2.2. Linear approximation to Coulomb correlations
Strong electron–electron correlations in a free electron gas were dealt with by Bohm and Pines by subjecting the many-electron Hamiltonian to a series of canonical transformations. These transformations result in weakly interacting elementary excitations (plasmons) which represent collective elementary excitations of the electron gas.
Using the second quantized notation [
20] for the electron creation (
) and annihilation operator (
c), the Schrödinger equation for an N-electron free electron gas is
where
with the subscripts
i,
j,
k,
l denoting the set of four one-electron quantum numbers and the arguments (
) denoting the four coordinates of the
electron, three of which being the space coordinates and the fourth being the spin coordinate. The operator f is a
single-electron operator, similar to that in Equation (
1), but consisting of only the kinetic energy terms, the electrons being
free. In Equation (2b), a typical spin-orbital is represented by
wherein the spin part is either
for
i.e., ↑, or
for
i.e., ↓.
The second quantized Hamiltonian in Equation (2a) is equivalently written as
or more compactly using the Dirac notation for the integrals as
Without compromising the above Hamiltonian in any way, we can place the
most part of the two-electron interactions
in Equation (
1) in a one-electron operator
by writing
or briefly as
where
is a one-electron operator and
is only a small part of the full Hamiltonian. The notations employed in Equations (4a) and (4b) are self-explanatory. Essentially, the N-electron Hamiltonian is re-written such that it can be approximated by (
), since
is small; we shall treat it perturbatively. The choice of the operator
F is so made that the total energy functional
is minimized.
When
is neglected, the unperturbed ground state wavefunction of the N-electron system is expressible as a determinant:
wherein
with
or
. Observe that the one-particle eigenvalue
is doubly degenerate with the spin-orbitals for spin up and down being linearly independent. In the second determinant in Equation (
6), we have only re-designated the single-electron spin-orbitals.
Re-designation of the single-electron spin-orbitals in Equation (
6) is appropriate for closed shell atoms for which the random-phase approximation is applicable. Wave functions of the
excited unperturbed states are also Nth order determinants made up of eigenfunctions of Equation (
7), but with one or more
. In an ordered set of spin-orbitals, let us denote
and
. Thus, we denote a typical spin-orbital in the ground state Slater determinant by
p and an excited state spin-orbital corresponding to a single excitation by
q. The operator
is diagonal with respect to one-electron functions, and
. The choice of the operator
which makes the energy functional (Equation (
5)) a minimal is the one for which the matrix element
as can be shown using a variational method under the
frozen orbital approximation; i.e., for the excited state we use a Slater determinant only the
spin-orbital replaced by the excited
—
all other spin-orbitals in the Slater determinant (8) retain their occupancies.
The one-electron Hartree–Fock equation satisfied by the SCF ground-state spin-orbitals is
The four-dimensional integration
in Equation (10a) includes integration over the three (continuous) space coordinates and the discrete summation over the spin-coordinate. The second and the third terms on the left-hand side of Equation (10a) involve two-electron terms; the second is the Coulomb term and the third is the exchange term. We consider non-ferromagnetic systems so that the number of spin-up and -down terms is equal. The energies
are doubly degenerate with two linearly independent functions corresponding to up- and down-spins. Carrying out the summation over the spin variable, Equation (10a) simplifies to
in which we have written the inter-electron Coulomb interaction as
Now,
hence the HF SCF equation becomes
We write the momentum-dependent energies as
or equivalently as
, since
. The free electron gas is the only many-electron system for which the HF SCF equation can be obtained analytically. The linearization of the Coulomb approximation was developed by Bohm and Pines [
17] in which the positive charges of the nuclei were considered to be spread out uniformly over a volume
V as jellium (
Figure 1). The electron gas is also spread out over the volume
V in which the electron wavefunction is box-normalized.
Adding the jellium potential
in the HF SCF equation we get
The attractive jellium potential (second term on the left-hand side of Equation (
13)) exactly cancels the electron–electron Coulomb repulsion term (third term). We are then left with
where
i.e.,
Using (i) box-normalized wavefunctions
and (ii)
the Hartree–Fock equation for the free-electron gas (with exchange) in the positive jellium potential becomes
with the exchange term given by
being the electron momentum at the highest occupied energy level; viz., Fermi level.
