ACS Publications. Most Trusted. Most Cited. Most Read
My Activity
CONTENT TYPES

Figure 1Loading Img
RETURN TO ISSUEPerspectiveNEXT

Measuring Electron Correlation: The Impact of Symmetry and Orbital Transformations

Cite this: J. Chem. Theory Comput. 2023, 19, 10, 2703–2720
Publication Date (Web):April 6, 2023
https://doi.org/10.1021/acs.jctc.3c00122

Copyright © 2023 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0.
  • Open Access

Article Views

6055

Altmetric

-

Citations

-
LEARN ABOUT THESE METRICS
PDF (2 MB)

Abstract

In this perspective, the various measures of electron correlation used in wave function theory, density functional theory and quantum information theory are briefly reviewed. We then focus on a more traditional metric based on dominant weights in the full configuration solution and discuss its behavior with respect to the choice of the N-electron and the one-electron basis. The impact of symmetry is discussed, and we emphasize that the distinction among determinants, configuration state functions and configurations as reference functions is useful because the latter incorporate spin-coupling into the reference and should thus reduce the complexity of the wave function expansion. The corresponding notions of single determinant, single spin-coupling and single configuration wave functions are discussed and the effect of orbital rotations on the multireference character is reviewed by analyzing a simple model system. In molecular systems, the extent of correlation effects should be limited by finite system size and in most cases the appropriate choices of one-electron and N-electron bases should be able to incorporate these into a low-complexity reference function, often a single configurational one.

This publication is licensed under

CC-BY 4.0.
  • cc licence
  • by licence

1. Electron Correlation

ARTICLE SECTIONS
Jump To

Electron correlation has often been called the “chemical glue” of nature due to its ubiquitous influence in molecules and solids. (1,2) The problem was studied in the smallest two-electron systems, such as He and H2 already in the 1920s, and these early efforts are discussed in detail elsewhere. (3−6) The correlation concept usually presupposes an independent particle model, typically Hartree–Fock (HF) mean-field theory, that serves as a reference compared to which the exact solution is correlated. The associated idea of correlation energy goes back to Wigner’s work in the 1930s on the uniform electronic gas and metals, (7,8) and it is nowadays commonly defined, following Löwdin, as the difference between the exact and the HF energy. (3) There is a vast literature on the subject and here we will confine ourselves to some of the most pertinent reviews. Early work on the correlation problem is reviewed by Löwdin in the 1950s (3) and more recently, (9) as well as by Sinanoğlu (10,11) within the context of his many-electron theory. The early perspective of McWeeny is a short but useful discussion of the problem in terms of density matrices. (12) Kutzelnigg, Del Re and Berthier penned an overview on the statistical measures of correlation, (13) and, as P. v. Herigonte, they also reviewed the situation in the seventies. (14) This was later followed up several times by Kutzelnigg. (15,16) Bartlett and Stanton not only discuss correlation effects in molecules, but also offer a tutorial on post-HF wave function methods. (17) Similarly educational is the chapter written by Knowles, Schütz and Werner (18) and the feature articles of Tew, Klopper, Helgaker (19) and Martin. (2) More recent work on the homogeneous electron gas is reviewed by Senatore and March (20) as well as Loos and Gill, (21) the former authors also discuss orbital-based and quantum Monte Carlo methods. For a review taking an entanglement-based approach to the correlation problem, we refer the reader to Chan. (22) There are also several book-length monographs dedicated to the subject by Fulde, (23) March (24) and Wilson, (25) as well as a number of collections. (26,27) In the remainder of this section, we will consider some selected topics about correlation effects before stating the scope of this perspective.
An important aspect of electron correlation is its strength and the various measures proposed to quantify it. For the homogeneous electron gas, where some analytical energy expressions can be obtained at least under some circumstances, the discussion is often concerned with the relative magnitudes of kinetic and potential energies in the limiting cases of high and low electron densities. While Wigner was able to obtain an expression for the correlation energy at the low-density limit, (7,8) in general, perturbation theory diverges already at second order, (28) a problem that can only be remedied partly at the high-density limit following Gell-Mann and Brueckner. (29) In the high-density limit, the kinetic energy dominates over the repulsive potential energy, the electrons are delocalized and behave like a gas, thus, an independent particle model should provide a good description. In the low-density limit, it is the Coulomb interactions responsible for correlation effects that dominate, forcing the electrons to localize on site (form a lattice). Interestingly, the low- and high-density limits have their analogues in molecular systems: (20) in the hydrogen molecule, the high-density case would correspond to short bond lengths and a delocalized molecular orbital (5,30−32) (MO) description, while the low-density limit would correspond to large internuclear separation where the electrons are localized around the nuclei, better described by a local valence bond (5,33−35) (VB) approach. While this is a useful conceptual link, it must be remembered that the homogeneous electron gas is a better description of metallic solids than it is of more inhomogeneous systems like molecules. In the next section, the contrast between solids and molecules with regard to strong correlation will be discussed before we turn our attention fully to molecular systems and examine the various measures used in wave function theory, density functional theory and quantum information theory.
From a wave function point of view, correlation strength is often associated with the nature of the reference function which is usually the HF state, as mentioned above. An interesting point that draws attention to the significance of the chosen reference is the observation that from a statistical point of view HF already contains some correlation introduced by antisymmetrization and the exclusion principle. (13) Following Kutzelnigg, Del Re and Berthier, (13) two variables may be called uncorrelated if
r 1 · r 2 i j = r 1 i r 2 j
(1)
where the quantities corresponding to the ith and jth particles in an N-electron system can be simply related to the usual expectation values as
r 1 = N r 1 i , r 1 · r 2 = N ( N 1 ) 2 r 1 · r 2 i j
(2)
Applying this to position vectors, the implication is that the positions of two electrons are uncorrelated. This is a weaker notion than that of independence, which is often taken to mean lack of correlation. Statistically, independence (or, correlation in the sense of dependence) could be defined (36) by requiring that the two-body cumulant λ2
λ 2 ( 1 , 1 ; 2 , 2 ) = γ 2 ( 1 , 1 ; 2 , 2 ) γ 1 ( 1 , 1 ) γ 1 ( 2 , 2 )
(3)
obtained from the one- and two-body reduced density matrices (1- and 2-RDM), γ1 and γ2, should vanish. The fulfilment of this condition would imply a relationship like eq 1 for any observable. This is the sense in which Wigner and Seitz (37,38) understood correlation, and this would correspond to a simple Hartree product of orbitals as a reference function, or, more generally, to a direct product of subsystem wave functions. Antisymmetrization then implies switching from this product reference function to an antisymmetric Slater determinant constructed from spin–orbitals. More generally, it is also possible to use generalized (antisymmetrized/wedge) product functions (39) to obtain the total wave function from antisymmetric group functions of subsystems. Using an antisymmetrized reference, it is customary to talk about Fermi and Coulomb correlation, the former being excluded from the conventional definition of the correlation energy. To make this explicit, the cumulant can be rewritten as
λ 2 ( 1 , 1 ; 2 , 2 ) = γ 2 ( 1 , 1 ; 2 , 2 ) γ 1 ( 1 , 1 ) γ 1 ( 2 , 2 ) + γ 1 ( 1 , 2 ) γ 1 ( 2 , 1 )
(4)
Here, the last term accounts for the exchange contribution, and thus, λ2 only contains the parts of γ2 that cannot be factorized in terms of γ1 in any way. In this sense, λ2 is the most general descriptor of correlation effects. It should also be noted that by using partial trace relations, (36) it can be shown that Tr(λ2) = Tr(γ12 – γ1), and hence correlation effects are also related to the idempotency of the 1-RDM. This notion goes beyond that of Wigner and Seitz and is closer to Löwdin’s idea of correlation. The significance of Fermi correlation in particular was discussed by several authors in the past, (12,13,19) and more recently in a pedagogical manner by Malrieu, Angeli and co-workers. (40) Nevertheless, as pointed out by Kutzelnigg and Mukherjee, (36) eq 4 is not identical to Löwdin’s definition either since it contains no reference to the Hartree–Fock state, apart from that fact that the closed-shell HF determinant, for example, would be enough to make λ2 vanish and would thus be uncorrelated in this sense. The definition in eq 4 is an intrinsic one in the sense that γ1 and γ2 can be calculated for any (reference) state. More generally, any second-quantized operator string can also be evaluated with respect to a general reference (vacuum) state using generalized normal order and the associated general form of Wick’s theorem that itself relies on n-body cumulants. (41,42)
The foregoing discussion indicates the role of the reference state in the definition of correlation effects and highlights the fact that reference states should conveniently incorporate certain correlation effects to make subsequent treatment easier. As antisymmetrization dictates using Slater determinants rather than simple orbital products, spin symmetry may be enforced by demanding that the reference be constructed as a fixed linear combination of determinants to account for spin symmetry, (19) and there might be other criteria. (22,43,44) In such cases, one may distinguish between correlation recovered by the reference and residual effects. From such an operational point of view, the question is whether the strongest correlation effects can already be captured at the reference level to ensure a qualitatively correct starting point. In practice, one may choose to build reference states and represent the wave function using the following N-electron basis states:
  • Determinant (DET): An antisymmetrized product of orbitals. They are eigenfunctions of Ŝz but not necessarily of Ŝ2.

  • Configuration State Function (CSF): A function that is an eigenstate of both Ŝz and Ŝ2. CSFs can be obtained from linear combinations of determinants, but there are at most as many CSFs as DETs for a given number of unpaired electrons and total spin.

  • Configuration (CFG): At an abstract level, a configuration is just a string of spatial orbital occupation numbers (0,1,2) for each orbital in a given N-electron wave function. By extension, a configuration is also the set of determinants or CSFs that share the same spatial occupation numbers, but differ in the spin–orbital occupation numbers.

We will discuss the role of this choice and offer a classification of wave function expansions and potential reference states based on the properties of the exact solution, which leads us to define what we call the multireference character of the wave function. We will also examine the role of the orbital basis and ask the question how orbital rotations influence this apparent multireference character in a simple model system. On the other hand, we are not concerned here with the practical construction of good reference states, even though the topics discussed undoubtedly have consequences in that regard. At this stage, we simply ask what the difference is between a canonical and a local orbital representation in terms of the apparent multireference character and what the chances are of incorporating the strongest correlation effects into the reference function, be that via enforcing spin symmetry or generating a new many-electron basis by rotating orbitals.

2. Solids versus Molecules

ARTICLE SECTIONS
Jump To

Strong correlation is one of those important but elusive concepts that is often used in discussions of the properties of molecular and solid-state systems. A brief survey of the literature that follows below will reveal that it is often equated with various other notions, such as the multireference character of the wave function, the failure of density functional theory (DFT) and quantum entanglement. It is not immediately clear how these various notions are related to one another or to strong correlation, or whether any one should be preferred over the other. On the other hand, it seems relatively clear that a large part of the discussion on strong correlation focuses on extended solid-state systems, where density functional and dynamical mean-field (45,46) approaches are commonly used, (47) although many-body methods (48) and quantum Monte Carlo techniques (49,50) are also available as well as a number of exactly solvable model problems. (51) The theoretical tools used to describe periodic solids of infinite extension and finite-size molecular systems can be quite different in ways that reflect the underlying physics. Complexity and locality may have different manifestations for these systems in terms of the number of parameters or the type of basis sets required to describe them. In terms of symmetry, the obvious difference is the presence of translational symmetry in solids. Furthermore, symmetry for an infinite number of degrees of freedom also makes spontaneous symmetry breaking possible in ground-state solutions. (52) For all these reasons, it seems reasonable to separate the discussion of electron correlation in solids from that in molecules.
As mentioned before, in the homogeneous electron gas model strong correlation effects are related to the low density limit. Wigner argued that at this limit the electrons are localized at lattice points that minimize the potential energy (7) and vibrate around the lattice points (8) with just the zero point energy demanded by the uncertainty principle. Leaving the vibrational part aside, the total energy of the system at very low densities has the form (8,24,53,54)
E = T ^ + V ^ , T ^ = α r 2 , V ^ = β r
(5)
where α and β are constants and r is the Fermi radius, which in this context is the radius of a sphere available for a single electron. If the density is low, r is large, and that means that the mean kinetic energy ⟨⟩ is negligible compared to the mean potential energy ⟨⟩, which is sometimes taken to be a criterion for strong correlation. Wigner was able to obtain a correlation energy formula by assuming a vanishingly small kinetic energy at the low-density limit (7) and interpolating between the low- and high-density limits, and considered the zero point kinetic energy, which would add a term proportional to r 3 / 2 in eq 5, only in a later study. (8,21,55) While the result works reasonably at density values actually present in some solids, some criticisms merit attention especially when application to molecules are concerned. While admitting the good agreement with measurements in metallic systems and similar results derived from plasma theory by Bohm and Pines, (56,57) Slater criticized (58) the assumption of uniform positive background, which should be an even more serious problem when considering more inhomogeneous systems, such as the dissociating H2 molecule. (3) We note in passing that it is possible to account for inhomogeneities in the electron gas, (59) but perhaps the conceptual simplicity of the homogeneous electron gas model is more relevant for the current discussion. Concerns were also raised by Löwdin (3) about the model not obeying the virial theorem (60) in the sense expected in molecular systems. It has since been shown that the uniform electron gas obeys the more general form of the virial theorem (61,62)
2 T ^ + V ^ = r d E d r
(6)
which means that in the low-density limit, the total energy E is not stationary (although the potential energy may be in the sense needed to find an optimal lattice). In molecular systems, on the other hand, stationary points of the total electronic energy are of special interest as in the case of H2 at equilibrium and infinite separations. In these cases, the virial theorem takes the form 2⟨⟩ + ⟨⟩ = 0, which does not allow the kinetic energy to be vanishingly small. (3) In this regard, see also the work of Ruedenberg (63) for an analysis of the formation of the chemical bond that does respect the virial theorem. For the electron gas, the main use of the virial theorem is actually the separation of kinetic and potential energy contributions in the correlation energy. (61) Thus, it may be concluded that the criterion of the relative magnitude of the mean kinetic and potential energy is appropriate within the homogeneous electron gas model, which is a better approximation of metallic solids than molecules. Nevertheless, extensions of the same basic idea are possible. Within the Hubbard model, (64) for example, correlation effects are commonly considered as being strong if the on-site electronic repulsion is larger than the bandwidth (resonance energy) or the hopping term, which favors antiferromagnetic coupling between the various sites. (23) Repulsion is a potential energy term, while the hopping integral has to do with the kinetic energy of electrons. We will discuss the appropriate extension to the molecular Hamiltonian in later sections.
As in this perspective we are more concerned with molecular systems, it is especially interesting that solids containing d or f electrons are considered strongly correlated, (23) as this would prompt one to look for strong correlation in molecular systems containing transition metals or lanthanides. Before leaving the discussion of solid state behind, it should also be recalled that sometimes the nature of the problem at hand calls for the combination of solid-state and molecular approaches, as in studying strongly correlated (bond-breaking) problems in surface chemistry via embedding approaches, which pose challenges of their own. (65) The foregoing discussion also underlines the problem of finding quantitative measures of strong correlation, or, perhaps more precisely, of the various other properties associated with it. After reviewing the various attempts made along these lines, we will expound our own perspective on strongly correlated systems in chemistry as viewed through the lens of the multireference character of the wave function expansion. We will use a simple metric that can be obtained from the exact solution and examine its properties using the hydrogen molecule as a model system. As mentioned before, this problem has been studied since the early days of quantum chemistry (3−6) and also has some practical relevance in that it can be taken as a model for the antiferromagnetic coupling in transition metal compounds such as a typical Cu(II) dimer. We note in passing that the hydrogen chain is also a common model problem in the study of extended systems. (66,67)

3. Wave Function Theory

ARTICLE SECTIONS
Jump To

Within the framework of wave function-based post-Hartree–Fock approaches, it is customary to distinguish between single and multireference methods to which the concepts of dynamic and static or nondynamic correlation also correspond. (11,17,18,68) Dynamic correlation is also called short-range correlation as it has to do with the electrostatic repulsion of electrons and the difficulty of describing the resulting Coulomb cusp (see this study (69) for a different perspective). When it comes to other correlation effects, some authors use static and nondynamic correlation interchangeably and sometimes refer to it as long-range correlation. (18,70) Others (17) prefer to reserve static correlation for effects induced by enforcing spin symmetry in order to construct a correct zeroth order description, and call those associated with other sources, such as bond breaking, nondynamical. Tew, Klopper and Helgaker follow a similar distinction except that they include spin symmetry in Fermi correlation and identify static correlation as the correlation associated with near degeneracy and thus needed for a correct zeroth order description. (19) Yet others (71) distinguish within static/nondynamical effects between those that can be captured by spin-unrestricted HF and those that cannot. A summary of more formal attempts to define static and dynamic correlations is also available. (72) The important conclusion from this discussion seems to be a point already raised by Wigner and Seitz (38) that the main difference is between correlation effects that arise from symmetry and those arising from interelectronic repulsion, regardless of how we choose to classify these further. Thus, within symmetry effects one might further distinguish between antisymmetry and spin symmetry, within repulsion effects between those associated with spatial (near) degeneracy and the rest.
Ignoring the question of spin symmetry for the moment, nondynamic/static correlation may be identified with strong correlation in the present context, and thus, it is also associated with multireference methods, where “reference” is a generic term for some function that serves as an initial state on which a more sophisticated Ansatz can be built. This also suggests that there are two ways of dealing with static correlation, either improving the reference or improving the wave function Ansatz built on it. In principle, single reference methods will yield the exact solution if all possible excitation classes are included in the Ansatz, and it was furthermore demonstrated that they can also be adapted to strong correlation by removing some of the terms from the working equations. (73) More pragmatically, one might ask at what excitation levels an Ansatz, such as coupled cluster (CC), needs to be truncated to yield acceptable results for some properties, (74,75) but this is costly on the one hand and does not offer a simple picture of strong correlation either. One way to improve the reference is to find the exact solution within at least a complete active space (CAS). (76) Unfortunately, this requires the construction of many-electron basis states the number of which scales exponentially with the size of the active space. However, as observed by Chan and Sharma, (77) this apparent exponential complexity does not always reflect the true complexity of the system since the principle of locality often forces a special structure on the wave function expansion. This entails that the amount of information needed to construct the wave function should be proportional to the system size, which agrees well with the observation that static correlation in molecules can typically be tackled with only a handful of terms in the wave function expansion. (18) As an illustration, the hydrogen molecule around the equilibrium bond length is well-described by a single-determinant built from canonical orbitals and associated with the configuration σg2, while at larger separations another contribution from σu2 becomes equally important. (18,77) In the latter limit, the wave function takes the following bideterminant form
Ψ 0 = 1 2 ( | i i ¯ | | a a ¯ | ) = 1 2 ( | μ ν ¯ | | μ ¯ ν | )
(7)
where i and a are the occupied and unoccupied molecular orbitals in the minimal basis HF solution, and the shorthand notation |ii| is used to denote determinants, with the overbar signifying a β spin. While these canonical orbitals are delocalized, μ and ν denote atomic orbitals, each localized on one of the hydrogens and orthonormal at large distances. As is well-known, (4,6,78) the molecular orbital (MO) and valence bond (VB) approaches offer equivalent descriptions of this wave function, but the locality of the VB picture leads to an interpretation in terms of “left–right” correlation, i.e., the phenomenon that at large internuclear separation an electron preferably stays close to the left nucleus if the other one is close to the right nucleus, and vice versa. This form of static correlation can be accounted for in a spin-unrestricted single-determinant framework leading to different orbitals for different spins beyond a certain bond distance, but, for molecular systems, approximations preserving the symmetries of the exact wave function are usually preferable. It is all the more interesting that the constrained-pairing mean-field theory of Scuseria and co-workers is designed to capture strong correlation (43) by breaking and restoring various symmetries, (79) while Kutzelnigg took a different route toward obtaining a similarly economic reference function by separating bond-breaking correlation effects via a generalized valence bond (GVB)/antisymmetrized product of strongly orthonormal geminals (APSG) approach. (44)
It is also apparent from these considerations that most traditional methods follow the “bottom-up” approach in which a systematic Ansatz is built on a simple reference function and then truncated using some simple criterion (e.g., the excitation level). The typical arsenal of this approach includes many body perturbation theory, CI and CC approaches in both the single and multireference cases. (17,18,80) The most popular of them is CC theory, (81) which was introduced into chemistry by Čižek, (82) Paldus and Shavitt, (83) was reformulated in terms of a variational principle by Arponen (84) and whose mathematical properties are still an active area of research. (85,86) Even more actively studied are its local variants (87,88) and its explicitly correlated, (89) excited state (90) and multireference (91) extensions. Beyond these methods, there is also room for a “top-down” approach in which one aims directly at the exact solution (full CI, full CC) and applies some strategy to neglect contributions that are less important while trying to save as much of the accuracy as possible. The relative sparsity illustrated above with the H2 example can be exploited in selected CI methods of the CIPSI (configuration interaction with perturbative selection through iterations) type (92) for much larger systems. The CIPSI method is the first of many selected CI (SCI) approaches that attempt to solve the problem of selecting relevant contributions in a deterministic fashion, but it is also possible to apply stochastic sampling processes, including QMC approaches (93) such as the FCIQMC method. (94,95) FCIQMC can be performed without approximation, in which case the method is exact but suffers from a sign problem in general. (96) Instead, the initiator approximation is typically applied, which involves truncating contributions between certain determinants with low weight. (97,98) These and other methods have been reviewed in more detail recently by Eriksen, (99) including Eriksen and Gauss’ own approach based on many body expansions. (100,101) Another approximate FCI method particularly suitable for strongly correlated systems, the density-matrix renormalization group (77,102) (DMRG) approach will be discussed later among entanglement-based approaches. We also note that some of these methods have been analyzed in terms of scaling and the gains due to the reduction in the number of important contributions in the wave function expansion (103−106) (see also some remarks on orbital rotations below).
It has been observed that multireference character, interpreted here as the fact that the wave function expansion contains more than one determinant of significant weight for some system, does not necessarily imply strong correlation, and the additional criteria of the breakdown of single reference perturbation theory and the appearance of degeneracies were suggested. (107) More explicitly, strong correlation could be said to occur whenever the perturbation (the difference between the full Hamiltonian and the Fockian) is larger than the gaps between the Hartree–Fock orbital energies (spectrum of the Fockian). (22) This could be regarded as an extension of the similar definition within the Hubbard model. For nonmetallic states, strong correlation would also be implied if the correlation energy is larger than the excitation energy to the first excited state of the same symmetry as the ground state. (108) Nevertheless, the multireference character can be used to characterize the exact wave function, even if its usefulness for approximate approaches may be questioned. This leads us toward a definition of multireference character via the full configuration interaction expansion
Ψ = I C I Φ I
(8)
where ΦI are some basis states, typically determinants. If any one coefficient CI dominates (see more later) the expansion, then the associated ΦI could be taken as the reference in a single reference (determinant) treatment. Thus, multireference character is associated with wave function coefficients CI which for orthonormal basis states are equivalent to the overlap between the exact state Ψ and the basis state ΦI. Unfortunately, this criterion is impractical to decide in advance whether a multireference treatment of a system is necessary, and other diagnostics, (109) such as the one relying on the norm of the singles vector (T1) in a CC calculation (110) truncated at the level of single and double excitations, were also found insufficient (109) or even misleading. (111) Metrics relying on natural orbital populations are more promising as they give information about the number of electrons uncoupled by correlation driven processes. (75,112,113) The deviation of the one-body reduced density matrix from idempotency has also been used as a measure (114,115) and well as measures associated directly with the two-body cumulant. (44,115,116) More pragmatically, the static correlation energy has also been defined as the exact correlation energy obtained from a minimal basis calculation involving valence electrons, the idea being that in such a calculation the most important configurations would be represented, although this approach also needs a hierarchy of approximations to be useful. (117) A few other schemes have also been suggested to decompose the correlation energy into dynamical and nondynamical parts using simple formulas, (70,115) and more recently, using cumulants. (115)
For the purposes of this perspective, the FCI-based criterion in eq 8 will be sufficient as it enables us to base our discussion on the properties of the exact wave function. While it is common to use determinants as basis states, this is not the only possibility. Using CSFs or CFGs for example would yield basis states that can be written as sums of determinants or constructed in some other way. (103,118) We will investigate their effects on the multireference character in a later section. Already in eq 7, while both the MO and VB descriptions are multideterminant expressions, multiple configurations only enter into the (canonical) MO picture. We will also examine the role of orbital transformations in defining the multireference character of the wave function. Again, in eq 7, the concept of left–right correlation seems to be connected more to the local VB picture than to the MO one. Our primary goal here is to comment on multireference character, rather than other notions about strong correlation, but before doing that, we briefly examine other approaches to the problem.

