Electron Correlation
Electron Correlation
Related terms:
Before demonstrating this, we first summarize the coset results by SO(6)/U(3). The
system Hamiltonian is:
Here, ei(i = 1, 2, 3) is the energy at the i orbital. . The number of paired electrons are
and
shows the paired electron numbers in orbitals j and k(≠ i). (Strictly speaking, a factor
of has to be multiplied.)
i = tan−1(−pi/qi) shows the phase angle. As time elapses, if i only takes certain specific
Table16.2 shows the pairing patterns and the phase angles derived from the Hückel
eigencoefficients (See Table 16.1) of the linear and ring lattices of three sites. Phase
0 and π indicate that the signs of the coefficients of the orbitals are of the same
sign or not, respectively. For instance, for the 1st level of the linear lattice, the actions
on the three sites are (0.25, 0.5, 0.25). Hence, there are two pairing configurations:
(0, 0.25, 0.25) and (0.25, 0.25, 0). The phase angles are 0.
Table 16.2. The pairing patterns and the phase angles derived from the Hückel
eigencoefficients (See Table 16.1) of the linear and ring lattices of three sites
Taking these pairing configurations and angles as the initial conditions and by
Hamilton's equations, we can obtain the evolving actions and angles (ni, i). Para-
meters are ei = 0,Vij = 1000 cm−1.Fig.16.10 shows that for a linear lattice, i are strictly
limited in certain values. This indicates that the electron pairing or correlation of the
Hückel orbitals of a linear lattice is very stable. Table16.2 also shows these angles
( 1, 2, 3). For the 1st level of ring lattice, (ni, i) shows quasi-periodic behavior while
the 2nd level shows the angles are limited to certain ranges as shown in Fig.16.11.
Linear and ring lattices are different in that in the ring lattice each site is two-sided
connected while in the linear lattice, the end points are only one-sided connected.
This difference results in that the electronic pairing in the ring lattice is more
unstable. For the linear lattice, the pairing is with fixed angles.
Fig.16.10. For the 1st level of linear lattice, i are strictly limited in certain values.
Fig.16.11. For the 2nd level of ring lattice, the angles are limited to certain ranges.
Calculation also shows that the increase of ei will expand the range of i, i.e., will
not be prone to the pairing stability while the increase of Vij will force i to certain
specific values, i.e., will enhance pairing or its stability. The unevenness of ei and Vij
will destroy the pairing. This is consistent with the physical intuition. Besides, we
have tried the non-Hückel orbitals. They just do not possess the pairing property.
In summary, from the viewpoint of coset dynamics, Hückel orbitals possess nontriv-
ial properties.
Abstract
Relativistic and electron correlation effects play an important role in the electronic
structure of molecules containing heavy elements (main group elements, transition
metals, lanthanide and actinide complexes). It is therefore mandatory to account for
them in quantum mechanical methods used in theoretical chemistry, when inves-
tigating for instance the properties of heavy atoms and molecules in their excited
electronic states. In this chapter we introduce the present state-of-the-art ab initio
spin-orbit configuration interaction methods for relativistic electronic structure cal-
culations. These include the various types of relativistic effective core potentials in the
scalar relativistic approximation, and several methods to treat electron correlation
effects and spin-orbit coupling. We discuss a selection of recent applications on
the spectroscopy of gas-phase molecules and on embedded molecules in a crystal
environment to outline the degree of maturity of quantum chemistry methods. This
also illustrates the necessity for a strong interplay between theory and experiment.
The essential part of non dynamical correlation energy for polyatomic molecules is
the “ left-right electron correlation ”, which is concerned with the ionic-covalent
balance within a given two-electron bond. Let us therefore discuss this type of
correlation.
(1)
This simple wave function, so called the Heitler-London (HL) wave function, was able
to account for about 66% of the bonding energy of H2, and performed a little better
than the rival MO method that appeared almost at the same time.