Whereas the HF SCF energy of an atom described by the Hamiltonian (Equation (
1)) is
that of an electron gas in the jellium potential of the positive charges is
where
with
and
If we now consider the entire physical volume under consideration to consist of spheres of radius
(Seitz parameter in Bohr units), each having
one unit of electron charge, then
then the K.E. contribution to the average HF ground state energy
per electron in a free-electron gas is
and the exchange correlation energy per electron is
Thus, for electron gas in the SCF jellium potential, the average Hartree–Fock energy per electron is
A first order perturbative treatment gives essentially the same result as above. electron–electron exchange interactions reduce the energy below that of the Sommerfeld gas in a positive jellium potential; exchange energy is negative.
In the mid-1950s, Bohm and Pines improvised on the above model by considering a random mutual displacement of the centers of the positive and negative charge densities (
Figure 2). In the jellium potential, these are coincident; their mutual displacement can be considered to have been triggered by a random event, but once displaced, the positive and negative charges are set in oscillations of the plasma as the system seeks its original configuration. Bohm and Pines modeled these oscillations using a harmonic oscillator potential, inclusive of a dispersive wavelength-dependent frequency of the plasma oscillations.
The Hamiltonian for
N electrons in a volume
V together with a uniform positive charge background jellium distribution is
where
represents the many-electron part,
represents the jellium background and
represents the interaction between the electrons and the jellium background.The
N-electron system in the jellium background potential constitutes an electrically neutral system, but the relative displacements of the positive and negative charges allow for plasma oscillations of the electron gas. A mathematical device using the coefficient
in the exponential terms in Equations (
31)–(
33) is introduced to avoid some divergences; solutions are finally sought in the limit
. As a result of carrying out the integrals in Equations (
32) and (
33), the Hamiltonian (Equation (
30)) turns out to be
which manifestly diverges in the limit
. This is commonly referred to as the
-divergence.
It is most convenient to: (i) use the second quantized representations of the Hamiltonian for the bulk electron gas in a uniform positive background jellium potential (Equation (
30)) using electron creation operator
and the annihilation operator
,
being the momentum (in units of
ℏ) and spin quantum numbers; (ii) employ the Fourier representation of the screened Coulomb interaction that appears in Equations (
31)–(
33); and finally (iii) seek the limits
(specifically in this order, with
). Using the three steps described, after some tedious algebra, one finds that terms corresponding to momentum transfer
in the two-electron interactions term for which
in the
part of the Hamiltonian cancels the abovementioned
-divergence, and along with the limits sought as per (iii), the Hamiltonian in the second quantized notation is
Scaling
and
allows us to write the Hamiltonian as
where we have introduced a dimensionless variable
,
being the Bohr radius. In the high density limit
, the second term in Equation (
37) can be treated using first order perturbation theory even if the electron–electron interactions in the second term are quite strong. The result in the first order turns out to be essentially the same as in Equation (
29), but higher order perturbation theory does not converge. Therefore, Bohm and Pines developed a non-perturbative approximation by carrying out canonical transformation of the Hamiltonian to represent pseudoparticles (elementary excitations of the many-electron gas) called
plasmons which represent collective oscillations of the electron gas. The approximation involves linearization of the Hamiltonian concomitant with the neglect of certain terms whose phases are random and hence cancelable. Prior to discussing the canonical transformation of the Hamiltonian, we briefly visit their earlier semi-classical model which helps build insight in the
linearization process and also in the
approximation involved in the concomitant cancellation of terms having
random phases.
In the semi-classical model, both the electron gas and the positive charge in the bulk medium are considered to be uniformly spread over the entire volume with their collective centers coincident. An incidental movement of the electron density
(
Figure 2) sets in oscillations of the electron gas described by the classical equation of motion:
wherein
denotes the surface charge density
, the static average volume charge density being written as
The zero-point energy of the plasma oscillations is
wherein the natural frequency of the plasma oscillations is
Using the Fourier decomposition of the inter-electron interaction,
the potential energy of the ith electron due to one electron charge
uniformly smeared throughout the box is
, i.e.,
.
