View metadata, citation and similar papers at core.ac.uk brought to you by CORE

provided by DSpace at Waseda University

Efficient Two-Component Relativistic Method to Obtain Electronic and Molecular Structures of Heavy-Element Systems

ͷిࢠঢ়ଶ͓Αͼ෼ࢠߏ଄ͷͨΊͷܥॏݩૉ ޮ཰తͳ 2 ੒෼૬ର࿦๏

February 2017

Waseda University Graduate School of Advanced Science and Engineering Department of Chemistry and Biochemistry, Research on Electronic State Theory

Yuya NAKAJIMA தౢɹ༟໵

Contents

Chapter 1 General introduction 1

References ...... 6

Chapter 2 Theoretical background 7

2.1 Hamiltonian ...... 7

2.2 The IODKH method ...... 9

2.3 The LUT scheme ...... 12

2.4 Spin-free and spin-dependent formalisms ...... 14

2.5 The FCP method ...... 16

2.6 The GHF method ...... 19

2.7 The Analytical energy derivative for the GHF method . . . 21

References ...... 24

Chapter 3 Analytical energy gradient for spin-free infinite-order

Douglas–Kroll–Hess method with local unitary transfor-

mation 25

3.1 Introduction ...... 25

3.2 Theory and implementation ...... 27

3.2.1 Energy gradient for IODKH ...... 27

3.2.2 Analytical derivative for the space transformation

matrices ...... 30

3.2.3 Energy gradient for LUT-IODKH ...... 32

3.2.4 Implementation ...... 33

i 3.3 Numerical assessments ...... 34

3.3.1 Computational details ...... 34

3.3.2 Numerical gradient values ...... 37

3.3.3 Accuracies of IODKH/ and LUT-IODKH/C methods 38

3.3.4 Computational cost of the LUT scheme ...... 45

3.3.5 Metal complexes ...... 47

3.3.6 Heavier analogues of ethylene ...... 48

3.3.7 Harmonic frequencies of diatomic molecules . . . . 49

3.4 Conclusion ...... 50

References ...... 52

Chapter 4 Implementation of spin-dependent relativistic analytical

energy gradient 57

4.1 Introduction ...... 57

4.2 Implementation ...... 60

4.3 Numerical assessments ...... 61

4.3.1 Computational details ...... 61

4.3.2 Accuracy of SD-IODKH and LUT-SD-IODKH . . . . 63

4.3.3 Computational cost of (LUT-)SD-IODKH method . . 67 − 4.3.4 Application in fac Ir(ppy3)...... 70 4.4 Conclusion ...... 72

References ...... 74

Chapter 5 Implementation of LUT-IODKH in GAMESS program 79

5.1 Introduction ...... 79

5.2 Implementation ...... 79

5.2.1 Relativistic correction in GAMESS program . . . . . 79

5.2.2 Threshold setting in the IOTC method ...... 80

ii 5.2.3 Integral evaluation at quadruple precision ...... 81

5.2.4 Combination with DC ...... 82

5.2.5 Input options for LUT-IOTC ...... 83

5.2.6 Major capabilities of LUT-IOTC ...... 83

5.3 Numerical assessment ...... 84

5.3.1 Computational details ...... 84

5.3.2 Total energies of heavy atoms and molecules . . . . 85

5.3.3 Threshold dependence of IOTC and LUT-IOTC . . . 86

5.3.4 Computational cost of LUT-IOTC ...... 87

5.4 Conclusion ...... 89

References ...... 91

Chapter 6 Relaxation of core orbitals in the frozen core potential treat-

ment 93

6.1 Introduction ...... 93

6.2 Theory and implementation ...... 94

6.2.1 FCP with relaxation of core electrons ...... 94

6.2.2 Implementation ...... 96

6.3 Numerical assessments ...... 98

6.3.1 Computational details ...... 98

6.3.2 Computational cost of FCP-CR ...... 99

6.3.3 Accuracy of FCP-CR ...... 103

6.3.4 Core ionization energy and core level shift ...... 104

6.3.5 Accuracy of an iterative procedure between valence

and core calculations ...... 107

6.4 Conclusion ...... 108

Appendix 6.A Dependence of core ionization potential energies 111

iii References ...... 113

Chapter 7 Relativistic effect on enthalpy of formation for transition

metal complexes 117

7.1 Introduction ...... 117

7.2 Computational details ...... 119

7.3 Results and discussion ...... 121

7.3.1 Accuracy of WFT ...... 121

7.3.2 Functional dependence ...... 123

7.3.3 Geometry difference between PP and AE methods . 124

7.3.4 Effect of the levels of relativistic Hamiltonians . . . . 124

7.3.5 Contribution of frozen core orbitals ...... 126

7.4 Conclusion ...... 128

References ...... 130

Chapter 8 General Conclusion 133

Acknowledgments 137

List of Achievements 139

iv List of abbreviations

2c two-component

4c four-component

AE all-electron

AO atomic orbital

BSS Barysz-Sadlej-Snijders

CA composite approach

CCSD singles and doubles

CCSD(T) coupled cluster singles, doubles, and perturbative triples

CIE core ionization energy

CLS core level shift

CPU central processing unit

DC divide-and-conquer

DFT density functional theory

DLU local approximation to the unitary decoupling transformation dmpe 1,2-bis(dimethylphosphino)ethane dOEI derivative of one-electron integral dppe 1,2-bis((pentafluorophenyl)phosphino)ethane dTEI derivative of two-electron integral

ECP effective core potential

ESC elimination of small components

F-dppe 1,2-bis-((pentafluorophenyl)phosphino)ethane

FCP frozen core potential

v FCP-CR frozen-core potential with relaxation of core electrons

FW Fouldy-Wouthuysen

GAMESS General Atomic and Molecular Electronic Structure System

GHF general Hartree–Fock

HF Hartree–Fock

HFR Hartree–Fock–Roothaan

IODK infinite-order Douglas–Kroll–Hess

IOTC infinite-order two-component

LUT local unitary tranformation

MAE mean absolute error

MaxE maximum error

MCP model core potential

MP model potential

MP2 second-order Møller–Plesset perturbation

NESC normalized elimination of the small component

NMR nuclear magnetic resonance

NR non-relativistic

OEI one-electron integral

PP pseudo-potential ppy 2–phenylpyridine

RA regular approximation

RECP relativistic effective core potential

RI resolution of identity

SAC-CI symmetry-adapted cluster configuration interaction

SCF self-consistent field

SD spin-dependent

SF spin-free

vi TEI two-electron integral

UT unitary transformation

WFT wave function theory

X2C exact two-component

vii

Chapter 1 General introduction

Quantum chemistry plays an essential role in the qualitative and/or quantitative prediction and analysis of energetics and molecular properties such as geometries, spectra, and reactivities. To date, the basic, commonly used equation of has been the NR Schrodinger¨ equation. In significant cases involving relativistic effects, such as orbital contractions and splittings in heavy-element systems, intersystem crossing, and core-electron related properties, relativistic effects are mostly accounted for through corrections to the NR treatment. On the other hand, using the Dirac equation, as a basic equation that satisfies the Lorentz invariance for electron motion, can account for relativistic effects. In 2002, Barysz et al. proposed a rigorous 2c method, referred to as the IODKH method1, that treats only electronic states. Seino and Nakai proposed the LUT technique that is able to perform efficient calculations without the loss of accuracy, using the original IODKH method2–6. The construction and extension of a theory based on both IODKH and LUT methods would enable the efficient treatment of relativistic effects using basic equations.

Another efficient method for the treatment of heavy elements is the ECP method, which reduces the number of electrons that are treated explicitly7. The

ECP method replaces the effect of core orbitals on the valence electrons with a potential. Constructing this potential to include relativistic effects provides a rel- ativistic treatment in a convenient manner. In 2014, Seino et al. proposed the FCP

1 method8, which describes core potentials by utilizing explicit core-orbital infor- mation obtained through atomic calculations. However, the effects of the chemical environment are not taken into account in the FCP method for molecules. This thesis extends the LUT-IODKH and FCP methods from a theoretical perspective, and applies the extended method to the determination of relativistic effects in significant systems. In order to perform geometry optimizations and frequency calculations, the analytical energy gradient for LUT-IODKH was developed and extended to the spin-dependent method9,10. This gradient method and the LUT scheme were then implemented in the GAMESS quantum chemical package11.

Furthermore, FCP-CR, which relaxes the core orbitals that were treated as frozen orbitals by FCP, was also developed12. These methods were utilized for enthalpy of formation calculations, in the gas phase, of transition metal complexes.

This thesis includes seven chapters, in addition this general introduction chapter (Chapter 1).

Chapter 2 summarizes the theoretical background of the 2c relativistic method, as well as the IODKH, LUT, and FCP methods. Moreover, the energy and gradient expressions for the GHF method are provided.

Chapter 3 extends the IODKH method to include analytical energy gradi- ents, and gradients combined with the LUT scheme. The energy gradient of the

IODKH Hamiltonian, with respect to nuclear coordinates, is analytically derived.

Numerical assessments of equilibrium bond lengths in diatomic molecules con- taining heavy elements indicate that the accuracy of this method is close to that of 4c. When compared to electron-correlation methods such as MP2 and DFT, good agreements with equilibrium bond lengths are observed. Especially for heavy-element systems, the simultaneous consideration of both electron corre- lation and relativistic effects was found to be required for accurate calculations.

For harmonic frequency and force constant calculations, relativistic effects were

2 found to contribute to 10% of the obtained values, even for first-row transition metals. In order to treat larger-sized molecules, the LUT technique is applied to the gradient calculation. The resulting CPU time required for relativistic unitary transformations in 1d- to 3d-silver clusters was found to be similar to those of NR methods, and scaled quasi-linearly with respect to the number of atoms.

Following on from the previous chapter, Chapter 4 extends and implements the SD energy gradient. The NR and SF relativistic methods are only described within real number space because no spin operator, represented by Pauli spin matrices, are included. To treat the spin operator explicitly, the GHF method, and its energy gradient with respect to nuclear coordinates, is derived from the viewpoint of its implementation within complex number space. In addition, the IODKH and LUT-IODKH methods are extended to the SD treatment. These enable efficient, self-consistent, relativistic geometry optimizations that include

SD relativistic effects. The atomic forces obtained by the analytical SD-IODKH gradient method are enough close to the corresponding numerical gradients in atomic units. Equilibrium bond lengths of diatomic molecules containing fifth- and sixth-row elements are close to those calculated using 4c methods. A com- parison of the SF and SD relativistic effects reveals that the contributions of these effects depend on the type of orbital partitioning within the bond. In addition, the CPU time for the relativistic transformation scales quasi-linearly with respect to the number of atoms.

Chapter 5 implements the LUT-IODKH method into the GAMESS quantum chemical package and assesses its performance numerically. GAMESS includes relativistic correction functionalities based on 2c methods such as NESC, RESC,

first- to third-order DKH, and IOTC. One of them, IOTC, is equivalent to IODKH from a theoretical point of view. Based on the original code for IOTC, the LUT technique, and its gradient extension, has been implemented in order to perform

3 LUT-IOTC calculations, such as excited-state calculations, in combination with

GAMESS functionalities. In order to verify the integrity of this implementation, the total energies of heavy atoms and molecules are calculated. Several of the

IOTC program-default settings were altered to obtain numerically reliable results.

Furthermore, when combined with the DC method, which has been implemented since 2009 in GAMESS, overall computational linear-scaling is achieved, from relativistic transformations to electron correlation method such as CCSD(T).

Chapter 6 describes studies into theoretically improving FCP for describing the core orbitals of molecules. Conventional FCP only describes valence orbitals of molecules accurately. In this chapter, FCP is extended by relaxing the core orbitals that have been treated as frozen orbitals, by means of valence-orbital potentials, for the accurate descriptions of molecular core orbitals. This enables accurate AE calculations using core potentials. Numerical assessments of gold clusters, composed of up to 86 atoms, indicate that FCP-CR is faster than AE. The orbital energies and total energies of coinage metal dimers are as accurate as those obtained by AE. Furthermore, because FCP-CR possesses core-orbital wavefunc- tions, it can provide core-electron related properties. Using these wavefunctions the CIEs of the 1s orbitals of C in vinyl acetate, and of the 4f orbitals of W in tungsten complexes, are qualitatively reproduced.

Chapter 7 discusses the evaluation of the gas-phase enthalpies of formation of transition metal complexes using the methods developed in the previous chapters.

These complexes include early and late transition metals, light and heavy met- als, low- and high-valency complexes, soft and hard ligands, and organometallic complexes. Contributions of relativistic effects to the enthalpies of formation, cal- culated using NR, ECP and FCP methods, are assessed and compared. In addition, the correlation between experimental and DFT-calculated data is examined.

Finally, future prospects are discussed after the results of the individual

4 chapters are briefly summarized.

5 References

(1) Barysz, M.; Sadlej, A. J. J. Chem. Phys. 2002, 116, 2696.

(2) Seino, J.; Nakai, H. J. Chem. Phys. 2012, 136, 244102.

(3) Seino, J.; Nakai, H. J. Chem. Phys. 2012, 137, 144101.

(4) Seino, J.; Nakai, H. J. Chem. Phys. 2013, 139, 034109.

(5) Seino, J.; Nakai, H. J. Comput. Chem. Jpn. 2014, 13,1.

(6) Seino, J.; Nakai, H. Int. J. Quantum Chem. 2015, 115, 253.

(7) Schwerdtfeger, P. Chem. Phys. Chem 2011, 12, 3143.

(8) Seino, J.; Tarumi, M.; Nakai, H. Chem. Phys. Lett. 2014, 592, 341.

(9) Nakajima, Y.; Seino, J.; Nakai, H. J. Chem. Phys. 2013, 139, 244107.

(10) Nakajima, Y.; Seino, J.; Nakai, H. J. Chem. Theory Comput. 2016, 15, 68.

(11) Schmidt, M. W.; Baldridge, K. K.; Boatz, J. A.; Elbert, S. T.; Gordon, M. S.;

Jensen, J. H.; Koseki, S.; Matsunaga, N.; Nguyen, K. A.; Su, S.; Windus, T. L.;

Dupuis, M.; Montgomery Jr., J. A. J. Comput. Chem. 1993, 14, 1347.

(12) Nakajima, Y.; Seino, J.; Nakai, H. Chem. Phys. Lett. 2016, 663, 97.

6 Chapter 2 Theoretical background

Relativistic quantum chemistry is based on the Dirac Hamiltonian, which con- tains electronic and positronic states. In the 2c relativistic treatment, the electronic state (or large component thereof), which is essential for chemistry, is only treated explicitly after several Hamiltonian manipulations, so as to ESCs and apply UTs that are directly related to the accuracy and efficiency of the approximations.

Moreover, the relativistic 2c Hamiltonian includes spin operators, or Pauli spin matrices, which require a generalized treatment of SCF-calculated wavefunctions in order to achieve highly accurate results. In this chapter, brief reviews of theoret- ical backgrounds are provided and include the IODKH Hamiltonian1, for accurate relativistic transformations, and its derivatives; the LUT2 and FCP3 schemes for efficient relativistic transformations without the loss of accuracy; and the GHF method, and its derivatives, for explicit use with spin operators. In addition, methods for splitting the SF and SD terms are also presented.

2.1 Dirac Hamiltonian

Relativistic quantum chemistry is based on the Dirac equation in order to include relativistic effects. The Dirac Hamiltonian is a one-particle Hamiltonian, which is

7 expressed as follows: ⎛ ⎞ ⎛ ⎞ ⎜ ⎟ ⎜ ⎟ ⎜ψ ⎟ ⎜ψ ⎟ ⎜ 1⎟ ⎜ 1⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ψ ⎟ ⎜ψ ⎟ ⎜ 2⎟ = ⎜ 2⎟ h4 ⎜ ⎟ E ⎜ ⎟ (2.1) ⎜ ⎟ ⎜ ⎟ ⎜ψ ⎟ ⎜ψ ⎟ ⎜ 3⎟ ⎜ 3⎟ ⎜ ⎟ ⎜ ⎟ ⎝ψ ⎠ ⎝ψ ⎠ 4 4 with ⎛  ⎞ ⎜ ⎟ ⎜ − ⎟ ⎜ V 0 cpz c px ipy ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ 0 Vc(p + ip ) −cp ⎟ = ⎜ x y z ⎟ , h4 ⎜  ⎟ (2.2) ⎜ ⎟ ⎜ cp c p − ip V − 2c2 0 ⎟ ⎜ z x y ⎟ ⎜ ⎟ ⎝ + − − 2 ⎠ c(px ipy) cpz 0 V 2c where V is a scalar potential, px, py, and pz are momentum operators along with x-, y-, and z-axis, respectively, and c is the speed of light. Note that atomic units ψ are used hereafter. The upper two wavefunction components in Eq. (2.1), 1 and ψ ψ 2, are referred to as the large components, while the lower two components, 3 ψ and 4, are referred to as the small components. These two types of wavefunction are interpreted to be the electronic and positronic wavefunctions, respectively.

Individual wavefunction components correspond to α- and β-spin, respectively.

Furthermore, using Pauli spin matrices, the Dirac Hamiltonian can be rewritten as follows: ⎛ ⎞ ⎜ σ · p ⎟ ⎜ V12 c ⎟ = ⎜ ⎟ h4 ⎜ ⎟ (2.3) ⎝ σ · p − 2 ⎠ c V12 2c 12 with  σ = σ ,σ ,σ , x y z (2.4) ⎛ ⎞ ⎜ ⎟ ⎜ ⎟ ⎜01⎟ σ = ⎜ ⎟ , (2.5) x ⎝⎜ ⎠⎟ 10 ⎛ ⎞ ⎜ ⎟ ⎜ − ⎟ ⎜0 i⎟ σ = ⎜ ⎟ , (2.6) y ⎝⎜ ⎠⎟ i 0

8 and ⎛ ⎞ ⎜ ⎟ ⎜ ⎟ ⎜10⎟ σ = ⎜ ⎟ , (2.7) z ⎝⎜ ⎠⎟ 0 −1

× where 1n is the n n identity matrix.

2.2 The IODKH method

As mentioned above, the Dirac equation includes both electronic and positronic information. In general chemistry however, more attention is given to the elec- tronic information, which, conceptually, leads to a decoupling of the electronic and positronic states. The Dirac Hamiltonian is block-diagonalized using a uni- tary transformation U, which is expressed as: ⎛ ⎞ ⎜ + ⎟ ⎜h 0 ⎟ = † = ⎜ 2 ⎟ , h2 U h4U ⎜ ⎟ (2.8) ⎝⎜ −⎠⎟ 0h2

+ − where h2 and h2 are the electronic and positronic Hamiltonians, respectively. Practical calculations use only the electronic Hamiltonian. In this thesis, the

IODKH method, which is one of the most accurate relativistic transformation schemes, is used. Many electronic, two-component, relativistic Hamiltonians are generally expressed by:

= + + ++ , , H2 h2 (i) g2 (i j) (2.9) i i>j

+ ++ where h2 and g2 are one- and two-particle Hamiltonians, respectively. For the + one-particle Hamiltonian, h2 , the IODKH Hamiltonian was adopted, and is given by: + = Ω† Ω, h2 G (2.10) where the second-step transformation, Ω, is defined by

Ω = + † −1/2, (12 Y Y) (2.11)

9 and the Hamiltonian transformed in Eq. (2.10), G, is expressed as

p2 = 2 + † + † + Π† σ · p σ · p Π, G p b12 Y − 12 Y M VM ( V ) (2.12) 1 ep with = + , M K(12 bpY) (2.13)

Π = α − −1 , K( b12 p Y) (2.14)

+ 1/2 ep 1 K = , (2.15) 2ep

= 1 , b + (2.16) ep 1  1/2 = + α2 2 , ep 1 p (2.17) and

α = 1/c. (2.18)

Here, p is the momentum operator and V is the sum of the nucleus-electron

= attraction operators with respect to an atom A, i.e., V A VA. The operator Y is obtained by the decoupling conditions:  + = α3 − −1 σ · p σ · p epY Yep pKbVK p K V bK  − − + α2 p 1Kσ · pVσ · pp 1KY − YKVK   (2.19) + α4 pKbVbKpY − YKbσ · pVσ · pbK  − + α3Y Kbσ · pVσ · pKp 1 − KVKbp Y.

The present thesis focuses on the relativistic effect on the one-particle interac-

++ tion. Hence, the two-particle Hamiltonian g2 is replaced as the NR Coulomb interaction: 1 g(i, j) = . (2.20) rij

10 This treatment is denoted as IODKH/C. It should be noted that relativistic two- particle interactions can be included using the Breit-Pauli approximation,4 as well as the FW5 and IODKH transformations.6

In the pragmatic evaluation of Eq. (2.10),the matrix transformation method and the RI technique proposed by Hess are adopted.7 The first step in the technique of Hess is the determination of the eigenvectors, {k},ofp2, which are expanded by the linear combination of the primitive functions {φ},

| = |φ , k r Crk (2.21) r ffi where the coe cient Crk is determined by diagonalizing a matrix whose element     φ  2  φ { } is r p s . The resulting (pseudo) eigenvectors, k , satisfy:        2  = ω δ , k p k k kk (2.22)

ω { } where k is a (pseudo) eigenvalue. The explicit matrix expressions, in k -space, for Eqs. (2.10)–(2.19) can be written as:  + † = Ω Ω , h Gk k k k (2.23) 2 kk kk k,k       − /   † 1 2  Ω = + , kk k  12 Y Y  k (2.24)          2   2  †  p  =   +   Gkk k p b12 k Ykk k  − 12  k Yk k 1 ep k ,k  (2.25) † † + | | +Π | σ · p σ · p | Π , Mkk k V k Mk k kk k V k k k k,k with   = + , Mkk Kk 1 bkpkYkk (2.26)  − Π = α − 1 , kk Kk bk pk Ykk (2.27)

11 and ⎡ ⎢  1 ⎢ − = ⎢α3 | | − 1 | σ · p σ · p | Ykk ⎢ pkKkbk k V k Kk p Kk k V k bk Kk + ⎣ k ek ek  − − + α2 1 | σ · p σ · p | 1 − | | pk Kk k K V k pk Kk Yk k Ykk Kk k V k Kk k  + α4 | | pkKkbk k V k bk Kk pk Yk k k − | σ · p σ · p | Ykk Kk bk k V k bk Kk )  − + α3 | σ · p σ · p | 1 Ykk Kk bk k V k Kk pk k,k   − | | . Kk k V k Kk bk pk Yk k

(2.28)

Here, the approximate RI relationship is inserted between the individual opera- tors, which is expressed as

|kk|≈1. (2.29) k  The matrix elements, which are the functions of p2, f p2 , such as the operators

K, b, and p operators in Eqs. (2.15)–(2.17) become:        2  = ω δ . k f (p ) k f ( k) kk (2.30)

2.3 The LUT scheme

The IODKH method achieves high accuracy compared to the 4c method although the computational cost associated with the relativistic transformation is consider- ably large. On the basis of the nature of unitary transformations and relativistic effects, approximated unitary transformations can be constructed in order to im- prove the efficiency. In this section, the concept of the LUT scheme and the

LUT-IODKH Hamiltonian are provided.

The LUT scheme arises from the localization of the relativistic effect. The

12 entire one-particle unitary transformation U can be approximated in a block- diagonal matrix form consisting of the subsystem contributions:

UTotal ≈ UA ⊕ UB ⊕ UC ··· , (2.31) where {A, B, C, ···}denotes the subsystems. Here, atomic partitioning is adopted.

The relativistic effect on the kinetic energy is dominant at each atom while the effect on the nuclear attraction is dominant not only at each atom, but also on its interactions within a small distances τ. The matrix representation of the one-particle Hamiltonian with basis function {χ} is expressed as         χA  +  χB ≈ χA  LUT  χB μ h2 ν μ h2 ν ⎧       ⎪   ⎪ A  + + NR  B ⎪ χμ  T + V + V  χν (A = B) ⎪  A C  ⎪  CA  ⎪    ⎨⎪   (2.32) χA  + + + + NR + NR  χB  , ≤ τ = ⎪ μ  V V T V  ν (A B RAB ) ⎪  A B C  ⎪   CA,B ⎪    ⎪   ⎪ A NR NR B ⎪ χμ  T + V  χν (A  B, R >τ) ⎩  C  AB C

NR NR where T and VC denote the standard NR kinetic and nuclear-attraction oper- ators: 1 TNR = p21 , (2.33) 2 2 and NR = . VC VC (2.34)

+ + T and VC are the relativistic operators given by: p2 + = Ω† 2 + † Ω, T p b12 Y − 12 Y (2.35) 1 ep and + = Ω† † + Π† σ · p σ · p Π Ω. VC M VCM ( VC ) (2.36)

13 Note that the operators Y, Ω, M, and Π are determined within individual sub- systems because U contains only single-subsystem information. Hence, the com- putational cost of the relativistic transformation using the LUT scheme scales linearly with respect to the system size.

2.4 Spin-free and spin-dependent formalisms

SF and SD terms are considered separately in standard relativistic quantum chem- istry. One of the reasons for this is that the SF formalism is a straightforward extension to NR theory. This subsection explains how the SF and SD formalisms are obtained.

In order to obtain the SF formalism, the spin-related terms containing Pauli spin matrices are separated by the Dirac relationship8,

σ · pVσ · p = pV · p + iσ · (pV × p) (2.37) where i denotes the imaginary unit. Since the dot products and cross products can be evaluated in {χ}-space, the explicit expression containing these terms is given by:     χ | p · p | χ = χ  · + · + ·  χ V pxV px pyV py pzV pz (2.38) and ⎛    ⎞ ⎜   ⎟ ⎜ 0 χ  (pV × p)  χ ⎟ χ | σ · p × p | χ = ⎜ y ⎟ i ( V ) ⎜    ⎟ ⎝ χ  p × p  χ ⎠ ( V )y 0 ⎛ ⎞     (2.39) ⎜     ⎟ ⎜ χ  (pV × p)  χ χ  (pV × p)  χ ⎟ + ⎜ z x ⎟ i ⎜     ⎟ ⎝ χ  p × p  χ − χ  p × p  χ ⎠ ( V )x ( V )z with       χ  p × p  χ = χ  · − ·  χ , ( V )x pyV pz pzV py (2.40)

      χ  p × p  χ = χ  · − ·  χ , ( V )y pxV pz pzV px (2.41)

14 and       χ  p × p  χ = χ  · − ·  χ . ( V )z pxV py pyV px (2.42)

The momentum operator p is performed on the basis set {χ}. Partial integrations of Eqs. (2.40)–(2.42) leads to:                 ∂χ   ∂χ ∂χ   ∂χ χ  (pV × p)  χ =  V  −  V  , (2.43) x ∂y ∂z ∂z ∂y                   ∂χ   ∂χ ∂χ   ∂χ χ  (pV × p)  χ =  V  −  V  , (2.44) y ∂x ∂z ∂z ∂x and                 ∂χ   ∂χ ∂χ   ∂χ χ  (pV × p)  χ =  V  −  V  . (2.45) z ∂x ∂y ∂y ∂x

In these equations, x, y, and z refer to electronic coordinates of electrons. In the SF formalism, the IODKH and LUT-IODKH Hamiltonians are expressed as follows:  + † = Ω Ω , h Gk k k k (2.46) 2 kk kk k,k       − /   † 1 2  Ω = + , kk k  12 Y Y  k (2.47)          2   2  †  p  =   +   Gkk k p b12 k Ykk k  − 12  k Yk k 1 ep k ,k  (2.48) † † + | | +Π | p · p | Π , Mkk k V k Mk k kk k V k k k k,k with   = + , Mkk Kk 1 bkpkYkk (2.49)  − Π = α − 1 , kk Kk bk pk Ykk (2.50)

15 and ⎡ ⎢  1 ⎢ − = ⎢α3 | | − 1 | p · p | Ykk ⎢ pkKkbk k V k Kk p Kk k V k bk Kk + ⎣ k ek ek  − − + α2 1 | p · p | 1 − | | pk Kk k V k pk Kk Yk k Ykk Kk k V k Kk k  + α4 | | pkKkbk k V k bk Kk pk Yk k k (2.51) − | p · p | Ykk Kk bk k V k bk Kk )  − + α3 | p · p | 1 Ykk Kk bk k V k Kk pk k,k   − | | , Kk k V k Kk bk pk Yk k and ⎧       ⎪   ⎪ A  + + NR  B ⎪ χμ  T + V + V  χν (A = B) ⎪  A C  ⎪  CA  ⎪        ⎨⎪   A  LUT  B χA  + + + + NR + NR  χB  , ≤ τ χμ h χν = ⎪ μ  V V T V  ν (A B RAB ) 2 ⎪  A B C  ⎪   CA,B ⎪    ⎪   ⎪ A NR NR B ⎪ χμ  T + V  χν (A  B, R >τ) ⎩  C  AB C (2.52) with + = Ω† † + Π† p · p Π Ω. VC M VCM ( VC ) (2.53)

2.5 The FCP method

The IODKH method, when combined with the LUT scheme, extends the appli- cability of relativistic quantum chemistry to the calculation of relatively large molecules without loss of accuracy. However, all-electron calculations still have problems that include large basis set requirements for heavy elements, prob- lematic convergence during SCF calculations, and the demand for high memory and disk capacities. For the further extension of relativistic quantum chemistry, decreasing the number of electrons treated explicitly offers a solution to these

16 problems. In this subsection, the FCP method, that provides a solution to the problems mentioned above, is explained.

All-electron calculations are based on the HFR equation:

F(D)C = SCE (2.54) with

Fμν(D) = Hμν + Gμν(D), (2.55) 1 Gμν(D) = Dλρ χμχρ|g|χνχλ− χμχρ|g|χλχν , (2.56) λρ 2 and occ = , Dμν 2 CμiCνi (2.57) i where F is the Fock matrix; D is the density matrix, C is the orbital coefficient matrix, S is the overlap matrix, E is the orbital energy matrix, H is the one-electron integral, G is the two-electron integral, g is the two-electron Hamiltonian, χ are basis functions, while “occ” denotes occupied orbitals, and C is the complex conjugate of C.

