08 Chapter 1

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 82

AB INITIO METHODS IN

COMPUTATIONAL
QUANTUM CHEMISTRY

Aneesh. M.H “ A theoretical study on the regioselectivity of electrophilic


reactions of heterosubstituted allyl systems ” Thesis. Department of Chemistry,
University of Calicut, 2012
CHAPTER I

AB INITIO METHODS IN COMPUTATIONAL


QUANTUM CHEMISTRY

Contents
1.1 Introduction to Computational Chemistry
1.2 Molecular Mechanics/Molecular Dynamics Methods
1.3 Electronic Structure Methods
1.4 An Overview of ab initio Methods
1.4.1 Hartree-Fock (HF) method
1.4.2 The HF limit and the concept of electron correlation
1.4.3 Post-HF methods
1.4.3.1 Configuration interaction methods
1.4.3.2 Perturbative theories
1.4.3.3 Coupled cluster methods
1.4.3.4 Multi-configuration (MC) methods
1.4.3.5 Composite methods
1.4.3.6 Methods for excited states
1.4.3.7 Multirefernece methods
1.4.3.8 r12 methods
1.5 Quantum Mechanics – Molecular Mechanics (QM-MM) Methods
1.6 Semi Empirical Methods
1.7 Density Functional Theory (DFT)
1.8 Basis Sets
1.8.1 Exponential type orbitals
1.8.2 Gaussian type orbitals (GTO)
1.8.2.1 Even - tempered basis sets
1.8.2.2 Well - tempered basis sets
1.8.2.3 Universal basis sets
1.8.2.4 Geometrical basis sets
1.8.2.5 Contracted basis sets
1.8.2.6 Segmented contractions
1.8.2.7 Polarization functions
1.8.2.8 Diffuse functions
1.8.2.9 STO-nG basis sets
1.8.2.10 Pople – style basis sets
1.8.2.11 General contractions
1.8.3 Basis set super position error (BSSE)
1.8.4 Two electron basis functions
1.9 Model Chemistry
Appendix 1: Atomic Units
REFERENCES

0
CHAPTER I
AB INITIO METHODS IN COMPUTATIONAL
QUANTUM CHEMISTRY

1.1 Introduction to Computational Chemistry

Although much of its discovery process is descriptive and qualitative,


chemistry is fundamentally a quantitative science. It serves a wide range of
human needs, activities and concerns. Chemistry attained the status of a
quantitative science through the development of one major discipline in
science, namely, quantum mechanics. The basic principles of quantum
mechanics were formulated in the 1920s through the works of Heisenberg,
Born, Schrödinger, Dirac and many others. Though, in principle, the
equations of quantum mechanics could be applied to any system to produce
exact quantitative results, in practice, it was not so. Exact solutions of the
equations were possible only for certain one electron systems. Even with the
advent of the so called approximation methods the situation didn’t improve
dramatically due to the formidable mathematics involved in solving the
equations. This is quite evident from the famous quote made by Dirac in
1928.

“The fundamental laws necessary for the mathematical treatment of large


parts of Physics and the whole of Chemistry are thus fully known, and the
difficulty lies only in the fact that application of these laws leads to equations
that are too complex to be solved.”

A dramatic leap from this almost stagnant situation was made possible
through the advent of digital computers in the 1950s. Efficient algorithms
were written which helped to solve the formidable mathematical equations
with relative ease. The benefits were evident not only in solving equations of
quantum mechanics but also in solving the equations of classical mechanics.

1
This fusion between radically different disciplines paved the way for a new
branch in Chemistry, Computational Chemistry. Since then computational
science in general has made and continues to make its rapid expansion, both
in terms of hardware and software technologies. Every step forward in
computational science has its beneficial impact on Chemistry. Thus,
computational chemistry can be broadly defined as the branch of chemistry
that uses computers to generate information that is complementary to
experimental data on the structures, properties and reactions of substances.

Computational chemistry has also sought to devise and to implement


quantitative algorithms for organizing massive amounts of data from the
laboratory and for predicting the course and extent of chemical phenomena in
situations that are difficult or even impossible to observe directly. Over the
last 50 years, methods have evolved from those that were used to study 1-and
2-atom systems in 1928, through those that were used to study 2- to 5-atom
systems in 1970, to the present programs that produce useful quantitative
results for molecules with up to 10 to 20 atoms. With cruder models of the
atom (for instance, simpler approximations as in molecular mechanics) it is
now possible to model biological molecules with thousands of atoms. As a
result of the revolutionary expansion in the breadth and capability of
theoretical and computational chemistry in recent decades, these fields have
acquired the power to resolve pressing problems both of a fundamental
scientific character and of clearly practical interest.1 Computational chemistry
also underpins rational drug design, contributes to the selection and synthesis
of new compounds and guides the design of catalysts.

Computational Chemistry has two broad areas: Molecular mechanics


/molecular dynamics methods which use the principles of classical
mechanics and electronic structure methods which use the principles of
quantum mechanics.

2
1.2 Molecular Mechanics/ Molecular Dynamics Methods

Molecular mechanics and molecular dynamics (MM/MD) refer to methods


for computing certain molecular properties, particularly molecular structure
and relative energy. They both typically use fairly simple potential energy
functions that are derived from classical mechanics (e.g., a parabolic function
to calculate the energy required to stretch or to compress a chemical bond). In
addition, they both rely on parameters that are derived either from experiment
(e.g., infrared spectroscopy and X-ray crystallography) or from quantum
mechanics-based calculations (e.g., high-level ab initio molecular orbital
calculations). A collection of potential energy functions and the associated
parameters that are employed for molecular mechanics/ molecular dynamics
calculations is frequently referred to as a "force field"; thus, calculations that
utilize the MM/MD approach are often referred to as empirical force field
calculations.

The molecular mechanics method is generally employed to compute the


relative energies of different geometries (conformations) of the same
molecule that arise from rotations about chemical bonds as well as relative
energies of intermolecular complexes. Often, energy minima are sought; thus,
the molecular mechanics method is frequently coupled with optimization
procedures. On the other hand, in the molecular dynamics method, Newton's
equations of motion are solved by using the gradient of the above-mentioned
potential energy function (force field) to compute the dynamic trajectory of a
molecule or of an ensemble of molecules. Both the MM and MD methods
have found widespread use in the modeling of bio-molecules, for which
quantum mechanical calculations are simply not practical due to the
overwhelming number of particles involved.

3
1.3 Electronic Structure Methods

These methods which use the principles of quantum mechanics can be


classified into two categories: ab initio methods and semi empirical (SE)
methods.

1.4 An Overview of ab initio Methods

The term ab initio comes from the Latin words for “from the beginning”. Ab
initio methods involve quantum mechanical calculations which are derived
directly from theoretical principles (only mathematical approximations are
involved), with no inclusion of experimental data. In this section, we attempt
to highlight the basic principles involved in the different ab initio methods
qualitatively without discussing much of the mathematical details. Nearly all
ab initio quantum chemical methods attempt to solve the eigen value problem
defined by the time – independent Schrödinger equation,

ĤΨ = EΨ (1.1)

where, Ψ is the total molecular wavefunction (a function of electronic and


nuclear coordinates) and Ĥ is the molecular Hamiltonian. For an ‘n’ electron
molecule having ‘M’ nuclei, Ĥ in SI units is given by equation 1.2a

−ℏ −ℏ −
Ĥ= ∇ + ∇ + +
2M 2 4 4 !
!"

#
+ (%. &')
4 $ #
#"

Ĥ in atomic units is given in equation 1.2b (see appendix 1 for details).


In the remainder of the thesis, atomic units (a.u.) are used.
*
−∇ −∇ − 1
(%. &,)
( )
Ĥ= + + + +
2M 2 ! $()
!" )>(

4
The molecular Hamiltonian is formidable enough to strike terror in the heart
of any quantum chemist. Fortunately, there exists a highly accurate
simplifying approximation of neglecting the kinetic Hamiltonian of nuclei
(the first sum of terms in equation 1.2) in comparison with the kinetic
Hamiltonian of the electrons - the Born-Oppenheimer approximation. 2 After
applying the above approximation equation 1.1 can be written as

. / + 011 )Ψ / = 3Ψ
(- / (1.3)

. / and 011 are given by equations 1.4 and 1.5


where, -

−∇ − 1
./=
- + + (%. 5)
2 !
!"

#
011 = (%. 6)
$ #
#"

E gives the total energy including nuclear repulsion under Born-Oppenheimer


approximation. The variables in equation 1.3 are the electronic coordinates.
The inter nuclear distances, $ # , in equation 1.3 are not treated as variables
but are kept fixed at some constant values (this is referred to as fixed nuclear
framework), while the electronic motions are calculated using equation 1.3.
Of course, there is infinite number of such fixed nuclear configurations each
with a constant 011 and for each of them we may solve the electronic
Schrödinger equation 1.3 to get a set of electronic wavefunctions and
corresponding electronic energies. Thus, the electronic wavefunctions and
energies depend parametrically on the nuclear configuration defined by 011 .
If 011 is omitted from 1.3, there is not going to be any change in the value of
Ψ / (It can be proved that the omission of a constant term, C, from the
Hamiltonian does not affect the wavefunctions and simply decreases each
energy eigen value by C). Thus if 011 is omitted, 1.3 becomes,
. /Ψ / = 3 /Ψ
- / (1.6)

5
such that
3 = 3 / + 011 ( 1.7
1.7 )
The Ψ / , thus obtained by solving equation 1.6 is taken to be the exact
wavefunction for practical purposes even though the equation involves the
Born-Oppenheimer approximation (keeping nuclei fixed in space with respect
to the electrons) and neglects all relativistic effects. The inter-electronic
repulsion terms, ‘rij’s, in the Hamiltonian (see equation 1.4) make exact
solution of the electronic Schrödinger equation impossible for many electron
systems.

1.4.1 Hartree - Fock (HF) method


The HF method occupies a central position as far as all ab initio techniques
are considered. All the other electronic structure methods (including the SE
methods and Density Functional Theory (DFT) based methods) stem from the
HF method for one or other reasons. In the HF approximation, the electronic
Hamiltonian (1.4) is replaced with an approximate Hamiltonian identical in
the first two terms, but with an effective potential VHF instead of the exact rij
terms.3,4 This means that the Coulombic electron – electron repulsion is not
specifically taken into account. However its net effect is included in the
calculation in an average manner through the central field approximation. The
method was made vastly more applicable with the introduction of the Roothan
formalism.5 In 1951, Roothan showed that there exists a procedure by which a
matrix representation of the HF problem may be solved iteratively by
approximating orbitals as linear combinations of a finite set of basis functions
(known functions used to represent an unknown function). In the same year,
Ruedenberg devised mathematical techniques for evaluating integrals over a
basis of atomic orbitals.6 The concept of basis sets introduces a degree of
chemical intuition into quantum chemistry by way of almost universally
known orbital concept. A summary of Roothan-Hall-HF formalism is
highlighted below as a separate segment.

6
Summary of the derivation of the Roothan-Hall equations

1. The total wavefunction Ψ of a closed shell molecule having n electrons is

AAAAAA
expressed as a Slater determinant of spin MOs (9i s) and is designated here as Ψ:;

AAAA(1) 9
9 (1) 9 (1) .. .. 9 B (1)
AAAA(2) 9
@ 9 (2) 9 (2) .. .. AAAAAA
9 B (2)@
AAAA AAAAAA
Ψ:; (1, 2, … , n) = 9 (3) 9 (3) 9 (3) .. .. 9 B (3) (
( 1.8
1.8 )
: : : : : :
!
@ : : : @
: : :
9 (D) 9AAAA(D) 9 AAAAAA
(D) . . . . 9 B (D)
where, 9 stands for ith MO with F spin and 9HG for ith MO with I spin.
Since each orbital can accommodate two electrons DB2 spatial orbitals only are required. 1,
2, 3…, n in parentheses stand for the coordinates of electrons 1, 2, 3, …, n respectively.

2. The electronic energy is given by the average value postulate (assuming Ψ:; to be

. / LΨ:;M
normalized).
< 3 >= KΨ:;L- ( 1.9
1.9 )
3. Substituting the Slater determinant Ψ:; (1.8) and the explicit form of the
. / {equation (1.4)}, the energy can be shown to be (after much
Hamiltonian -
algebraic manipulation)
YB YB YB
& & &

QRS = & TUU + (&VUW − XUW ) (%. %Z)


U % U %W %

where,

- = [9 (1)\ \9 (1)` + [9 (1)\ \9 (1)` (%. %%)


] ∇_^ ]ab
c^b

d ! = [9 (1)9 (1)\ c \9! (2)9! (2)` (1.12)


(1.12)
e_

1
f ! = [9 (1)9! (2)g g9 (2)9! (1)` (%. %h)

We can assign the following physical meanings to each of the above terms (1.11 to
1.13). - corresponds to the electronic energy of a single electron moving simply under the
attraction of a nuclear core with all the other electrons stripped away. d ! which is called the
coulomb integral represents the electrostatic repulsion between an electron in 9 and another
in 9! . f ! , which is called the exchange integral can be regarded as a kind of correction to

7
d ! reducing the effect of d ! . We could consider the summed (2d ! − f ! ) terms to be the
true coulombic repulsion corrected to electron spin.
Substituting equations 1.11 to 1.13 in 1.10, 3 / can be conveniently written as a sum of
three terms
3 / = i + 01 + 0 (%. %5)

where,
i ; kinetic energy due to all the n electrons.(resulting from the first term in 1.11)
01 ; potential energy due to nuclear electronic coulombic attraction (resulting from the
second term in 1.11) and
0 ; potential energy due to inter electronic repulsion (due to the combined effect of
coulomb and exchange integrals.

4. The HF method looks for those MOs (9 s) that minimize the variational integral
(equation 1.9). Minimizing QRS in equation 1.10 with respect to 9 s [to find the best
9 s ] gives a set ‘n’ HF one electron equations.
jk (1)9 (1) = 9 (1) 1.15)
1.15)
(1.15)

where, jk is the Fock operator which contains additional terms due to electron spin
compared to the Hamiltonian operator and is a one electron energy.

jk (m) = - + { 2do! (m) − f


.! (m)} (%. %q)
!

do! is the coulomb operator and f


.! is the exchange operator.

equations. This is because of the fact that the Fock operator jk (m) itself depends on
5. Unlike the Schrödinger equation, the HF equations are not quite eigenvalue

9 : in a true eigenvalue equation the operator can be written without reference to


the function on which it acts. As such the HF equations were not very useful for
molecular calculations for mainly two reasons: they do not prescribe mathematically
viable procedure for getting the initial guesses for the MO functions 9 and the
wavefunction may be so complicated that they contribute nothing to a qualitative
understanding of the electron distribution. A key development that helped to make
feasible the calculation of accurate SCF wavefunctions was Roothan’s proposal to

8
expand the spatial orbitals 9 s as linear combinations of a set of one electron basis
functions. Substituting into the HF equations the Roothan-Hall linear combination
of ‘m’ basis functions χs

9 = ts χs (%. %u)
s

Substituting for the MO’s (in 1.15) the linear combinations of basis functions (1.17)
gives the Roothan Hall equations (after some mathematical manipulations) which
can be written in the matrix form as,
j j .. j w t t .. t w
j j .. j w t t .. t w
v : : : : xy : : : : z
jw jw . . jww tw tw . . tww
{ { .. { w t t .. t w 0 .. 0
{ { .. { w t t .. t w 0 .. 0
=v : : : : xy : : : : zv : : x (%. %})
: 0
{w {w . . {ww tw tw . . tww 0 0 . . ww

or F C = S C ε

Where, F is the Fock matrix consisting of the one electron Fock operators, C matrix
consisting of the coefficients in the linear combinations of basis functions in the
different MOs and S is the overlap matrix and ε is the diagonal matrix

The procedure for diagonalizing the fock matrix F and extracting the MO coefficients
and eigen values

1. The overlap matrix S is calculated and used to calculate an orthogonalising matrix


~ ]½
2. ~ ]½ is used to convert F to F’
The transformed Fock matrix F’ ‘satisfies F’ = C’ ε (C )-1

The overlap matrix S is readily calculated, so if F can be calculated, it can be


transformed to F’ , which can be diagonalised to give C’ and ε [ε yields the MO
energy levels “εi” s]
3. Transformations of C’ to C gives the coefficients csi in the expansion of the MOs
9 s in terms of the basis functions χs s.

9
From a practical standpoint, the central field approximation and the basis set
approximation allow an initial guess at the electron configuration (usually a
Slater determinant) of the chemical system to produce a guess at the effective
potential, which can then be used as an improvement to the guess of the
orbitals and so on. The iterative calculation attains a self consistency, once the
orbitals and the potential no longer change. The built-in advantage of the HF
method is that since it involves a variational calculation, there is a guarantee
that the energy computed is never going to be less than the actual ground state
energy of the system.

At first, HF-SCF calculations restricted the energies of α and β spin orbitals to


be identical as implied by most simple models of chemical systems. This type
of calculation is called restricted HF calculation (RHF). Although RHF
wavefunction is generally used for closed-shell states, two different
approaches are widely used for open-shell states. In the Restricted open-shell
Hartree Fock (ROHF) method, electrons that are paired with each other are
given the same spatial function. For example, the ROHF wavefunction for the
ground state of Li atom is given by the Slater determinant |1• AAA
1• 2•|. The
interaction of the 2sα and 1sα electrons differs from that between the 2sα and
1sβ electrons and it seems to be reasonable to give the two 1s electrons
slightly different spatial orbitals (1s and 1s’). This is the methodology in
selecting an unrestricted HF (UHF) wavefunction. Thus in the example of the
Li atom the UHF wavefunction is |1• AAAA 2•|, where 1s≠1s’.7 The UHF
1•′
wavefunction gives a slightly lower energy than the ROHF wavefunction and
is much more useful in predicting esr spectra. The main problem with the
UHF wavefunction is that it is not an eigen function of the square of the spin
operator, •̂ , whereas, the true wavefunction and the ROHF wavefunction are
eigen functions of •̂ .

10
1.4.2 The HF limit and the concept of electron correlation
The conceptual underpinnings of HF-SCF are both its strength and its
downfall. Because of the central field imposed by the HF potential, each inter
– electronic interaction is not explicitly taken into account, instead it is treated
only in an average manner. Therefore, there is a finite probability that two
electrons will occupy the same point in space, which is unrealistic. For high
accuracy, it is necessary to describe properly how the motion of each electron
is affected by that of the other electrons. This is the electron correlation
problem8, which is often the single largest source of error in quantum
chemical calculations. There are two types of correlation: Fermi correlation
arising from symmetry and Coulomb correlation arising from electron –
electron repulsion. Higher levels of theory attempts to resolve this issue by
recovering the correlation energy, which is defined as
Ecorrelation = Eactual – EHF-limit (1.19
(1.19)
19)
where, EHF-limit is the HF energy of the system obtained with an infinite basis
set. Ecorrelation is always negative because the HF energy is an upper bound to
the exact energy (as guaranteed by the variation theorem). Most of the
standard high-level techniques in quantum chemistry utilize the HF
approximation as a starting point and then attempt to correlate electrons by
more rigorous computational methods.