Hence, the potential energy due to
all the electrons is:
where
, except for
. The term corresponding to
cancels the positive jellium; hence the potential energy of the
electron due to all the electrons
and the positive background is
Now, in terms of the electron field operators, the total number of electrons is
and the electron density is
wherein we have used the Fourier expansion of the charge density with the Fourier components being given by
Identifying the force on the electron force as the negative gradient of the potential in Equation (41b), we arrive at the semi-classical equation of motion for the harmonic oscillator
which translates to the equation of motion for density fluctuations of the Fourier components in Equation (
43):
The first term in Equation (
45) is quadratic in
k. It can be ignored if
; i.e., if one limits
k to be small. The ‘upper bound’ on the wave number, denoted by
, of the plasma oscillations is
Now, the integral of the charge density over the entire space adds up to the total number of electrons N, i.e.,
corresponding to
Using the fact that the Fourier expansion of the charge density
with
and
In the third term on the right-hand side of Equation (
45), we have
and
, which involve oscillatory terms consisting of phase factors of modulus unity. It is like carrying out a sum of vectors in a complex plane whose directions are
random, and one expects this to be a zero-sum addition. Thus: (i) neglecting the first term (enabled by placing an upper limit on
k) and (ii)
linearizing Equation (
45) (i.e.,
neglecting the quadratic terms, taking advantage of the
random-phases), we obtain
which essentially is an equation of motion for a simple harmonic oscillator. Quantized collective excitations of this system are elementary excitations. They are pseudo-particles called plasmons. The frequency of plasma oscillations is
The zero-point energy of the plasma oscillations is
, where
In the Bohm–Pines method of canonical transformation of the Hamiltonian discussed below, the significance of the approximation involving linearization of the Hamiltonian concomitant with the neglect of terms having random phases gets further accentuated.
We have seen that the Hamiltonian for a bulk electron gas in a uniform positive background jellium potential is
and noting that the
term adds up to
N, the total number of electrons, we arrive at
where the last form is obtained by adding and subtracting the term corresponding to the
.
The quantum problem to be solved for the above Hamiltonian is
Bohm and Pines recognized that the classical model which yielded plasma oscillations described by Equation (
49) would be an approximation to a quantum model. One ought to seek a transformation of the Hamiltonian (Equations (51a) and (51b)) such that plasma oscillations appear explicitly as a set of
Hamiltonians for simple harmonic oscillators for various
values limited by Equation (
46). They therefore proposed canonical transformations
of the Hamiltonian in Equations (51a) and (51b) to a
new set of generalized coordinates Q and momenta P such that the new quantum Hamiltonian would have the
form
which is characteristic of the Hamiltonian for a simple harmonic oscillator represented by the Hamiltonian (in units of
)
The transformation we seek is not inspired by actual measurements of the new coordinates and momenta; it is motivated only by seeking the form in Equation (53a). Hence, the operators
Q &
P need not necessarily be Hermitian. The Bohm–Pines strategy consists of starting with an
auxiliary Hamiltonian
with
We do not demand the operators
Q &
P to be Hermitian. Instead, by choosing
and
we see on recognizing that the summation over
and that over
is equivalent considering the symmetry in the momentum space that,
is Hermitian, even if
Q and
P are not. The wavefunction depends only on the original set of electron coordinates; it cannot depend on any additional degrees of freedom. It is therefore judicious to employ subsidiary conditions
however, limited by
. The derivative operator is the momentum,
hence
and
which is just the uncertainty relation for canonically conjugate coordinates and momenta. Use of Equation (56c) in Equation (54a) ensures that
and the Hamiltonian
describes the same quantum system. We seek a transformation affected by an operator
with
and
which gives
and we see that the transformation is unitary. It follows that
and the
component of the operator is
Essentially, under the transformation under consideration, the operators
remain invariant, but
and
are different. We now ask what the transformed Hamiltonian,
is. After some tedious algebra, it turns out to be
where
and
The new Hamiltonian has a manifestly complicated form. The term
K (Equation (
65)) is quadratic in the new coordinates and has
random phases which would cancel out in a
linearization process, as explained earlier in the context of the classical model and arrived at Equation (
49). Linearization of the
makes it possible to drop the operator K and justifies the term
random-phase approximation. The rest of the Hamiltonian is
in which
(Equation (
63)) represents a set of
quasi-particles interacting via
short-range screened-Coulomb potential and given by
and
is the self-energy of the electron gas.
is accounted for by a further canonical transformation of the Hamiltonian (in which
K is ignored) written in terms of transformed coordinates and momenta. Using the
random-phase approximation concomitant with
linearization of the transformed Hamiltonian (i.e., neglect of quadratic terms in the newer set of coordinates), the
term gets dropped, but in the process, the first two terms get somewhat modified, and the new approximate Hamiltonian becomes
where
and
expresses a weak
k-dependent dispersion of the plasma frequency.