The density matrix is divided into core and valence contributions as follows:

D = DC + DV (2.58) with core C = Dμν 2 CμcCνc (2.59) c and valence V = . Dμν 2 CμvCνv (2.60) v In the case of basis sets that segmented and contracted, the contributions from core and valence AOs to the core-density matrix, are large (μ and ν in core AOs) and small (μ or ν in valence AOs), respectively. The core AOs contribute moderately

17 to the valence-density matrix. In the case of generally contracted basis sets, the core and valence density matrices are negligibly affected by valence and core

AOs, respectively, since the individual core and valence regions are sufficiently described by core and valence AOs. When generally contracted basis sets are adopted, the entire density matrix can be approximately represented in a block- diagonalized form, as follows:

D ≈ dC ⊕ dV, (2.61) where dC and dV are the core- and valence-contributing density matrices, respec- tively, which include only core and valence AO indices. Since the individual atomic core-density matrices contribute mainly to the molecular core-density ma- trix, the introduction of the direct sum approximation to the core-density matrix sufficiently reproduces the original molecular core-density matrix:

C ≈ ˜ C = ˜ C ⊕ ˜ C ⊕··· D D DA DB (2.62) and C ≈ ˜ C = ˜ C ⊕ ˜ C ⊕···, d d dA dB (2.63)

C where {A, B, ···}are individual atoms. The atomic core-density matrices, D˜ and C d˜ , correspond to the core-density matrices obtained from the individual atomic calculations, which are frozen for molecular calculations.

These two approximations lead to the Huzinaga-Cantu equation9, which is expressed as:

fV(dV)cV = sVcVεV (2.64) with   V V V V C C = μν + − + − + μν, , fμν(d ) H 2Jμν Kμν 2Jμν,A Kμν,A P A (2.65) A A

18 valence         V V V V V   V V 1 V V   V V 2Jμν − Kμν = dλρ χμ χρ g χν χλ − χμ χρ g χλ χν , (2.66) λ,ρ 2 core          C C C V C   V C 1 V C   C V 2Jμν, − Kμν, = d˜λρ χμ χρ g χν χλ − χμ χρ g χλ χν , (2.67) A A A 2 λ,ρ∈A and

core        = − ε χV  ϕ ϕ  χV . Pμν,A 2 c,A μ c,A c,A ν (2.68) c∈A  ϕ ε C C Here, c,A is the core orbital, c,A is the core orbital energy, J and K are Coulomb and exchange terms for the core region obtained by atomic all-electron calcula- tions, JV and KV are Coulomb and exchange terms for the valence region, and P represents projection/shift terms. Finally, the total, core, and valence energies are expressed as follows:

ETotal = EC + EV (2.69) with    C − C + C 1 nA ZB nB nB ZA EC = tr DC H + F DC + (2.70) 2 RAB A

C where ZA is the nuclear charge, nA is the number of core electrons in atom A, and

RAB is the distance between atoms A and B.

2.6 The GHF method

The explicit consideration of the effect of SD in quantum chemical calculation requires a description of spin mixing. The standard HF method only describes the spin parallel to the z-axis. In order to describe an arbitrary direction of

19 electronic spin, the GHF method should be adopted. Hence, in this section, the energy expression using the GHF method is presented.

In the GHF method, the wave function is expressed using a linear- combination of α- and β-spin orbitals:

α β ω ψ = φμα + φμβ = φμω, i Cμi Cμi Cμi (2.73) μ μ ω=α,β μ where C is the coefficient matrix, ω is the spin index, and φ is the spacial orbital.

The HF energy is generally expressed as:    1     EHF = [i | h | i] + ij ij − ij ji + V . (2.74) 2 nuc i i,j

Here, EHF denotes the HF electronic energy, i and j are the electronic indices,     [i | h | i] is a one-electron Hamiltonian matrix, and ij ij and ij ji are the NR two- electron interactions, i.e., Coulomb and exchange terms, respectively, while Vnuc is the nuclear repulsion potential. Note that the one-electron Hamiltonian contains the spin operator, or Pauli spin matrices in the case of the relativistic Hamiltonian.

Using Eq. (2.73), the energy expression for the one-electron Hamiltonian matrix can be rewritten as follows:   ωω   [i | h | i] = Dμν φμω  h  φνω (2.75) i ω=α,β ω=α,β μ,ν with

ωω = ω ω , Dμν CμiCνi (2.76) i where D is the density matrix. Likewise, the Coulomb and exchange terms can be obtained:    1  1 ωω ττ  ij ij = Dμν Dλρ φμωφνω φλτφρτ (2.77) 2 2 i,j ω,ω τ,τ μ,ν,λ,ρ and    1  1 ωτ τω  ij ji = DμλDνρ φμωφντ φλτφρω , (2.78) 2 2 i,j ω,ω τ,τ μ,ν,λ,ρ

20 where τ and τ are spin indices. By the integration of Eqs. (2.77) and (2.78) over spin functions, the entire two-electron integrals become:    1 ωω ττ ωτ τω  Dμν Dλρ δωω δττ − DμλDνρ δωτ δτω φμφν φλφρ , (2.79) 2 ω,ω τ,τ μ,ν,λ,ρ   where δ is the Kronecker delta and φμφν  φλφρ is the two-electron integral in a given basis set. Finally the GHF energy can be provided:   HF ωω   E = Dμν φμω  h  φνω ω,ω μ,ν    1 ωω ττ ωτ τω  + Dμν Dλρ δωω δττ − DμλDνρ δωτ δτω φμφν φλφρ (2.80) 2 ω,ω τ,τ μ,ν,λ,ρ

+ . Vnuc

2.7 The Analytical energy derivative for the GHF method

In the previous section, the energy expression for the GHF method, which con- tains a summation over electronic spins, is provided. To perform the geometry optimization based on the GHF method requires a differentiated GHF expression.

This section presents the analytical energy gradient for the GHF method.

The total energy from the GHF method is expressed by Eq. (2.80). Differ- entiating the total energy with respect to a nuclear coordinate RA provides the

21 following expression: ⎛ ⎞ ⎛ ⎞ ∂ ωω ∂ ωω ∂ HF ⎜ hμν ⎟ ⎜ Dμν ⎟ E = ωω ⎜ ⎟ + ⎜ ⎟ ωω ∂ Dμν ⎝ ∂ ⎠ ⎝ ∂ ⎠ hμν RA ω,ω μ,ν RA ω,ω μ,ν RA    1 ωω ττ ωτ τω ∂  + δ δ − δ δ φ φ  φ φ Dμν Dλρ ωω ττ DμλDνρ ωτ τω ∂ μ ν λ ρ 2 ω,ω τ,τ μ,ν,λ,ρ RA ⎛ ⎞ ⎜∂ ωω ⎟   1 ⎜ Dμν ττ ωτ τω ⎟  + ⎜ δ δ − δ δ ⎟ φ φ  φ φ ⎝ ∂ Dλρ ωω ττ DμλDνρ ωτ τω ⎠ μ ν λ ρ 2 ω,ω τ,τ μ,ν,λ,ρ RA ⎛ ττ ⎞ ⎜ ∂D ⎟   1 ⎜ ωω λρ ωτ τω ⎟  + ⎜ δωω δττ − δωτ δτω ⎟ φμφν  φλφρ ⎝Dμν ∂ DμλDνρ ⎠ 2 ω,ω τ,τ μ,ν,λ,ρ RA ⎛ ωτ ⎞ ⎜ ∂D ⎟   1 ⎜ ωω ττ μλ τω ⎟  + ⎜ δ δ − δ δ ⎟ φ φ  φ φ ⎝Dμν Dλρ ωω ττ ∂ Dνρ ωτ τω ⎠ μ ν λ ρ 2 ω,ω τ,τ μ,ν,λ,ρ RA ⎛ ⎞ ⎜ ∂ τω ⎟   1 ⎜ ωω ττ ωτ Dνρ ⎟  + ⎜ δ δ − δ δ ⎟ φ φ  φ φ ⎝Dμν Dλρ ωω ττ Dμλ ∂ ωτ τω ⎠ μ ν λ ρ 2 ω,ω τ,τ μ,ν,λ,ρ RA ∂ + Vnuc . ∂ RA (2.81)

The GHF-SCF Fock matrix, and its derivative for molecular orbitals at i = j under orthonormal conditions are as follows: ⎛ ⎞ ⎜ ⎟   ωω ωω ⎜ ττ τ ω ⎟  = + ⎜ δωω δττ − δωτ δτω ⎟ φμφν  φλφρ Fμν hμν ⎝⎜Dλρ CλiCjρ ⎠⎟ (2.82) τ,τ λ,ρ i,j and ⎛ ⎞ ⎛ ⎞ ⎜∂ ω ⎟ ∂ ωω ⎜ C μ ⎟ ⎜ Sμν ⎟ ⎜ i ⎟ ωω ω + + ω ⎜ ⎟ ω = , ⎝⎜ ∂ ⎠⎟ Sμν Ciν c.c. Ciμ ⎝ ∂ ⎠ Ciν 0 (2.83) μ,ν RA μ,ν RA where S is the overlap matrix and c.c. is the complex conjugate. Eq. (2.81) can be

22 rewritten as: ⎛ ⎞ ∂ ωω ∂ HF ⎜ hμν ⎟ E = ωω ⎜ ⎟ ∂ Dμν ⎝ ∂ ⎠ RA ω,ω RA    1 ωω ττ ωτ τω ∂  + δ δ − δ δ φ φ  φ φ Dμν Dλρ ωω ττ DμλDνρ ωτ τω ∂ μ ν λ ρ 2 ω,ω τ,τ μ,ν,λ,ρ RA ⎛ ⎞ ωω ⎜∂ μν ⎟ ∂ ωω ⎜ S ⎟ Vnuc + ⎜ ⎟ δ + , Wμν ⎝ ∂ ⎠ ωω ∂ ω,ω RA RA (2.84)

ε where W is the energy-weighted density matrix with the orbital energy i,

ωω = ε ω ω . Wμν iCμiCiν (2.85) i

23 References

(1) Barysz, M.; Sadlej, A. J. J. Chem. Phys. 2002, 116, 2696.

(2) Seino, J.; Nakai, H. J. Chem. Phys. 2012, 136, 244102.

(3) Seino, J.; Tarumi, M.; Nakai, H. Chem. Phys. Lett. 2014, 592, 341.

(4) Bethe, H. A.; Salpeter, E. E., Quantum Mechanics of One- and Two-electron

Atoms; Springer, Berlin: 1957.

(5) Foldy, L. L.; Wouthuysen, S. A. Phys. Rev. 1950, 78, 29.

(6) Seino, J.; Hada, M. Chem. Phys. Lett. 2008, 461, 327.

(7) Hess, B. A. Phys. Rev. A 1985, 32, 756.

(8) Reiher, M.; Wolf, A., Relativistic Quantum Chemistry; Wiley-VCH, Germany:

2009.

(9) Bonifacic, V.; Huzinaga, S. J. Chem. Phys. 1974, 60, 2779.

24 Chapter 3 Analytical energy gradient for spin-free infinite-order Douglas–Kroll–Hess method with local unitary transformation

3.1 Introduction

Reliable predictions of geometries and properties for molecules including heavy elements are some of the most important tasks in quantum chemical calcula- tions. Notably, in case of calculations for heavy (and super heavy) elements and their compounds, the inclusion of relativistic effect is essential because electrons, especially core electrons, have considerably high speeds. For example, bond contractions/elongations due to relativistic shrinking/expansion of molecular or- bitals1–4 were observed in several diatomic molecules such as halogen dimers and

PtAu.5,6 The evaluations of first- and higher-order derivatives of the molecular energy enables computations of optimized geometries, various molecular prop- erties such as molecular vibrations observed in infrared and Raman spectra and chemical shifts observed in NMR spectra. Accordingly, the development of ana- lytical energy derivatives with the relativistic Hamiltonian is desired in order to extend the applicability of the quantum chemical method.

A highly accurate treatment in relativistic quantum chemistry is based on the 4c relativistic theory, which adopts a one-particle Dirac operator that satisfies

25 Lorentz invariance only in terms of the motion of each electron. Numerical applications of the analytical energy gradient method with the 4c treatments confirmed the high accuracy in determining molecular geometries,1 and electric and magnetic properties.7,8 Such applications, however, were suitable only in the case small- or medium-sized molecules because of the practical problems involving computational costs and/or variational conditions9 due to the existence of small-component Dirac spinors or positronic-state spinors. There have been several attempts to resolve these problems.10–12

An alternative approach is based on the 2c relativistic theory, which treats only electronic-state (or large-component) information. Recent developments in the 2c treatment of many-electron molecular systems have exhibited a higher de- gree of accuracy, in comparison with 4c treatment.13–18 The energy gradient meth- ods for accurate 2c relativistic schemes are under development. The 2c relativistic theories are categorized into two types. The first approach uses an equation for

ESC, such as RA,19–21 RESC,22 and NESC.23 The analytical energy gradient schemes have been made available for the zeroth-order RA and its scaling schemes,24–26

RESC,27 and NESC.28,29 The second approach uses the decoupling matrix trans- formation of 4c Dirac Hamiltonian and includes approaches such as the exact quasi-relativistic method.30 In this type of method we separate the electronic and positronic Hamiltonian from 4c Dirac Hamiltonian or Fock matrix using unitary transformation. The widely used schemes are the FW transformation31 and DKH transformation.32 These were extended to higher-order DKH33–36 and IODKH37 methods. The molecular property calculation for this type of scheme is based on the second-order38–40 and extended35,41,42 DKH Hamiltonian.

Nakai and co-workers have worked on the development of an accurate and efficient 2c relativistic scheme, termed LUT, at the one- and two-particle SF-

IODKH levels.43,44 The LUT method utilizes the locality of the relativistic effect.

26 Accordingly, the computational costs become drastically lower than those of con- ventional methods and scale linearly with respect to the system size with a small prefactor. In addition, the LUT scheme clarifies the primary relativistic effect in molecular systems, because the deviations in total energy are less than milli- hartrees. Recently, the LUT scheme has been combined with the linear-scaling

DC based on HF45–47 and electron correlation48–53 methods, such as MP2 and

CCSD.54 The combination of LUT and DC techniques leads to the first approach that achieves overall (quasi-)linear scaling with a small prefactor for relativistic electron correlation calculations.

In this chapter, the author derives the analytical energy gradient of the one- particle IODKH Hamiltonian for the accurate calculations. Furthermore, the

LUT scheme is introduced for efficient calculations for large-sized molecules.

Theoretical aspects are explained in Section 3.2. In Section 3.3, the present scheme is numerically assessed from the viewpoints in accuracy and efficiency. The concluding remarks are finally provided in Section 3.4.

3.2 Theory and implementation 3.2.1 Energy gradient for IODKH

In the HF method, the first derivative of the electronic energy with respect to a coordinate RA of a nucleus A is written as ⎛  ⎞ ⎜∂ + ⎟ ∂ ⎜ h μν ⎟ ∂ μλ||νρ ∂ ∂ μν E ⎜ 2 ⎟ 1 ( ) Vnuc S = μν ⎜ ⎟+ μν λρ + − ε μν , ∂ D ⎜ ∂ ⎟ D D ∂ ∂ iD ∂ RA ⎝ RA ⎠ 2 RA RA RA μ,ν μ,ν,λ,ρ i μ,ν (3.1) where μ, ν, λ, and ρ are the indices of {χ};(μλ||νρ), the antisymmetric electron {χ} repulsion integral in ; D, the density matrix; Vnuc, the nuclear repulsion poten- tial; ε, the orbital energy; and S, the overlap matrix. In the expression of Eq. (3.1), ∂ μλ||νρ /∂ ∂ /∂ ∂ /∂ ( ) RA, Vnuc RA, and Sμν RA are identical to those in the NR theory.

27 The IODKH Hamiltonian in {χ} and that in {k} have the following relation

+ {χ} = † † + { } , h2 ( ) d X h2 ( k ) Xd (3.2) with

X = ΓΛ, (3.3)

ffi + {χ} + { } where d denotes the contraction coe cients. h2 ( ) and h2 ( k ) are the IODKH Hamiltonians in {χ} and {k}, respectively. Γ is the transformation matrix from space {χ} to the orthonormalized space, while Λ is the transformation matrix { } + {χ} from the orthonormalized space to k . The analytical derivative for h2 ( ) is expressed as ∂ + {χ} ∂ † ∂ + { } ∂ h2 ( ) = † X + + † h2 ( k ) + † + X , ∂ d ∂ h2 X X ∂ X X h2 ∂ d (3.4) RA RA RA RA with ∂ + { } ∂Ω† ∂ ∂Ω h2 ( k ) = Ω + Ω† G Ω + Ω† . ∂ ∂ G ∂ G∂ (3.5) RA RA RA RA

Here, we can obtain the derivatives of the matrix elements of Ω and G by differ- entiating individual matrix elements with respect to {k} in Eqs. (2.24)–(2.28),      − / † ∂Ω  1 † 3 2 ∂Y † ∂Y  kk =  − + +  , ∂ k  12 Y Y ∂ Y Y ∂  k (3.6) RA 2 RA RA and           ∂ ∂ ∂  2  Gkk  2  †  p  =   +   ∂ ∂ k p b12 k ∂ Ykk k  − 12  k Yk k RA RA RA 1 ep k ,k  ∂ † + | | ∂ Mkk k V k Mk k (3.7) RA k ,k  ∂ † + Π | σ · p σ · p | Π , ∂ kk k V k k k RA k ,k with

∂ ∂   ∂   Mkk Kk = + + , ∂ ∂ 1 bkpkYkk Kk ∂ bkpkYkk (3.8) RA RA RA

28   ∂Π ∂ ∂ kk Kk − − = α − 1 + α − 1 , ∂ ∂ bk pk Ykk Kk ∂ bk pk Ykk (3.9) RA RA RA and ∂ ∂ ∂ Ykk 1 ek ek = − − Ykk Ykk ∂ + ∂ ∂ RA ek ek RA RA ∂   + α3 | | ∂ pkKkbk k V k Kk RA ∂  3 −1 − α p K k | σ · pVσ · p | k b K ∂R k k k k A  ∂ − − + α2 1 | σ · p σ · p | 1 ∂ pk Kk k V k pk Kk Yk k RA k ∂ − α2 | | ∂ (Ykk Kk k V k Kk ) RA k ∂   + α4 | | ∂ pkKkbk k V k bk Kk pk Yk k RA k ∂ − α4 | σ · p σ · p | ∂ (Ykk Kk bk k V k bk Kk ) RA k  ∂ − + α3 | σ · p σ · p | 1 ∂ Ykk Kk bk k V k Kk pk Yk k RA k k ∂   − α3 | | . ∂ Ykk Kk k V k Kk bk pk Yk k RA k k (3.10)

Note that vectors such as K, b, and p in Eqs. (3.6)–(3.10) have the information about molecular structure through the function {k}. Thus, the elements should be differentiated as ⎛ ⎞ ∂ ⎜ ∂p2 ⎟ Kk = −1α2 −3 −1 ⎜ k ⎟ , ∂ ek Kk ⎝∂ ⎠ (3.11) RA 8 RA ⎛ ⎞ ∂ ⎜ ∂p2 ⎟ bk = −1α2 −1 2 ⎜ k ⎟ , ∂ ek bk ⎝∂ ⎠ (3.12) RA 2 RA ⎛ ⎞ ∂ ⎜ ∂p2 ⎟ ek = 1α2 −1 ⎜ k ⎟ , ∂ ek ⎝∂ ⎠ (3.13) RA 2 RA and ⎛ ⎞ − ∂p 1 ⎜ ∂p2 ⎟ k = −1 −3 ⎜ k ⎟ . ∂ pk ⎝∂ ⎠ (3.14) RA 2 RA

29 3.2.2 Analytical derivative for the space transformation matrices

In order to obtain explicit derivatives of the space transformation matrix X and

∂ 2/∂ pk RA, the same procedure as that in Ref. [38] is adopted. First, the procedure for the derivative of Γ in Eq. (3.3) is explained. Γ is determined by the condition

ΓTSΓ = 1, (3.15) with the overlap matrix S. In more detail, S is first diagonalized as

WTSW = s. (3.16)

Γ is obtained using the unitary transformation W and the eigenvalues of s

− / Γ = Ws 1 2. (3.17)

The explicit expression of the derivative of Γ is as follows:

∂Γ ∂ ∂ = W −1/2 − 1 −3/2 s . ∂ ∂ s Ws ∂ (3.18) RA RA 2 RA

The derivative of s is given by

∂ ∂ T ∂ ∂ s = W + T S + T W ∂ ∂ SW W ∂ W W S ∂ (3.19) RA R A RA RA ∂ = T S − , , W ∂ W [PW s]− (3.20) RA with ∂ = T W . PW W ∂ (3.21) RA

Here, square brackets represent the anti-commuter/commuter of matrices as

[A, B]± = AB ± BA. (3.22)

ff ∂ /∂ Note that the o diagonal elements of s RA and the diagonal elements of PW are zero because s is a diagonal matrix and PW is an antisymmetric matrix. From Eq.

30 (3.20) and these conditions, the matrix expression of PW is obtained as T ∂ /∂ W ( S RA)W kk = − δ . (PW) (1 kk ) (3.23) kk − sk sk ∂ /∂ ∂ /∂ Thus, s RA is obtained by substituting Eq. (3.23) to Eq. (3.20). Since W RA is also given by ∂ W = , ∂ WPW (3.24) RA ∂Γ/∂ ∂ /∂ ∂ /∂ RA is determined from Eq. (3.18) using s RA and W RA.

The procedure for the derivative of Λ in Eq is explained. (3.3). The first step is a transformation of TNR from {φ} to the orthonormalized space as

† T({φ}) = Γ TNRΓ. (3.25)

By diagonalizing T({φ}), Λ is obtained:

† T({k}) = Λ T({φ})Λ. (3.26)

Here, the derivative of T({k}) is expressed as

∂ { } ∂ {φ} T( k ) = Λ† T( ) Λ − , { } , ∂ ∂ [PV T( k )]− (3.27) RA RA with ∂Λ = Λ† . PV ∂ (3.28) RA ∂ { } /∂ ∂ /∂ Note that T( k ) RA has a relationship with p RA:

∂ ∂ T = p . ∂ 2∂ (3.29) RA RA ∂ {φ} /∂ ff To obtain T( ) RA, Eq. (3.25) is di erentiated with respect to RA : ∂T({φ}) ∂ ∂ = −1/2 TW −1/2 − 1 −1 s , {χ} , ∂ s ∂ s s ∂ T( ) (3.30) RA RA 2 RA + where = T NR . TW W T W (3.31)

31 Eq. (3.30) is rewritten using the derivative of TW:

∂ ∂ NR TW = T T − , . ∂ W ∂ W [PW TW]− (3.32) RA RA

For the properties of Eq. (3.27), offdiagonal elements on the left-hand side in { } Eq. (3.27) are zero because T( k ) is diagonal. The diagonal elements of PV are zero due to the antisymmetry of PV. These properties provide  Λ† ∂ {φ} /∂ Λ T( ) RA kk = − δ . (PV) (1 kk ) (3.33) kk { } − { } T( k )k T( k )k

Finally, the derivative of Λ becomes

∂Λ = Λ . ∂ PV (3.34) RA

Using the derivatives of Γ and Λ, the derivative of the space transformation matrix X can be obtained:

∂ ∂Γ ∂Λ X = Λ + Γ . ∂ ∂ ∂ (3.35) RA RA RA

3.2.3 Energy gradient for LUT-IODKH

This subsection explains the analytical derivative of the LUT-SF-IODKH Hamil- tonian for efficient gradient evaluations. The derivative expression of the LUT-

SF-IODKH matrix of Eq. (2.32) is given by ⎧     ⎪   ⎪ ∂   ⎪ χA  NR  χB = ⎪∂ μ  VC  ν (A B) ⎪ RA   ⎪ CA ⎪ ⎪ ∂     ⎪ χA  + + +  χB     ⎪ μ V V ν ∂   ⎨⎪∂R  A B  χA  LUT  χB = ⎪ A     ∂ μ h2 ν ⎪ ∂   RA ⎪ A  NR NR  B ⎪+ χμ  T + V  χν (A  B, R ≤ τ) ⎪ ∂  C  AB ⎪ RA  , ⎪   C A B   ⎪   ⎪ ∂   ⎪ χA  NR + NR  χB  , >τ, ⎩∂ μ  T VC  ν (A B RAB ) RA   C (3.36)

32 where + ∂V ∂ ∂ C = Ω† † VC + Π† p · p Π Ω. ∂ M ∂ M ∂ VC (3.37) RA RA RA

Note that the IODKH transformation is only applied to the nucleus-electron in- teraction for different atoms with a small distance τ.IfA = B, the derivative of + + the IODKH transformed matrices such as T and VC become zero because the transformations of the LUT-IODKH method have closure forms within individ-  ≤ τ + ual atoms. If A B with RAB , the nuclear attraction elements VC for the ff LUT-IODKH transformation are obtained by di erentiation with respect to RA as well as the NR kinetic terms TNR and the NR nuclear attraction ones VNR.

3.2.4 Implementation

This subsection presents the scheme for evaluating the energy gradient with the

(LUT-) IODKH Hamiltonian. Table 3.1 summarizes the procedures for (a) NR, (b)

IODKH, and (c) LUT-IODKH. For the NR/C case, the density matrix D and orbital energy ε obtained from energy calculation are preserved in the energy gradient calculation. Then the derivatives of NR one-electron integrals TNR and VNR are calculated in contracted formalism.

For the conventional IODKH case, integrals and momentum factors in uncon-

NR NR Ω Π −1 tracted formalism such as S, T , V , , Y, G, h, M, , K, b, ep, and p as well as the density matrix D and orbital energy ε obtained from energy calculation are preserved in energy gradient calculation. Furthermore, the derivatives of NR and relativistic integral, S, TNR, VNR, and σ · pVσ · p, are calculated in uncontracted formalism. Using S and TNR, the derivatives of transformation matrices Γ and

Λ are calculated using Eqs. (3.15)-(3.34). At the same time, the derivatives of kinematic factors of Eqs. (3.11)-(3.14) can be calculated. Using kinematic factors and their derivatives, the derivatives of Ω, G, and Y in {k} are calculated using

Eqs. (3.6), (3.7), and (3.10), respectively. Finally the derivative of relativistic one-

33 particle Hamiltonian in Eq. (3.5) is obtained by back transformation into {χ} using the derivatives of Γ and Λ.

For the LUT-IODKH case, the procedure becomes simpler than the procedure Ω Π of the conventional IODKH. First, the transformation factors , Y, M, , K, b, ep, − and p 1 in each subsystem as well as D and ε are preserved in energy calculation.

Then, the NR integrals for TNR and VNR are calculated in contracted formalism.

In order to obtain the transformed integrals for interatomic interactions with ≤ τ NR σ · p σ · p RAB in Eq. (3.36), the derivatives of V and V are calculated in uncontracted formalism. After that, the transformation is performed using Eq.

(3.37). Thus, the derivative of LUT-IODKH Hamiltonian matrix can be obtained.

Note that this complicated procedure involving Eqs. (3.15)-(3.34) is unnecessary in the case of the LUT-IODKH scheme. In addition, the range of transformation is extremely small. Consequently, the LUT-IODKH scheme is suitable for the efficient calculation for analytical derivative of one-electron integral.