1.4.3 Post - HF methods


Generally, there are two classes of methods employed to recover the majority
of correlation energy - variational and perturbational. Variational methods
treat the exact solution as a linear combination of discrete solution sets, while
perturbative methods separate the problem into an exactly solvable part and a
difficult part with no general analytic solution. The most popular variational
post-HF method is called the Configuration interaction (CI) and the two
important perturbative post-HF methods are the Many Body Perturbation
Theory (MBPT) and the Coupled Cluster (CC) method.

11
1.4.3.1 Configuration interaction methods

Configuration interaction (CI) methods begin by noting that exact


wavefunction Ψ cannot be expressed as a single determinant as HF theory
assumes.9,10 The standard CI method relies on partitioning the exact
wavefunction Ψ into a selected collection of Ψ s and their corresponding
coefficients ci’s with a view to including the interaction energy due to excited

Ψ = t0 Ψ0 + ∑m tm Ψm (%. &Z)
states.

where, Ψ0 is the HF determinant (equation 1.8) and each Ψi is a Slater


determinantal antisymmetric function corresponding to an excited state
configuration obtained by replacing one or more occupied orbitals 9 s within
the HF determinant with a virtual orbital. If all the possible excited
configurations are considered, the calculation is referred to as a full CI (FCI).
Such a calculation is variational and size consistent,* but extremely expensive
and therefore impractical for all but very smallest systems.

Practical CI methods augment the HF by adding only a limited set of


substitutions, truncating the CI expansion at some level of substitution. For
example, the CI-singles (CIS) method11 adds single excitations (a virtual
orbital, say, 9… , replaces an occupied orbital 9 within the HF determinant) to
the HF determinant, CI-doubles (CID) adds double excitations (two occupied
orbitals are replaced by virtual orbitals), CI-singles doubles (CISD) adds
singles and doubles excitations, CI-singles doubles triples (CISDT) adds
singles, doubles and triples excitations and so on.12,13,14 The Davidson
correction can be used to estimate a correction to the CISD energy to account
for higher excitations.15 A disadvantage of all these limited CI variants is that
they are not size consistent. CISD, which is typically augmented with triple
and/or quadruple excitations in some fashion is known as the Quadratic
16
Configuration Interaction (QCISD) and was developed to correct this

12
deficiency. QCISD (T) adds triple substitutions to QCISD, providing an even
greater accuracy.16 Similarly, QCISD (TQ) adds both triples and quadruples
from the full expansion to QCISD.

1.4.3.2 Perturbative theories


Many – body perturbation theory (MBPT) was the first of the perturbative
theories to evolve. MBPT is a way to account for electron correlation by
treating it as a perturbation to the HF wave function. It is a rather
straightforward application of simple perturbation theory. Using MBPT, we

Ĥ| †Ψ〉=
† Ĥ0 L †9 〉† + 0k L †9 9 … 9 〉† = E| †Ψ〉†
wish to solve the eigen value problem

0 1 2 D
(1.21)
(1.21)
where, we know the eigen functions and eigen values of Ĥ0
Ĥ 0 9 0 = E0 9 0 (1.22
(1.22)
22)
)
and will systematically improve the eigen functions and eigen values of Ĥ0
such that they approach those of Ĥ. This can be accomplished by producing a
set of separable equations that represent various orders of a Maclaurin series:
by solving each equation, the exact energy can be determined by summing the
values
E = E0 + E(1) + E(2) + …… + E(n) (1.
(1.23
(1.23)
23)
17
In the Møller – Plesset perturbation theory (generally referred to by the
acronym MPn, where, n is the order at which the perturbation theory is
truncated), Ĥ0 is defined as the sum of the one electron Fock operators.
Ĥ0 = Σ Fi (1.24
(1.24)
24)
where, Fi is the Fock operator acting on the ith electron.

“MP0” would use the electronic energy obtained by simply summing the HF
one electron energies (first sum in equation 1.10). This ignores inter-
electronic repulsion except for refusing to allow more than two electrons in
the same spatial MO. “MP1” corresponds to “MP0” corrected with the
Coulomb and exchange integrals J and K, i.e., “MP1” is just the HF energy.

13
The variant of MP theory which is truncated at the second order is known as
“MP2”.18,19,20,21,22 MP2 remains a very useful tool in a computational
chemist’s tool box, for it can successfully model a wide variety of systems
and optimize geometries quite accurately. Higher level MP orders (MP323,24,
MP425, and MP526) are available for cases where the second order solution of
MP2 is inadequate. In practice, however, only MP4 sees wide use. MP3 is
usually not sufficient to handle cases where MP2 does poorly, and it seldom
offers improvements over MP2 which are commensurate with the additional
computational cost. Since MPn methodology is not variational, it is possible
that such methods (especially MP2) overestimate the correlation energy.
However, this rarely happens in practice because of the basis set limitations.

At the MP4 level, integrals involving triply and quadruply excited


determinants appear. The evaluation of the terms involving triples is the most
costly (triples doubles the time) and scales as N7. If we ignore the triples, the
method scales more favourably and this choice is typically abbreviated
MP4(SDQ).27,28

1.4.3.3 Coupled cluster methods


One of the mathematically elegant techniques for estimating the electron
correlation energy is the coupled cluster theory.29,30,31Since its introduction in
the late 1960s, it has emerged as perhaps the most reliable method for the
prediction of molecular properties. The theory begins by suggesting that the
exact wave function may be described as
Ψ = eT Ψ 0 (1.25
(1.25)
25)
in which, Ψ0, is a suitable reference function, usually the HF function. ‘T’ is
an excitation operator called the cluster operator. The effect of eT operator is
to express Ψ as a linear combination of Slater determinants that include Ψ0
and all possible excitations of electrons from occupied to virtual spin orbitals.
T = T1 + T2 + ……. + Tn (1.26)

14
The exponential form of the operator makes coupled cluster (CC) theory
unique among all techniques in quantum chemistry. Operating on the HF
wave function with T is equivalent to doing a full CI. But CC method is
superior to CI when we try to truncate T at some point. For example, suppose
that we consider only the double excitation operator, i.e., T =T2, then the
method is called coupled cluster doubles (CCD) or coupled-pair many-
‰_
electron theory (CPMET).32 ΨCCD = ΨHF
‰__ ‰_Š
= (1+ i + + + ……) ΨHF (1.27)
! ‹!

The first two terms, (1 + T2), define the CID method and the remaining terms
involve products of excitation operators. Each application of T2 generates
double excitations, so the square of T2, generates quadruple excitations and
the cube of T2 generates hextuple excitations and so on, thus making it size
consistent. In fact, the failure to include these excitations makes CID non size
consistent. Studies show that CCD theory gives results close to those of a
Møller-Plesset perturbation treatment to fourth order in the space of double
and quadruple substitutions MP4(DQ).

Methods for analytic evaluation of the CCD gradient have been developed by
Bartlett et.al.33 Approximations to the CCD method include, the linearized
‰_
CCD (LCCD), which uses the approximation ≈ T1 + T2), the coupled
electron-pair approximation (CEPA) and independent electron-pair
approximation (IEPA).

The next step in improving the CCD method is to include the operator T1 and
take T = T1 + T2 in eT: this gives the CC singles and doubles (CCSD)
method.34,35,36With T = T1 + T2 + T3, one obtains the CC singles, doubles,
triples (CCSDT) method.37 CCSDT calculations give very accurate results for
correlation energies but are very demanding computationally and are only
feasible for small molecules with small basis sets. Instead of the very
demanding CCSDT, one often performs the CC perturbative triples

15
{CCSD(T)} which uses the perturbation theory to compute the connected
triples.38 It is one of the most popular ab initio methods. One goal of ab initio
quantum chemistry is to be predictive. Predictive quantum chemistry requires
a very accurate inclusion of the essential effects of electron correlation. CC
theory offers a novel, elegant approach for correlation that has had a dramatic
impact on the field in the past two decades and is destined to have important
impact in the future.39 The application of orbital based CC theory to the
calculation of molecular energies and properties is recently reviewed by
Helgaker et.al.40

Hybrid methods incorporating the spirits of CI and perturbative methods are


also reported. The CIS(D) which uses configuration interaction singles with
perturbative doubles is an example.41,42 A modification of the CIS(D)
abbreviated as PR-CIS(Ds) (partially renormalized variant of configuration
interaction singles with perturbative doubles and extra singles) has been
suggested recently. 43

Two factors that should be mentioned with post-HF calculations are of


whether a method is size-consistent and/or whether it is variational. A size
consistent quantum mechanical method is one for which the energy and hence
the energy error in the calculation increase in proportion to the size of the
molecule. It is important whenever calculations on molecules of substantially
different sizes are to be compared. A special case of size consistency is that of
infinitely separated systems. When dealing with such a system, a method is
said to be size-consistent if it gives the energy of a collection of ‘n’ widely
separated atoms or molecules as being ‘n’ times the energy of one of them.
HF& FCI - size-consistent and variational.
MPn - size-consistent but not variational.
CID - not size-consistent but variational.
CCD - size-consistent but not variational.

16
1.4.3.4 Multi-configuration (MC) methods
In configuration interaction (CI), the trial wavefunction is written as a linear
combination of determinants with the expansion coefficients determined by
requiring that the energy should be a minimum. The MO’s used for building
the excited state Slater determinants are taken from a HF calculation and held
fixed. The MCSCF44,45,46,47,48 on the other hand, can be considered as a CI,
where not only the coefficients in the linear combination of determinants are
optimized by the variational principle, but also the MO’s used for
constructing the determinants are made optimum. The MCSCF optimization
is iterative just like the SCF procedure (if the “multiconfiguration” is only
one, it is simply HF). Since the number of MCSCF iterations required for
achieving convergence tends to increase with the number of configurations
included, the size of MCSCF wavefunction that can be treated is somewhat
smaller than that for CI methods.49 MCSCF methods are rarely used for
calculating large fractions of the correlation energy. The orbital relaxation
usually does not recover much electron correlation, it is more efficient to
include additional determinants and keep the MOs fixed (CI) if the interest is
just in obtaining a large fraction of the correlation energy.
An example where MCSCF method can be used to good effect is in dealing
with the structure of ozone.

O O
O O O O
Figure 1.1: Resonating structures of ozone
Ozone has two non-equivalent resonating structures. A RHF single
determinant wavefunction is inadequate in this case. The simplest MCSCF for
ozone will contain two configurational state functions (CSFs) with the
optimum MOs and configurational weights determined by the variational
principle. The CSFs entering an MCSCF expansion are pure spin states, and

17
MCSCF wavefunctions therefore do not suffer from the problem of spin
contamination.

The major problem with MCSCF methods is selecting the necessary


configurations to include the property of interest. One of the popular
approaches is the complete active space SCF (CASSCF) method50 {also
called full optimized reaction space (FORS)}. Here the selection of
configurations is done by partitioning the MOs into active and inactive
spaces. The active MOs will typically be of some of the highest occupied and
some of the lowest unoccupied MOs from a RHF calculation. The active MOs
either have 2 or 0 electrons, i.e. always doubly occupied or empty. Within the
active MOs a full CI is performed and all the proper symmetry adapted
configurations are included in the MCSCF optimization. Which MOs to
include in the active space must be decided manually, by considering the
problem at hand and the computational expense. A common notation is [n,m]
– CASSCF, indicating that ‘n’ electrons are distributed in all possible ways in
‘m’ orbitals.

As for any full CI expansion, the CASSCF becomes unmanageably large even
for quite small active spaces. A variation of CASSCF is the restricted active
space SCF (RASSCF) method.51. Here the active MOs are further divided into
three sections, RAS1 (MOs which are doubly occupied in the HF reference
determinant) RAS2 (has both occupied and unoccupied orbitals) and RAS3
(MOs which are empty in the HF determinant).

RAS3

CAS All excitations RAS2

RAS1

Figure 1.2: Partitioning of MOs in CASSCF and RASSCF methods

18
An FCI in RAS2, a maximum of two electrons excited from RAS1 and a
maximum of two electrons excited to the RAS3 space. In essence, a typical
RASSCF procedure thus generates configurations of a full CI in a small
number of MOs (RAS2) and a CISD in a somewhat larger MO space (RAS1
and RAS2). The number of singlet CSFs generated for an [n,n] – CASSCF
wavefunction shows a factorial increase.

Table 1.1: No. of CSFs as a function of no. of electrons in a CASSCF method

N No. of CSFs
2 3
4 20
6 175
8 1764
10 19404
12 226512
14 2760615

1.4.3.5 Composite methods

The highest level ab initio techniques are not applicable (in a practical point
of view) to even medium sized molecules. Two different classes of
approximations (compromise) to overcome this difficulty are basis set
extrapolation techniques and the design of composite methods (which include
some kind of empiricism). The former, which is more accurate among the two
classes, is based on CCSD(T) calculations using very large correlation
consistent basis sets extrapolated to the complete basis set limit with addition
of corrections for some smaller effects not included in the calculations such as
core-valence effects, relativistic effects and atomic spin-orbit effects. This
type of approach is limited to smaller molecules because of the use of very
large basis sets. The later methods are widely used for the calculation of
thermo-chemical data. They combine methods with a high level of theory and

19
a small basis set with methods that employ lower levels of theory with larger
basis sets. They are commonly used to calculate thermodynamic quantities
such as enthalpies of formation, atomization energies, ionization energies and
electron affinities. They aim for chemical accuracy which is usually defined
as within 1 kcal/mol of the experimental value. The most popular among them
are the Gaussian-n (Gn) theories, which employ a set of calculations with
different levels of accuracy and basis sets with a goal of approaching the
exact energy. In the Gn approach, a high level correlation calculation {e.g.
QCISD(T) and CCSD(T)} with a moderate sized basis set is combined with
energies from lower level calculations (e.g. MP4 and MP2) with larger basis
sets to approximate the energies of more expensive calculations. In addition,
several molecule-independent empirical parameters {higher level correction
terms (HLC terms)} are included to estimate remaining deficiencies,
assuming that they are systematic. The different steps involved in the G152,
G253, G354 and G455 theories are listed below.
An intermediate approach, referred to as correlation consistent composite
approach (ccCA) with no parametrization has recently been introduced.56
Other composite techniques related to Gn methods include, the complete basis
set (CBS) methods57 and the multi coefficient methods.58,59

Steps involved in G2 theory


1. E+ = E[MP4/6-311+G(d,p)] – E[MP4/6-311G(d,p)]

2. E2df = E[MP4/6-311G(2df,p)] – E[MP4/6-311G(d,p)]

3. EQCI = E[QCISD(T)/ 6-311G(d,p)] – E[MP4/6-311G(d,p)]

4. E+3df,2p = E[MP2/6-311+G(3df,2p)] – E[MP2/6-311G(2df,p)] –


E[MP2/6-311+G(d,p)] + E[MP2/6-311G(d,p)

5. ∆3 HLC = - 4.81 Nβ – 0.19 Nα


6. 3 G2 = E[MP4/6-311G(d,p)] + E+ + E2df + EQCI + E+3df,2p +
∆3 HLC

20
Steps involved in G3 theory
1. HF/6-31G(d) geometry optimization
2. Zero point vibrational energy (ZPVE) from HF/6-31G(d) frequencies
3. MP2(Full)/ 6-31G(d) geometry optimization (all subsequent
calculations use this geometry)
4. E[MP4/ 6-31+G(d)] - E[MP4/ 6-31G(d)]
5. E[MP4/ 6-31G(2df,p)] - E[MP4/ 6-31G(d)]
6. E[QCISD(T)/ 6-31G(d)] - E[MP4/ 6-31G(d)]
7. E[MP2(Full)/G3 large] - E[MP2/ 6-31G(2df,p)] - E[MP2/ 6-31+G(d)] -
E[MP2/ 6-31G(d)]
8. EHLC = – 0.006386 (number of valence electron pairs) – 0.002977
(number of unpaired valence electrons).
9. EG3 = 0.8929 (2) + E[MP4/ 6-31G(d)] + (4) + (5) + (6) + (7) + (8)

Steps involved in G4 theory


1. B3LYP/6-31G(2df,p) geometry optimization
2. ZPVE from B3LYP/6-31G(2df,p) frequencies scaled by a factor 0.9854
3. EHF/aug-cc-pVnZ = EHF/limit + B ]α• where, α is an adjustable parameter and
‘n’ the number of contractions in the valence shell of the basis set.
4. E+ = E[MP4/ 6-31+G(d)] - E[MP4/ 6-31G(d)]

5. E(2df,p) = E[MP4/ 6-31G(2df,p)] - E[MP4/ 6-31G(d)]

6. ECC = E[CCSD(T)/ 6-31G(d)] - E[MP4/ 6-31G(d)]

7. E(G3 large XP) = E[MP2(Full)/G3 large XP)] - E[MP2/ 6-31G(2df,p)] -


E[MP2/ 6-31+G(d)] + E[MP2/ 6-31G(d)]
8. E(combined) = E[MP4/ 6-31G(d)] + E+ + E(2df,p) + ECC + E(G3 large
XP) + {EHF/limit – E[HF/G3 large XP]}+ E(SO), where, E(SO) is the
spin-orbit correction.

9. Ee(G4) = E(combined) - EHLC, where, EHLC is the same as that of G3


10. E0(G4) = Ee(G4) + E(ZPE).

21
The rationale behind these model chemistries is determined by a fitting to a
library of compounds called the test set. A test set is a compilation of accurate
experimental data. The first in this series was the G2 test set of 125 energies.53
This was followed by the G2/97 (301 energies)60, G3/99 (376 energies)61 and
G3/05 (454 energies)62 test sets. Each succeeding test set includes energies
from the preceding test sets and additional species of larger sizes and of
different types. The test sets contain thermochemical data such as enthalpies
of formation, ionization potentials, electron affinities and proton affinities
chosen based on listed accuracy of + 1 kcal/mol or better. The latest test set
G3/05 contains 270 enthalpies of formation, 105 ionisation energies, 63
electron affinities, 10 proton affinities and 6 hydrogen bonded complexes.
When the G3 theory was originally published, it was assessed on the G2/97
test set and was found to have an average absolute deviation of 1.02 kcal/mol
from experiment. The two succeeding test sets, G3/99 and G3/05 gave
average absolute deviations of 1.07 and 1.13 kcal/mol respectively for the G3
theory. G4 theory provides significant improvement and this is evident from
the average absolute deviation of 0.83 kcal/mol from experiment.