(2)
The physical constitution of the Hartree-Fock wave function appears most clearly by
expanding the MO determinant of eq 2 as a linear combination of determinants
constructed from pure AO’s, eq 3:
(3)
Here the first two determinants are the determinantal form of the Heitler-London
function (eq 1), and represent a purely covalent interaction between the atoms. The
remaining determinants represent zwitterionic structures, H−H+ and H+H−, and con-
tribute 50% to the wave function. The same constitution holds for any interatomic
distance. This weight of the ionic structures is clearly too much at equilibrium
distance, and becomes absurd at infinite separation where the ionic component
is expected to drop to zero. Qualitatively, this can be corrected by including a
second configuration where both electrons occupy the antibonding orbital, u, i.e.
the doubly excited configuration. The more elaborate wave function CI is shown in
eq. 4, where C1 and C2 are coefficients of the two MO configurations:
(4)
This is the essence of the configuration interaction (CI) method. When both the
coefficients of the configurations and their orbitals are optimized simultaneously
in flexible basis sets, the method is called multi-configuration SCF (MCSCF). The
doubly excited configuration in eq. 4 also involves too a 50:50 mixture of covalent
and ionic components but with a negative sign between them. Consequently the
combination of the two configurations deletes the excess ionic character of HF,
thereby leading to the wave function in eq 5. The corrected wave function displays a
qualitatively correct behavior, with an optimal covalent/ionic ratio of typically 80:20
at equilibrium distance [16] all the way to 100:0 at infinite separation.
(5)
The early VB point of view was based solely on the purely covalent HL wave function.
In this wave function the electrons are never allowed to approach each other and
therefore their electron repulsion is minimized and their Coulomb correlation is at
maximum. Thus, while the Hartree-Fock model has no electron correlation, giving
equal weight to covalent and ionic structures, the early VB models overestimated the
correlation. The true situation is about half-way in-between. In the same way as the
Hartree-Fock wave function is improved by CI, the purely covalent VB function can
be improved by admixture of ionic structures as in eq 5, in which the coefficients
and μ would be directly optimized in the VB framework. Both improved models thus
lead to wave functions that are strictly equivalent and physically correct, even though
their initial expressions appear entirely different. This statement can be generalized:
Since both ab initio VB and ab initio MO theories exploit a subspace of the same
configuration space, the VB and MO wave functions of a given electronic structure
are mutually interconvertible and become equivalent when both theories are driven
to their higher level of refinement.
(6)
(7)
The Coulson-Fischer proposal gave rise to the “separated electron pair theory” which
was initiated by Hurley, Lennard-Jones and Pople [18]. Its further development
by Goddard [19], resulted in the “Generalized Valence Bond” (GVB) method. In
the latter method, each bond in a polyatomic molecule is considered as a pair of
non-orthogonal and spin-coupled orbitals, as in the HL wave function. The different
GVB pairs can in turn be constrained to be mutually orthogonal, without much
loss in numerical accuracy. Much as the Coulson-Fischer orbitals, each GVB orbital
is centered on one atom with delocalization tails on the neighboring atoms. The
resulting GVB wave function that formally displays purely covalent bonds implicitly
contains ionic structures, necessary for a reasonable description of the bonds. The
most popular version of GVB theory is the so-called “perfect pairing” approximation,
which considers a single spin-coupling scheme in which spin pairing is restricted to
the electrons and orbitals of the bonded atoms in the Lewis structure. For example,
in methane, the perfectly paired GVB wave function couples the electrons of each
sp3 hybrid of carbon to the hydrogen that faces it. In fact, for this case of 8 electrons
in 8 orbitals there are 14 possible spin coupling schemes in all rigor. As such, the
perfectly paired GVB approximation constitutes a tremendous simplification of the
wave function, often with no serious loss of accuracy. The closely related method is
the Spin- Coupled (SC) theory of Gerratt, Raimondi and Cooper [20]. This method
removes any orthogonality restrictions and consider all possible spin-coupling
schemes between the singly occupied orbitals. Note that the shape of the orbitals
(e.g. sp3-Mike in the carbon atom of methane) and their degrees of delocalization
are not a priori imposed, but naturally arise from the optimization of the orbitals for
self-consistency. The lone pairs can be treated either as doubly occupied localized
orbitals, or as pairs of strongly overlapping singly occupied orbitals.
but this feature is not taken into account in the wave functions of Coulson-Fisher
type. To include all non dynamical electron correlation, one should abandon the
Coulson-Fisher idea and go back to VB structures constructed with strictly atomic
orbitals, without any delocalization tails, and generate all possible VB structures,
allowing their coefficients and orbitals to be optimized simultaneously.
