2.2. Atomic State Function and Configuration State Functions
Relativistic atomic structure theory in
grasp treats the many-electron atom within Furry’s bound state interaction picture [
75] of quantum electrodynamics. The formalism resembles that of non-relativistic atomic theory but is based on a Hamiltonian that uses the Dirac operator
to describe the relativistic dynamics of electrons in the field of other charged particles, as described in
Section 2.3 below.
The program structure draws from the one of the non-relativistic program
atsp2K [
28].
grasp assumes that the wave function of an atomic state
, with
being its identifying label,
J the total angular momentum quantum number,
the total magnetic quantum number and
its parity, is approximated by an atomic state function (ASF),
, which is a linear combination of configuration state functions (CSFs) in
-coupling
For each CSF
, the multi-index label
contains all the needed information on its structure, i.e., the constituent subshells with their symmetry labels and the way their angular momenta are coupled to each other, as described in
Section 2.4. The main difference with the non-relativistic program is the need to use four-component spinor orbital functions (
5) to build the configurational state functions (CSF) with accompanying technical machinery to express the variational equations in computable form.
grasp can perform both variational MCDHF and configuration–interaction (CI) calculations. The latter assume that all data on the CSFs, radial orbital functions and angular coupling coefficients have previously been computed. The CI approach is therefore limited to the eigenvalue problem
where
is the matrix of one of the chosen Hamiltonians of
Section 2.3 in the CSF space spanned by a given set
and
is the eigenvector corresponding to the eigenvalue
E, the total energy. Note that the magnetic quantum number
is irrelevant for field-free atoms, as assumed for the MCDHF and CI approaches considered in the present section. The mixing coefficients
appearing in (
6) are therefore a priori
-independent. However, the electronic magnetic quantum number
might become relevant and affect the electronic structure when external perturbations and/or hyperfine effects due to the coupling of the electronic
and nuclear
angular momenta are considered, as described in
Section 3.2.
The MCDHF approach determines the orbital components
of the one-electron orbitals (
5) spanning the CSF space, permitting the numerical construction of the CSFs. The radial functions
and
are defined on a grid
where
,
B = 0.05, and
are parameters that define the default grid. The default grid is usually sufficient for light to medium heavy systems. For the heaviest systems, it may be necessary to use a few thousand points, as is discussed in the accompanying
grasp manual §13.5. The radial functions are obtained by solving a set of differential equations (see
Section 2.7) with finite difference methods [
26,
73]. Just as the wave functions of a one-electron equation, the orbitals in
grasp make up an orthonormal set.
2.3. Dirac-Coulomb, Dirac-Coulomb-Breit Hamiltonians and QED Corrections
In
grasp, the Hamiltonian used for self-consistent calculations, see
Section 2.7, is the Dirac–Coulomb Hamiltonian
The nuclear potential
results from a nuclear charge density given by a two-parameter Fermi distribution function [
76]. The nuclear charge, the two parameters of the Fermi distribution, the mass of the nucleus, the nuclear spin and the magnetic dipole and quadrupole moments are, in the
grasp code suite, defined by the program
rnucleus and saved in a file
isodata (see accompanying manual §2.2 and §8.1).
Corrections to the above Hamiltonian can be included in configuration interaction calculations; see
Section 2.8. To represent magnetic interactions and retardation effects [
77,
78,
79], one can add the so-called transverse photon interaction, which is correct to order of
.
In this expression, the
∇-operators act only on
, while
represents the energy of the virtual photon exchanged between two electrons, as described in QED [
78]. In the low-photon energy limit, when
, the expression (
10) reduces to the Breit interaction
Adding the transverse photon interaction to the Dirac–Coulomb Hamiltonian gives the Dirac–Coulomb–Breit Hamiltonian
Additional important QED contributions are the self-energy (SE) correction and the vacuum polarization (VP). The SE correction, which is the result of emission and absorption of a virtual photon by the same electron, is given as a sum of one-electron corrections weighted by the fractional occupation numbers of the one-electron orbitals in the wave function. The VP correction, which is related to the creation and annihilation of virtual electron–positron pairs in the field of the nucleus, can be described by a correction to the Coulomb potential. The above QED terms [
80,
81,
82,
83] are included in the configuration interaction
rci code, as implemented in the original
grasp program [
78], to yield the final Hamiltonian
Nuclear recoil effects [
68], hyperfine interactions [
66], and symmetry breaking interactions with external magnetic fields [
84] are included in first-order perturbation theory; see
Section 3.