1.4.3.6 Methods for excited states


Ground state wave functions can very often be expressed in terms of a single
Slater determinant formed from variationally optimized MO’s, with possible
accounting for electron correlation taken thereafter. However, the problem of
variational collapse typically prevents an equivalent SCF description for
excited states. That is, any attempt to optimize the occupied MO’s with
respect to the energy will necessarily return the wave function to that of the
ground state.63 This limitation of traditional quantum chemistry led
researchers to develop new techniques for electronically excited states. This is
absolutely essential since many reactions are photochemically activated
(leading to electronically excited states) rather than thermally activated. ab
initio methods used to simulate electronically excited states include CIS,

22
time–dependent HF and Green function methods. Even though a large variety
of efficient quantum chemical methods is available for electronic ground-state
calculations, the situation is much more difficult when excited-state energy
surfaces and surface crossings should be treated. Analytic energy derivatives
and nonadiabatic coupling vectors are crucial tools for the determination of
energy minima, saddle points and conical intersections.
1.4.3.7 Multireference methods

Chemically speaking, a multireference system is a molecule or reaction in


which the one electron approximation is not only quantitatively but also
qualitatively wrong. Practically, all reactions that involve bond breaking are
of this type, as are some low lying excited states of many molecules. In such
situations, the typical HF wavefunction is not a good description of the
system, and something more complex is necessary to reproduce the chemistry.
The CI methods consider only CSFs generated by exciting electrons from a
single determinant. This corresponds to having a HF type wavefunction as the
reference. However, a MCSCF wavefunction may also be chosen as the
reference. In that case a CISD involves excitations of one or two electrons out
of all the determinants which enter the MCSCF. This method is referred to as
a multi-reference configuration interaction [MRCI] method.64 Compared to
the single reference CISD, the number of configurations is increased by a
factor roughly equal to the number of configurations included in the MCSCF.
Large scale MRCI wavefunctions [many configurations in the MCSCF] can
generate very accurate wavefunctions, but are also computationally very
intensive. Variations of MRCI include the MRCISD where single and double
excitations are also considered. Since MRCI methods truncate the CI
expansion, they are not size-consistent. State specific MRCC (SS-MRCC)
methods applying the principles of coupled cluster theory on the
multireference determinants has also been reported.65 They are shown to be

23
size consistent. The size consistency of various multireference methods is
analysed by Ruttink et.al.66

1.4.3.8 r12 methods


The Hartree-Fock (HF) approximation uses an effective potential (a smooth
potential as a consequence of the central field approximation) to replace the
inter-electronic terms (rijs) in the Hamiltonian. It has been long acknowledged
that the true Hamiltonian form displays singularities in the coulomb potential
at the points where electrons would exist at the same position. These are
called “coalescence points’’, and they imply that there will be irregularities in
the first and higher derivatives of the wavefunction with respect to the inter-
electronic distance r12 as shown by Kato in 1957.67 A smooth potential (like
that used in HF) will not intrinsically model this ‘cusp’ and therefore will
show slow convergence behavior. All electron correlation methods based on
expanding the N-electron wavefunction in terms of Slater determinants built
from the HF determinants suffer from this problem of agonizingly slow
convergence. Literally millions or billions of determinants are required for
obtaining results which in an absolute sense are close to the exact results. It
would therefore seem natural that the interelectronic distance would be a
necessary variable for describing electron correlation. Methods based on this
idea has been known since the early days of quantum mechanics. Hylleraas
used a trial wavefunction consisting of an orbital product times an expansion

∑•/w t•/w (
in electron coordinates, such as
Ž( , ) = ]•e ce
]•_ c_
+ )

( − )
/ w
(1.28)
for attacking the two electron He atom problem, variationally
optimized Ž( , ) by hand and yielded an energy within ≈ 0.6 kcal/mol of
the exact non-relativistic value (it would require ≈ 30 terms in a CI
calculation and a computer to match that feat). Calculations incorporating
more complicated expansions have achieved results approaching the exact
-12
non-relativistic energy of He atom to within 10 kcal/mol.68 Hylleraas type

24
wavefunctions used for H2 molecule produces an energy within 10-9 a.u.,
which is more accurate than that can be determined experimentally.69
Unfortunately, Hylleraas
raas type wavefunctions became impractical for systems
containing more than 3-4
3 4 electrons. For such systems, ways to explicitly
include terms in the wavefunctions which are linear in the interelectronic
70
distances have been developed by Kutzelnigg and coworkers.
cowor The first order
correlation to the HF wavefunction involves only doubly excited determinants
(as in MP2 theory). In r12 methods additional terms are included which
essentially are the HF determinants multiplied with rij factors. For example,
Žce_ = ‘’“ + ∑ !…• ” !…• ‘…•
! + ∑!–! ! ‘’“ (1.29)

Such r12 wavefunctions may then be used in connection with the CI, MBPT or
CC methods. These methods show significant promise and are in
i frequent use
for achieving greater accuracy. The recent versions of explicitly correlated
coupled-cluster (CC-R12
R12 or F12) methods include CCSD-R12
R12 (up to double
excitations)71, CCSDT-R12
CCSDT R12 (up to triple excitations) and CCSDTQ-R12
CCSDTQ (up
to quadruple excitations)72. Using the Slater-type exp( r12) as a
type function exp(−
correlation function, a CC-R12
CC method can provide the aug-cc
cc-pV5Z-quality
results of the conventional CC method of the same excitation rank using only
the aug-cc-pVTZ
pVTZ basis set.

1.5 Quantum Mechanics


anics- Molecular Mechanics (QM-MM)
MM) Methods
Force field methods are inherently unable to describe the details of bond
breaking/ forming reactions, since there is an extensive rearrangement of the
electrons, which is neglected in the classical model. If the system of interest is
too large to treat entirely by electronic structure methods, there are two
possible approximate
imate methods that can be used. In some cases the system can
be ‘pruned’ to a size that can be treated, by replacing ‘unimportant’’ parts of
the molecule by smaller model groups, e.g. substitution of a hydrogen or
methyl group for a phenyl ring. For studying enzymes, however, it is usually

25
assumed that the whole system is important for holding the active size in the
proper arrangement and the ‘backbone’ conformation may change during the
reaction. Hybrid methods have been designed for modeling such cases, where
the active size is calculated by electronic structure methods (usually semi-
empirical, low level ab initio or DFT methods), while the backbone is
calculated by a force field method. Such methods are often denoted QM-MM
methods.73 The new ONIOM (our own n-layered integrated molecular orbital
and molecular mechanics) approach has been proposed and shown to be
successful in reproducing benchmark calculations and experimental results.74
The main problem with QM-MM schemes is deciding how the two parts
should be connected. Partial charges on the MM atoms can be incorporated
into the electronic HF equations, analogously to nuclear charges (i.e. adding
Vne-like terms to the one electron matrix elements, and the QM atoms thus
feel the electric potential due to all the MM atoms.

1.6 Semi Empirical Methods


Semi Empirical (SE) methods were developed to treat large molecules
because of the difficulties met with in applying the ab initio methods to such
molecules. SE methods also attempt a quantum mechanical description of
electrons based on similar principles as ab initio, but with many (more)
approximations built into the equations to make calculations go faster. Some
of the commonly used approximations include the zero differential overlap
approximation (i.e. overlap matrix is reduced to unit matrix), setting one-
electron integrals involving three centres to zero and neglecting many three
and four centre 2-electron integrals. These approximations considerably
decrease the complexities in the computation process. However, to
compensate for these approximations, parameterization (design of
computational equations or input parameters) based on experimental
(empirical) data is often done. The advantages include; calculations are faster
than ab initio, larger systems can be handled and the results often agree well

26
with experimental values because of parameterization. But the major
disadvantage of such methods is that they are only reliable for systems for
which a method is parameterized.
Some of the widely used SE methods include HUCKEL method, Austin
Model 1 (AM1), Parametric Method 3 (PM3) and methods based on neglect
of differential overlap (MNDO, CNDO, MINDO, CINDO etc.).

1.7 Density Functional Theory (DFT)


One of the most important and difficult problems in chemistry is the accurate
calculation of molecular properties from first principles. Chemists aim to
calculate the energy of molecules to within chemical accuracy (<1 kcal/mol),
which is required to predict reaction rates. Exact first-principles calculations
of molecular properties using the wavefunction theory (solving the
Schrödinger equation using the FCI method and computing the 3N
dimensional wavefunction) are currently intractable because their
computational cost grows exponentially with both the number of atoms and
basis set size. The difficulty of solving this problem exactly has led to the
development of several approximate methods for first-principles quantum
chemistry. The Density Functional Theory (DFT) is one among them. The
foundation of DFT lies in the fact that it is not necessary to solve the
Schrödinger equation and determine the 3N dimensional wave function in
order to compute the ground state energy of a system. Instead the energy and
other properties can be expressed as functionals of electron density; a three
dimensional entity. Since the computational method used in the present work
is DFT based, a detailed account of DFT methods and concepts is included in
the second chapter.
1.8 Basis Sets
Another approximation involved in all ab initio methods is the introduction of
a basis set (see Roothan-Hall-HF formalism). A basis set is defined as a set of
known functions used to represent an unknown function such as a molecular

27
orbital. The quantum chemistry literature has a plethora of basis sets and a
comprehensive review is beyond the scope of the present writing. In this
section, a simplified introduction to basis sets is attempted with a view to
highlighting a historical sketch of the development of basis sets and the merits
and demerits of different types of basis sets in use. Special care has been
given to include the original reference as far as possible.
The first step in any ab initio method of atoms or molecules is the selection of
a trial atomic or molecular wavefunction, which is usually built from a set of
one-electron one-center functions (usually called atomic orbitals) and then
making improvements (depending on the method in use) to produce the best
possible final wavefunction. Since molecular properties are computed from
this function it is always very crucial to select the initial guess in a judicious
manner. There is controversy that whether the individual functions in a basis
set can be called atomic orbitals (AOs) or not. Strictly speaking, AOs are
solutions of the HF equations for the atom, i.e. wave functions for a single
electron in the atom. Anything else is not really an atomic orbital. Later on,
the term AO was replaced by "basis function". Hence, a basis set may be
defined as a collection of basis functions (not really AOs) used to construct an
AO or a MO of a multi electron system. Thus, it can be generally represented
as:
9 = ts χs (%. &—)
s

here • varies from 1 to m. Here, 9 is the ith MO, ts are the coefficients of
linear combination, χs is the sth basis function and ‘m’ is the number of basis
functions.

The choice of a basis set must be guided by the consideration of two mutually
opposing factors - desired accuracy in the results and computational costs.
The basis set should be flexible enough to produce good results over a wide
range of molecular geometries and sufficiently small so that the problem is

28
computationally easy and economically reasonable. The enormous
computational cost required when a large basis set is used forces us to
truncate the basis set at some convenient point on practical grounds. The
truncation of the basis set is an important source of error in molecular
electronic structure calculations. Most recent work has recognized the need
for a more systematic approach to the problem of constructing basis sets with
the aim of reducing the basis set truncation error in molecular electronic
structure calculations. In practice, most popular computer programs for ab
initio calculations contain internally defined basis sets from which the user
must select an appropriate one, always creating a chance of a subjective error.
It may be possible in the future to have programs make an informed decision
about the choice of basis set based on the results of thousands of previous
calculations which are accessible in a data base, but that time has not yet
come. The responsibility still rests with the program user. The available basis
functions can be grouped into two major classes - Exponential Type Orbitals
(ETOs) and Gaussian Type Orbitals (GTOs).

1.8.1 Exponential type orbitals


The only exactly known wavefunctions are the hydrogenic AOs which are
one electron functions and can be written in the form,
Ψ ,/,w ( , θ, φ) = $ ,/ ( )˜/,w (θ, φ) (1.30)
where, the $ ,/ ( ) is the radial part which consists of the Lauguerre
polynomials and the ˜/,w (θ, φ) is the angular part which is a spherical
harmonics function. Naturally, one will be tempted to go for these functions
(with appropriate modifications) when looking for a set of functions to
describe the wavefunction of a multi-electron atom or a molecule. The multi-
centre integral evaluations using these kinds of functions (as required by HF-
Roothan-Hall formalism) are mathematically formidable. In 1930, Slater

29
proposed modifications of the hydrogen wavefunctions by replacing the radial
part by a much simpler function which has the functional form,
∗ ›½) ∗]
$ ( ; D, ζ) = 2ζ( n(2D∗ )!p]½ ]ζœ
(1.31)
a]•
where, n*(the effective quantum number) and ζ = (where, Z being the

atomic number and S being the screening constant) are parameters determined
by Slater’s rules.75 He formulated the rules by fitting of mathematical
expressions to the numerical data for the radial distributions for the electrons
in many electron atoms.76,77
During the same period Zener showed that the wave functions for the atoms
Be-Ne, can be written as simple analytic expressions with several parameters
whose best values being determined by the variation method.78 Afterwards,
Roothaan and Bagus wrote an SCF code for atoms under the LCAO
approximation. They introduced the functions
χ(r, θ, φ; ζ, n, l, m) = N ]
˜/,w (θ, φ) ]ζœ
(1.32)
where, n, l and m being the principal, azimuthal and magnetic quantum
numbers, respectively and ζ is called the orbital exponent. The £ values are
determined by performing variational HF calculations for atoms using the
exponents as variational parameters. The exponent values which give the
lowest energy are the best – at least for the atom. In general, these functions
are named Slater type orbitals (STOs). They satisfy the nuclear cusp condition
because of their exponential relationship with the nucleus-electron distance.
This fact allows the STO basis functions to reproduce correctly the behaviour
of electrons in the regions near the nucleus. They also decay properly in the
regions far away from the nucleus, in a similar way to that of the hydrogen
atomic orbitals. Thus STOs show excellent behavior in the near and far
regions of the atomic nucleus. A basis set constructed using STO basis
functions is known as an STO basis set. For the atoms from He to Kr, basis
sets of STO functions were optimized with respect to the atomic energy by

30
Clementi.79 This work was improved and extended until the Xe atom by
Clementi and Roetti80 and until the Rn atom by McLean and McLean.81 They
also pointed out that if a single STO (per sub shell) fails to reproduce the
exact chemistry, the quality of the results can be improved by representing
each sub shell by proper linear combinations of more than one STO s. For
example a 2s sub shell in an atom can be represented using two STO s as,

STO basis sets can be classified as minimal or single zeta (SZ), double zeta
(DZ), Triple zeta (TZ) and so on depending on the number of STOs used to
represent each sub shell in an atom or a molecule.
Minimal basis set: a set containing only enough functions to contain all the
electrons of the neutral atom(s).
DZ basis set: there are two STOs for each sub shell, such that they have
identical n and l but different ζ values. The basis function with larger ζ value
is a tighter function and the one with smaller exponent is more diffuse. A DZ
basis set offers more flexibility to the wavefunction and can be used to
represent different types of bonding better than a SZ basis set (say σ and π
bonds formed by the same atom in a molecule).

TZ basis set: Three times as many functions as in the minimal basis set.
Quadruple zeta (QZ) and Quintuple zeta (5Z, not QZ) are also created on
similar arguments.
Split-valence basis set: Often it takes too much effort to calculate a double-
zeta for every orbital. Since only valence electrons are involved in bonding,
doubling can be restricted to only the valence orbitals. This is the principle
behind a split-valence basis set. Split valence basis sets can further be
categorized as valence double zeta (VDZ), valence triple zeta (VTZ) etc.
based on number of STO s used to represent the valence orbitals. For

31
example, the VDZ basis set uses a SZ for the core electrons and a DZ for
valence electrons.
Unfortunately, the initial and successful work developed in the atomic
calculations with STO functions was almost stopped for a long period. The
exponential functions on r2 (named Gaussian type orbitals or GTO) started to
become the standard basis sets employed in the calculation of the molecular
electronic structure. This regression in the use of the STOs was due to the fact
that such orbital products on different atoms were difficult to manipulate for
the evaluation of two-centre integrals (due to cumbersome orbital translations
involving slowly convergent infinite sums).
Later, efforts were undertaken to solve the problems associated with integral
evaluation and the interest in STOs was renewed (of course due to their
superiority compared with GTO s). Earlier reported basis sets6 were improved
by re-optimizing the exponents taking advantage of the present computational
resources and new optimization algorithms.82

Optimization criteria different from the energy minimization have also been
proposed and used for obtaining basis sets. Such methods include the
maximum overlap between the functions of a large reference basis set and the
functions of the optimized basis set83, the minimal geometrical distance
between the sub space associated to the reference basis set and the optimized
one84, the minimization of the mono and bi electronic contributions to the
total atomic energy85 and the Generator Coordinate HF method.86 These types
of basis sets have been applied to the calculations of molecular systems.

The basis set optimization demands a high computational cost which


increases with the size of the considered basis set. To overcome this difficulty
alternative ways for generating basis sets with a limited lower number of
variational parameters were proposed. Slater functions with a common
exponent for different principal quantum numbers was one of them. That is

32
1s, 2s and 2p sub shells were represented as linear combinations of STOs with
the same exponent. The use of only one exponent reduces the number of
integrals to be evaluated in the electronic structure calculations. Such STO
basis sets are known as Single Exponent Slater Functions (SESF).87 Another
attempt in this direction was the introduction of STO basis sets with non-
integer quantum numbers. Roothaan-Hartree-Fock calculations were carried
out for the ground states of the atoms from helium to xenon,88 for the singly
charged cations Li+ to Cs+ and anions H- to I-89 and for the atoms Cs (Z = 55)
through Lr (Z = 103)90 using a minimal basis set of Slater-type functions
whose principal quantum numbers were allowed to take variationally optimal
non-integer values. The resulting energies are substantially superior to those
obtained previously under the usual restriction that principal quantum
numbers be positive integers. Considering the improvement in the total
energy produced by the non-integer quantum numbers and the decreasing of
variational parameters as in SESF basis sets, a mixed strategy has been
proposed. For the atoms He to Ar, a new type of double-zeta basis set that
combines the use of non-integer principal quantum numbers and the use of
common exponents in Slater-type functions has been suggested.91 The new
double-zeta non-integer SESFs (NSESF) basis sets result in an improvement
of the energy and a reduction of the computational time.

An alternate method which uses exponential type orbitals (ETO) other than
STOs has been proposed and found to produce significant improvement.
Modified Bessel functions (BTO), has been used for this purpose by Filter
and Steinborn.92 The main properties of these BTO functions are related with
the computation of the integrals. A two center distribution can be written as a
multiple one-center distribution, whereas a four-center integral can be
considered the sum of two-center Coulomb integrals. These two features
together with the simple form of their Fourier transform allow to evaluate all
the required molecular integrals. At DZ level, BTOs yield almost the same

33
accuracy as STOs, while at SZ level, STO basis sets are significantly better
than BTOs.

Recently, a method called Coulomb operator resolutions was suggested which


enables the exponential type orbital translations to be completely avoided in
ab initio molecular electronic structure calculations.93 Coulomb operator
resolutions provide an excellent approximation that reduces the two electron-
multi center integrals to a sum of one electron overlap-like integral products,
which involve orbitals on at most two centers. The convergence using this
procedure has been shown to rapid in all cases.94 Four center STO molecular
integrals have been solved using this method for the H2 molecule dimer (van
der Waals complex) in a very recent report.95 Numerical values for the four
center terms in the H2 molecule dimer agree well with complete ab initio
results obtained using very large Gaussian basis sets. These recent works offer
promising breakthroughs in basis set theory and it is hoped that more
powerful basis sets based on STOs will be developed in future which will be
useful to reproduce the chemistry more accurately at a lower cost.