3.3 Numerical assessments 3.3.1 Computational details

This section presents the assessment of the performance of the analytical energy gradient for the IODKH and LUT-IODKH schemes. We adopted the spin-free for- malism of IODKH/C: the one-electron Dirac Hamiltonian and two-electron non- relativistic Coulomb interaction. For LUT, the threshold for cutoff of relativistic interaction, τ, was set to 3.5 Å. For comparison, the spin-free 4c Dirac-Coulomb and non-relativistic (NR/C) methods were also calculated. The calculations of parts of small-component integrals, i.e., (SS|SS)-type ones, were skipped in the 4c method.55

Numerical tests were performed for various closed-shell diatomic molecules: = hydrogen halides HX, halogen dimers X2 (X F, Cl, Br, I, and At), coinage metal

34 Table 3.1 Procedures for computing the analytical energy gradient of (a) NR, (b) IODKH, and (c) LUT-IODKH. (a) NR (b) IODKH (c) LUT-IODKH ε NR NR Ω ε D, , S, T , V , , Y, G, h, ε Ω Π −1 (I) From energy calculation D, −1 D, , , Y, M, , K, b, ep, p M, Π, K, b, ep, p (for whole subsystem) (for each system)

∂TNR ∂VNR ∂TNR ∂VNR (II) Derivatives of NR one-electrton integrals ∂ , ∂ – ∂ , ∂ RA RA RA RA (except for interatomic interac- ≤ τ tion with RAB ) Derivatives of NR one-electron integrals ∂S ∂TNR ∂VNR ∂(pV·p) ∂VNR ∂(pV·p) (III) for realtivistic transformation (uncon- – ∂ , ∂ , ∂ , ∂ ∂ , ∂ RA RA RA RA RA RA 35 tracted form) (for interatomic interaction ≤ τ with RAB ) ∂Γ ∂Λ (IV) Derivatives of transformation matrices – ∂ , ∂ [from Eqs. (3.18) and (3.34)] – RA RA ∂ ∂ −1 + ∂K ∂b ep p ∂V (V) Derivatives of relativistic one-electron integrals – ∂ , ∂ , ∂ , ∂ [from Eqs. (3.11)-(3.14)] ∂ [from Eq. (3.37)] RA RA RA RA RA ∂Y (for interatomic interaction ∂ [from Eq. (3.10)] ≤ τ RA with RAB ) ∂M ∂Π ∂ , ∂ [from Eqs. (3.8) and (3.9)] RA RA ∂Ω ∂G ∂ , ∂ [from Eqs. (3.6) and (3.7)] RA RA ∂(μλ||νρ) (VI) Derivatives of NR two-electron integral ∂ RA (VII) Energy gradient calculation from Eq. (3.1). = hydrides MH, and coinage metal dimers M2 (M Cu, Ag, and Au). In addition, one-, two-, and three-dimentional silver clusters Agn were calculated to investi- ffi gate the e ciency of the LUT scheme. Figure 3.1 illustrates the structures of Ann = , , ··· , clusters. For one-dimensional silver clusters with Agn (n 1 2 10), the Ag-

Ag bond distance used here was 2.706 Å. For two- and three-dimensional silver clusters, a face-centered cubic structure with the experimental lattice constant,

4.086 Å56, was adopted. The unit cells for two- and three-dimensional clusters contain five and 14 atoms, respectively. We adopted several sizes of Ann clusters; the total numbers of atoms in the two-dimensional clusters were 5, 13, 25, 41, 61,

85, 113, 145, and 181, which correspond to 1 × 1, 2 × 2, ···, and 9 × 9 unit cells, respectively. The total numbers of atoms in the three-dimensional clusters were

14, 23, 32, 38, 53, 63, 74, 88, 123, and 172, which correspond to 1 × 1 × 1, 2 × 1 × 1,

3×1×1, 2×2×1, 3×2×1, 2×2×2, 3×3×1, 3×2×2, 3×3×2, and 3×3×3. For more practical systems, 20 metal complexes including the fourth–sixth row-transition metals were examined. Heavier analogues of ethylene were employed to investi- gate the relativistic effect in geometry. For harmonic frequency calculation, MH and M2 were utilized. The calculations were performed at the HF, MP2, and density functional the- ory (DFT) with B3LYP functional levels.57 The basis sets for all-electron calculation in diatomic molecules and 20 metal complexes are summarized in Table 3.2. All basis sets were used in uncontracted form. For the 4c method, the kinetically bal- anced small-component basis sets were adopted. MCP/NOSeC-V-TZP58–68 and

Stuttgart/Dresden (SDD69–72) type effective core potential are used in RECP. In

SDD calculation, the first–third row atoms were described with cc-pVTZ.73,74 The

IODKH energy gradient method with/without the LUT scheme was implemented in the GAMESS program.75 The 4c calculations were performed in the DIRAC12 program.76 The Gaussian09 program was used in SDD calculations.77

36 Figure 3.1 Geometry of (a) one-dimensional silver cluster Ag10, (b) two-dimensional silver × × × cluster Ag13 (2 2), and (c) three-dimensional silver cluster Ag63 (2 2 2).

Table 3.2 Basis sets for diatomic molecules and 20 metal complexes used at the levels of HF, MP2, and B3LYP in NR, IODKH, and large component of 4c. Period Element Description Basis sets 1H(6s3p2d) Ref. [78] 2 C,N,O,F (12s8p3d2 f ) Refs. [59, 79] 3 P, Cl (16s11p4d2 f ) Refs. [60, 80] 4 Ti, V (19s13p7d3 f ) Refs. [62, 80] Cr (20s13p8d3 f ) Refs. [62, 80] Co, Ni (20s12p9d3 f ) Refs. [62, 80] Cu (20s12p8d3 f ) Refs. [62, 80] Zn (20s13p9d3 f ) Refs. [62, 80] Br (20s15p10d2 f ) Refs. [63, 79] 5 Zr (22s17p13d3 f ) Refs. [64, 79] Nb (22s17p12d3 f ) Refs. [64, 79] Mo (22s16p11d3 f ) Refs. [64, 79] Ru (21s16p12d3 f ) Refs. [64, 79] Ag (22s16p12d3 f ) Refs. [64, 79] Cd (22s16p13d3 f ) Refs. [64, 79] I (22s18p13d2 f ) Refs. [63, 79] 6 Ta, W, Re (26s22p16d12 f ) Refs. [66, 79] Os, Pt, Au, Hg (26s21p16d12 f ) Refs. [66, 79] At (26s23p17d13 f ) Refs. [65, 79]

3.3.2 Numerical gradient values

This subsection investigates the accuracy of the analytical gradient values for

IODKH/C and LUT-IODKH/C Hamiltonians by comparing with the numerical

37 ones in HAt, At2, AuH, and Au2 molecules. Table 3.3 compares between the ana- lytical and numerical energy gradient values calculated by the NR/C, IODKH/C,

LUT-IODKH/C, and 4c at the HF level. We examined the gradient values in 1.0,

1.5, 2.0, and 3.0 times of the experimental bond length, re, which were 1.742, 3.046,

81 5 82 83 1.5324, and 2.4719 Å for HAt ,At2 , AuH , and Au2 , respectively. The numer- ical gradient calculations were performed using a central difference method with

0.01 Å stepsize. The differences between the analytical gana and numerical gnum energy gradient values, that is, Δgana, are shown in parentheses, which represent the correctness of the analytical scheme. The deviations from 4c, Δg4c, are given in square brackets, which represent the two-electron relativistic effect of contribution of (SS|SS)-type integrals.

In IODKH/C, the absolute deviations from the numerical results, i.e., Δgana, are less than 0.0004 Hartree/Bohr. Such small deviations confirm the correctness of the present implementation for the analytical energy gradient of IODKH/C.

| Δ 4c | The deviations from 4c, i.e., g , are comparatively large in At2 and ff | Au2. This is due to the lack of two-electron relativistic e ect in 2c or (SS SS)-type integral in 4c. In LUT-IODKH/C, | Δgana | are also small. The largest deviation is 0.0001 Hartree/Bohr. In addition, the absolute values of differences between

IODKH/C and LUT-IODKH/C, which are given in | ΔgLUT |, are less than 0.0003

Hartree/Bohr. This means that the LUT treatment evaluates the relativistic energy gradient without loss of accuracy.

3.3.3 Accuracies of IODKH/C and LUT-IODKH/C methods

This subsection examines the accuracy of the analytical energy gradient scheme for IODKH/C Hamiltonian. Table 3.4 shows the bond lengths (Å) of diatomic molecules calculated by the NR/C, IODKH/C, and 4c methods at the HF level.

The difference Δrrel of IODKH/C and 4c methods from NR/C are given in square

38 / / / / / Table 3.3 Analytical and or numerical gradient values (Hartree Bohr) of HAt, At2, AuH, and Au2 calculated by NR C, IODKH C, LUT-IODKH C, and 4c at HF level. NR/C IODKH/C LUT-IODKH/C4c Mole. lengtha) gana gana gnum Δganab) Δg4cc) gana gnum Δganab) ΔgLUTd) gana

HAt 1.0re 0.0113 0.0192 0.0195 (-0.0003) [-0.0002] 0.0194 0.0193 ( 0.0001) [ 0.0002] 0.0194 1.5re 0.0675 0.0657 0.0654 ( 0.0003) [ 0.0001] 0.0655 0.0655 ( 0.0000) [-0.0002] 0.0656 2.0re 0.0379 0.0366 0.0365 ( 0.0001) [-0.0001] 0.0365 0.0366 (-0.0001) [-0.0001] 0.0367 e) 3.0re 0.0098 0.0096 0.0097 (-0.0001) – 0.0096 0.0096 ( 0.0000) [ 0.0000] N. C. At2 1.0re 0.0256 0.0304 0.0305 (-0.0001) [-0.0110] 0.0304 0.0305 (-0.0001) [ 0.0000] 0.0414 1.5re 0.0273 0.0259 0.0255 ( 0.0004) [-0.0047] 0.0256 0.0256 ( 0.0000) [-0.0003] 0.0306 2.0re 0.0082 0.0079 0.0078 ( 0.0001) [-0.0027] 0.0078 0.0078 ( 0.0000) [-0.0001] 0.0106 39 3.0re 0.0019 0.0019 0.0015 ( 0.0004) [-0.0012] 0.0019 0.0020 (-0.0001) [ 0.0000] 0.0031 AuH 1.0re 0.0823 0.0135 0.0135 ( 0.0000) [ 0.0011] 0.0135 0.0135 ( 0.0000) [ 0.0000] 0.0124 1.5re 0.0336 0.0540 0.0540 ( 0.0000) [ 0.0000] 0.0540 0.0541 (-0.0001) [ 0.0000] 0.0540 2.0re 0.0313 0.0365 0.0365 ( 0.0000) [-0.0001] 0.0365 0.0365 ( 0.0000) [ 0.0000] 0.0367 3.0re 0.0149 0.0138 0.0138 ( 0.0000) [-0.0002] 0.0138 0.0138 ( 0.0000) [ 0.0000] 0.0140 Au2 1.0re 0.0798 0.0305 0.0304 ( 0.0001) [ 0.0117] 0.0305 0.0305 ( 0.0000) [ 0.0000] 0.0188 1.5re 0.0170 0.0270 0.0270 ( 0.0000) [-0.0053] 0.0268 0.0269 (-0.0001) [-0.0002] 0.0323 2.0re 0.0127 0.0134 0.0133 ( 0.0001) [-0.0031] 0.0133 0.0133 ( 0.0000) [-0.0001] 0.0165 3.0re 0.0045 0.0036 0.0037 (-0.0001) [-0.0013] 0.0036 0.0035 ( 0.0001) [ 0.0000] 0.0048 a) re denotes the experimental bond length taken from [5, 81–83]. b) Difference in analytical gradient value from the numerical one is given in parentheses. c) Difference in analytical gradient value from 4c is given in square brackets. d) Difference between IODKH and LUT-IODKH in analytical gradient value is given in square brackets. e) SCF calculations were not converged. brackets, representing the relativistic effect accounted by IODKH/C and 4c. The deviations Δr4c of IODKH/C from 4c are shown in parentheses, indicating the accuracy of the IODKH/C treatment.

ff Δ rel The relativistic e ect on bond lengths, r , is negative except for F2 and Cl2.

The absolute values increase as heavier elements are added to the systems. For example, the values in halogen dimers are 0.000, 0.001, -0.002, -0.012, and -0.036

Å for F2,Cl2,Br2,I2, and At2, respectively. This might be the common tendency in the relativistic effect. Furthermore, the deviations in halogen molecules are smaller than those in coinage metal molecules. This is because the coinage metal molecules have σ-bonding formed by each s valence orbital, which is largely shrunken by the relativistic effect in comparison with σ-bonding formed by p valence orbitals seen in halogen molecules.

The deviations of IODKH/C from 4c, Δr4c, are comparatively small. The largest deviation is 0.010 Å for At2. This indicates that the spin-free two-electron relativistic corrections reasonably demonstrate the bond lengths of elements, at least, up to the sixth row of the periodic table.

Table 3.5 shows the bond lengths (Å) of diatomic molecules calculated by the NR/C and IODKH/C at the HF, MP2, and B3LYP levels. The differences Δrcorr of NR/C and IODKH/C methods with respect to the corresponding HF results are given in square brackets, representing the correlation effect accounted by

NR/C and IODKH/C. The deviations Δrexp of NR/C and IODKH/C methods from reference are shown in parentheses, indicating the accuracy of the NR/C and

IODKH/C treatment. In additions, the mean absolute error (MAE) and the mean error (ME) are given for Δrexp.

In the MP2 results, the correlation effect on bond lengths, Δrcorr, of IODKH/C shows the same tendency as NR/C except for HBr and HAt. Hydrogen halides except for HAt and halogen dimers have positive values, whereas coinage metal

40 = = / / Table 3.4 Bond lengths (Å) of HX, X2 (X F, Cl, Br, I, and At), MH, and M2 (M Cu, Ag, and Au) molecules obtained by NR C, IODKH C and 4c methods at HF level. NR IODKH/C4c Mole. rNR rIODKH Δrrela) Δr4cb) r4c Δrrela) HF 0.897 0.896 [-0.001] (-0.001) 0.897 [ 0.000] HCl 1.264 1.263 [-0.001] (-0.001) 1.264 [ 0.000] HBr 1.470 1.404 [-0.066] (-0.001) 1.405 [-0.065] HI 1.608 1.600 [-0.008] ( 0.000) 1.600 [-0.008] HAt 1.711 1.684 [-0.027] ( 0.001) 1.683 [-0.028] F 1.328 1.328 [ 0.000] ( 0.000) 1.328 [ 0.000]

41 2 Cl2 1.975 1.976 [ 0.001] ( 0.000) 1.976 [ 0.001] Br2 2.274 2.272 [-0.002] (-0.002) 2.274 [ 0.000] I2 2.675 2.663 [-0.012] (-0.002) 2.665 [-0.010] At2 2.890 2.854 [-0.036] ( 0.010) 2.844 [-0.046] CuH 1.568 1.540 [-0.028] ( 0.000) 1.540 [-0.028] AgH 1.778 1.700 [-0.078] ( 0.001) 1.699 [-0.079] AuH 1.828 1.568 [-0.260] (-0.001) 1.569 [-0.259] Cu2 2.440 2.401 [-0.039] (-0.001) 2.402 [-0.038] Ag2 2.818 2.706 [-0.112] ( 0.000) 2.706 [-0.112] Au2 2.924 2.604 [-0.320] ( 0.002) 2.602 [-0.322] a) Relativistic correction is in square brackets. b) Difference in bond length from that obtained using 4c is in parentheses. halides and coinage metal dimers have negative values. The deviations of NR/C from reference, Δrexp, increase as the heavier elements are present in the molecules.

On the other hand, the deviations of IODKH/C from reference remain approx- imately constant. For example, the values in coinage metal halides are -0.042,

-0.034, and -0.037 Å for CuH, AgH, and AuH, respectively. Furthermore, MAE and ME of IODKH/C are relatively smaller than those of NR/C. This indicates that both correlation and relativistic effects are essential for the molecules containing heavier elements.

In the B3LYP results, the correlation effects on bond lengths, Δrcorr,of

IODKH/C and NR/C show similar trends except for HBr. The deviations of

/ Δ exp NR C from reference, r , are positive except for F2. The values increase as the elements in the system are heavier. For example, the values in coinage dimers are 0.062, 0.151, and 0.332 Å for Cu2,Ag2, and Au2, respectively. The deviations / of IODKH C are positive except for F2 and CuH and slightly increases in the case of heavier elements. MAE and ME of IODKH/C are smaller than those of NR/C because both correlation and relativistic effects are considered.

Table 3.6 shows the bond length (Å) of diatomic molecules calculated by the

IODKH/C and LUT-IODKH/C at the HF, MP2, and B3LYP levels. The deviation

ΔrLUT of the LUT-IODKH/C method from IODKH/C is shown in parentheses, indicating the accuracy of the LUT-IODKH/C treatment.

The deviations are both positive and negative. The largest absolute deviations in HF, MP2, and B3LYP are 0.014 Å for At2, 0.013 Å for At2, and 0.009 Å for

HAt, respectively. Since MAE and ME are reasonably small, the LUT-IODKH/C treatment can sufficiently describe the relativistic effect for estimating the bond lengths.

42 = = / Table 3.5 Bond lengths (Å) of HX, X2 (X F, Cl, Br, I, and At), MH, and M2 (M Cu, Ag, and Au) molecules calculated by the NR C and IODKH/C at HF, MP2 and B3LYP level. HF MP2 B3LYP Exptl. Ref. NR IODKH NR IODKH NR IODKH Mole. rNR rIODKH rNR Δrcorra) Δrexpb) rIODKH Δrcorra) Δrexpb) rNR Δrcorra) Δrexpb) rIODKH Δrcorra) Δrexpb) HF 0.897 0.896 0.917 [ 0.020] (-0.001) 0.917 [ 0.021] (-0.001) 0.922 [ 0.025] ( 0.004) 0.922 [ 0.026] ( 0.004) 0.9176 [83] HCl 1.264 1.263 1.270 [ 0.006] (-0.005) 1.270 [ 0.007] (-0.005) 1.281 [ 0.017] ( 0.006) 1.281 [ 0.018] ( 0.006) 1.2749 [83] HBr 1.470 1.404 1.412 [-0.058] (-0.003) 1.410 [ 0.006] (-0.005) 1.425 [-0.045] ( 0.010) 1.423 [ 0.019] ( 0.008) 1.4146 [83] HI 1.608 1.600 1.611 [ 0.003] ( 0.001) 1.603 [ 0.003] (-0.007) 1.624 [ 0.016] ( 0.014) 1.618 [ 0.018] ( 0.008) 1.6099 [83] HAt 1.711 1.684 1.714 [ 0.003] — 1.682 [-0.002] — 1.725 [ 0.014] — 1.696 [ 0.012] — — —

43 F2 1.328 1.328 1.400 [ 0.072] (-0.012) 1.400 [ 0.072] (-0.012) 1.398 [ 0.070] (-0.014) 1.398 [ 0.070] (-0.014) 1.4119 [83] Cl2 1.975 1.976 1.985 [ 0.010] (-0.003) 1.984 [ 0.008] (-0.004) 2.012 [ 0.037] ( 0.024) 2.012 [ 0.036] ( 0.024) 1.9879 [83] Br2 2.274 2.272 2.284 [ 0.010] ( 0.003) 2.281 [ 0.009] (-0.000) 2.316 [ 0.042] ( 0.035) 2.313 [ 0.041] ( 0.032) 2.2811 [83] I2 2.675 2.663 2.677 [ 0.002] ( 0.011) 2.669 [ 0.006] ( 0.003) 2.708 [ 0.033] ( 0.042) 2.702 [ 0.039] ( 0.036) 2.6663 [83] At2 2.890 2.854 2.892 [ 0.002] — 2.860 [ 0.006] — 2.919 [ 0.029] — 2.883 [ 0.029] — — — CuH 1.568 1.540 1.451 [-0.117] (-0.015) 1.424 [-0.116] (-0.042) 1.482 [-0.086] ( 0.016) 1.459 [-0.081] (-0.007) 1.4658 [82] AgH 1.778 1.700 1.662 [-0.116] ( 0.044) 1.584 [-0.116] (-0.034) 1.697 [-0.081] ( 0.079) 1.625 [-0.075] ( 0.007) 1.6179 [82] AuH 1.828 1.568 1.709 [-0.119] ( 0.177) 1.495 [-0.073] (-0.037) 1.746 [-0.082] ( 0.214) 1.538 [-0.030] ( 0.006) 1.5324 [82] Cu2 2.440 2.401 2.240 [-0.200] ( 0.021) 2.203 [-0.198] (-0.016) 2.281 [-0.159] ( 0.062) 2.248 [-0.153] ( 0.029) 2.2192 [84] Ag2 2.818 2.706 2.591 [-0.227] ( 0.061) 2.505 [-0.201] (-0.025) 2.681 [-0.137] ( 0.151) 2.595 [-0.111] ( 0.065) 2.5303 [85] Au2 2.924 2.604 2.710 [-0.214] ( 0.238) 2.468 [-0.136] (-0.004) 2.804 [-0.120] ( 0.332) 2.547 [-0.057] ( 0.075) 2.4719 [83] MAE 0.042 0.014 0.072 0.023 ME 0.037 -0.013 0.070 0.020 a) Difference in bond length from the corresponding HF result is given in squre brakets. b) Difference in bond length from the experimental result is given in parentheses. = = Table 3.6 Bond length (Å) of HX, X2 (X F, Cl, Br, I, and At), MH, and M2 (M Cu, Ag, and Au) molecules calculated by the IODKH and LUT-IODKH at HF, MP2 and B3LYP level. HF MP2 B3LYP Molecule rIODKH rLUT ΔrLUTa) rIODKH rLUT ΔrLUTa) rIODKH rLUT ΔrLUTa) HF 0.896 0.896 ( 0.000) 0.917 0.917 ( 0.000) 0.922 0.922 ( 0.000) HCl 1.263 1.264 ( 0.001) 1.270 1.270 ( 0.000) 1.281 1.281 ( 0.000) HBr 1.404 1.404 ( 0.000) 1.410 1.410 ( 0.000) 1.423 1.423 ( 0.000) HI 1.600 1.601 ( 0.001) 1.603 1.604 ( 0.001) 1.618 1.618 ( 0.000) HAt 1.684 1.682 (-0.002) 1.682 1.689 ( 0.007) 1.696 1.705 ( 0.009)

F2 1.328 1.328 ( 0.000) 1.400 1.400 ( 0.000) 1.398 1.398 ( 0.000) 44 Cl2 1.976 1.976 ( 0.000) 1.984 1.985 ( 0.001) 2.012 2.013 ( 0.001) Br 2.272 2.273 ( 0.001) 2.281 2.281 ( 0.000) 2.313 2.313 ( 0.000)

I2 2.663 2.664 ( 0.001) 2.669 2.667 (-0.002) 2.702 2.700 (-0.002) At2 2.854 2.840 (-0.014) 2.860 2.847 (-0.013) 2.883 2.879 (-0.004) CuH 1.540 1.541 ( 0.001) 1.424 1.425 ( 0.001) 1.459 1.460 ( 0.001) AgH 1.700 1.700 ( 0.000) 1.584 1.584 ( 0.000) 1.625 1.626 ( 0.001) AuH 1.568 1.570 ( 0.002) 1.495 1.491 (-0.004) 1.538 1.536 (-0.002)

Cu2 2.401 2.401 ( 0.000) 2.203 2.203 ( 0.000) 2.248 2.247 (-0.001) Ag2 2.706 2.705 (-0.001) 2.505 2.506 ( 0.001) 2.595 2.595 ( 0.000) Au2 2.604 2.600 (-0.004) 2.468 2.462 (-0.006) 2.547 2.540 (-0.007) MAE 0.002 0.003 0.002 ME -0.001 -0.001 0.000 a) Difference in bond length from that obtained using the IODKH method. 3.3.4 Computational cost of the LUT scheme

This subsection investigates the efficiency of the analytical energy gradient in the

LUT scheme using multi-dimensional silver clusters Agn. Figure 3.2 shows the system-size dependence of the CPU time for the one-electron analytical energy gradient calculation in the one-electron IODKH and LUT-IODKH transforma- = tions of one-dimensional silver clusters Agn (n 1, 2, ..., 10). The vertical axis stands for the CPU time measured with the Xeon X5470/3.33 GHz (Quad core) on a single core in seconds while the horizontal axis shows the system size n.

The computational cost of the LUT-IODKH transformation is small with a small prefactor compared with the conventional IODKH transformation. The scalings

. . are O(n1 31) and O(n2 52), respectively, which are in reasonable agreement with theoretical ones, namely, O(n1) and O(n3). In addition, Figure 3.3 shows the sys- tem size dependence of the CPU time in the LUT-IODKH transformations of two- and three- dimensional silver clusters. In multi-dimensional systems, the LUT

. treatment scales quasi-linear: O(n1 21) for both two- and three-dimensional cases.

Thus, the LUT scheme for the analytical energy gradient method can efficiently evaluate the gradient for any chemical compound in the relativistic realm.

Table 3.7 shows the CPU time for the seven steps in the analytical energy gradient calculations at the HF level in an Ag10 cluster: OEI, UT of OEI, an initial guess obtained with the Huckel¨ calculation, TEI, SCF, dOEI, and dTEI. The NR,

IODKH, and LUT-IODKH Hamiltonian were used. The percentage of the CPU time for each step is given in parentheses. In energy calculations from OEI to

SCF, the time for the unitary transformation is considerably reduced in LUT-

IODKH compared with IODKH. The conventional IODKH requires 53.31% for the unitary transformation in the analytical energy gradient calculation. The LUT-

IODKH method, however, can reduce CPU time. Note that the total CPU time

45 60000

50000 IODKH LUT-IODKH 40000

2.52 30000 O(n ) CPU time [s] 20000

10000 O(n1.31) 0 0246810 n Figure 3.2 System-size dependence of CPU time (s) of the one-electron analytical energy = , , ··· , gradient in Agn (n 1 2 10) obtained with conventional IODKH and LUT-IODKH methods. A Xeon X5470/3.33GHz (Quad Core) processor was used on a single core.

10.0

9.0 two-dimension 8.0 three-dimension 7.0 6.0 O(n1.21) 5.0 4.0 O(n1.21) CPU time [s] 3.0 2.0 1.0 0 0 50 100 150 200 n Figure 3.3 System-size dependence of CPU time (s) of the one-electron analytical energy gradient in multi-dimensional silver clusters Agn obtained with LUT-IODKH methods. A Xeon X5470/3.33GHz (Quad Core) processor was used on a single core.

46 Table 3.7 CPU time (s) for seven steps in Hartree-Fock calculation obtained with NR,

IODKH, and LUT-IODKH in Ag10 linear cluster: OEI, UT of OEI, initial guess (Guess) calculated with Huckel¨ scheme, TEI, SCF, dOEI, and dTEI. A Quad core Xeon/3.3 GHz processor was used on a single core. NR IODKH LUT-IODKH Time % Time % Time % OEI 0.22 ( 0.01) 0.22 ( 0.00) 0.23 ( 0.01) UT of OEI 0.00 ( 0.00) 41902.32 ( 41.56) 12.54 ( 0.31) Guess 254.07 ( 6.52) 262.55 ( 0.26) 267.26 ( 6.50) TEI 886.03 ( 22.74) 928.95 ( 0.92) 886.19 ( 21.57) SCF 1242.38 ( 31.89) 2446.54 ( 2.43) 1328.72 ( 32.34) dOEI 2.31 ( 0.06) 53755.54 ( 53.31) 15.91 ( 0.39) dTEI 1510.97 ( 38.78) 1536.95 ( 1.52) 1598.19 ( 38.89) Total 3895.98 (100.00) 100833.07 (100.00) 4109.04 (100.00) for the LUT-IODKH method is close to those of the NR method, indicating that the LUT-IODKH scheme achieve both high accuracy close to the conventional

IODKH method as well as the 4c one and low computational cost comparable with NR method.

3.3.5 Metal complexes

This subsection provides the geometry parameters for 20 transition-metal com- plexes. Table 3.8 shows the complexes calculated by the NR, RECP, MCP, and

LUT-IODKH methods at the B3LYP level. A-B indicates the bond length between

A and B. The differences Δrrel of the RECP, MCP, and LUT-SF-IODKH methods from the NR method are given in brackets, representing the relativistic effect ac- counted by the RECP, MCP, and LUT-IODKH methods. Figure 3.4 shows the difference in the bond length from that obtained using the NR method. The bond lengths in the sixth-row metal complexes calculated by relativistic methods are shorter than that of the NR method. Especially in the twelfth group, Zn, Cd, and

Hg, the bond lengths are shorter than those in the other groups.

47 Table 3.8 Bond lengths (Å) for the 20 transition-metal complexes calculated by NR, RECP, MCP, and LUT-IODKH at the B3LYP level. Entry Bond NR RECP MCP LUT-IODKH

TiOCl2 Ti-O 1.598 1.593 1.601 1.595 Ti-Cl 2.239 2.232 2.237 2.236

TiF4 Ti-F 1.754 1.749 1.755 1.752 VO(OEt)3 V-O 1.577 1.576 1.587 1.576 V-OEt 1.768 1.763 1.768 1.764

Cr(CO)6 Cr-C 1.926 1.920 1.931 1.920 Co(CO)4H Co-H 1.472 1.475 1.478 1.468 Co-C(ax.) 1.822 1.802 1.804 1.811 Co-C(eq.) 1.809 1.816 1.824 1.798

Ni(PF3)4 Ni-P 2.131 2.128 2.119 2.115 Ni(CO)4 Ni-C 1.847 1.832 1.820 1.830 ZnMe2 Zn-C 1.958 1.945 1.933 1.942 ZrF4 Zr-F 1.906 1.906 1.907 1.905 ZrBr4 Zr-Br 2.495 2.495 2.494 2.488 NbCl5 Nb-Cl(ax.) 2.342 2.339 2.342 2.336 Nb-Cl(eq.) 2.234 2.289 2.298 2.288

MoBr4 Mo-Br 2.400 2.406 2.402 2.390 RuO4 Ru-O 1.693 1.682 1.687 1.681 CdEt2 Cd-C 2.201 2.153 2.171 2.158 TaCl5 Ta-Cl(ax.) 2.361 2.337 2.304 2.306 Ta-Cl(eq.) 2.312 2.288 2.264 2.260

WOCl4 W-O 1.721 1.696 1.689 1.700 W-Cl(ax.) 2.431 2.404 2.365 2.405 W-Cl(eq.) 2.299 2.266 2.241 2.264

Re(CO)5Br Re-C(ax.) 1.994 1.944 1.931 1.944 Re-C(eq.) 2.077 2.023 2.006 2.022 Re-Br 2.703 2.656 2.662 2.652

OsO4 Os-O 1.737 1.697 1.684 1.700 cis-Platin Pt-N 2.195 2.106 2.057 2.086 Pt-Cl 2.370 2.307 2.294 2.303

HgMe2 Hg-C 2.247 2.123 2.111 2.118

3.3.6 Heavier analogues of ethylene

This subsection provides the geometry parameters for heavier analogues of ethy- = lene H2MMH2 (M C, Si, Ge, Sn, and Pb). Figure 3.5 shows M-M bond length and H-M-H angle calculated by the NR and LUT-SF-IODKH methods at the HF level. M-M bond length is longer in both methods as the heavier atom is in- cluded. In the case of Pb, the differences in the bond length and angle between

48 0.10 RECP 0.05 MCP LUT-IODKH 0.00 [Å] rel r ¨ -0.05

-0.10 Forth period Fifth period Sixth period -0.15 Ti-F V-O Zr-F Ni-P Pt-N Ti-O W-O Ni-C Cr-C Zn-C Pt-Cl Ti-Cl Cd-C Os-O Ru-O Co-H Hg-C Zr-Br Re-Br V-OEt Mo-Br Re-C (ap.) Re-C (eq.) Co-C (ap.) Co-C (eq.) W-Cl (ap.) W-Cl (eq.) Ta-Cl (ap.) Ta-Cl (eq.) Nb-Cl (ap.) Nb-Cl (eq.) Bond Figure 3.4 Difference in bond length calculated by the RECP, MCP, and LUT-SF-IODKH from that obtained using the NR method.

the LUT-SF-IODKH method and the NR one are largest in the other cases since the hybridization of s and p orbitals in valence ones is difficult due to the rel- ativistic effect, which increases their orbital energy difference. In addition, the optimized geometry obtained by the LUT-IODKH method is close to PbH2 dimer rather than H2PbPbH2 molecule. Molecular orbitals for individual PbH2 dimer obtained by the LUT-IODKH method are not hybridized but the same as that of PbH2 monomer, while molecular orbitals obtained by the NR method are hy- bridized and contributed to M-M bonding.