2.4. Building Configuration State Functions
Adopting the following notation for a subshell containing
w equivalent electrons
a general relativistic configuration consists of
m groups of equivalent electrons
where
is the occupation number of the relativistic subshell
i [
30].
The relativistic Hamiltonian for an N-electron system commutes with the total angular momentum operator , and the solutions to the wave equation can be taken as eigenfunctions of and with as good quantum numbers. Atoms are fermionic systems, and wave functions are required to be antisymmetric with respect to permutations of the electron co-ordinates. Not all states arising from the angular momentum coupling are permitted, because the Pauli exclusion principle selects only fermionic states within each subshell of equivalent electrons. A configuration state function (CSF) is the simplest approximation of a many-electron wave function being both antisymmetric and an eigenfunction of and . Such a CSF can be built
By using the well-known vector coupling techniques of angular momentum theory [
85] to couple sequentially, from left to right, the subshell angular momenta
,
associated with the
m antisymmetric subshell wave functions
Antisymmetrize the resulting coupled products through the permutations restricted to the exchange of electron coordinates involving
different subshells [
86].
In each wave function (
15),
is the seniority number [
87,
88], making it possible to discriminate states arising from the same relativistic subshell
configuration having the same
J-value. As shown in
Table 2 extracted from [
89], the seniority
and
are needed to discriminate the two states
(or
) arising from
, i.e., from any subshell
or
. If the seniority classification fails, i.e., if two levels appear with the same
values, additional quantum numbers
are usually introduced to unambiguously designate the state considered [
90]. A systematic approach to classify the states uses the theory of Lie groups [
90] in which
contains the irreducible representations labels of the invoked chain of groups. Unfortunately, a complete classification scheme remains an open problem [
91]. Therefore, from a practical point of view, we will restrict
to a simple number,
, as suggested in [
89,
92]. As revealed by
Table 2—see also Table A.5 of [
73]—this case arises for configurations
or
for which the same seniority,
, is assigned to the two allowed
(or
) values.
Due to the current restrictions of the
grasp2018 package “
occupied subshells with are restricted to a maximum of two electrons”,
will never be seen in the output file
rcsf.out produced by the
rcsfgenerate program; see accompanying manual §3.2. However, we deliberately keep the
notation, since the spin-angular library [
93] allows us to deal with the
subshell without any occupation restriction.
The above two-step procedure leads to the most general form of a CSF [
30]
In this notation,
is the compact representation of a CSF, collecting all the needed information for each subshell together with all the intermediate quantum numbers that unambiguously define the CSF. The parity
is often omitted in the CSF notation since it can be easily deduced from the collection of angular momenta hidden in
, using
.
Each subshell wave function (
15) can be built using a recursive coupling method in terms of fractional parentage coefficients (CFPs) [
30,
73,
86]. In addition to the seniority classification, there exists an alternative representation of the same subshell wave function provided by the quasi-spin formalism [
91,
94,
95] that offers many advantages explored in the spin-angular algebra [
93,
96]. In the quasi-spin classification scheme, also used in
Table 2, the subshell wave function (
15) is rewritten as
where
carries the same information as
through the following relations [
73,
91],
A nice property of the quasi-spin operator
Q is that the ladder operators
connect subshell wave functions with occupation numbers
w and
having the same seniority
. The Wigner–Eckart (WE) theorem can be applied in the space of quasi-spin for all individual subshell states much in the same way as for
J-space, allowing an efficient reduction and factorization of matrix elements and CFP matrices [
96].
