraints: molecular dynamics of n-alkanes. Journal

of Computational Physics 23, 327 (1977).

277. Rys, J., Dupuis, M. & King, H. F. Computation of electron repulsion integrals using the

rys quadrature method. J. Comput. Chem. 4, 154 (1983).

278. Schaeffer, N. Efficient spherical harmonic transforms aimed at pseudospectral numerical

simulations. Geochem. Geophys. 14, 751 (2013).

279. Scheubert, K., Hufsky, F. & Bocker, S. Computational mass spectrometry for small

molecules. J. Cheminf. 5, 12 (2013).

280. Schrodinger, E. Quantisierung als Eigenwertproblem. Ann. Phys. (Berl). 384, 361 (1926).

281. Schrodinger, E. An Undulatory Theory of the Mechanics of Atoms and Molecules. Phys.

Rev. 28, 1049 (1926).

282. Schwenke, D. W. The extrapolation of one-electron basis sets in electronic structure

calculations: How it should work and how it can be made to work. J. Chem. Phys. 122,

014107 (2005).

283. Shavitt, I. & Bartlett, R. J. Many-body methods in chemistry and physics: MBPT and

coupled-cluster theory, Cambridge molecular science series (Cambridge University Press,

Cambridge, 2009).

284. Shen, H., Duhrkop, K., Bocker, S. & Rousu, J. Metabolite identification through multiple

kernel learning on fragmentation trees. Bioinformatics 30, i157 (2014).

285. Sheridan, R. P. & Kearsley, S. K. Why do we need so many chemical similarity search

methods? Drug Discov. Today 7, 903 (2002).

286. Simonite, T. Moore’s Law is Dead. Now What? https://www.technologyreview.com/

s/601441/moores-law-is-dead-now-what/ (2017).

287. Slater, J. C. Atomic Shielding Constants. Phys. Rev. 36, 57 (1930).

288. Snyder, L. C. & Basch, H. Molecular wave functions and properties: (ed Basch, H.)

(Wiley-Interscience, New York, 1972).

289. Spackman, M. A. & Maslen, E. N. Chemical-properties from the promolecule. J. Phys.

Chem. 90, 2020 (1986).

217

COMPLETE BIBLIOGRAPHY

290. Spackman, M. A. Molecules in crystals. Phys. Scripta 87, 048103 (2013).

291. Spackman, M. A. & Jayatilaka, D. Hirshfeld surface analysis. Crystengcomm 11, 19

(2009).

292. Spackman, M. A. & McKinnon, J. J. Fingerprinting intermolecular interactions in molec-

ular crystals. Crystengcomm 4, 378 (2002).

293. Spackman, M. A. & Byrom, P. G. A novel definition of a molecule in a crystal. Chem.

Phys. Lett. 267, 215 (1997).

294. Spackman, M. A. & Jayatilaka, D. Hirshfeld surface analysis. CrystEngComm 11, 19.

295. Spackman, P. R., Thomas, S. P. & Jayatilaka, D. High Throughput Profiling of Molecular

Shapes in Crystals. Sci. Rep. 6, 22204 (2016).

296. Stahura, F. L. & Bajorath, J. New methodologies for ligand-based virtual screening. Cur.

Pharm. Des. 11, 1189 (2005).

297. Staroverov, V. N., Scuseria, G. E., Tao, J. & Perdew, J. P. Comparative assessment of

a new nonempirical density functional: Molecules and hydrogen-bonded complexes. J.

Chem. Phys. 119, 12129 (2003).

298. Stein, S. E. An integrated method for spectrum extraction and compound identification

from gas chromatography/mass spectrometry data. J. Am. Soc. Mass Spectrom. 10, 770

(1999).

299. Stein, S. E. Estimating probabilities of correct identification from results of mass spectral

library searches. J. Am. Soc. Mass Spectrom. 5, 316 (1994).

300. Stein, S. E. & Scott, D. R. Optimization and testing of mass spectral library search

algorithms for compound identification. J. Am. Soc. Mass Spectrom. 5, 859 (1994).

301. Stephens, P. J., Devlin, F. J., Chabalowski, C. F. & Frisch, M. J. Ab Initio Calculation of

Vibrational Absorption and Circular Dichroism Spectra Using Density Functional Force

Fields. J. Phys. Chem. 98, 11623 (1994).

302. Stewart, R. F. Small Gaussian Expansions of Slater-Type Orbitals. J. Chem. Phys. 52,

431 (1970).