3.3.7 Harmonic frequencies of diatomic molecules

This subsection provides the harmonic frequencies for heavy diatomic molecules = MH and M2 (M Cu, Ag, and Au). Table 3.9 shows the harmonic frequencies for heavy diatomic molecules by the NR and LUT-IODKH methods at the DFT with PBE0 functional. The difference between NR and LUT-IODKH, Δrel, is given in brackets indicates the relativistic effect, while the difference Δ of LUT-IODKH from experimental value is given in square brackets. The Hessian matrices were

49 3.5 120 (a) (b) 3.0 115 2.5 110 ]

Å 2.0 105 1.5 M-M [ H-M-H [°] 100 1.0 NR NR 0.5 95 LUT-IODKH LUT-IODKH 0.0 90 CSiGeSn Pb CSiGe Sn Pb Figure 3.5 Geometry parameters of (a) bond length (M-M) and (b) bond angle (H-M-H) for = heavier analogues of ethylene H2MMH2 (M C, Si, Ge, Sn, and Pb).

= Table 3.9 Harmonic frequencies of MH and M2 (M Cu, Ag, and Au) obtained by the NR and LUT-IODKH method at the DFT with PBE0 functional. Molecule NR Δrel LUT-IODKH Δ Exptl. CuH 1886.92 (−81.31) 1968.23 [26.97] 1941.26 AgH 1600.98 (−165.97) 1766.95 [7.05] 1759.90 AuH 1610.00 (−645.62) 2255.62 [−49.39] 2305.01 . − . . − . . Cu2 242 85 ( 20 12) 262 97 [ 1 58] 264 55 . − . . − . . Ag2 158 94 ( 21 32) 180 26 [ 12 14] 192 40 . − . . − . . Au2 118 02 ( 61 27) 179 29 [ 11 61] 190 90 obtained by numerical differentiations of analytical gradient values. The rela- tivistic effects on harmonic frequencies are more than 8% even at lighter diatomic molecule of Cu2;AtAu2, the contribution reaches to 38%. Resultant deviation in LUT-IODKH may be originated mainly from the limitation of electron correlation in DFT and partly from the SD relativistic and two-electron relativistic effects.

3.4 Conclusion

In this Chapter, the analytical energy gradient for the LUT-IODKH method is provided. Since the size of the relativistic transformation matrix is smaller than that of the conventional IODKH method and the definition is closure within in- dividual atoms, the derivative of the LUT-IODKH Hamiltonian can be obtained

50 efficiently. Numerical assessments were performed using several heavy-atom di- atomic molecules for accuracy, of which bond lengths were in good agreement with both the 4c method at the HF level and the experimental values at the MP2 and DFT with B3LYP levels. Efficiency for the derivative calculation of the LUT-

IODKH Hamiltonian is high compared to that of the conventional IODKH method using multi-dimensional silver clusters. This scheme achieved quasi-linear scal- ing in one-, two-, and three-dimensional clusters with small prefactors, which enables the efficient calculation in as much CPU time a geometry optimization as that of the NR method. In addition, 20 metal complexes and heavier analogues of ethylene were optimized as more practical system, which indicated that the relativistic effect plays an important role in bonding. As well as those bond- ing, in terms of the harmonic frequency, the relativistic effect is non-negligibly contributed even at the forth row element.

51 References

(1) Pyykko,¨ P. Chem. Rev. 1988, 88, 563.

(2) Dyall, K.; Fægri, K., Introduction to Relativistic Quantum Chemistry; Oxford

University Press: 2007.

(3) Reiher, M.; Wolf, A., Relativistic Quantum Chemistry; Wiley-VCH, Germany:

2009.

(4) Autschbach, J. J. Chem. Phys. 2012, 136, 150902.

(5) Visscher, L.; Dyall, K. G. J. Chem. Phys. 1996, 104, 9040.

(6) Abe, M.; Mori, S.; Nakajima, T.; Hirao, K. Chem. Phys. 2005, 311, 129.

(7) Salek, P.; Helgaker, T.; Saue, T. Chem. Phys. 2005, 311, 187.

(8) Komorovsky,´ S.; Repsisky,´ M.; Malkina, O. L.; Malkin, V. G.; Malkin, I. O.;

Kaupp, M. J. Chem. Phys. 2008, 128, 104101.

(9) Schwarz, W. H. E.; Wechsei-Trakowski, E. Chem. Phys. Lett. 1982, 85, 94.

(10) Nakajima, T.; Yanai, M.; Hirao, K. J. Comput. Chem. 2002, 23, 847.

(11) Liu, W.; Peng, D. J. Chem. Phys. 2006, 125, 044102.

(12) Kelly, M. S.; Shiozaki, T. J. Chem. Phys. 2013, 138, 204113.

(13) Samzow, R.; Hess, B. A.; Jansen, G. J. Chem. Phys. 1992, 96, 1227.

(14) Nakajima, T.; Hirao, K. J. Chem. Phys. 2003, 119, 4105.

(15) Van Wullen,¨ C.; Michauk, C. J. Chem. Phys. 2005, 123, 204113.

(16) Seino, J.; Hada, M. Chem. Phys. Lett. 2007, 442, 134.

(17) Seino, J.; Hada, M. Chem. Phys. Lett. 2008, 461, 327.

(18) Sikkema, J.; Visscher, L.; Saue, T.; llias,ˇ M. J. Chem. Phys. 2009, 131, 124116.

(19) Chang, C.; Pelissier, M.; Durand, P. Phys. Scr. 1986, 34, 394.

(20) Van Lenthe, E.; Baerends, E. J.; Snijders, J. G. J. Chem. Phys. 1993, 99, 4597.

(21) Van Lenthe, E.; Baerends, E. J.; Snijders, J. G. J. Chem. Phys. 1994, 101, 9783.

(22) Nakajima, T.; Hirao, K. Chem. Phys. Lett. 1999, 302, 383.

52 (23) Dyall, K. G. J. Chem. Phys. 1997, 106, 9618.

(24) Van Lenthe, E.; Ehlers, A.; Baerends, E.-J. J. Chem. Phys. 1999, 110, 8943.

(25) Van Lenthe, J. H.; Faas, S.; Snijders, J. G. Chem. Phys. Lett. 2000, 328, 107.

(26) Filatov, M.; Cremer, D. J. Chem. Phys. 2003, 118, 6741.

(27) Fedrov, D. G.; Nakajima, T.; Hirao, K. Chem. Phys. Lett. 2001, 335, 183.

(28) Filatov, M.; Cremer, D. Chem. Phys. Lett. 2003, 370, 647.

(29) Zou, W.; Filatov, M.; Cremer, D. J. Chem. Phys. 2011, 134, 244117.

(30) Kutzelnigg, W.; Liu, W. J. Chem. Phys. 2005, 123, 241102.

(31) Foldy, L. L.; Wouthuysen, S. A. Phys. Rev. 1950, 78, 29.

(32) Douglas, M.; Kroll, N. M. Ann. Phys. 1974, 82, 89.

(33) Nakajima, T.; Hirao, K. J. Chem. Phys. 2000, 113, 7786.

(34) Barysz, M. J. Chem. Phys. 2001, 114, 9315.

(35) De Jong, W. A; Harrison, R. J.; Dixon, D. A. J. Chem. Phys. 2001, 114, 48.

(36) Reiher, M.; Wolf, A. J. Chem. Phys. 2004, 121, 10945.

(37) Barysz, M.; Sadlej, A. J. J. Chem. Phys. 2002, 116, 2696.

(38) Nasluzov, V. A.; Rosch,¨ N. Chem. Phys 1996, 210, 413.

(39) Malkina, I.; Malkina, O. L.; Malkin, V. G. Chem. Phys. Lett. 2002, 361, 231.

(40) Matveev, A. V.; Nasluzov, V. A.; Rosch,¨ N. Int. J. Quantum Chem. 2007, 107,

3236.

(41) Neese, F.; Wolf, A.; Fleig, T.; Reiher, M.; Hess, B. A. J. Chem. Phys. 2005, 122,

204107.

(42) Mastalerz, R.; Barone, G.; Lindh, R.; Reiher, M. J. Chem. Phys. 2007, 127,

074105.

(43) Seino, J.; Nakai, H. J. Chem. Phys. 2012, 136, 244102.

(44) Seino, J.; Nakai, H. J. Chem. Phys. 2012, 137, 144101.

(45) Yang, W. Phys. Rev. Lett. 1991, 66, 1438.

(46) Yang, W.; Lee, T.-S. J. Chem. Phys. 1995, 103, 5674.

53 (47) Akama, T.; Kobayashi, M.; Nakai, H. J. Comput. Chem. 2007, 28, 2003.

(48) Kobayashi, M.; Akama, T.; Nakai, H. J. Chem. Phys. 2006, 125, 204106.

(49) Kobayashi, M.; Imamura, Y.; Nakai, H. J. Chem. Phys. 2007, 127, 074103.

(50) Kobayashi, M.; Nakai, H. J. Chem. Phys. 2008, 129, 044103.

(51) Kobayashi, M.; Nakai, H. Int. J. Quantum Chem. 2009, 109, 2227.

(52) Kobayashi, M.; Nakai, H. J. Chem. Phys. 2009, 131, 114108.

(53) Kobayashi, M.; Nakai, H. Phys. Chem. Chem. Phys. 2012, 14, 7629.

(54) Seino, J.; Nakai, H. J. Chem. Phys. 2013, 139, 034109.

(55) Visscher, L. Theor. Chem. Acc. 1997, 98, 68.

(56) Suh, I.-K.; Ohta, H.; Waseda, Y. J. Mater. Sci. 1988, 23, 757.

(57) Becke, A. D. J. Chem. Phys. 1993, 98, 5648.

(58) Sakai, Y.; Miyoshi, E.; Klobukowski, M.; Huzinaga, S. J. Chem. Phys. 1997,

106, 8084.

(59) Noro, T.; Sekiya, M.; Koga, T. Theor. Chem. Acc. 1997, 98, 25.

(60) Sekiya, M.; Noro, T.; Koga, T.; Matsuyam, H. J. Mol. Struct. (THEOCHEM))

1998, 451, 51.

(61) Miyoshi, E.; Sakai, Y.; Tanaka,K.; Masamura, M. J. Mol. Struct. (THEOCHEM)

1998, 451, 73.

(62) Noro, T.; Sekiya, M.; Koga, T.; Matsuyama, H. Theor. Chem. Acc. 2000, 104,

146.

(63) Sekiya, M.; Noro, T.; Osanai, Y.; Koga, T. Theor. Chem. Acc. 2001, 106, 297.

(64) Osanai, Y.; Sekiya, M.; Noro, T.; Koga, T. Mol. Phys. 2003, 101, 65.

(65) Noro, T.; Sekiya, M.; Osanai, Y.; Miyoshi, E.; Koga, T. J. Chem. Phys. 2003,

119, 5142.

(66) Osanai, Y.; Noro, T.; Miyoshi, E.; Sekiya, M.; Koga, T. J. Chem. Phys. 2004,

120, 6408.

54 (67) Osanai, Y.; Mon, M. S.; Noro, T.; Mori, H.; Nakashima, H.; Klobukowski, M.;

Miyoshi, E. Chem. Phys. Lett. 2008, 452, 210.

(68) Mori, H.; Ueno-Noto, K.; Osanai, Y.; Noro, T.; Fujiwara, T.; Klobukowski,

M.; Miyoshi, E. Chem. Phys. Lett. 2009, 476, 317.

(69) Dolg, M.; Wedig, U.; Stoll, H.; Preuss, H. J. Chem. Phys. 1987, 86, 866.

(70) Figgen, D.; Rauhut, G.; Dolg, M.; Stoll, H. Chem. Phys. 2005, 311, 227.

(71) Peterson, K. A.; Figgen, D.; Dolg, M.; Stoll, H. J. Chem. Phys. 2007, 126,

124101.

(72) Figgen, D.; Peterson, K. A.; Dolg, M.; Stoll, H. J. Chem. Phys. 2009, 130,

164108.

(73) Dunning, Jr., T. H. J. Chem. Phys. 1989, 90, 1007.

(74) Woon, D. E.; Thom. H. Dunning, J. J. Chem. Phys. 1993, 98, 1358.

(75) Schmidt, M. W.; Baldridge, K. K.; Boatz, J. A.; Elbert, S. T.; Gordon, M. S.;

Jensen, J. H.; Koseki, S.; Matsunaga, N.; Nguyen, K. A.; Su, S.; Windus, T. L.;

Dupuis, M.; Montgomery Jr., J. A. J. Comput. Chem. 1993, 14, 1347.

(76) DIRAC, a relativistic ab initio electronic structure program, Release

DIRAC12 (2012), written by H. J. Aa. Jensen, R. Bast, T. Saue, and

L. Visscher, with contributions from V. Bakken, K. G. Dyall, S. Dubillard,

U. Ekstrom,¨ E. Eliav, T. Enevoldsen, T. Fleig, O. Fossgaard, A. S. P. Gomes,

T. Helgaker, J. K. Lærdahl, Y. S. Lee, J. Henriksson, M. Ilias,ˇ Ch. R. Jacob,

S. Knecht, S. Komorovsky,´ O. Kullie, C. V. Larsen, H. S. Nataraj, P. Norman,

G. Olejniczak, J. Olsen, Y. C. Park, J. K. Pedersen, M. Pernpointner, K. Ruud,

P. Sałek, B. Schimmelpfennig, J. Sikkema, A. J. Thorvaldsen, J. Thyssen,

J. van Stralen, S. Villaume, O. Visser, T. Winther, and S. Yamamoto (see

http://www.diracprogram.org).

(77) Frisch, M. J. et al. Gaussian 09 Revision A.2., Gaussian Inc. Wallingford CT

2009.

55 (78) Noro, T.; Sekiya, M.; Koga, T. Theor. Chem. Acc. 2003, 109, 85.

(79) Tatewaki, H.; Koga, T. Chem. Phys. Lett. 2000, 328, 473.

(80) Koga, T.; Tatewaki, H.; Shimazaki, T. Chem. Phys. Lett. 2000, 328, 473.

(81) Han, Y.-K.; Bae, C.; Son, S.-K.; Lee, Y. S. J. Chem. Phys. 2000, 112, 2684.

(82) Seto, J. Y.; Morbi, Z.; Charron, F.; Lee, S. K.; Bernath, P. F.; Roy, R. J. L. J.

Chem. Phys. 1999, 110, 11756.

(83) Huber, K. P.; Herzberg, G., Chemistry Spectra and Molecular Structure IV.

Constants of Diatomic Molecules; Van Nostrand, New York: 1979.

(84) Ram, R. S.; Jarman, C. N.; Bernath, P. F. J. Mol. Spectrosc. 1992, 156, 468.

(85) Kramer,¨ H.-G.; Beutel, V.; Weyers, K.; Demtroder,¨ W. Chem. Phys. Lett. 1992,

193, 331.

56 Chapter 4 Implementation of spin-dependent relativistic analytical energy gradient

4.1 Introduction

A practical computational method that includes high-level relativistic effects has become one of the attracting goals for theoretical chemists. A fundamental treat- ment is based on a 4c Dirac theory, which adopts a one-particle Dirac operator.

For many electron systems, the 4c theory normally adopts the NR Coulomb in- teraction1 or a lower order of quantum electrodynamic effects such as the Breit2 or Gaunt3 interaction. Although the 4c relativistic calculations succeeded in ac- curately predicting molecular geometries4 and electronic/magnetic properties5,6 for small molecules, including at most 50 atoms, their application to more real- istic systems is difficult. Variational collapse arising from the existence of small- component Dirac spinors and/or positronic-state spinors has been a computational problem in the 4c calculations. The kinetic balance condition, which is a technique to resolve the variational problem, requires a large number of basis functions for the small-component spinors in comparison to those for the large-component spinors. To address the expensive computational cost in the 4c calculations, the basis spinor scheme has been proposed by Nakajima et al.7, which uses com- mon expansion coefficients in each spinor. Recently, Kelley and Shiozaki have improved the algorithm using density fitting technique combined with a highly

57 parallelized implementation for a large-scale calculation and have extended the algorithm to an analytical nuclear gradients.8,9

An alternative approach is a 2c relativistic theory, which treats only electronic- state (or large-component) information. Recent development in 2c treatment for many-electron systems has exhibited high accuracy, even when compared to the

4c theory. In particular, the Dirac-exact approaches, which give equivalent results to the 4c treatments, have been proposed for the one-electron Dirac Hamiltonian such as NESC10, X2C11, and IODKH12 methods, and the many-electron Dirac-

Coulomb(-Gaunt) ones13–18. Expensive computations for the decoupling between electronic and positronic states were also accelerated by introducing the approx- imate local transformations19–26. Among the approximations, LUT by Seino and co-workers and DLU by Reiher and co-workers can accurately and efficiently reproduce the results of the original 2c treatments since they include not only the diagonal relativistic components but also off-diagonal ones between adjacent atoms for the one-electron Dirac Hamiltonian. The LUT scheme, furthermore, was successfully extended to the many-electron Dirac-Coulomb Hamiltonian22,23.

Analytical energy gradient schemes for relativistic theories are essential for the geometry optimization of heavy-element systems. The relativistic effects are roughly categorized into SF and SD effects. The SF effect directly contributes to bond contractions/elongations due to the relativistic shrinking/expansion of molecular orbitals4,27–29. The SD effect, which indirectly contributes to molecular structures due to energy-level splittings of degenerate orbitals, is non-negligible in heavy-element chemistry. For example, the bond lengths are shortened/elongated

30 31 by approximately 0.1 Å in At2 and 0.2 Å in both (113)H and (115)H by includ- ing the SD relativistic effect. A contribution of a two-electron spin-orbit coupling is significant to describe molecular properties such as magnetic shielding con- stants32,33, excitation energies34–36, and spectroscopic constants37. The analytical

58 energy gradient schemes for the SF relativistic effect have been made available in

Dirac-exact 2c approaches such as NESC38–40, X2C41,42, and IODKH43, as well as in the approximate 2c approaches such as RA schemes44–46, RESC method47, and the

finite-order DKH methods48–54. Note that these three Dirac-exact 2c approaches give identical results although they need different procedures to obtain energies and properties. Recently, the analytical energy gradient for the 2c NESC method with a screened nucleus potential using effective nuclear charges has been pro- posed by Zou et al.55 The scheme has shown that a contribution of a spin-orbit coupling to bond lengths is small in general, but relatively large for van der Waals complexes. The progress of the analytical energy derivative in the relativistic quantum chemistry has been well described in recent review papers40,56,57.

From the computational point of view, the SF effect is more easily taken into account compared to the SD effect because modifications from the NR quantum chemical program are limited. In fact, the SF relativistic theory for the one-electron

Dirac Hamiltonian requires only the implementation related to the one-electron integrals. On the contrary, the implementation of the SD relativistic theory needs a considerable change of the NR program. For example, many variables such as orbital coefficients, one- and two-electron integrals for the SD operators, and Fock matrices, should be treated as complex numbers. In consequence, the quantum chemical programs that can deal with the analytical energy gradient of the SD relativistic theories are very limited.

In this Chapter, the author implemented a computational program for the analytical energy gradient, as well as the energy at the SD relativistic level. The previous Chapter derived the formulation of the analytical energy gradient for the IODKH method with/without the LUT treatment, but the author implemented only the SF computation. The local approximation schemes are effective to de- scribe chemical bonds in both light- and heavy-element systems, especially for

59 comparatively strong bonds such as covalent bond, metallic bond, and ionic bond because the schemes reproduce the total energy obtained by corresponding con- ventional methods will less than 1 millihartree deviations. Note that an extension of the approximate local transformations to other Dirac-exact methods are possi- ble, whose derivation and implementation might be easy. The IODKH method denoted here is equivalent to IOTC12 method or BSS58,59 method. The present chapter is organized as follows. Implementation is explained in Section 4.2. Nu- merical assessments in terms of both accuracy and efficiency are presented in

Section 4.3. Finally, concluding remarks are given in Section 4.4.

4.2 Implementation

The theoretical aspects of (LUT-)IODKH is the same as Section 3.2. This section presents the implementation of evaluating the analytical energy gradient of the two-component (LUT-)IODKH method. To clarify the difference in implementa- tion among NR, SF, and SD relativistic methods, we show the evaluation processes for all the methods in detail. Table 4.1 summarizes the procedures of energy and energy gradient calculations and data types. In the energy calculations, the pro- cedures are divided into three parts: one-electron integral (including TNR, VNR,

σ · pVσ · p, matrix transformation, and relativistic transformation), two-electron integral, and SCF (including construction and diagonalization of Fock matrix). In the gradient calculations, the procedures are also divided into three part: one- electron integral, two-electron integral, and the energy derivative defined in Eq.

(2.84).

In detail for the NR and SF calculations, all integrals, their derivatives without relativistic integrals, and transformations are evaluated in a real form. For the SF-

IODKH calculation, steps for the spin-free parts of the σ ·pVσ ·p integrals, matrix transformation, relativistic transformation, and these derivatives are added to

60 the procedure in NR. On the other hand, the LUT-IODKH method simplifies the calculation using a relativistic transformation determined in each atom in the energy calculation.

For the SD calculations, the spin-dependent parts of the σ · pVσ · p integrals in a complex form are added to the SF procedure. Furthermore, the atomic orbital size increases to twice its original size in order to describe the α- and β-spin mixing.

Therefore, the relativistic transformation, SCF procedure, and these derivatives require complex calculations with twice the number of basis sets.

4.3 Numerical assessments 4.3.1 Computational details

This section examines the performance of the analytical energy gradients for the

2c IODKH and LUT-IODKH methods in terms of accuracy and efficiency. As the

2c relativistic Hamiltonian, the one-electron SD-IODKH Hamiltonian and two- electron Coulomb interaction, denoted by SD-IODKH, were adopted. In the LUT scheme, the threshold for cutoff of relativistic interaction, τ, was set to 3.5 Å. For comparison, the 4c Dirac-Coulomb, (LUT-) SF-IODKH, and NR methods were also employed.

Twelve closed-shell diatomic molecules containing fifth- and sixth-period atoms were calculated as numerical tests: InH, SnO, Sb2, HI, I2, AuH, Au2, TlH,

PbO, Bi2, HAt, and At2. Five van der Waals complexes, Ne2,Ar2,Kr2,Xe2, and ff Hg2, were used in order to test cuto radius. Furthermore, multi-dimensional silver clusters were employed to investigate the efficiency in LUT-SD-IODKH. = , ··· , The Ag-Ag bond length in one-dimensional (1D) Agn clusters (n 2 10) / was fixed at 2.706 Å. In two- and three-dimensional (2D 3D) Agn clusters, an experiment lattice constant of 4.806 Å with a face-centered cubic structure was

60 adopted . The total numbers of atoms in the 2D Agn clusters were 5, 8, 11, 13,

61 Table 4.1 Procedures of energy and energy gradient calculations and data types in NR, (LUT-)SF-IODKH, and (LUT-)SD-IODKH. SF SD NR IODKH LUT-IODKH IODKH LUT-IODKH Energy One-electron integral TNR real real real real real VNR real real real real real σ · pVσ · p – real real complex complex Matrix transformation X – real real (atom) real real (atom) Relativistic transformation – real real complex complex 62 / Two-electron integral 1 rij real real real real real SCF Fock real real real complex complex Diagonalization real real real complex complex Energy gradient Derivative of one-electron integral TNR real real real real real VNR real real real real real σ · pVσ · p – real real complex complex Matrix transformation X – real – real – Relativistic transformation – real real complex complex / Derivative of two-electron integral 1 rij real real real real real Summation for total gradient real real real complex complex 14, 17, 18, 23, 25, 28, 32, 39, 41, and 50, which correspond to 1 × 1, 1 × 2, 1 × 3,

2 × 2, 1 × 4, 1 × 5, 2 × 3, 2 × 4, 3 × 3, 2 × 5, 3 × 4, 3 × 5, 4 × 4, and 4 × 5 unit cells, respectively. The total numbers in the 3D Agn clusters were 14, 23, 32, 41, and 53, which correspond to 1×1×1, 1×1×2, 1×3×1, 1×1×4, and 1×2×3, respectively. − For a practical calculation, fac Ir(ppy)3 was also examined. All the NR, SF, and SD calculations were performed at the RHF, RHF, and GHF levels, respectively. For the diatomic molecules and Agn clusters, the Sapporo- DZP-201261 and DKH3-Gen-TK/NOSeC-V-TZP62–64 basis sets with uncontracted forms were employed in the H atom and other atoms, respectively. The basis sets used for van der Waals complexes were DKH3-Gen-TK/NOSeC-V-TZP in an uncontracted form for noble gas atoms (Ne2,Ar2,Kr2, and Xe2). For Hg2, two basis sets were adopted; DKH3-Gen-TK/NOSeC-V-TZP in an uncontracted form with a large exponent (> 1.2 × 109) and SARC-DKH in an uncontracted form with

< . × 6 − a moderate exponent ( 1 8 10 ). In fac Ir(ppy)3, the contracted DKH3-Gen- TK/NOSeC-V-TZP for Ir and cc-pVDZ65 for H, C, and N atoms were adopted.

− 66 Note that in the NR calculation for fac Ir(ppy)3, the basis set for Ir, TZP , was adopted. The analytical energy gradient for the (LUT-)SD-IODKH method was implemented in the GAMESS program67. The DIRAC12 program was utilized for the 4c calculations68.

4.3.2 Accuracy of SD-IODKH and LUT-SD-IODKH

This section examines the accuracy of the analytical energy gradient method for the 2c IODKH method with/without the LUT scheme. Table 4.2 shows the analytical and numerical gradient values (in hartree/bohr) at equilibrium lengths of re obtained by NR, SD-IODKH, LUT-SD-IODKH, and 4c at the HF level. Here,

= 31 the experimental bond lengths were used for re, namely, re 1.742 Å for HAt ,

69 70 71 3.046 Å for At2 , 1.5324 Å for AuH , and 2.4719 Å for Au2 . Furthermore, 1.5re

63 and 2.0re for AuH and HAt were also calculated. The numerical gradient values were obtained using a five point numerical differentiation method at the 0.001

Å step size. The inherent accuracy of this numerical differentiation is a fourth-

− order accuracy. The convergence threshold of the SCF calculation is 5.0 × 10 6 for the density changes. A quadruple precision routines for evaluating one-electron integrals including overlap and kinetic energy, and for diagonalizing the integral matrices were used to keep accuracy in the Hess’ numerical technique72. The deviation of the analytical gradient with respect to the numerical gradient, Δgana, shows the reliability of the present implementation.

The maximum absolute deviations of NR, SD-IODKH, LUT-SD-IODKH, and

4c are 0.000002, 0.000013, 0.000008, and 0.000002, respectively. The deviations of

SD-IODKH and LUT-SD-IODKH are slightly larger than that of NR. This is due to the resolution-of-identity (RI) approximation, which is essential for Hess’ numer- ical calculation72 in the IODKH method. Considering these theoretical demands, the results indicate the correct implementation of the present analytical energy gradients in SD-IODKH and LUT-SD-IODKH within the Hess approximation.

Table 4.3 presents the bond lengths (Å) of diatomic molecules containing

fifth- and sixth-period elements calculated by the NR, SF-IODK/C, SD-IODKH,

LUT-SD-IODKH, and 4c methods. The deviations of the NR, SF-IODKH, and

SD-IODKH methods from 4c, Δr4c, indicates the fully relativistic effect, the SD relativistic effect in one-electron terms, and the fully relativistic effect in two- electron terms, respectively. The deviation of SD-IODKH from SF-IODKH, ΔrSD, is shown in square brackets, representing the SD relativistic effect in one-electron terms. The deviation of LUT-SD-IODKH from SD-IODKH, ΔrLUT, indicates the accuracy of the LUT scheme.