2.5. Second Quantization and Composite Tensor Operators
With complicated CSF structures, second quantization is the most convenient way to handle the mathematics of CSFs and their matrix elements. An extensive
-coupling treatment, adequate for
grasp users who have no need for the technical details, can be found in ([
73], §6.8, pp. 368–390). Other treatments, mainly focusing on
-coupling, are the original lectures of Brian Judd [
94] and Rudzikas’s monograph [
91] giving the Vilnius way of doing things. The
rangular code from [
96] is based on the quasi-spin constructions, replacing the older formulation following Fano [
86] used in earlier versions of the
atsp2K [
28] and
grasp [
43] packages.
We define the electron creation operator
as the operator that generates the Dirac orbital when acting on the vacuum state
of the electron field
This is destroyed by the annihilation operator
so that
Creation and annihilation operators anti-commute so that
Thus, interchanging neighboring pairs of operators introduces a sign change, for example
(implying
) so that the algebra of creation and annihilation operators is what we need for manipulating antisymmetric states obeying the Pauli exclusion principle: one electron only in each orbital state. Products
in which the operators are all different therefore represent
n-electron determinants from which we can generate more complex
n-electron states.
Subshells of Dirac electrons are useful subdivisions of CSFs, as the set of
states
is
equivalent in the sense that each of them can be expressed in terms of the others under a rotation of axes. Focusing on a single subshell, we can drop
n and consider the
(
) equivalent antisymmetric states
In the language of group theory ([
73], §6.8.2), this belongs to a reducible representation
(
w factors). This can be decomposed as a direct sum, a Clebsch–Gordan series, of irreducible representations
, which can be used to characterize the subshell states—see the list in
Table 2. These can be further distinguished by the seniority number
, which is defined for each subshell configuration as the lowest value of
w for which a particular
J appears. As explained above, the few cases in which
fails to identify the subshell state uniquely are labeled
in this table.
The operators
can be regarded ([
73], §6.8.3) as defining irreducible tensor operators,
of rank
j. From these, we can build composite tensor operators using angular momentum theory such as
and more complex operators. Omitting the
rank in the notation of creation and annihilation operators acting on the same subshell (
), the three operators
satisfy the commutation relations of the
quasi-spin vector operator , enabling us to classify subshell states in terms of the quantum numbers
of (
18) and (19), which are equivalent to the classification
in terms of the seniority scheme. The main use of the quasi-spin classification is that
is related to the occupation number
w so that relations between states of the subshell configurations
and
can be expressed entirely in terms of a
-symbol in quasi-spin.
2.6. Calculation of Matrix Elements
The above construction of the CSFs makes it possible to derive analytical expressions for matrix elements of one- and two-electron interaction operators. A one-electron irreducible tensor operator has the form
where
k and
q are tensor indices. Examples of one-electron tensor operators are the first part of the Dirac–Coulomb Hamiltonian and the operators describing the interactions with the nuclear dipole and quadrupole moments.
A two-electron scalar operator has the generic form
where
is a scalar two-body interaction between two electrons
where
is the scalar product of two tensors
and
acting on electron coordinates
i and
j. Examples are the Coulomb interaction, the Breit interaction and the specific mass shift interaction, arising from the nuclear recoil effect.
The matrix element of an irreducible tensorial operator between two CSFs,
and
, can be factorized thanks to the WE theorem [
85],
where
is a reduced matrix element independent of the
M-quantum numbers and of the
q tensorial component. Another version of the WE theorem can be found in Brink and Satchler [
97]
Equating the l.h.s of (
28) and (
29) provides a relation between the two definitions of the reduced matrix elements (RMEs)
Many formulas in various papers may appear different due to this ambiguity. In the present work, we adopt Edmonds’ formulation (
28) of the WE theorem [
85] that also fits with Racah [
98,
99], Judd [
100], Cowan [
101], or Rudzikas [
91], while Brink–Satchler WE theorem (
29) is coherent with Rose [
102], except for an extra
phase factor (As pointed out by Judd [
103], such a phase factor leaves the relation between RMEs phase-free when
k is integral, as is almost always the case.
k is indeed integer (even or odd) for all irreducible tensorial operators representing physical quantities, invariant under a
rotation. However, as observed by Judd [
94] and Rudzikas [
91], phase systems can be crucial for second quantization operators.).