1.8.2 Gaussian type orbitals (GTO)


The use of STOs is still limited to atomic and diatomic systems where high
accuracy is required and to certain semi empirical methods where all the three
and four center integrals are neglected. Boys and McWeeny first advocated
the use of Gaussian-type basis functions on the practical ground that all of the
integrals required for a molecular calculation could be easily and efficiently
evaluated.96,97 The functional form of a Gaussian Type Orbital (GTO) in
cartesian coordinates is
χ(x, y, z; ζ, i, j, k) = N ª « ! ¬ • ]ζœ_
(1.33)
(1.33)
(1.33)

where, N is a normalization constant, ζ is called " the exponent". The i, j, and


k are not quantum numbers but non-negative integers (indices) that dictate the
nature of the orbital in a cartesian sense and = ª +« +¬ . The sum of the

34
indices, L=i+j+k, is used analogously to the angular momentum quantum
number for atoms, to mark the functions as s-type (L=0), p-type (L=1), d-type
(L=2) f-type (L=3) etc. When all three of these indices are zero, the GTO has
spherical symmetry, and is called an s-type GTO. When exactly one of the
indices is one, the function has axial symmetry about a single cartesian axis
and is called a p-type GTO. There are three possible choices named as px, py
and pz GTO s. When the sum of the indices is two, the orbital is called a d-
type GTO. But here there are six possible combinations of index values (i,j,k)
that can sum to two. This leads to six cartesian pre-factors x2, y2, z2, xy, xz
and yz. These six functions are called the cartesian d functions. In the solution
of the Schrödinger equation for the H atom, only five functions of the d-type
are required to span all possible values of the z-component of orbital angular
momentum for l=2. These functions are usually xy, xz, yz, (x2 - y2) and (2z2 -
x2 - y2). The first three of these canonical d functions are common with the
cartesian d functions, while the later two can be derived as linear
combinations of the cartesian d functions. A remaining combination that can
be formed from the cartesian d functions is ª +« +¬ , which is actually an
s-type GTO (since it has spherical symmetry). Different Gaussian basis sets
adopt different conventions with respect to their d functions; some use all six
Cartesian d functions, others prefer to reduce the total basis set size and use
the five linear combinations. The extra function is used to represent an s
orbital of higher angular momentum than that of the d functions (since this
combination has the same exponent as the other d-type functions and a d
orbital is more diffuse than an s orbital). Examination of f-type functions
shows that there are 10 possible cartesian Gaussians, which introduce 4px, 4py
and 4pz type contamination. However, there are only 7 linearly independent f-
type functions. This is a major headache since some programs remove these
spurious functions and some do not. Of course, the results obtained with all

35
possible cartesian Gaussians will be different from those obtained with a
reduced set.98

Cartesian GTOs are used in molecular calculations because the multicenter


integrals are easily evaluated due to the Gaussian theorem that allows to
express the product of two Gaussian functions centered in two different points
of space as another Gaussian function centered in a third point located on the
line that joins the two initial points.

]
The main difference of a GTO from an STO is that , the pre-exponential
factor, is dropped, the r in the exponential function is squared, and angular
momentum part is a simple function of cartesian coordinates. Calling these
functions GTOs is probably a misnomer, since they are not really orbitals.
They are simpler functions. In recent literature, they are frequently called
]
Gaussian primitives. The absence of factor restricts single Gaussian
primitive to approximating only 1s, 2p, 3d, 4f ... orbitals. It was done for
practical reasons, namely, for fast integral calculations. However,
combinations of Gaussians are able to approximate correct nodal properties of
atomic orbitals by taking them with different signs. Because a Gaussian has
the wrong behavior both near the nucleus and far from the nucleus, it was
clear that many more Gaussians would be required to describe an atomic
orbital than if Slater- type orbital (STO) basis functions were used. To ease
this problem, several Gaussian primitives are often grouped together to form
what are known as contracted GTOs (CGTOs). The term contraction means
"a linear combination of Gaussian primitives {or more often called primitive
GTO (PGTO)s} to be used as basis function.". A CGTO can be represented as

χ(CGTO) = ” χ (±²i³) (%. h5)

36
Such a basis function will have its coefficients and exponents fixed { ” and ζi
in χ (±²i³)}. Early work on contracted functions dealt with Gaussian lobe
functions in which, for example, a ‘p’ orbital was approximated by
differences of ‘s’ orbitals slightly displaced from each other.99 For higher
angular momentum, the number of terms required, and the loss of accuracy in
computing the integrals, made the method unmanageable. Clementi and Davis
extended the use of “contracted” functions to include Cartesian Gaussians.100
Although individual integrals with Cartesian Gaussians are somewhat more
complicated than with Gaussian lobes, the ease in extending the basis set to
higher angular momentum has made this procedure the method of choice.
Fortunately, the ratio of the number of Gaussians to the number of STO’s
required to obtain comparable accuracy is not as large as originally believed.
Although 4 Gaussians are needed to get within 1 mH (1 mH = 6.27 kcal/mol)
of the exact energy for the hydrogen atom, for atoms further down the
periodic table, such as argon, the ratio is about 2.61. CGTOs have other
advantages as well and on practical grounds they have been used extensively.
So a detailed discussion of CGTOs is presented later in this section.

An important task in the use of GTOs (whether they are used as such or as
CGTOs) is to derive a good set of PGTOs. Gaussian primitives are usually
obtained from quantum calculations on atoms (i.e. Hartree-Fock or Hartree-
Fock plus some correlated calculations, e.g. CI). Typically, the exponents (ζ)
are varied until the lowest total energy of the atom is achieved (as in the
optimization of STOs). The first optimized Gaussian set for atomic SCF
energies was published by Huzinaga.101 Later, van Duijneveldt extended
Huzinaga's work increasing the sizes of the basis set analyzed up to (14s
9p).102 Ever since, over the years several GTO sets have been optimized. The
related papers are consolidated in a table elsewhere.103 The methodology used
in these works was based on the variational determination of the exponents

37
with respect to the HF atomic energies. The (10s 6p) and (14s 9p) basis imply
a space of 16 and 23 dimensions, respectively, in which a local energy
minimum is looked for. The difficulties found in the optimization of
individual exponents of large basis sets are associated to the smoothness or
non existing minima of the energy surface. Therefore, no individually
optimized exponent sets larger than van Duijneveldt’s have been produced.

1.8.2.1 Even-tempered basis sets

When the basis set becomes large, the optimization problem is no longer easy.
The basis functions start to become linearly dependent and the energy
becomes a very flat function of the exponents. Furthermore, the multiple local
minima problem is encountered. An analysis of the basis sets which have
been optimized by variational methods reveals that the ratio between two
successive exponents is nearly a constant. Taking this ratio to be constant
reduces the optimization problem to only two parameters for each type of
basis function independently of the size of the set. Such basis sets were
labeled even-tempered basis sets with the ith exponent given as ζ = FI
where F ”D´ I are optimized for a given type of function (s, p, d etc.) and
nuclear charge (or atom).104 The advantage of such basis sets is that it is easy
to generate a sequence of basis sets which are guaranteed to converge towards
a complete basis. This is useful if the attempt is to extrapolate a given
property to the basis set limit. It was later discovered that the optimum
F ”D´ I costants to a good approximation can be written as functions of the
size of the basis set, N (the number of primitives) allowing extrapolation to
very large basis sets without the need for re-optimization.105

1.8.2.2 Well-tempered basis sets


Even tempered basis sets have the same ratio between exponents over the
whole range. From chemical considerations it is usually preferable to cover
the valence region better than the core region. This may be achieved by well-

38
tempered basis sets.106 The idea is similar to the even-tempered basis sets, the
exponents are generated by a suitable formula containing only a few
parameters to be optimized. The exponents in a well-tempered basis of size N
(number of PGTOs) are generated as
δ
ζ = FI ]
µ1 + ¶ · ¸ ¹, where, i = 1, 2,…,N (1.35)
1

The four parameters are F, I, ¶ ”D´ δ are optimized for each atom. The
exponents are the same for all types of angular momentum functions. Thus s,
p, d and higher angular momentum functions have the same radial part.

1.8.2.3 Universal basis sets


Conventionally the basis sets employed in electronic structure calculations
have a small number of functions in order to restrict the computational
requirements. However, for achieving high accuracy the basis set size should
be large to provide the much needed flexibility. Silver and Nieuwpoort proved
that optimization becomes a less important step in deriving larger basis sets.
Thus they introduced the concept of universal basis set as the basis set that
might satisfactorily describe several atoms.107 Following this idea Wilson and
coworkers suggested the use of even-tempered STO and GTO basis sets large
enough to reproduce the atomic energies of the elements belonging to the first
row.108 For example, in their initial work they generated a universal STO
basis set, nine ‘s’ and six ‘p’ functions, for the B to Ne atoms. The atomic
energies obtained thus have an accuracy of 0.5-1.5 mhartrees. The great
advantage of these universal basis sets is the transferability of the integrals.
As a consequence the use of the same basis set for each atom in the molecular
system and the computation and storage of integrals is simplified and reduced.
Such universal Gaussian basis sets may substantially reduce the
computational work required for the calculation of molecular integrals in ab
initio MO calculations.

39
1.8.2.4 Geometrical basis sets
Clementi and Corongiuz combined the ideas of even-tempered exponents and
universal basis sets along with six constraints to produce a new type of basis
set, which they called “geometrical”, for use in large molecule calculations.109
All atoms from H to Sr are represented by the same set of exponents
(although differing numbers of s, p, and d functions appear). In a generator
coordinate version of the HF equations, the obtained equations are
numerically integrated. From the discretization points used to perform this
integration Trisic et al. defined the exponents of the Gaussian basis sets. Since
there is no need for a variational optimization of the basis functions, the same
discretization points are valid for different atoms, and from them universal
Gaussian basis sets are defined. The accuracy obtained for the atomic energies
of Li to Ne atoms varies between 24 and 100 mhartrees, with respect to
corresponding STO basis sets.

1.8.2.5 Contracted basis sets


One disadvantage of all energy optimized basis sets is the fact that they
primarily depend on the wavefunction in the region of the inner shell
electrons. The 1s electrons account for a large part of the total energy and
minimizing the energy will tend to make the basis set optimum for the core
electrons, and less than optimum for the valence electrons. However,
chemistry is mainly dependent on the valence electrons. Furthermore, many
properties (for example, polarisability depend on the wavefunction “tail” (far
from the nucleus), which energetically, is unimportant. This fact that many
basis functions go into describing the energetically important, but chemically
unimportant, core electrons is the foundation for contracted basis sets.
Contraction is especially useful for orbitals describing the inner (core)
electrons, since they require a relatively large number of functions for
representing the wavefunction cusp near the nucleus, and further more are

40
largely independent of the environment. Contracting a basis set will always
increase the energy, since it is a restriction of the number of variational
parameters and makes the basis set less flexible, but will reduce the
computational cost significantly. A CGTO is defined as a linear combination
of PGTOs (see equation.1.34). The molecular orbitals (the initial guess at the
molecular wavefunction as required by the Roothan- Hall procedure) will be
chosen as linear combinations of the CGTOs whose coefficients will be
obtained in the electronic molecular calculation.

” χ (º²i³) (%. hq)


χ(MO) =

However, neither the exponents of the PGTOs nor the coefficients involved in
the CGTO definition are modified during the calculation. The cost of a
computational calculation depends on the number of PGTOs and CGTOs in
the following way. For gradient searches for optimum structures (geometry
optimization) the cost can be dominated by the time to do the basic integrals.
The computational effort (i.e. "CPU time") for calculating integrals in the
Hartree-Fock procedure depends upon the 4th power in the number of PGTOs.
This advocates smaller primitive sets for geometry optimization. Calculations
of the spectrum and other properties by perturbation theory or configuration
interaction depend on the number of CGTOs, but are relatively independent of
the number of PGTOs. Hence, in this step there is a strong motivation for
using larger numbers of primitives to produce better contracted basis
functions. Also, the storage required for integrals (when direct SCF is not
used) is proportional to the number of basis functions i.e. CGTOs (not
primitives). Frequently the disk storage and not the CPU time is a limiting
factor. So it is always advisable to do the geometry search with a smaller
basis set and then calculate the energy, at the selected geometries, with a more
elaborate basis.

41
Two main schemes of contraction have been proposed: the segmented
contraction110 and the general contraction.111 In segmented contraction, every
PGTO only contributes to one CGTO, which implies that each PGTO has
only a significant weight in the description of only one atomic orbital. In the
general contraction scheme, however, all the given PGTOs in the basis set of
a given ‘l’ symmetry contribute to all the CGTOs of this symmetry. The
segmented contractions are far more popular than general contractions. The
reason for their popularity is not that they are better, but simply, that the most
popular ab initio packages do not implement efficient integral calculations
with general contractions. The computer code to perform integral calculations
with general contractions is much more complex than that for the segmented
case.

1.8.2.6 Segmented contractions

The segmented basis sets are usually structured in such a way that the most
diffuse primitives (primitives with the smallest exponent) are left
uncontracted (i.e. one primitive per basis function). More compact primitives
(i.e. those with larger exponents) are taken with their coefficients from atomic
Hartree-Fock calculations and one or more contractions are formed. Then the
contractions are renormalized. The primitives are chosen from different shells
of functions to which the Cartesian Gaussians are grouped. The s-shell is a
collection of s type Gaussians; p-shell is a collection of p-type Gaussians; d-
shell is a collection of d-type Gaussians; and so on. Of course, combining
primitives belonging to different shells within the same contraction does not
make sense because primitives from different shells are orthogonal. But many
basis sets use the same exponents for functions corresponding to the same
principal quantum number, i.e., electronic shell. For the group of basis
functions in which s and p type functions share the same exponents, the term
sp shell is used.

42
The early Gaussian contractions were obtained by a least square fit to STOs.
The number of CGTOs (not primitives) used for representing a single STO
(i.e. zeta) was a measure of the goodness of the set. Based on the number of
CGTOs used for representing a single STO, the basis sets can be classified as
minimal or single zeta (SZ), double zeta (DZ), triple zeta (TZ), quadruple zeta
(QZ), etc. (similar to the classification of STO basis sets).

In the minimal basis set (i.e. SZ) only one basis function (CGTO) per STO is
used. DZ sets have two CGTOs per STO, TZ sets have three CGTOs per STO
and so on. Since valence orbitals of atoms are more affected by forming a
bond than the inner (core) orbitals, more CGTOs are assigned frequently to
describe valence orbitals. This prompted development of split-valence (SV)
basis sets, i.e., basis sets in which more CGTOs are used to describe valence
orbitals than core orbitals. That more basis functions are assigned to valence
orbitals does not mean the valence orbitals incorporate more primitives
(PGTOs). Frequently, the core orbitals are long contractions consisting of
many primitive Gaussians to represent well the "cusp" of s type function at
the position of the nucleus. Split valence basis sets can further be categorized
as valence double zeta (VDZ), valence triple zeta (VTZ) etc. based on
number of CGTO s used to represent the valence orbitals. For example, the
VDZ basis set uses one CGTO per core electron and two CGTOs per valence
electron.

1.8.2.7 Polarization functions

Most Gaussian primitive sets are constructed by optimization of the Hartree-


Fock energy of the atom. As already remarked, this choice will place heavy
emphasis on representing the core orbitals, as these orbitals contribute most of
the total energy of an atom. The most obvious defect is that atomic
calculations only define functions with the same L as the occupied orbitals,
e.g., s and p for carbon. But during molecule formation these atomic orbitals

43
undergo distortion. As atoms are brought close together, their charge
distribution causes a polarization effect (the positive charge is drawn to one
side while the negative charge is drawn to the other) which distorts the shape
of the atomic orbitals. In this case, 's' orbitals begin to have a little of the 'p'
flavor and 'p' orbitals begin to have a little of the 'd' flavor. This is not
satisfactorily represented using atomic Gaussian sets. For use in a molecular
calculation, these sets may be supplemented with additional functions of
higher angular momentum.
momentum Such functions are called polarization functions.

(a)

(b)

Figure 1. 3: (a) mixing of ‘s’ and ‘p’ (b) mixing of ‘p’ and ‘d’

The exponents for polarization functions cannot be derived from HF


calculations for the atom, since they are not populated. However, they can be
estimated from variational calculations on atoms with correlated
wavefunctions or from variational HF calculations on molecular systems
where the HF energy depends on polarization
polarization functions. In practice, however,
these exponents are also estimated "using well established rules of thumb or
by explicit optimization".

Augmentation with polarization functions is denoted by the letter ‘P’ in the


"zeta" terminology. Thus,
Thus DZP means double-zeta
zeta plus polarization, TZP
stands for triple-zeta
zeta plus polarization, etc. Sometimes the number of

44
polarization functions is given, e.g. TZDP, TZ2P, TZ+2P stands for triple-
zeta plus double polarization.

1.8.2.8 Diffuse functions

In chemistry, we are mainly concerned with the valence electrons which


interact with other molecules. However, many of the basis sets we have talked
about previously concentrate on the main energy located in the inner shell
electrons. This is the main area under the wave function curve. In the graphic
below, this area is that to the left of the dotted line. Normally the tail (the area
to the right of the dotted line), is not really a factor in calculations.

Figure 1.4: General nature of a plot of wavefunction vs distance from the nucleus

However, when an atom is in an anion or in an excited state, the loosely bond


electrons, which are responsible for the energy in the tail of the wave
function, become much more important. To compensate for this area, we can
use diffuse functions. These basis sets utilize very small exponents to clarify
the properties of the tail.

45
Figure 1.5: The effect of adding diffuse functions on an‘s’ and on a ‘p’ orbital.

1.8.2.9 STO-nG Basis sets


The most popular minimal basis sets are the STO-nG
STO nG (where n denotes the
number of primitives in the contraction) developed by Pople and coworkers.
These sets were obtained by least square fit of the combination of n Gaussian
functions to a Slater type orbital of the same type with ζ = 1.0, For this set
additional constraint is used, that exponents of corresponding gaussian
primitives are the same for basis functions describing orbitals with the same
principal quantum number (e.g. the same primitives are used for 2s and 2p
function). Then, these exponents are multiplied by the square of zeta in the
Slater orbital which described best the set of molecules. The STO-3G
STO (i.e. 3
primitives per each function) is the most widely used set. Originally
formulated for first row elements112 it was later extended to second
second-row113,
third row114 andd fourth row115 main group elements. It has also been applied
row transition metals.116 The main attraction of this basis,
to first- and second-row
other than its small size, is its effectiveness in predicting geometries. Pople
reports that the mean absolute deviation
deviation from experiment for SCF bond
lengths in several dozen molecules containing H, C, N, O, and F is 0.030 A0.
Moreover, the literature contains hundreds of other comparisons with
experiment, most of which fall in a similar range. Such agreement may appear
remarkable for a minimal basis, until it is realized that the large STO-3G
STO basis

46
set superposition error helps cancel other defects to produce reasonable bond
lengths.
Notation used for other contractions

For other sets a more complicated notation needs to be used to specify the
number of primitives and contractions explicitly. For each atom in the
molecule, the parentheses ( ) embrace the number of primitives that are given
in the order of angular momentum quantum number. Square brackets [ ] are
used to specify the number of resulting contractions. For example: (12s,9p,1d)
means 12 primitives on s-shell, 9 primitives on p-shell, and 1 primitive on d-
shell. This is sometimes abbreviated even further by skipping the shell
symbols (12,9,1). The [5,4,1] means that s-shell has 5 contractions, p-shell
has 4 contractions and d-shell has 1 contraction. To denote how contractions
were performed, the following notation is frequently used: (12,9,1) [5,4,1]
or (12,9,1)/[5,4,1] or (12s,9p,1d) [5s,4p,1d]. This means that 12 s-type
primitives were contracted to form 5 s-type contractions, 9 p-primitives were
contracted to 4 basis functions and 1 d-primitive was used as a basis function
by itself. The statement “9 p-primitives were contracted to 4 basis functions"
actually means that 12 basis functions were created. Each p-type basis
functions has 3 variants: px ,py and pz which differ in their cartesian part (i.e.,
angular part). The same is true for d-, f-, and higher angular momentum
functions. The notation above does not say how many primitives are used in
each contraction. The more elaborate notation explicitly lists the number of
primitives in each contraction. For example: (63111,4311,1) means that there
are 5 s-type contractions consisting of 6, 3, 1, 1 and 1 primitives, respectively.
The p-shell consists of 4 basis functions with 4, 3, 1 and 1 primitives, and d-
shell has 1 uncontracted primitive.