To calculate electron correlation energies and densities to very high accuracy (e.g.,
for developing new correlation functionals for use in density functional theory) it
is of benefit to be able to calculate the HF wavefunction accurately for any nuclear
charge, including at low, noninteger values to incorporate the long-range behavior
of electronic motion. Therefore we set about implementing HF theory for the singlet
state of two-electron atoms within the clamped nucleus approximation, Eq. (9). Our
goal was twofold (i) to provide a balanced description of the electron correlation
effects in two-electron atoms, and (ii) to exploit the elegant series solution method
to avoid costly explicit integration if possible.
To complement our three-body work, the HF wavefunction, HF, was taken as the
product
(59)
(60)
(61)
It was found that the series solution method could be used to calculate the one--
electron matrix elements arising from . This lead to a five-term recursion relation,
between the coefficients C(q) in the expansion (60), of the form51:
(62)
This recursion relation represents a set of linear equations for the determination of
the coefficients C(q), and the vanishing of their determinant yields the energy
eigenvalues for the core Hamiltonian. It was not possible to use the series solution
method for the two-electron integrals (J and K) as terms arise that do not satisfy the
Laguerre orthogonality condition, (4). However, an analytical solution was possible
by expressing the Laguerre polynomials in the form (2) and using the standard
integral:
(63)
With just a 20-term wavefunction and a single variational parameter, it was shown
that energies were in excellent agreement with the most accurate literature values
and offered increased accuracy for all the heliogenic cations considered, i.e., Z =
3–18. Furthermore, the wavefunctions were shown to provide accurate nucleus–-
electron cusps and expectation values of the interparticle distances for all systems
considered.51
Additionally, we reported the first HF energy for a two-electron system with Z = 0.911
028 224 and calculated the critical nuclear charge for the Hartree–Fock method,
which was found to be .031 177 528. This critical nuclear charge is greater than that
of the hydride ion and reinforces the crucial role electron correlation effects play in
stabilizing anionic systems.51
Of particular interest is the Coulomb hole. The Coulomb hole was defined by Coul-
son and Nielson52 as the difference in the distribution function of the interelectronic
distance r12 for the correlated wavefunction and the Hartree–Fock wavefunction. The
intracule density is defined as,
(64)
It measures the radial correlation between two like-charged particles where r12 is
the distance between them. The intracule distribution function, D(r) = 4πr2h(r), is
normalized to unity such that . The difference between the intracule distribution
functions generated with the statistically independent and uncorrelated HF approx-
imation and the explicitly correlated FC method defines the Coulomb hole, i.e.,
(65)
where ΔD(r) is calculated as the numerical difference between DFC(r) and DHF(r). The
Coulomb hole for the hydride ion is shown in Fig 7. It shows that the most probable
separation (rmax, the maximum in the intracule) shifts from 2.43 a0 in the mean-field
approach of HF to 3.17 a0 as result of electron correlation, and the corresponding
probability at these values is reduced by over 8%. This demonstrates that HF theory
is qualitatively and quantitatively wrong for anions and affirms the importance of
electron correlation.
Fig. 7. The Coulomb hole (solid line), calculated as the difference between the
intracule distribution function DFC(r) (dashed line) and DHF(r) (dotted line) for the
singlet ground state of H−.
(38)
where Dz is the zth component of the electric dipole operator, | i> is the initial state,
and the matrix elements can be understood from the approximate form of the total
(correlated) wavefunction of the resonance state:
The may contain contributions from the scattering continuum of the same channel.
We concluded that “it is reasonable to expect that in reality these optimum condi-
tions cannot be satisfied exactly and thus physical spectra must contain some sin-
gle-particle excitation character even in regions where the spectrum is characterized
mainly by a single, strongly absorbing peak” [143, p. L262].