We have another subsidiary condition, similar to Equation (56c):
The kinetic energy part in the newer Hamiltonian is diminished by the factor
; actual calculations show that
and hence the kinetic energy part is reduced by about 8%. The long-range part of the interaction is what leads to the plasma oscillations corresponding to the first curly bracket in Equation (
69).
denotes the short-range screened-Coulomb interaction between the new pseudo-particles, which are elementary excitations called plasmons. The Hartree–Fock approximation accounted for only the static part of the density fluctuations of the collective behavior of an electron gas. The frozen-orbital approximation that leads us to the Koopmans theorem highlights this approximation which limits it to the neglect of the Coulomb correlations. The method of canonical transformation of the Hamiltonian enables us address the Coulomb correlations albeit in an approximate manner by systematic and straightforward interpretation of Equation (
69) in which the first curly bracket represents the collective oscillations of the electron gas resulting from the long-range part of the Coulomb interaction. A quantum of these oscillations is the plasmon. The second curly bracket represents the Hamiltonian for the screened Coulomb interaction, and the third represents the self-energy of the electron gas.
The Bohm–Pines method elucidates the physical content of the
random-phase approximation (RPA) and the linearization process it involves. There are other methods of arriving at the RPA, such as the Equation of Motion method [
13] and the Greens function method [
11]. The approximation is equivalent to summing over all the ring diagrams (along with the diagrams for the exchange interaction corresponding to each Coulomb term) in Feynman diagrammatic perturbation theory. Another equivalent approach to the RPA(E) results from the linearization of the Time-Dependent Hartree–Fock (TDHF) method developed by Dalgarno and Victor [
14] and Amusia [
15], and its relativistic version, namely the linearized Time Dependent Dirac–Hartree–Fock (TDDHF, often briefly denoted as TDDF) developed by Johnson and Lin [
4]. In the next section, we summarize the linearized TDHF/TDDF approximations.
2.3. Linearization of TDHF and That of TDDF Formalism
The Hartree–Fock self-consistent field (HF-SCF) method accounts for correlations in many-electron dynamics that result by demanding that a many-fermion wavefunction must be anti-symmetric with respect to every exchange of pairs of the elementary particles. These correlations are therefore equivalently referred to as exchange correlations or as statistical (Fermi–Dirac) correlations. The Pauli Exclusion Principle automatically follows from it; hence, they are also sometimes called the Pauli correlations. The HF-SCF, however, only accounts for a static average of the density fluctuations of the many-electron system and thus leaves out what are known as
Coulomb correlations. In the previous section, we discussed the RPA which employs a
linearization technique and provides for a very successful methodology to account for the Coulomb correlations. We now proceed to discuss the RRPA [
4], which is essentially based on
linearization of the Time-Dependent Dirac–Hartree–Fock (TD-DHF, or just TDDF) family of coupled integro-differential equations. The linearized TDHF (RPAE) approximation [Amusia] was the precursor to the RRPA; it employs the
linearization of the TDHF family of equations. The RPAE employs two-component spin-orbitals obtained as SCF solutions to the non-relativistic Schrödinger equation for the many-electron system. These spin-orbitals go into the construct of the Slater determinantal single-configuration wavefunctions. The RRPA is the relativistic extension of the RPAE. It employs four-component bi-spinor SCF solutions to the Dirac equation for the many-electron system which appear in the Slater determinant. The bi-spinors (spin-orbitals) are, however, admittedly time-dependent allowing for density fluctuations of the many-electron system. The resulting TD-DHF (TDHF) equations are non-linear; making a
linear approximation to the TD-DHF (TDHF) equations result in the RRPA (RPAE).