47
1.8.2.10 Pople-style basis sets

A large number of multiple zeta basis sets were also developed by Pople and
coworkers with a view to overcoming the shortcomings of the STO-nG basis
sets. A different convention was adopted in naming these basis sets. The basis
set structure is given for the whole molecule, rather than for a particular atom.
The notation also emphasizes the split valence (SV) nature of these sets.
Symbols like n-ijG or n-ijkG are used which can be encoded as: n - number
of primitives for the inner shells; ij or ijk - number of primitives for
contractions in the valence shell. The ij notations describe sets of valence
double zeta quality and ijk describe sets of valence triple zeta quality.
Generally, Pople-style basis sets, the s and p contractions belonging to the
same "electron shell" (i.e. corresponding formally to the same principal
quantum number n) are folded into a sp-shell. In this case, number of s-type
and p-type primitives is the same, and they have identical exponents.
However, the coefficients for s- and p-type contractions are different.

For example, in the 4-31G basis, for hydrogen (hydrogen has only valence
electrons!) (4s) [2s], for first row atoms (8s,4p) [3s,2p] and for 2nd
row (12s,8p) [4s,3p] contraction scheme is adopted.

Pople's basis sets augmented with polarization functions are labeled by adding
asterisk(s) (*). For example, if polarization functions are added to only heavy
atoms (other than H atoms) the n-ijG* or n-ijkG* notation is used. If H atoms
are also augmented with p-functions n-ijG** or n-ijkG** notation is used.
Similarly, augmenting with diffuse functions are represented using single ‘+’
(if diffuse functions are added only to heavy atoms) and double ‘++’ (if
diffuse functions are added to all atoms). The notation n-ij+G, or n-ijk+G is
used when 1 diffuse s-type and p-type Gaussians with the same exponents are
added to a standard basis set on heavy atoms. The notation n-ij++G, or n-

48
ijk++G is used when 1 diffuse s-type and p-type Gaussians on heavy atoms
and 1 diffuse s-type Gaussian on hydrogens are included. In short, the
symbol, 6-31++G** can be encoded as a Pople-style DZ basis set augmented
with polarization functions and diffuse functions on all atoms. Further, a
CGTO comprising of 6 PGTOs is used for the core orbitals and two CGTOs
(one a contraction of 3 PGTOs and the other uncontracted) are used for the
valence orbitals.
Some of the Pople-style split valence basis sets are listed here with a view to
highlighting the original papers. They include 4-21G117, 3-21G (for the first
row118 and the second row119 elements), 3-21G*120, 3-21+G and 3-21++G121,
4-31G (for first row122 and second row123 elements), 4-31+G,124 6-31G125,
6-31G*126 (for first and second row elements), 6-31+G127, 6-31+G*128,
6-311G**129
Table 1.2: Number of basis functions [CGTOs] and number of PGTOs for the
molecule C3H4ClLi for the Pople-style basis set 6-31+G(d).

No. of basis functions No.of basis functions


Total CGTOs per each

(CGTOs) due to (PGTOs) due to


No of atoms in the

No of atoms in the
Occupied subshell

Polarization

Total PGTOs
type of atoms
Diffuse

Diffuse
Polar
Total

Total
VDZ

VDZ
molecule

molecule
Atom

H 1s 2 0 0 2 4 8 4 0 0 4 4 16
C 1s,2s,2p 9 6 4 19 3 57 22 6 4 32 3 96
Li 1s,2s,2p 9 6 4 19 1 19 22 6 4 32 1 32
Cl 1s,2s,2p,3s,3p 13 6 4 23 1 23 46 6 4 56 1 56
Grand Total 107 Grand Total 200

High quality contracted basis sets other than Pople-style basis sets are also
available in literature. They include the Dunning DZP130, Dunning/Hay VDZ,
MINI-i, MIDI-i, MAXI-I and Mc Lean/Chandler basis sets.131 Over the years
several basis sets for transition metals have also been developed. Contracted
basis sets developed prior to 1986 are reviewed by Davidson and Feller.132

49
1.8.2.11 General contractions
The Atomic Natural Orbitals (ANO) and correlation consistent (cc) basis sets
are of the general contraction scheme.
Atomic natural orbitals
One popular way of obtaining general contraction coefficients is from Atomic
Natural Orbital (ANO).133 The idea in these basis sets is to contract a large
PGTO set to a fairly small number of CGTOs by using natural orbitals from a
correlated calculation on the free atom (typically, at the CISD level). The
natural orbitals are those which diagonalize the density matrix and the eigen
values are called orbital occupation numbers (number of electrons in the
orbital). For an RHF calculation, ANOs would be identical to the canonical
orbitals with occupation numbers 0 or 2. When a correlated wavefunction is
used the occupation numbers may have any value between 0 and 2.The ANO
contraction selects the important combination of the PGTOs from the
magnitude of the occupation numbers. The main feature of ANO contraction
is that it more or less “automatically” generates balanced basis sets.
Correlation consistent (cc) basis sets
The primary disadvantage of ANO basis sets is that a very large number of
PGTOs is necessary for converging towards the basis set limit. But cc basis
sets134 yield results comparable to those for the ANO sets with smaller sets of
primitives. The cc part of the name implies that the exponents and contraction
coefficients are variationally optimized not only for HF calculations but for
calculations including electron correlation. These basis sets are designed such
that functions which contribute similar amounts of correlation energy are
included at the same stage, independently of the function type. For example,
the first ‘d’ function provides a large energy lowering, but the contribution
from a second ‘d’ function is similar to that from the first ‘f’ function. The
energy lowering from a third ‘d’ function is similar to that from the second ‘f’
function and the first ‘g’ function. Addition of polarization functions should

50
be done in the order 1d, 2d1f, 3d2f1g and so on. Several different types of cc
basis sets are available. Two examples are listed in the following table.
Table 1.3: No. of CGTOs and PGTOs in cc basis sets
Number of Number of
Basis set Acronym PGTOs - Heavy CGTOs - Heavy
atom/H atom atom/H atom
Correlation consistent polarized
cc-pVDZ 9s,4p,1d/4s,1p 3s,2p,1d/2s,1p
valence double zeta
Correlation consistent polarized
cc-pDTZ 10s,5p,2d,1f/5s,2p,1d 4s,3p,2d,1f/3s,2p,1d
valence triple zeta

A step up in terms of quality increases each type of function by one and adds
a new type of higher order polarization function. They do not give good
description of the anions, but this problem has been solved by enlarging the
basis set with additional diffuse functions. The energy optimized cc basis sets
can be augmented by additional diffuse functions,135 adding the prefix aug- to
the acronym. The augmentation consists of adding one extra function with a
smaller exponent for each angular momentum function. Thus, aug-cc-pVDZ
has additionally 1s, 1p and 1d functions while, aug-cc-pVTZ has additionally
1s, 1p, 1d and 1f functions.

The cc basis sets may also be augmented by additional tight functions (with
large exponents) if the interest is in recovering core-core and core-valence
electron correlation. For example,
cc-pCVDZ has additionally 1s and 1p tight functions,
cc-pCVTZ has additionally 2s, 2p and 1d tight functions,
cc-pCVQZ has additionally 3s, 3p, 2d and 1f tight functions and
cc-pCV5Z has additionally 4s, 4p, 3d, 2f and 1g tight functions.136

The main advantage of ANO and cc basis sets is the ability to generate a
sequence of basis sets which converges towards the basis set limit. The major
difficulty in using the cc basis sets is that each step up in quality almost
doubles the number of basis functions.

51
1.8.3 Basis set superposition error (BSSE)

The non-completeness of the basis sets employed in molecular calculations


has a significant impact in the computation of molecular dissociation and
atomization energies. This non-completeness of the basis sets arises from the
truncation of the basis set functions introduced by the LCAO approximation.
For illustrating BSSE, consider the dissociation
» /¼s
() ½¾¾¿ ( + )
The dissociation energy is usually given by the expression

À3 = 3 # − (3 + 3# ) (1.37)
(1.37)

For evaluating ∆3, three computations have to be done (one each for A, B and
AB). In AB, however, the basis set of A is available to B and that of B is
available to A. So A and B in AB, in effect, have bigger basis sets than do
isolated A and B. Thus 3 # is more accurately estimated than either 3 Á 3# .
Thus ∆3 is overestimated if an incomplete basis set is employed. This error
due the incompleteness of basis sets is termed as BSSE. There are two ways
to deal with BSSE – the counterpoise method137 and the saturation method. In
the counterpoise method the energies of isolated A and B are improved by the
introduction of extra basis functions, frequently known as ghost orbitals
(which are basis functions not accompanied by atoms – spirits without
bodies). This method gives only an approximate value of BSSE and is not
uniquely defined for species of more than two components. The second way
to handle BSSE is to saturate with basis functions. If the basis set in the
atomic calculation of A and B is large enough, the atomic energies will be
near the HF limit and the improvement of the basis set with other functions
will not modify the total energy. In this case, the basis set is named a
saturated basis set, thus eliminating BSSE. This also becomes cumbersome
for bigger molecules.

52
1.8.4 Two electron basis functions

Another innovation in the development of basis set chemistry is the


introduction of two electron basis functions. It is well known that the two
electron correlation ‘cusp’ (the occurrence of two electrons at the same point
in space) is treated poorly by expansions in products of one electron
functions. The limits of what can be achieved using one electron functions is
already approaching and one of the ways to overcome this difficulty is to
consider explicitly in the wavefunction two electron basis functions. Such
functions have a Gaussian dependence on the inter electronic distance
(Gaussian-type geminals). Persson and Taylor have shown that only a few
such functions are enough to dramatically reduce the error in the correlation
energy.138

1.9 Model Chemistry

A unique combination of a computational method and a basis set is called


Model Chemistry. The selection of an appropriate Model Chemistry is
the key to the success of a computational research work. It should be
guided by the different factors described in the previous sections of this
chapter.

53
Appendix 1

ATOMIC UNITS

A system of units that is widely used in atomic and molecular calculations is


atomic units (a.u.). In this system four basic units are defined as (equations
A.1 to A.4).
Â
ℏ= =1 (A.1)
Ã

= ”•• ÁÄ Åℎ Ç tÅ ÁD = 1 (A.2)

= tℎ” È ÁÄ Åℎ Ç tÅ ÁD = 1 (A.3)

4 =É mÅmÊmÅÅ« = 1 (A.4)

For example, the Hamiltonian operator, Ĥ, of an ‘n’ electron molecule


having ‘M’ nuclei in SI units is given by equation A.5 (see equation 1.2a).

−ℏ −ℏ −
Ĥ= ∇ + ∇ + +
2M 2 4 4 !
!"

+ (Ë. 6)
#
4 $ #
#"

This Hamiltonian can be converted to one in a.u. by substituting the four


basic units defined above (equation A.6; see also equation 1.2b).

−∇ − ∇ − 1
Ĥ= + + + + (Ë. q)
#
2M 2 ! $ #
!" #"

Ĥ in a.u. is independent of any physical constants. It turns out that all other
physical quantities can be expressed in terms of the four base units

(the bohr radius, ” ) in a.u. is given by equation A.7.


(equations A.1 to A.4). For example, the radius of the first orbit in H atom

” = = 1 ”. Ì Á 1 –Áℎ
2
4 0 ℏ
w 2
(A.7)

54
Similarly, the atomic unit of energy is called a hartree (denoted as H) and is
given by equation A.8.

= 1 ”. Ì. Á 1 -
w 4
3= _ 2
0 ℏ
_ (A.8)
Í

The atomic units bohr and hartree are called derived units because they are
expressed in terms of the four base units defined by A.1 to A.4.

One hartree can be expressed in other units such as kcal mol-1, kJ mol-1 & eV
(see A.9).

1 - = 625.51 Ît”Ç ÁÇ ] = 2627.38 Îd ÁÇ ] = 27.21 0 (A.9)

55
REFERENCES
[1] Mathematical Challenges from Theoretical/Computational Chemistry, http://
books.nap.edu/catalog/4886.html
[2] M. Born; J. R. Oppenheimer, Ann. Phys., 46, 618 (1934).
[3] D. R. Hartree, Proc. Cambridge Philos. Soc., 24, 328 (1928).
[4] V. A. Fock, Z.Phys., 15, 126 (1930).
[5] C. C. J. Roothan, “New developments in molecular orbital theory” Rev.Mod. Phys.,
23, 69-89 (1951).
[6] (a) K. Ruedenberg, “On the Three- and Four-center integrals in molecular quantum
mechanics” J. Chem. Phys., 19, 1433 (1951).
(b) K. Ruedenberg, “A study of two-center integrals useful in calculations on
molecular structure. II. The two-center exchange integrals” J. Chem. Phys., 19,
1459 (1951).
[7] J. A. Pople; R. K. Nesbet, “Self-Consistent orbitals for radicals” J. Chem. Phys., 22,
571 (1954).
[8] David P. Tew; Wim Klopper; Trygve Helgaker, “Electron Correlation: The many
body problem at the heart of chemistry” J. Comput. Chem., 28, 1307-1320 (2007).
[9] E. A. Hylleraas, “Über den Grundzustand des Heliumatoms” Z. Physik, 48, 46
(1928).
[10] A. Meckler, “Electronic energy levels of molecular oxygen” J. Chem. Phys., 21,
1750 (1953).
[11] J. B. Foresman; M. Head-Gordon; J. A. Pople; M. J. Frisch, “Toward a systematic
molecular orbital theory for excited states” J. Phys. Chem., 96, 135 (1992).
[12] J. A. Pople; R. Seeger; R. Krishnan, “Variational configuration interaction
methods and comparison with perturbation theory” Int. J. Quantum Chem., Suppl.
Y-11, 149 (1977).
[13] K. Raghavachari; H. B. Schlegel; J. A. Pople, “Derivative studies in configuration-
interaction theory” J. Chem. Phys., 72, 4654 (1980).
[14] K. Raghavachari; J. A. Pople, “Calculation of one-electron properties using limited
configuration-interaction techniques” Int. J. Quantum Chem., 20, 1067 (1981).
[15] S. R. Langhoff; E. R. Davidson, “Configuration interaction calculations on the
nitrogen molecule” Int. J. Quantum Chem., 8, 61 (1974).
[16] J. A. Pople; M. Head-Gordon; K. Raghavachari, “Quadratic configuration
interaction. A general technique for determining electron correlation energies”
J. Chem. Phys., 87, 5968 (1987).
[17] C. Møller; M. S. Plesset, “Note on an approximation treatment for many-electron
systems” Phys. Rev., 46, 618 (1934)
[18] M. Head-Gordon; J. A. Pople; M. J. Frisch, “MP2 energy evaluation by direct
methods” Chem. Phys. Lett., 153, 503 (1988).

56
[19] S. Saebø; J. Almlöf, “Avoiding the integral storage bottleneck in LCAO calculations
of electron correlation” Chem. Phys. Lett., 154, 83 (1989).
[20] M. J. Frisch; M. Head-Gordon; J. A. Pople,“Direct MP2 gradient method” Chem.
Phys. Lett., 166, 275 (1990).
[21] M. J. Frisch; M. Head-Gordon; J. A. Pople, “Semi-direct algorithms for the MP2
energy and gradient” Chem. Phys. Lett., 166, 281 (1990).
[22] M. Head-Gordon; T. Head-Gordon, “Analytic MP2 frequencies without fifth order
storage: theory and application to bifurcated hydrogen bonds in the water
hexamer” Chem. Phys. Lett., 220, 122 (1994).
[23] J. A. Pople; J. S. Binkley; R. Seeger, “Theoretical models incorporating electron
correlation” Int. J. Quantum Chem., Suppl. Y-10, 1 (1976).
[24] J. A. Pople; R. Seeger; R. Krishnan, “Variational configuration interaction methods
and comparison with perturbation theory” Int. J. Quantum Chem., Suppl. Y-11,
149 (1977).
[25] K. Raghavachari; J. A. Pople, “Approximate 4th-order perturbation-theory of
electron correlation energy” Int. J. Quantum Chem., 14, 91 (1978).
[26] K. Raghavachari; J. A. Pople; E. S. Replogle; M. Head-Gordon, “Fifth order Møller-
Plesset Perturbation Theory: Comparison of existing correlation methods and
implementation of new methods correct to fifth order” J. Phys. Chem., 94,
5579 (1990).
[27] G. W. Trucks; J. D. Watts; E. A. Salter; R. J. Bartlett, “Analytical MBPT(4)
Gradients” Chem. Phys. Lett., 153, 490 (1988).
[28] G. W. Trucks; E. A. Salter; C. Sosa; R. J. Bartlett, “Theory and implementation of
the MBPT Density Matrix: An application to One-Electron properties” Chem.
Phys. Lett., 147, 359 (1988).
[29] J. Čížek, J. Chem. Phys., 45, 4526 (1966).
[30] J. Čížek, “On the use of the cluster expansion and the technique of diagrams in
calculations of correlation effects in atoms and molecules” Adv. Chem. Phys., 14,
35 (1969).
[31] J. Čížek; J. Paldus, “Correlation problems in atomic and molecular systems III.
Rederivation of the coupled-pair many-electron theory using the traditional quantum
chemical methods” Int. J. Quant. Chem., 5, 359 (1971).
[32] J. A. Pople; R. Krishnan; H. B. Schlegel; J. S. Binkley, “Electron correlation
theories and their application to the study of simple reaction potential surfaces”
Int. J. Quantum Chem., 14, 545 (1978).
[33] D. Watts; G. W. Trucks; R. J. Bartlett, “Coupled-cluster, unitary coupled cluster and
MBPT(4) open-shell analytical gradient methods” Chem. Phys. Lett., 164, 502 (1989).
[34] G. D. Purvis; R. J. Bartlett, “A full coupled-cluster singles and doubles model–the
inclusion of disconnected triples” J. Chem. Phys., 76, 1910 (1982).
[35] G. E. Scuseria; C. L. Janssen; H. F. Schaefer III, “An efficient reformulation of the
closed-shell coupled cluster single and double excitation (CCSD) equations”
J. Chem. Phys., 89, 7382 (1988).