4. Density Functional Theory

ARTICLE SECTIONS
Jump To

In their recent paper, Perdew et al. remark that strong correlation is sometimes taken to mean “everything that density functional theory gets wrong”. (119) Most practical DFT approaches rely on the Kohn–Sham procedure which in most cases corresponds to a single-determinantal implementation. Unlike the Hartree–Fock solution, however, the Kohn–Sham determinant is not an approximation to the exact wave function; rather, it is a tool to obtain the ground-state density or spin densities. There seem to be two opinions in the literature on what this signifies. Perdew et al. argue that as a consequence, the Kohn–Sham determinant need not exhibit the symmetries of the Hamiltonian as those will be reflected in the density in a more indirect fashion. This also implies that despite the appearance of a mean-field approach, Kohn–Sham DFT is in principle able to account for correlation effects. (120) Others argue (121−123) that spin-adapted reference functions should be used at least in some open-shell cases which may be multideterminantal, as it is done in the spin-adapted Hartree–Fock case in wave function theory. In any case, a single determinant may not always suffice and the question remains to what extent DFT can treat strongly correlated systems. In trying to answer this question, Perdew et al. (120) distinguish between strong correlation arising in condensed matter models and static correlation originating in near or exact degeneracies, such as the closing HOMO–LUMO gap in the H2 bond dissociation process. On problems of the latter type, they note the typical failure of common semilocal approximations.
While multideterminental techniques have been introduced in DFT, (121−123) it is more common to rely on broken symmetry DFT (78,124−126) in which multiplet energies are obtained from a combination of single-determinant energies. This scheme may or may not hold exactly, depending on the specific case. (121,126) Perhaps the simplest example of spin-symmetry breaking is the fact that the spin-unrestricted approach yields only one of the determinants |μν̅| or |μ̅ν| in eq 7 which can be interpreted as a resonance structure in the valence bond sense. Although such determinants are not pure spin states, they are degenerate. From such broken symmetry solutions and monodeterminantal pure spin states, it is often possible to calculate the energies of multideterminantal spin states: in H2, the appropriate combination of the energy of the broken symmetry solution and the triplet energy yields the energy of the excited-state singlet. (78) The use of this simple two-level model in the description of magnetic coupling is analyzed in detail elsewhere. (127) The more general broken symmetry (BS) DFT procedure starts with a mixture of canonical orbitals obtained from a closed-shell calculation and optimizes the mixing angles during the iterative solution to obtain a mixture of multiple configurations. (123) A similar idea underlies the permuted orbital (PO) approach where the broken symmetry states are generated by permuting orbitals, typically the HOMO and the LUMO, (123) and can be used to construct multideterminant spin eigenfunctions. Based on the capabilities of the BS and PO approaches, Cremer proposes a DFT-oriented multireference typology: (123) Type 0 wave functions need a single configuration or determinant and the spin-unrestricted approach can handle them; Type I consists of a single configuration with multiple determinants for which the PO approach is assigned; Type II is the case of multiple configurations each with a single determinant to be treated by BS-DFT; Type III systems with multiple configurations each with multiple determinants are the most difficult and can be handled by explicitly multireference methods, although possibly the PO approach might work. While we find the distinction between determinant and configuration/CSF references commendable, in the cases studied by Cremer there was no need to distinguish between configurations and CSFs, a feature that we will come back to in our own classification. As for interpreting broken symmetry solutions, it should be recalled that while the exact functional for finite systems is not symmetry broken, approximations are. Strong correlation in the exact wave function describes “fluctuations” in which the spins on the different nuclei are interchanged (|μν̅| vs |μ̅ν|). (128) Symmetry breaking in approximate methods is in fact taken to reveal strong correlations (119) with a mechanism associated with it: at large distances, when H2 becomes a pair of H atoms, an infinitesimal perturbation can reduce the wave function and the spin densities to a symmetry broken form, in which an electron with a given spin is localized around one H atom. (119,128) All this is in line with Anderson’s famed essay (129) on emergent phenomena on different scales involving increasingly larger numbers of particles, though it also implies that symmetry breaking may be more relevant for solids than for molecules. (52,119)
It may be worth pointing out that the physical contents of broken symmetry wave functions can be conveniently “read” by means of the corresponding orbital transformation (COT). (130) This transformation orders the spin-up and spin-down orbitals into pairs of maximum similarity. This results in a wave function that can be read analogously to a generalized valence bond (GVB) wave function (131,132) consisting of doubly occupied orbitals, singlet coupled electron pairs in two different orbitals and unpaired (spin-up) electrons. Obviously BS-DFT can properly represent the singlet coupled pairs, but the analysis shows that the spatially nonorthogonal pseudosinglet pairs obtained from the COT serve a similar purpose, namely to variationally adjust the ionic- and neutral components of the wave function. (127) This way of reading BS wave functions has found widespread application in the field of metal-radical coupling (e.g., examples (133−135)). A more thorough review is available. (136)
The theorems due to Hohenberg and Kohn (137) form the basis of the variational formulation of DFT, although questions about their precise meaning (138) remain pertinent enough to motivate further research, (139,140) not to mention the fact that these theorems do not generalize easily to excited states. (141) The favorable linear scaling property of DFT approaches is often derived from Kohn’s principle of nearsightedness, (142,143) i.e., that under certain conditions the density at some point is mostly determined by the external potential in a local region around that point. Practical approaches to density functional theory often consist of proposing new functional forms (144) in lieu of the unknown exact density functional of the energy. In contrast, the energy is a known functional of the 2-RDM, (114,145,146) although enforcing constraints (147) to actually optimize this functional is not trivial. Fortunately, the energy can also be written as a functional of the 1-RDM, (148) and different variants of density matrix functional theory are usually parametrized in terms of the 1-RDM and the 2-body cumulant. (149−152) Like DFT, this approach also relies on approximate functionals, but it manages to overcome some of the failings of the former. The appropriate constrains for the 1-RDM itself (153) are relatively easy to enforce, although the reconstructed 2-RDM should then still satisfy its own N-representability conditions. (151) One might then construct functionals (154) inspired by wave function approaches, such as geminal theories, or use auxiliary quantities to parametrize the cumulant. (155) In density cumulant functional theory (150,156) (DCFT), the functional is given in terms of the best idempotent estimate to the 1-RDM and the 2-body cumulant, and the critical N-representability conditions affect only the relatively small part due to the latter. Connections with traditional DFT can also be established, e.g., in the DCFT case, an analogue of the DFT exchange-correlation functional can be obtained by neglecting the Coulomb terms in which the nonidempotent part of the 1-RDM occur. (150) The nearsightedness principle applies to the 1-RDM as well, (142) and density matrix functionals have the advantage over DFT that the kinetic, Coulomb and exchange energies are all known functionals of the 1-RDM, only the correlation energy functional remains unknown. (157) Excited state applications of RDM functionals have also been considered. (152,158) A collection of studies on the various aspects of RDM-based methods is available, (159) while the relationship between wave function based methods, DFT and RDM functionals is discussed in Kutzelnigg’s review. (157) In the current context, it is relevant to mention that some density matrix functionals have been applied with success to strongly correlated problems. (152,155,160−164)
So far we have only considered functional approaches in which the variable is the density or the RDMs, but it is also possible to set up a variational formulation (165−167) for functionals of Green’s functions, which themselves can be regarded as generalizations of RDMs with their own cumulant expansion. (168) All three of these quantities form the basis for embedding theories enabling the treatment of large systems. (169) Dynamical mean-field theory (45,46,170) is a much applied method in solid state physics that can be derived from such a variational approach and has also been investigated from the perspective of molecular quantum chemistry. (171,172) Another method that falls into this category is the popular GW approximation of Hedin. (173−175) The properties of these and other similar methods have also been analyzed from the perspective of strong correlation, (170,172,176,177) excited states (174,178) and symmetry breaking. (177,179) Although both have some success in multireference systems, GW approaches are typically applied to weakly correlated systems while dynamical mean-field approaches have succeeded where DFT or other Green’s function approaches failed. (170,172,176)
It emerges from this discussion that the DFT concept of strong correlation is related on the one hand to its monodeterminant nature and to symmetry breaking on the other. Measures have been proposed along these lines. Thus, Yang and co-workers relate static correlation to fractional spins, and present an extension to DFT covering fractional spins including a measure based on this extension. (180,181) Grimme and Hansen offer a measure as well as a visualization tool for static correlation effects based on fractional occupation numbers obtained from finite temperature DFT calculations. (182) Head-Gordon and co-workers use spin (and sometimes spatial) symmetry breaking as a criterion and suggest a measure based on spin contamination, (107) which may be related to natural orbital occupation numbers, a criterion also used by others. (123) Still other authors use checks based on the idempotency of the one-body density, which is essentially used here as a check on the monodeterminant nature of the wave function under consideration. (114,183)

5. Quantum Information Theory

ARTICLE SECTIONS
Jump To

Quantum information theory offers a different perspective on the same phenomenon, albeit one that can be related to more traditional quantum chemical concepts. (108) The starting point is a two-level quantum system, a qubit, that serves as an analogue of classical bits. We take spin-up and spin-down states to form the qubit in the following. A spin-up state can be labeled |0⟩ and a spin-down state as |1⟩, and a qubit can be any linear combination of these two basis states, (184) |m⟩ = α|0⟩ + β|1⟩. If the spin of an electron prepared in such a state is measured in a field oriented along an axis parallel to the spin-up and spin-down basis states, then the probability of finding the electrons in state |0⟩ or |1⟩ is |α|2 and |β|2, respectively. Along the same lines, one may prepare two electrons. Assuming at first that the two electrons do not interact before the measurement, the possible two-electron states may be described in the basis of the product states |0⟩ ⊗ |0⟩, |0⟩ ⊗ |1⟩, |1⟩ ⊗ |0⟩ and |1⟩ ⊗ |1⟩. They remain distinguishable and they are also independent in the sense that measuring the first qubit yields results that only depend on how the first electron was prepared and is not influenced by the second electron. Imagine now bringing spin-1/2 particles close enough together so that their spins align in an antiparallel fashion corresponding to the ground state and then separating them again far enough so that one is located at site A the other at site B. The wave function that describes this state is
| Ψ 0 = 1 2 ( | 0 A | 1 B | 1 A | 0 B )
(9)
In classical computing, the value of the first bit in a two-bit string does not yield information about the value of the second bit, in the same way as the independently prepared quantum systems do not. The two-qubit quantum state in eq 9 is different because measuring the spin only at site A also reveals what the result of the measurement will be at site B. Thus, the two qubits at site A and B are entangled (185) in the sense that the result of a measurement at A determines that at B. It is also often said that they are correlated, although entanglement is only a special kind of quantum correlation. (186) This is different from the definition of strong correlation in terms of perturbation strength, but has the advantage that it is more easily quantifiable. (22) Note that eq 9 describes a system of distinguishable spin-1/2 particles, such as an electron and a positron, or two electrons at a large distance. Although electrons are identical particles, in cases like the dissociating H2 molecule, each electron can be identified as the one localized around a specific nucleus if the two nuclei are separated enough so that the electronic wavefunctions no longer overlap. (187) Comparing eq 9 to the VB wave function of H2 at large bond distances in eq 7 and assuming that μ is at site A and ν at site B reveals that
| μ ν ¯ | = 1 2 ( | 0 A | 1 B | 1 B | 0 A ) , | μ ¯ ν | = 1 2 ( | 1 A | 0 B | 0 B | 1 A )
(10)
since the tensor product here is ordered w.r.t. the particle labels, e.g., μ(2)α(2)ν(1)β(1) → |1⟩B ⊗ |0⟩A. Thus, eq 7 is the antisymmetrized version of eq 9. In a sense, this indicates another kind of symmetry breaking. In spin-unrestricted approaches discussed in the DFT context, only one of the determinants |μν̅| or |μ̅ν| in eq 7 are present at large bond distances, i.e., it is known which spin is at site A (or B) but not which particle carries that spin. On the other hand, eq 9 breaks antisymmetry as a natural consequence of the assumption of distinguishability: in this case it is known which particle is at site A (or B), but not the spin of that particle, which is the question a subsequent measurement can help resolve. As this discussion suggests, the established notion of entanglement (184) assumes distinguishable subsystems, although extensions to indistinguishable particle systems exist. (188) This should perhaps be kept in mind when comparing with quantum chemical notions, where indistinguishability of electrons in molecules is assumed (cf. also the discussion on product functions and generalized product functions (39) earlier). Still, the resonance structures of VB/GVB theories, for example, assume a degree of distinguishability and entanglement may be a useful tool in dealing with them. (22) It is also worth mentioning that when measuring the spin of a single electron as discussed above, it is always possible to rotate the measurement field so that, say, the spin-up state is measured with probability 1. This is not possible for the singlet described in eq 9, since the expectation value of one of the spins along any axis is zero. Due to this rotational invariance, the invariance of entanglement measures with respect to the choice of basis is often emphasized, (185) but this should not be confused with orbital rotations to be discussed below.
It may be said that the VB wave function is created by identifying the entangled local states (orbitals) associated with chemical bonds. In fact, just as traditional wave function approaches start from the independent particle model and build more accurate Ansätze by incorporating the most important excitations, one might start a similar hierarchy around independent subsystems as a starting point which explores the Hilbert space by encoding more entanglement. (22) This leads to matrix product states (MPS) in general and the density matrix renormalization group (DMRG) approach in particular. (22) The seniority-based CC theory of Scuseria and co-workers (189) can also be interpreted on a quantum information theory basis. (190,191) This approach is built on orbital pairs rather than single orbitals with seniority being the number of unpaired electrons in a determinant that can be used to select the determinants for inclusion in a wave function expansion. The claim is that strong correlation can be treated by low-seniority contributions in a suitable orbital basis, (189) and it is possible to set up an entanglement-based measure to be used as a cost function. (191)
One measure of entanglement is based on the von Neumann entropy (the “quantum Shannon entropy”) of a state and the quantum relative entropy (184) which provides a measure of the distinguishability of two states. (186) The relative entropy of entanglement can be given as (22)
S = i σ i 2 log 2 σ i 2
(11)
in terms of the singular values σi corresponding to a Schmidt decomposition: (184) if A and B are distinguishable subsystems of the whole system, then a pure state |Ψ⟩ can be expanded in the basis of functions of the type |ΨA⟩ ⊗ |ΨB
| Ψ = A B C A B | Ψ A | Ψ B
(12)
where |ΨA⟩, |ΨB⟩ are basis states in A and B, and the singular values σi are those of the expansion coefficient matrix CAB. If only one singular value is nonzero, the entropy is zero and |Ψ⟩ is separable; otherwise, it is entangled, and it is maximally entangled if all singular values are the same. (22,185) Observe that entanglement has to do with the characterization of subsystems and depends on the partitioning of the system. For example, in eq 9, it is the two subsystems at sites A and B that are entangled because the total state cannot be written in the simple form |mA ⊗ |nB. For further considerations on superselection rules and classical and quantum correlation types, see. (186,191) If the local states are chosen to be single orbitals, then we may speak about (single) orbital entropy (190,192) and use the eigenvalues of the orbital reduced density matrix instead of σi2 in the definition of the entropy above, with the summation running over unoccupied, singly occupied spin-up and spin-down, and doubly occupied states. Orbital entropy has been used to measure multiconfigurational character (193) and as a basis for automatic active space selection. (194) It has also been applied to the H2 model system (193) and to a number of other bond formation processes. (193,195) See also another entanglement-based study on 2-electron systems by Huang and Kais. (196) The dependence of entanglement-based measures on the orbital basis has been noted in the literature (186,193) and they have also been used to set up a qualitative distinction between nondynamic, static, dynamic and dispersion correlation effects (190,197) in the sense of Bartlett and Stanton. (17) In earlier approaches, natural orbital occupation numbers have also been substituted for σi2 to calculate the von Neumann entropy as a measure of correlation strength. (198) Recently, entropy-based measures have even served as a basis of a geometric interpretation and an upper bound to the correlation energy. (199) Finally, we note that the two-orbital entropy can be defined in an analogous manner, and serves as the basis for orbital interaction (mutual information) measures (190,192,200) that can be used to find an orbital ordering that ensures good convergence in DMRG by placing strongly interacting ones near each other. (200)
So far we have discussed the application of quantum information theoretical concepts in quantum chemistry. However, these concepts also underlie the theory of quantum computing in which chemistry has also been identified as one of the potential areas of breakthrough, (201) as exemplified by resource estimates in the study of Reiher et al. on FeMoco. (202) The latter compound served as an example of a strongly correlated system that would especially benefit from the fact that quantum computers are quantum systems themselves that need not explicitly generate the FCI coefficients in eq 8. Thus, we will also briefly mention recent developments in this field. In particular, the need that the initial states should have a good overlap with the target state for certain quantum algorithms to succeed is well-known. (203) The fact that the preparation of spin eigenfunctions on quantum computers has been investigated recently (204,205) can be seen as a step in this direction and it also underpins our efforts in the upcoming sections to provide a more nuanced classification of reference functions than the usual single vs multideterminant one. As far as the potential gains in scaling are concerned, a recent study suggests that the exponential nature of quantum advantage in quantum chemistry (206) cannot be taken for granted in most cases, although the authors are careful to point out that quantum computers may still be very useful in quantum chemistry. Thus, the search is on for specific chemical systems where quantum computers could be most beneficially applied. (207−209)

6. Symmetry Considerations

ARTICLE SECTIONS
Jump To

We have already distinguished between DETs, CSFs and CFGs as various choices of basis states. It also follows from the previous discussion that symmetry plays a significant role in the classification of correlation strength. To get a better grip on these notions, let us remind ourselves about how an N-electron wave function is built from one electron functions (orbitals) and how various symmetries may be taken into account in the construction. The simplest approach is to approximate the N-electron wave function as a Hartree product (HPD) of N orbitals and enforce various symmetries. In general, the operator encodes a symmetry of the Hamiltonian Ĥ if [Ĥ, ] = 0, and thus the exact wave function is a simultaneous eigenstate of Ĥ and . If an approximate wave function does not have a particular symmetry, an appropriate projection operator ÔQ can be constructed that would produce an eigenfunction of with eigenvalue Q. We are not concerned here with all possible symmetries (e.g., the particle number operator, time reversal), but may limit ourselves to the following scheme:
HPD A ^ DET O ^ S CSF ( CFG ) O ^ R SA‐CSF
(13)
where antisymmetrization ( A ^ ) produces a Slater determinant (DET), spin projection (ÔS) a configuration state function (CSF), or, more generally, a number of CSFs belonging to the same configuration (CFG, i.e., a sequence of spatial occupation numbers), while spatial symmetry can also be enforced via ÔR of some symmetry operation R, producing a spatial symmetry-adapted CSF (SA-CSF). We will briefly examine these in the following. It only remains to note here that symmetry adaptation in general leads to higher energies in a variational procedure since it introduces constraints in the variational manifold, a phenomenon sometimes referred to as Löwdin’s dilemma. (210)
The original purpose of introducing the Slater determinant (211) was to find an N-electron basis that automatically satisfies the exclusion principle, or, rather, its generalization requiring a Fermionic wave function to be antisymmetric. For this purpose, the HPD is constructed from the product of spin orbitals ϕi(xi), which itself is the product of a spatial orbital φi(ri) and a spin function σi(si), and the antisymmetrization A ^ is carried out by permuting the coordinates xi = (ri, si) and summing over the permutations multiplied by their parity
Φ = A ^ i N ϕ i ( x i ) = A ^ i N φ i ( r i ) i N σ i ( s i )
(14)
Since A ^ is a projector multiplied by a constant, i.e., A ^ 2 = N ! A ^ , Φ is an eigenfunction of A ^ . The correlation introduced by this formulation is usually illustrated (19,40) using the case of two parallel electrons, where the spatial part φ1(r12(r2) – φ2(r11(r2) is obviously zero at r1 = r2. It has been observed that this Fermi hole is in fact nonlocal, because the spatial part also vanishes if φ1(r1) = φ1(r2) = ± φ2(r1) = ± φ2(r2). (40) Moreover, antisymmetrization also introduces anticorrelation in the opposite-spin case, which has been used to explain the fact that same-spin double excitations contribute less to the energy in post-Hartree–Fock calculations than opposite-spin ones. (40)
Given that the Slater determinant is a much simpler entity than a general group theoretical construction of the wave function, it could be claimed with some plausibility that “Slater had slain the “Gruppenpest”,” (212) i.e., the pestilence of groups that threatened the conceptual clarity of quantum mechanics in his view. Were it only about the antisymmetry requirement, Slater’s approach would be preferable to general group theoretical presentations given that it is much more transparent, a fact also responsible to a large extent for the establishment of the Slater determinant as the “default choice” as reference state. However, the question of spin adaptation still remains and is in some sense a natural extension of antisymmetrization. A spin-adapted CSF may be written as
Θ = A ^ i N φ i ( r i ) X S ( s 1 , ... , s N )
(15)
where we have retained the simplest Hartree product in the spatial part, but assumed a generalized spin part XS which is an eigenfunction of Ŝ2 with total spin S. Since XS can be written as a linear combination of spin function products of the type appearing in eq 14, it also follows that Θ can be written as a linear combination of determinants, although CSFs can be constructed directly without determinants as intermediates. (103,118) The general construction of XS remains an elaborate group theoretical exercise. Matsen’s spin-free approach (213) developed in the 1960s relied on the observation that enforcing the SU(2) spin symmetry in the spin part of the wave function is equivalent to representing the N-particle spatial part in terms of the irreducible representations of the symmetric group SN of permutations. (118) On the other hand, the spin-free representation of the Hamiltonian is given in terms of the generators of the unitary group U(n) in the n-orbital basis. The latter observation is the basis of the unitary group approach (214) (UGA) and its graphical application (215) (GUGA), which was recently discussed in detail by Dobrautz et al. (216) and the advantages of which have also been exploited in a number of publications. (104−106) We will return to some aspects of spin-adaptation and its consequences for the correlation problem in the next section, and similar considerations will also occur in the section on the role of orbitals.
Finally, we shall briefly consider the problem of spatial symmetry since it does appear in discussions on strong correlation. (107) The prototypical example is the He2+ system (217,218) which dissociates into a He atom and a He+ ion. At large distances, two electrons are localized on one of the He nuclei and one electron around the other. In the minimal basis, this means that the Hartree–Fock orbital associated with the two electrons has a larger spatial extent than the other one, and they do not reflect the point group symmetry of He2+. Moreover, there are two symmetry broken solutions depending on whether the two electrons are localized around nucleus A or B, corresponding in essence to two resonance structures. It is possible to enforce spatial symmetry by constructing orbitals of the type φA ± φB, where φA and φB are orbitals localized around A and B. A determinant built from such orbitals could be expanded in terms of the resonance structures mentioned, but would also include high-energy ionic configurations, giving rise to an artificially large energy at long bond distances. Thus, this is the spatial equivalent of the dissociation problem of the H2 molecule in which breaking spin (and spatial) symmetry leads to a lower energy. Usually, these problems are regarded multiconfigurational, (218) because more than one configuration is needed to give the correct energy without breaking symmetry. Unfortunately, at the level of approximate wave functions (often used as a reference), enforcing spatial symmetry may easily lead to additional complications. Since infinitesimal changes may disrupt the symmetric arrangement of the nuclei, this may lead to discontinuities in the energy or the need to use suboptimal symmetry-adapted orbitals. (219) Thus, using SA-CSFs may introduce more conceptual uncertainty about symmetry-related correlation effects than simply using DETs or even CSFs (CFGs). We will not consider SA-CSFs here any further, but the review of Davidson and Borden (220) discusses the issue of spatial symmetry breaking in all its aspects, be it “real or artifactual”, as they put it, and a series of papers by Čížek and Paldus on the HF stability conditions (221−223) also devotes considerable space to this problem.
It has been noted before that the various categories for correlation are operational, i.e., what counts as static or nondynamical correlation is practically whatever is covered by choice at the level of the reference calculation. Symmetry on the other hand introduces exact degeneracies that are known a priori and reference functions can be constructed to account for them. Thus, in line with the general thrust of many discussions on the subject, it seems worthwhile to distinguish between correlation effects due to various kinds of symmetry and the rest. We may call these symmetry effects static correlation. (12,17) In the next section, we will discuss accounting for spin symmetry in the basis states. The remaining correlation effects may be weak or strong depending on the weight of the symmetry-adapted reference state in the final calculation. In this regard, near degeneracies are of interest since they do not arise from the symmetry effects discussed so far. The case of the Be atom is instructive: (19) due to the near degeneracy of the 2s and 2p orbitals, the 1s22s2 configuration does not completely dominate the wave function expansion, the 1s22p2 configurations must also be considered. Such nondynamical effects are not captured by symmetry restrictions but are usually still included in the reference calculation, leaving any residual correlation effects, i.e., dynamic correlation, to be recovered by the wave function Ansatz built on that reference. Handy and co-workers (70) even go so far as to suggest that even this “angular” correlation in Be, so-called because s and p orbitals are of different angular types, might be regarded as dynamical based on a model in which explicit dependence on the interelectronic distance appears. The Be example also indicates that the weight of the reference state is of crucial importance and what is regarded as nondynamical or strong depends on a threshold value set on this weight. The dependence of weight-based measures on the orbital space will also be addressed in a subsequent section using a simple bond breaking process as an example, another situation in which strong correlation effects are typically expected.

