Next Article in Journal / Special Issue
Fully Conjugated Heteroatomic Non- and Quasi-Alternant Polyradicals
Previous Article in Journal / Special Issue
An Electronic Structural Analysis of O2-Binding Dicopper Complex: Insights from Spin Magnetism and Molecular Orbitals
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fragmentation-Based Linear-Scaling Method for Strongly Correlated Systems: Divide-and-Conquer Hartree–Fock–Bogoliubov Method, Its Energy Gradient, and Applications to Graphene Nano-Ribbon Systems

1
Department of Chemistry, Faculty of Science, Hokkaido University, Sapporo 060-0810, Japan
2
Institute for Chemical Reaction Design and Discovery (WPI-ICReDD), Hokkaido University, Sapporo 001-0021, Japan
*
Author to whom correspondence should be addressed.
Chemistry 2025, 7(2), 46; https://doi.org/10.3390/chemistry7020046
Submission received: 29 January 2025 / Revised: 1 March 2025 / Accepted: 4 March 2025 / Published: 18 March 2025

Abstract

:
This study introduces a fragmentation-based linear-scaling method for strongly correlated systems, specifically the divide-and-conquer Hartree–Fock–Bogoliubov (DC-HFB) approach. Two energy gradient formulations of the DC-HFB method are derived and implemented, enabling efficient optimization of molecular geometries in large systems. This method is applied to graphene nanoribbons (GNRs) to explore their geometries and polyradical characters. Numerical results demonstrate that the present DC-HFB method has the potential to treat the static electron correlation and predict diradical character in GNRs, offering new avenues for studying large-scale strongly correlated systems.

1. Introduction

A simple yet accurate description of static electron correlation, especially in large systems, is still a challenging problem in electronic structure theory. Static electron correlation is often closely related to chemical structure. For example, a bond-alternating polyene chain has a weaker static correlation than a chain with a uniform bond length. Transition-state structures often suffer from stronger static correlation than stable structures. Polyacenes [1] and graphene nanoribbons (GNRs) [2] exhibit different strengths of static correlation depending on their size, shape, and topology (sheet, moebius, or twisted) [3]. In these systems, the strength of static correlation is related to diradical character [4]. Diradical character was originally defined in the context of the valence configuration interaction wavefunction [5]. A definition based on the natural orbital occupation numbers obtained from the unrestricted Hartree–Fock (UHF) density matrix was also proposed [6,7]. In particular, Nakano has conducted detailed research on the relationship between diradical character and excited-state properties such as optical response [8,9,10].
A standard method for describing static electron correlation is the active-space self-consistent field (SCF) method, typified by the complete active-space SCF (CASSCF) method [11]. These methods require the specification of an active space that defines strongly correlated orbitals and electrons. Although the recent advent of the density matrix renormalization group method [12,13,14,15] allows much larger active spaces to be handled than before, its application is still limited to medium-sized systems.
Another way to account for static correlation is based on the two-electron wavefunction, called geminal [16,17,18]. These schemes, including generalized valence bond methods [19], only treat two opposite-spin electrons in a variational way, and additionally, dynamical correlation may be incorporated in a perturbational manner [20,21]. Wang et al. applied the block-correlated coupled cluster method to a generalized valence bond reference to account for intergeminal electron correlation [22]. Although the formal computational scaling for these methods can be lower than the CASSCF method with a large active space, the simultaneous optimization of orbitals and CI coefficients usually causes serious convergence problems. Note that the geminal method is also related to pair-coupled cluster theories [23,24].
An alternative approach to static correlation problems inspired by the geminal-based method is the Hartree–Fock–Bogoliubov (HFB) method reformulated for quantum chemical problems [25]. Since the HFB wavefunction is related to the antisymmetrized geminal power wavefunction by particle-number projection [26], the HFB method can effectively describe static electron correlation, although the obtained as-is HFB wavefunction is contaminated by different-particle-number wavefunctions. The formal computational scaling for the HFB method is similar to that of the Hartree–Fock (HF) method with doubled basis functions. However, we have proposed a method to apply the linear-scaling divide-and-conquer (DC) scheme [27,28,29] to the HFB method [30].
To predict molecular geometry from quantum chemical calculations, the energy gradient must be formulated and implemented in computational code [31]. One of the authors (MK) has formulated the HFB energy gradient [32]. Note that Chai and coworkers have proposed thermally assisted occupation density functional theory (TAO-DFT) [33,34,35]. This method can consider static electron correlation by introducing fictitious electronic temperature, and its energy gradient is available. On the other hand, in the DC method, the energy gradient was proposed in the earliest Yang and Lee paper in the HF and DFT frameworks [27]. Since this DC energy gradient expression includes an approximation, MK and coworkers delived the exact DC-HF/DFT energy gradient and proposed an improved approximate energy gradient expression [36].
In this paper, we derived two energy gradient expressions for the DC-HFB method in accordance with refs. [27,36]. The accuracy of these two gradient expressions is investigated through calculations of GNRs including polyacenes. In addition, the strength of static electron correlation is discussed based on the natural orbitals of the HFB wavefunction and their occupation numbers, which are closely related to diradical, tetraradical, and hexaradical characters. Although the simultaneous consideration of dynamical electron correlation as well as static correlation is necessary for a quantitative discussion, we focus here only on static correlation for a basic evaluation of the method. The organization of this paper is as follows: The theoretical perspective of the presented method is given in Section 2. The numerical assessment of GNR calculations is presented in Section 3. Finally, Section 4 gives concluding remarks.