57
[36] G. E. Scuseria; H. F. Schaefer III, “Is coupled cluster singles and doubles (CCSD)
more computationally intensive than quadratic configuration-interaction(QCISD)?”
J. Chem. Phys., 90, 3700 (1989).
[37] J. Noga; R. J. Bartlett, “The full CCSDT model for molecular electronic structure”
J. Chem. Phys., 86, 7041 (1987).
[38] K. Raghavachari; G. W. Trucks; J. A. Pople; M. Head-Gordon, “A fifth-order
perturbation comparison of electron correlation theories” Chem. Phys. Lett., 157,
479 (1989).
[39] R. J. Bartlett, “Coupled-cluster approach to molecular structure and spectra: A step
toward predictive quantum chemistry” J. Phys. Chem., 93, 1697 (1989).
[40] T. Helgaker; T.A. Ruden; P. Jørgensen; J. Olsen; W. Klopper, “A priori calculation
of molecular properties to chemical accuracy” J. Phys. Organic Chem., 17, 913
(2004).
[41] M. Head-Gordon; R. J. Rico; M. Oumi; T. J. Lee, “A doubles correction to
electronic excited-states from configuration-interaction in the space of single
substitutions” Chem. Phys. Lett., 219,21 (1994).
[42] M. Head-Gordon; D. Maurice; M. Oumi, “A Perturbative correction to restricted
open-shell configuration-interaction with single substitutions for excited-states of
radicals” Chem. Phys. Lett., 246, 114 (1995).
[43] Y. Mochizuki; K. Tanaka, “Modification for spin-adapted version of configuration
interaction singles with perturbative doubles” Chem. Phys. Lett., 443, 389 (2007).
[44] D. Hegarty; M. A. Robb, “Application of unitary group-methods to configuration-
interaction calculations” Mol. Phys., 38, 1795 (1979).
[45] R. H. A. Eade; M. A. Robb, “Direct minimization in MC SCF theory - the Quasi-
Newton method” Chem. Phys. Lett., 83, 362 (1981).
[46] H. B. Schlegel; M. A. Robb, “MC SCF gradient optimization of the H2CO→ H2+CO
transition structure” Chem. Phys. Lett., 93, 43 (1982).
[47] M. J. Frisch; I. N. Ragazos; M. A. Robb; H. B. Schlegel, “An evaluation of 3 Direct
MC- SCF procedures” Chem. Phys. Lett., 189, 524 (1992).
[48] N. Yamamoto; T. Vreven; M. A. Robb; M. J. Frisch; H. B. Schlegel, “A Direct
derivative MC-SCF procedure” Chem. Phys. Lett., 250, 373 (1996).
[49] Frank Jensen, “Introduction to computational chemistry” John Wiley & Sons, pp
101-117 (1999).
[50] B. O. Ross; P. R. Taylor; P. E. M. Siegbahn, “ A complete active space SCF method
(CASSCF) using a density matrix formulated super-CI approach” Chem. Phys., 48,
157 (1980).
[51] J. Olsen; B. O. Roos; P. Jørgensen; H. J. A. Jensen, “Determinant based
configuration interaction algorithms for complete and restricted configuration
interaction spaces” J. Chem. Phys., 89, 2185 (1998).
[52] J. A. Pople; M. Head-Gordon; D. J. Fox; K. Raghavachari; L. A. Curtiss, “Gaussian
‐1 theory: A general procedure for prediction of molecular energies” J. Chem.
Phys., 90, 5622 (1989).

58
[53] L. A. Curtiss; K. Raghavachari; G. W. Trucks; J. A. Pople, “Gaussian-2 theory for
molecular energies of first- and second-row compounds” J. Chem. Phys., 94, 7221
(1991).
[54] L. A. Curtiss; K. Raghavachari; P. C. Redfern; V. Rassolov; J. A. Pople, “Gaussian-
3 (G3) theory for molecules containing first and second-row atoms” J. Chem. Phys.,
109, 7764 (1998).
[55] L. A. Curtiss; P. C. Redfern; K. Raghavachari, “Gaussian-4 theory” J. Chem. Phys.,
126, 084108 (2007).
[56] (a) N. J. De Yonker; T. R. Cundari; A. K. Wilson, “The correlation consistent
composite approach (ccCA): An alternative to the Gaussian-n methods” J. Chem.
Phys., 124, 114104 (2006).
(b) N. J. De Yonker; T. Grimes; S. Yockel; A. Dinescu; B. Mintz; T. R. Cundari; A.
K. Wilson, “The correlation-consistent composite approach: Application to the
G3/99 test set" J. Chem. Phys., 125, 104111 (2006).
[57] (a) J. W. Ochterski; G. A. Petersson; J. A. Montgomery, “A complete basis set model
chemistry. V. Extensions to six or more heavy atoms” J. Chem. Phys., 104,2598 (1996).
(b) J. A. Montgomery Jr.; M. J. Frisch; J. W. Ochterski; G. A. Petersson, “A
complete basis set model chemistry. VI. Use of density functional geometries and
frequencies” J. Chem. Phys., 110, 2822 (1999).
(c) J. A. Montgomery Jr.; M. J. Frisch; J. W. Ochterski; G. A. Petersson, “A
complete basis set model chemistry. VII. Use of the minimum population localization
method” J. Chem. Phys., 112, 6532 (2000).
[58] P. L. Fast; M. L. Sánchez; D. G. Truhlar, “Multi-coefficient Gaussian-3 method for
calculating potential energy surfaces” Chem. Phys. Lett., 306, 407, (1999).
[59] B. J. Lynch; D. G. Truhlar, “Robust and affordable multicoefficient methods for
thermochemistry and thermochemical kinetics: The MCCM/3 suite and SAC/3” J.
Phys. Chem. A, 107, 3898 (2003).
[60] L. A. Curtiss; K. Raghavachari; P. C. Redfern; J. A. Pople, “Assessment of
Gaussian-2 and density functional theories for the computation of enthalpies of
formation” J. Chem. Phys., 106, 1063 (1997).
[61] L. A. Curtiss; P. C. Redfern; K. Raghavachari; J. A. Pople, “Assessment of
Gaussian-3 and density functional theories for a larger experimental test set”
J. Chem. Phys., 112, 7374 (2000).
[62] L. A. Curtiss; P. C. Redfern; K. Raghavachari, “Assessment of Gaussian-3 and
density-functional theories on the G3/05 test set of experimental energies” J. Chem.
Phys., 123, 124107 (2005).
[63] C. J. Cramer, “Essentials of computational chemistry: Theories and Models” 2nd
Edn., John Wiley & Sons, p493 (2004).
[64] A. Venkatnathan; A. B. Szilva; D. Walter; R. J. Gdanitz; E. A. Carter, “Size
extensive modification of local multi-reference configuration interaction” J. Chem.
Phys., 120,1693 (2004).
[65] S. Chattopadhyay; U. S. Mahapatra; B. Datta; D. Mukherjee, “State-specific multi-
reference coupled electron-pair approximation like methods: Formulation and
molecular applications” Chem. Phys. Lett., 357, 426 (2002).

59
[66] J. A. R. Paul, “On the size consistency of multi-reference CEPA methods” Collection
of Czechoslovak Chemical Communications, 70, 638 (2005).
[67] T. Kato, “On the eigenfunctions of many-particle systems in quantum mechanics”
Comm. Pure Appl. Math., 10, 151 (1957).
[68] G. W. F. Drake; Z. Van, “Variational eigenvalues for the S states of helium” Chem.
Phys. Lett., 229, 486 (1994).
[69] W. Kolos; L. Wolniewicz, “Improved theoretical ground-state energy of the
hydrogen molecule” J. Chem. Phys., 49, 404 (1968).
[70] W. Kutzelnigg; W. Klopper, “Wave functions with terms linear in the interelectronic
coordinates to take care of the correlation cusp. I. General theory” J. Chem. Phys.,
94, 1985 (1991).
[71] T. Shiozaki; M. Kaniya; S. Hirata; E. F. Valeev, “Explicitly correlated coupled-
cluster singles and doubles method based on complete diagrammatic equations”
J. Chem. Phys., 129, 071101 (2008).
[72] T. Shiozaki; M. Kaniya; S. Hirata; E. F. Valeev, “Higher-order explicitly correlated
coupled-cluster methods” J. Chem. Phys., 130, 054101 (2009).
[73] M. J. Field; P. A. Bash; M. Karplus, “A combined quantum mechanical and
molecular mechanical potential for molecular dynamics simulations” J. Comput.
Chem., 11, 700 (1990).
[74] M. Svensson; S. Humbel; R. D. J. Froese; T. Matsubara; S. Sieber; K. Morokuma,
“ONIOM: A multilayered integrated MO + MM method for geometry optimizations
and single point energy predictions” J. Phys. Chem., 100, 19357 (1996).
[75] C. M. Quinn, “Computational quantum chemistry: An interactive guide to basis set
theory” Academic Press, p 17 (2002).
[76] J. C. Slater, “Atomic shielding constants” Phys. Rev., 36, 57 (1930).
[77] J. C. Slater, “Analytic atomic wave functions” Phys. Rev., 42, 33 (1932).
[78] C. Zener, “Analytic atomic wave functions” Phys. Rev., 36, 51 (1930).
[79] E. Clementi, “Tables of atomic functions” IBM J. Res. Develop. Suppl., (1965).
[80] E. Clementi and C. Roetti, At. Data Nucl. Data Tables, 14, 177 (1974).
[81] A.D. McLean and R.S. Mclean, At. Data Nucl. Data Tables, 26, 197 (1981).
[82] (a) T. Koga; H. Tatewaki; A. J. Thakkar, “Roothaan-Hartree-Fock wave functions
for atoms with Z≤54,” Phys. Rev. A, 47, 4510 (1993).
(b) T. Koga; Y. Seki; A.J. Thakkar; H. Tatewaki, “Roothaan-Hartree-Fock wave
functions for ions with N< or =54” J. Phys. B: At. Mol. Opt.Phys.,26,2529 (1993).
(c) T. Koga; S. Watanabe; K. Kanayama; R. Yasuda; A.J. Thakkar,“Improved
Roothaan-Hartree-Fock wave functions for atoms and ions with N<=54” J. Chem.
Phys., 103, 3000 (1995).
(d) C.F. Bunge; J.A. Barrientos; A.V. Bunge; J.A. Cogordan, “Hartree-Fock and
Roothaan-Hartree-Fock energies for the ground states of He through Xe” Phys.
Rev. A, 46, 3691 (1992).
[83] E. Francisco; L. Seijo; L. Pueyo, “The maximum overlap method: A general and

60
efficient scheme for reducing basis sets. Application to the generation of
approximate AO's for the 3d transition metal atoms and ions” J. Solid State Chem.,
63, 391 (1986).
[84] (a) J. M. G. Vega; J. F. Rico; J. I. F. andez-Alonso, “Quantum chemical applications
of the theory of the distance between subspaces” J. Mol. Struct. (THEOCHEM).,
184, 1 (1989).
(b) J. M. G. Vega; B. Miguel, “Approximate STO functions for the first-row
transition metal atoms” J. Comput. Chem., 12, 1172 (1991).
(c) J. M. G. Vega; B. Miguel, “Slater functions for Y to Cd atoms by the distance
between subspaces” J. Solid State Chem., 116, 275 (1995).
[85] (a) J.A. Sordo; L. Pueyo, “A quality test for the Hartree-Fock-Roothaan SCF wave
functions” Int. J. Quantum Chem., 28, 687 (1985).
(b) J .A. Sordo; L. Pueyo, “High-quality Hartree-Fock-Roothaan wave functions for
molecular calculations” J. Mol. Struct. (THEOCHEM) 120, 9 (1985).
[86] J. R. Mohallem; M. Trsic, “A universal Gaussian basis set for atoms Li through Ne
based on a generator coordinate version of the Hartree-Fock equations” J. Chem.
Phys., 86, 5043 (1987).
[87] (a) J. M. G. Vega; B. Miguel, “Single-exponent Slater function expansions of the He
atom 1s orbital and its isoelectronic series” Chem. Phys. Lett., 207, 270 (1993).
(b) J. M. G. Vega; B. Miguel, “Orbitals expanded in Slater functions with single-
exponent by shell and by subshell” Int. J. Quantum Chem., 51, 397 (1994).
[88] T. Koga; K. Kanayana; A. J. Thakkar, “Theoretical and computational
developments: Noninteger principal quantum numbers increase the efficiency of
Slater-type basis sets” Int. J. Quantum Chem., 62, 1 (1997).
[89] T. Koga; Y. Seki; A. J. Thakkar; H. Tatewaki, “Noninteger principal quantum
numbers increase the efficiency of Slater-type basis sets: Singly charged cations and
anions” J. Phys. B: At. Mol. Opt. Phys., 30, 1623 (1997).
[90] T. Koga; K. Kanayana, “Noninteger principal quantum numbers increase the
efficiency of Slater-type basis sets: heavy atoms” Chem. Phys. Lett., 266, 123
(1997).
[91] T. Koga; J. M. G. Vega; B. Miguel, “Double-zeta Slater-type basis sets with
noninteger principal quantum numbers and common exponents” Chem. Phys. Lett.,
283, 97 (1998).
[92] E. Filter; E.O. Steinborn, “Extremely compact formulas for molecular two-center
one-electron integrals and Coulomb integrals over Slater-type atomic orbitals”
Phys. Rev. A, 18, 1 (1978).
[93] S. A. Vagranov; A. T. B. Gilbert; E. Duplaxes; P. M. W. Gill, “Resolutions of the
Coulomb operator” J. Chem. Phys., 128, 201104 (2008).
[94] P. M. W. Gill: A. T. B. Gilbert, “Resolutions of the Coulomb operator: II. The
Laguerre Generator” Chem. Phys., 86, 356 (2009).
[95] P. E. Hoggan, “Four-center Slater-type orbital molecular integrals without orbital
translations” Int. J. Quant. Chem., 110, 98 (2010).

61
[96] S. F. Boys, “Electronic wave functions. I. A general method of calculation for the
stationary states of any molecular system” Proc. R. Soc. London, A200, 542 (1950).
[97] R. McWeeny,“Gaussian approximations to wave functions” Nature, 166, 21 (1950).
[98] C. J. Cramer, “Essentials of computational chemistry: Theories and models” 2nd
Edn., John Wiley & Sons, p167 (2004).
[99] (a) E. Clementi, “ab initio computations in atoms and molecules” IBM J. Res. and
Dev., 9, 2 (1965).
(b) J. L. Whitten, “Gaussian Lobe function expansions of Hartree-Fock solutions
for the First‐Row atoms and Ethylene” J. Chem. Phys., 44, 359 (1966).
[100] E. Clementi; D. R. Davis, “Electronic structure of large molecular systems”
J. Comput. Phys., 1, 223 (1967).
[101] S. Huzinaga; “Gaussian-Type functions for Polyatomic systems” J. Chem. Phys.,
42, 1293 (1965).
[102] F.B. van Duijneveldt, IBM Technical Research Report, RS 945 (1971).
[103] L. A. Montero; L. A. Diaz and R. Bader (eds.), “Introduction to advanced topics of
computational chemistry” Editorial de la Universidad de La Habana, p 52 (2003).
[104] (a) C. M. Reeves; M. C. Harrison, “Use of Gaussian Functions in the calculation of
wave functions for small molecules. II. The Ammonia molecule” J. Chem. Phys., 39,
11 (1963).
(b) M. W. Schmidt; K. Ruedenberg, “Effective convergence to complete orbital
bases and to the atomic Hartree–Fock limit through systematic sequences of
Gaussian primitives” J. Chem. Phys., 71, 3951 (1979).
[105] (a) D. Feller; K. Ruedenberg, “Cluster expansion of the wave function” Theor.
Chim. Acta, 71, 231 (1979).
(b) M. W. Schmidt; K. Ruedenberg,. “Effective convergence to complete orbital
bases and to the atomic Hartree-Fock limit through systematic sequences of
Gaussian primitives” J. Chem. Phys., 3951, 71 (1979).
[106] (a) S. Huzinaga; M. Klobukowski; H. Tatewaki, “The well-tempered GTF basis sets
and their applications in the SCF calculations on N2, CO, Na2, and P2” Can. J.
Chem., 63, 1812 (1985).
(b) S. Huzinaga; M. Klobukowski; H. Tatewaki, “Well-tempered GTF basis sets for
the atoms K through Xe” Chem. Phys. Lett., 120, 509 (1985)
(c) S. Huzinaga; B. Miguel, “A comparison of the geometrical sequence formula
and the well-tempered formulas for generating GTO basis orbital exponents” Chem.
Phys. Lett., 175,289 (1990).
(d) S. Huzinaga; B. Miguel, “A comparison of the geometrical sequence formula
and the well-tempered formulas for generating GTO basis orbital exponents” Chem.
Phys. Lett., 75, 289 (1990).
(e) M. Klobukowski, “Systematic sequences of well-balanced Gaussian basis sets”
Can. J. Chem., 72, 1741 (1994).
[107] D. M. Silver; W. C. Nieuwpoort, “Two-centre model potential energy calculations
for the 2Σ and 2Π states of Li+2, Na+2, K+2, Rb+2 and Cs+2” Chem. Phys. Lett.,
57, 421 (1978).
[108] (a) D. M. Silver; S. Wilson; W.C. Nieuwpoort, “Universal basis sets and transf-

62
erability of integrals” Int. J. Quantum Chem., 14, 635 (1978).
(b) S. Wilson, “Systematic sequences of even-tempered Gaussian primitives in
electron correlation studies using many-body perturbation theory” Theor. Chim.
Acta, 57, 53 (1980).
(c) S. Wilson, “Universal systematic sequence of even-tempered Gaussian primitive
functions in electronic correlation studies” Theor. Chim. Acta, 58, 31 (1980).
[109] (a) E. Clementi; G. Corongiu, “Geometrical basis set for molecular computations”
Chem. Phys. Lett., 90, 359 (1982).
(b) E. Clementi and G. Corongiu, IBM Tech. Rep. POK-11 (1982).
[110] (a) T. H. Dunning Jr., “Gaussian basis functions for use in molecular calculations.
I. Contraction of (9s5p) atomic basis sets for the first‐row atoms” J. Chem. Phys.,
53, 2823 (1970).
(b) W. J. Hehre; R. F. Stewart; J. A. Pople, “Self‐Consistent Molecular‐Orbital
methods. I. Use of Gaussian expansions of Slater‐Type atomic orbitals” J. Chem.
Phys., 51, 2657 (1969).
[111] R. C. Raenetti, “General contraction of Gaussian atomic orbitals: Core, valence,
polarization and diffuse basis sets; Molecular integral evaluation” J. Chem. Phys.,
58, 4452 (1973).
[112] W. J. Hehre; R. F. Stewart; J. A. Pople, “Self-Consistent Molecular-Orbital methods.
I. Use of Gaussian Expansions of Slater-Type atomic orbitals” J. Chem. Phys., 51,
2657 (1969).
[113] W. J. Hehre; R. Ditchfield; R. F. Stewart; J. A. Pople, “Self-Consistent molecular
orbital methods. IV. Use of Gaussian Expansions of Slater- Type orbitals. Extension
to Second-Row molecules” J. Chem. Phys., 52, 2769 (1970).
[114] W. J. Pietro; B. A. Levi; W. J. Hehre; R. F. Stewart, “Molecular orbital theory of
the properties of inorganic and organometallic compounds. 1. STO-NG basis sets for
third-row main-group elements” Inorg. Chem., 19, 2225 (1980).
[115] W. J. Pietro; E. S. Blurock; R. F. Hout; W. J. Hehre; D. J. DeFrees; R.F. Stewart,
“Molecular orbital theory of the properties of inorganic and organometallic
compounds. 2. STO-NG basis sets for fourth-row main-group elements” Inorg.
Chem., 20, 3650 (1981).
[116] W. J. Pietro; W. J. Hehre , “Molecular orbital theory of the properties of inorganic
and organometallic compounds. 3. STO-3G basis sets for first- and second-row
transition metals” J. Comp. Chem., 4, 241 (1983).
[117] P. Pulay; G. Fogarasi; F. Pang; J. E. Boggs, “Systematic ab initio gradient
calculation of molecular geometries, force constants, and dipole moment
derivatives” J. Am. Chem. Soc., 101, 2550 (1979).
[118] J. S. Binkley; J. A. Pople; W. J. Hehre, “Self-consistent molecular orbital methods.
21. Small split-valence basis sets for first-row elements,” J. Am. Chem. Soc., 102,
939 (1980).
[119] M. S. Gordon; J. S. Binkley; J. A. Pople; W. J. Pietro; W. J. Hehre, “Self-consistent
molecular-orbital methods. 22. Small split-valence basis sets for second-row
elements” J. Am. Chem. Soc., 104, 2797 (1982).
[120] W. J. Pietro; M. M. Francl; W. J. Hehre; D. J. DeFrees; J. A. Pople; J. S. Binkley,