The reduced matrix element of a one-electron operator between the CSFs, in turn, can be expressed as a weighted sum over the active orbitals of one-electron reduced matrix elements
The reduced matrix element
depends only on the orbitals and on the nature of the operator. It can be further reduced to a radial integral times a matrix element involving the spherical spinors of the two orbitals. The spin-angular coefficients
contain all information about the configuration and the angular couplings. They can be expressed in terms of CFPs and recoupling coefficients necessary to match subshell states in the two CSFs [
104,
105]. These coefficients are computed by routines of the
grasp librang90 library, which are then called by the
rangular program for MCDHF approach, the
rci code for CI calculations, the
rbiotransform and
rtransition programs for transition properties, the
rhfs and
hfszeeman95 codes for hyperfine parameters and magnetic interactions, the
ris4 programs for isotope shifts and the
rdensity program for radial electron densities and natural orbitals, see accompanying manual §1.3 and §2.2.
For scalar two-electron operators, the application of the WE theorem (
28) is trivial
Similarly to (
31), the two-body reduced matrix elements can be expressed as [
73]
with
The summations over
in (
33) are running over the relativistic subshells
while the summation over
in (
26) runs over the electron coordinates
r. The effective interaction strength,
, is specific to the nature of the interaction and involves only the active orbitals [
73,
106]. It can be written in terms of a radial double integral
and factors involving matrix elements of the spherical spinors of the active orbitals. The spin-angular
coefficients are computed by the routines of the
librang90 library, which are then called by the
rangular program for MCDHF approach, the
rci program for CI calculations, and the
ris4 program for isotope shifts; compare to the accompanying manual §2.2.
The Coulomb interaction
where
is the renormalized spherical harmonic, which fits with the generic form (
27), whereas the Gaunt and Breit interactions involve composite tensor operators,
, with more convoluted radial parts [
73]. The magnetic interactions are therefore more complex but can similarly be reduced to angular factors multiplied by two-electron radial integrals (see Sections 6.4 and 6.5 of [
73] for the decomposition and for individual interaction strengths, respectively).
In second quantization, operators such as the angular momentum
are expressed in the form
operating on the ket (
23) to the right and on the corresponding bra to the left. Matrix elements using CSFs as in (
16) require detaching an
active electron orbital, say
from a subshell, matching one of the kets,
, in (
36). Actions are similar with the bra. This is accomplished ([
73], eq. (6.8.30)) by separating
using
where
where
is a CFP. The latter can also be expressed in terms of reduced matrix elements of creation and annihilation operators.
The construction (
16) just couples the subshell states in the order chosen by the CSF generator. Decoupling one electron from a ket subshell introduces a
-symbol, a product of the active electron creation operator, and a CFP in the subshell position. Moving the active creation operator to the end of the subshell sequence nearest to the interaction operator where it selects the right matrix element involves interchanging with other creation operators, introducing a factor
at each interchange. The interchanges introduce an overall phase factor, while the replacement of the original subshell state by the parent left behind changes the CSF coupling scheme, requiring a recoupling coefficient. Further discussion may be found in ([
73], §6.9).
A powerful spin-angular algebra based on angular momentum theory, on the concept of the irreducible tensorial sets, second-quantization in a coupled tensorial form, quasi-spin formalism and Wick’s theorem, reduced (in quasi-spin space) coefficients of fractional parentage and on a generalized graphical method, is used in
grasp to evaluate the
and
coefficients appearing in (
31) and (
33), as described in detail in [
89,
96,
106,
107]. The corresponding spin-angular library, implemented in
grasp2018, is fully documented by Gaigalas [
93] in the present Special Issue. Interesting illustrations on how this spin-angular algebra is applied for some physical quantities can be found in [
108,
109] in
- and
-coupling, respectively.