2. Theory

2.1. Restricted DC-HFB Method

Here, we focus on the restricted HFB method for simplicity [32]. Therefore, we only consider singlet systems, including open-shell singlets. The generalization to the unrestricted HFB scheme (i.e., non-singlet systems) will be straightforward. Throughout this paper, Greek-letter indices { μ , ν , } refer to basis functions [i.e., atomic orbitals (AOs) that can be non-orthogonal] and { i , j , } refer to quasiparticle orbitals with negative eigenvalues. The restricted HFB energy can be expressed as a functional of the density matrix D ¯ D and the pairing matrix K ¯ K = K :
E HFB = 2 μ ν h μ ν D ¯ ν μ + μ ν λ ρ ( 2 μ ν | λ ρ μ ν | ρ λ ) D ¯ λ μ D ¯ ρ ν ζ μ ν λ ρ μ ν | λ ρ K ¯ μ ν K ¯ λ ρ + 2 Λ N μ ν S μ ν D ¯ ν μ ,
where μ ν | λ ρ = μ ( r 1 ) ν ( r 2 ) r 12 1 λ ( r 1 ) ρ ( r 2 ) d r 1 d r 2 , h is a one-electron integral matrix, S represents the overlap matrix of the non-orthogonal basis functions, and ζ is a scaling factor introduced by Staroverov and Scuseria [25]. Λ is a chemical potential that is tuned to ensure that the number of electrons in the system is maintained at 2 N . The density and pairing matrices constitute the generalized density matrix R ¯ :
R ¯ = D ¯ K ¯ K ¯ S 1 D ¯ .
Variation in the energy (1) with respect to the generalized density matrix (2) leads to the restricted HFB Hamiltonian
H = F ˜ Δ Δ F ˜
where F ˜ is the shifted Fock matrix,
F ˜ μ λ = h μ λ Λ S μ λ + ν ρ ( 2 μ ν | λ ρ μ ν | ρ λ ) D ¯ ρ ν = h ˜ μ λ + ν ρ ( 2 μ ν | λ ρ μ ν | ρ λ ) D ¯ ρ ν ,
with h ˜ = h Λ S , and Δ is the pairing-field matrix,
Δ μ ν = ζ λ ρ μ ν | λ ρ K ¯ λ ρ .
The eigenvectors of the HFB Hamiltonian (3), ( Y i T X i T ) T ,
F ˜ Δ Δ F ˜ Y i X i = S 0 0 S Y i X i ε i ,
represent the quasiparticle orbitals with negative eigenvalues ( ε i 0 ), which are connected to the density and pairing matrices as follows:
D ¯ μ ν = i M Y μ i Y ν i ,
K ¯ μ ν = i M Y μ i X ν i = i M X μ i Y ν i .
Here, M represents the number of quasiparticle orbitals. Note that the generalized density matrix R ¯ of the converged HFB solution is idempotent, namely,
D ¯ S D ¯ D ¯ = K ¯ S K ¯ ,
D ¯ S K ¯ = K ¯ S D ¯ .
In the DC method, the entire system is spatially divided into disjoint subsystems called the central region, where the set of AOs in subsystem α is denoted by S ( α ) , and the union of S ( α ) becomes the set of basis functions over the entire system denoted by T as follows:
S ( α ) S ( β ) = α β ,
α S ( α ) = T .
When constructing MOs for subsystem α , the AOs of atoms adjacent to the central region, called the buffer region, denoted by B ( α ) , are taken into consideration in addition to S ( α ) . The union of the central and buffer regions of subsystem α is called the localization region, in which the set of AOs is denoted by L ( α ) :
S ( α ) B ( α ) L ( α ) .
In the DC-HFB method [30], the generalized density matrix of Equation (2) is approximated as the sum of subsystem contributions:
D ¯ μ ν D ¯ μ ν DC = α P μ ν α D ¯ μ ν α ,
K ¯ μ ν K ¯ μ ν DC = α P μ ν α K ¯ μ ν α ,
where P α is the partition matrix defined by
P μ ν α = 1 ( μ S ( α ) ν S ( α ) ) 1 / 2 ( μ S ( α ) ν B ( α ) , or vice versa ) 0 ( otherwise ) .
Using the submatrices of F ˜ and Δ for L ( α ) , denoted by F ˜ α and Δ α , the subsystem density and pairing matrices are obtained with the subsystem quasiparticle orbitals obtained as the eigenvectors of the following subsystem HFB equation:
F ˜ α Δ α Δ α F ˜ α Y i α X i α = S α 0 0 S α Y i α X i α ε i α ,
as
D ¯ μ ν α = i M α Y μ i α Y ν i α ,
K ¯ μ ν α = i M α Y μ i α X ν i α = i M α X μ i α Y ν i α ,
with the number of quasiparticle orbitals in subysystem α , M α . Another expression that is consistent with the DC-HF method with a finite electronic temperature of 1 / ( k B β ) is the following:
D ¯ μ ν α = p f β ε p α Y μ p α Y ν p α ,
K ¯ μ ν α = p f β ε p α Y μ p α X ν p α = p f β ε p α X μ p α Y ν p α ,
where f β ( x ) = 1 + exp ( β x ) 1 is a Fermi distribution function. Note that the summation over p runs all the quasiparticle orbitals regardless of the sign of the eigenvalues.