For DFT and static electron correlation (MCSCF), the same formal expression of
Eq. (2) is used. For DFT, exchange-correlation functionals are added to each n-mer
Hamiltonian, whereas for MCSCF the multiconfigurational wave function is used
and the corresponding MCSCF equations are solved. In the latter case, one fragment
is chosen to be of MCSCF type and all other fragments are treated with RHF. The
dimer calculations use the MCSCF wave function only if one of the monomers
composing the dimer is of MCSCF type. In all other cases, dimers are handled using
RHF. The MCSCF active space definition is the same in MCSCF monomers and
dimers.
(16)
The correlated energy EFMOn-corr is added to EFMOn-RHF (n = 2 or 3). The n-mer con-
tributions EcorrX to the total correlation energy, where X is I, IJ or IJK, are computed
using integrals and Fock matrices computed in the preceding RHF. That is, they
contain the external electrostatic potential and the projection operators added to
one-electron integrals, otherwise usual ab initio expressions for individual n-mer
correlation energies are used.
Although the formal equation is shared for uncorrelated and correlated total en-
ergies, the accuracy of the two is quite different. The errors in the former arise from
the neglect of exchange and restricting density distribution within n-mers; the errors
in the latter are mostly due to restricting the short-ranged correlation within n-mers.
In the case of CI, in the present state of development one has to follow the multi-
layer treatment described below, and the excited states are computed just for one
fragment in the external field of other fragments, and no expression similar to Eq.
(2) is used.
The following notation for the FMO method is introduced: FMOn-M, where M is
the wave function based on the n-body expansion. For example, FMO2-RHF implies
using the two-body expansion with the RHF wave function and FMO3-CCSD(T)
is the three-body method with the CCSD(T) wave function. The fragment size is
sometimes specified as FMOn-M/m, where m is the number of residues or water
molecules per fragment.
> Read full chapter
(152)
Classic papers
The first calculations incorporating electron correlation in an atom (helium) were
published by Egil Andersen Hylleraas in an article “Neue Berechnung der Energie des
Heliums im Grundzustande, sowie des tiefsten Terms von Ortho-Helium”, Zeitschrift für
Physik, 54 (1929) 347. ★The first calculations with electron correlation for molecules
were performed by Walter Heitler and Fritz Wolfgang London in a paper “Wechsel-
wirkung neutraler Atome und homöopolare Bindung nach der Quantenmechanik”
published in Zeitschrift für Physik, 44 (1927) 455. Formation of the covalent bond (in
H2) could be correctly described only after the electron correlation has been included.
June 30, 1927, when Heitler and London submitted the paper, is the birth date of
quantum chemistry. ★Later, significantly more accurate results were obtained for
the hydrogen molecule by Hubert M. James and Albert S. Coolidge in an article “The
Ground State of the Hydrogen Molecule”, Journal of the Chemical Physics, 1 (1933) 825,
and a contemporary reference point for that molecule are papers by Włodzimierz
Kołos and Lutosław Wolniewicz, among others an article entitled “Potential Energy
Curves for the States of the Hydrogen Molecule” published in Journal of Chemical
Physics, 43 (1965) 2429. ★Christian Møller and Milton S. Plesset in Physical Review,
46 (1934) 618 published a paper “Note on an Approximation Treatment for Many--
Electron Systems”, where they presented a perturbational approach to electron
correlation. ★The first calculations with the Multi-Configurational Self-Consistent
Field (MC SCF) method for atoms was published by Douglas R. Hartree, William
Hartree and Bertha Swirles in a paper “Self-Consistent Field, Including Exchange and
Superposition of Configurations, with some Results for Oxygen”, Philosophical Trans-
actions of the Royal Society (London), A 238 (1939) 229, and the general MC SCF
theory was presented by Roy McWeeny in a work “On the Basis of Orbital Theories”,
Proceedings of the Royal Society (London), A 232 (1955) 114. ★As a classic paper in
electronic correlation we also consider an article by Per-Olov Löwdin “Correlation
Problem in Many-Electron Quantum Mechanics” in Advances in Chemical Physics,
2 (1959) 207. ★The idea of the Coupled Cluster (CC) method was introduced by
Fritz Coester in a paper in Nuclear Physics, 7 (1958) 421 entitled “Bound States of a
Many-Particle System”. Jiří Čí žek introduced the (diagrammatic) CC method into
electron correlation theory in the paper “On the Correlation Problem in Atomic
and Molecular Systems. Calculation of Wavefunction Components in Ursell-type
Expansion Using Quantum-Field Theoretical Methods” published in the Journal
of Chemical Physics, 45 (1966) 4256. ★The book edited by Oktay Sinanoğlu and
Keith A. Brueckner “Three Approaches to Electron Correlation in Atoms”, Yale Univ.