63
“Self-consistent molecular orbital methods.24. Supplemented small split-valence
basis sets for second-row elements” J. Am. Chem. Soc., 104, 5039 (1982).
[121] T. Clark; J. Chandrasekhar; G. W. Spitznagel; P. v. R. Schleyer, “Efficient diffuse
function-augmented basis sets for anion calculations. III. The 3-21+G basis set for
first-row elements, Li-F” J. Comp. Chem., 4, 294 (1983).
[122] R. Ditchfield; W. J. Hehre; J. A. Pople, “Self-Consistent Molecular-Orbital
methods. IX. An extended Gaussian-Type basis for Molecular-Orbital studies of
organic molecules” J. Chem. Phys., 54, 724 (1971).
[123] W. J. Hehre; W. A. Lathan, “Self-Consistent Molecular Orbital Methods. XIV. An
extended Gaussian-Type basis for molecular orbital studies of organic molecules.
inclusion of second row elements” J. Chem. Phys., 56, 5255 (1972).
[124] J. Chandrasekhar; J. G. Andrade; P.v.R. Schleyer, “Efficient and accurate
calculation of anion proton affinities” J. Am. Chem. Soc., 103, 5609 (1981).
[125] W. J. Hehre; R. Ditchfield; J. A. Pople, “Self-Consistent Molecular Orbital Methods.
XII. Further extensions of Gaussian-Type Basis Sets for use in molecular orbital
studies of organic molecules” J. Chem. Phys., 56, 2257 (1972).
[126] M. M. Francl; W. J. Pietro; W. J. Hehre, “Self-consistent molecular orbital methods.
XXIII. A polarization-type basis set for second-row elements” J. Chem. Phys., 77,
3654 (1982).
[127] R. H. Nobes; W. R. Rodwell; Leo Radom, “The 6-31G++ basis set: An economical
basis set for correlated wavefunctions” J. Comp. Chem., 3, 561 (1982).
[128] G. W. Spitznagel; T. Clark; J. Chandrasekhar; P. v. R. Schleyer, “Stabilization of
methyl anions by first-row substituents. The superiority of diffuse function-
augmented basis sets for anion calculations” J. Comput. Chem., 3, 363 (1982).
[129] (a) T. H. Dunning, “Gaussian basis functions for use in molecular calculations. I.
Contraction of (9s5p) atomic basis sets for the first-row atoms” J. Chem. Phys., 53,
2823 (1970).
(b) M. J. Frisch; J. A. Pople, “Self-consistent molecular orbital methods 25.
Supplementary functions for Gaussian basis sets” J. Chem. Phys., 80, 3265 (1984).
[130] T. H. Dunning, “Gaussian Basis Functions for use in molecular calculations. I.
Contraction of (9s5p) atomic basis sets for the First-row atoms” J. Chem. Phys., 53,
2823 (1970).
[131] A. D. McLean; G. S. Chandler, “Contracted Gaussian basis sets for molecular
calculations. I. Second row atoms, Z=11–18” J. Chem. Phys., 72, 5639 (1980).
[132] E. R. Davidson; D. Feller, “Basis set selection for molecular calculations” Chem.
Rev., 86, 861 (1986).
[133] J. Almlof; P. R. Taylor,“General contraction of Gaussian basis sets. I. Atomic
natural orbitals for First‐ and Second‐row atoms” J. Chem.Phys., 86, 4070 (1987).
[134] (a) T. H. Dunning Jr.,“Gaussian basis sets for use in correlated molecular
calculations. I. The atoms boron through neon and hydrogen” J. Chem. Phys., 90,
1007 (1989).
(b) A. K. Wilson; T. Mourik; T. H. Dunning Jr., "Gaussian basis sets for use in
correlated molecular calculations. VI. Sextuple zeta correlation consistent basis sets

64
for boron through neon” J. Mol. Struct., 339-349 (1996).
[135] (a) R. A. Kendall; T. H. Dunning Jr.; R. J. Harrison,“Electron affinities of the
first‐row atoms revisited. Systematic basis sets and wave functions” J. Chem. Phys.,
96, 6796 (1992).
(b) R. A. Woon; T. H. Dunning Jr.,“Gaussian basis sets for use in correlated
molecular calculations. III. The atoms aluminum through argon” J. Chem. Phys., 98,
1358 (1993).
[136] F. Jensen, “Introduction to Computational Chemistry” John Wiley & Sons, p162
(1999).
[137] W. Kultzelnigg; W. Kopper, “Wave functions with terms linear in the inter
electronic coordinates to take care of the correlation cusp. I. General theory”
J. Chem. Phys., 94, 1985 (1991).
[138] (a) B. J. Persson; P. R. Taylor, “Accurate quantum‐chemical calculations: The use of
Gaussian‐type geminal functions in the treatment of electron correlation” J. Chem.
Phys., 105, 5915 (1996).
(b) B.J. Persson; P.R. Taylor, “Molecular integrals over Gaussian-type geminal
basis functions” Theoret. Chem. Acc., 97, 240 (1997).

65
Reference

[1] Mathematical Challenges from Theoretical/Computational Chemistry,


http:// books.nap.edu/catalog/4886.html.

[2] M. Born; J. R. Oppenheimer, Ann. Phys., 46, 618 (1934).

[3] D. R. Hartree, Proc. Cambridge Philos. Soc., 24, 328 (1928).

[4] V. A. Fock, Z.Phys., 15, 126 (1930).

[5] C. C. J. Roothan, “New Developments in Molecular Orbital Theory”


Rev. Mod. Phys., 23, 69-89 (1951).

[6] (a) K. Ruedenberg, “On the Three- and Four-Center Integrals in


Molecular Quantum Mechanics” J. Chem. Phys., 19, 1433 (1951).
(b) K. Ruedenberg, “A Study of Two-Center Integrals Useful in
Calculations on Molecular Structure. II. The Two-Center Exchange
Integrals” J. Chem. Phys., 19, 1459 (1951).

[7] J. A. Pople; R. K. Nesbet, “Self-Consistent Orbitals for Radicals” J.


Chem. Phys., 22, 571 (1954).

[8] David P. Tew; Wim Klopper; Trygve Helgaker, “Electron Correlation:


The Many Body Problem at the Heart of Chemistry” J. Comput. Chem.,
28, 1307-1320 (2007).

[9] E. A. Hylleraas, “Über den Grundzustand des Heliumatoms” Z. Physik,


48, 469 (1928).

[10] A. Meckler, “Electronic Energy Levels of Molecular Oxygen” J.


Chem. Phys. 21, 1750 (1953).

[11] J. B. Foresman; M. Head-Gordon; J. A. Pople; M. J. Frisch, “Toward a


Systematic Molecular Orbital Theory for Excited States” J. Phys. Chem.,
96, 135 (1992).

66
[12] J. A. Pople; R. Seeger; R. Krishnan, “Variational Configuration
Interaction Methods and Comparison with Perturbation Theory” Int. J.
Quantum Chem., Suppl. Y-11, 149 (1977).

[13] K. Raghavachari; H. B. Schlegel; J. A. Pople, “Derivative studies in


configuration- interaction theory” J. Chem. Phys., 72, 4654 (1980).

[14] K. Raghavachari; J. A. Pople, “Calculation of one-electron properties


using limited configuration-interaction techniques” Int. J. Quantum
Chem., 20, 1067 (1981).

[15] S. R. Langhoff; E. R. Davidson, “Configuration interaction calculations


on the nitrogen molecule” Int. J. Quantum Chem. 8, 61 (1974).

[16] J. A. Pople; M. Head-Gordon; K. Raghavachari, “Quadratic


configuration interaction. A general technique for determining electron
correlation energies” J. Chem. Phys., 87, 5968 (1987).

[17] C. Møller; M. S. Plesset, “Note on an Approximation Treatment for


Many-Electron Systems” Phys. Rev., 46, 618 (1934).

[18] M. Head-Gordon; J. A. Pople; M. J. Frisch, “MP2 energy evaluation by


direct methods” Chem. Phys. Lett., 153, 503 (1988).

[19] S. Saebø; J. Almlöf, “Avoiding the integral storage bottleneck in LCAO


calculations of electron correlation” Chem. Phys. Lett., 154, 83 (1989).

[20] M. J. Frisch; M. Head-Gordon; J. A. Pople,“Direct MP2 gradient


method” Chem. Phys. Lett., 166, 275 (1990).

[21] M. J. Frisch; M. Head-Gordon; J. A. Pople, “Semi-direct algorithms for


the MP2 energy and gradient” Chem. Phys. Lett., 166, 281 (1990).

[22] M. Head-Gordon; T. Head-Gordon, “Analytic MP2 Frequencies Without


Fifth Order Storage: Theory and Application to Bifurcated Hydrogen
Bonds in the Water Hexamer” Chem. Phys. Lett., 220, 122 (1994).

[23] J. A. Pople; J. S. Binkley; R. Seeger, “Theoretical Models Incorporating


Electron Correlation” Int. J. Quantum Chem., Suppl. Y-10, 1 (1976).

67
[24] J. A. Pople; R. Seeger; R. Krishnan, “Variational Configuration
Interaction Methods and Comparison with Perturbation Theory” Int. J.
Quantum Chem., Suppl. Y-11, 149 (1977).

[25] K. Raghavachari; J. A. Pople, “Approximate 4th-order perturbation-


theory of electron correlation energy” Int. J. Quantum Chem., 14, 91
(1978).

[26] K. Raghavachari; J. A. Pople; E. S. Replogle; M. Head-Gordon, “Fifth


Order Møller-Plesset Perturbation Theory: Comparison of Existing
Correlation Methods and Implementation of New Methods Correct to
Fifth Order” J. Phys. Chem., 94, 5579 (1990).

[27] G. W. Trucks; J. D. Watts; E. A. Salter; R. J. Bartlett, “Analytical


MBPT(4) Gradients” Chem. Phys. Lett., 153, 490 (1988).

[28] G. W. Trucks; E. A. Salter; C. Sosa; R. J. Bartlett, “Theory and


Implementation of the MBPT Density Matrix: An Application to One-
Electron Properties” Chem. Phys. Lett., 147, 359 (1988).

[29] J. Čížek, J. Chem. Phys., 45, 4526 (1966).

[30] J. Čížek, “On the Use of the Cluster Expansion and the Technique of
Diagrams in Calculations of Correlation Effects in Atoms and
Molecules” Adv. Chem. Phys., 14, 35 (1969).

[31] J. Čížek; J. Paldus, “Correlation problems in atomic and molecular


systems III. Rederivation of the coupled-pair many-electron theory
using the traditional quantum chemical methods” Int. J. Quant.
Chem., 5, 359 (1971).

[32] J. A. Pople; R. Krishnan; H. B. Schlegel; J. S. Binkley, “Electron


Correlation Theories and Their Application to the Study of Simple
Reaction Potential Surfaces” Int. J. Quantum Chem., 14, 545 (1978).

[33] D. Watts; G. W. Trucks; R. J. Bartlett, “Coupled-cluster, unitary


coupled-cluster and MBPT(4) open-shell analytical gradient methods”
Chem. Phys. Lett., 164, 502 (1989).

[34] G. D. Purvis; R. J. Bartlett, “A full coupled-cluster singles and doubles


model–the inclusion of disconnected triples” J. Chem. Phys., 76, 1910
(1982).

68
[35] G. E. Scuseria; C. L. Janssen; H. F. Schaefer III, “An efficient
reformulation of the closed-shell coupled cluster single and double
excitation (CCSD) equations” J. Chem. Phys., 89, 7382 (1988).

[36] G. E. Scuseria; H. F. Schaefer III, “Is coupled cluster singles and


doubles (CCSD) more computationally intensive than quadratic
configuration-interaction(QCISD)?” J. Chem. Phys., 90, 3700 (1989).

[37] J. Noga; R. J. Bartlett, “The full CCSDT model for molecular electronic
structure” J. Chem. Phys., 86, 7041 (1987).

[38] K. Raghavachari; G. W. Trucks; J. A. Pople; M. Head-Gordon, “A fifth-


order perturbation comparison of electron correlation theories” Chem.
Phys. Lett., 157, 479 (1989).

[39] R. J. Bartlett, “Coupled-cluster approach to molecular structure and


spectra: a step toward predictive quantum chemistry” J. Phys.
Chem., 93, 1697 (1989).

[40] T. Helgaker; T.A. Ruden; P. Jørgensen; J. Olsen; W. Klopper,


“A priori calculation of molecular properties to chemical accuracy”
J. Phys. Organic Chem., 17, 913 (2004).

[41] M. Head-Gordon; R. J. Rico; M. Oumi; T. J. Lee, “A Doubles


Correction to Electronic Excited-States from Configuration-Interaction
in the Space of Single Substitutions” Chem. Phys. Lett., 219,21 (1994).

[42] M. Head-Gordon; D. Maurice; M. Oumi, “A Perturbative Correction to


Restricted Open-Shell Configuration-Interaction with Single
Substitutions for Excited-States of Radicals” Chem. Phys. Lett., 246,
114 (1995).

[43] Y. Mochizuki; K. Tanaka, “Modification for spin-adapted version of


configuration interaction singles with perturbative doubles” Chem.
Phys. Lett., 443, 389 (2007).

[44] D. Hegarty; M. A. Robb, “Application of unitary group-methods to


configuration- interaction calculations” Mol. Phys., 38, 1795 (1979).

[45] R. H. A. Eade; M. A. Robb, “Direct minimization in MC SCF theory -


the Quasi-Newton method” Chem. Phys. Lett., 83, 362 (1981).

69
[46] H. B. Schlegel; M. A. Robb, “MC SCF gradient optimization of the
H2CO→ H2+CO transition structure” Chem. Phys. Lett., 93, 43 (1982).

[47] M. J. Frisch; I. N. Ragazos; M. A. Robb; H. B. Schlegel, “An Evaluation


of 3 Direct MC- SCF Procedures” Chem. Phys. Lett., 189, 524 (1992).

[48] N. Yamamoto; T. Vreven; M. A. Robb; M. J. Frisch; H. B. Schlegel, “A


Direct Derivative MC-SCF Procedure” Chem. Phys. Lett., 250, 373
(1996).

[49] Frank Jensen, “Introduction to Computational Chemistry” John Wiley &


Sons, pp 101-117 (1999).

[50] B. O. Ross; P. R. Taylor; P. E. M. Siegbahn, “ A complete active space


SCF method (CASSCF) using a density matrix formulated super-CI
approach” Chem. Phys., 48, 157 (1980).

[51] J. Olsen; B. O. Roos; P. Jørgensen; H. J. A. Jensen, “Determinant based


configuration interaction algorithms for complete and restricted
configuration interaction spaces” J. Chem. Phys., 89, 2185 (1998).

[52] J. A. Pople; M. Head-Gordon; D. J. Fox; K. Raghavachari; L. A. Curtiss,


“Gaussian‐1 theory: A general procedure for prediction of molecular
energies” J. Chem. Phys. 90, 5622 (1989).

[53] L. A. Curtiss; K. Raghavachari; G. W. Trucks; J. A. Pople, “Gaussian-2


theory for molecular energies of first- and second-row compounds”
J. Chem. Phys., 94, 7221 (1991).

[54] L. A. Curtiss; K. Raghavachari; P. C. Redfern; V. Rassolov; J. A. Pople,


“Gaussian-3 (G3) theory for molecules containing first and second-row
atoms” J. Chem. Phys., 109, 7764 (1998).

[55] L. A. Curtiss; P. C. Redfern; K. Raghavachari, “Gaussian-4 theory” J.


Chem. Phys.. 126, 084108 (2007).

[56] (a) N. J. De Yonker; T. R. Cundari; A. K. Wilson, “The correlation


consistent composite approach (ccCA): An alternative to the
Gaussian-n methods” J. Chem. Phys., 124, 114104 (2006).
(b) N. J. De Yonker; T. Grimes; S. Yockel; A. Dinescu; B. Mintz; T. R.
Cundari; A. K. Wilson, “The correlation-consistent composite

70
approach: Application to the G3/99 test set" J. Chem. Phys., 125,
104111 (2006).

[57] (a) J. W. Ochterski; G. A. Petersson; J. A. Montgomery, “A complete


basis set model chemistry. V. Extensions to six or more heavy
atoms” J. Chem. Phys., 104, 2598 (1996).
(b) J. A. Montgomery Jr.; M. J. Frisch; J. W. Ochterski; G. A. Petersson,
“A complete basis set model chemistry. VI. Use of density functional
geometries and frequencies” J. Chem. Phys., 110, 2822 (1999).
(c) J. A. Montgomery Jr.; M. J. Frisch; J. W. Ochterski; G. A. Petersson,
“A complete basis set model chemistry. VII. Use of the minimum
population localization method” J. Chem. Phys., 112, 6532 (2000).

[58] P. L. Fast; M. L. Sánchez; D. G. Truhlar, “Multi-coefficient Gaussian-3


method for calculating potential energy surfaces” Chem. Phys. Lett.,
306, 407, (1999).

[59] B. J. Lynch; D. G. Truhlar, “Robust and Affordable Multicoefficient


Methods for Thermochemistry and Thermochemical Kinetics: The
MCCM/3 Suite and SAC/3” J. Phys. Chem. A, 107, 3898 (2003).

[60] L. A. Curtiss; K. Raghavachari; P. C. Redfern; J. A. Pople, “Assessment


of Gaussian-2 and density functional theories for the computation of
enthalpies of formation” J. Chem. Phys., 106, 1063 (1997).

[61] L. A. Curtiss; P. C. Redfern; K. Raghavachari; J. A. Pople, “Assessment


of Gaussian-3 and density functional theories for a larger experimental
test set” J. Chem. Phys., 112, 7374 (2000).

[62] L. A. Curtiss; P. C. Redfern; K. Raghavachari, “Assessment of Gaussian-


3 and density-functional theories on the G3/05 test set of experimental
energies” J. Chem. Phys., 123, 124107 (2005).

[63] C. J. Cramer, “Essentials of Computational Chemistry: Theories and


Models” 2nd Edn., John Wiley & Sons, p493 (2004).

[64] A. Venkatnathan; A. B. Szilva; D. Walter; R. J. Gdanitz; E. A. Carter,


“Size Extensive Modification of local multi-reference configuration
interaction” J. Chem. Phys., 120,1693 (2004).

[65] S. Chattopadhyay; U. S. Mahapatra; B. Datta; D. Mukherjee, “State-


specific multi-reference coupled electron-pair approximation like

71
methods: formulation and molecular applications” Chem. Phys. Lett.,
357, 426 (2002).