7. Representing the Wave Function

ARTICLE SECTIONS
Jump To

It has already been mentioned that DETs are only eigenfunctions of Ŝz, while CSFs are also eigenfunctions of the total spin. The total number of DETs (80) associated with a CFG and with a given spin-projection MS and number of unpaired electrons N is
f N M = ( N 1 2 N M S )
(16)
from which the CSFs with different possible total spin values MSS can also be built. The number of CSFs (80) with S = MS is
f N S = ( N 1 2 N S ) ( N 1 2 N S 1 )
(17)
which is obviously less than the number of determinants. In fact, summing over all possible S values yields the number of determinants. (80)
To help ground these notions, consider some simple examples collected in Table 1, and discussed in detail elsewhere. (80) Taking the example of two electrons in two orbitals, the electron pair may be placed on the same orbital as in the configurations 20 and 02, or on different orbitals as in the configuration 11. In the former two cases, a single CSF consisting of a single determinant is enough to describe a closed-shell singlet, hence the success of determinants in many situations in chemistry. In the 11 case, a singlet and a triplet CSF can be built from two determinants, though the other two triplet states need only a single determinant. A slightly more complicated example involves distributing three electrons among three orbitals. For the configuration 111, the S = 3/2 cases yield nothing conceptually new, they are described by CSFs consisting of a single or multiple determinants. On the other hand, the S = 1/2, MS = 1/2 subspace can only be spanned by two degenerate CSFs, themselves consisting of multiple determinants. Since the reference function is often obtained at the Hartree–Fock level, it is worth recalling how this discussion affects such calculations. In more commonly occurring cases, the spin-adapted HF wave function can be constructed from either a single determinant, as in the typical closed-shell case, or from a single CSF, as in a low-spin open-shell case. (224) In those open-shell cases, such as S = 1/2, MS = 1/2 for 111, where a configuration cannot be uniquely described by a single CSF, (225) Edwards and Zerner recommend in a footnote that a HF calculation using a chosen CSF should be followed by a configuration interaction (CI) calculation. (225) This might minimally include only the CSFs belonging to the CFG in question, e.g., the two degenerate ones in the 111 case, possibly with or without orbital optimization, though the question remains whether such a small CI calculation would accurately reflect the exact solution. Indeed, Edwards and Zerner recommend CI in the active space, i.e., presumably including multiple configurations that might produce the same total and projected spin (e.g., 201 in the three-electron case). Nevertheless, the expansion coefficients of the CSFs of a single CFG will be of interest here for analytic purposes.
Table 1. Simple Examples of N-Electron Basis States, including Configurations (CFG), Configuration State Functions (CSF) and Determinants (DET)a
S MS CFG CSF DET
0 0 20 |11̅| |11̅|
0 0 02 |22̅| |22̅|
0 0 11 (|12̅|−|1̅2|) |12̅|, |1̅2|
1 1 11 |12| |12|
1 0 11 (|12̅|+|1̅2|) |12̅|, |1̅2|
1 –1 11 |1̅2̅| |1̅2̅|
3/2 3/2 111 |123| |123|
3/2 1/2 111 (|1̅23|+|12̅3|+|123̅|) |1̅23|, |12̅3|, |123̅|
1/2 1/2 111 (2|123̅|−|12̅3|−|1̅23|), |1̅23|, |12̅3|, |123̅|
a

Spatial orbitals are labelled simply as 1, 2, or 3 in DETs and CSFs, while the corresponding occupation numbers (0, 1 or 2) are indicated in CFGs.

It is worth noting that already at the HF level spin-restricted open-shell CSFs are considerably harder to optimize than the unrestricted determinants. This has led to several ways of circumventing the direct use of CSFs. A simplified version (130) of Löwdin’s projection operator (226) may be used to remove the spin-components that lie nearest to the desired state from the spin-unrestricted determinant, but in any case, spin-projection leads to artifacts if applied after optimization and falls short of size-consistency if made part of it. (227) More recent projection approaches constrain the eigenvalues of the spin-density to obtain results that are identical to the more cumbersome conventional spin-restricted approach. (228) Even more pragmatically, quasi-restricted approaches of various kinds have been proposed to obtain orbitals resembling restricted ones from some simple recipe. (229,230) More complicated Ansätze are usually obtained from reference states by (ideally spin-adapted) excitation (or ionization) operators, (214) and in a similar vein, it is also possible to define spin-flip operators that change the spin state of the reference. (231) In simple cases, such spin-flip operators may even be spin-adapted themselves. (232) Sometimes symmetry-broken solutions may be used creatively, e.g., to obtain some property of spin-states as in PO/BS-DFT discussed above, or in combining unrestricted determinants for quasi-spin-adaptation, (233) sometimes considerations about the CSF expansion are employed innovatively, e.g., to obtain singlet–triplet gaps from bideterminantal MS = 0 states (see Table 1) in cases in which methods with a single determinant bias encounter multideterminant situations. (234)
At this stage, there is some ambiguity whether we associate multireference with the requirement that the exact solution is dominated by more than a single determinant, CSF or configuration. In a normalized FCI expansion of the type in eq 8, a set of basis states dominates the expansion if their cumulative weight, ∑P |CP|2 is larger than the rest of the weights combined, i.e., 1 – ∑P |CP|2. Assuming knowledge of the exact solution, the following categories are proposed (cf. ref (123)):
  • Single Determinant (SD) Expansion: The weight of a single determinant dominates, and hence a single-determinant HF calculation is enough for a reference.

  • Single Spin-Coupling (SS) Expansion: It consists of a linear combination of determinants whose relative weights are fixed but the weight of the resulting CSF dominates the expansion. An restricted open-shell calculation should be a good reference.

  • Single Configurational (SC) Expansion: Such an expansion has multiple CSFs belonging to the same configuration. An expansion would fall into this category if the cumulative weight of CSFs of a single configuration dominates. A restricted open-shell calculation followed by a minimal CI calculation to determine the relative weight of the CSFs involved might be a good reference.

  • Multiconfiguration (MC) Expansion: No configuration alone dominates, and hence this is the true multireference case since no single reference function of the kinds discussed before can be found.

Note that the single reference (SR) categories above may overlap, since e.g., the closed-shell ground state of a two-electron system is at the same time SD, SS and SC. An open-shell singlet is SS and SC, but not SD. The three-electron doublet that has two CSFs would be SC but not SS or SD, if the cumulative weight of those CSFs dominate the expansion. All these are examples of various SR expansions. The simplest example of an MR expansion would be a combination of the type 02 + 20, as in the case of the dissociating H2 molecule. In the next section, we shall discuss how the choice of orbital space impacts this classification. Once spin-coupling is taken care of, an important source of correlation effects is addressed and the question remains how strong the remaining correlation effects are. This reflects the distinction between static and nondynamic correlation as used by Bartlett and Stanton, (17) with static correlation being eliminated by spin-adaptation. In this regard, this generalization of the excitation schemes can be compared with entanglement (22) and seniority (189) approaches which also aim at finding a good reference based on different criteria.

8. The Role of the Orbital Space

ARTICLE SECTIONS
Jump To

The idea that the weight of the reference function in an approximate CI calculation could be used as an indication of multireference character is an old one, but it was also observed long ago that due to the bias of canonical orbitals toward the reference (HF) state, it is of limited use. (110) Furthermore, it was also suggested that such CI calculations should be complete within at least an active space for this metric to be of use. (109) In that case, the leading CAS coefficient in the basis of natural orbitals using determinants or CSFs was found to be a sensible metric. (235,236) The role of the orbital space has been discussed in connection to most approaches discussed in this perspective, (109,123,186,189,193) and orbital localization techniques form a part of practical DMRG (237) calculations as well as reducing the complexity of CAS (238) calculations. Within the DMRG context, orbital optimization techniques have been recently suggested as a means of reducing the multireference character of the wave function. (239) A similar result has been observed in SCI methods, where orbital optimization leads to a more rapid convergence to the FCI solution. (240) The recent work of Li Manni and co-workers on GUGA based FCIQMC (104−106) and its application as a stochastic CASCI solver (241) shed even more light on the role of the orbital space. This methodology has the following features relevant for the current discussion: a) using GUGA to calculate spin-adapted basis states efficiently; b) using a unitary orbital transformation to compress the wave function expansion; c) the combined use of unitary and symmetric group approaches means that the ordering of orbitals also affects the complexity of the wave function. The technique has been used to reduce the number of CSFs expected from the exponentially scaling combinatorial formula to a small number of CSFs, sometimes only a single one. (106) Moreover, not only is the wave function more compact, but the Hamiltonian assumes a convenient block-diagonal form that allows for the selective targeting of excited states. (106) Interestingly, it was also found that the best orbital ordering for a spin-adapted CSF is not identical to the best site-ordering in the DMRG sense. (242) More recent work has also attempted to place the orbital ordering process on a more systematic basis exploiting the commutation properties of cumulative spin operators and the Hamiltonian. (243) Here, the role of the orbital space will be illustrated via a simple example that will remain within a more traditional CAS/FCI context.
In this perspective, we are concerned with the a posteriori characterization of states rather than finding an a priori indicator of multireference character, and hence we will assume that an FCI solution is available. We will illustrate the role of orbital spaces by considering the H2 bond dissociation problem using canonical and localized orbitals. (244) We will define the multireference character M as
M = 1 P | C P | 2 1 + P | C P | 2
(18)
where the summation runs over the DETs or CSFs belonging to the dominant CFG. This measure will converge to 1 for infinitely many coefficients assuming they are all equal, i.e., the worst MR case. In SR cases, M should be close to 0, and in the H2 case M = 1/3 corresponds to two configurations of equal weight. In practice, one might choose a different threshold on the cumulative weight of the configuration. For example, requiring the dominant cumulative weights to be above 80% leads to M ≤ 1/9. If such a dominant configuration exists (and if it were known in advance), it would be the ideal reference function on which a correlated Ansatz could be built.
Figure 1 shows the value of M at different internuclear separations D. In the canonical basis, a single configuration 20 is dominant around the equilibrium distance, while at infinite separation the CFGs 20 and 02 contribute equally. Thus, in the latter case, the expansion is multiconfigurational as well as multideterminantal. Evaluating M in the basis of local orbitals yields to the opposite results. The local orbitals are each essentially localized on one atom, and at long bond distance, a single CSF of the 11 configurations can describe the dissociation. Thus, the expansion is SC and SS, but not SD. At shorter bond lengths, the localized expansion is much less appropriate as seen from the rising multireference character. Changing from the canonical to the local orbital basis has a similar effect on entanglement-based metrics: (193) in the canonical basis entanglement changes between its minimal and maximal values as D grows and the other way around in the local basis since in the latter all occupation patterns are equally likely at small D. This ties in also with our earlier remarks on the relationship between VB and MO approaches.

Figure 1

Figure 1. Multireference character M as a function of internuclear distance D in the H2 dissociation problem for the FCI/STO-1G ground state. The blue curve corresponds to canonical orbitals (can), the red one to localized orbitals (loc).

It has been argued earlier that an important source of correlation effects may be eliminated by spin adaptation. Yet, because of Löwdin’s dilemma, the constrained reference energy may increase and this could in turn lead to an increase in the correlation energy. In the H2 example, the spin-restricted closed-shell reference function has been found qualitatively adequate around the equilibrium bond length, but at large distances it leads to a catastrophic increase in the magnitude of the reference and correlation energies. Using canonical orbitals, the remedy involves combining the 20 and 02 configurations. Thus, it seems that enforcing spin symmetry makes the problem more complicated by requiring a multiconfigurational treatment whereas the spin-unrestricted single determinant also converges to the qualitatively correct energy at large bond lengths. This is because it reduces to one of the symmetry broken determinants |μν̅| or |μ̅ν| in this limit. In general, these broken symmetry determinants produce an energy that lies between the open shell singlet in eq 7 and triplet states that can also be constructed from them, and as such, they are not a good approximation to either. At large bond distances, however, the singlet–triplet gap closes and the energy converges to the correct limit while the wave function is not the exact one. (18,23) While symmetry breaking may be desirable under some circumstances, for finite molecular systems symmetric solutions are in general preferred. In the open-shell singlet case, for example, spin-adaptation is indispensable for a qualitatively accurate description. It is here that orbital rotations may play some role: by localizing the orbitals at large distances, it is possible to convert the multiconfigurational representation of the dissociating H2 molecule in the canonical basis into a single configurational open-shell singlet in the local basis.
The fact that the multireference character depends on the orbital representation warns against assigning the multiconfigurational nature of the wave function expansion as a physical property of a state of the system. At the very least, one should also demand that there be no unitary transformation originating in the orbital space that removes the multireference character of an expansion completely. Of course, the analysis as presented here is not a practical recipe for computation, since it assumes knowledge of the FCI solution, although it may be used fruitfully in interpreting less costly wave functions of the CASCI or CASSCF type. Furthermore, it does underline the potential of orbital rotations in reducing the multireference character. This important subject will be analyzed in more detail in a forthcoming publication.

9. Outlook

ARTICLE SECTIONS
Jump To

The distinctions we introduced among single-determinantal, single spin-coupling, single configurational and multiconfigurational expansions should have a relevance in any chemical system, as also pointed out by other authors. (17,19) These categories form a clear hierarchy of SR measures in the sense that they are increasingly less restrictive: SD implies SS and both SD and SS imply SC. Unlike in extended systems, where symmetry broken solutions are often preferred, symmetry-adapted wave functions are often needed to reliably predict certain chemical properties of molecules. Again, unlike in periodic solids, strongly interacting regimes are usually easier to localize, and hence the complexity of the wave function can often be reduced. If spin coupling effects can be incorporated into a single reference function, then the remaining correlation effects might be weak, and in cases when they are not, such as in certain regions of bond dissociation curves, they can be typically taken care of by small active space calculations. As observed by Malrieu et al., (245) the fact that one can always build a small complete active space representation by placing two electrons in a pair of localized bonding and antibonding orbitals that accounts for such correlation effects is perhaps the best unbiased confirmation of Lewis’ idea of the chemical bond consisting of an electron pair. (246) We have discussed the effect of orbital rotations on the multireference character in a simple bond dissociation process and pointed out that such orbital rotations have the potential of reducing it. If by strong correlation we also mean the large size of the active space necessary to remedy these nondynamical effects, then it seems likely that most molecular systems are not strongly correlated, with the possible exception of multisite antiferromagnetically coupled systems. (107,247−251) Thus, the complexity of most finite chemical systems should be well below exponential and should in most cases involve only a few configurations. Devising an appropriate metric that can help find these a priori is a challenge, but measures based on leading coefficients, density matrices, natural orbitals and orbital entanglement have been used for this purpose in the literature.

Author Information

ARTICLE SECTIONS
Jump To

Acknowledgments

ARTICLE SECTIONS
Jump To

The authors thank Earl T. Cambell, Giovanni Li Manni and Pavel Pokhilko for useful discussions on the paper.

References

ARTICLE SECTIONS
Jump To

This article references 251 other publications.