Press, New Haven and London, 1970, contains several reprints of the papers which
cleared the path towards the coupled-cluster method. ★A derivation of the coupled
cluster equations (for interacting nucleons) was presented by Herman Kümmel
and Karl-Heinz Lührmann, Nuclear Physics, A191 (1972) 525 in a paper entitled
“Equations for Linked Clusters and the Energy Variational Principle”.
We have learnt, from the example given at the beginning of this chapter, that
the “genetic defect” of mean field methods is, that they describe electrons and
completely ignore the fact that they are close or far away from each other. For
example, in the two-electron case previously considered when we established the
coordinates of electron 1, electron 2 has a certain distribution of the probability
density. This distribution does not change when the electron 1 moves to a different
position. This means that the electrons “are not afraid” to get close to each other,
although they should, since when electrons are close the energy increases (Fig.
10.1.a, b).
Fig. 10.1. Absence of electronic correlation in the helium atom as seen by the
Hartree–Fock method. Visualization of the cross-section of the square of the wave
function (probability density distribution) describing electron 2 within the plane xy
provided electron 1 is located in a certain point in space: a) at (−1, 0, 0); b) at (1, 0, 0).
Note, that in both cases the conditional probability density distributions of electron 2 are
identical. This means electron 2 does not react to the motion of electron 1, i.e. there
is no correlation whatsoever of the electronic motions (when the total wave function
is the Hartree–Fock one).
The explicitly correlated wave function (we will get to it in a moment) has the interelec-
tronic distance built in its mathematical form. We may compare this to making the
electrons wear spectacles.6 Now they avoid each other. One of my students said that
it would be the best if the electrons moved apart to infinity. Well, they cannot. They
are attracted by the nucleus (energy gain), and being close to it, are necessarily close
to each other too (energy loss). There is a compromise to achieve.
Classic Papers
The first calculations with electron correlation for molecules were performed by
Walter Heitler and Fritz Wolfgang London in a paper called “Wechselwirkung neutraler
Atome und homöopolare Bindung nach der Quantenmechanik,” published in Zeitschrift
für Physik, 44, 455 (1927). The covalent bond (in the hydrogen molecule) could be cor-
rectly described only after the electron correlation was included. June 30, 1927, when
Heitler and London submitted this paper, is the birth date of quantum chemistry.
The first calculations incorporating electron correlation in an atom (helium) were
published by Egil Andersen Hylleraas in an article called “Neue Berechnung der Energie
des Heliums im Grundzustande, sowie des tiefsten Terms von Ortho-Helium,” published
in Zeitschrift für Physik, 54, 347 (1929). Later, significantly more accurate results
were obtained for the hydrogen molecule by Hubert M. James and Albert S. Coolidge
in an article called “The ground state of the hydrogen molecule,” published in the Journal
of the Chemical Physics, 1, 825 (1933), and a contemporary reference point for that
molecule are several papers by Włodzimierz Kołos and Lutosław Wolniewicz, among
which was an article entitled “Potential energy curves for the states of the hydrogen
molecule” published in the Journal of Chemical Physics, 43, 2429 (1965). Christian
Møller and Milton S. Plesset in Physical Review, 46, 618 (1934), published a paper
called “Note on an approximation treatment for many-electron systems,” where they
presented a perturbational approach to electron correlation. The first calculations
with the Multi-configurational self-consistent field (MC SCF) method for atoms was
published by Douglas R. Hartree, his father, William Hartree, and Bertha Swirles in
a paper called “Self-consistent field, including exchange and superposition of configu-
rations, with some results for oxygen,” Philosophical Transactions of the Royal Society
(London), A238, 229 (1939), and the general MC SCF theory was presented by Roy
McWeeny in a work called “On the basis of orbital theories,” Proceedings of the Royal
Society (London), A232, 114 (1955). As a classic paper in electronic correlation, we
also recommend an article by Per-Olov Löwdin, “Correlation problem in many-electron
quantum mechanics,” published in Advances in Chemical Physics, 2, 207 (1959). The
idea of the coupled cluster (CC) method was introduced by Fritz Coester in a paper
in Nuclear Physics, 7, 421 (1958), entitled “Bound states of a many-particle system.”