2.2. DC-HFB Energy Gradient with Respect to Atomic Coordinate

According to ref. [32], the differentiation of Equation (1) with respect to atomic coordinate Q gives
E HFB Q = 2 μ ν ( h μ ν Λ S μ ν ) Q D ¯ ν μ + μ ν λ ρ ( 2 μ ν | λ ρ μ ν | ρ λ ) Q D ¯ λ μ D ¯ ρ ν ζ μ ν λ ρ μ ν | λ ρ Q K ¯ μ ν K ¯ λ ρ + 2 μ ν ( h μ ν Λ S μ ν ) D ¯ ν μ Q + 2 μ ν λ ρ ( 2 μ ν | λ ρ μ ν | ρ λ ) D ¯ λ μ Q D ¯ ρ ν 2 ζ μ ν λ ρ μ ν | λ ρ K ¯ μ ν Q K ¯ λ ρ + 2 Λ Q N μ ν S μ ν D ¯ ν μ = G H F Q + 2 tr F ˜ D ¯ Q + Δ K ¯ Q .
The first term, G H F Q , represents the Hellmann–Feynman (H–F) force that relates to the derivatives of the Hamiltonian matrix elements, and the second term is the so-called Pulay force. For a variational solution, the evaluation of D ¯ and K ¯ derivatives with respect to Q can be avoided by using the idempotency condition of the generalized density matrix, R ¯ , as [32]
2 tr F ˜ D ¯ Q + Δ K ¯ Q = 2 tr W ¯ S Q ,
where W ¯ is an energy-weighted density matrix defined by
W ¯ μ ν = i M ε i Y μ i Y ν i .
Note that the eigenvalues of Equation (6) are { ε i } .
Similarly to the DC-HF energy gradient [36], the Pulay force of the DC-HFB energy gradient cannot be simplified to the form of Equation (23) because the DC-HFB density and pairing matrices are not variationally determined. However, as Yang and Lee [27] proposed, the Pulay force can be approximated in a similar manner to the DC density and pairing matrices of Equations (14) and (15) as
W ¯ μ ν W ¯ μ ν DC = α P μ ν α W ¯ μ ν α ,
W ¯ μ ν α = i M α ε i α Y μ i α Y ν i α ,
Therefore, the DC-HFB energy gradient can be approximated as the following:
E DC - HFB Q G H F Q 2 tr W ¯ DC S Q .
According to ref. [36], this formula is referred to as the YL formula throughout this paper.
Alternatively, the Pulay force in the DC-HFB energy can be reformulated as the following:
2 tr F ˜ D ¯ DC Q + Δ K ¯ DC Q = 2 α μ S ( α ) F ˜ α D ¯ α Q + Δ α K ¯ α Q μ μ .
By analogy with the formulation in ref. [36], we assume here that the subsystem’s generalized density matrix,
R ¯ α = D ¯ α K ¯ α K ¯ α ( S α ) 1 D ¯ α ,