  1. 1
    Kurth, S.; Perdew, J. P. Role of the exchange–correlation energy: Nature’s glue. Int. J. Quantum Chem. 2000, 77, 814818,  DOI: 10.1002/(SICI)1097-461X(2000)77:5<814::AID-QUA3>3.0.CO;2-F
  2. 2
    Martin, J. M. Electron Correlation: Nature’s Weird and Wonderful Chemical Glue. Isr. J. Chem. 2022, 62, e202100111,  DOI: 10.1002/ijch.202100111
  3. 3
    Löwdin, P.-O. Correlation Problem in Many-Electron Quantum Mechanics I. Review of Different Approaches and Discussion of Some Current Ideas. In Advances in Chemical Physics; Interscience: New York, 1959; Vol. 2; Chapter 7, pp 207322.
  4. 4
    Kotani, M.; Ohno, K.; Kayama, K. Quantum Mechanics of Electronic Structure of Simple Molecules. In Encyclopedia of Physics/Handbuch der Physik; Springer: Berlin, 1961; Vol. 37.2, Chapter 1, pp 1172.
  5. 5
    Slater, J. C. Quantum Theory of Molecules and Solids. Vol. 1: Electronic Structure of Molecules; McGraw-Hill: New York, 1963.
  6. 6
    Kutzelnigg, W. Einführung in die theoretische Chemie; Wiley-VCH: Weinheim, 2002; Vol. 2.
  7. 7
    Wigner, E. On the interaction of electrons in metals. Phys. Rev. 1934, 46, 1002,  DOI: 10.1103/PhysRev.46.1002
  8. 8
    Wigner, E. Effects of the electron interaction on the energy levels of electrons in metals. Trans. Faraday Soc. 1938, 34, 678685,  DOI: 10.1039/tf9383400678
  9. 9
    Löwdin, P.-O. The historical development of the electron correlation problem. Int. J. Quantum Chem. 1995, 55, 77102,  DOI: 10.1002/qua.560550203
  10. 10
    Sinanoğlu, O. Many-Electron Theory of Atoms and Molecules. Proc. Nat. Acad. Sci. 1961, 47, 12171226,  DOI: 10.1073/pnas.47.8.1217
  11. 11
    Sinanoğlu, O. Many-Electron Theory of Atoms, Molecules and Their Interactions. In Advances in Chemical Physics; John Wiley & Sons, Ltd: London, 1964; Vol. 6, Chapter 7, pp 315412.
  12. 12
    McWeeny, R. The nature of electron correlation in molecules. Int. J. Quantum Chem. 1967, 1, 351359,  DOI: 10.1002/qua.560010641
  13. 13
    Kutzelnigg, W.; Del Re, G.; Berthier, G. Correlation coefficients for electronic wave functions. Phys. Rev. 1968, 172, 4959,  DOI: 10.1103/PhysRev.172.49
  14. 14
    von Herigonte, P. Electron correlation in the seventies. In Struct. Bonding (Berlin); Springer: New York, 1972; Vol. 12, Chapter 1, pp 147.
  15. 15
    Kutzelnigg, W.; von Herigonte, P. Electron correlation at the dawn of the 21st century. In Adv. Quantum Chem.; Academic Press: San Diego, 2000; Vol. 36, pp 185229.
  16. 16
    Kutzelnigg, W. Theory of electron correlation. In Explicitly Correlated Wave Functions in Chemistry and Physics; Springer Science & Business Media: Dordrecht, 2003; Chapter 1, pp 390.
  17. 17
    Bartlett, R. J.; Stanton, J. F. Applications of Post-Hartree─Fock Methods: A Tutorial. In Reviews in Computational Chemistry; VCH Publishers: New York, 1994; Chapter 2, pp 65169.
  18. 18
    Knowles, P. J.; Schütz, M.; Werner, H.-J. Ab Initio Methods for Electron Correlation in Molecules. In Modern Methods and Algorithms of Quantum Chemistry; Grotendorst, J., Ed.; NIC: Jülich, 2000; Vol. 1, pp 69151.
  19. 19
    Tew, D. P.; Klopper, W.; Helgaker, T. Electron correlation: The many-body problem at the heart of chemistry. J. Comput. Chem. 2007, 28, 13071320,  DOI: 10.1002/jcc.20581
  20. 20
    Senatore, G.; March, N. Recent progress in the field of electron correlation. Rev. Mod. Phys. 1994, 66, 445479,  DOI: 10.1103/RevModPhys.66.445
  21. 21
    Loos, P.-F.; Gill, P. M. The uniform electron gas. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2016, 6, 410429,  DOI: 10.1002/wcms.1257
  22. 22
    Chan, G. K.-L. Low entanglement wavefunctions. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2012, 2, 907920,  DOI: 10.1002/wcms.1095
  23. 23
    Fulde, P. Electron correlations in molecules and solids; Springer-Verlag: Berlin, 1995.
  24. 24
    March, N. H. Electron correlation in molecules and condensed phases; Springer Science & Business Media: New York, 1996.
  25. 25
    Wilson, S. Electron Correlation in Molecules; Dover: Mineola, 2007.
  26. 26
    Advances in Chemical Physics; Vol. 14; Lefebvre, R., Moser, C., Eds.; John Wiley & Sons, Ltd: London, 1969 Correlation Effects in Atoms and Molecules.
  27. 27
    Electron Correlation in the Solid State; March, N. H., Ed.; Imperial Collage Press: London, 1999.
  28. 28
    Raimes, S. Many-electron Theory; North-Holland: Amsterdam, 1972.
  29. 29
    Gell-Mann, M.; Brueckner, K. A. Correlation energy of an electron gas at high density. Phys. Rev. 1957, 106, 364368,  DOI: 10.1103/PhysRev.106.364
  30. 30
    Hund, F. Zur Deutung der Molekelspektren. Z. Phys. 1927, 40, 742764,  DOI: 10.1007/BF01400234
  31. 31
    Mulliken, R. S. The assignment of quantum numbers for electrons in molecules. I. Phys. Rev. 1928, 32, 186222,  DOI: 10.1103/PhysRev.32.186
  32. 32
    Lennard-Jones, J. E. The electronic structure of some diatomic molecules. Trans. Faraday Soc. 1929, 25, 668686,  DOI: 10.1039/tf9292500668
  33. 33
    Heitler, W.; London, F. Wechselwirkung neutraler Atome und homöopolare Bindung nach der Quantenmechanik. Z. Phys. 1927, 44, 455472,  DOI: 10.1007/BF01397394
  34. 34
    Pauling, L. The nature of the chemical bond. Application of results obtained from the quantum mechanics and from a theory of paramagnetic susceptibility to the structure of molecules. J. Am. Chem. Soc. 1931, 53, 13671400,  DOI: 10.1021/ja01355a027
  35. 35
    Coulson, C. A.; Fischer, I. Notes on the molecular orbital treatment of the hydrogen molecule. Philos. Mag. 1949, 40, 386393,  DOI: 10.1080/14786444908521726
  36. 36
    Kutzelnigg, W.; Mukherjee, D. Cumulant expansion of the reduced density matrices. J. Chem. Phys. 1999, 110, 28002809,  DOI: 10.1063/1.478189
  37. 37
    Wigner, E.; Seitz, F. On the Constitution of Metallic Sodium. Phys. Rev. 1933, 43, 804810,  DOI: 10.1103/PhysRev.43.804
  38. 38
    Wigner, E.; Seitz, F. On the Constitution of Metallic Sodium. II. Phys. Rev. 1934, 46, 509524,  DOI: 10.1103/PhysRev.46.509
  39. 39
    McWeeny, R. The density matrix in many-electron quantum mechanics I. Generalized product functions. Factorization and physical interpretation of the density matrices. Proc. R. Soc. A 1959, 253, 242259,  DOI: 10.1098/rspa.1959.0191
  40. 40
    Giner, E.; Tenti, L.; Angeli, C.; Malrieu, J.-P. The “Fermi hole” and the correlation introduced by the symmetrization or the anti-symmetrization of the wave function. J. Chem. Phys. 2016, 145, 124114,  DOI: 10.1063/1.4963018
  41. 41
    Kutzelnigg, W.; Mukherjee, D. Normal order and extended Wick theorem for a multiconfiguration reference wave function. J. Chem. Phys. 1997, 107, 432449,  DOI: 10.1063/1.474405
  42. 42
    Evangelista, F. A. Automatic derivation of many-body theories based on general Fermi vacua. J. Chem. Phys. 2022, 157, 064111,  DOI: 10.1063/5.0097858
  43. 43
    Tsuchimochi, T.; Scuseria, G. E. Strong correlations via constrained-pairing mean-field theory. J. Chem. Phys. 2009, 131, 121102,  DOI: 10.1063/1.3237029
  44. 44
    Kutzelnigg, W. Separation of strong (bond-breaking) from weak (dynamical) correlation. Chem. Phys. 2012, 401, 119124,  DOI: 10.1016/j.chemphys.2011.10.020
  45. 45
    Georges, A.; Kotliar, G. Hubbard model in infinite dimensions. Phys. Rev. B 1992, 45, 64796483,  DOI: 10.1103/PhysRevB.45.6479
  46. 46
    Georges, A.; Kotliar, G.; Krauth, W.; Rozenberg, M. J. Dynamical mean-field theory of strongly correlated fermion systems and the limit of infinite dimensions. Rev. Mod. Phys. 1996, 68, 13125,  DOI: 10.1103/RevModPhys.68.13
  47. 47
    Anisimov, V.; Izyumov, Y. Electronic Structure of Strongly Correlated Materials; Springer: Berlin, 2010.
  48. 48
    Kuramoto, Y. Quantum Many-Body Physics; Springer: Tokyo, 2020.
  49. 49
    Foulkes, W. M. C.; Mitas, L.; Needs, R. J.; Rajagopal, G. Quantum Monte Carlo simulations of solids. Rev. Mod. Phys. 2001, 73, 3383,  DOI: 10.1103/RevModPhys.73.33
  50. 50
    Zhang, S.; Malone, F. D.; Morales, M. A. Auxiliary-field quantum Monte Carlo calculations of the structural properties of nickel oxide. J. Chem. Phys. 2018, 149, 164102,  DOI: 10.1063/1.5040900
  51. 51
    Exactly solvable models of strongly correlated electrons; Korepin, V. E., Essler, F. H., Eds.; World Scientific: Singapore, 1994.
  52. 52
    Gross, D. J. The role of symmetry in fundamental physics. Proc. Nat. Acad. Sci. 1996, 93, 1425614259,  DOI: 10.1073/pnas.93.25.14256
  53. 53
    Raimes, S. The wave mechanics of electrons in metals; North-Holland: Amsterdam, 1963.
  54. 54
    Fulde, P. Solids with weak and strong electron correlations. In Electron Correlation in the Solid State; Imperial Collage Press: London, 1999; Chapter 2, pp 47102.
  55. 55
    Coldwell-Horsfall, R. A.; Maradudin, A. A. Zero-Point Energy of an Electron Lattice. J. Math. Phys. 1960, 1, 395404,  DOI: 10.1063/1.1703670
  56. 56
    Bohm, D.; Pines, D. A Collective Description of Electron Interactions. I. Magnetic Interactions. Phys. Rev. 1951, 82, 625634,  DOI: 10.1103/PhysRev.82.625
  57. 57
    Pines, D. Electron Interaction in Metals. In Solid State Physics; Academic Press: New York, 1955; Vol. 1, pp 367450.
  58. 58
    Slater, J. C. The electronic structure of solids. In Encyclopedia of Physics/Handbuch der Physik; Springer: Berlin, 1956; Vol. 19, pp 1136.
  59. 59
    Brueckner, K. A. The Correlation Energy of a Non-Uniform Electron Gas. In Advances in Chemical Physics; John Wiley & Sons, Ltd: London, 1969; Vol. 14, Chapter 7, pp 215236.
  60. 60
    Slater, J. C. The Virial and Molecular Structure. J. Chem. Phys. 1933, 1, 687691,  DOI: 10.1063/1.1749227
  61. 61
    March, N. H. Kinetic and Potential Energies of an Electron Gas. Phys. Rev. 1958, 110, 604605,  DOI: 10.1103/PhysRev.110.604
  62. 62
    Argyres, P. N. Virial Theorem for the Homogeneous Electron Gas. Phys. Rev. 1967, 154, 410413,  DOI: 10.1103/PhysRev.154.410
  63. 63
    Ruedenberg, K. The Physical Nature of the Chemical Bond. Rev. Mod. Phys. 1962, 34, 326376,  DOI: 10.1103/RevModPhys.34.326
  64. 64
    Hubbard, J. Electron correlations in narrow energy bands. Proc. R. Soc. A 1963, 276, 238257,  DOI: 10.1098/rspa.1963.0204
  65. 65
    Janesko, B. G. Strong correlation in surface chemistry. Mol. Simul. 2017, 43, 394405,  DOI: 10.1080/08927022.2016.1261136
  66. 66
    Motta, M.; Ceperley, D. M.; Chan, G. K.-L.; Gomez, J. A.; Gull, E.; Guo, S.; Jiménez- Hoyos, C. A.; Lan, T. N.; Li, J.; Ma, F.; Millis, A. J.; Prokof’ev, N. V.; Ray, U.; Scuseria, G. E.; Sorella, S.; Stoudenmire, E. M.; Sun, Q.; Tupitsyn, I. S.; White, S. R.; Zgid, D.; Zhang, S. Towards the Solution of the Many-Electron Problem in Real Materials: Equation of State of the Hydrogen Chain with State-of-the-Art Many-Body Methods. Phys. Rev. X 2017, 7, 031059,  DOI: 10.1103/PhysRevX.7.031059
  67. 67
    Motta, M.; Genovese, C.; Ma, F.; Cui, Z.-H.; Sawaya, R.; Chan, G. K.-L.; Chepiga, N.; Helms, P.; Jiménez-Hoyos, C.; Millis, A. J.; Ray, U.; Ronca, E.; Shi, H.; Sorella, S.; Stoudenmire, E. M.; White, S. R.; Zhang, S. Ground-State Properties of the Hydrogen Chain: Dimerization, Insulator-to-Metal Transition, and Magnetic Phases. Phys. Rev. X 2020, 10, 031058
  68. 68
    Sinanoğlu, O.; Tuan, D. F.-t. Many-Electron Theory of Atoms and Molecules. III. Effect of Correlation on Orbitals. J. Chem. Phys. 1963, 38, 17401748,  DOI: 10.1063/1.1776948
  69. 69
    Prendergast, D.; Nolan, M.; Filippi, C.; Fahy, S.; Greer, J. Impact of electron–electron cusp on configuration interaction energies. J. Chem. Phys. 2001, 115, 16261634,  DOI: 10.1063/1.1383585
  70. 70
    Mok, D. K.; Neumann, R.; Handy, N. C. Dynamical and nondynamical correlation. J. Phys. Chem. 1996, 100, 62256230,  DOI: 10.1021/jp9528020
  71. 71
    Hollett, J. W.; Gill, P. M. The two faces of static correlation. J. Chem. Phys. 2011, 134, 114111,  DOI: 10.1063/1.3570574
  72. 72
    Benavides-Riveros, C. L.; Lathiotakis, N. N.; Marques, M. A. Towards a formal definition of static and dynamic electronic correlations. Phys. Chem. Chem. Phys. 2017, 19, 1265512664,  DOI: 10.1039/C7CP01137G
  73. 73
    Bulik, I. W.; Henderson, T. M.; Scuseria, G. E. Can single-reference coupled cluster theory describe static correlation?. J. Chem. Theory Comput. 2015, 11, 31713179,  DOI: 10.1021/acs.jctc.5b00422
  74. 74
    Karton, A.; Rabinovich, E.; Martin, J. M.; Ruscic, B. W4 theory for computational thermochemistry: In pursuit of confident sub-kJ/mol predictions. J. Chem. Phys. 2006, 125, 144108,  DOI: 10.1063/1.2348881
  75. 75
    Hait, D.; Tubman, N. M.; Levine, D. S.; Whaley, K. B.; Head-Gordon, M. What levels of coupled cluster theory are appropriate for transition metal systems? A study using near-exact quantum chemical values for 3d transition metal binary compounds. J. Chem. Theory Comput. 2019, 15, 53705385,  DOI: 10.1021/acs.jctc.9b00674
  76. 76
    Roos, B. O.; Taylor, P. R.; Sigbahn, P. E. A complete active space SCF method (CASSCF) using a density matrix formulated super-CI approach. Chem. Phys. 1980, 48, 157173,  DOI: 10.1016/0301-0104(80)80045-0
  77. 77
    Chan, G. K.-L.; Sharma, S. The density matrix renormalization group in quantum chemistry. Annu. Rev. Phys. Chem. 2011, 62, 465481,  DOI: 10.1146/annurev-physchem-032210-103338
  78. 78
    Mouesca, J.-M. Density functional theory-broken symmetry (DFT-BS) methodology applied to electronic and magnetic properties of bioinorganic prosthetic groups. In Metalloproteins; Springer: New York, 2014; pp 269296.
  79. 79
    Scuseria, G. E.; Jiménez-Hoyos, C. A.; Henderson, T. M.; Samanta, K.; Ellis, J. K. Projected quasiparticle theory for molecular electronic structure. J. Chem. Phys. 2011, 135, 124108,  DOI: 10.1063/1.3643338
  80. 80
    Helgaker, T.; Jorgensen, P.; Olsen, J. Molecular electronic-structure theory; John Wiley & Sons: Chichester, 2000.
  81. 81
    Shavitt, I.; Bartlett, R. J. Many-body methods in chemistry and physics: MBPT and coupled-cluster theory; Cambridge University Press: Cambridge, 2009.
  82. 82
    Čížek, J. On the Correlation Problem in Atomic and Molecular Systems. Calculation of Wavefunction Components in Ursell-Type Expansion Using Quantum-Field Theoretical Methods. J. Chem. Phys. 1966, 45, 42564266,  DOI: 10.1063/1.1727484
  83. 83
    Paldus, J.; Čížek, J.; Shavitt, I. Correlation Problems in Atomic and Molecular Systems. IV. Extended Coupled-Pair Many-Electron Theory and Its Application to the BH3 Molecule. Phys. Rev. A 1972, 5, 5067,  DOI: 10.1103/PhysRevA.5.50
  84. 84
    Arponen, J. Variational principles and linked-cluster exp S expansions for static and dynamic many-body problems. Ann. Phys. 1983, 151, 311382,  DOI: 10.1016/0003-4916(83)90284-1
  85. 85
    Faulstich, F. M.; Laestadius, A.; Legeza, O.; Schneider, R.; Kvaal, S. Analysis of the Tailored Coupled-Cluster Method in Quantum Chemistry. SIAM J. Numer. Anal. 2019, 57, 25792607,  DOI: 10.1137/18M1171436
  86. 86
    Csirik, M. A.; Laestadius, A. Coupled-cluster theory revisited. arXiv:2105.13134 2021, DOI: 10.48550/arXiv.2105.13134 .
  87. 87
    Schütz, M.; Werner, H.-J. Low-order scaling local electron correlation methods. IV. Linear scaling local coupled-cluster (LCCSD). J. Chem. Phys. 2001, 114, 661681,  DOI: 10.1063/1.1330207
  88. 88
    Riplinger, C.; Neese, F. An efficient and near linear scaling pair natural orbital based local coupled cluster method. J. Chem. Phys. 2013, 138, 034106,  DOI: 10.1063/1.4773581
  89. 89
    Shiozaki, T.; Valeev, E. F.; Hirata, S. Explicitly correlated coupled-cluster methods. In Annual Reports in Computational Chemistry; Elsevier: Amsterdam, 2009; Vol. 5, Chapter 6, pp 131148.
  90. 90
    Izsák, R. Single-reference coupled cluster methods for computing excitation energies in large molecules: The efficiency and accuracy of approximations. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2020, 10, e1445,  DOI: 10.1002/wcms.1445
  91. 91
    Lyakh, D. I.; Musiał, M.; Lotrich, V. F.; Bartlett, R. J. Multireference nature of chemistry: The coupled-cluster view. Chem. Rev. 2012, 112, 182243,  DOI: 10.1021/cr2001417
  92. 92
    Huron, B.; Malrieu, J. P.; Rancurel, P. Iterative perturbation calculations of ground and excited state energies from multiconfigurational zeroth-order wavefunctions. J. Chem. Phys. 1973, 58, 57455759,  DOI: 10.1063/1.1679199
  93. 93
    Austin, B. M.; Zubarev, D. Y.; Lester Jr, W. A. Quantum Monte Carlo and related approaches. Chem. Rev. 2012, 112, 263288,  DOI: 10.1021/cr2001564
  94. 94
    Booth, G. H.; Thom, A. J.; Alavi, A. Fermion Monte Carlo without fixed nodes: A game of life, death, and annihilation in Slater determinant space. J. Chem. Phys. 2009, 131, 054106,  DOI: 10.1063/1.3193710
  95. 95
    Petruzielo, F. R.; Holmes, A. A.; Changlani, H. J.; Nightingale, M. P.; Umrigar, C. J. Semistochastic Projector Monte Carlo Method. Phys. Rev. Lett. 2012, 109, 230201,  DOI: 10.1103/PhysRevLett.109.230201
  96. 96
    Spencer, J. S.; Blunt, N. S.; Foulkes, W. M. The sign problem and population dynamics in the full configuration interaction quantum Monte Carlo method. J. Chem. Phys. 2012, 136, 054110,  DOI: 10.1063/1.3681396
  97. 97
    Cleland, D.; Booth, G. H.; Alavi, A. Communications: Survival of the fittest: Accelerating convergence in full configuration-interaction quantum Monte Carlo. J. Chem. Phys. 2010, 132, 041103,  DOI: 10.1063/1.3302277
  98. 98
    Cleland, D. M.; Booth, G. H.; Alavi, A. A study of electron affinities using the initiator approach to full configuration interaction quantum Monte Carlo. J. Chem. Phys. 2011, 134, 024112,  DOI: 10.1063/1.3525712
  99. 99
    Eriksen, J. J. The shape of full configuration interaction to come. J. Phys. Chem. Lett. 2021, 12, 418432,  DOI: 10.1021/acs.jpclett.0c03225
  100. 100
    Eriksen, J. J.; Gauss, J. Many-body expanded full configuration interaction. I. Weakly correlated regime. J. Chem. Theory Comput. 2018, 14, 51805191,  DOI: 10.1021/acs.jctc.8b00680
  101. 101
    Eriksen, J. J.; Gauss, J. Many-body expanded full configuration interaction. II. Strongly correlated regime. J. Chem. Theory Comput. 2019, 15, 48734884,  DOI: 10.1021/acs.jctc.9b00456
  102. 102
    White, S. R. Density matrix formulation for quantum renormalization groups. Phys. Rev. Lett. 1992, 69, 28632866,  DOI: 10.1103/PhysRevLett.69.2863
  103. 103
    Chilkuri, V. G.; Neese, F. Comparison of many-particle representations for selected-CI I: A tree based approach. J. Comput. Chem. 2021, 42, 9821005,  DOI: 10.1002/jcc.26518
  104. 104
    Li Manni, G.; Dobrautz, W.; Alavi, A. Compression of spin-adapted multiconfigurational wave functions in exchange-coupled polynuclear spin systems. J. Chem. Theory Comput. 2020, 16, 22022215,  DOI: 10.1021/acs.jctc.9b01013
  105. 105
    Li Manni, G. Modeling magnetic interactions in high-valent trinuclear [Mn3(IV)O4]4+ complexes through highly compressed multi-configurational wave functions. Phys. Chem. Chem. Phys. 2021, 23, 1976619780,  DOI: 10.1039/D1CP03259C
  106. 106
    Li Manni, G.; Dobrautz, W.; Bogdanov, N. A.; Guther, K.; Alavi, A. Resolution of low-energy states in spin-exchange transition-metal clusters: Case study of singlet states in [Fe(III)4S4] cubanes. J. Phys. Chem. A 2021, 125, 47274740,  DOI: 10.1021/acs.jpca.1c00397
  107. 107
    Shee, J.; Loipersberger, M.; Hait, D.; Lee, J.; Head-Gordon, M. Revealing the nature of electron correlation in transition metal complexes with symmetry breaking and chemical intuition. J. Chem. Phys. 2021, 154, 194109,  DOI: 10.1063/5.0047386
  108. 108
    Benavides-Riveros, C. L.; Lathiotakis, N. N.; Schilling, C.; Marques, M. A. Relating correlation measures: The importance of the energy gap. Phys. Rev. A 2017, 95, 032507,  DOI: 10.1103/PhysRevA.95.032507
  109. 109
    Jiang, W.; DeYonker, N. J.; Wilson, A. K. Multireference character for 3d transition-metal-containing molecules. J. Chem. Theory Comput. 2012, 8, 460468,  DOI: 10.1021/ct2006852
  110. 110
    Lee, T. J.; Rice, J. E.; Scuseria, G. E.; Schaefer, H. F. Theoretical investigations of molecules composed only of fluorine, oxygen and nitrogen: determination of the equilibrium structures of FOOF, (NO)2 and FNNF and the transition state structure for FNNF cis-trans isomerization. Theor. Chim. Acta 1989, 75, 8198,  DOI: 10.1007/BF00527711
  111. 111
    Liakos, D. G.; Neese, F. Interplay of Correlation and Relativistic Effects in Correlated Calculations on Transition-Metal Complexes: The (Cu2O2)2+ Core Revisited. J. Chem. Theory Comput. 2011, 7, 15111523,  DOI: 10.1021/ct1006949
  112. 112
    Head-Gordon, M. Characterizing unpaired electrons from the one-particle density matrix. Chem. Phys. Lett. 2003, 372, 508511,  DOI: 10.1016/S0009-2614(03)00422-6
  113. 113
    Ramos-Cordoba, E.; Matito, E. Local Descriptors of Dynamic and Nondynamic Correlation. J. Chem. Theory Comput. 2017, 13, 27052711,  DOI: 10.1021/acs.jctc.7b00293
  114. 114
    Löwdin, P.-O. Quantum theory of many-particle systems. I. Physical interpretations by means of density matrices, natural spin-orbitals, and convergence problems in the method of configurational interaction. Phys. Rev. 1955, 97, 14741489,  DOI: 10.1103/PhysRev.97.1474
  115. 115
    Ramos-Cordoba, E.; Salvador, P.; Matito, E. Separation of dynamic and nondynamic correlation. Phys. Chem. Chem. Phys. 2016, 18, 2401524023,  DOI: 10.1039/C6CP03072F
  116. 116
    Juhász, T.; Mazziotti, D. A. The cumulant two-particle reduced density matrix as a measure of electron correlation and entanglement. J. Chem. Phys. 2006, 125, 174105,  DOI: 10.1063/1.2378768
  117. 117
    Crittenden, D. L. A Hierarchy of Static Correlation Models. J. Phys. Chem. A 2013, 117, 38523860,  DOI: 10.1021/jp400669p
  118. 118
    Pauncz, R. Spin Eigenfunctions: Construction and Use; Plenum Press: New York, 1979.
  119. 119
    Perdew, J. P.; Ruzsinszky, A.; Sun, J.; Nepal, N. K.; Kaplan, A. D. Interpretations of ground-state symmetry breaking and strong correlation in wavefunction and density functional theories. Proc. Nat. Acad. Sci. 2021, 118, e2017850118,  DOI: 10.1073/pnas.2017850118
  120. 120
    Perdew, J. P.; Ruzsinszky, A.; Constantin, L. A.; Sun, J.; Csonka, G. I. Some fundamental issues in ground-state density functional theory: A guide for the perplexed. J. Chem. Theory Comput. 2009, 5, 902908,  DOI: 10.1021/ct800531s
  121. 121
    Filatov, M.; Shaik, S. Spin-restricted density functional approach to the open-shell problem. Chem. Phys. Lett. 1998, 288, 689697,  DOI: 10.1016/S0009-2614(98)00364-9
  122. 122
    Gagliardi, L.; Truhlar, D. G.; Li Manni, G.; Carlson, R. K.; Hoyer, C. E.; Bao, J. L. Multiconfiguration pair-density functional theory: A new way to treat strongly correlated systems. Acc. Chem. Res. 2017, 50, 6673,  DOI: 10.1021/acs.accounts.6b00471
  123. 123
    Cremer, D. Density functional theory: coverage of dynamic and non-dynamic electron correlation effects. Mol. Phys. 2001, 99, 18991940,  DOI: 10.1080/00268970110083564
  124. 124
    Ziegler, T.; Rauk, A.; Baerends, E. J. On the calculation of multiplet energies by the Hartree-Fock-Slater method. Theor. Chim. Acta 1977, 43, 261271,  DOI: 10.1007/BF00551551
  125. 125
    Noodleman, L. Valence bond description of antiferromagnetic coupling in transition metal dimers. J. Chem. Phys. 1981, 74, 57375743,  DOI: 10.1063/1.440939
  126. 126
    Daul, C. Density functional theory applied to the excited states of coordination compounds. Int. J. Quantum Chem. 1994, 52, 867877,  DOI: 10.1002/qua.560520414
  127. 127
    Neese, F. Definition of corresponding orbitals and the diradical character in broken symmetry DFT calculations on spin coupled systems. J. Phys. Chem. Solids 2004, 65, 781785,  DOI: 10.1016/j.jpcs.2003.11.015
  128. 128
    Gunnarsson, O.; Lundqvist, B. I. Exchange and correlation in atoms, molecules, and solids by the spin-density-functional formalism. Phys. Rev. B 1976, 13, 42744298,  DOI: 10.1103/PhysRevB.13.4274
  129. 129
    Anderson, P. W. More is different: broken symmetry and the nature of the hierarchical structure of science. Science 1972, 177, 393396,  DOI: 10.1126/science.177.4047.393
  130. 130
    Amos, A.; Hall, G. Single determinant wave functions. Proc. R. Soc. A 1961, 263, 483493,  DOI: 10.1098/rspa.1961.0175
  131. 131
    Ladner, R. C.; Goddard III, W. A. Improved Quantum Theory of Many-Electron Systems. V. The Spin-Coupling Optimized GI Method. J. Chem. Phys. 1969, 51, 10731087,  DOI: 10.1063/1.1672106
  132. 132
    Bobrowicz, F. W.; Goddard, W. A. The Self-Consistent Field Equations for Generalized Valence Bond and Open-Shell Hartree─Fock Wave Functions. In Methods of Electronic Structure Theory; Springer: New York, 1977; Chapter 4, pp 79127.