The fully relativistic effect on bond lengths, Δr4c of NR, is mostly positive whereas I2,Bi2, and At2 are opposite. The absolute values are large in AuH and

64 / / Table 4.2 Analytical and or numerical gradient values (hartree bohr) of HAt, At2, AuH, and Au2 calculated by NR, SD-IODKH, LUT-SD-IODKH, and 4c at the HF level. NR SD-IODKH LUT-SD-IODKH 4c Molecules Length g ΔggΔggΔggΔg

AuH 1.0re 0.082251 0.000001 0.012531 0.000013 0.012518 0.000007 0.011711 0.000000 65 1.5re 0.033559 0.000000 0.054765 0.000007 0.054768 0.000001 0.054581 0.000000 2.0re 0.031298 0.000000 0.036769 0.000000 0.036770 0.000002 0.036850 0.000000 HAt 1.0re 0.011267 0.000002 0.008891 0.000009 0.008908 0.000001 0.009392 0.000000 − 1.5re 0.067515 0.000001 0.054690 0.000005 0.054686 0.000000 0.055145 0.000001 2.0re 0.037875 0.000000 0.000126 0.000000 0.000126 0.000008 0.031849 0.000000 Au2 1.0re 0.079816 0.000002 0.028908 0.000006 0.028880 0.000004 0.029315 0.000000 − − At2 1.0re 0.025557 -0.000001 0.007775 0.000002 0.007796 0.000000 0.008351 0.000002 Au2: 0.263 Å for AuH and 0.330 Å for Au2. The large deviations are reduced by the SF relativistic effect. Actually, Δr4c of SF-IODKH are 0.007 Å for AuH and

0.008 Å for Au2. This indicates that AuH and Au2 have an s-bonding nature formed by each s valence orbital, which is largely contracted by the relativistic effect. However, the absolute values |Δr4c| of SF-IODKH exhibit large differences, − − especially in Bi2 and At2: 0.095 Å for Bi2 and 0.129 Å for At2. On the other hand, |Δr4c| of SD-IODKH are small in all molecules. This indi- cates that the SD relativistic effect in heavy-element systems is important when the bond has non-zero angular momentum such as p, d, and f orbitals. The con- tributions of two-electron relativistic effects can be small in bond lengths of the systems used here, although the contributions sometimes can be important.55 Fur- ff thermore, TlH has a negative value by the SD relativistic e ect, even though Bi2, ff ff HAt, and At2 have positive values. The SD relativistic e ect gives two di erent contributions to bond length: a splitting of degenerate orbitals and an overlap of orbitals by a second-order spin-orbit interaction. The former is when p, d, { } { } { } and f orbitals are split into p1/2,p3/2 , d3/2,d5/2 , and f5/2,f7/2 , respectively.

The orbitals with less (more) total angular momentum are stabilized and con- tracted (destabilized and expanded). The latter contributes to the stabilization of a molecule by an overlap of orbitals with the same total angular momentum σ π π σ and spatial symmetry, i.e., 1/2g and 1/2g or 1/2u and 1/2u, which are split by a spin-orbit interaction. In the case of TlH, the former contribution is dominant σ because two electrons occupy the 1/2g orbital, which is stabilized by the SD rela- ff tivistic e ect. In the case of Bi2, HAt, and At2, the latter contribution is dominant because electrons occupy bonding and/or anti-bonding orbitals. In this situation, the SD relativistic effect stabilizes a molecule largely when the bond length is long, although the opposite is true for the SF relativistic effect. As a result, the bonds are elongated by the SD relativistic effect.

66 In all cases, ΔrLUT is considerably small. This represents that the LUT treat- ment can sufficiently describe both SF and SD relativistic effects for estimating the bond lengths.

Note that in systems whose bond lengths are longer such as van der Waals complexes, the cutoff radius including nearest neighbor atoms is set carefully in order to keep accuracy. Table 4.4 shows the bond lengths (Å) of van der Waals complexes obtained by the LUT-IODKH and IODKH methods at the MP2 level.

Two cutoff radii set to 3.5 and 5.0 Å. In noble gas dimers, the deviations of LUT-

IODKH from IODKH with 5.0 Å of cutoff radius were within 0.001 Å. On the / other hand, in Hg2, the deviations were 0.013 for DKH3-Gen-TK NOSeC-V-TZP and 0.006 for SARC-DKH. The large deviations are related with the large exponent values because the deviations increase when the exponent value increase. One reason of the deviations can be due to the dropping terms in the LUT scheme.

Another reason is due to the numerical error by large values in the integrals when we adopted a basis sets with an extremely large exponent value. These results indicate that van der Waals radius should be included at least in cutoff radius for keeping accuracy when we use the LUT scheme in a geometry optimization.

4.3.3 Computational cost of (LUT-)SD-IODKH method

This subsection examines the computational cost of the analytical energy gradient for the SD-IODKH method with/without the LUT scheme on 1D, 2D, and 3D silver clusters. Figure 4.1 shows the system-size dependence of the central processing unit (CPU) time for the analytical gradient of the one-electron integrals in 1D, 2D, and 3D silver clusters obtained by the SD-IODKH and LUT-SD-IODKH methods.

The horizontal axis indicates the total numbers of the atoms n. The vertical axis indicates the CPU time in seconds. A Xeon E5-2690/2.90 GHz (Octa core) processor on a single core was utilized to measure the CPU time. In 1D cluster, the CPU time

67 Table 4.3 Bond lengths (Å) of diatomic molecules containing fifth- and sixth-period elements calculated by the NR, SF-IODKH, SD-IODKH, LUT-SD-IODKH, and 4c methods at the HF level. IODKH NR 4c Molecule SF SD LUT-SD rNR Δr4c rSF Δr4c rSD Δr4c ΔrSD rLUT ΔrLUT r4c InH 1.856 0.012 1.845 0.001 1.843 −0.001 −0.002 1.843 0.000 1.844 SnO 1.798 0.008 1.789 −0.001 1.790 0.000 0.001 1.790 0.000 1.790 68 − Sb2 2.460 0.018 2.436 0.006 2.443 0.001 0.007 2.443 0.000 2.442 HI 1.608 0.005 1.600 −0.003 1.603 0.000 0.003 1.603 0.000 1.603 − − I2 2.675 0.003 2.664 0.014 2.679 0.001 0.015 2.679 0.000 2.678 AuH 1.828 0.263 1.572 0.007 1.569 0.004 −0.003 1.568 −0.001 1.565 − − Au2 2.924 0.330 2.602 0.008 2.593 0.001 0.009 2.593 0.000 2.594 TlH 1.932 0.069 1.890 0.027 1.859 −0.004 −0.031 1.859 0.000 1.863 PbO 1.890 0.019 1.864 −0.007 1.873 0.002 0.009 1.873 0.000 1.871 − − Bi2 2.647 0.015 2.567 0.095 2.667 0.005 0.100 2.667 0.000 2.662 HAt 1.711 0.001 1.683 −0.027 1.712 0.002 0.029 1.712 0.000 1.710 − − At2 2.890 0.081 2.842 0.129 2.976 0.005 0.134 2.978 0.002 2.971 Table 4.4 Bond lengths (Å) of van der Waals complexes obtained by the LUT-IODKH and IODKH methods at the MP2 level. LUT-IODKH Moleucle τ = 3.5 τ = 5.0 IODKH Δ

Ne2 3.278 3.278 3.278 0.000 Ar2 3.913 3.913 3.913 0.000 Kr2 4.170 4.087 4.087 0.000 Xe2 4.613 4.559 4.558 0.001 Hg2 3.495 3.495 3.501 0.006 Hg2 3.509 3.551 3.538 0.013 a) Results of the DKH3-Gen-TK/NOSeC-V-TZP basis set in an uncontracted form. b) Results of the SARC-DKH basis set in an uncontracted form.

. in SD-IODKH is long; the scaling is O(n2 98), which corresponds to an order in a matrix multiplication. On the other hand, the LUT scheme can drastically reduce

. the CPU time; the scaling is quasi-linear, O(n1 33). The results also represent that the LUT-SD-IODKH method achieves quasi-linear scaling for 2D and 3D clusters:

. . O(n1 35) for 2D and O(n1 36) for 3D.

Table 4.5 shows the CPU time for an analytical energy gradient calculation of Ag10. The NR, SF-IODKH, LUT-SF-IODKH, SD-IODKH, and LUT-SD-IODKH methods were adopted. The calculation is divided into six steps: namely, one- electron integrals including the 2c transformation (OEI), an initial guess (Guess) by Huckel¨ scheme, TEI, the SCF procedure, derivatives of OEI including the 2c transformation (dOEI), and dTEI. The percentage of CPU time in each step is given in parentheses.

In single-point energy calculations including OEI, Guess, TEI, and SCF, SF- and SD-IODKH for OEI require long CPU times in comparison with NR: 0.12 s for NR, 75.72 s for SF-IODKH, and 1487.83 s for SD-IODKH. The long CPU times for OEI are reduced by the LUT schemes: 4.47 s for LUT-SF-IODKH and 24.69 s for LUT-SD-IODKH. In the SCF step, the CPU time at the SD level is ≥10 times as much as that at the SF level. This is because the GHF method for the SD effect

69 7000 70 (a) (b) 6000 60 SD-IODKH 2d 5000 LUT-SD-IODKH 50 3d 4000 40 O(n1.36) 3000 O(n2.98) 30 CPU time [s] CPU time [s] 2000 20 O(n1.35) 1000 10 O(n1.33) 0 0 020246810 0 40 60 n n Figure 4.1 System-size dependence of CPU time (s) of the one-electron analytical energy gradient obtained with the SD-IODKH and LUT-SD-IODKH methods: (a) 1D Agn cluster / and (b) 2D and 3D Agn cluster. A Xeon E5-2690 2.90 GHz (Octa Core) processor was used on a single core.

requires a calculation with a complex form and twice as large basis sets as the

RHF one.

In gradient calculations, including dOEI and dTEI, SF- and SD-IODKH for dOEI demand large CPU time in comparison to NR: 1.11 s for NR, 2652.28 s for SF-

IODKH, and 61247.55 s for SD-IODKH. The LUT scheme reduces the CPU times:

129.36 s for LUT-SF-IODKH and 263.35 s for LUT-SD-IODKH. Furthermore, in dTEI, the CPU times at the SD level are longer than those at the SF one. This is because additional CPU times are required to multiply the density matrix with the

GHF formalism to the derivative of TEI, although the integral calculations are the same. These results indicate that the LUT scheme resolves one of the bottlenecks, i.e., long CPU times in the OEI and dOEI steps.

− 4.3.4 Application in fac Ir(ppy3)

This subsection discusses the performance of the LUT-SF-IODKH and LUT-SD-

IODKH methods in fac-Ir(ppy)3. It should be noted that this molecule cannot be treated without the LUT technique for the relativistic calculation. Table 4.6 shows

70 Table 4.5 CPU time for six steps in the restricted or general Hartree-Fock calculation obtained with NR, SF-IODKH, LUT-SF-IODKH, SD-

IODKH, and LUT-SD-IODKH in Ag10 linear cluster: OEI, initial guess (Guess) calculated with the Huckel¨ scheme, TEI, SCF, dOEI, and dTEI. A Xeon E5-2690/2.90 GHz processor (Octa core) was used on a single core. NR SF-IODKH LUT-SF-IODKH SD-IODKH LUT-SD-IODKH Time (s) % Time (s) % Time (s) % Time (s) % Time (s) % 71 OEI 0.12 (0.00) 75.72 (1.35) 4.47 (0.15) 1487.83 (1.95) 24.69 (0.18) Guess 20.55 (0.75) 20.63 (0.37) 19.52 (0.67) 20.45 (0.03) 20.39 (0.15) TEI 675.33 (24.66) 725.13 (12.91) 687.98 (23.65) 690.26 (0.91) 693.53 (5.04) SCF 817.20 (29.84) 904.83 (16.11) 814.73 (28.01) 10507.45 (13.79) 10687.16 (77.74) dOEI 1.11 (0.04) 2652.28 (47.21) 129.36 (4.45) 61427.55 (80.64) 263.35 (1.92) dTEI 1224.57 (44.71) 1238.95 (22.06) 1253.02 (43.07) 2046.20 (2.69) 2058.81 (14.98) Total 2738.88 (100.00) 5617.54 (100.00) 2909.08 (100.00) 76179.74 (100.00) 13747.93 (100.00) Table 4.6 Geometric parameters of Ir(ppy)3 obtained by NR, LUT-SF-IODKH and LUT-SD- IODKH at the Hartree-Fock level. Parameters NR SF SD ΔSF ΔSD ΔRel Ir-N 2.225 2.198 2.149 −0.027 −0.049 −0.076 Ir-C 2.031 2.051 1.964 0.020 −0.087 −0.067 C-Ir-N 172.2 175.4 173.23.2 −2.21.0 C-Ir-C 95.396.992.51.6 −4.4 −2.8 N-Ir-N 97.796.596.5 −1.20.0 −1.2 the geometric parameters calculated at the Hartree-Fock level of theory. A-B and

A-B-C denote bond lengths between atoms A and B and angles among atoms A, B, and C, respectively. ΔSF is the difference of the geometrical parameters between

NR and LUT-SF-IODKH, indicating the SF relativistic effect on the geometry.

ΔSD is the difference of the geometrical parameters between LUT-SF-IODKH and

LUT-SD-IODKH, indicating the SD relativistic effect. ΔRel is the summation of

ΔSF and ΔSD, indicating that total relativistic effect. Note that we used the nearest neighboring atom from the central metal Ir to estimate the bond lengths and angles. ff In the bond lengths of fac-Ir(ppy)3, the SF relativistic e ect is smaller than the SD relativistic effect. For Ir-N and Ir-C, {ΔSF, ΔSD} are {−0.027, −0.049} and {0.020,

−0.087} in Å, respectively. On the other hand, the three bond angles, C-Ir-N,

C-Ir-C, and N-Ir-N, are slightly affected by both SF and SD effects. For C-Ir-N,

C-Ir-C, and N-Ir-N, {ΔSF, ΔSD} are {3.2, −2.2}, {1.6, −4.4}, and {−1.2, 0.0} in degrees, respectively.

4.4 Conclusion

The present study develops the analytical energy gradient for the GHF method based on the 2c IODKH method with/without the efficient LUT scheme in a 2c relativistic transformation. The accuracy of the present method was assessed using several diatomic molecules containing fifth- and sixth-period elements.

72 The optimized bond lengths are in good agreement with those of the 4c method.

Furthermore, the deviations of the LUT scheme are less than 0.003 Å in bond length. The efficiency of the LUT scheme was also examined using 1D, 2D, and

3D silver clusters. The CPU time for a one-electron relativistic transformation scales quasi-linear in these clusters. A practical calculation using a fac-Ir(ppy)3 molecule was performed to estimate the accuracy and efficiency of the proposed method. The results show that the SD relativistic effect shortens the bond length by approximately 0.1 Å. In addition, although the SD relativistic program is much more complicated than that of NR, the LUT-SD-IODKH scheme demands only about twice as much CPU time as NR. The present analytical energy gradient method for LUT-IODKH with the SD relativistic effect was found to be applicable to practical systems.

73 References

(1) Moss, R. E., Advanced Molecular Quantum Mechanics; Chapman and Hall,

London: 1973.

(2) Breit, G. Phys. Rev. 1929, 34, 553.

(3) Gaunt, J. A. Proc. R. Soc. Lond. A 1929, 124, 163.

(4) Pyykko,¨ P. Chem. Rev. 1988, 88, 563.

(5) Salek, P.; Helgaker, T.; Saue, T. Chem. Phys. 2005, 311, 187.

(6) Komorovsky,´ S.; Repsisky,´ M.; Malkina, O. L.; Malkin, V. G.; Malkin, I. O.;

Kaupp, M. J. Chem. Phys. 2008, 128, 104101.

(7) Nakajima, T.; Yanai, M.; Hirao, K. J. Comput. Chem. 2002, 23, 847.

(8) Kelly, M. S.; Shiozaki, T. J. Chem. Phys. 2013, 138, 204113.

(9) Shiozaki, T. J. Chem. Theory Comput. 2013, 9, 4300.

(10) Dyall, K. G. J. Chem. Phys. 1997, 106, 9618.

(11) Liu, W.; Peng, D. J. Chem. Phys. 2006, 125, 044102.

(12) Barysz, M.; Sadlej, A. J. J. Chem. Phys. 2002, 116, 2696.

(13) Samzow, R.; Hess, B. A.; Jansen, G. J. Chem. Phys. 1992, 96, 1227.

(14) Nakajima, T.; Hirao, K. J. Chem. Phys. 2003, 119, 4105.

(15) Van Wullen,¨ C.; Michauk, C. J. Chem. Phys. 2005, 123, 204113.

(16) Seino, J.; Hada, M. Chem. Phys. Lett. 2007, 442, 134.

(17) Seino, J.; Hada, M. Chem. Phys. Lett. 2008, 461, 327.

(18) Sikkema, J.; Visscher, L.; Saue, T.; llias,ˇ M. J. Chem. Phys. 2009, 131, 124116.

(19) Wahlgren, U.; Schimmelpfennig, B.; Jusuf, S.; Stromsnes,¨ H.; Gropen, O.;

Maron, L. Chem. Phys. Lett. 1998, 287, 525.

(20) Thar, J.; Kirchner, B. J. Chem. Phys. 2009, 130, 124103.

(21) Seino, J.; Nakai, H. J. Chem. Phys. 2012, 136, 244102.

(22) Seino, J.; Nakai, H. J. Chem. Phys. 2012, 137, 144101.

74 (23) Seino, J.; Nakai, H. J. Chem. Phys. 2013, 139, 034109.

(24) Seino, J.; Nakai, H. Int. J. Quantum Chem. 2015, 115, 253.

(25) Seino, J.; Nakai, H. J. Comput. Chem. Jpn. 2014, 13,1.

(26) Peng, D.; Reiher, M. J. Chem. Phys. 2012, 136, 244108.

(27) Dyall, K.; Fægri, K., Introduction to Relativistic Quantum Chemistry; Oxford

University Press: 2007.

(28) Reiher, M.; Wolf, A., Relativistic Quantum Chemistry; Wiley-VCH, Germany:

2009.

(29) Autschbach, J. J. Chem. Phys. 2012, 136, 150902.

(30) Wang, Z.; Wang, F. Phys. Chem. Chem. Phys. 2013, 15, 17922.

(31) Han, Y.-K.; Bae, C.; Son, S.-K.; Lee, Y. S. J. Chem. Phys. 2000, 112, 2684.

(32) Vaara, J.; Ruud, K.; Vahtras, O.; Ågren, H.; Jokisaari, J. J. Chem. Phys. 1998,

109, 1212.

(33) Malkina, O.; Schimmelpfennig, B.; Kaupp, M.; Hess, B.; Chandra, P.;

Wahlgren, U.; Malkin, V. Chem. Phys. Lett. 1998, 296, 93.

(34) Launila, O.; Schimmelpfennig, B.; Fagerli, H.; Gropen, O.; Taklif, A. G.;

Wahlgren, U. J. Mol. Spectrosc. 1997, 186, 131.

(35) Fagerli, H.; Schimmelpfennig, B.; Gropen, O.; Wahlgren, U. J. Mol. Struct.

(THEOCHEM) 1998, 451, 227.

(36) Rakowitz, F.; Marian, C. M.; Schimmelpfennig, B. Phys. Chem. Chem. Phys.

2000, 2, 2481.

(37) Malmqvist, P. Å.; Roos, B. O.; Schimmelpfennig, B. Chem. Phys. Lett. 2002,

357, 230.

(38) Filatov, M.; Cremer, D. Chem. Phys. Lett. 2003, 370, 647.

(39) Zou, W.; Filatov, M.; Cremer, D. J. Chem. Phys. 2011, 134, 244117.

(40) Filatov, M.; Zou, W.; Cremer, D. Int. J. Quantum Chem. 2014, 114, 993.

(41) Cheng, L.; Gauss, J. J. Chem. Phys. 2011, 135, 084114.

75 (42) Cheng, L.; Gauss, J.; Stanton, J. F. J. Chem. Phys. 2013, 139, 054105.

(43) Nakajima, Y.; Seino, J.; Nakai, H. J. Chem. Phys. 2013, 139, 244107.

(44) Van Lenthe, E.; Ehlers, A.; Baerends, E.-J. J. Chem. Phys. 1999, 110, 8943.

(45) Van Lenthe, J. H.; Faas, S.; Snijders, J. G. Chem. Phys. Lett. 2000, 328, 107.

(46) Filatov, M.; Cremer, D. J. Chem. Phys. 2003, 118, 6741.

(47) Fedrov, D. G.; Nakajima, T.; Hirao, K. Chem. Phys. Lett. 2001, 335, 183.

(48) Nasluzov, V. A.; Rosch,¨ N. Chem. Phys 1996, 210, 413.

(49) De Jong, W. A; Harrison, R. J.; Dixon, D. A. J. Chem. Phys. 2001, 114, 48.

(50) Malkina, I.; Malkina, O. L.; Malkin, V. G. Chem. Phys. Lett. 2002, 361, 231.

(51) Neese, F.; Wolf, A.; Fleig, T.; Reiher, M.; Hess, B. A. J. Chem. Phys. 2005, 122,

204107.

(52) Matveev, A. V.; Nasluzov, V. A.; Rosch,¨ N. Int. J. Quantum Chem. 2007, 107,

3236.

(53) Mastalerz, R.; Barone, G.; Lindh, R.; Reiher, M. J. Chem. Phys. 2007, 127,

074105.

(54) Autschbach, J.; Peng, D.; Reiher, M. J. Chem. Theory Comput. 2012, 8, 4239.

(55) Zou, W.; Filatov, M.; Cremer, D. J. Chem. Phys. 2015, 142, 214106.

(56) Cheng, L.; Stopkowicz, S.; Gauss, J. Int. J. Quantum Chem. 2014, 114, 1108.

(57) Cremer, D.; Zou, W.; Filatov, M. WIREs Comput. Mol. Sci. 2014, 4, 436.

(58) Barysz, M.; Sadlej, A. J.; Snijders, J. G. Int. J. Quantum Chem. 1997, 65, 225.

(59) Barysz, M. J. Chem. Phys. 2001, 114, 9315.

(60) Suh, I.-K.; Ohta, H.; Waseda, Y. J. Mater. Sci. 1988, 23, 757.

(61) Noro, T.; Sekiya, M.; Koga, T. Theor. Chem. Acc. 2003, 109, 85.

(62) Sekiya, M.; Noro, T.; Osanai, Y.; Koga, T. Theor. Chem. Acc. 2001, 106, 297.

(63) Osanai, Y.; Noro, T.; Miyoshi, E.; Sekiya, M.; Koga, T. J. Chem. Phys. 2004,

120, 6408.

(64) Tatewaki, H.; Koga, T. Chem. Phys. Lett. 2000, 328, 473.

76 (65) Dunning Jr., T. H. J. Chem. Phys. 1989, 90, 1007.

(66) Martins, L. S. C.; Jorge, F. E.; Machado, S. F. Mol. Phys. 2015, 113, 3578.

(67) Schmidt, M. W.; Baldridge, K. K.; Boatz, J. A.; Elbert, S. T.; Gordon, M. S.;

Jensen, J. H.; Koseki, S.; Matsunaga, N.; Nguyen, K. A.; Su, S.; Windus, T. L.;

Dupuis, M.; Montgomery Jr., J. A. J. Comput. Chem. 1993, 14, 1347.

(68) DIRAC, a relativistic ab initio electronic structure program, Release

DIRAC12 (2012), written by H. J. Aa. Jensen, R. Bast, T. Saue, and

L. Visscher, with contributions from V. Bakken, K. G. Dyall, S. Dubillard,

U. Ekstrom,¨ E. Eliav, T. Enevoldsen, T. Fleig, O. Fossgaard, A. S. P. Gomes,

T. Helgaker, J. K. Lærdahl, Y. S. Lee, J. Henriksson, M. Ilias,ˇ Ch. R. Jacob,

S. Knecht, S. Komorovsky,´ O. Kullie, C. V. Larsen, H. S. Nataraj, P. Norman,

G. Olejniczak, J. Olsen, Y. C. Park, J. K. Pedersen, M. Pernpointner, K. Ruud,

P. Sałek, B. Schimmelpfennig, J. Sikkema, A. J. Thorvaldsen, J. Thyssen,

J. van Stralen, S. Villaume, O. Visser, T. Winther, and S. Yamamoto (see

http://www.diracprogram.org).

(69) Visscher, L.; Dyall, K. G. J. Chem. Phys. 1996, 104, 9040.

(70) Seto, J. Y.; Morbi, Z.; Charron, F.; Lee, S. K.; Bernath, P. F.; Roy, R. J. L. J.

Chem. Phys. 1999, 110, 11756.

(71) Huber, K. P.; Herzberg, G., Chemistry Spectra and Molecular Structure IV.

Constants of Diatomic Molecules; Van Nostrand, New York: 1979.

(72) Hess, B. A. Phys. Rev. A 1985, 32, 756.

77

Chapter 5 Implementation of LUT-IODKH in GAMESS program

5.1 Introduction

Relativistic effects are essential for highly accurate calculations of molecular struc- tures and properties for molecules containing heavy elements. To describe the relativistic effects, 2c relativistic schemes are among the most practical methods.

In fact, most quantum chemical packages contain various 2c relativistic schemes.

For example, the GAMESS program package1 implements NESC2, RESC3, and

first- to third-order DKH4–9, and IOTC10 schemes. In particular, IOTC gives results equivalent to the 4c relativistic calculations in one-electron systems. This chapter explains the implementation of LUT-IOTC in GAMESS. The present chapter is organized as follows. The implementation is explained in Section 5.2. Numer- ical assessments for correct implementations are performed in Section 5.3. The chapter conclusion is stated in Section 5.4.

5.2 Implementation 5.2.1 Relativistic correction in GAMESS program

This subsection explains the relativistic correction in GAMESS program from the implementation viewpoint. Relativistic correction is applied in a one-electron in- tegral routine, ONEEI. After NR integrals such as overlap integrals, kinetic energy

79 integrals, and nuclear-attraction integrals, are evaluated, so-called Darwin-like in- tegrals (pVp) are evaluated using a partial integration technique in the subroutine called PVPINT. These four types of integrals are then used in relativistic correc- tion routines in order to implement the Hess’ numerical technique, followed by subroutines for 2c relativistic transformation: HSANDTR for NESC, RESCX for

RESC, QRELX for DKH, and QRELX IOTC for IOTC. All the relativistic correc- tions are summed up and added to a total one-electron Hamiltonian and saved on a disk.

5.2.2 Threshold setting in the IOTC method

This subsection explains threshold used in the IOTC method. IOTC is one of the most rigorous 2c methods due to a complete block-diagonalization of the 4c Dirac

Hamiltonian. Resulting electronic Hamiltonian is free from a positronic informa- tion. This separation is based on the condition represented by “R-operator” or

“Y-operator” equation in Eq. (2.28), whose solution is obtained by an iterative procedure adopted in the original paper10. Note that this equation, which is cate- gorized as a Riccati-type equation, can be solved without iteration as proposed by

Ke¸dziera and Barysz11. For its convergence, the maximum difference of individ- ual Y matrix elements should be less than a threshold. In GAMESS, the default

− threshold was set to 1.0 × 10 7.

Another threshold in IOTC is a linear dependency of basis sets used in the

Hess’ numerical evaluation. In the Hess’ technique, eigenvectors of p2-space are spanned by a linear combination of given basis sets for calculation. To keep the accuracy of the RI approximation, internally uncontracted basis sets are adopted, which means that uncontracted basis sets are utilized in the linear combination for the eigenvectors before a contraction is performed at the end of the IOTC subroutine. The contracted basis sets are automatically uncontracted in GAMESS,

80 which generates the same exponents and angular momentum primitive basis functions without considering a linear dependency of the resultant basis functions due to a simplicity of the uncontraction process. For getting stable results, extra

− default threshold for a linear dependency was set to 1.0 × 10 6 in the original version of GAMESS.

5.2.3 Integral evaluation at quadruple precision

This subsection explains the integral evaluation at a quadruple precision with a relativistic basis set. Most of the codes in quantum chemistry adopt double precision for real variables, because the precision of approximately 16 decimal digits is sufficient for obtaining accurate results. The accuracy depends on basis sets used in a calculation, which consist of exponents and contraction coefficients.

The GAMESS program uses only Gaussian basis functions. In terms of integral evaluation, larger exponents included in the basis set can present a more critical problem for the accuracy. For example, when the exponent value is larger than

1.0×1010, kinetic energy integrals can exceed the range of a double precision float- ing point format. This is the case for relativistic basis sets describing a relativistic contraction of s-type orbitals. Such basis demands the use of a quadruple preci- sion floating point format that requires longer CPU time compared to a double precision because it potentially has a 32 decimal digits representation.

The GAMESS program had a routine for kinetic energy integration at a quadruple precision named QSANDT. This routine is designed for other rela- tivistic methods to perform the Hess’ numerical evaluation without the loss of accuracy. All of the kinetic energy and overlap integrals are simultaneously eval- uated with a quadruple precision in the routine, which are required to be partially separated in LUT-IOTC because the LUT technique defines the relativistic unitary transformation within an atom. For improving the performance of LUT with a

81 quadruple precision, new subroutines, QSANDTL and QSANDTLUT, are imple- mented. The former routine is for the evaluation within an atom while the latter is for the counterpart. These subroutines correspond to the first term of the first line and the first term of the third line in the r. h. s. of Eq. (2.36), respectively.

5.2.4 Combination with DC

This subsection explains the connection with the DC method. The DC method is one of the large-scale molecular theories proposed by Yang12, which divides a molecule into several localization regions. One of the outstanding features of DC is an introduction of a buffer region around a central region for taking environmental effects into account. This controls a balance between the accuracy and efficiency of

DC. From the early 2000s, Nakai group has extended its applicability not only to the HF and DFT methods13 but also to post-HF methods such as MP and CC14–19.

In terms of implemented DC codes in GAMESS, the DC subroutine works like a wrapper routine at both SCF and post-HF steps, meaning that ordinary subroutines are called from it with no additional computation. To combine LUT-

IOTC and DC methods, one only pays attention to the size of basis functions in a localization region automatically determined in the DC routines before in- tegral evaluations, because no relativistic methods had been supported in the original GAMESS package. As mentioned in the previous subsection, LUT uses the internally uncontracted basis functions and evaluates one-electron integrals in uncontracted forms, followed by the contraction at the end of LUT. In addition, the NR integrals in Eq. (2.32) are evaluated using contracted basis functions, but in other relativistic methods, all of one-electron integrals are evaluated in uncon- tracted form. This difference is responsible for mismatching the dimensions of the localization region, which is determined before entering the one-electron integral evaluation.