303. Sukumar, N. A Matter of Density (Wiley, New Jersey, 2012).

304. Sylvetsky, N., Peterson, K. A., Karton, A. & an M. L. Martin, J. Toward a W4-F12 ap-

proach: Can explicitly correlated and orbital-based ab initio CCSD(T) limits be reconciled?

J. Chem. Phys. 144, 214101 (2016).

305. Szabo, A. & Ostlund, N. S. Modern Quantum Chemistry (Courier Corporation, London,

1989).

218

COMPLETE BIBLIOGRAPHY

306. Tajti, A. et al. HEAT: High accuracy extrapolated ab initio thermochemistry. J. Chem.

Phys. 121, 11599 (2004).

307. Tanaka, K. X-ray analysis of wavefunctions by the least-squares method incorporating

orthonormality. I. General formalism. Acta Crystallogr. A: Crystal Physics 44, 1002

(1988).

308. Tangelder, J. W. H. & Veltkamp, R. C. A survey of content based 3D shape retrieval

methods. Multimed. Tools Appl. 39, 441. issn: 1573-7721 (2007).

309. Tao, J., Perdew, J. P., Staroverov, V. N. & Scuseria, G. E. Climbing the Density Functional

Ladder: Nonempirical Meta21Generalized Gradient Approximation Designed for Molecules

and Solids. Phys. Rev. Lett. 91, 146401 (2003).

310. Ten-no, S. & Noga, J. Explicitly correlated electronic structure theory from R12/F12

ansatze. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2, 114 (2012).

311. Tentscher, P. R. & Arey, J. S. Geometries and Vibrational Frequencies of Small Radicals:

Performance of Coupled Cluster and More Approximate Methods. J. Chem. Theory

Comput. 8, 2165 (2012).

312. The HDF Group. Hierarchical data format version 5 http://www.hdfgroup.org/HDF5.

313. Thiel, W. MNDO99, version 7.0 2004.

314. Thiel, W. Theoretical Chemistry—Quo Vadis? Angew. Chem. Int. Ed. 50, 9216 (2011).

315. Thomas, S. P., Jayatilaka, D. & Guru Row, T. N. S[three dots, centered]O chalcogen

bonding in sulfa drugs: insights from multipole charge density and X-ray wavefunction of

acetazolamide. Phys. Chem. Chem. Phys. 17, 25411 (2015).

316. Thornber, C. W. Isosterism and molecular modification in drug design. Chem. Soc. Rev.

8, 563 (1979).

317. Truhlar, D. G. Basis-set extrapolation. Chem. Phys. Lett. 294, 45 (1998).

318. Turner, M. J., McKinnon, J. J., Jayatilaka, D. & Spackman, M. A. Visualisation and

characterisation of voids in crystalline materials. CrystEngComm 13, 1804 (2011).

319. Vainio, M. J., Puranen, J. S. & Johnson, M. S. ShaEP: Molecular Overlay Based on

Shape and Electrostatic Potential. J. Chem. Inf. Model. 49, 492 (2009).

320. Valeev, E. F., Allen, W. D., Hernandez, R., Sherrill, C. D. & Schaefer, H. F. On the

accuracy limits of orbital expansion methods: Explicit effects of k-functions on atomic

and molecular energies. J. Chem. Phys. 118, 8594 (2003).

321. Valkenborg, D., Mertens, I., Lemiere, F., Witters, E. & Burzykowski, T. The isotopic

distribution conundrum. Mass Spectrom. Rev. 31, 96 (2012).

219

COMPLETE BIBLIOGRAPHY

322. Van der Walt, S., Colbert, S. C. & Varoquaux, G. The NumPy Array: A Structure for

Efficient Numerical Computation. Comput. Sci. Eng. 13, 22 (2011).

323. Varandas, A. J. C. Basis-set extrapolation of the correlation energy. J. Chem. Phys. 113,

8880 (2000).

324. Vardi, M. Y. Science has only two legs. Commun. ACM 53, 5 (2010).

325. Verheyen, E. et al. Design of zeolite by inverse sigma transformation. Nature Mat. 11,

1059 (2012).

326. Voronoi, G. Nouvelles applications des parametres continus a la theorie des formes

quadratiques. Premier memoire. Sur quelques proprietes des formes quadratiques positives

parfaites. J. reine angew. Math. 1908, 97 (1908).

327. Vosko, S. H., Wilk, L. & Nusair, M. Accurate spin-dependent electron liquid correlation

energies for local spin density calculations: a critical analysis. Can. J. Phys. 58, 1200

(1980).

328. Wan, K. X., Vidavsky, I. & Gross, M. L. Comparing similar spectra: from similarity index

to spectral contrast angle. J. Am. Soc. Mass Spectrom. 13, 85 (2002).