  133. 133
    Ghosh, P.; Bill, E.; Weyhermüller, T.; Neese, F.; Wieghardt, K. Noninnocence of the Ligand Glyoxal-bis (2-mercaptoanil). The Electronic Structures of [Fe(gma)]2,[Fe(gma)(py)].py,[Fe(gma)(CN)]1–/0,[Fe(gma)I], and [Fe(gma)(PR3)n] (n= 1, 2). Experimental and Theoretical Evidence for “Excited State” Coordination. J. Am. Chem. Soc. 2003, 125, 12931308,  DOI: 10.1021/ja021123h
  134. 134
    Herebian, D.; Wieghardt, K. E.; Neese, F. Analysis and interpretation of metal-radical coupling in a series of square planar nickel complexes: correlated ab initio and density functional investigation of [Ni(LISQ)2](LISQ= 3,5-di-tert-butyl-o-diiminobenzosemiquinonate(1-)). J. Am. Chem. Soc. 2003, 125, 1099711005,  DOI: 10.1021/ja030124m
  135. 135
    Chłopek, K.; Muresan, N.; Neese, F.; Wieghardt, K. Electronic Structures of Five-Coordinate Complexes of Iron Containing Zero, One, or Two π-Radical Ligands: A Broken-Symmetry Density Functional Theoretical Study. Chem. Eur. J. 2007, 13, 83908403,  DOI: 10.1002/chem.200700897
  136. 136
    Neese, F. Prediction of molecular properties and molecular spectroscopy with density functional theory: From fundamental theory to exchange-coupling. Coord. Chem. Rev. 2009, 253, 526563,  DOI: 10.1016/j.ccr.2008.05.014
  137. 137
    Hohenberg, P.; Kohn, W. Inhomogeneous electron gas. Phys. Rev. 1964, 136, B864B871,  DOI: 10.1103/PhysRev.136.B864
  138. 138
    Kutzelnigg, W. Density functional theory in terms of a Legendre transformation for beginners. J. Mol. Struct.: THEOCHEM 2006, 768, 163173,  DOI: 10.1016/j.theochem.2006.05.012
  139. 139
    Lieb, E. H. Density functionals for Coulomb systems. Int. J. Quantum Chem. 1983, 24, 243277,  DOI: 10.1002/qua.560240302
  140. 140
    Penz, M.; Tellgren, E. I.; Csirik, M. A.; Ruggenthaler, M.; Laestadius, A. structure of the density-potential mapping. Part I: Standard density-functional theory. arXiv:2211.16627 2022, DOI: 10.48550/arXiv.2211.16627 .
  141. 141
    Casida, M. E.; Huix-Rotllant, M. Progress in time-dependent density-functional theory. Annu. Rev. Phys. Chem. 2012, 63, 287323,  DOI: 10.1146/annurev-physchem-032511-143803
  142. 142
    Kohn, W. Density Functional and Density Matrix Method Scaling Linearly with the Number of Atoms. Phys. Rev. Lett. 1996, 76, 31683171,  DOI: 10.1103/PhysRevLett.76.3168
  143. 143
    Prodan, E.; Kohn, W. Nearsightedness of electronic matter. Proc. Nat. Acad. Sci. 2005, 102, 1163511638,  DOI: 10.1073/pnas.0505436102
  144. 144
    Li, L.; Burke, K. Recent developments in density functional approximations. In Handbook of Materials Modeling: Methods: Theory and Modeling; Springer: Cham, 2020; pp 213226.
  145. 145
    Mayer, J. E. Electron Correlation. Phys. Rev. 1955, 100, 15791586,  DOI: 10.1103/PhysRev.100.1579
  146. 146
    Bopp, F. Ableitung der Bindungsenergie vonN-Teilchen-Systemen aus 2-Teilchen-Dichtematrizen. Z. Phys. 1959, 156, 348359,  DOI: 10.1007/BF01461233
  147. 147
    Mazziotti, D. A. Structure of Fermionic Density Matrices: Complete N-Representability Conditions. Phys. Rev. Lett. 2012, 108, 263002,  DOI: 10.1103/PhysRevLett.108.263002
  148. 148
    Gilbert, T. L. Hohenberg-Kohn theorem for nonlocal external potentials. Phys. Rev. B 1975, 12, 21112120,  DOI: 10.1103/PhysRevB.12.2111
  149. 149
    Müller, A. Explicit approximate relation between reduced two- and one-particle density matrices. Phys. Lett. A 1984, 105, 446452,  DOI: 10.1016/0375-9601(84)91034-X
  150. 150
    Kutzelnigg, W. Density-cumulant functional theory. J. Chem. Phys. 2006, 125, 171101,  DOI: 10.1063/1.2387955
  151. 151
    Piris, M.; Ugalde, J. M. Perspective on natural orbital functional theory. Int. J. Quantum Chem. 2014, 114, 11691175,  DOI: 10.1002/qua.24663
  152. 152
    Pernal, K., Giesbertz, K. J. H. Reduced Density Matrix Functional Theory (RDMFT) and Linear Response Time-Dependent RDMFT (TD-RDMFT). In Density-Functional Methods for Excited States; Springer: Cham, 2016; pp 125183.
  153. 153
    Coleman, A. J. Structure of fermion density matrices. Rev. Mod. Phys. 1963, 35, 668686,  DOI: 10.1103/RevModPhys.35.668
  154. 154
    Cioslowski, J.; Pernal, K.; Buchowiecki, M. Approximate one-matrix functionals for the electron–electron repulsion energy from geminal theories. J. Chem. Phys. 2003, 119, 64436447,  DOI: 10.1063/1.1604375
  155. 155
    Piris, M. A natural orbital functional based on an explicit approach of the two-electron cumulant. Int. J. Quantum Chem. 2013, 113, 620630,  DOI: 10.1002/qua.24020
  156. 156
    Simmonett, A. C.; Wilke, J. J.; Schaefer, H. F., III; Kutzelnigg, W. Density cumulant functional theory: First implementation and benchmark results for the DCFT-06 model. J. Chem. Phys. 2010, 133, 174122,  DOI: 10.1063/1.3503657
  157. 157
    Kutzelnigg, W. Density Functional Theory (DFT) and ab-initio Quantum Chemistry (AIQC). Story of a difficult partnership. In Trends and Perspectives in Modern Computational Science; Brill: Leiden, 2006; pp 2362.
  158. 158
    Copan, A. V.; Sokolov, A. Y. Linear-response density cumulant theory for excited electronic states. J. Chem. Theory Comput. 2018, 14, 40974108,  DOI: 10.1021/acs.jctc.8b00326
  159. 159
    Cioslowski, J., Ed. Many-electron densities and reduced density matrices; Springer Science & Business Media: New York, 2000.
  160. 160
    Gidofalvi, G.; Mazziotti, D. A. Active-space two-electron reduced-density-matrix method: Complete active-space calculations without diagonalization of the N-electron Hamiltonian. J. Chem. Phys. 2008, 129, 134108,  DOI: 10.1063/1.2983652
  161. 161
    Mullinax, J. W.; Sokolov, A. Y.; Schaefer, H. F., III Can density cumulant functional theory describe static correlation effects?. J. Chem. Theory Comput. 2015, 11, 24872495,  DOI: 10.1021/acs.jctc.5b00346
  162. 162
    Lathiotakis, N. N.; Helbig, N.; Rubio, A.; Gidopoulos, N. I. Local reduced-density-matrix-functional theory: Incorporating static correlation effects in Kohn-Sham equations. Phys. Rev. A 2014, 90, 032511,  DOI: 10.1103/PhysRevA.90.032511
  163. 163
    Piris, M. Global Method for Electron Correlation. Phys. Rev. Lett. 2017, 119, 063002,  DOI: 10.1103/PhysRevLett.119.063002
  164. 164
    Gibney, D.; Boyn, J.-N.; Mazziotti, D. A. Density Functional Theory Transformed into a One-Electron Reduced-Density-Matrix Functional Theory for the Capture of Static Correlation. J. Phys. Chem. Lett. 2022, 13, 13821388,  DOI: 10.1021/acs.jpclett.2c00083
  165. 165
    Martin, P. C.; Schwinger, J. Theory of many-particle systems. I. Phys. Rev. 1959, 115, 13421373,  DOI: 10.1103/PhysRev.115.1342
  166. 166
    Luttinger, J. M.; Ward, J. C. Ground-state energy of a many-fermion system. II. Phys. Rev. 1960, 118, 14171427,  DOI: 10.1103/PhysRev.118.1417
  167. 167
    Baym, G.; Kadanoff, L. P. Conservation laws and correlation functions. Phys. Rev. 1961, 124, 287299,  DOI: 10.1103/PhysRev.124.287
  168. 168
    Ziesche, P. Cumulant expansions of reduced densities, reduced density matrices, and Green’s functions. In Many-Electron Densities and Reduced Density Matrices; Springer: New York, 2000; Chapter 3, pp 3356.
  169. 169
    Sun, Q.; Chan, G. K.-L. Quantum embedding theories. Acc. Chem. Res. 2016, 49, 27052712,  DOI: 10.1021/acs.accounts.6b00356
  170. 170
    Kotliar, G.; Savrasov, S. Y.; Haule, K.; Oudovenko, V. S.; Parcollet, O.; Marianetti, C. Electronic structure calculations with dynamical mean-field theory. Rev. Mod. Phys. 2006, 78, 865951,  DOI: 10.1103/RevModPhys.78.865
  171. 171
    Lin, N.; Marianetti, C.; Millis, A. J.; Reichman, D. R. Dynamical mean-field theory for quantum chemistry. Phys. Rev. Lett. 2011, 106, 096402,  DOI: 10.1103/PhysRevLett.106.096402
  172. 172
    Zgid, D.; Chan, G. K.-L. Dynamical mean-field theory from a quantum chemical perspective. J. Chem. Phys. 2011, 134, 094115,  DOI: 10.1063/1.3556707
  173. 173
    Hedin, L. New method for calculating the one-particle Green’s function with application to the electron-gas problem. Phys. Rev. 1965, 139, A796A823,  DOI: 10.1103/PhysRev.139.A796
  174. 174
    Aryasetiawan, F.; Gunnarsson, O. The GW method. Rep. Prog. Phys. 1998, 61, 237312,  DOI: 10.1088/0034-4885/61/3/002
  175. 175
    Reining, L. The GW approximation: content, successes and limitations. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2018, 8, e1344,  DOI: 10.1002/wcms.1344
  176. 176
    Phillips, J. J.; Zgid, D. Communication: The description of strong correlation within self-consistent Green’s function second-order perturbation theory. J. Chem. Phys. 2014, 140, 241101,  DOI: 10.1063/1.4884951
  177. 177
    Pokhilko, P.; Zgid, D. Interpretation of multiple solutions in fully iterative GF2 and GW schemes using local analysis of two-particle density matrices. J. Chem. Phys. 2021, 155, 024101,  DOI: 10.1063/5.0055191
  178. 178
    Blase, X.; Duchemin, I.; Jacquemin, D. The Bethe–Salpeter equation in chemistry: relations with TD-DFT, applications and challenges. Chem. Soc. Rev. 2018, 47, 10221043,  DOI: 10.1039/C7CS00049A
  179. 179
    Pokhilko, P.; Iskakov, S.; Yeh, C.-N.; Zgid, D. Evaluation of two-particle properties within finite-temperature self-consistent one-particle Green’s function methods: Theory and application to GW and GF2. J. Chem. Phys. 2021, 155, 024119,  DOI: 10.1063/5.0054661
  180. 180
    Cohen, A. J.; Mori-Sánchez, P.; Yang, W. Fractional spins and static correlation error in density functional theory. J. Chem. Phys. 2008, 129, 121104,  DOI: 10.1063/1.2987202
  181. 181
    Cohen, A. J.; Mori-Sánchez, P.; Yang, W. Insights into current limitations of density functional theory. Science 2008, 321, 792794,  DOI: 10.1126/science.1158722
  182. 182
    Grimme, S.; Hansen, A. A practicable real-space measure and visualization of static electron-correlation effects. Angew. Chem., Int. Ed. 2015, 54, 1230812313,  DOI: 10.1002/anie.201501887
  183. 183
    Muechler, L.; Badrtdinov, D. I.; Hampel, A.; Cano, J.; Rösner, M.; Dreyer, C. E. Quantum embedding methods for correlated excited states of point defects: Case studies and challenges. Phys. Rev. B 2022, 105, 235104,  DOI: 10.1103/PhysRevB.105.235104
  184. 184
    Nielsen, M. A.; Chuang, I. L. Quantum Computation and Quantum Information; Cambridge University Press: Cambridge, 2010.
  185. 185
    Almeida, M.; Omar, Y.; Rocha Vieira, V. Introduction to entanglement and applications to the simulation of many-body quantum systems. In Strongly Correlated Systems, Coherence And Entanglement; World Scientific: Singapore, 2007; Chapter 19, pp 525547.
  186. 186
    Ding, L.; Knecht, S.; Zimborás, Z.; Schilling, C. Quantum Correlations in Molecules: from Quantum Resourcing to Chemical Bonding. Quantum Sci. Technol. 2023, 8, 015015,  DOI: 10.1088/2058-9565/aca4ee
  187. 187
    Omar, Y. Particle statistics in quantum information processing. Int. J. Quantum Inf. 2005, 3, 201205,  DOI: 10.1142/S021974990500075X
  188. 188
    Benatti, F.; Floreanini, R.; Franchini, F.; Marzolino, U. Entanglement in indistinguishable particle systems. Phys. Rep. 2020, 878, 127,  DOI: 10.1016/j.physrep.2020.07.003
  189. 189
    Henderson, T. M.; Bulik, I. W.; Stein, T.; Scuseria, G. E. Seniority-based coupled cluster theory. J. Chem. Phys. 2014, 141, 244104,  DOI: 10.1063/1.4904384
  190. 190
    Boguslawski, K.; Tecmer, P. Orbital entanglement in quantum chemistry. Int. J. Quantum Chem. 2015, 115, 12891295,  DOI: 10.1002/qua.24832
  191. 191
    Ding, L.; Zimboras, Z.; Schilling, C. Quantifying Electron Entanglement Faithfully. arXiv:2207.03377 2022, DOI: 10.48550/arXiv.2207.03377 .
  192. 192
    Legeza, Ö.; Sólyom, J. Optimizing the density-matrix renormalization group method using quantum information entropy. Phys. Rev. B 2003, 68, 195116,  DOI: 10.1103/PhysRevB.68.195116
  193. 193
    Stein, C. J.; Reiher, M. Measuring multi-configurational character by orbital entanglement. Mol. Phys. 2017, 115, 21102119,  DOI: 10.1080/00268976.2017.1288934
  194. 194
    Stein, C. J.; Reiher, M. Automated selection of active orbital spaces. J. Chem. Theory Comput. 2016, 12, 17601771,  DOI: 10.1021/acs.jctc.6b00156
  195. 195
    Boguslawski, K.; Tecmer, P.; Barcza, G.; Legeza, O.; Reiher, M. Orbital entanglement in bond-formation processes. J. Chem. Theory Comput. 2013, 9, 29592973,  DOI: 10.1021/ct400247p
  196. 196
    Huang, Z.; Kais, S. Entanglement as measure of electron–electron correlation in quantum chemistry calculations. Chem. Phys. Lett. 2005, 413, 15,  DOI: 10.1016/j.cplett.2005.07.045
  197. 197
    Boguslawski, K.; Tecmer, P.; Legeza, O.; Reiher, M. Entanglement measures for single-and multireference correlation effects. J. Phys. Chem. Lett. 2012, 3, 31293135,  DOI: 10.1021/jz301319v
  198. 198
    Ziesche, P. Correlation strength and information entropy. Int. J. Quantum Chem. 1995, 56, 363369,  DOI: 10.1002/qua.560560422
  199. 199
    Ghosh, K. J.; Kais, S.; Herschbach, D. R. Geometrical picture of the electron–electron correlation at the large-D limit. Phys. Chem. Chem. Phys. 2022, 24, 92989307,  DOI: 10.1039/D2CP00438K
  200. 200
    Rissler, J.; Noack, R. M.; White, S. R. Measuring orbital interaction using quantum information theory. Chem. Phys. 2006, 323, 519531,  DOI: 10.1016/j.chemphys.2005.10.018
  201. 201
    Cao, Y.; Romero, J.; Olson, J. P.; Degroote, M.; Johnson, P. D.; Kieferová, M.; Kivlichan, I. D.; Menke, T.; Peropadre, B.; Sawaya, N. P. D.; Sim, S.; Veis, L.; Aspuru-Guzik, A. Quantum Chemistry in the Age of Quantum Computing. Chem. Rev. 2019, 119, 1085610915,  DOI: 10.1021/acs.chemrev.8b00803
  202. 202
    Reiher, M.; Wiebe, N.; Svore, K. M.; Wecker, D.; Troyer, M. Elucidating reaction mechanisms on quantum computers. Proc. Nat. Acad. Sci. 2017, 114, 75557560,  DOI: 10.1073/pnas.1619152114
  203. 203
    Tubman, N. M.; Mejuto-Zaera, C.; Epstein, J. M.; Hait, D.; Levine, D. S.; Huggins, W.; Jiang, Z.; McClean, J. R.; Babbush, R.; Head-Gordon, M.; Whaley, K. B. Postponing the orthogonality catastrophe: efficient state preparation for electronic structure simulations on quantum devices. arXiv:1809.05523 2018, DOI: 10.48550/arXiv.1809.05523 .
  204. 204
    Carbone, A.; Galli, D. E.; Motta, M.; Jones, B. Quantum circuits for the preparation of spin eigenfunctions on quantum computers. Symmetry 2022, 14, 624,  DOI: 10.3390/sym14030624
  205. 205
    Lacroix, D.; Ruiz Guzman, E. A.; Siwach, P. Symmetry breaking/symmetry preserving circuits and symmetry restoration on quantum computers. Eur. Phys. J. A 2023, 59, 3,  DOI: 10.1140/epja/s10050-022-00911-7
  206. 206
    Lee, S.; Lee, J.; Zhai, H.; Tong, Y.; Dalzell, A. M.; Kumar, A.; Helms, P.; Gray, J.; Cui, Z.-H.; Liu, W.; Kastoryano, M.; Babbush, R.; Preskill, J.; Reichman, D. R.; Campbell, E. T.; Valeev, E. F.; Lin, L.; Chan, G. K.-L. Is there evidence for exponential quantum advantage in quantum chemistry. arXiv:2208.02199 2022, DOI: 10.48550/arXiv.2208.02199 .
  207. 207
    Tazhigulov, R. N.; Sun, S.-N.; Haghshenas, R.; Zhai, H.; Tan, A. T.; Rubin, N. C.; Babbush, R.; Minnich, A. J.; Chan, G. K.-L. Simulating models of challenging correlated molecules and materials on the Sycamore quantum processor. PRX Quantum 2022, 3, 040318,  DOI: 10.1103/PRXQuantum.3.040318
  208. 208
    Amsler, M.; Deglmann, P.; Degroote, M.; Kaicher, M. P.; Kiser, M.; Kühn, M.; Kumar, C.; Maier, A.; Samsonidze, G.; Schroeder, A.; Streif, M.; Vodola, D.; Wever, C. Quantum-enhanced quantum Monte Carlo: an industrial view. arXiv:2301.11838 2023, DOI: 10.48550/arXiv.2301.11838 .
  209. 209
    Rubin, N. C.; Berry, D. W.; Malone, F. D.; White, A. F.; Khattar, T.; Sicolo, A. E. D.; Kühn, S.; Kaicher, M.; Lee, M.; Babbush, J. Fault-tolerant quantum simulation of materials using Bloch orbitals. arXiv:2302.05531 2023, DOI: 10.48550/arXiv.2302.05531 .
  210. 210
    Lykos, P.; Pratt, G. W. Discussion on The Hartree-Fock Approximation. Rev. Mod. Phys. 1963, 35, 496501,  DOI: 10.1103/RevModPhys.35.496
  211. 211
    Slater, J. C. The Theory of Complex Spectra. Phys. Rev. 1929, 34, 12931322,  DOI: 10.1103/PhysRev.34.1293
  212. 212
    Slater, J. C. Solid-state and molecular theory: a scientific biography; Wiley-Interscience: New York, 1975.
  213. 213
    Matsen, F. Spin-free quantum chemistry. In Adv. Quantum Chem.; Interscience: New York, 1964; Vol. 1, pp 59114.
  214. 214
    Paldus, J. The Unitary Group for the Evaluation of Electronic Energy Matrix Elements. In The Unitary Group for the Evaluation of Electronic Energy Matrix Elements; Springer: Berlin, 1981; pp 150.
  215. 215
    Shavitt, I. The Graphical Unitary Group Approach and its Application to Direct Configuration Interaction Calculations. In The Unitary Group for the Evaluation of Electronic Energy Matrix Elements; Springer: Berlin, 1981; pp 5199.
  216. 216
    Dobrautz, W.; Smart, S. D.; Alavi, A. Efficient formulation of full configuration interaction quantum Monte Carlo in a spin eigenbasis via the graphical unitary group approach. J. Chem. Phys. 2019, 151, 094104,  DOI: 10.1063/1.5108908
  217. 217
    McLean, A. D.; Lengsfield, B. H.; Pacansky, J.; Ellinger, Y. Symmetry breaking in molecular calculations and the reliable prediction of equilibrium geometries. The formyloxyl radical as an example. J. Chem. Phys. 1985, 83, 35673576,  DOI: 10.1063/1.449162
  218. 218
    Ayala, P. Y.; Schlegel, H. B. A nonorthogonal CI treatment of symmetry breaking in sigma formyloxyl radical. J. Chem. Phys. 1998, 108, 75607567,  DOI: 10.1063/1.476190
  219. 219
    Manne, R. Brillouin’s theorem in Roothaan’s open-shell SCF method. Mol. Phys. 1972, 24, 935944,  DOI: 10.1080/00268977200102061
  220. 220
    Davidson, E. R.; Borden, W. T. Symmetry breaking in polyatomic molecules: real and artifactual. J. Phys. Chem. 1983, 87, 47834790,  DOI: 10.1021/j150642a005
  221. 221
    Čížek, J.; Paldus, J. Stability Conditions for the Solutions of the Hartree–Fock Equations for Atomic and Molecular Systems. Application to the Pi-Electron Model of Cyclic Polyenes. J. Chem. Phys. 1967, 47, 39763985,  DOI: 10.1063/1.1701562
  222. 222
    Čížek, J.; Paldus, J. Stability Conditions for the Solutions of the Hartree–Fock Equations for Atomic and Molecular Systems. III. Rules for the Singlet Stability of Hartree–Fock Solutions of π-Electronic Systems. J. Chem. Phys. 1970, 53, 821829,  DOI: 10.1063/1.1674065
  223. 223
    Paldus, J.; Čížek, J. Stability Conditions for the Solutions of the Hartree-Fock Equations for Atomic and Molecular Systems. VI. Singlet-Type Instabilities and Charge-Density-Wave Hartree-Fock Solutions for Cyclic Polyenes. Phys. Rev. A 1970, 2, 22682283,  DOI: 10.1103/PhysRevA.2.2268
  224. 224
    Davidson, E. R. Spin-restricted open-shell self-consistent-field theory. Chem. Phys. Lett. 1973, 21, 565567,  DOI: 10.1016/0009-2614(73)80309-4
  225. 225
    Edwards, W. D.; Zerner, M. C. A generalized restricted open-shell Fock operator. Theor. Chim. Acta 1987, 72, 347361,  DOI: 10.1007/BF01192227
  226. 226
    Löwdin, P.-O. Quantum Theory of Many-Particle Systems. III. Extension of the Hartree-Fock Scheme to Include Degenerate Systems and Correlation Effects. Phys. Rev. 1955, 97, 15091520,  DOI: 10.1103/PhysRev.97.1509
  227. 227
    Mayer, I. The spin-projected extended Hartree-Fock method. In Adv. Quantum Chem.; Academic Press: New York, 1980; Vol. 12, pp 189262.
  228. 228
    Tsuchimochi, T.; Scuseria, G. E. Communication: ROHF theory made simple. J. Chem. Phys. 2010, 133, 141102,  DOI: 10.1063/1.3503173
  229. 229
    Rittby, M.; Bartlett, R. J. An open-shell spin-restricted coupled cluster method: application to ionization potentials in nitrogen. J. Phys. Chem. 1988, 92, 30333036,  DOI: 10.1021/j100322a004
  230. 230
    Neese, F. Importance of direct spin-spin coupling and spin-flip excitations for the zero-field splittings of transition metal complexes: A case study. J. Am. Chem. Soc. 2006, 128, 1021310222,  DOI: 10.1021/ja061798a
  231. 231
    Casanova, D.; Krylov, A. I. Spin-flip methods in quantum chemistry. Phys. Chem. Chem. Phys. 2020, 22, 43264342,  DOI: 10.1039/C9CP06507E
  232. 232
    Roemelt, M.; Neese, F. Excited states of large open-shell molecules: an efficient, general, and spin-adapted approach based on a restricted open-shell ground state wave function. J. Phys. Chem. A 2013, 117, 30693083,  DOI: 10.1021/jp3126126
  233. 233
    Izsák, R. Second quantisation for unrestricted references: formalism and quasi-spin-adaptation of excitation and spin-flip operators. Mol. Phys. 2022, e2126802,  DOI: 10.1080/00268976.2022.2126802
  234. 234
    Veis, L.; Antalik, A.; Legeza, O.; Alavi, A.; Pittner, J. The intricate case of tetramethyleneethane: A full configuration interaction quantum Monte Carlo benchmark and multireference coupled cluster studies. J. Chem. Theory Comput. 2018, 14, 24392445,  DOI: 10.1021/acs.jctc.8b00022
  235. 235
    Sears, J. S.; Sherrill, C. D. Assessing the Performance of Density Functional Theory for the Electronic Structure of Metal-Salens: The 3d0-Metals. J. Phys. Chem. A 2008, 112, 34663477,  DOI: 10.1021/jp711595w
  236. 236
    Sears, J. S.; Sherrill, C. D. Assessing the performance of Density Functional Theory for the electronic structure of metal-salens: the d2-metals. J. Phys. Chem. A 2008, 112, 67416752,  DOI: 10.1021/jp802249n
  237. 237
    Olivares-Amaya, R.; Hu, W.; Nakatani, N.; Sharma, S.; Yang, J.; Chan, G. K.-L. The ab-initio density matrix renormalization group in practice. J. Chem. Phys. 2015, 142, 034102,  DOI: 10.1063/1.4905329
  238. 238
    Pandharkar, R.; Hermes, M. R.; Cramer, C. J.; Gagliardi, L. Localized Active Space-State Interaction: a Multireference Method for Chemical Insight. J. Chem. Theory Comput. 2022, 18, 65576566,  DOI: 10.1021/acs.jctc.2c00536
  239. 239
    Máté, M.; Petrov, K.; Szalay, S.; Legeza, Ö. Compressing multireference character of wave functions via fermionic mode optimization. J. Math. Chem. 2023, 61, 362375,  DOI: 10.1007/s10910-022-01379-y
  240. 240
    Smith, J. E. T.; Mussard, B.; Holmes, A. A.; Sharma, S. Cheap and Near Exact CASSCF with Large Active Spaces. J. Chem. Theory Comput. 2017, 13, 54685478,  DOI: 10.1021/acs.jctc.7b00900
  241. 241
    Dobrautz, W.; Weser, O.; Bogdanov, N. A.; Alavi, A.; Li Manni, G. Spin-Pure Stochastic-CASSCF via GUGA-FCIQMC Applied to Iron–Sulfur Clusters. J. Chem. Theory Comput. 2021, 17, 56845703,  DOI: 10.1021/acs.jctc.1c00589
  242. 242
    Dobrautz, W.; Katukuri, V. M.; Bogdanov, N. A.; Kats, D.; Li Manni, G.; Alavi, A. Combined unitary and symmetric group approach applied to low-dimensional Heisenberg spin systems. Phys. Rev. B 2022, 105, 195123,  DOI: 10.1103/PhysRevB.105.195123
  243. 243
    Li Manni, G.; Kats, D.; Liebermann, N. Resolution of Electronic States in Heisenberg Cluster Models within the Unitary Group Approach. ChemRxiv 2022, DOI: 10.26434/chemrxiv-2022-rfmhk-v2 .
  244. 244
    Pipek, J.; Mezey, P. G. A fast intrinsic localization procedure applicable for abinitio and semiempirical linear combination of atomic orbital wave functions. J. Chem. Phys. 1989, 90, 49164926,  DOI: 10.1063/1.456588
  245. 245
    Malrieu, J.-P.; Guihéry, N.; Calzado, C. J.; Angeli, C. Bond electron pair: Its relevance and analysis from the quantum chemistry point of view. J. Comput. Chem. 2007, 28, 3550,  DOI: 10.1002/jcc.20546
  246. 246
    Lewis, G. N. The atom and the molecule. J. Am. Chem. Soc. 1916, 38, 762785,  DOI: 10.1021/ja02261a002
  247. 247
    Chen, H.; Lai, W.; Shaik, S. Multireference and Multiconfiguration Ab Initio Methods in Heme-Related Systems: What Have We Learned So Far?. J. Phys. Chem. B 2011, 115, 17271742,  DOI: 10.1021/jp110016u
  248. 248
    Li, Z.; Guo, S.; Sun, Q.; Chan, G. K.-L. Electronic landscape of the P-cluster of nitrogenase as revealed through many-electron quantum wavefunction simulations. Nat. Chem. 2019, 11, 10261033,  DOI: 10.1038/s41557-019-0337-3
  249. 249
    Khedkar, A.; Roemelt, M. Modern multireference methods and their application in transition metal chemistry. Phys. Chem. Chem. Phys. 2021, 23, 1709717112,  DOI: 10.1039/D1CP02640B
  250. 250
    Tarrago, M.; Römelt, C.; Nehrkorn, J.; Schnegg, A.; Neese, F.; Bill, E.; Ye, S. Experimental and Theoretical Evidence for an Unusual Almost Triply Degenerate Electronic Ground State of Ferrous Tetraphenylporphyrin. Inorg. Chem. 2021, 60, 49664985,  DOI: 10.1021/acs.inorgchem.1c00031
  251. 251
    Han, R.; Luber, S.; Li Manni, G. Magnetic Interactions in a [Co(II)3Er(III)(OR)4] Model Cubane Through Forefront Multiconfigurational Methods. ChemRxiv 2023, DOI: 10.26434/chemrxiv-2023-xd0wv .