82 5.2.5 Input options for LUT-IOTC

This subsection describes the input option settings of LUT-IOTC that are available in a user input file GAMESS uses a namelist style input. LUT related namelists are

CONTRL and RELWFN. CONTRL namelist is used for specifying chemical control data (fundamental job options), such as types of wavefunctions and calculations, and RELWFN is used for specifying a relativistic correction, such as a floating- point type, threshold values, and detailed settings for individual 2c methods.

In CONTRL namelist, “relwfn=lut-iotc” option turns on the LUT functionality for both energy and gradient calculation cases. In RELWFN namelist, one new option ʠ tauʡ sets a threshold for relativistic transformation range in Eq. (2.32).

The default values of “tau” is 3.5 Å. Note that theoretical differences between

IOTC and LUT-IOTC still exist when the energy difference seems to be converged numerically because a unitary transformation of LUT is defined within an atom independent from “tau”.

5.2.6 Major capabilities of LUT-IOTC

This subsection describes capabilities of the implemented LUT-IOTC in GAMESS.

Generally, because the 2c SF relativistic treatment replaces only one- electron integrals, all NR supported methods are straightforwardly connected with LUT-

IOTC not only for energy but also for gradient calculations. This is true for HF,

DFT, post-HF, and excited-state methods. Note that a semi-numerical scheme using differences between analytical gradient values is supported for a frequency calculation. However, methods such as the Morokuma analysis that are using the basis set information and/or a unique subsystem different from LUT are not combined with LUT. An extra programing is required to use those methods.

Figure 5.1 shows the sample input file for the geometry optimization for cisplatin using LUT-IOTC at DFT (B3LYP) level of theory. Similar to NR and other 2c

83 $contrl runtype=optimization scftype=rhf relwfn=lut-iotc dfttyp=b3lyp $end $basis gbasis=SPKrTZC $end $data Sample input of cis-Platin Cnv 2 Pt 78.0 0.00000 0.00000 -0.12244 Cl 17.0 0.00000 1.81826 1.46113 N 7.0 0.00000 1.59816 -1.56285 H 1.0 0.82604 1.63808 -2.15006 H 1.0 0.00000 2.39652 -0.92766 $end

Figure 5.1 Sample input of geometry optimization for cis-Platin.

relativistic methods, LUT-IOTC uses a molecular symmetry, for example, C2v in cisplatin, for reducing computational cost.

5.3 Numerical assessment 5.3.1 Computational details

This section examines the performance of LUT-IOTC from the viewpoint of ac- + curacy and efficiency compared with NR and IOTC. The total energies of Fm99 ,

Rn, and Au were calculated for verifying integrity of implementation because + an atomic electronic energy is independent from the LUT approximation. Fm99 calculation used Cartesian basis functions at a quadruple precision in evaluating kinetic energy and overlap integrals followed by a matrix diagonalization. Rn and

Au calculations used spherical harmonic basis functions in double precision and quadruple precision formats in those integrals and diagonalization, respectively. − A hypothetical heavy molecule, S Sb(I)(SeH)(SnH3), was employed to con- firm molecular total energy with Cartesian basis functions in double precision formats. The threshold dependence of the Y operator iterative solution was in-

. × −7 . × −15 vestigated using Au2. The threshold changed from 1 0 10 to 1 0 10 with

−1 10 step size. The bond length of Au2 was fixed at 3.1751 Å, which is equal to

84 6.0 bohr. For efficiency in large-scale calculations combined with the DC method, = , , ..., (HF)n zigzag molecules (n 1 2 100) were employed. The bond length of HF was fixed at 0.97 Å, while the inter-molecular distance, HF···HF was set to 1.53 ◦ ffi Å. The bond angle for atoms H, F, and H was 120 . For e ciency, fac-Ir(ppy)3 was adopted.

The atomic energies and threshold dependence were calculated at the HF level. The efficiency was tested at the HF, MP2, and CCSD levels of theory.

A quadruple precision was used for basis functions with exponents larger than

1.0 × 109 in order to maintain the accuracy of integrals. For Fm, uncontracted

50s basis set was used. The well-tempered basis set20,21, (28s24p18d12f), and universal basis set22, (32s29p20d15f), were employed for Rn and Au, respectively. − 23–25 For S Sb(I)(SeH)(SnH3), the Sapporo-DKH3-DZP-2012 basis sets were used for all elements. The relativistic first-order polarized basis set26 was used for Au for threshold dependency. The basis sets for both H and F was 6-311G(d,p)27.In

− / 28 fac Ir(ppy)3, the contracted DKH3-Gen-TK NOSeC-V-TZP for Ir and cc-pVDZ − for H, C, and N atoms were adopted. The NR calculation for fac Ir(ppy)3 adopted the basis set for Ir, TZP29. All the calculations were performed using the GAMESS program.

5.3.2 Total energies of heavy atoms and molecules

This subsection examines the correct implementation of LUT-IOTC with dou- ble/quadruple precision and Cartesian/spherical basis functions. Table 5.1 shows the total energies of heavy elements obtained by IOTC and LUT-IOTC at the HF level. The basis and format represent basis functions and floating point precision adopted, respectively. The energy difference ΔE between LUT-IOTC and IOTC verifies the LUT implementation integrity.

All of the total energies obtained by LUT-IOTC and IOTC agreed well with

85 Table 5.1 Total energies (hartree) of heavy-element atoms and a heavy molecule obtained by IOTC and LUT-IOTC at the HF level. basis format IOTC LUT-IOTC ΔE + Fm99 Cart qp −5939.195136 −5939.195136 0.000000 Rn sph dp −23563.597363 −23563.597363 0.000000 Au sph qp −19012.301060 −19012.301060 0.000000 − − − S Sb(I)(SeH)(SnH3) Cart dp 22593.075520 22593.075374 0.000146

+ each other. There were no differences in total energies for Fm99 , Rn, and Au

− calculated to the precision of 10 6 hartree. As compared with the conventional

IOTC, both Cartesian and spherical basis functions produced no errors in calcu- lations using the LUT routine, in which the Hess’ numerical evaluation technique was modified to run with LUT. In addition to this, as stated in the subsection

5.2.3, the integral evaluation routine at quadruple precision was modified. These algorithmic changes did not introduce any change in the resultant total energies. − Δ For S Sb(I)(SeH)(SnH3), E was 0.146 millihartree due to the local approx- imation in a unitary transformation. This energy difference between IOTC and

LUT-IOTC is quite reasonable according to previous works. These results indicate that the LUT scheme is implemented correctly in total energy calculations.

5.3.3 Threshold dependence of IOTC and LUT-IOTC

This subsection investigates the total energy dependence on the internal threshold in the iterative procedure of IOTC. Figure 5.2 shows the threshold dependence of iterative solution for Y operator in IOTC for total energy of Au2 at the HF level. The horizontal axis shows a threshold value of convergence criteria in the iterative solution while the vertical axis shows the energy difference (ΔE) between total

− − − energies at 10 n and 10 (n 1) (n = 8, 9, ..., 15) thresholds.

In both IOTC and LUT-IOTC, ΔE converged smoothly with an increase in

− threshold parameter n. The default threshold value in GAMESS (10 7) resulted in a slightly lower energy than those obtained with more restrictive thresholds.

86 0.0025 IOTC 0.0020 LUT-IOTC

0.0015 [mhartree]

E 0.0010 ¨

0.0005

0 10-8 10-9 10-10 10-11 10-12 10-13 10-14 10-15 Threshold Figure 5.2 Threshold dependence of the iterative solution for Y operator in total energy of

Au2 obtained with IOTC and LUT-IOTC at the HF level.

− At 10 11, ΔE becomes considerably smaller than 0.0001 millihartree. Considering the applicability of LUT-IOTC in large scale molecules, the error in the calculated energy should be less than 1 kcal/mol. The obtained threshold dependence results indicate that for more reliable energy determination required in calculations with

− LUT-IOTC the upper bound for the threshold should be 1.0 × 10 9 or less.

5.3.4 Computational cost of LUT-IOTC

This subsection examines the computational cost of energy and gradient calcula- tions for the LUT-IOTC method with/without the DC scheme. Figure 5.3 shows the system-size dependence of CPU times in (HF)n molecules at the HF, MP2, and CCSD levels. The horizontal axis indicates the total numbers of the molecules n. The vertical axis indicates the CPU time in seconds. The estimated scaling of CPU time with n is shown next to the graph. A Xeon E5-2637v3/3.50 GHz processor (Quad core) on a single core was utilized to measure the CPU time. The scalings for all levels of the theory using the DC scheme is quasi-linear with a

87 n5.36 100000 n3.89 DC-CCSD DC-MP2 n1.23 3.03 10000 DC-HF n CCSD MP2 1000 HF n1.28 n1.09 100 CPU time [s]

10

1 1 2461020 40 60 100 n

Figure 5.3 System-size dependence of CPU time (s) in (HF)n molecules at the HF, MP2, and CCSD levels with and without DC combined with LUT-IOTC. A Xeon E5-2637v3/3.50 GHz processor (Quad core) was used on a single core.

...... small prefactor: n3 03 to n1 28 in HF, n3 89 to n1 09 in MP2, and n5 36 to n1 23 in CCSD.

This result is the same as that discussed for the NR Hamiltonian.

Table 5.2 shows the CPU time percentages (in columns 3 and 5) of individual steps for one-cycle geometry optimization of Ir(ppy)3 with the NR and LUT-IOTC methods. The total CPU times for both methods are close to each other: 14671.86 s for NR and 11501.70 s for LUT-IOTC, because the LUT scheme requires small time to evaluate the relativistic transformation based on the nature of localization in relativistic effects. Moreover, the individual percentages of the calculation steps are similar. The bottleneck of a geometry optimization in both NR and

LUT-IOTC is the evaluation of derivatives of two-electron integrals. These results indicate that there are no drawbacks in adopting the LUT-IOTC method based on computational cost considerations.

88 Table 5.2 CPU time and percentages of seven steps in geometry optimization for fac-Ir(ppy)3 at the HF level with NR and LUT-IOTC. A Xeon X5680/3.33 GHz processor (Hexa core) was used on a single core. NR LUT-IOTC Time (s) % Time (s) % OEI 2.26 0.02 9.16 0.08 Guess 2.34 0.02 2.25 0.02 TEI 3091.65 21.07 3017.26 26.23 SCF 2635.95 17.97 3227.26 28.06 d-rel. – – 11.05 0.10 dOEI 22.77 0.16 350.61 3.05 dTEI 8916.89 60.78 4884.11 42.46 Total 14671.86 100.00 11501.70 100.00

5.4 Conclusion

In this chapter, details of implementation of 2c relativistic LUT-IOTC method in the GAMESS program package were explained from the coding perspective. The previous GAMESS package had the functionality of relativistic corrections such as NESC, RESC, first- to third-order DKH, and IOTC. Based on these implemen- tations, especially the implementation for IOTC, which is theoretically equivalent to IODKH, energy and gradient evaluations for the LUT-IOTC method were implemented to improve the efficiency of the relativistic calculations. Several routines were carefully updated to use the LUT scheme, such as QRELX IOTC for atomic unitary transformations. Other routines required for LUT were newly added to GAMESS: driver routine LUTX and gradient driver routine LUTGX. The large-scale molecular theory, DC, was supported by relativistic corrections with

LUT-IOTC for the first time in GAMESS. The compatibility of LUT-IOTC with other GAMESS functionality stems from the fact that only one-electron integrals are treated in LUT-IOTC. Numerical assessments were performed to validate the implementation of LUT-IOTC in GAMESS and evaluate its efficiency by perform- ing test calculations for heavy atoms, heavy molecule, HF zigzag clusters, and

89 Ir(ppy)3. All of the functionality worked properly based on the comparison with the conventional IOTC and NR results. The GAMESS program including the LUT technique was released in August, 2016.

90 References

(1) Schmidt, M. W.; Baldridge, K. K.; Boatz, J. A.; Elbert, S. T.; Gordon, M. S.;

Jensen, J. H.; Koseki, S.; Matsunaga, N.; Nguyen, K. A.; Su, S.; Windus, T. L.;

Dupuis, M.; Montgomery Jr., J. A. J. Comput. Chem. 1993, 14, 1347.

(2) Dyall, K. G. J. Chem. Phys. 1997, 106, 9618.

(3) Nakajima, T.; Hirao, K. Chem. Phys. Lett. 1999, 302, 383.

(4) Foldy, L. L.; Wouthuysen, S. A. Phys. Rev. 1950, 78, 29.

(5) Douglas, M.; Kroll, N. M. Ann. Phys. 1974, 82, 89.

(6) Nakajima, T.; Hirao, K. J. Chem. Phys. 2000, 113, 7786.

(7) Barysz, M. J. Chem. Phys. 2001, 114, 9315.

(8) De Jong, W. A; Harrison, R. J.; Dixon, D. A. J. Chem. Phys. 2001, 114, 48.

(9) Reiher, M.; Wolf, A. J. Chem. Phys. 2004, 121, 10945.

(10) Barysz, M.; Sadlej, A. J. J. Chem. Phys. 2002, 116, 2696.

(11) Ke¸dziera, D.; Barysz, M. Chem. Phys. Lett. 2007, 446.

(12) Yang, W. Phys. Rev. Lett. 1991, 66, 1438.

(13) Akama, T.; Kobayashi, M.; Nakai, H. J. Comput. Chem. 2007, 28, 2003.

(14) Kobayashi, M.; Akama, T.; Nakai, H. J. Chem. Phys. 2006, 125, 204106.

(15) Kobayashi, M.; Imamura, Y.; Nakai, H. J. Chem. Phys. 2007, 127, 074103.

(16) Kobayashi, M.; Nakai, H. J. Chem. Phys. 2008, 129, 044103.

(17) Kobayashi, M.; Nakai, H. Int. J. Quantum Chem. 2009, 109, 2227.

(18) Kobayashi, M.; Nakai, H. J. Chem. Phys. 2009, 131, 114108.

(19) Kobayashi, M.; Nakai, H. Phys. Chem. Chem. Phys. 2012, 14, 7629.

(20) Huzinaga, S.; Miguel, B. Chem. Phys. Lett. 1990, 175, 289.

(21) Huzinaga, S.; Klobukowski, M. Chem. Phys. Lett. 1993, 212, 260.

(22) Malli, G. L.; Silva, A. B.F. D.; Ishikawa, Y. Phys. Rev. A 1993, 47, 143.

(23) Yamamoto, H.; Matsuoka, O. Bull. Univ. Electro. Comm. 1992, 5, 23.

91 (24) Noro, T.; Sekiya, M.; Koga, T. Theor. Chem. Acc. 2003, 109, 85.

(25) Noro, T.; Sekiya, M.; Koga, T. Theor. Chem. Acc. 2012, 131, 1124.

(26) Kello, V.; Sadlej, A. J. Theor. Chem. Acc. 1996, 94, 93.

(27) Krishnan, R.; Binkley, J. S.; Seeger, R.; Pople, J. A. J. Chem. Phys. 1980, 72, 650.

(28) Dunning Jr., T. H. J. Chem. Phys. 1989, 90, 1007.

(29) Martins, L. S. C.; Jorge, F. E.; Machado, S. F. Mol. Phys. 2015, 113, 3578.

92 Chapter 6 Relaxation of core orbitals in the frozen core potential treatment

6.1 Introduction

The development of a highly accurate and efficient scheme to treat any elements in the periodic table is one of the most significant problems in quantum chem- istry. One rigorous scheme is the 4c formalism based on the Dirac equation, which satisfies the Lorentz invariance for the motion of electrons1. This scheme automatically includes both the spin-free and spin-dependent relativistic effects.

Recent theoretical and computational developments in the 4c method have en- larged its applicability for treating large molecules2–4. An alternative relativistic scheme is a 2c formalism that uses only electronic states. In the last decade, sev- eral schemes have been proposed, such as NESC5, X2C6, and IODKH7 methods, which provide the same energies as the four-component method. From the view- point of efficiency in 2c formalisms, the locality of the relativistic effect has been considered8–16. Seino and co-workers have also proposed a scheme, termed LUT, for both one- and two-electron relativistic transformations10,11. Furthermore, the

LUT scheme was combined with the DC method to achieve an overall linear- scaling technique from integral calculations to post-HF calculations12–14. These improvements can reduce the gap between non-relativistic and highly accurate relativistic methods in AE calculations.

93 A widely used method to reduce the number of electrons treated explicitly is the ECP method. The ECP method, which includes the PP17–23 and model potential (MP)24–27 methods, replaces the effect of the core electrons with a core potential. The ECP method includes part of the relativistic effects in the core potential. The PP method introduces a smooth and nodeless potential function in the core region. On the other hand, the MP method uses a nodal potential function to describe the valence orbital in the AE treatment. Seino et al has also proposed another scheme for reducing the number of electrons treated explicitly, which is termed the FCP method28. The FCP method is based on the same equation as the MP method but the core electrons obtained from atomic calculations are treated explicitly. Another advantage of the FCP method is a seamless connection to the AE method with an accurate 2c relativistic Hamiltonian. The ECP and

FCP methods are quite effective for determining valence electronic states and valence properties. However, the ECP method has difficulties in describing core electron states and their properties because the core potential is obtained by atomic calculations without the consideration of any chemical environment effect.

In the present chapter, the author proposes a novel scheme based on the FCP method to describe core properties using the relaxations of core electrons. This chapter is organized as follows; the theory and implementation are described in

Section 6.2; numerical assessments are performed in Section 6.3; and concluding remarks are given in Section 6.4.

6.2 Theory and implementation 6.2.1 FCP with relaxation of core electrons

This subsection provides the theory for core electron relaxations based on the

FCP scheme. The FCP method is based on the Huzinaga–Cantu equation29, which separates electrons into core and valence regions. Here, a closed-shell

94 system is discussed for simplicity, though extension to an open-shell system is straightforward. The AE HFR equation and Fock matrix are given by (2.54) and

Eq. (2.55), respectively. The density matrix D is separated into core and valence contributions as Eqs. (2.58)-(2.60).

In the valence FCP method, two approximations are adopted. First, the total density matrix D is approximated in a block-diagonal form, Eq. (2.61). This approximation is effective if a generally contracted basis set is adopted. Second, the molecular core density matrix DC is approximated as the direct sum of the atomic contributions in Eq. (2.63). Using these two approximations, the valence- only HFR equation is expressed as

fV(dV)cV = sVcVεV (6.1) with       V V 1 V V V  V V 1 V V  V V fμν =hμν + dλρ χμ χλ χν χρ − χμ χλ χρ χν 2 λ,ρ 2        1 C V C  V C 1 V C  C V + d˜ χμ χλ χν χρ − χμ χλ χρ χν 2 A λρ 2 (6.2) λ,ρ A

C − μ ν 2 s lwlmsm l,m and core C = ε ∗ , wμν ccμccνc (6.3) c χ χ |χ χ ε where μ λ ν ρ is the two-electron integral, S is the overlap matrix, c is the { } ffi core orbital energy, and cμc is the core orbital coe cient. The last term is a projection operator that prevents electrons from falling into core orbitals. The valence energy in the FCP scheme is defined as    − C − C ZA n ZB n V 1 V V V A B E = dμν hμν + fμν + (6.4) 2 RAB μ,ν A

C where ZA, nA, and RAB are the nuclear charge, the number of core electrons in atom A, and the internuclear distance between atoms A and B, respectively.

95 After the valence SCF calculation, the core electron relaxations are performed.

For the relaxation, the equation is derived using the above first approximation of

Eq. (2.61) in the FCP method. A similar derivation for the valence-only equation is performed to obtain a core-only equation. The resulting equation is given by

fC(dC)cC = sCcCεC (6.5) with       C C 1 C C C  C C 1 C C  C C fμν =hμν + dλρ χμ χλ χν χρ − χμ χλ χρ χν 2 λ,ρ 2       + 1 V χCχV  χCχV − 1 χCχV  χVχC dλρ μ λ ν ρ μ λ ρ ν (6.6) 2 λ,ρ 2

V − μ ν 2 s lwlmsm l,m and

V = ε ∗ . wμν vcμvcνv (6.7) v

Using the fixed valence orbitals obtained from the valence FCP calculation, the core-only equation in Eq. (6.5) is solved self-consistently to determine the relaxed core orbitals and their energies.

A core energy is defined as    − V − V ZA n ZB n C 1 C C C A B E = dμν hμν + fμν + , (6.8) 2 RAB μ,ν A

V where nA is the number of valence electrons in atom A.

6.2.2 Implementation

This subsection gives the implementation of FCP-CR. FCP-CR consists of three steps: an atomic calculation, a valence calculation, and a core calculation as shown in Table 6.1. In atomic calculations, each atomic AE calculation included in a target

96 Table 6.1 Calculation procedures for FCP-CR. (I) Atom (I.1) OEIs and TEIS for each atom (I.2) AE HF calculation for each atom (I.3) Save core orbital, density matrix, and orbital energy (II) Valence (II.1) Core-valence TEIs in Eq. (6.2) (II.2) Construct core potential (II.3) Construct projection operator matrices in Eq. (6.2) (II.4) OEIs and TEIs for valence (II.5) Add projection operator matrix and core potential (II.6) Solve valence-only equation of Eq. (6.1) (III) Core (III.1) Construct fixed valence potential in Eq. (6.6) (III.2) Construct projection operator matrices in Eq. (6.6) (III.3) OEIs and TEIs for core (III.4) Add projection operator matrix and fixed valence potential (III.5) Solve core-only equation of Eq. (6.5) molecule is performed to determine the atomic core orbitals and orbital energies.

These are then used in the atomic core density matrix in Eq. (2.58) and the core potential and projection matrices in Eq. (6.2). In a valence calculation, the valence- only equation in Eq. (6.1) is solved using the information for the atomic core electrons. Here, valence–valence two-electron integrals, which require smaller basis set dimensions than AE, are evaluated. In addition, the computational scaling for calculations of the core–valence two-electron integrals is theoretically reduced from O(n4)toO(n3). Finally, in a core calculation, the SCF calculation based on the core-only equation in Eq. (6.6) is performed using a fixed valence potential, which is determined using Eq. (6.6). Note that the core–core and core– valence two-electron integrals with smaller computational costs than the total two-electron integral are evaluated. Throughout this Letter, the calculations of the core–valence two-electron integrals for Eq. (6.6) are skipped because these integrals, which are the same as those used in Eq. (6.2), are stored on a disk.

97 6.3 Numerical assessments 6.3.1 Computational details

This section examines the performance of the FCP-CR scheme. The LUT-IODKH

Hamiltonian was applied to all the calculations. Numerical assessments were performed to investigate the accuracy and efficiency of the FCP-CR scheme. The

Core, valence, and total HF energies, as well as the orbital energies, of coinage = metal dimers M2 (M Cu, Ag, and Au) were calculated for accuracy. The bond

30,31 lengths were 2.55, 2.89, and 2.89 Å for Cu2,Ag2, and Au2, respectively . The CIEs and CLSs of C 1s and W 4f orbitals were calculated to investigate differences between core electrons in various chemical environments at the HF and SAC-

CI levels32,33. Vinyl acetate was employed for CIEs and CLSs of C 1s orbitals, whose geometry was optimized at the DFT with the B3LYP functional using cc-

34 = pVTZ basis set . In addition, W(CO)4L(L dppe, dmpe, and F-dppe) was adopted for CIEs and CLSs of W 4f orbitals. The threshold sets (hartree) for the

− − perturbation selection35 were 1.0 × 10 6 for SAC and 1.0 × 10 7 for SAC-CI in

. × −6 . × −7 vinyl acetate and 5 0 10 for SAC and 5 0 10 for SAC-CI in W(CO)4L with a non-direct algorithm. Their geometries were optimized at DFT with the B3LYP functional using Sapporo-DZP-2012 basis set for main group elements36 and the

SDD relativistic small-core effective core potential for W37. Gold clusters were adopted to investigate the efficiency. In these clusters, an experimental lattice constant of 4.062 Å with a face-centered cubic structure was adopted38. The total numbers of atoms in the Aun clusters were 14, 32, 50, 68, and 86, which correspond to 1 × 1 × 1, 1 × 1 × 3, 1 × 1 × 5, 1 × 1 × 7, 1 × 1 × 9 unit cells, respectively. In AE, the maximum number of the atoms in a cluster was 50. All the results were compared to those of the AE method.

The basis sets were DKH3-Gen-TK/NOSeC-V-TZP39–41 for C in vinyl acetate,

98 Cu, Ag, W, and Au; Sapporo-DZP-2012 for H and O in vinyl acetate; and 6-31G42,43 for H, C, O, F, and P in tungsten complexes. In gold clusters the minimal third- order DKH basis set44 was adopted. Note that a polarization function was added to the basis sets of the C and P atoms nearest to the central tungsten atom of these complexes. The valence orbitals were set to {2s,2p} for C, {3s,3p,3d,4s} for Cu,

{4s,4p,4d,5s} for Ag, {5s,5p,5d} for W, and {5s,5p,5d,6s} for Au.

HF calculation was performed using our in-house relativistic quantum chem- ical program. A modified version of the Gaussian09 suite of programs45 was employed for DFT and SAC-CI calculations.

6.3.2 Computational cost of FCP-CR

This subsection examines the efficiency of FCP-CR. Table 6.2 shows the CPU times of the HF calculation obtained with FCP-CR and AE for a Au50 cluster at the following four steps: core or valence potential (Potential), OEI, TEI, and

SCF procedure. Valence is the CPU time for the valence electron calculation, which corresponds to FCP, whereas core is the CPU time for the relaxation of core electrons. All is a sum over both the valence and core CPU times. A Xeon

E5-2690/2.90 GHz (Octa-core) processor was used on a single core.

The total CPU time of FCP-CR (all) was 34.4% of that for the AE case: 70816.3 and 203983.8 s for FCP-CR and AE, respectively. The total CPU time was mainly reduced by the TEI and SCF calculations. In the TEI calculation, the number of integrals was decreased by treating only valence or core electrons. The SCF calculation was performed with a smaller dimension basis set in FCP-CR com- pared with AE. In addition, the CPU time for valence was dominant in FCP-CR compared with that of core. = Figure 6.1 shows the system-size dependence of the CPU time in Aun (n

14, 32, 50, 68, and 86). The horizontal axis shows the number of gold atoms in

99 25000 AE 20000 FCP-CR

15000 O(n2.71) O(n2.28)

CPU times [s] 10000

5000

0 020406080 100 n = Figure 6.1 System-size dependence of CPU time (s) in Aun (n 14, 32, 50, 68, and 86) ob- tained with the AE and FCP-CR methods. A Xeon E5-2690/2.90 GHz (octa-core) processor was used on a single core.

the cluster. The vertical axis shows the total CPU time of FCP-CR and AE. FCP-

. CR reduces the CPU time in comparison with AE. The scalings are O(n2 71) and

. O(n2 28) for AE and FCP-CR, respectively.

To investigate the scaling for FCP-CR, Figure 6.2 shows the system-size de- pendence of the CPU time for dominant steps: TEI, the SCF procedure, and construction of a potential (Potential). The CPU times were measured at (a) AE,

(b) FCP-CR for valence electrons (equal to FCP), and (c) FCP-CR for core electrons.

The scaling of the CPU time in each step is given in Figure 6.2. In the AE, the

. . TEI and SCF scalings are O(n2 50) and O(n4 32). Thus, the scaling of the total CPU

. time becomes O(n2 71). In FCP-CR, the CPU time for the valence calculation is larger than that for the core calculation. This is because the dimension of the basis set is larger. Furthermore, the TEI steps in the valence and core calculations are

. . dominant. The scalings are O(n2 53) and O(n2 18) for valence and core, respectively,

. and resulting scaling of the total CPU time is O(n2 28).

100 Table 6.2 CPU time for four steps in the Hartree–Fock calculation obtained with FCP-CR and AE for a Au50 cluster: construction of core or valence potential (Potential), OEI, TEI, and SCF procedure. A Xeon E5-2690/2.90 GHz processor (octa-core) was used on a single core. The percentages of individual steps are given in parentheses. FCP-CR Valence Core All AE 101 Step Time (s) % Time (s) % Time (s) % Time (s) % Potential 20120.4 (44.5) 175.9 (0.7) 20296.3 (28.9) 0.0 (0.0) OEI 328.4 (0.7) 719.0 (2.9) 1047.4 (1.5) 985.1 (0.5) TEI 24363.1 (53.8) 23991.0 (96.2) 48354.1 (68.9) 149132.0 (73.1) SCF 440.0 (1.0) 48.5 (0.2) 488.5 (0.7) 53866.7 (26.4) Total 45251.9 (100.0) 24934.4 (100.0) 70186.3 (100.0) 203983.8 (100.0) 150000 150000 150000 TEI (a)TEI (b)TEI (c) 120000 SCF 120000 SCF 120000 SCF Potential Potential

90000 O(n2.50) 90000 90000

O(n2.53) O(n2.18) 60000 60000 60000 CPU time [s] CPU time [s] CPU time [s] 102

O(n2.09) 30000 O(n4.32) 30000 30000 O(n2.29) O(n3.46) O(n2.28) 0 0 0 02040 60 80 100 02040 6080 100 0 2040 60 80 100 n n n = Figure 6.2 System-size dependence of CPU time (s) for dominant steps with the AE and FCP-CR methods in Aun (n 14, 32, 50, 68, and 86): two-electron integral (TEI), SCF procedure, and construction of potential (Potential). (a) AE; (b) valence part in FCP-CR; and (c) core part in FCP-CR. A Xeon E5-2690/2.90 GHz (octa-core) processor was used on a single core. 6.3.3 Accuracy of FCP-CR

This subsection examines the accuracy of FCP-CR. Table 6.3 shows the core, va- lence, and total HF energies (in hartree) obtained with the LUT-IODKH Hamilto- = nian for M2 (M Cu, Ag, and Au), as well as the results of the FCP and molecular AE methods. The deviations from the molecular AEs are shown in parentheses.

The energy components of the molecular AE are evaluated using the procedure described in a previous study28.