Jiří Čížek introduced the (diagrammatic) CC method into electron correlation theory
in a paper “On the correlation problem in atomic and molecular systems. Calculation
of wavefunction components in Ursell-type expansion using quantum-field theoretical
methods,” published in the Journal of Chemical Physics, 45, 4256 (1966). The book
“Three Approaches to Electron Correlation in Atoms” (Yale University Press, New
Haven, CT, and London; 1970), edited by Oktay Sinanoğlu and Keith A. Brueckner,
contains several reprints of the papers that cleared the path toward the CC method.
A derivation of the CC equations for interacting nucleons was presented by Herman
Kümmel and Karl-Heinz Lührmann, Nuclear Physics, A191, 525 (1972), in a paper
entitled “Equations for linked clusters and the energy variational principle.”
There is, however, one requirement that we believe to be a natural one for any
method. Namely,
any reliable method when applied to a system composed of very distant (i.e., non-in-
teracting) subsystems should give the energy, which is a sum of the energies for the
individual subsystems. A method having this feature is known as size consistent.a
Before we consider other methods, let us check whether our fundamental method
(i.e., the Hartree-Fock method) is size consistent or not.
Hartree-Fock Method
As shown on p. 417, the Hartree-Fock electronic energy reads as , while the total
energy is , where the last term represents a constant repulsion of the nuclei. When
the intersubsystem distances are infinite (they are then non-interacting), one can
divide the spinorbitals into non-overlapping sets , where means the molecular
spinorbital is localized on the subsystem and represents a Hartree-Fock spinorbital
of molecule , etc. Then, in the limit of large distances (symbolized by stands for the
operator of the interaction of the nuclei of molecule B with an electron, while , with
representing the nuclear repulsion within molecule A, and denotes the Hartree-Fock
energy of molecule A):
The zeros in the above formula appeared instead of the terms that vanish because
of the Coulombic interaction of the objects that are farther and farther from one
another. For example, in the mixed terms , the spinorbitals and belong to different
molecules, all integrals of the type vanish because they correspond to the Coulomb
interaction of electron 1, with the probability density distribution in molecule ,
and electron 2, with the distribution centered on molecule . Such an interaction
vanishes as the inverse of the distance; i.e., it goes to zero in the limit under
consideration. The integrals vanish even faster because they correspond to the
Coulombic interaction of with and each of these distributions itself vanishes
exponentially if the distance goes to infinity. Hence, all the mixed terms tend to
zero.
Thus,
We have learned, from the example given at the beginning of this chapter, that the
“genetic defect” of the mean field methods is that they describe electrons that ignore
the fact that they are close to or far from each other. For instance, in the two-electron
case previously considered, where we established the coordinates of electron ,
electron has a certain distribution of the probability density. This distribution does not
change when electron moves to a different position. This means that the electrons are
not “afraid” to get close to each other, although they should, since when electrons
are close, the energy increases (Fig. 10.1a,b).
Figure 10.1. Absence of electronic correlation in the helium atom as seen by the
Hartree-Fock method. Visualization of the cross-section of the square of the wave
function (probability density distribution) describing electron 2 within the plane ,
provided that electron 1 is located in a certain point in space: (a) at ; b) at . Note that
in both cases, the conditional probability density distributions of electron 2 are identical.
This means electron 2 does not react to the motion of electron 1; i.e., there is no
correlation whatsoever of the electronic motions (when the total wave function is the
Hartree-Fock one).
The explicitly correlated wave function (which we will explain in a moment) has the
interelectronic distance built in its mathematical form. We may compare this to making
the electrons wear spectacles.6 Now they avoid each other. One of my students said
that it would be the best if the electrons moved apart to infinity. Well, they cannot.
They are attracted by the nucleus (energy gain), and, being close to it, must be close
to each other too (energy loss). There is a compromise to achieve.