Cited By

ARTICLE SECTIONS
Jump To

This article has not yet been cited by other publications.

  • Abstract

    Figure 1

    Figure 1. Multireference character M as a function of internuclear distance D in the H2 dissociation problem for the FCI/STO-1G ground state. The blue curve corresponds to canonical orbitals (can), the red one to localized orbitals (loc).

  • References

    ARTICLE SECTIONS
    Jump To

    This article references 251 other publications.

    1. 1
      Kurth, S.; Perdew, J. P. Role of the exchange–correlation energy: Nature’s glue. Int. J. Quantum Chem. 2000, 77, 814818,  DOI: 10.1002/(SICI)1097-461X(2000)77:5<814::AID-QUA3>3.0.CO;2-F
    2. 2
      Martin, J. M. Electron Correlation: Nature’s Weird and Wonderful Chemical Glue. Isr. J. Chem. 2022, 62, e202100111,  DOI: 10.1002/ijch.202100111
    3. 3
      Löwdin, P.-O. Correlation Problem in Many-Electron Quantum Mechanics I. Review of Different Approaches and Discussion of Some Current Ideas. In Advances in Chemical Physics; Interscience: New York, 1959; Vol. 2; Chapter 7, pp 207322.
    4. 4
      Kotani, M.; Ohno, K.; Kayama, K. Quantum Mechanics of Electronic Structure of Simple Molecules. In Encyclopedia of Physics/Handbuch der Physik; Springer: Berlin, 1961; Vol. 37.2, Chapter 1, pp 1172.
    5. 5
      Slater, J. C. Quantum Theory of Molecules and Solids. Vol. 1: Electronic Structure of Molecules; McGraw-Hill: New York, 1963.
    6. 6
      Kutzelnigg, W. Einführung in die theoretische Chemie; Wiley-VCH: Weinheim, 2002; Vol. 2.
    7. 7
      Wigner, E. On the interaction of electrons in metals. Phys. Rev. 1934, 46, 1002,  DOI: 10.1103/PhysRev.46.1002
    8. 8
      Wigner, E. Effects of the electron interaction on the energy levels of electrons in metals. Trans. Faraday Soc. 1938, 34, 678685,  DOI: 10.1039/tf9383400678
    9. 9
      Löwdin, P.-O. The historical development of the electron correlation problem. Int. J. Quantum Chem. 1995, 55, 77102,  DOI: 10.1002/qua.560550203
    10. 10
      Sinanoğlu, O. Many-Electron Theory of Atoms and Molecules. Proc. Nat. Acad. Sci. 1961, 47, 12171226,  DOI: 10.1073/pnas.47.8.1217
    11. 11
      Sinanoğlu, O. Many-Electron Theory of Atoms, Molecules and Their Interactions. In Advances in Chemical Physics; John Wiley & Sons, Ltd: London, 1964; Vol. 6, Chapter 7, pp 315412.
    12. 12
      McWeeny, R. The nature of electron correlation in molecules. Int. J. Quantum Chem. 1967, 1, 351359,  DOI: 10.1002/qua.560010641
    13. 13
      Kutzelnigg, W.; Del Re, G.; Berthier, G. Correlation coefficients for electronic wave functions. Phys. Rev. 1968, 172, 4959,  DOI: 10.1103/PhysRev.172.49
    14. 14
      von Herigonte, P. Electron correlation in the seventies. In Struct. Bonding (Berlin); Springer: New York, 1972; Vol. 12, Chapter 1, pp 147.
    15. 15
      Kutzelnigg, W.; von Herigonte, P. Electron correlation at the dawn of the 21st century. In Adv. Quantum Chem.; Academic Press: San Diego, 2000; Vol. 36, pp 185229.
    16. 16
      Kutzelnigg, W. Theory of electron correlation. In Explicitly Correlated Wave Functions in Chemistry and Physics; Springer Science & Business Media: Dordrecht, 2003; Chapter 1, pp 390.
    17. 17
      Bartlett, R. J.; Stanton, J. F. Applications of Post-Hartree─Fock Methods: A Tutorial. In Reviews in Computational Chemistry; VCH Publishers: New York, 1994; Chapter 2, pp 65169.
    18. 18
      Knowles, P. J.; Schütz, M.; Werner, H.-J. Ab Initio Methods for Electron Correlation in Molecules. In Modern Methods and Algorithms of Quantum Chemistry; Grotendorst, J., Ed.; NIC: Jülich, 2000; Vol. 1, pp 69151.
    19. 19
      Tew, D. P.; Klopper, W.; Helgaker, T. Electron correlation: The many-body problem at the heart of chemistry. J. Comput. Chem. 2007, 28, 13071320,  DOI: 10.1002/jcc.20581
    20. 20
      Senatore, G.; March, N. Recent progress in the field of electron correlation. Rev. Mod. Phys. 1994, 66, 445479,  DOI: 10.1103/RevModPhys.66.445
    21. 21
      Loos, P.-F.; Gill, P. M. The uniform electron gas. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2016, 6, 410429,  DOI: 10.1002/wcms.1257
    22. 22
      Chan, G. K.-L. Low entanglement wavefunctions. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2012, 2, 907920,  DOI: 10.1002/wcms.1095
    23. 23
      Fulde, P. Electron correlations in molecules and solids; Springer-Verlag: Berlin, 1995.
    24. 24
      March, N. H. Electron correlation in molecules and condensed phases; Springer Science & Business Media: New York, 1996.
    25. 25
      Wilson, S. Electron Correlation in Molecules; Dover: Mineola, 2007.
    26. 26
      Advances in Chemical Physics; Vol. 14; Lefebvre, R., Moser, C., Eds.; John Wiley & Sons, Ltd: London, 1969 Correlation Effects in Atoms and Molecules.
    27. 27
      Electron Correlation in the Solid State; March, N. H., Ed.; Imperial Collage Press: London, 1999.
    28. 28
      Raimes, S. Many-electron Theory; North-Holland: Amsterdam, 1972.
    29. 29
      Gell-Mann, M.; Brueckner, K. A. Correlation energy of an electron gas at high density. Phys. Rev. 1957, 106, 364368,  DOI: 10.1103/PhysRev.106.364
    30. 30
      Hund, F. Zur Deutung der Molekelspektren. Z. Phys. 1927, 40, 742764,  DOI: 10.1007/BF01400234
    31. 31
      Mulliken, R. S. The assignment of quantum numbers for electrons in molecules. I. Phys. Rev. 1928, 32, 186222,  DOI: 10.1103/PhysRev.32.186
    32. 32
      Lennard-Jones, J. E. The electronic structure of some diatomic molecules. Trans. Faraday Soc. 1929, 25, 668686,  DOI: 10.1039/tf9292500668
    33. 33
      Heitler, W.; London, F. Wechselwirkung neutraler Atome und homöopolare Bindung nach der Quantenmechanik. Z. Phys. 1927, 44, 455472,  DOI: 10.1007/BF01397394
    34. 34
      Pauling, L. The nature of the chemical bond. Application of results obtained from the quantum mechanics and from a theory of paramagnetic susceptibility to the structure of molecules. J. Am. Chem. Soc. 1931, 53, 13671400,  DOI: 10.1021/ja01355a027
    35. 35
      Coulson, C. A.; Fischer, I. Notes on the molecular orbital treatment of the hydrogen molecule. Philos. Mag. 1949, 40, 386393,  DOI: 10.1080/14786444908521726
    36. 36
      Kutzelnigg, W.; Mukherjee, D. Cumulant expansion of the reduced density matrices. J. Chem. Phys. 1999, 110, 28002809,  DOI: 10.1063/1.478189
    37. 37
      Wigner, E.; Seitz, F. On the Constitution of Metallic Sodium. Phys. Rev. 1933, 43, 804810,  DOI: 10.1103/PhysRev.43.804
    38. 38
      Wigner, E.; Seitz, F. On the Constitution of Metallic Sodium. II. Phys. Rev. 1934, 46, 509524,  DOI: 10.1103/PhysRev.46.509
    39. 39
      McWeeny, R. The density matrix in many-electron quantum mechanics I. Generalized product functions. Factorization and physical interpretation of the density matrices. Proc. R. Soc. A 1959, 253, 242259,  DOI: 10.1098/rspa.1959.0191
    40. 40
      Giner, E.; Tenti, L.; Angeli, C.; Malrieu, J.-P. The “Fermi hole” and the correlation introduced by the symmetrization or the anti-symmetrization of the wave function. J. Chem. Phys. 2016, 145, 124114,  DOI: 10.1063/1.4963018
    41. 41
      Kutzelnigg, W.; Mukherjee, D. Normal order and extended Wick theorem for a multiconfiguration reference wave function. J. Chem. Phys. 1997, 107, 432449,  DOI: 10.1063/1.474405
    42. 42
      Evangelista, F. A. Automatic derivation of many-body theories based on general Fermi vacua. J. Chem. Phys. 2022, 157, 064111,  DOI: 10.1063/5.0097858
    43. 43
      Tsuchimochi, T.; Scuseria, G. E. Strong correlations via constrained-pairing mean-field theory. J. Chem. Phys. 2009, 131, 121102,  DOI: 10.1063/1.3237029
    44. 44
      Kutzelnigg, W. Separation of strong (bond-breaking) from weak (dynamical) correlation. Chem. Phys. 2012, 401, 119124,  DOI: 10.1016/j.chemphys.2011.10.020
    45. 45
      Georges, A.; Kotliar, G. Hubbard model in infinite dimensions. Phys. Rev. B 1992, 45, 64796483,  DOI: 10.1103/PhysRevB.45.6479
    46. 46
      Georges, A.; Kotliar, G.; Krauth, W.; Rozenberg, M. J. Dynamical mean-field theory of strongly correlated fermion systems and the limit of infinite dimensions. Rev. Mod. Phys. 1996, 68, 13125,  DOI: 10.1103/RevModPhys.68.13
    47. 47
      Anisimov, V.; Izyumov, Y. Electronic Structure of Strongly Correlated Materials; Springer: Berlin, 2010.
    48. 48
      Kuramoto, Y. Quantum Many-Body Physics; Springer: Tokyo, 2020.
    49. 49
      Foulkes, W. M. C.; Mitas, L.; Needs, R. J.; Rajagopal, G. Quantum Monte Carlo simulations of solids. Rev. Mod. Phys. 2001, 73, 3383,  DOI: 10.1103/RevModPhys.73.33
    50. 50
      Zhang, S.; Malone, F. D.; Morales, M. A. Auxiliary-field quantum Monte Carlo calculations of the structural properties of nickel oxide. J. Chem. Phys. 2018, 149, 164102,  DOI: 10.1063/1.5040900
    51. 51
      Exactly solvable models of strongly correlated electrons; Korepin, V. E., Essler, F. H., Eds.; World Scientific: Singapore, 1994.
    52. 52
      Gross, D. J. The role of symmetry in fundamental physics. Proc. Nat. Acad. Sci. 1996, 93, 1425614259,  DOI: 10.1073/pnas.93.25.14256
    53. 53
      Raimes, S. The wave mechanics of electrons in metals; North-Holland: Amsterdam, 1963.
    54. 54
      Fulde, P. Solids with weak and strong electron correlations. In Electron Correlation in the Solid State; Imperial Collage Press: London, 1999; Chapter 2, pp 47102.
    55. 55
      Coldwell-Horsfall, R. A.; Maradudin, A. A. Zero-Point Energy of an Electron Lattice. J. Math. Phys. 1960, 1, 395404,  DOI: 10.1063/1.1703670
    56. 56
      Bohm, D.; Pines, D. A Collective Description of Electron Interactions. I. Magnetic Interactions. Phys. Rev. 1951, 82, 625634,  DOI: 10.1103/PhysRev.82.625
    57. 57
      Pines, D. Electron Interaction in Metals. In Solid State Physics; Academic Press: New York, 1955; Vol. 1, pp 367450.
    58. 58
      Slater, J. C. The electronic structure of solids. In Encyclopedia of Physics/Handbuch der Physik; Springer: Berlin, 1956; Vol. 19, pp 1136.
    59. 59
      Brueckner, K. A. The Correlation Energy of a Non-Uniform Electron Gas. In Advances in Chemical Physics; John Wiley & Sons, Ltd: London, 1969; Vol. 14, Chapter 7, pp 215236.
    60. 60
      Slater, J. C. The Virial and Molecular Structure. J. Chem. Phys. 1933, 1, 687691,  DOI: 10.1063/1.1749227
    61. 61
      March, N. H. Kinetic and Potential Energies of an Electron Gas. Phys. Rev. 1958, 110, 604605,  DOI: 10.1103/PhysRev.110.604
    62. 62
      Argyres, P. N. Virial Theorem for the Homogeneous Electron Gas. Phys. Rev. 1967, 154, 410413,  DOI: 10.1103/PhysRev.154.410
    63. 63
      Ruedenberg, K. The Physical Nature of the Chemical Bond. Rev. Mod. Phys. 1962, 34, 326376,  DOI: 10.1103/RevModPhys.34.326
    64. 64
      Hubbard, J. Electron correlations in narrow energy bands. Proc. R. Soc. A 1963, 276, 238257,  DOI: 10.1098/rspa.1963.0204
    65. 65
      Janesko, B. G. Strong correlation in surface chemistry. Mol. Simul. 2017, 43, 394405,  DOI: 10.1080/08927022.2016.1261136
    66. 66
      Motta, M.; Ceperley, D. M.; Chan, G. K.-L.; Gomez, J. A.; Gull, E.; Guo, S.; Jiménez- Hoyos, C. A.; Lan, T. N.; Li, J.; Ma, F.; Millis, A. J.; Prokof’ev, N. V.; Ray, U.; Scuseria, G. E.; Sorella, S.; Stoudenmire, E. M.; Sun, Q.; Tupitsyn, I. S.; White, S. R.; Zgid, D.; Zhang, S. Towards the Solution of the Many-Electron Problem in Real Materials: Equation of State of the Hydrogen Chain with State-of-the-Art Many-Body Methods. Phys. Rev. X 2017, 7, 031059,  DOI: 10.1103/PhysRevX.7.031059
    67. 67
      Motta, M.; Genovese, C.; Ma, F.; Cui, Z.-H.; Sawaya, R.; Chan, G. K.-L.; Chepiga, N.; Helms, P.; Jiménez-Hoyos, C.; Millis, A. J.; Ray, U.; Ronca, E.; Shi, H.; Sorella, S.; Stoudenmire, E. M.; White, S. R.; Zhang, S. Ground-State Properties of the Hydrogen Chain: Dimerization, Insulator-to-Metal Transition, and Magnetic Phases. Phys. Rev. X 2020, 10, 031058
    68. 68
      Sinanoğlu, O.; Tuan, D. F.-t. Many-Electron Theory of Atoms and Molecules. III. Effect of Correlation on Orbitals. J. Chem. Phys. 1963, 38, 17401748,  DOI: 10.1063/1.1776948
    69. 69
      Prendergast, D.; Nolan, M.; Filippi, C.; Fahy, S.; Greer, J. Impact of electron–electron cusp on configuration interaction energies. J. Chem. Phys. 2001, 115, 16261634,  DOI: 10.1063/1.1383585
    70. 70
      Mok, D. K.; Neumann, R.; Handy, N. C. Dynamical and nondynamical correlation. J. Phys. Chem. 1996, 100, 62256230,  DOI: 10.1021/jp9528020
    71. 71
      Hollett, J. W.; Gill, P. M. The two faces of static correlation. J. Chem. Phys. 2011, 134, 114111,  DOI: 10.1063/1.3570574
    72. 72
      Benavides-Riveros, C. L.; Lathiotakis, N. N.; Marques, M. A. Towards a formal definition of static and dynamic electronic correlations. Phys. Chem. Chem. Phys. 2017, 19, 1265512664,  DOI: 10.1039/C7CP01137G
    73. 73
      Bulik, I. W.; Henderson, T. M.; Scuseria, G. E. Can single-reference coupled cluster theory describe static correlation?. J. Chem. Theory Comput. 2015, 11, 31713179,  DOI: 10.1021/acs.jctc.5b00422
    74. 74
      Karton, A.; Rabinovich, E.; Martin, J. M.; Ruscic, B. W4 theory for computational thermochemistry: In pursuit of confident sub-kJ/mol predictions. J. Chem. Phys. 2006, 125, 144108,  DOI: 10.1063/1.2348881
    75. 75
      Hait, D.; Tubman, N. M.; Levine, D. S.; Whaley, K. B.; Head-Gordon, M. What levels of coupled cluster theory are appropriate for transition metal systems? A study using near-exact quantum chemical values for 3d transition metal binary compounds. J. Chem. Theory Comput. 2019, 15, 53705385,  DOI: 10.1021/acs.jctc.9b00674
    76. 76
      Roos, B. O.; Taylor, P. R.; Sigbahn, P. E. A complete active space SCF method (CASSCF) using a density matrix formulated super-CI approach. Chem. Phys. 1980, 48, 157173,  DOI: 10.1016/0301-0104(80)80045-0
    77. 77
      Chan, G. K.-L.; Sharma, S. The density matrix renormalization group in quantum chemistry. Annu. Rev. Phys. Chem. 2011, 62, 465481,  DOI: 10.1146/annurev-physchem-032210-103338
    78. 78
      Mouesca, J.-M. Density functional theory-broken symmetry (DFT-BS) methodology applied to electronic and magnetic properties of bioinorganic prosthetic groups. In Metalloproteins; Springer: New York, 2014; pp 269296.
    79. 79
      Scuseria, G. E.; Jiménez-Hoyos, C. A.; Henderson, T. M.; Samanta, K.; Ellis, J. K. Projected quasiparticle theory for molecular electronic structure. J. Chem. Phys. 2011, 135, 124108,  DOI: 10.1063/1.3643338
    80. 80
      Helgaker, T.; Jorgensen, P.; Olsen, J. Molecular electronic-structure theory; John Wiley & Sons: Chichester, 2000.
    81. 81
      Shavitt, I.; Bartlett, R. J. Many-body methods in chemistry and physics: MBPT and coupled-cluster theory; Cambridge University Press: Cambridge, 2009.
    82. 82
      Čížek, J. On the Correlation Problem in Atomic and Molecular Systems. Calculation of Wavefunction Components in Ursell-Type Expansion Using Quantum-Field Theoretical Methods. J. Chem. Phys. 1966, 45, 42564266,  DOI: 10.1063/1.1727484
    83. 83
      Paldus, J.; Čížek, J.; Shavitt, I. Correlation Problems in Atomic and Molecular Systems. IV. Extended Coupled-Pair Many-Electron Theory and Its Application to the BH3 Molecule. Phys. Rev. A 1972, 5, 5067,  DOI: 10.1103/PhysRevA.5.50
    84. 84
      Arponen, J. Variational principles and linked-cluster exp S expansions for static and dynamic many-body problems. Ann. Phys. 1983, 151, 311382,  DOI: 10.1016/0003-4916(83)90284-1
    85. 85
      Faulstich, F. M.; Laestadius, A.; Legeza, O.; Schneider, R.; Kvaal, S. Analysis of the Tailored Coupled-Cluster Method in Quantum Chemistry. SIAM J. Numer. Anal. 2019, 57, 25792607,  DOI: 10.1137/18M1171436
    86. 86
      Csirik, M. A.; Laestadius, A. Coupled-cluster theory revisited. arXiv:2105.13134 2021, DOI: 10.48550/arXiv.2105.13134 .
    87. 87
      Schütz, M.; Werner, H.-J. Low-order scaling local electron correlation methods. IV. Linear scaling local coupled-cluster (LCCSD). J. Chem. Phys. 2001, 114, 661681,  DOI: 10.1063/1.1330207
    88. 88
      Riplinger, C.; Neese, F. An efficient and near linear scaling pair natural orbital based local coupled cluster method. J. Chem. Phys. 2013, 138, 034106,  DOI: 10.1063/1.4773581
    89. 89
      Shiozaki, T.; Valeev, E. F.; Hirata, S. Explicitly correlated coupled-cluster methods. In Annual Reports in Computational Chemistry; Elsevier: Amsterdam, 2009; Vol. 5, Chapter 6, pp 131148.
    90. 90
      Izsák, R. Single-reference coupled cluster methods for computing excitation energies in large molecules: The efficiency and accuracy of approximations. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2020, 10, e1445,  DOI: 10.1002/wcms.1445
    91. 91
      Lyakh, D. I.; Musiał, M.; Lotrich, V. F.; Bartlett, R. J. Multireference nature of chemistry: The coupled-cluster view. Chem. Rev. 2012, 112, 182243,  DOI: 10.1021/cr2001417
    92. 92