The difference in the total energy of FCP is less than 0.09 hartree: 0.001430,

0.08759, and 0.007563 hartree for Cu2,Ag2, and Au2, respectively. For each molecule, the difference in the core part is smaller than that of the valence part because the core regions are similar to that in the atomic state. On the other hand, the total energies of FCP-CR give small deviations compared with FCP: 0.000188,

0.000117, and 0.001538 hartree for Cu2,Ag2, and Au2, respectively. The deviations in the valence energy are speculated to be canceled out by those in the core energy because the core electrons are relaxed self-consistently under the frozen valence potential.

Figure 6.3 shows the orbital energy differences (in hartree) from those of the

AE method in coinage metal dimers (Cu2,Ag2, and Au2) obtained by FCP and FCP-CR. The vertical axis shows the difference from AE in the orbital energy,

Δε. The horizontal axis indicates the orbital species from the core to the valence regions. The dotted line indicates the boundary between the core and valence regions.

The values Δε in the valence orbitals of FCP-CR are the same as those of FCP because these schemes are common until the valence electrons are solved self- consistently. The mean absolute deviations of both schemes are 0.04, 0.04, and

0.03 eV for Cu2,Ag2, and Au2, respectively, which is consistent with the results

103 of a previous study28.

The Δε values in the core orbitals of FCP-CR are smaller than those of FCP.

The mean absolute deviations of FCP-CR are 0.04, 0.0, and 0.02 eV for Cu2.Ag2, and Au2, respectively, while those of FCP are 0.24, 0.22, and 0.04 eV for Cu2,Ag2, and Au2, respectively. Thus, the core orbitals can be properly relaxed under a molecular valence potential.

6.3.4 Core ionization energy and core level shift

This subsection discusses the accuracies of the CIEs and CLSs. In HF calculations, the CIEs and CLSs were estimated from the orbital energies based on Koopmans’ theorem. In SAC-CI calculations, these values were directly obtained.

Table 6.5 shows the CIEs (in eV) of C 1s and W 4f in vinyl acetate and W(CO)4L complexes, respectively. Vinyl acetate has four types of C 1s with different chem- ical environments: COO, CHO, CH3, and CH2. In addition, W(CO)4L has three types of W 4f with different ligands: dppe, dmpe, and F-dppe. The CLSs (in eV), which are shown in parentheses, were estimated using the differences from the reference values. As references, CH3 and dppe were adopted for vinyl acetate and

W(CO)4L, respectively. The experimental values for vinyl acetate and W(CO)4L were taken from Refs. [46] and [47], respectively.

In FCP, the CIEs at the HF level were always the same value in different chemical environments: 308.298 eV and 47.540 eV for vinyl acetate and W(CO)4L, respectively. On the other hand, in FCP-CR, the CIEs depended on the chemical environment. The CIEs and CLSs were close to those of AE, within a deviation of 0.5 eV. Furthermore, the electron correlation effect, which is the difference be- tween HF and SAC-CI, contributed to reproducing the values of the experimental

CIEs. The deviations from the experimental CIEs were less than 5.7 and 6.0 eV for vinyl acetate and W(CO)4L, respectively. Note that the deviations from the

104 = Table 6.3 Comparison of core, valence, and total HF energies (in hartree) of M2 (M Cu, Ag, and Au) obtained by the FCP, FCP-CR, and molecular AE methods. The deviations from the molecular AEs are given in parentheses. FCP FCP-CR Molecular AE − . − . − . − . Cu2 Core 2913 534822 ( 0 000063) 2913 533204 (0.001555) 2913 534759 Valence −392.856146 (−0.001367) −392.856146 (−0.001367) −392.854779 105 Total −3306.390968 (−0.001430) −3306.389350 (0.000188) −3306.389538 − . − . − . − . Ag2 Core 10332 790532 ( 0 000150) 10332 704656 (0.085726) 10332 790382 Valence −293.119642 (−0.085609) −293.119642 (−0.085609) −293.034033 Total −10625.910174 (−0.085759) −10625.824298 (0.000117) −10625.824415 − . − . − . − . Au2 Core 37750 131480 (0.001660) 37750 137505 ( 0 004365) 37750 133140 Valence −272.254111 (0.005903) −272.254111 (0.005903) −272.260014 Total −38022.385591 (0.007563) −38022.391616 (0.001538) −38022.393154 0.004

0.002 0.000

-0.002 Core orbitals Valence orbitals Core orbitals -0.004 Valence

106 orbitals Core orbitals

ǻİ>KDUWUHH@ -0.006 -0.008

FCP Valence orbitals -0.010 FCP-CR -0.012 1s2pı(3s) ı(4s) 1s 3dı(4s) ı(5s) 1s 4f ı(5s) ı(6s) Figure 6.3 Differences of between the FCP and FCP-CR orbital energies (in hartree) and those of the AE method in coinage metal dimers at the

Hartree–Fock level: (a) Cu2, (b) Ag2, and (c) Au2. The orbitals are sorted in ascending order of energy. experimental values are mainly caused by the incompleteness of the basis sets, which is shown in Appendix 6.A.

In vinyl acetate, the tendency is explained by the characteristics of the atoms or bonds surrounding a target atom. For COO, the carbon has bonds with two electronegative oxygen atoms, which leads to a decrease of the electron density at C. For CHO, the shift is smaller than that of COO because CHO has one C–O

2 bond. For CH2, the carbon has one double bond with a hybridized sp orbital. The hybridization expands the valence orbitals compared with the sp3 hybridization in CH3, which also leads to a decrease of the electron density at C.

In W(CO)4L, the tendency is explained by the characteristics of the ligands. For dppe, the shift is smaller than that of dmpe because of the screening capabil- ity of a core hole in the ligands. The π electrons of the phenylphosphine group are likely polarized in comparison with the σ electrons of the methylphosphine group, which leads to easy screening of core holes47. For F-dppe, the pentafluo- rophenylphosphine group includes highly electronegative F, and thus the electron density at W decreases.

6.3.5 Accuracy of an iterative procedure between valence and core calculations

This subsection examines the effect of an iterative self-consistent procedure be- tween the valence and core calculations. FCP-CR calculates core orbitals under a valence potential only once, after valence orbitals are calculated under an atomic core potential. However, the valence orbitals could be more accurate if the va- lence orbitals were recalculated under the core potential obtained by FCP-CR.

Here, a procedure is iteratively performed between the valence and core calcu- lations based on Eqs. (6.1), (6.2), (6.5), and (6.6). Table 6.4 shows the iteration = number dependence in total energy (hartree) of Au2. Iteration 0 and 1 indi-

107 Table 6.4 Iteration number dependence in total energy (in hartree) of Au2 obtained by the valence–core iterative procedure. Iteration Energy ΔE ΔEAE 0 (FCP) −38022.385591 – 0.007563 1 (FCP-CR) −38022.391616 −0.006025 0.001537 2 −38022.391693 −0.000077 0.001461 3 −38022.391697 −0.000004 0.001457 4 −38022.391698 −0.000001 0.001456 5 −38022.391698 0.000000 0.001456 cate FCP and FCP-CR, respectively. ΔE is the deviation in the total energy from the previous value. ΔEAE is the deviation from the reference value obtained by molecular AE; the reference value is −38022.393154 hartree. The total energy was steeply converged with a small iteration number. The largest deviation in ΔEAE is

0.001456 hartree at the fifth iteration. This deviation is due to the approximation of the basis set in Eq. (2.61). Furthermore, the largest deviation in ΔE is −0.006025 hartree at the first iteration, which is the difference between FCP and FCP-CR.

This result indicates that FCP-CR, which requires only one calculations of core orbitals under a valence potential, is enough to describe the core electrons in a chemical environment.

6.4 Conclusion

This chapter proposed the FCP-CR scheme for relaxation of core electrons based on FCP at the IODKH level. The core-only HFR equation is derived using the same procedure as that for a valence-only one. The molecular core orbitals and their energies are computed under the frozen valence molecular potential, which is constructed on the fly. Numerical assessments were performed with coinage metal dimers and small gold clusters to examine accuracy and efficiency. The

CIEs and CLSs of C 1s in vinyl acetate and W 4f in three tungsten complexes were calculated with FCP-CR at the HF and SAC-CI levels. The results suggest that

108 Table 6.5 Core ionization energies (in eV) obtained by FCP, FCP-CR, and AE of C 1s and W 4f for vinyl acetate and three tungsten complexes, respectively. Experimental values are taken from Ref. [46] for vinyl acetate and Ref. [47] for the tungsten complexes. Core level shifts from the

reference molecules (CH3 or dppe) are given in parentheses. HF SAC-CI Molecule type FCP FCP-CR AE FCP-CR AE Exptl. vinyl acetate COO 308.298 310.249 310.103 298.617 298.475 294.94 (0.000) (3.951) (3.938) (3.870) (3.860) (3.510) CHO 308.298 307.700 307.553 295.726 295.585 292.32 109 (0.000) (1.402) (1.388) (0.979) (0.970) (0.890)

CH3 308.298 306.298 306.165 294.747 294.615 291.43 (0.000) (0.000) (0.000) (0.000) (0.000) (0.000)

CH2 308.298 305.634 305.509 293.181 293.061 290.54 (0.000) (−0.664) (−0.656) (−1.566) (−1.554) (−0.890)

W(CO)4L dppe 47.540 47.589 48.033 36.274 36.558 31.118 (0.000) (0.000) (0.000) (0.000) (0.000) (0.000) dmpe 47.540 47.588 48.030 36.963 37.236 31.296 (0.000) (−0.001) (−0.003) (0.689) (0.678) (0.178) F-dppe 47.540 48.316 48.773 37.403 37.742 31.931 (0.000) (0.727) (0.740) (1.129) (1.184) (0.813) FCP-CR sufficiently reproduces the corresponding AE results with high efficiency.

The proposed scheme has a high potential for becoming an alternative and efficient

AE method, without loss of accuracy.

110 Appendix 6.A Dependence of core ionization potential energies

Core ionization potential energies were examined by the standard SAC-CI calcu- ff lations without a symmetry (C1) as well as using di erent basis sets: cc-pCVXZ

(X = D, T, and Q) and special basis sets designed for a core ionization potential in the paper written by Kuramoto and Ehara [J. Chem. Phys. 122 (2005) 014304].

Table 6.6 shows the core ionization energies (eV) of C 1s for vinyl acetate. Δ shows ff Δ the di erence from the experimental values. In the C1 symmetry, is almost the Δ same as the CS symmetry in Section 3. In the cc-pCVXZ types of basis sets, gets systematically smaller as the cardinal number X increases. For cc-pCVQZ basis set, the maximum deviation is 1.34 eV compared with 4.93 eV for cc-pCVDZ and

1.55 for cc-pCVTZ. Moreover, the special basis set gives the best results among these basis sets. The maximum deviation is 1.04 eV although the size of the basis sets is smaller than cc-pCVQZ, because the special basis set includes the differ- ential basis function for the primitive s-type of function. These indicate that the proper basis set selection requires for reproducing the experimental values.

111 Table 6.6 Core ionization energies (in eV) of C 1s for vinyl acetate with C1 symmetry, cc-pCVXZ(X = D, T, and Q), and special basis sets. Δ Δ Δ Δ Δ type C1 cc-pCVDZ cc-pCVTZ cc-pCVQZ Special Exptl. 112 COO 298.492 3.55 299.871 4.93 296.491 1.55 296.284 1.34 295.984 1.04 294.94 CHO 295.701 3.38 297.536 5.22 293.600 1.28 293.428 1.11 293.012 0.69 292.32

CH3 294.523 3.09 295.851 4.42 292.521 1.09 292.330 0.90 291.866 0.44 291.43 CH2 293.247 2.73 295.822 5.28 291.644 1.10 291.499 0.96 291.021 0.48 290.54 References

(1) Moss, R. E., Advanced Molecular Quantum Mechanics; Chapman and Hall,

London: 1973.

(2) Nakajima, T.; Yanai, M.; Hirao, K. J. Comput. Chem. 2002, 23, 847.

(3) Kelly, M. S.; Shiozaki, T. J. Chem. Phys. 2013, 138, 204113.

(4) Shiozaki, T. J. Chem. Theory Comput. 2013, 9, 4300.

(5) Dyall, K. G. J. Chem. Phys. 1997, 106, 9618.

(6) Liu, W.; Peng, D. J. Chem. Phys. 2006, 125, 044102.

(7) Barysz, M.; Sadlej, A. J. J. Chem. Phys. 2002, 116, 2696.

(8) Wahlgren, U.; Schimmelpfennig, B.; Jusuf, S.; Stromsnes,¨ H.; Gropen, O.;

Maron, L. Chem. Phys. Lett. 1998, 287, 525.

(9) Thar, J.; Kirchner, B. J. Chem. Phys. 2009, 130, 124103.

(10) Seino, J.; Nakai, H. J. Chem. Phys. 2012, 136, 244102.

(11) Seino, J.; Nakai, H. J. Chem. Phys. 2012, 137, 144101.

(12) Seino, J.; Nakai, H. J. Chem. Phys. 2013, 139, 034109.

(13) Seino, J.; Nakai, H. Int. J. Quantum Chem. 2015, 115, 253.

(14) Seino, J.; Nakai, H. J. Comput. Chem. Jpn. 2014, 13,1.

(15) Peng, D.; Reiher, M. J. Chem. Phys. 2012, 136, 244108.

(16) Peng, D.; Middendorf, N.; Weigend, F.; Reiher, M. J. Chem. Phys. 2013, 138,

184105.

(17) Phillips, J. C.; Kleinman, L. Phys. Rev. 1959, 116, 287.

(18) Weeks, J. D.; Rice, S. A. J. Chem. Phys. 1968, 49, 2741.

(19) Hamann, D. R.; Schluter,¨ M.; Chiang, C. Phys. Rev. Lett. 1979, 43, 1494.

(20) Hay, P. J.; Wadt, W. R. J. Chem. Phys. 1985, 82, 270.

(21) Vanderbilt, D. Phys. Rev. B 1990, 41, 7892.

113 (22) Lassonen, K.; Pasquarello, A.; Car, R.; Lee, C.; Vanderbilt, D. Phys. Rev. B

1993, 47, 10142.

(23) Figgen, D.; Rauhut, G.; Dolg, M.; Stoll, H. Chem. Phys. 2005, 311, 227.

(24) Sakai, Y.; Miyoshi, E.; Klobukowski, S.; Huzinaga, M. J. Comput. Chem. 1987,

8, 226.

(25) Rakowitz, F.; Marian, C. M.; Seijo, L. J. Chem. Phys. 1999, 111, 10346.

(26) Tsuchiya, T.; Nakajima, T. N.; Hirao, K.; Seijo, L. Chem. Phys. Lett. 2002, 361,

334.

(27) Zeng, T.; Klobukowski, M. J. Chem. Phys. 2009, 130, 204107.

(28) Seino, J.; Tarumi, M.; Nakai, H. Chem. Phys. Lett. 2014, 592, 341.

(29) Huzinaga, S; Cantu, A. A. J. Chem. Phys. 1971, 55, 5543.

(30) Nautiyal, T.; Youn, S.; Kim, K. Phys. Rev. B 2003, 68, 033407.

(31) Nazin, G.; Qiu, X.; Ho, W. Phys. Rev. Lett. 2003, 90, 216110.

(32) Nakatsuji, H.; Hirao, K. J. Chem. Phys. 1978, 68, 2053.

(33) Nakatsuji, H. Chem. Phys. Lett. 1979, 67, 334.

(34) Dunning, Jr., T. H. J. Chem. Phys. 1989, 90, 1007.

(35) Nakatsuji, H. Chem. Phys. 1983, 75, 425.

(36) Noro, T.; Sekiya, M.; Koga, T. Theor. Chem. Acc. 2012, 131, 1124.

(37) Figgen, D.; Peterson, K. A.; Dolg, M.; Stoll, H. J. Chem. Phys. 2009, 130,

164108.

(38) Haas, P.; Tran, F.; Blaha, P. Phys. Rev. B 2009, 79, 085104.

(39) Tatewaki, H.; Koga, T. Chem. Phys. Lett. 2000, 328, 473.

(40) Sekiya, M.; Noro, T.; Osanai, Y.; Koga, T. Theor. Chem. Acc. 2001, 106, 297.

(41) Osanai, Y.; Noro, T.; Miyoshi, E.; Sekiya, M.; Koga, T. J. Chem. Phys. 2004,

120, 6408.

(42) Hehre, W. J.; Ditchfield, R.; Pople, J. A. J. Chem. Phys. 1972, 56, 2257.

(43) Dill, J. D.; Pople, J. A. J. Chem. Phys. 1975, 62, 2921.

114 (44) Tsuchiya, T.; Abe, M.; Nakajima, T.; Hirao, K. J. Chem. Phys. 2001, 115, 4463.

(45) Frisch, M. J. et al. Gaussian 09 Revision A.2., Gaussian Inc. Wallingford CT

2009.

(46) Bakke, A. A.; Chen, A. W.; Jolly, W. L. J. Electron Spectrosc. Relat. Phenom.

1980, 20, 333.

(47) Crotti, C.; Farnetti, E.; Celestino, T.; Stener, M.; Fontana, S. Organometallics

2004, 23, 5219.

115

Chapter 7 Relativistic effect on enthalpy of formation for transition metal complexes

7.1 Introduction

Relativistic effects play an essential role in the accurate description of chemical phenomena in molecules containing transition metals. The relativistic effects are mainly divided into SF and SD relativistic effects. In general, the SF relativistic effect contracts and expands s/p and d/f orbitals, respectively. The SD relativis- tic effect splits orbitals according to their angular momentum numbers such as ff p1/2 and p3/2 and d5/2 and d7/2. These changes a ect geometries and reactivities of molecules. For describing the SF relativistic effect efficiently, PP methods are widely utilized for transition metal calculations because PPs decrease the compu- tational cost of electronic state calculations because of introduced core potentials.

Core potentials replace the effect of core orbitals in molecules with potentials fit- ted to both of AE atomic valence orbitals and energies or AE atomic transitions in valence space1. Several ECPs have been proposed such as LANL2DZ2–4, SBKJC5, and SDD6–9. The model core potential and the ab initio model potential are alter- natives to core potentials because the basic equation for constructing the potential is different from that of ECPs10. The FCP11 method is another method that directly uses atomic orbitals for describing core potentials; the computational cost is still

117 higher than that for other PPs but less than AE treatments.

The thermochemical properties are difficult to precisely determine experi- mentally at the equivalent accuracy among all transition metal complexes. Many computational studies have been performed on this topic using the DFT and post-HF methods12–28, time-dependent DFT29,30, Gaussian-4 method31, and the correlation consistent composite method32–38. A theoretically accurate calcula- tion is accomplished for first and second transition metal complexes. The MAE is less than 3 kcal/mol. However, realizing the sufficient accuracy in thermochemical properties for the third transition metal complexes is difficult by means of theoret- ical calculations owing to many effects and phenomena such as low-lying excited states, the spin-orbit interaction, and the electron correlation. While relativistic effects are considered in these transition metal calculations, the accuracy depends on the PPs. The reasons for the accuracy differences between PPs are varied: a reference relativistic method for constructing potentials, optimized geometry of molecules, and need to be relativistic effects taken into account in calculations.

Note that PPs give reliable and accurate results compared with all-electron rela- tivistic methods when a core size is carefully chosen39–41.

Seino and co-workers have developed and employed an accurate and effi- cient relativistic method based on the IODKH method with LUT since 2010 not only for energy calculations42,43 but also molecular properties such as geometry optimizations44,45. These methods consider enough relativistic effect to calculate all the element at the same accuracy. In addition, FCP fixes the core orbitals ar- bitrarily, meaning that individual core orbital contributions to a target property can be examined. This chapter focuses on the relativistic effect on the enthalpy of formation for transition metal molecules using the AE, ECP, and FCP methods.

Section 7.2 provides computational details. In Section 7.3, the accuracy of WFT and DFT, and relativistic effect for the enthalpy of formation of transition metal

118 complexes are discussed. Finally, concluding remarks of this chapter are provided in Section 7.4.

7.2 Computational details

Twelve diatomic molecules including those composed of first to third transition metals were investigated to assess the accuracy of the AE method: TiO, VO, NiF,

YO, ZrO, ZrF2, MoO, AgH, TaO, WO, WF, and HgH, which are assigned the code numbers, S1 to S12, respectively. The spin multiplicities of these diatomic molecules set to the most stable state: singlet for ZrO, AgH, and WO, doublet for NiF, YO, TaO, WF, and HgH, triplet for TiO, ZrF2, and MoO, and quartet for VO. 23 transition-metal complexes were investigated, which are in the sub- set of transition-metal complexes used in previous assessments by Cundari et

46 47 al. and Mori et al. :TiF4, TiOCl2, VO(OEt)3, Cr(CO)6, Mn(CO)5Br, Fe(CO)5,

Ferrocene, Co(CO)4H, Ni(CO)4, ZnMe2, ZrBr4, ZrF4, NbCl5, Mo(CO)5py, MoBr4,

RuO4, cis-PdCl2(NCMe)2, CdEt2, TaCl5, WOCl4, OsO4, cis-Platin, and HgMe2, which are assigned the code numbers, C1 to C23, respectively.

The gas-phase enthalpies of formation ΔH were calculated as follows:

+ + ΔH =ΔEelec +ΔEvib rot trans +ΔEZPE, (7.1)

+ + where ΔEelec is electronic energy, ΔEvib rot trans is sum of the vibrational, rotational, and translational energies, and ΔEZPE is the zero-point energy. Experimental atomization enthalpy was adopted. ΔEelec was obtained at WFT for HF, MP2,

CCSD, and CCSD(T). CA was employed for estimating electronic energies of transition-metal molecules using the formula as follows:

CA = / + / − / , E E[MP2 QZ-QZ] (E[CCSD(T) X1Z-X2Z] E[MP2 X1Z-X2Z]) (7.2) where E[MP2/QZ-QZ] is the total energy at MP2 with quadruple-ζ basis sets for transition metals (shown in former QZ) and quadruple-ζ basis sets for other

119 / / elements (shown in later QZ), E[CCSD(T) X1Z-X2Z] and E[MP2 X1Z-X2Z] are the total energy at CCSD(T) and at MP2, respectively. Here, X1 and X2 mean a cardinal number of the basis sets, D for 2, T for 3, and Q for 4. Furthermore, in order to examine the functional dependence on DFT, five types of the DFT functional were employed with quadruple-ζ quality of basis sets for all elements: BLYP, B3LYP,

M06, ωB97XD, and LC-ωPBE.

The effect of spin-orbit interaction ΔESO was calculated as the difference in the total energy of SD and SF methods at the MP2 level using triple-ζ quality of basis sets,

ΔESO =ΔE[MP2/uTZ-TZ] − ΔE[MP2/uTZ-TZ]. (7.3)

Here uTZ denotes a triple-ζ quality basis set in uncontracted form was used.

The geometry was optimized with the DFT with B3LYP-D functional using

DKH3-Gen-TK-NOSeC-V-TZP for transition metals and cc-pVDZ-DK for other elements. The thermal corrections and zero-point energy at 298.15 K were taken into account for the gas-phase enthalpy of formation. All of the AE energies were obtained by the fourth-order DKH Hamiltonian with spin-orbit corrections using a quadruple-ζ quality of the same basis set family as used in the geometry optimization unless otherwise stated.

The different levels of relativistic Hamiltonians were utilized for evaluating the higher-order relativistic effect in the enthalpy of formation: NR, RESC, DKH1,

DKH2, and DKH3 Hamiltonians. Three types of pseudopotentials were adopted for comparison with AE: LANL2DZ, SBKJC, and SDD. The standard accompa- nying basis sets were adopted for LANL2DZ and SBKJC; cc-pVTZ-PP basis sets were adopted for SDD. The contribution of frozen-core orbitals to the enthalpy of formation was estimated by the FCP method.

120 7.3 Results and discussion 7.3.1 Accuracy of WFT

This subsection examines the accuracy of WFT for determining the enthalpy of formation of transition metal diatomic molecules. The left panel of Figure

7.1 shows the deviation of enthalpy of formation in transition metal diatomic molecules from experimental values at the MP2, CCSD, and CCSD(T) levels of theory. The horizontal axis indicates the code number. The vertical axis indicates the deviation from the experimental values.

The deviation from experimental value in MP2 was larger that those in CCSD and CCSD(T); the MAEs were 9.65, 7.46, and 3.21 kcal/mol, respectively. CCSD overestimated the enthalpy of formation in all cases except for AgH and HgH. The largest deviation was 24.50 kcal/mol in MoO. Furthermore in the case of CCSD(T), the largest deviation is less than 10 kcal/mol, which is in good agreement with experimental values among the three methods. These results indicate that the electronic correlation energy beyond doubles is important for accurate estimation of the enthalpy of formation.

The right panel of Figure 7.1 shows the basis set dependence of the enthalpy of formation at the composite approach. The horizontal axis indicates the code number while the vertical axis indicates the deviation from the experimental values. The difference in CA(TZ-DZ), CA(QZ-DZ), and CA(QZ-TZ) is the basis set for transition metals and other elements. TZ-DZ means triple-ζ quality of the basis set for transition metals and double-ζ quality of the basis set for other elements.

Similarly,QZ-DZ and QZ-TZ respectively denote quadruple- and double-ζ quality of the basis set and quadruple- and triple-ζ quality of the basis set for transition metals and other elements. For comparison, CCSD(T) result with QZ-QZ basis sets is shown in the right panel of Figure 7.1.

121 30 15 MP2 (a) (b) CCSD 10 20 CCSD(T)

5 10 0 0 -5 CA(TZ-DZ) -10 -10 CA(TZ-TZ) CA(QZ-DZ) CCSD(T)(QZ-QZ) Deviation from experimental vlaues [kcal/mol] Deviation from experimental values [kcal/mol] -20 -15 S1S4 S9 S12 S1 S4 S9 S12 Code number Code number Figure 7.1 Deviation from experimental values (kcal/mol) of enthalpy of formation for transition metal diatomic molecules (a) at the MP2, CCSD, and CCSD(T) levels of theory and (b) for the CA method with different combinations of basis sets.

Table 7.1 MaxE and MAE (kcal/mol) of enthalpy of formation for transition metal complexes at HF, MP2, and CA methods. HF MP2 CA MaxE 630.59 213.89 33.05 MAE 247.42 82.46 14.89

The basis set dependence is small in the CA method. MAEs are 4.36, 3.81, and

3.41 kcal/mol for CA(TZ-DZ), CA(QA-DZ), and CA(QZ-TZ), respectively. Con- sidering the balance of accuracy and efficiency, CA(TZ-DZ) is the most reasonable method.

Table 7.1 shows the MaxE and MAE of enthalpy of formation for 23 transition metal complexes at the HF, MP2, and CA levels. Both of MaxE and MAE were improved as the electron correlation effect is taken into account: 630.6 and 247.4 kcal/mol for HF, 213.9 and 82.5 kcal/mol for MP2, and 33.1 and 14.9 kcal/mol for

CA. Similar to the transition metal diatomic molecules, the electron correlation effect from a perturbation triple excitation, (T), was significant in the enthalpy of formation.

122 Table 7.2 MaxE and MAE (kcal/mol) of enthalpy of formation for transition metal complexes at the DFT with BLYP, B3LYP, M06, ωB97XD, and LC-ωPBE functionals. BLYP B3LYP M06 ωB97XD LC-ωPBE diatomic molecules MaxE 16.016.616.634.261.0 MAE 6.27.86.111.011.9 metal complexes MaxE 35.066.634.137.154.2 MAE 12.225.212.816.623.1

7.3.2 Functional dependence

This subsection investigates the functional dependence of the enthalpy of for- mation for the transition metal molecules. Table 7.2 shows the MaxE and MAE of six functionals of the enthalpy of formation for transition metal molecules at

DFT with BLYP, B3LYP, M06, ωB97XD, and LC-ωPBE. For diatomic molecules, the pure and hybrid functionals, BLYP, B3LYP, and M06, are in good agreement with the experimental values while range separated functionals, ωB97XD and

LC-ωPBE, have relatively large deviations: 11.0 and 11.9 kcal/mol for ωB97XD and LC-ωPBE, respectively. On the other hand, for transition metal complexes, the tendency was different from that for the diatomic molecules. BLYP, M06, and

ωB97XD functionals have smaller MaxE and MAE: 35.0 and 12.2 kcal/mol, respec- tively, for BLYP; 34.1 and 12.8 kcal/mol, respectively, for M06; and 37.1 and 16.6 kcal/mol, respectively, for ωB97XD. The B3LYP functional gives the largest MaxE and MAE among these five functionals, 66.6 and 25.2 kcal/mol, indicating that the appropriate description of transition metal complexes is difficult using the same functional that achieved the accurate description of electronic states in light main elements. LC-ωPBE gives the largest MaxE and MAE: 54.2 and 23.1 kcal/mol, respectively. Using the M06 functional, experimental values of the gas-phase enthalpy of formation are reproduced within a 15 kcal/mol deviation.

123 7.3.3 Geometry difference between PP and AE methods

This subsection examines the geometry difference between the PP and AE meth- ods. Figure 7.2 shows the difference in bond lengths between the central metal and the nearest-neighbor atom from those of LUT-IODKH for transition metal complexes obtained by LANL2DZ, SBKJC, and SDD. The horizontal axis shows the types of bonds and code numbers. The vertical axis shows the bond length difference from LUT-IODKH. All three core potentials are in good agreement with those of the AE method in bond length less than 0.05 Å difference except for three bond lengths obtained by LANL2DZ. For SDD, more than 85% of the bond length differences are less than 0.01 Å, which are the best results among these potentials.