2.7. Multiconfiguration Dirac–Hartree–Fock
In the MCDHF method, the wave function of an atomic state
is approximated by the atomic state function (
6). For spectrum calculations, the orbital optimization often targets simultaneously several atomic states that may belong to different
J symmetries,
(suppressing the
and parity
quantum numbers for brevity). In this scheme, the different ASFs belonging to the same
J (
) are chosen to be orthonormal, so that
where
is the column vector collecting the mixing coefficients
for a given ASF. The energy of the atomic state
is
where the Hamiltonian reduced matrix
has the elements
The reduced matrix elements (
40) can, as discussed above, be expressed in terms of angular coefficients and radial integrals [
30,
73]
where now the factors involving matrix elements of the spherical spinors are absorbed in the angular coefficients. The one-electron radial integrals are given by
The two-electron radial double integrals, the so-called Slater integrals, are given by
with
defined by
In the integral, and are the smaller and larger of r and s, respectively.
Multiconfiguration methods are energy driven. Introducing the radial orthonormality condition
and applying the variational principle on the statistically weighted energy functional of the targeted states
where Lagrange multipliers
are introduced to ensure the orthonormality of the orbitals, yields the equations for the radial functions
, see [
30,
73]
The potential
consists of three terms
where the variations of the
integrals weighted with the angular coefficients and the state averaged expansion coefficients contribute to
. Variations of the other
integrals and the off-diagonal
integrals, again weighted with the angular coefficients and the state-averaged expansion coefficients, contribute to the exchange operator
[
110].
is the generalized occupation number of orbital
a and
are energy parameters related to the Lagrange multipliers.
In B-spline methods [
111], once off-diagonal Lagrange multipliers have been dealt with, the radial functions
,
, are obtained as the solution of an eigenvalue problem. However, the differential equation method used by
grasp, where the radial functions are defined as a vector of values on a grid, requires that the Dirac equation be rewritten as a pair of first-order differential equations [
33,
73,
112]. In the first equation, we introduce
as a non-homogeneous term of the differential equation. Similarly, for the second equation, we let
. Furthermore, it is customary to have the coefficient of the highest derivative be unity, so that
As a last step, the first equation should also be multiplied by , which considerably changes the symmetry. In the self-consistent iterative method, , , are computed from current estimates and the orbital energy , and updated radial functions are solutions of the differential equation.
For bound states, it is required that the solutions are square integrable. A necessary, but not sufficient, condition is that the radial amplitudes vanish as
and
. For a point charge nuclear model, a rigorous analysis of the radial behavior at
leads to power series of the form
where
and
are determined by coupled linear algebraic equations. For a more realistic model of the nuclear charge distribution, other conditions apply [
73]. The Equation (
47), together with the accompanying boundary conditions, are solved iteratively on the radial grid (
8) by a finite difference self-consistent field (SCF) procedure. Briefly, given initial estimates of the radial functions, the Hamiltonian matrix with elements (
40) is constructed and diagonalized to give expansion coefficients of the CSFs for each of the ASFs. Improved estimates of the radial functions are then obtained by solving Equation (
47). The two last steps are repeated until the energy of the states and/or the radial functions do not change anymore.
It is desirable to optimize all orbitals simultaneously, which is sometimes referred to as the “full variational” (FV) approach. However, due to numerical convergence issues, the MCDHF method often employs a layer-by-layer (LBL) strategy [
113,
114] (see
Section 4.4, and also in §3.5 of the manual [
72]), in which only the newly introduced orbitals for the layer considered are optimized while the remaining ones are kept frozen. In this context, a layer is a set of new orbitals to be optimized consisting of one orbital per angular momentum symmetry. The LBL approach is attractive as the computation time for each new layer is much shorter than the corresponding computation time of the FV approach. The price to pay for the LBL strategy is a larger active set of correlation orbitals to compensate for the lost degrees of freedom.