      Huron, B.; Malrieu, J. P.; Rancurel, P. Iterative perturbation calculations of ground and excited state energies from multiconfigurational zeroth-order wavefunctions. J. Chem. Phys. 1973, 58, 57455759,  DOI: 10.1063/1.1679199
    93. 93
      Austin, B. M.; Zubarev, D. Y.; Lester Jr, W. A. Quantum Monte Carlo and related approaches. Chem. Rev. 2012, 112, 263288,  DOI: 10.1021/cr2001564
    94. 94
      Booth, G. H.; Thom, A. J.; Alavi, A. Fermion Monte Carlo without fixed nodes: A game of life, death, and annihilation in Slater determinant space. J. Chem. Phys. 2009, 131, 054106,  DOI: 10.1063/1.3193710
    95. 95
      Petruzielo, F. R.; Holmes, A. A.; Changlani, H. J.; Nightingale, M. P.; Umrigar, C. J. Semistochastic Projector Monte Carlo Method. Phys. Rev. Lett. 2012, 109, 230201,  DOI: 10.1103/PhysRevLett.109.230201
    96. 96
      Spencer, J. S.; Blunt, N. S.; Foulkes, W. M. The sign problem and population dynamics in the full configuration interaction quantum Monte Carlo method. J. Chem. Phys. 2012, 136, 054110,  DOI: 10.1063/1.3681396
    97. 97
      Cleland, D.; Booth, G. H.; Alavi, A. Communications: Survival of the fittest: Accelerating convergence in full configuration-interaction quantum Monte Carlo. J. Chem. Phys. 2010, 132, 041103,  DOI: 10.1063/1.3302277
    98. 98
      Cleland, D. M.; Booth, G. H.; Alavi, A. A study of electron affinities using the initiator approach to full configuration interaction quantum Monte Carlo. J. Chem. Phys. 2011, 134, 024112,  DOI: 10.1063/1.3525712
    99. 99
      Eriksen, J. J. The shape of full configuration interaction to come. J. Phys. Chem. Lett. 2021, 12, 418432,  DOI: 10.1021/acs.jpclett.0c03225
    100. 100
      Eriksen, J. J.; Gauss, J. Many-body expanded full configuration interaction. I. Weakly correlated regime. J. Chem. Theory Comput. 2018, 14, 51805191,  DOI: 10.1021/acs.jctc.8b00680
    101. 101
      Eriksen, J. J.; Gauss, J. Many-body expanded full configuration interaction. II. Strongly correlated regime. J. Chem. Theory Comput. 2019, 15, 48734884,  DOI: 10.1021/acs.jctc.9b00456
    102. 102
      White, S. R. Density matrix formulation for quantum renormalization groups. Phys. Rev. Lett. 1992, 69, 28632866,  DOI: 10.1103/PhysRevLett.69.2863
    103. 103
      Chilkuri, V. G.; Neese, F. Comparison of many-particle representations for selected-CI I: A tree based approach. J. Comput. Chem. 2021, 42, 9821005,  DOI: 10.1002/jcc.26518
    104. 104
      Li Manni, G.; Dobrautz, W.; Alavi, A. Compression of spin-adapted multiconfigurational wave functions in exchange-coupled polynuclear spin systems. J. Chem. Theory Comput. 2020, 16, 22022215,  DOI: 10.1021/acs.jctc.9b01013
    105. 105
      Li Manni, G. Modeling magnetic interactions in high-valent trinuclear [Mn3(IV)O4]4+ complexes through highly compressed multi-configurational wave functions. Phys. Chem. Chem. Phys. 2021, 23, 1976619780,  DOI: 10.1039/D1CP03259C
    106. 106
      Li Manni, G.; Dobrautz, W.; Bogdanov, N. A.; Guther, K.; Alavi, A. Resolution of low-energy states in spin-exchange transition-metal clusters: Case study of singlet states in [Fe(III)4S4] cubanes. J. Phys. Chem. A 2021, 125, 47274740,  DOI: 10.1021/acs.jpca.1c00397
    107. 107
      Shee, J.; Loipersberger, M.; Hait, D.; Lee, J.; Head-Gordon, M. Revealing the nature of electron correlation in transition metal complexes with symmetry breaking and chemical intuition. J. Chem. Phys. 2021, 154, 194109,  DOI: 10.1063/5.0047386
    108. 108
      Benavides-Riveros, C. L.; Lathiotakis, N. N.; Schilling, C.; Marques, M. A. Relating correlation measures: The importance of the energy gap. Phys. Rev. A 2017, 95, 032507,  DOI: 10.1103/PhysRevA.95.032507
    109. 109
      Jiang, W.; DeYonker, N. J.; Wilson, A. K. Multireference character for 3d transition-metal-containing molecules. J. Chem. Theory Comput. 2012, 8, 460468,  DOI: 10.1021/ct2006852
    110. 110
      Lee, T. J.; Rice, J. E.; Scuseria, G. E.; Schaefer, H. F. Theoretical investigations of molecules composed only of fluorine, oxygen and nitrogen: determination of the equilibrium structures of FOOF, (NO)2 and FNNF and the transition state structure for FNNF cis-trans isomerization. Theor. Chim. Acta 1989, 75, 8198,  DOI: 10.1007/BF00527711
    111. 111
      Liakos, D. G.; Neese, F. Interplay of Correlation and Relativistic Effects in Correlated Calculations on Transition-Metal Complexes: The (Cu2O2)2+ Core Revisited. J. Chem. Theory Comput. 2011, 7, 15111523,  DOI: 10.1021/ct1006949
    112. 112
      Head-Gordon, M. Characterizing unpaired electrons from the one-particle density matrix. Chem. Phys. Lett. 2003, 372, 508511,  DOI: 10.1016/S0009-2614(03)00422-6
    113. 113
      Ramos-Cordoba, E.; Matito, E. Local Descriptors of Dynamic and Nondynamic Correlation. J. Chem. Theory Comput. 2017, 13, 27052711,  DOI: 10.1021/acs.jctc.7b00293
    114. 114
      Löwdin, P.-O. Quantum theory of many-particle systems. I. Physical interpretations by means of density matrices, natural spin-orbitals, and convergence problems in the method of configurational interaction. Phys. Rev. 1955, 97, 14741489,  DOI: 10.1103/PhysRev.97.1474
    115. 115
      Ramos-Cordoba, E.; Salvador, P.; Matito, E. Separation of dynamic and nondynamic correlation. Phys. Chem. Chem. Phys. 2016, 18, 2401524023,  DOI: 10.1039/C6CP03072F
    116. 116
      Juhász, T.; Mazziotti, D. A. The cumulant two-particle reduced density matrix as a measure of electron correlation and entanglement. J. Chem. Phys. 2006, 125, 174105,  DOI: 10.1063/1.2378768
    117. 117
      Crittenden, D. L. A Hierarchy of Static Correlation Models. J. Phys. Chem. A 2013, 117, 38523860,  DOI: 10.1021/jp400669p
    118. 118
      Pauncz, R. Spin Eigenfunctions: Construction and Use; Plenum Press: New York, 1979.
    119. 119
      Perdew, J. P.; Ruzsinszky, A.; Sun, J.; Nepal, N. K.; Kaplan, A. D. Interpretations of ground-state symmetry breaking and strong correlation in wavefunction and density functional theories. Proc. Nat. Acad. Sci. 2021, 118, e2017850118,  DOI: 10.1073/pnas.2017850118
    120. 120
      Perdew, J. P.; Ruzsinszky, A.; Constantin, L. A.; Sun, J.; Csonka, G. I. Some fundamental issues in ground-state density functional theory: A guide for the perplexed. J. Chem. Theory Comput. 2009, 5, 902908,  DOI: 10.1021/ct800531s
    121. 121
      Filatov, M.; Shaik, S. Spin-restricted density functional approach to the open-shell problem. Chem. Phys. Lett. 1998, 288, 689697,  DOI: 10.1016/S0009-2614(98)00364-9
    122. 122
      Gagliardi, L.; Truhlar, D. G.; Li Manni, G.; Carlson, R. K.; Hoyer, C. E.; Bao, J. L. Multiconfiguration pair-density functional theory: A new way to treat strongly correlated systems. Acc. Chem. Res. 2017, 50, 6673,  DOI: 10.1021/acs.accounts.6b00471
    123. 123
      Cremer, D. Density functional theory: coverage of dynamic and non-dynamic electron correlation effects. Mol. Phys. 2001, 99, 18991940,  DOI: 10.1080/00268970110083564
    124. 124
      Ziegler, T.; Rauk, A.; Baerends, E. J. On the calculation of multiplet energies by the Hartree-Fock-Slater method. Theor. Chim. Acta 1977, 43, 261271,  DOI: 10.1007/BF00551551
    125. 125
      Noodleman, L. Valence bond description of antiferromagnetic coupling in transition metal dimers. J. Chem. Phys. 1981, 74, 57375743,  DOI: 10.1063/1.440939
    126. 126
      Daul, C. Density functional theory applied to the excited states of coordination compounds. Int. J. Quantum Chem. 1994, 52, 867877,  DOI: 10.1002/qua.560520414
    127. 127
      Neese, F. Definition of corresponding orbitals and the diradical character in broken symmetry DFT calculations on spin coupled systems. J. Phys. Chem. Solids 2004, 65, 781785,  DOI: 10.1016/j.jpcs.2003.11.015
    128. 128
      Gunnarsson, O.; Lundqvist, B. I. Exchange and correlation in atoms, molecules, and solids by the spin-density-functional formalism. Phys. Rev. B 1976, 13, 42744298,  DOI: 10.1103/PhysRevB.13.4274
    129. 129
      Anderson, P. W. More is different: broken symmetry and the nature of the hierarchical structure of science. Science 1972, 177, 393396,  DOI: 10.1126/science.177.4047.393
    130. 130
      Amos, A.; Hall, G. Single determinant wave functions. Proc. R. Soc. A 1961, 263, 483493,  DOI: 10.1098/rspa.1961.0175
    131. 131
      Ladner, R. C.; Goddard III, W. A. Improved Quantum Theory of Many-Electron Systems. V. The Spin-Coupling Optimized GI Method. J. Chem. Phys. 1969, 51, 10731087,  DOI: 10.1063/1.1672106
    132. 132
      Bobrowicz, F. W.; Goddard, W. A. The Self-Consistent Field Equations for Generalized Valence Bond and Open-Shell Hartree─Fock Wave Functions. In Methods of Electronic Structure Theory; Springer: New York, 1977; Chapter 4, pp 79127.
    133. 133
      Ghosh, P.; Bill, E.; Weyhermüller, T.; Neese, F.; Wieghardt, K. Noninnocence of the Ligand Glyoxal-bis (2-mercaptoanil). The Electronic Structures of [Fe(gma)]2,[Fe(gma)(py)].py,[Fe(gma)(CN)]1–/0,[Fe(gma)I], and [Fe(gma)(PR3)n] (n= 1, 2). Experimental and Theoretical Evidence for “Excited State” Coordination. J. Am. Chem. Soc. 2003, 125, 12931308,  DOI: 10.1021/ja021123h
    134. 134
      Herebian, D.; Wieghardt, K. E.; Neese, F. Analysis and interpretation of metal-radical coupling in a series of square planar nickel complexes: correlated ab initio and density functional investigation of [Ni(LISQ)2](LISQ= 3,5-di-tert-butyl-o-diiminobenzosemiquinonate(1-)). J. Am. Chem. Soc. 2003, 125, 1099711005,  DOI: 10.1021/ja030124m
    135. 135
      Chłopek, K.; Muresan, N.; Neese, F.; Wieghardt, K. Electronic Structures of Five-Coordinate Complexes of Iron Containing Zero, One, or Two π-Radical Ligands: A Broken-Symmetry Density Functional Theoretical Study. Chem. Eur. J. 2007, 13, 83908403,  DOI: 10.1002/chem.200700897
    136. 136
      Neese, F. Prediction of molecular properties and molecular spectroscopy with density functional theory: From fundamental theory to exchange-coupling. Coord. Chem. Rev. 2009, 253, 526563,  DOI: 10.1016/j.ccr.2008.05.014
    137. 137
      Hohenberg, P.; Kohn, W. Inhomogeneous electron gas. Phys. Rev. 1964, 136, B864B871,  DOI: 10.1103/PhysRev.136.B864
    138. 138
      Kutzelnigg, W. Density functional theory in terms of a Legendre transformation for beginners. J. Mol. Struct.: THEOCHEM 2006, 768, 163173,  DOI: 10.1016/j.theochem.2006.05.012
    139. 139
      Lieb, E. H. Density functionals for Coulomb systems. Int. J. Quantum Chem. 1983, 24, 243277,  DOI: 10.1002/qua.560240302
    140. 140
      Penz, M.; Tellgren, E. I.; Csirik, M. A.; Ruggenthaler, M.; Laestadius, A. structure of the density-potential mapping. Part I: Standard density-functional theory. arXiv:2211.16627 2022, DOI: 10.48550/arXiv.2211.16627 .
    141. 141
      Casida, M. E.; Huix-Rotllant, M. Progress in time-dependent density-functional theory. Annu. Rev. Phys. Chem. 2012, 63, 287323,  DOI: 10.1146/annurev-physchem-032511-143803
    142. 142
      Kohn, W. Density Functional and Density Matrix Method Scaling Linearly with the Number of Atoms. Phys. Rev. Lett. 1996, 76, 31683171,  DOI: 10.1103/PhysRevLett.76.3168
    143. 143
      Prodan, E.; Kohn, W. Nearsightedness of electronic matter. Proc. Nat. Acad. Sci. 2005, 102, 1163511638,  DOI: 10.1073/pnas.0505436102
    144. 144
      Li, L.; Burke, K. Recent developments in density functional approximations. In Handbook of Materials Modeling: Methods: Theory and Modeling; Springer: Cham, 2020; pp 213226.
    145. 145
      Mayer, J. E. Electron Correlation. Phys. Rev. 1955, 100, 15791586,  DOI: 10.1103/PhysRev.100.1579
    146. 146
      Bopp, F. Ableitung der Bindungsenergie vonN-Teilchen-Systemen aus 2-Teilchen-Dichtematrizen. Z. Phys. 1959, 156, 348359,  DOI: 10.1007/BF01461233
    147. 147
      Mazziotti, D. A. Structure of Fermionic Density Matrices: Complete N-Representability Conditions. Phys. Rev. Lett. 2012, 108, 263002,  DOI: 10.1103/PhysRevLett.108.263002
    148. 148
      Gilbert, T. L. Hohenberg-Kohn theorem for nonlocal external potentials. Phys. Rev. B 1975, 12, 21112120,  DOI: 10.1103/PhysRevB.12.2111
    149. 149
      Müller, A. Explicit approximate relation between reduced two- and one-particle density matrices. Phys. Lett. A 1984, 105, 446452,  DOI: 10.1016/0375-9601(84)91034-X
    150. 150
      Kutzelnigg, W. Density-cumulant functional theory. J. Chem. Phys. 2006, 125, 171101,  DOI: 10.1063/1.2387955
    151. 151
      Piris, M.; Ugalde, J. M. Perspective on natural orbital functional theory. Int. J. Quantum Chem. 2014, 114, 11691175,  DOI: 10.1002/qua.24663
    152. 152
      Pernal, K., Giesbertz, K. J. H. Reduced Density Matrix Functional Theory (RDMFT) and Linear Response Time-Dependent RDMFT (TD-RDMFT). In Density-Functional Methods for Excited States; Springer: Cham, 2016; pp 125183.
    153. 153
      Coleman, A. J. Structure of fermion density matrices. Rev. Mod. Phys. 1963, 35, 668686,  DOI: 10.1103/RevModPhys.35.668
    154. 154
      Cioslowski, J.; Pernal, K.; Buchowiecki, M. Approximate one-matrix functionals for the electron–electron repulsion energy from geminal theories. J. Chem. Phys. 2003, 119, 64436447,  DOI: 10.1063/1.1604375
    155. 155
      Piris, M. A natural orbital functional based on an explicit approach of the two-electron cumulant. Int. J. Quantum Chem. 2013, 113, 620630,  DOI: 10.1002/qua.24020
    156. 156
      Simmonett, A. C.; Wilke, J. J.; Schaefer, H. F., III; Kutzelnigg, W. Density cumulant functional theory: First implementation and benchmark results for the DCFT-06 model. J. Chem. Phys. 2010, 133, 174122,  DOI: 10.1063/1.3503657
    157. 157
      Kutzelnigg, W. Density Functional Theory (DFT) and ab-initio Quantum Chemistry (AIQC). Story of a difficult partnership. In Trends and Perspectives in Modern Computational Science; Brill: Leiden, 2006; pp 2362.
    158. 158
      Copan, A. V.; Sokolov, A. Y. Linear-response density cumulant theory for excited electronic states. J. Chem. Theory Comput. 2018, 14, 40974108,  DOI: 10.1021/acs.jctc.8b00326
    159. 159
      Cioslowski, J., Ed. Many-electron densities and reduced density matrices; Springer Science & Business Media: New York, 2000.
    160. 160
      Gidofalvi, G.; Mazziotti, D. A. Active-space two-electron reduced-density-matrix method: Complete active-space calculations without diagonalization of the N-electron Hamiltonian. J. Chem. Phys. 2008, 129, 134108,  DOI: 10.1063/1.2983652
    161. 161
      Mullinax, J. W.; Sokolov, A. Y.; Schaefer, H. F., III Can density cumulant functional theory describe static correlation effects?. J. Chem. Theory Comput. 2015, 11, 24872495,  DOI: 10.1021/acs.jctc.5b00346
    162. 162
      Lathiotakis, N. N.; Helbig, N.; Rubio, A.; Gidopoulos, N. I. Local reduced-density-matrix-functional theory: Incorporating static correlation effects in Kohn-Sham equations. Phys. Rev. A 2014, 90, 032511,  DOI: 10.1103/PhysRevA.90.032511
    163. 163
      Piris, M. Global Method for Electron Correlation. Phys. Rev. Lett. 2017, 119, 063002,  DOI: 10.1103/PhysRevLett.119.063002
    164. 164
      Gibney, D.; Boyn, J.-N.; Mazziotti, D. A. Density Functional Theory Transformed into a One-Electron Reduced-Density-Matrix Functional Theory for the Capture of Static Correlation. J. Phys. Chem. Lett. 2022, 13, 13821388,  DOI: 10.1021/acs.jpclett.2c00083
    165. 165
      Martin, P. C.; Schwinger, J. Theory of many-particle systems. I. Phys. Rev. 1959, 115, 13421373,  DOI: 10.1103/PhysRev.115.1342
    166. 166
      Luttinger, J. M.; Ward, J. C. Ground-state energy of a many-fermion system. II. Phys. Rev. 1960, 118, 14171427,  DOI: 10.1103/PhysRev.118.1417
    167. 167
      Baym, G.; Kadanoff, L. P. Conservation laws and correlation functions. Phys. Rev. 1961, 124, 287299,  DOI: 10.1103/PhysRev.124.287
    168. 168
      Ziesche, P. Cumulant expansions of reduced densities, reduced density matrices, and Green’s functions. In Many-Electron Densities and Reduced Density Matrices; Springer: New York, 2000; Chapter 3, pp 3356.
    169. 169
      Sun, Q.; Chan, G. K.-L. Quantum embedding theories. Acc. Chem. Res. 2016, 49, 27052712,  DOI: 10.1021/acs.accounts.6b00356
    170. 170
      Kotliar, G.; Savrasov, S. Y.; Haule, K.; Oudovenko, V. S.; Parcollet, O.; Marianetti, C. Electronic structure calculations with dynamical mean-field theory. Rev. Mod. Phys. 2006, 78, 865951,  DOI: 10.1103/RevModPhys.78.865
    171. 171
      Lin, N.; Marianetti, C.; Millis, A. J.; Reichman, D. R. Dynamical mean-field theory for quantum chemistry. Phys. Rev. Lett. 2011, 106, 096402,  DOI: 10.1103/PhysRevLett.106.096402
    172. 172
      Zgid, D.; Chan, G. K.-L. Dynamical mean-field theory from a quantum chemical perspective. J. Chem. Phys. 2011, 134, 094115,  DOI: 10.1063/1.3556707
    173. 173
      Hedin, L. New method for calculating the one-particle Green’s function with application to the electron-gas problem. Phys. Rev. 1965, 139, A796A823,  DOI: 10.1103/PhysRev.139.A796
    174. 174
      Aryasetiawan, F.; Gunnarsson, O. The GW method. Rep. Prog. Phys. 1998, 61, 237312,  DOI: 10.1088/0034-4885/61/3/002