For SBKJC, the bond length differences are relatively larger than the other core potentials. For LANL2DZ, three bond lengths show relatively larger differences from those of the AE method: 0.0820, 0.0655, and 0.1050 Å for Zn-C, Cd-C, and

Hg-C, respectively. These elements are in the twelfth group which has the 6s oc- cupied orbitals, indicating that the relativistic contraction of the orbitals may not be sufficiently described. These results indicate that PPs provide more reasonable bond length between heavy and light atoms compared with the AE method.

7.3.4 Effect of the levels of relativistic Hamiltonians

This subsection discusses the effect of the levels of relativistic Hamiltonians. Table

7.3 shows the MAE for the enthalpy of formation obtained by five types of NR and relativistic Hamiltonians. The IODKH enthalpy of formation is used as the reference value. For first transition metals, all Hamiltonians are in good agreement with the IODKH result. All relativistic Hamiltonians give less than

0.3 kcal/mol MAE. MAE is 7.05 kcal/mol without the relativistic effect because the relativistic effect on the first transition metal is relatively small. For the second transition metals, the MAE of NR is more than two orders of magnitude larger

124 0.15 LANL2DZ

SBKJC Hg-C 0.10 SDD Zn-C Cd-C

0.05 Ru-O V-OEt Mn-Br Pd-N Pt-N Pt-Cl Zr-F Zn-Br Mo-C W-Cl Os-O Nb-Cl V-O Mo-Br Ta-Cl W-O Mn-C (ax.) Ti-F Ti-O Ti-Cl Pd-Cl Mn-C (eq.) Cr-C Co-H Co-C (ax.) Co-C (eq.) Fe-C Fe-C Ni-C 0 Difference from LUT-IODKH [Å] -0.05 C1 C11C19 C23 Code number Figure 7.2 Difference in bond length (Å) between central transition metal and nearest neighbor atom from LUT-IODKH obtained by SDD, SBKJC, and LANL2DZ.

Table 7.3 MAE (kcal/mol) for enthalpy of formation obtained by different levels of non- and relativistic Hamiltonians: NR, RESC, DKH1, DKH2, and DKH3. transition metal NR RESC DKH1 DKH2 DKH3 First 7.05 0.05 0.21 0.01 0.00 Second 160.00 6.88 1.98 0.03 0.00 Third 385.58 1.93 5.60 0.32 0.05 All 157.59 2.54 2.03 0.08 0.01 than those of all four relativistic Hamiltonians: 6.88, 1.98, 0.03, and 0.00 kcal/mol, respectively. RESC and DKH1 give larger MAE than DKH2 and DKH3 due to the lack of inclusion of the higher-order relativistic effect. For third transition metals, a similar tendency is shown by NR, RESC, and DKH1 although those of

RESC and DKH1 are comparable. DKH2 and DKH3 have a small but relatively larger MAE than those for first- and second-row transition metals because in third-row transition metals, the higher-order relativistic correction contributes to the enthalpy of formation. These indicate that errors due to the use of difference

Hamiltonians are far smaller compared to all the other approximations made for the lighter elements.

Figure 7.3 shows the contribution of the spin-orbit coupling for the enthalpy

125 30

25 HF MP2 20

15

10 5 0

-5

-10 Effect of spin-orbit interaction [kcal/mol] -15 C1 C11C19 C23 Code number Figure 7.3 Contribution of the spin-orbit interaction (kcal/mol) for the enthalpy of formation at the HF and MP2 levels of theory.

of formation at the HF and MP2 levels of theory. The horizontal axis indicates the code number of transition metal complexes. The vertical axis indicates the contribution of spin-orbit coupling (kcal/mol) in the gas-phase enthalpy of for- mation. At the HF level, the spin-orbit effect in the third-row transition metal is larger than that in the other transition metals, while at the MP2 level, the effect is comparable for all transition metals but different from those of HF. Thus, in third transition metals, the spin-orbit effect significantly contributes to the enthalpy of formation.

7.3.5 Contribution of frozen core orbitals

This subsection examines the contribution of frozen core approximation in the pseudopotential method to the enthalpy of formation. Table 7.4 shows the MaxE and MAE of enthalpy of formation (kcal/mol) obtained by LANL2DZ, SBKJC,

SDD, and FCP methods at the HF and MP2 levels of theory. In the HF method,

126 Table 7.4 MaxE and MAE (kcal/mol) of enthalpy of formation for transition metal complexes at LANL2DZ, SBKJC, SDD, and FCP methods. LANL2DZ SBKJC SDD FCP HF MaxE 303.22 38.14 15.62 6.73 MAE 20.96 9.03 5.36 0.55 MP2 MaxE 50.99 43.08 35.85 53.14 MAE 20.66 16.96 15.03 15.46 the contributions of the frozen core approximation are relatively small; MAEs were 20.96, 9.03, 5.36, and 0.55 kcal/mol for LANL2DZ, SBKJC, SDD, and FCP, respectively. In particular, FCP showed the smallest MAE because the core poten- tials are determined by the IODKH atomic calculation. The resultant deviation is improved by an orbital relaxation technique48.

In the MP2 method, all of MAEs were close to or larger than those of HF: 20.66 kcal/mol for LANL2DZ, 16.96 kcal/mol for SBKJC, 15.03 kcal/mol for SDD, and

15.46 kcal/mol for FCP. This is because the basis sets are smaller than those of AE using core potentials. The resultant virtual orbitals are not adequately described compared to AE.

Figure 7.4 shows the contribution of individual orbitals to the enthalpy of / formations (kcal mol) in TaCl5 and OsO4. The horizontal axis shows the outermost frozen core orbitals. The vertical axis shows the contribution of individual orbitals to the enthalpy of the formation. In third transition metals, the 4f orbital is frozen in most of the PP methods such as SDD, LANL2DZ, and SBKJC.

For both molecules, the inner orbitals have small contributions to the enthalpy of formation because these orbitals rarely participate in the valence bonding, meaning that orbitals maintain their atomic nature sufficiently to satisfy the frozen orbital approximation. On the other hand, 4f, 5s, and 5p contributions are more than two times larger than those of the other orbitals. These orbitals are subvalence and valence orbitals affected by bonding. The frozen core approximation using

127 60

TaCl5 OsO 40 4

20

0

Contribution of individual orbitals [kcal/mol] -20 1s 2s 2p 3s 3p 3d 4s 4p 4d 4f 5s 5p Orbital / Figure 7.4 Orbital contribution for the enthalpy of formations (kcal mol) in TaCl5 and OsO4 at DFT with the M06 functional.

atomic orbitals is no longer a good model in this situation although 4f orbitals are frozen in most of the PP methods except for the small-core Stuttgart PPs.

7.4 Conclusion

In this chapter, the relativistic effect on the enthalpy of formation for transition metal complexes is discussed using the AE and PP methods. Two test sets, 12 di- atomic molecules containing transition metals and 23 transition metal complexes, were used. The WFT results are in good agreement with experimental enthalpy of formation; they show less than 15 kcal/mol deviation considering the electron correlation. The electron correlation at the CCSD(T) level of theory is important for the enthalpy of formation in transition metal molecules. The geometry dif- ference between PP and AE was small. When the molecule contains 4d or 5d transition metals, the higher-order relativistic theory whose effect is included in

128 the second- or higher-order DKH method should be utilized for accurate calcu- lation. The contribution of the spin-orbit effect to the enthalpy of formation of

5d transition metal complexes is more than 10 kcal/mol. The frozen core orbital approximation in PP methods have a relatively large contribution to the enthalpy of formation. These results indicate that understanding the relativistic effect and the range of core orbitals necessary to describe the target reaction and molecular property values is significant for accurate and efficient theoretical prediction.

129 References

(1) Schwerdtfeger, P. Chem. Phys. Chem 2011, 12, 3143.

(2) Hay, P. J.; Wadt, W. R. J. Chem. Phys. 1985, 82, 270.

(3) Hay, P. J.; Wadt, W. R. J. Chem. Phys. 1985, 82, 284.

(4) Hay, P. J.; Wadt, W. R. J. Chem. Phys. 1985, 82, 299.

(5) Stevens, W. J.; Krauss, M.; Basch, H.; Jasien, P. G. Can. J. Chem. 1992, 70, 612.

(6) Dolg, M.; Wedig, U.; Stoll, H.; Preuss, H. J. Chem. Phys. 1987, 86, 866.

(7) Figgen, D.; Rauhut, G.; Dolg, M.; Stoll, H. Chem. Phys. 2005, 311, 227.

(8) Peterson, K. A.; Figgen, D.; Dolg, M.; Stoll, H. J. Chem. Phys. 2007, 126,

124101.

(9) Figgen, D.; Peterson, K. A.; Dolg, M.; Stoll, H. J. Chem. Phys. 2009, 130,

164108.

(10) Huzinaga, S; Cantu, A. A. J. Chem. Phys. 1971, 55, 5543.

(11) Seino, J.; Tarumi, M.; Nakai, H. Chem. Phys. Lett. 2014, 592, 341.

(12) Buzko, V. Y.; Sukhno, I. V.; Buzko, M. B.; Subbotina, J. O. Int. J. Quantum

Chem. 2006, 106, 2236.

(13) Furche, F.; Perdew, J. P. J. Chem. Phys. 2006, 124, 044103.

(14) Riley, K. E.; Merz, Jr., K. M. J. Phys. Chem. A 2007, 111, 6044.

(15) Yang, Y.; Weaver, M. N.; Merz, Jr., K. M. J. Phys. Chem. A 2009, 113, 9843.

(16) Karlicky,´ F.; Otyepka, M. J. Chem. Theory Comput. 2011, 7, 2876.

(17) Hughes, T. F.; Harvey, J. N.; Friesner, R. A. Phys. Chem. Chem. Phys. 2012, 14,

7724.

(18) Jiang, W.; Laury, M. L.; Powell, M.; Wilson, A. K. J. Chem. Theory Comput.

2012, 8, 4102.

(19) Shil, S.; Bhattacharya, D.; Sarkar, S.; Misra, A. J. Phys. Chem. A 2013, 117,

4945.

130 (20) Weaver, M. N.; Merz, Jr., K. M.; Ma, D.; Kim, H. J.; Gagliardi, L. J. Chem.

Theory Comput. 2013, 9, 5277.

(21) Sparta, M.; Jensen, V. R.; Børve, K. J. Mol. Phys. 2013, 111, 1599.

(22) Weymuth, T.; Couzijn, E. P. A.; Chen, P.; Reiher, M. J. Chem. Theory Comput.

2014, 10, 3092.

(23) Matczak, P.; Wojtulewski, S. J. Mol. Model 2015, 21, 41.

(24) Weymuth, T.; Reiher, M. Int. J. Quantum Chem. 2015, 115,9.

(25) Wang, J.; Liu, L.; Wilson, A. K. J. Chem. Phys. A 2016, 120, 737.

(26) Gurtu, S.; Rai, S.; Ehara, M.; Priyakumar, U. D. Theor. Chem. Acc. 2016, 135,

93.

(27) Waitt, C.; Ferrara, N. M.; Eshuis, H. J. Chem. Theory Comput. 2016, 12, 1542.

(28) Skrˇ ´ıba, A.; Jasˇ´ık, J.; Andris, E.; Roithova,´ J. Organometallics 2016, 35, 990.

(29) Latouche, C.; Skouteris, D.; Palazzetti, F.; Barone, V. J. Chem. Theory Comput.

2015, 11, 3281.

(30) Alipour, M. Theor. Chem. Acc. 2016, 135, 67.

(31) Mayhall, N. J.; Raghavachari, K.; Redfern, P. C.; Curtiss, L. A. J. Phys. Chem.

A 2009, 113, 5170.

(32) DeYonker, N. J.; Peterson, K. A.; Steyl, G.; Wilson, A. K.; Cundari, T. R. J.

Phys. Chem. A 2007, 111, 11269.

(33) Williams, T. G.; Wilson, A. K. J. Chem. Phys. 2008, 129, 054198.

(34) DeYonker, N. J.; Williams, T. G.; Imel, A. E.; Cundari, T. R.; Wilson, A. K. J.

Chem. Phys 2009, 131, 024106.

(35) Laury, M. L.; DeYonker, N. J.; Jiang, W.; Wilson, A. K. J. Chem. Phys. 2011,

135, 214103.

(36) Jiang, W.; DeYonker, N. J.; Determan, J. J.; Wilson, A. K. J. Chem. Phys. A

2012, 116, 870.

(37) Laury, M. L.; Wilson, A. K. J. Chem. Theory Comput. 2013, 9, 3939.

131 (38) Manivasagam, S.; Laury, M. L.; Wilson, A. K. J. Chem. Phys. A 2015, 119, 6867.

(39) Schwerdtfeger, P.; Fischer, T.; Dolg, M.; Igel-Mann, G.; Nicklass, A.; Stoll, H.;

Haaland, A. J. Chem. Phys. 1995, 102, 2050.

(40) Leininger, T.; Nicklass, A.; Stoll, H.; Dolg, M.; Schwerdtfeger, P. J. Chem.

Phys. 1996, 102, 1052.

(41) Schwerdtfeger, P.; Brown, J. R.; Laerdahl, J. K.; Stoll, H. J. Chem. Phys. 2000,

113, 7110.

(42) Seino, J.; Nakai, H. J. Chem. Phys. 2012, 136, 244102.

(43) Seino, J.; Nakai, H. J. Chem. Phys. 2012, 137, 144101.

(44) Nakajima, Y.; Seino, J.; Nakai, H. J. Chem. Phys. 2013, 139, 244107.

(45) Nakajima, Y.; Seino, J.; Nakai, H. Chem. Phys. Lett. 2016, 663, 97.

(46) Cundari, T. R.; Leza, H. A. R.; Grimes, T.; Steyl, G.; Waters, A.; Wilson, A. K.

Chem. Phys. Lett. 2005, 401, 58.

(47) Mori, H.; Zeng, T.; Klobukowski, M. Chem. Phys. Lett. 2012, 521, 150.

(48) Nakajima, Y.; Seino, J.; Nakai, H. J. Chem. Theory Comput. 2016, 15, 68.

132 Chapter 8 General Conclusion

In this thesis, the accurate and efficient analytical energy gradient for IODKH with/without LUT was derived, assessed, extended to the SD formalism, and implemented in GAMESS. Furthermore, an efficient PP method, based on FCP, was extended to improve accuracy such that it approaches AE. In Chapter 3 the analytical gradient of IODKH, in the SF formalism, was derived and combined with the LUT scheme to extend the applicability of this method to large-scale molecules while reducing the computational cost associated with relativistic uni- tary transformations in energy and gradient calculations. Based on the derivation of LUT-IODKH, Chapter 4 implemented the SD formalism using basic GHF the- ory, because the spinor is represented by a linear combination of α- and β-spins that are capable of describing spin mixing through the explicit spin operator included in the SD Hamiltonian. This implementation also required all spin- operator-related values to be expressed in complex number space. In Chapter 5,

LUT-SF-IODKH energy and gradient methods were implemented in the GAMESS quantum chemical package, one of the de facto, standard, quantum chemical pro- grams, in order to combine the LUT-SF-IODKH method with many of the func- tionalities in GAMESS, such as correlation methods, excited-state calculations, and large-scale molecular methods, such as the DC method. Chapter 6 improved the FCP method by relaxing the core orbitals in molecules through the introduc- tion of the frozen-valence potential formed by the valence molecular orbitals.

133 Using all of the LUT-IODKH and FCP methods developed in this work, Chapter

7 investigated the relativistic effect on the gas-phase enthalpies of formation of transition metal complexes.

The results in Chapters 3, 4, and 5 indicate that the LUT-IODKH method is highly accurate and highly efficient, not only in SF, but also in the SD formalism.

Bond lengths of diatomic molecules were in reasonable agreement with those obtained by 4c methods at the HF level, as well as experimental bond lengths, and those obtained using DFT and correlation methods. Numerical assessment of multi-dimensional silver clusters confirmed a quasi-linear scaling of the rel- ativistic unitary transformation in all the cases tested. Moreover, for coding purposes, due to the nature of relativistic Gaussian basis functions that possess large exponent values, 2c relativistic methods such as LUT-IODKH require a vari- able precision format, that includes double or quadruple precision. In addition,

LUT-IOKH was combined with the linear-scaling DC technique; an overall linear- scaling calculation confirmed that the computational cost was equivalent to NR, with small prefactors.

Chapter 6 confirmed that the relaxation of core orbitals worked reasonably well from the viewpoints of accuracy and efficiency, without loss of wavefunction information. The computational costs of FCP-CR calculations on gold clusters of up to 86 atoms were less than half of those obtained by AE methods. In addition,

FCP-CR had improved accuracy when compared with conventional FCP, in both total and orbital energies. The relaxed core orbitals of FCP-CR, in combination with SAC-CI, provided good core properties such as CIEs. Furthermore, itera- tive core-valence relaxation solutions showed that the first iteration considerably improved core orbital descriptions.

Chapter 7 suggested that the inclusion of relativistic effects is required in order to provide accurate gas-phase enthalpies of the formation, even if only first-row

134 transition metals are included; the contribution was more than 7 kcal/mol. In terms of geometries, frequently used PPs such as LANL2DZ and SDD provided bond lengths that were in good agreement with those calculated by AE, differing by less than 0.05 Å in many cases, although SBKJC provided relatively large deviations.

In addition, the orbitals usually frozen in PPs were shown to contribute non- negligibly to the enthalpies of formation.

This study broadened the applicability of LUT-IODKH to basic molecular property calculations at both SF and SD levels, and combined LUT-IODKH with

FCP-CR to provide a highly efficient, quasi-all-electron method, without loss of accuracy. In addition, relativistic effects are sufficiently included compared to the 4c relativistic method. Relativistic quantum chemistry can now advance from the theoretical development phase, to one in which it can assist realistic applications, including the treatment of complicated systems containing heavy elements in various spin-states, while providing detailed assessments of special characteristics such as electronic states.

The author hopes that this thesis will accelerate the development of novel materials through the predictive application of the methodology described.

135

Acknowledgments

तͷ͝ࢦಋͷ΋ͱͰߦΘΕ·ͨ͠ɻڭ͸ૣҴాେֶԽֶɾੜ໋ԽֶՊதҪߒາڀຊݚ

ͷਐΊํ΍ߟ͑ํɼΘ͔Γ΍͍͢ൃදͷڀࣨ഑ଐҎདྷͷ 6 ೥ؒɼݚڀतʹ͸ݚڭதҪ

௖͖·ͨ͠ɻͯ͑ڭʹऀͱͯ͠ͷ͋ΓํͳͲΛ೤৺ڀ๏ɼ࿦จࣥචɼݚํ

͋ʹतʹ͸ɼຊ࿦จͷ৹ࠪڭतɼҪଜߟฏڭɼಉԽֶɾੜ໋ԽֶՊݹ઒ߦ෉ͨ·

ͨΓ෭ࠪΛ຿Ίͯ௖͖·ͨ͠ɻઌੜํʹ͸ɼ͝ଟ๩ͳ͕Β΋ஸೡͳࠪಡΛͯ͠௖͖ɼ

க͠·͢ɻँײॏͳ͝ࢦಋΛࣀΓ·ͨ͠ɻ৺ΑΓو

ࣨͷڀΛਐΊΔʹ͋ͨΓɼଟ͘ͷਓͨͪʹ͝ॿݴΛ௖͖·ͨ͠ɻݚڀຊ࿦จͷݚ

ελοϑͷํʑɼ·ͨઌഐํʹ͸ɼ೔ࠒΑΓେม͓ੈ࿩ʹͳΓ·ͨ͠ɻಛʹɼຊ࿦จͷ

શମʹΘͨΓ·ͯ͠ɼLUT-IODKH ๏͓Αͼ FCP ๏ͷ։ൃऀͰ͋Γɼॴଐ൝ͷϦʔ

·಺༰΍ϓϩάϥϜ࣮૷ɼֶձͰͷൃදʹࢸΔڀμʔͰ͋ͬͨਗ਼໺३࢘ത࢜ΑΓɼݚ

ɼ੺ڭतɼ٠஑ಹ໌ത࢜ɼখྛਖ਼ਓॿڭͰɼ༷ʑͳ͝ࢦಋΛ௖͖·ͨ͠ɻ·ͨɼࠓଜয়

഼߁߂ത࢜ɼੴ઒ರ೭ത࢜ɼԦᜅത࢜ɼ੢ଜ޷࢙തेޒ஌ࢠത࢜ɼ੢ᖒ޺ߊത࢜ɼؒ

க͠·͢ɻँײʹ࢜ɼখ໺७Ұത࢜ɼେӽণथത࢜ɼ٢઒෢࢘ത࢜ɼपݐ඾ത࢜ͷօ༷

ത࢜՝ఔʹਐֶ͠ɼͱ͖ʹ͸ܹ͍ٞ͠࿦Λઓʹڞ඙म࢜͸ɼڡੜͰ͋Δத໺ظಉ

Θͤͳ͕Β΋ɼ੾᛭ୖຏ͠ͳ͕Β੒௕͢Δ͜ͱ͕Ͱ͖·ͨ͠ɻ༏लͳޙഐॾ܅ʹ΋ܙ

கँײΕɼಛʹ଎ਫխੜम࢜ʹ͸ɼެࢲʹΘͨΓେม͓ੈ࿩ʹͳΓ·ͨ͠ɻ৺͔Β·

͠·͢ɻ

࠷ޙʹɼࢲΛܦࡁత·ͨਫ਼ਆతʹࢧ͑ଓ͚ͯ͘Ε·ͨ͠Ո଒ɼಛʹɼ෕ɼल࣏ͱɼ

க͠·͢ɻँײʹࢠܙ฼ɼ஌

137

List of Achievements

Original Articles

1. “Relativistic effect on enthalpy of formation for transition-metal complexes”

Y. Nakajima, J. Seino, and H. Nakai, Chem. Phys. Lett. in press.

2. “Relativistic Frozen Core Potential Scheme with Relaxation of Core Electrons”

Y. Nakajima, J. Seino, M. Hayami, and H. Nakai, Chem. Phys. Lett. 663,97

(2016).

3. “Implementation of efficient two-component relativistic method using local

unitary transformation to GAMESS program”

Y. Nakajima, J. Seino, and H. Nakai, J. Comput. Chem. Japan 15, 68 (2016).

4. “Implementation of Analytical Energy Gradient of Spin-Dependent General

Hartree–Fock Method Based on the Infinite-order Douglas–Kroll–Hess Rela-

tivistic Hamiltonian with Local Unitary Transformation”

Y. Nakajima, J. Seino, and H. Nakai, J. Chem. Theory and Comput. 12, 2181

(2016).

5. “Analytical Energy Gradient Based on Spin-Free Infinite-Order Douglas–Kroll–

Hess Method with Local Unitary Transformation”

Y. Nakajima, J. Seino, and H. Nakai, J. Chem. Phys. 139, 244107, (2013).

139 Conference Presentation International Conference

1. “Implementation of efficient infinite-order two-component relativistic scheme

into GAMESS”

Y.Nakajima, J. Seino, and H. Nakai, The Seventh Asia-Pacific Conference of Theoret-

ical and (APCTCC7), SC2, Kaohsiung (Taiwan), January

2016 (Oral).

2. “Geometries and Molecular properties of heavy main-group molecules based

on two-component relativistic scheme”

Y. Nakajima, J. Seino, and H. Nakai, The International Chemical Congress of Pacific

Basin Societies (Pacifichem 2015), PIIb-23, Honolulu, Hawaii (USA), December

2015 (Poster).

3. “Analytical gradient for spin-dependent infinite-order Douglas–Kroll–Hess

method with local unitary transformation”

Y. Nakajima, J. Seino, and H. Nakai, 11th International Conference on Relativistic

Effects in Heavy-Element Chemistry and Physics (REHE-2014), Smolenice Castle

(Slovak Republic), September 2014 (Poster).

4. “Efficient geometry optimization using accurate two-component relativistic

Hamiltonian with local unitary transformation scheme”

Y. Nakajima, J. Seino, and H. Nakai, 5th JCS International Symposium on Theoret-

ical Chemistry, PIIb-23, Nara (Japan), December 2013 (Poster).

5. “Efficient geometry optimization based on infinite-order Douglas-Kroll-Hess

method with local unitary transformation scheme”

140 Y. Nakajima, J. Seino, and H. Nakai, 6th Asian Pacific Conference of Theoretical

and Computational Chemistry, PS5(S-B), Gyeongiu (Korea), July 2013 (Poster).

Domestic Conference

”ॴϢχλϦʔม׵Λ༻͍ͨޮ཰తͳ 2 ੒෼૬ର࿦๏ͷ GAMESS ΁ͷ࣮૷ہ“ .1

,ژ೥ձ, 1O08, ౦قதౢ༟໵, ਗ਼໺३࢘, தҪߒາ, ೔ຊίϯϐϡʔλԽֶձ 2016 य़

2016 ೥ 6 ݄ (ޱ಄ൃද).

”ࢉܭදΛ໢ཏ͢Δޮ཰తͳ૬ର࿦తྔࢠԽֶظप“ .2

.(಄ൃදޱ) ݄ ೥ 11 2015 ,ژதౢ༟໵, ୈ 2 ճిࢠঢ়ଶཧ࿦γϯϙδ΢Ϝ, O02, ౦

”ಓ؇࿨ͷߟྀيౚ݁಺֪ϙςϯγϟϧ๏ͷ֦ுɿ಺֪“ .3

ޱ) ݄ ೥ 9 2015 ,ژதౢ༟໵, ਗ਼໺३࢘ɼதҪߒາɼୈ 9 ճ෼ࢠՊֶ౼࿦ձ, 2E15, ౦

಄ൃද).

”ࢉʹର͢Δ૬ର࿦ޮՌܭભҠۚଐࡨମͷੜ੒Τϯλϧϐʔ“ .4

2014 ,ژதౢ༟໵, ਗ਼໺३࢘, தҪߒາ, ୈ 4 ճ CSJ ԽֶϑΣελ 2014, P8-002, ౦

೥ 10 ݄ (ϙελʔൃද).

”(ࢉ (2ܭॏݩૉԽ߹෺ʹର͢Δ૬ର࿦తߏ଄࠷దԽ“ .5

݄ தౢ༟໵, ਗ਼໺३࢘, தҪߒາ, ୈ 17 ճཧ࿦Խֶ౼࿦ձ, 1P33, ໊ݹ԰, 2014 ೥ 5

(ϙελʔൃද).

”ࢉܭॏݩૉԽ߹෺ʹର͢Δ૬ର࿦తߏ଄࠷దԽ“ .6

౎, 2013 ೥ 9 ݄ (ϙژ ,தౢ༟໵, ਗ਼໺३࢘, தҪߒາ, ୈ 7 ճ෼ࢠՊֶ౼࿦ձ, 4P092

ελʔൃද).

ॴϢχλϦʔม׵Λ༻͍ͨແݶ࣍ Douglas-Kroll ๏ʹΑΔߴ଎ͳߏ଄࠷దԽखہ“ .7

๏ͷ։ൃ (2)”

141 ݄ ೥ 9 2012 ,ژதౢ༟໵, ਗ਼໺३࢘, தҪߒາ, ୈ 6 ճ෼ࢠՊֶ౼࿦ձ, 1P-099, ౦

(ϙελʔൃද).

”ߴਫ਼౓ 2 ੒෼૬ର࿦๏ʹΑΔߏ଄࠷దԽɿݩૉઓུʹ޲͚ͨ৽ͨͳऔΓ૊Έ“ .8

2012 ,ژதౢ༟໵, ਗ਼໺३࢘, தҪߒາ, ୈ 6 ճ෼ࢠՊֶձγϯϙδ΢Ϝ, P0007, ౦

೥ 6 ݄ (ϙελʔൃද).

ॴϢχλϦʔม׵Λ༻͍ͨແݶ࣍ Douglas-Kroll ๏ʹΑΔߴ଎ͳߏ଄࠷దԽखہ“ .9

๏ͷ։ൃ”

தౢ༟໵, ਗ਼໺३࢘, தҪߒາ, ୈ 15 ճཧ࿦Խֶ౼࿦ձ, 2P11, ઋ୆, 2012 ೥ 5 ݄ (ϙ

ελʔൃද).

”ແݶ࣍ Douglas-Kroll ๏ʹΑΔߏ଄࠷దԽख๏ͷ։ൃ“ .10

,ژ೥ձ, 1P07, ౦قதౢ༟໵, ਗ਼໺३࢘, தҪߒາ, ೔ຊίϯϐϡʔλԽֶձ 2012 य़

2012 ೥ 5 ݄ (ϙελʔൃද).

͘ղੳతΤωϧΪʔඍ෼๏ͷ։ൃͱͦͷߴ଎ͮجʹແݶ࣍ Douglas-Kroll ม׵๏“ .11

Խ”

೥ձ, 1A4-33, ਆಸ઒, 2012 ೥قதౢ༟໵, ਗ਼໺३࢘, தҪߒາ, ೔ຊԽֶձୈ 92 य़

3 ݄ (ϙελʔൃද).

Awards

1. Superior Oral Award

“Implementation of efficient infinite-order two-component relativistic scheme

into GAMESS”

The Seventh Asia-Pacific Conference of Theoretical and Computational Chem-

istry, 2016 ೥ 1 ݄.

142 2. ༏लϙελʔ৆

”(ࢉ (2ܭॏݩૉԽ߹෺ʹର͢Δ૬ର࿦తߏ଄࠷దԽ“

.݄ ୈ 17 ճཧ࿦Խֶ౼࿦ձ, 2014 ೥ 6

ୈ 3 ճؔࠜ٢࿠৆ .3

”༻ࢉʹର͢Δղੳతඍ෼๏ͷ։ൃͱͦͷԠܭ૬ର࿦ణྔࢠԽֶ“

.݄ Ҵాେֶઌਐཧ޻ֶ෦Խֶɾੜ໋ԽֶՊҴԽձ, 2014 ೥ 3ૣ

143