[66] J. A. R. Paul, “On the Size Consistency of Multi-reference CEPA


Methods” Collection of Czechoslovak Chemical Communications, 70,
638 (2005).

[67] T. Kato, “On the eigenfunctions of many-particle systems in quantum


mechanics” Comm. Pure Appl. Math., 10, 151 (1957).

[68] G. W. F. Drake; Z. Van, “Variational eigenvalues for the S states of


helium” Chem. Phys. Lett., 229, 486 (1994).

[69] W. Kolos; L. Wolniewicz, “Improved Theoretical Ground-State Energy


of the Hydrogen Molecule” J. Chem. Phys., 49, 404 (1968).

[70] W. Kutzelnigg; W. Klopper, “Wave functions with terms linear in the


interelectronic coordinates to take care of the correlation cusp. I.
General theory” J. Chem. Phys., 94, 1985 (1991).

[71] T. Shiozaki; M. Kaniya; S. Hirata; E. F. Valeev, “Explicitly correlated


coupled-cluster singles and doubles method based on complete
diagrammatic equations” J. Chem. Phys., 129, 071101 (2008).

[72] T. Shiozaki; M. Kaniya; S. Hirata; E. F. Valeev, “Higher-order explicitly


correlated coupled-cluster methods” J. Chem. Phys. 130, 054101
(2009).

[73] M. J. Field; P. A. Bash; M. Karplus, “A combined quantum mechanical


and molecular mechanical potential for molecular dynamics
simulations” J. Comput. Chem., 11, 700 (1990).

[74] M. Svensson; S. Humbel; R. D. J. Froese; T. Matsubara; S. Sieber; K.


Morokuma, “ONIOM: A Multilayered Integrated MO + MM Method
for Geometry Optimizations and Single Point Energy Predictions” J.
Phys. Chem., 100, 19357 (1996).

[75] C. M. Quinn, “Computational Quantum Chemistry: an interactive guide


to basis set theory” Academic Press, p 17 (2002).

[76] J. C. Slater, “Atomic Shielding Constants” Phys. Rev., 36, 57 (1930).

72
[77] J. C. Slater, “Analytic Atomic Wave Functions” Phys. Rev., 42, 33
(1932).

[78] C. Zener, “Analytic Atomic Wave Functions” Phys. Rev., 36, 51 (1930).

[79] E. Clementi, “Tables of atomic functions” IBM J. Res. Develop. Suppl.


(1965).

[80] E. Clementi and C. Roetti, At. Data Nucl. Data Tables 14, 177 (1974).

[81] A.D. McLean and R.S. Mclean, At. Data Nucl. Data Tables 26, 197
(1981).

[82] (a) T. Koga; H. Tatewaki; A. J. Thakkar, “Roothaan-Hartree-Fock wave


functions for atoms with Z≤54,” Phys. Rev. A., 47, 4510 (1993).
(b) T. Koga; Y. Seki; A.J. Thakkar; H. Tatewaki, “Roothaan-Hartree-
Fock wavefunctions for ions with N< or =54” J. Phys. B: At. Mol.
Opt. Phys., 26, 2529 (1993).
(c) T. Koga; S. Watanabe; K. Kanayama; R. Yasuda; A.J. Thakkar,
“Improved Roothaan-Hartree-Fock wave functions for atoms and
ions with N<=54” J. Chem. Phys., 103, 3000 (1995).
(d) C.F. Bunge; J.A. Barrientos; A.V. Bunge; J.A. Cogordan, “Hartree-
Fock and Roothaan-Hartree-Fock energies for the ground states of
He through Xe” Phys. Rev. A., 46, 3691 (1992).

[83] E. Francisco; L. Seijo; L. Pueyo, “The maximum overlap method: A


general and efficient scheme for reducing basis sets. Application to the
generation of approximate AO's for the 3d transition metal atoms and
ions” J. Solid State Chem., 63, 391 (1986).

[84] (a) J. M. G. Vega; J. F. Rico; J.I. F. andez-Alonso, “Quantum chemical


applications of the theory of the distance between subspaces” J.
Mol. Struct (Theochem)., 184, 1 (1989).
(b) J. M. G. Vega; B. Miguel, “Approximate STO functions for the first-
row transition metal atoms” J. Comput. Chem., 12, 1172 (1991).
(c) J. M. G. Vega; B. Miguel, “Slater Functions for Y to Cd Atoms by
the Distance between Subspaces” J. Solid State Chem., 116, 275
(1995).

[85] (a) J.A. Sordo; L. Pueyo, “A quality test for the Hartree-Fock-Roothaan
SCF wave functions” Int. J. Quantum Chem., 28, 687 (1985).

73
(b) J .A. Sordo; L. Pueyo, “High-quality Hartree-Fock-Roothaan wave
functions for molecular calculations” J. Mol. Struct. (Theochem)
120, 9 (1985).

[86] J. R. Mohallem; M. Trsic, “A universal Gaussian basis set for atoms Li


through Ne based on a generator coordinate version of the Hartree-
Fock equations” J. Chem. Phys., 86, 5043 (1987).

[87] (a) J. M. G. Vega; B. Miguel, “Single-exponent Slater function


expansions of the He atom 1s orbital and its isoelectronic series”
Chem. Phys. Lett., 207, 270 (1993).
(b) J. M. G. Vega; B. Miguel, “Orbitals expanded in slater functions
with single-exponent by shell and by subshell” Int. J. Quantum
Chem., 51, 397 (1994).

[88] T. Koga; K. Kanayana; A.J. Thakkar, “Theoretical and Computational


Developments: Noninteger principal quantum numbers increase the
efficiency of Slater-type basis sets” Int. J. Quantum Chem., 62, 1 (1997).

[89] T. Koga; Y. Seki; A.J. Thakkar; H. Tatewaki, “Noninteger principal


quantum numbers increase the efficiency of Slater-type basis sets: singly
charged cations and anions” J. Phys. B: At. Mol. Opt. Phys., 30, 1623
(1997).

[90] T. Koga; K. Kanayana, “Noninteger principal quantum numbers


increase the efficiency of Slater-type basis sets: heavy atoms” Chem.
Phys. Lett., 266, 123 (1997).

[91] T. Koga; J.M. G. Vega; B. Miguel, “Double-zeta Slater-type basis sets


with noninteger principal quantum numbers and common exponents”
Chem. Phys. Lett., 283, 97 (1998).

[92] E. Filter; E.O. Steinborn, “Extremely compact formulas for molecular


two-center one-electron integrals and Coulomb integrals over Slater-
type atomic orbitals” Phys. Rev. A., 18, 1 (1978).

[93] S. A. Vagranov; A. T. B. Gilbert; E. Duplaxes; P. M. W. Gill,


“Resolutions of the Coulomb operator” J. Chem. Phys., 128, 201104
(2008).

[94] P. M.W. Gill: A. T. B. Gilbert, “Resolutions of the Coulomb operator: II.


The Laguerre Generator” Chem. Phys., 86, 356 (2009).

74
[95] P. E. Hoggan, “Four-center Slater-type orbital molecular integrals
without orbital translations” Int. J. Quant. Chem., 110, 98 (2010).

[96] S. F. Boys, “Electronic Wave Functions. I. A General Method of


Calculation for the Stationary States of Any Molecular System” Proc. R.
SOC. London, A200, 542 (1950).

[97] R. McWeeny, “Gaussian Approximations, to Wave Functions” Nature.,


166, 21 (1950).

[98] C. J. Cramer, “Essentials of Computational Chemistry: Theories and


Models” 2nd Edn., John Wiley & Sons, p167 (2004).

[99] (a) E. Clementi, IBMJ. Res. and Dev., “ab Initio Computations in Atoms
and Molecules” 9, 2 (1965).
(b) J. L. Whitten, “Gaussian Lobe Function Expansions of Hartree-
Fock Solutions for the First‐Row Atoms and Ethylene” J. Chem.
Phys. 44, 359 (1966).

[100] E. Clementi; D. R. Davis, “Electronic structure of large molecular


systems” J. Comput. Phys., 1, 223 (1967).

[101] S. Huzinaga; “Gaussian-Type Functions For Polyatomic Systems” J.


Chem. Phys., 42, 1293 (1965).

[102] F.B. van Duijneveldt, IBM Technical Research Report, RS 945 (1971).

[103] L. A. Montero, L. A. Diaz and R. Bader (eds.), “Introduction to


Advanced Topics of Computational Chemistry” Editorial de la
Universidad de La Habana, p 52 (2003)

[104] (a) C.M. Reeves; M.C. Harrison, “Use of Gaussian Functions in the
Calculation of Wave functions for Small Molecules. II. The
Ammonia Molecule” J. Chem. Phys., 39, 11 (1963).
(b) M.W. Schmidt; K. Ruedenberg, “Effective convergence to
complete orbital bases and to the atomic Hartree–Fock limit
through systematic sequences of Gaussian primitives” J. Chem.
Phys., 71, 3951 (1979).

[105] (a) D. Feller; K. Ruedenberg, “Cluster expansion of the wave


function” Theor. Chim. Acta., 71, 231 (1979).

75
(b) M. W. Schmidt; K. Ruedenberg,. “Effective convergence to
complete orbital bases and to the atomic Hartree-Fock limit
through systematic sequences of Gaussian primitives” J. Chem.
Phys., 3951, 71 (1979).

[106](a) S. Huzinaga; M. Klobukowski; H. Tatewaki, “The well-tempered


GTF basis sets and their applications in the SCF calculations on
N2, CO, Na2, and P2” Can. J. Chem., 63, 1812 (1985).
(b) S. Huzinaga; M. Klobukowski; H. Tatewaki, “Well-tempered GTF
basis sets for the atoms K through Xe” Chem. Phys. Lett., 120,
509 (1985).
(c) S. Huzinaga; B. Miguel, “A comparison of the geometrical
sequence formula and the well-tempered formulas for generating
GTO basis orbital exponents” Chem. Phys. Lett., 175,289 (1990).
(d) S. Huzinaga; B. Miguel, “A comparison of the geometrical
sequence formula and the well-tempered formulas for generating
GTO basis orbital exponents” Chem. Phys. Lett., 75, 289 (1990).
(e) M. Klobukowski, “Systematic sequences of well-balanced Gaussian
basis sets” Can. J. Chem., 72, 1741 (1994).

[107] D.M. Silver; W.C. Nieuwpoort, “Two-centre model potential energy


calculations for the 2Σ and 2Π states of Li+2, Na+2, K+2, Rb+2 and
Cs+2” Chem. Phys. Lett., 57, 421 (1978).

[108] (a) D.M. Silver; S. Wilson; W.C. Nieuwpoort, “Universal basis sets
and transferability of integrals” Int.t. J. Quantum Chem., 14, 635
(1978).
(b) S. Wilson, “Systematic sequences of even-tempered Gaussian
primitives in electron correlation studies using many-body
perturbation theory” Theor. Chim. Acta 57, 53 (1980).
(c) S. Wilson, “Universal systematic sequence of even-tempered
Gaussian primitive functions in electronic correlation studies”
Theor. Chim. Acta 58, 31 (1980).

[109] (a) E. Clementi; G. Corongiu, “Geometrical basis set for molecular


computations” Chem. Phys. Lett., 90, 359 (1982).
(b) E. Clementi and G. Corongiu, IBM Tech. Rep. POK-11 (1982).

[110] (a) T.H. Dunning Jr., “Gaussian Basis Functions for Use in Molecular
Calculations. I. Contraction of (9s5p)Atomic Basis Sets for the
First‐Row Atoms” J. Chem. Phys., 53, 2823 (1970).

76
(b) W.J. Hehre; R.F. Stewart; J.A. Pople, “Self‐Consistent
Molecular‐Orbital Methods. I. Use of Gaussian Expansions of
Slater‐Type Atomic Orbitals” J. Chem. Phys., 51, 2657 (1969).

[111] R.C. Ra_enetti, “General contraction of Gaussian atomic orbitals:


Core, valence, polarization and diffuse basis sets; Molecular integral
evaluation” J. Chem. Phys., 58, 4452 (1973).

[112] W. J. Hehre; R. F. Stewart; J. A. Pople, “Self-Consistent Molecular-


Orbital Methods. I. Use of Gaussian Expansions of Slater-Type Atomic
Orbitals” J. Chem. Phys., 51, 2657 (1969).

[113] W. J. Hehre; R. Ditchfield; R. F. Stewart; J. A. Pople, “Self-Consistent


Molecular Orbital Methods. IV. Use of Gaussian Expansions of Slater-
Type Orbitals. Extension to Second-Row Molecules” J. Chem.
Phys., 52, 2769 (1970).

[114] W.J. Pietro; B. A. Levi; W. J. Hehre; R. F. Stewart, “Molecular orbital


theory of the properties of inorganic and organometallic compounds. 1.
STO-NG basis sets for third-row main-group elements” Inorg.
Chem., 19, 2225 (1980).

[115] W. J. Pietro; E. S. Blurock; R. F. Hout; W. J. Hehre; D. J. DeFrees;


R.F. Stewart, “Molecular orbital theory of the properties of inorganic
and organometallic compounds. 2. STO-NG basis sets for fourth-row
main-group elements” Inorg. Chem., 20, 3650 (1981).

[116] W.J. Pietro; W. J. Hehre , “Molecular orbital theory of the properties


of inorganic and organometallic compounds. 3. STO-3G basis sets for
first- and second-row transition metals” J. Comp. Chem., 4, 241
(1983).

[117] P. Pulay; G. Fogarasi; F. Pang; J. E. Boggs, “Systematic ab initio


gradient calculation of molecular geometries, force constants, and
dipole moment derivatives” J. Am. Chem. Soc., 101, 2550 (1979).

[118] J. S. Binkley; J. A. Pople; W.J. Hehre, “Self-consistent molecular


orbital methods. 21. Small split-valence basis sets for first-row
elements,” J. Am. Chem. Soc., 102, 939 (1980).

[119] M. S. Gordon; J. S.Binkley; J. A. Pople; W.J. Pietro; W. J. Hehre,


“Self-consistent molecular-orbital methods. 22. Small split-valence

77
basis sets for second-row elements” J. Am. Chem. Soc., 104, 2797
(1982).

[120] W. J. Pietro; M. M. Francl; W. J. Hehre; D. J. DeFrees; J. A. Pople; J.


S. Binkley, “Self-consistent molecular orbital methods. 24.
Supplemented small split-valence basis sets for second-row elements”
J. Am. Chem. Soc., 104, 5039 (1982).

[121] T. Clark; J. Chandrasekhar; G. W. Spitznagel; P. V. R. Schleyer,


“Efficient diffuse function-augmented basis sets for anion calculations.
III. The 3-21+G basis set for first-row elements, Li-F” J. Comp.
Chem.., 4, 294 (1983).

[122] R. Ditchfield; W. J. Hehre; J. A. Pople, “Self-Consistent Molecular-


Orbital Methods. IX. An Extended Gaussian-Type Basis for Molecular-
Orbital Studies of Organic Molecules” J. Chem. Phys., 54,
724 (1971).

[123] W. J. Hehre; W. A. Lathan, “Self-Consistent Molecular Orbital


Methods. XIV. An Extended Gaussian-Type Basis for Molecular
Orbital Studies of Organic Molecules. Inclusion of Second Row
Elements” J. Chem. Phys., 56, 5255 (1972).

[124] J. Chandrasekhar; J. G. Andrade; P. R. Schleyer, “Efficient and


accurate calculation of anion proton affinities” J. Am. Chem.
Soc., 103, 5609 (1981).

[125] W. J. Hehre; R. Ditchfield; J. A. Pople, “Self—Consistent Molecular


Orbital Methods. XII. Further Extensions of Gaussian-Type Basis Sets
for Use in Molecular Orbital Studies of Organic Molecules” J. Chem.
Phys., 56, 2257 (1972).

[126] M. M. Francl; W. J. Pietro; W. J. Hehre, “Self-consistent molecular


orbital methods. XXIII. A polarization-type basis set for second-row
elements” J. Chem. Phys., 77, 3654 (1982).

[127] R. H. Nobes; W. R. Rodwell; Leo Radom, “The 6-31G++ basis set: An


economical basis set for correlated wavefunctions” J. Comp. Chem..,
3, 561 (1982).

[128] G. W. Spitznagel; T. Clark; J. Chandrasekhar; P. V. R. Schleyer,


“Stabilization of methyl anions by first-row substituents. The

78
superiority of diffuse function-augmented basis sets for anion
calculations” J. Comput. Chem., 3, 363 (1982).

[129] (a) T. H. Dunning, “Gaussian Basis Functions for Use in Molecular


Calculations. I. Contraction of (9s5p) Atomic Basis Sets for the
First-Row Atoms” J. Chem. Phys., 53, 2823 (1970).
(b) M. J. Frisch; J.A. Pople, “Self-consistent molecular orbital methods
25. Supplementary functions for Gaussian basis sets” J. Chem.
Phys., 80, 3265 (1984).

[130] T. H. Dunning, “Gaussian Basis Functions for Use in Molecular


Calculations. I. Contraction of (9s5p) Atomic Basis Sets for the First-
Row Atoms” J. Chem. Phys., 53, 2823 (1970).

[131] A. D. McLean; G. S. Chandler, “Contracted Gaussian basis sets for


molecular calculations. I. Second row atoms, Z=11–18” J. Chem.
Phys., 72, 5639 (1980).

[132] E. R. Davidson; D. Feller, “Basis Set Selection for Molecular


Calculations” Chem. Rev., 86, 861 (1986).

[133] J. Almlof; P.R. Taylor, “General contraction of Gaussian basis sets. I.


Atomic natural orbitals for first‐ and second‐row atoms” J. Chem.
Phys., 86, 4070 (1987).

[134] (a)T.H. Dunning Jr., “Gaussian basis sets for use in correlated
molecular calculations. I. The atoms boron through neon and
hydrogen” J. Chem. Phys., 90, 1007 (1989).
(b) A. K. Wilson; T. Mourik; T. H. Dunning Jr., "Gaussian basis sets
for use in correlated molecular calculations. VI. Sextuple zeta
correlation consistent basis sets for boron through neon” J. Mol.
Struct., 339-349 (1996).

[135](a) R.A. Kendall; T.H. Dunning Jr.; R.J. Harrison, “Electron affinities
of the first‐row atoms revisited. Systematic basis sets and wave
functions” J. Chem. Phys., 96, 6796 (1992).
(b) R.A. Woon; T.H. Dunning Jr., “Gaussian basis sets for use in
correlated molecular calculations. III. The atoms aluminum through
argon” J. Chem. Phys., 98, 1358 (1993).

[136] F. Jensen, “Introduction to Computational Chemistry” John Wiley &


Sons, p162 (1999).

79
[137] W. Kultzelnigg; W. Kopper, “Wave functions with terms linear in the
interelectronic coordinates to take care of the correlation cusp. I.
General theory” J. Chem. Phys., 94, 1985 (1991).

[138](a)B.J. Persson; P.R. Taylor, “Accurate quantum‐chemical calculations:


The use of Gaussian‐type geminal functions in the treatment of
electron correlation” J. Chem. Phys., 105, 5915 (1996).
(b)B.J. Persson; P.R. Taylor, “Molecular integrals over Gaussian- type
geminal basis functions” Theoret. Chem. Acc., 97, 240 (1997).

80

You might also like