    175. 175
      Reining, L. The GW approximation: content, successes and limitations. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2018, 8, e1344,  DOI: 10.1002/wcms.1344
    176. 176
      Phillips, J. J.; Zgid, D. Communication: The description of strong correlation within self-consistent Green’s function second-order perturbation theory. J. Chem. Phys. 2014, 140, 241101,  DOI: 10.1063/1.4884951
    177. 177
      Pokhilko, P.; Zgid, D. Interpretation of multiple solutions in fully iterative GF2 and GW schemes using local analysis of two-particle density matrices. J. Chem. Phys. 2021, 155, 024101,  DOI: 10.1063/5.0055191
    178. 178
      Blase, X.; Duchemin, I.; Jacquemin, D. The Bethe–Salpeter equation in chemistry: relations with TD-DFT, applications and challenges. Chem. Soc. Rev. 2018, 47, 10221043,  DOI: 10.1039/C7CS00049A
    179. 179
      Pokhilko, P.; Iskakov, S.; Yeh, C.-N.; Zgid, D. Evaluation of two-particle properties within finite-temperature self-consistent one-particle Green’s function methods: Theory and application to GW and GF2. J. Chem. Phys. 2021, 155, 024119,  DOI: 10.1063/5.0054661
    180. 180
      Cohen, A. J.; Mori-Sánchez, P.; Yang, W. Fractional spins and static correlation error in density functional theory. J. Chem. Phys. 2008, 129, 121104,  DOI: 10.1063/1.2987202
    181. 181
      Cohen, A. J.; Mori-Sánchez, P.; Yang, W. Insights into current limitations of density functional theory. Science 2008, 321, 792794,  DOI: 10.1126/science.1158722
    182. 182
      Grimme, S.; Hansen, A. A practicable real-space measure and visualization of static electron-correlation effects. Angew. Chem., Int. Ed. 2015, 54, 1230812313,  DOI: 10.1002/anie.201501887
    183. 183
      Muechler, L.; Badrtdinov, D. I.; Hampel, A.; Cano, J.; Rösner, M.; Dreyer, C. E. Quantum embedding methods for correlated excited states of point defects: Case studies and challenges. Phys. Rev. B 2022, 105, 235104,  DOI: 10.1103/PhysRevB.105.235104
    184. 184
      Nielsen, M. A.; Chuang, I. L. Quantum Computation and Quantum Information; Cambridge University Press: Cambridge, 2010.
    185. 185
      Almeida, M.; Omar, Y.; Rocha Vieira, V. Introduction to entanglement and applications to the simulation of many-body quantum systems. In Strongly Correlated Systems, Coherence And Entanglement; World Scientific: Singapore, 2007; Chapter 19, pp 525547.
    186. 186
      Ding, L.; Knecht, S.; Zimborás, Z.; Schilling, C. Quantum Correlations in Molecules: from Quantum Resourcing to Chemical Bonding. Quantum Sci. Technol. 2023, 8, 015015,  DOI: 10.1088/2058-9565/aca4ee
    187. 187
      Omar, Y. Particle statistics in quantum information processing. Int. J. Quantum Inf. 2005, 3, 201205,  DOI: 10.1142/S021974990500075X
    188. 188
      Benatti, F.; Floreanini, R.; Franchini, F.; Marzolino, U. Entanglement in indistinguishable particle systems. Phys. Rep. 2020, 878, 127,  DOI: 10.1016/j.physrep.2020.07.003
    189. 189
      Henderson, T. M.; Bulik, I. W.; Stein, T.; Scuseria, G. E. Seniority-based coupled cluster theory. J. Chem. Phys. 2014, 141, 244104,  DOI: 10.1063/1.4904384
    190. 190
      Boguslawski, K.; Tecmer, P. Orbital entanglement in quantum chemistry. Int. J. Quantum Chem. 2015, 115, 12891295,  DOI: 10.1002/qua.24832
    191. 191
      Ding, L.; Zimboras, Z.; Schilling, C. Quantifying Electron Entanglement Faithfully. arXiv:2207.03377 2022, DOI: 10.48550/arXiv.2207.03377 .
    192. 192
      Legeza, Ö.; Sólyom, J. Optimizing the density-matrix renormalization group method using quantum information entropy. Phys. Rev. B 2003, 68, 195116,  DOI: 10.1103/PhysRevB.68.195116
    193. 193
      Stein, C. J.; Reiher, M. Measuring multi-configurational character by orbital entanglement. Mol. Phys. 2017, 115, 21102119,  DOI: 10.1080/00268976.2017.1288934
    194. 194
      Stein, C. J.; Reiher, M. Automated selection of active orbital spaces. J. Chem. Theory Comput. 2016, 12, 17601771,  DOI: 10.1021/acs.jctc.6b00156
    195. 195
      Boguslawski, K.; Tecmer, P.; Barcza, G.; Legeza, O.; Reiher, M. Orbital entanglement in bond-formation processes. J. Chem. Theory Comput. 2013, 9, 29592973,  DOI: 10.1021/ct400247p
    196. 196
      Huang, Z.; Kais, S. Entanglement as measure of electron–electron correlation in quantum chemistry calculations. Chem. Phys. Lett. 2005, 413, 15,  DOI: 10.1016/j.cplett.2005.07.045
    197. 197
      Boguslawski, K.; Tecmer, P.; Legeza, O.; Reiher, M. Entanglement measures for single-and multireference correlation effects. J. Phys. Chem. Lett. 2012, 3, 31293135,  DOI: 10.1021/jz301319v
    198. 198
      Ziesche, P. Correlation strength and information entropy. Int. J. Quantum Chem. 1995, 56, 363369,  DOI: 10.1002/qua.560560422
    199. 199
      Ghosh, K. J.; Kais, S.; Herschbach, D. R. Geometrical picture of the electron–electron correlation at the large-D limit. Phys. Chem. Chem. Phys. 2022, 24, 92989307,  DOI: 10.1039/D2CP00438K
    200. 200
      Rissler, J.; Noack, R. M.; White, S. R. Measuring orbital interaction using quantum information theory. Chem. Phys. 2006, 323, 519531,  DOI: 10.1016/j.chemphys.2005.10.018
    201. 201
      Cao, Y.; Romero, J.; Olson, J. P.; Degroote, M.; Johnson, P. D.; Kieferová, M.; Kivlichan, I. D.; Menke, T.; Peropadre, B.; Sawaya, N. P. D.; Sim, S.; Veis, L.; Aspuru-Guzik, A. Quantum Chemistry in the Age of Quantum Computing. Chem. Rev. 2019, 119, 1085610915,  DOI: 10.1021/acs.chemrev.8b00803
    202. 202
      Reiher, M.; Wiebe, N.; Svore, K. M.; Wecker, D.; Troyer, M. Elucidating reaction mechanisms on quantum computers. Proc. Nat. Acad. Sci. 2017, 114, 75557560,  DOI: 10.1073/pnas.1619152114
    203. 203
      Tubman, N. M.; Mejuto-Zaera, C.; Epstein, J. M.; Hait, D.; Levine, D. S.; Huggins, W.; Jiang, Z.; McClean, J. R.; Babbush, R.; Head-Gordon, M.; Whaley, K. B. Postponing the orthogonality catastrophe: efficient state preparation for electronic structure simulations on quantum devices. arXiv:1809.05523 2018, DOI: 10.48550/arXiv.1809.05523 .
    204. 204
      Carbone, A.; Galli, D. E.; Motta, M.; Jones, B. Quantum circuits for the preparation of spin eigenfunctions on quantum computers. Symmetry 2022, 14, 624,  DOI: 10.3390/sym14030624
    205. 205
      Lacroix, D.; Ruiz Guzman, E. A.; Siwach, P. Symmetry breaking/symmetry preserving circuits and symmetry restoration on quantum computers. Eur. Phys. J. A 2023, 59, 3,  DOI: 10.1140/epja/s10050-022-00911-7
    206. 206
      Lee, S.; Lee, J.; Zhai, H.; Tong, Y.; Dalzell, A. M.; Kumar, A.; Helms, P.; Gray, J.; Cui, Z.-H.; Liu, W.; Kastoryano, M.; Babbush, R.; Preskill, J.; Reichman, D. R.; Campbell, E. T.; Valeev, E. F.; Lin, L.; Chan, G. K.-L. Is there evidence for exponential quantum advantage in quantum chemistry. arXiv:2208.02199 2022, DOI: 10.48550/arXiv.2208.02199 .
    207. 207
      Tazhigulov, R. N.; Sun, S.-N.; Haghshenas, R.; Zhai, H.; Tan, A. T.; Rubin, N. C.; Babbush, R.; Minnich, A. J.; Chan, G. K.-L. Simulating models of challenging correlated molecules and materials on the Sycamore quantum processor. PRX Quantum 2022, 3, 040318,  DOI: 10.1103/PRXQuantum.3.040318
    208. 208
      Amsler, M.; Deglmann, P.; Degroote, M.; Kaicher, M. P.; Kiser, M.; Kühn, M.; Kumar, C.; Maier, A.; Samsonidze, G.; Schroeder, A.; Streif, M.; Vodola, D.; Wever, C. Quantum-enhanced quantum Monte Carlo: an industrial view. arXiv:2301.11838 2023, DOI: 10.48550/arXiv.2301.11838 .
    209. 209
      Rubin, N. C.; Berry, D. W.; Malone, F. D.; White, A. F.; Khattar, T.; Sicolo, A. E. D.; Kühn, S.; Kaicher, M.; Lee, M.; Babbush, J. Fault-tolerant quantum simulation of materials using Bloch orbitals. arXiv:2302.05531 2023, DOI: 10.48550/arXiv.2302.05531 .
    210. 210
      Lykos, P.; Pratt, G. W. Discussion on The Hartree-Fock Approximation. Rev. Mod. Phys. 1963, 35, 496501,  DOI: 10.1103/RevModPhys.35.496
    211. 211
      Slater, J. C. The Theory of Complex Spectra. Phys. Rev. 1929, 34, 12931322,  DOI: 10.1103/PhysRev.34.1293
    212. 212
      Slater, J. C. Solid-state and molecular theory: a scientific biography; Wiley-Interscience: New York, 1975.
    213. 213
      Matsen, F. Spin-free quantum chemistry. In Adv. Quantum Chem.; Interscience: New York, 1964; Vol. 1, pp 59114.
    214. 214
      Paldus, J. The Unitary Group for the Evaluation of Electronic Energy Matrix Elements. In The Unitary Group for the Evaluation of Electronic Energy Matrix Elements; Springer: Berlin, 1981; pp 150.
    215. 215
      Shavitt, I. The Graphical Unitary Group Approach and its Application to Direct Configuration Interaction Calculations. In The Unitary Group for the Evaluation of Electronic Energy Matrix Elements; Springer: Berlin, 1981; pp 5199.
    216. 216
      Dobrautz, W.; Smart, S. D.; Alavi, A. Efficient formulation of full configuration interaction quantum Monte Carlo in a spin eigenbasis via the graphical unitary group approach. J. Chem. Phys. 2019, 151, 094104,  DOI: 10.1063/1.5108908
    217. 217
      McLean, A. D.; Lengsfield, B. H.; Pacansky, J.; Ellinger, Y. Symmetry breaking in molecular calculations and the reliable prediction of equilibrium geometries. The formyloxyl radical as an example. J. Chem. Phys. 1985, 83, 35673576,  DOI: 10.1063/1.449162
    218. 218
      Ayala, P. Y.; Schlegel, H. B. A nonorthogonal CI treatment of symmetry breaking in sigma formyloxyl radical. J. Chem. Phys. 1998, 108, 75607567,  DOI: 10.1063/1.476190
    219. 219
      Manne, R. Brillouin’s theorem in Roothaan’s open-shell SCF method. Mol. Phys. 1972, 24, 935944,  DOI: 10.1080/00268977200102061
    220. 220
      Davidson, E. R.; Borden, W. T. Symmetry breaking in polyatomic molecules: real and artifactual. J. Phys. Chem. 1983, 87, 47834790,  DOI: 10.1021/j150642a005
    221. 221
      Čížek, J.; Paldus, J. Stability Conditions for the Solutions of the Hartree–Fock Equations for Atomic and Molecular Systems. Application to the Pi-Electron Model of Cyclic Polyenes. J. Chem. Phys. 1967, 47, 39763985,  DOI: 10.1063/1.1701562
    222. 222
      Čížek, J.; Paldus, J. Stability Conditions for the Solutions of the Hartree–Fock Equations for Atomic and Molecular Systems. III. Rules for the Singlet Stability of Hartree–Fock Solutions of π-Electronic Systems. J. Chem. Phys. 1970, 53, 821829,  DOI: 10.1063/1.1674065
    223. 223
      Paldus, J.; Čížek, J. Stability Conditions for the Solutions of the Hartree-Fock Equations for Atomic and Molecular Systems. VI. Singlet-Type Instabilities and Charge-Density-Wave Hartree-Fock Solutions for Cyclic Polyenes. Phys. Rev. A 1970, 2, 22682283,  DOI: 10.1103/PhysRevA.2.2268
    224. 224
      Davidson, E. R. Spin-restricted open-shell self-consistent-field theory. Chem. Phys. Lett. 1973, 21, 565567,  DOI: 10.1016/0009-2614(73)80309-4
    225. 225
      Edwards, W. D.; Zerner, M. C. A generalized restricted open-shell Fock operator. Theor. Chim. Acta 1987, 72, 347361,  DOI: 10.1007/BF01192227
    226. 226
      Löwdin, P.-O. Quantum Theory of Many-Particle Systems. III. Extension of the Hartree-Fock Scheme to Include Degenerate Systems and Correlation Effects. Phys. Rev. 1955, 97, 15091520,  DOI: 10.1103/PhysRev.97.1509
    227. 227
      Mayer, I. The spin-projected extended Hartree-Fock method. In Adv. Quantum Chem.; Academic Press: New York, 1980; Vol. 12, pp 189262.
    228. 228
      Tsuchimochi, T.; Scuseria, G. E. Communication: ROHF theory made simple. J. Chem. Phys. 2010, 133, 141102,  DOI: 10.1063/1.3503173
    229. 229
      Rittby, M.; Bartlett, R. J. An open-shell spin-restricted coupled cluster method: application to ionization potentials in nitrogen. J. Phys. Chem. 1988, 92, 30333036,  DOI: 10.1021/j100322a004
    230. 230
      Neese, F. Importance of direct spin-spin coupling and spin-flip excitations for the zero-field splittings of transition metal complexes: A case study. J. Am. Chem. Soc. 2006, 128, 1021310222,  DOI: 10.1021/ja061798a
    231. 231
      Casanova, D.; Krylov, A. I. Spin-flip methods in quantum chemistry. Phys. Chem. Chem. Phys. 2020, 22, 43264342,  DOI: 10.1039/C9CP06507E
    232. 232
      Roemelt, M.; Neese, F. Excited states of large open-shell molecules: an efficient, general, and spin-adapted approach based on a restricted open-shell ground state wave function. J. Phys. Chem. A 2013, 117, 30693083,  DOI: 10.1021/jp3126126
    233. 233
      Izsák, R. Second quantisation for unrestricted references: formalism and quasi-spin-adaptation of excitation and spin-flip operators. Mol. Phys. 2022, e2126802,  DOI: 10.1080/00268976.2022.2126802
    234. 234
      Veis, L.; Antalik, A.; Legeza, O.; Alavi, A.; Pittner, J. The intricate case of tetramethyleneethane: A full configuration interaction quantum Monte Carlo benchmark and multireference coupled cluster studies. J. Chem. Theory Comput. 2018, 14, 24392445,  DOI: 10.1021/acs.jctc.8b00022
    235. 235
      Sears, J. S.; Sherrill, C. D. Assessing the Performance of Density Functional Theory for the Electronic Structure of Metal-Salens: The 3d0-Metals. J. Phys. Chem. A 2008, 112, 34663477,  DOI: 10.1021/jp711595w
    236. 236
      Sears, J. S.; Sherrill, C. D. Assessing the performance of Density Functional Theory for the electronic structure of metal-salens: the d2-metals. J. Phys. Chem. A 2008, 112, 67416752,  DOI: 10.1021/jp802249n
    237. 237
      Olivares-Amaya, R.; Hu, W.; Nakatani, N.; Sharma, S.; Yang, J.; Chan, G. K.-L. The ab-initio density matrix renormalization group in practice. J. Chem. Phys. 2015, 142, 034102,  DOI: 10.1063/1.4905329
    238. 238
      Pandharkar, R.; Hermes, M. R.; Cramer, C. J.; Gagliardi, L. Localized Active Space-State Interaction: a Multireference Method for Chemical Insight. J. Chem. Theory Comput. 2022, 18, 65576566,  DOI: 10.1021/acs.jctc.2c00536
    239. 239
      Máté, M.; Petrov, K.; Szalay, S.; Legeza, Ö. Compressing multireference character of wave functions via fermionic mode optimization. J. Math. Chem. 2023, 61, 362375,  DOI: 10.1007/s10910-022-01379-y
    240. 240
      Smith, J. E. T.; Mussard, B.; Holmes, A. A.; Sharma, S. Cheap and Near Exact CASSCF with Large Active Spaces. J. Chem. Theory Comput. 2017, 13, 54685478,  DOI: 10.1021/acs.jctc.7b00900
    241. 241
      Dobrautz, W.; Weser, O.; Bogdanov, N. A.; Alavi, A.; Li Manni, G. Spin-Pure Stochastic-CASSCF via GUGA-FCIQMC Applied to Iron–Sulfur Clusters. J. Chem. Theory Comput. 2021, 17, 56845703,  DOI: 10.1021/acs.jctc.1c00589
    242. 242
      Dobrautz, W.; Katukuri, V. M.; Bogdanov, N. A.; Kats, D.; Li Manni, G.; Alavi, A. Combined unitary and symmetric group approach applied to low-dimensional Heisenberg spin systems. Phys. Rev. B 2022, 105, 195123,  DOI: 10.1103/PhysRevB.105.195123
    243. 243
      Li Manni, G.; Kats, D.; Liebermann, N. Resolution of Electronic States in Heisenberg Cluster Models within the Unitary Group Approach. ChemRxiv 2022, DOI: 10.26434/chemrxiv-2022-rfmhk-v2 .
    244. 244
      Pipek, J.; Mezey, P. G. A fast intrinsic localization procedure applicable for abinitio and semiempirical linear combination of atomic orbital wave functions. J. Chem. Phys. 1989, 90, 49164926,  DOI: 10.1063/1.456588
    245. 245
      Malrieu, J.-P.; Guihéry, N.; Calzado, C. J.; Angeli, C. Bond electron pair: Its relevance and analysis from the quantum chemistry point of view. J. Comput. Chem. 2007, 28, 3550,  DOI: 10.1002/jcc.20546
    246. 246
      Lewis, G. N. The atom and the molecule. J. Am. Chem. Soc. 1916, 38, 762785,  DOI: 10.1021/ja02261a002
    247. 247
      Chen, H.; Lai, W.; Shaik, S. Multireference and Multiconfiguration Ab Initio Methods in Heme-Related Systems: What Have We Learned So Far?. J. Phys. Chem. B 2011, 115, 17271742,  DOI: 10.1021/jp110016u
    248. 248
      Li, Z.; Guo, S.; Sun, Q.; Chan, G. K.-L. Electronic landscape of the P-cluster of nitrogenase as revealed through many-electron quantum wavefunction simulations. Nat. Chem. 2019, 11, 10261033,  DOI: 10.1038/s41557-019-0337-3
    249. 249
      Khedkar, A.; Roemelt, M. Modern multireference methods and their application in transition metal chemistry. Phys. Chem. Chem. Phys. 2021, 23, 1709717112,  DOI: 10.1039/D1CP02640B
    250. 250
      Tarrago, M.; Römelt, C.; Nehrkorn, J.; Schnegg, A.; Neese, F.; Bill, E.; Ye, S. Experimental and Theoretical Evidence for an Unusual Almost Triply Degenerate Electronic Ground State of Ferrous Tetraphenylporphyrin. Inorg. Chem. 2021, 60, 49664985,  DOI: 10.1021/acs.inorgchem.1c00031
    251. 251
      Han, R.; Luber, S.; Li Manni, G. Magnetic Interactions in a [Co(II)3Er(III)(OR)4] Model Cubane Through Forefront Multiconfigurational Methods. ChemRxiv 2023, DOI: 10.26434/chemrxiv-2023-xd0wv .

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

You’ve supercharged your research process with ACS and Mendeley!

STEP 1:
Click to create an ACS ID

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

MENDELEY PAIRING EXPIRED
Your Mendeley pairing has expired. Please reconnect