Although
grasp focuses on atomic bound states, the MCDHF equations can generate continuum orbitals provided different boundary conditions at
.
grasp ASFs can then be used as target states in collision calculations in codes such as
R-matrix [
73,
115]. In the
grasp suite of programs, the MCDHF equations are solved by the
rmcdhf program, which reads the nuclear parameters, the CSFs list, as generated by the
rcsfgenerate program and the necessary angular data, produced by
rangular, from disk files. The initial estimates of the radial orbitals for the SCF procedure are generated by the
rwfnestimate program and can be taken as screened hydrogenic functions, functions from a Thomas–Fermi calculation or converted non-relativistic orbitals; see the accompanying manual §3.3. For an overview of the program and file flows for the MCDHF calculations, see Figures 1 and 2 of the manual.
2.8. Configuration Interaction
The MCDHF calculations are used to generate an orbital basis. Given this basis, the final wave functions for the targeted states are obtained in relativistic configuration interaction (CI) calculations based on the Dirac–Coulomb–Breit and QED Hamiltonian (
13). For that step, the CSF expansions (
6) may be extended in comparison with those used in the previous MCDHF calculations to capture higher-order correlation contributions or to test the adequacy of correlation models. In CI calculations, the matrix elements of the CSFs are computed using the methods described in
Section 2.6. The expansion coefficients of the CSFs in the targeted states are then obtained by diagonalizing the Hamiltonian matrix. The number of CSFs depends on the shell structure of the atomic system in question as well as on the model for electron correlation. Even for moderate calculations, the number of zeroes in the interaction exceeds the non-zero values, and sparse matrix methods are used. For accurate calculations, a large number of CSFs are required, leading to very large matrices. To handle these large matrices, the CSFs can a priori be divided into two groups. Let
M be the size of the total CSF space. The first group,
P, with
m elements (
) contains CSFs that account for the major parts of the wave functions. The second group,
Q, with
elements contains CSFs that represent minor corrections. Allowing interaction between CSFs in group
P, interaction between CSFs in group
P and
Q and diagonal interactions between CSFs in
Q gives a matrix
where
. The restriction of
to diagonal elements results in a huge reduction in the total number of matrix elements and corresponding computational time. The assumptions of the approximation and the connections to the method of deflation in numerical analysis are discussed in [
30]. The structure of (
51) is reminiscent of the second-order Brillouin–Wigner perturbation theory [
116,
117,
118]. This form of the CI matrix has been available in the non-relativistic and relativistic multiconfiguration codes for a long time [
41,
119], where the
P and
Q groups were named “zeroth-order” and “first-order” sets, respectively. Our MCDHF/CI methods are therefore referred to as ‘Zero-First’ methods [
120], i.e., ZF-MCDHF or ZF-CI.
grasp uses sparse matrix methods for storing the interaction matrix in which only non-zero off-diagonal elements are stored by column for the upper matrix, taking advantage of matrix symmetry. The order of the CSFs is not important except when initial estimates of the eigenvectors are required, in which case up to 4000 of the first CSFs are considered, depending on a program parameter. Special iterative eigensolvers are used that determine only selected eigenvalues in the lower portion of the spectrum [
121] and matrix–vector multiplication with sparse matrices. Much of the efficiency of variational methods relies on the sparsity of the interaction matrix. In the
grasp suite of programs, the CI calculations are performed by the
rci program. Nuclear parameters, the CSFs list, as generated by
rcsfgenerate and radial orbitals, as produced by
rmcdhf, are read from disk files. Angular data needed to compute the Hamiltonian matrix elements are computed on the fly by calls to the routines of the
librang90 library. The program and file flows associated with a CI calculation are displayed in Figures 1 and 2 in the accompanying manual. In
grasp, the program
rcsfzerofirst is used to partition the CSF expansion in zero- and first-order sets. The ZF method is not the default mode but is only used to handle very large expansions and matrices, as discussed in detail in section 14 of the manual [
72] that is entirely devoted to strategies for ZF-MCDHF and ZF-CI.