The time-independent DHF equations for an
N-electron closed-shell atomic system are
where
represent the four-component bispinor,
represents the Dirac Hamiltonian,
represent the DHF eigenvalue and
represents the DHF potential, given by
and the prime denotes the argument over which integration is carried out. Solutions to the DHF equations are best represented by a Slater determinant
where
with
The electron densities of the DHF one-electron bispinors (spin-orbitals) represent only a time-average since the DHF model ignores electron correlations. Due to the electron correlations in the initial and the final state of a transition affected by what may be represented by an interaction operator
with the positive and negative frequency driving terms respectively denoted by
being the vector potential of the electromagnetic field, the DHF orbitals must be represented by time-dependent functions described by
The dots
… at the end in Equation (
79) represent higher harmonics. If we
rebuild the (Time-Dependent) Dirac–Hartree–Fock scheme with all the higher harmonics, we obtain Non-Linear Time-Dependent Dirac–Hartree–Fock equations. Dalgarno and Victor [
14] proposed the RPA linearization of the (non-relativistic TD-HF equations by dropping the higher harmonics. Following a similar logic, Johnson, Lin and Dalgarno [
2,
3,
4,
5] introduced the very same linearization in the TD-DHF system of coupled integro-differential equations for the orbitals
.The orbitals
represent perturbation of the DHF orbitals due to the positive frequency part of the perturbation, and the orbitals
represent perturbation of the DHF orbitals due to the negative frequency part. The linearized Time-Dependent Dirac–Hartree–Fock (L-TslatdetD-DHF) equations are
with
which includes the Coulomb correlations that are omitted in the DHF method. The factors
in Equation (
80) are the Lagrange’s variational multipliers, introduced in the algebraic equations to ensure orthogonality of the perturbed orbitals
with respect to the unperturbed ones
. Omission of the driving terms
gives us the fundamental RRPA equations:
The eigenvalues of the RRPA equations provide the linear approximation to the excitation spectra in both the discrete and the continuum. The positive and negative components
of the eigenfunctions describe the correlations which are omitted in the DHF formalism, respectively, in the excited state (both discrete and continuum) and in the initial states. The amplitude of transition from an initial state to the excited state described by the RRPA function
corresponding to the frequency
brought about by the interaction (Equation (76)) is
i.e.,
It is a very important property of the (R)RPA equations that the transition matrix element is invariant under gauge transformations of the electromagnetic potentials. In actual calculations, one often employs truncated RRPA in which only the most important interchannel coupling is used. This leads to a slight disagreement in the estimation of the transition matrix elements in the length gauge and in the velocity gauge.
It is very convenient to have a pictorial representation of the correlations that are addressed in a many-electron theory. The diagrammatic representation developed by Feynman, first presented in the Spring of 1948 at the Pocono Conference, is suitable for our purpose. In AMO sciences, we (normally) do not work with positrons, but there are ‘hole’ states which are vacant states normally occupied by electrons. Thus we represent evolution of atomic states by vertical solid lines with reference to a time-axis going from the bottom to the top (left to right is an alternative convention). The atomic state lines are sometimes referred to as the ‘trunk’ of the diagram. A vertex in the diagram represents an intersection of a photon wavy line and the trunk. Particle lines point upwards and hole lines point downwards. Summing over only the ring graphs, as shown by Gell-Mann and Brueckner [
21] has precisely the same effect as the linearization approximation that results in the random-phase approximation introduced by Bohm and Pines. Electron correlations are interpreted by recognizing that electrons exchange virtual photons which mediate the interaction between the electrons. The electromagnetic interaction is treated at the level of quantum theory. A positron is anelectron propagating backward in time.
Figure 3 shows some of the lowest order diagrams [
22] which contribute to the RRPA matrix elements.
Time-forward diagrams represent correlations in the final state in which configuration-interaction in the continuum is taken care of. In the RRPA, this corresponds to interchannel coupling. The time-backward diagrams represent correlations in the initial state. Variants of the RRPA that offer some advantages include the R-MCTD (Relativistic Multi-Configuration Tamm–Dancoff) method [
23] and the RRPA-with-relaxation (RRPA-R) method [
24]. The particle and the hole creation and annihilation operators that are referenced in the ring diagrams (
Figure 3) are defined with respect to a vacuum consisting of a closed-shell Hartree–Fock (or Dirac–Hartree–Fock) fermion system, such as that described by Equation (
8).
Two-particle two-hole correlations are included, accommodating both creation and annihilation of a particle-hole pair. The focus of this article is to discuss the linearization approximation involved in the RPA and illustrate a few applications of the RRPA; hence, we omit an elaboration of the RRPA-R and R-MCTD techniques.