329. Wan, P. T. C. et al. Mechanism of activation of the RAF-ERK signaling pathway by

oncogenic mutations of B-RAF. Cell 116, 855 (2004).

330. Wang, J., Wolf, R. M., Caldwell, J. W., Kollman, P. A. & Case, D. A. Development

and testing of a general amber force field. Journal of Computational Chemistry 25, 1157

(2004).

331. Wang, J., Wolf, R. M., Caldwell, J. W., Kollman, P. A. & Case, D. A. Development and

testing of a general amber force field. J. Comput. Chem. 25, 1157 (2004).

332. Wang, S. Y. & Schaefer, H. F. The small planarization barriers for the amino group in

the nucleic acid bases. J. Chem. Phys. 124, 044303 (2006).

333. Warr, W. A. Balancing the needs of the recruiters and the aims of the educators in Book

of Abstracts, 218th ACS National Meeting (New Orleans, 1999).

334. Watts, J. D., Gauss, J. & Bartlett, R. J. Coupled-cluster methods with noniterative triple

excitations for restricted open-shell hartree-fock and other general single determinant

reference functions - energies and analytical gradients. J. Chem. Phys. 98, 8718 (1993).

335. Weber, W. & Thiel, W. Orthogonalization corrections for semiempirical methods. Theor.

Chem. Acc. 103, 495 (2000).

336. Weigend, F. & Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and

quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys.

Chem. Chem. Phys. 7, 3297 (2005).

220

COMPLETE BIBLIOGRAPHY

337. Werner, H.-J. et al. MOLPRO, version 2012.1, a package of ab initio programs see. 2012.

338. Werner, H.-J., Knowles, P. J., Knizia, G., Manby, F. R. & Schuetz, M. Molpro: a general-

purpose quantum chemistry program package. Wiley Interdiscip. Rev. Comput. Mol. Sci.

2, 242 (2012).

339. Weyl, H. The Classical Groups: Their Invariants and Representations 8th ed. (Princeton

University Press, 1973).

340. Wiberg, K. B. Basis set effects on calculated geometries: 6-311++G** vs. aug-cc-pVDZ.

J. Comput. Chem. 25, 1342 (2004).

341. Wigner, E. & Seitz, F. On the Constitution of Metallic Sodium. Phys. Rev. 43, 804

(1933).

342. Wiley Registry of Mass Spectral Data, 11th Edition http://wiley.com/WileyCDA/

WileyTitle/productCd-1119171016.html (2017).

343. Willett, P. Enhancing the Effectiveness of Ligand-Based Virtual Screening Using Data

Fusion. QSAR Comb. Sci. 25, 1143 (2006).

344. Willett, P., Barnard, J. M. & Downs, G. M. Chemical Similarity Searching. J. Chem. Inf.

Model. 38, 983 (1998).

345. Wilson, A. K. & Dunning, T. H. Benchmark calculations with correlated molecular wave

functions .10. Comparison with ”exact” MP2 calculations on Ne, HF, H2O, and N-2. J.

Chem. Phys. 106, 8718 (1997).

346. Wishart, D. S. et al. DrugBank: a comprehensive resource for in silico drug discovery and

exploration. Nucleic acids research 34, D668 (2006).

347. Xie, Y. M., Scuseria, G. E., Yates, B. F., Yamaguchi, Y. & Schaefer, H. F. Methylnitrene

- theoretical predictions of its molecular-structure and comparison with the conventional

c-n single bond in methylamine. J. Am. Chem. Soc. 111, 5181 (1989).

348. Xu, D. & Li, H. Geometric moment invariants. Pattern Recognit. 41, 240 (2008).

349. Yaghi, O. M. et al. Reticular synthesis and the design of new materials. Nature 423, 705

(2003).

350. Yanai, T., Fann, G. I., Gan, Z., Harrison, R. J. & Beylkin, G. Multiresolution quantum

chemistry in multiwavelet bases: Analytic derivatives for Hartree–Fock and density

functional theory. J. Chem. Phys. 121, 2866 (2004).

351. Zhao, Y. & Truhlar, D. G. The M06 suite of density functionals for main group thermo-

chemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition

elements: two new functionals and systematic testing of four M06-class functionals and

12 other functionals. Theor. Chem. Acc. 120, 215 (2008).

221

COMPLETE BIBLIOGRAPHY

352. Zupan, J. & Gasteiger, J. Neural networks: A new method for solving chemical problems

or just a passing phase? Anal. Chim. Acta 248, 1 (1991).

222