also satisfies the idempotency condition, leading to
D ¯ α S α D ¯ α D ¯ α = K ¯ α S α K ¯ α ,
D ¯ α S α K ¯ α = K ¯ α S α D ¯ α .
Furthermore, we regard that the subsystem’s generalized density matrices, { R ¯ α } , are variationally determined. Then, any simultaneous solution of the derivatives of Equations (30) and (31),
D ¯ α Q S α D ¯ α + D ¯ α S α Q D ¯ α + D ¯ α S α D ¯ α Q D ¯ α Q = K ¯ α Q S α K ¯ α K ¯ α S α Q K ¯ α K ¯ α S α K ¯ α Q ,
D ¯ α Q S α K ¯ α + D ¯ α S α Q K ¯ α + D ¯ α S α K ¯ α Q = K ¯ α Q S α D ¯ α + K ¯ α S α Q D ¯ α + K ¯ α S α D ¯ α Q ,
gives the same force [37]. A simple set of solutions is the following [32]:
D ¯ α Q = K ¯ α S α Q K ¯ α D ¯ α S α Q D ¯ α .
K ¯ α Q = K ¯ α S α Q D ¯ α D ¯ α S α Q K ¯ α .
Applying Equations (34) and (35) to the Pulay force of Equation (28) and considering the subsystem HFB equation of Equation (17) leads to
2 α μ S ( α ) F ˜ α D ¯ α Q + Δ α K ¯ α Q μ μ = 2 α μ S ( α ) S α W ¯ α S α Q D ¯ α + S α V ¯ α S α Q K ¯ α μ μ ,
where V ¯ α is the subsystem energy-weighted pairing matrix defined by
V ¯ μ ν α = i M α ε i α X μ i α Y ν i α .
Finally, the following approximated DC-HFB energy gradient can be obtained:
E DC - HFB Q G H F Q 2 tr S Q α A α + B α ,
where
A α = D ¯ α [ L ( α ) × S ( α ) ] S α [ S ( α ) × L ( α ) ] W ¯ α ,
B α = K ¯ α [ L ( α ) × S ( α ) ] S α [ S ( α ) × L ( α ) ] V ¯ α .
Here, the superscript [ L ( α ) × S ( α ) ] represents the submatrix whose row and column indices are L ( α ) and S ( α ) , respectively. This formula is referred to as the KKASN formula [36] throughout this paper. Similarly to the DC-HF energy gradient [36], the A α and B α matrices in Equation (38) can be Hermitized.
In the case of the DC-HF energy gradient [36], the KKASN gradient generally performs better than the YL one. Namely, the deviation in the gradient values from the standard HF method was smaller, and the optimized geometry was better reproduced in the KKASN method than in the YL method.

3. Numerical Application to Graphene Nanoribbons

A single-layer graphene sheet shows characteristic electronic properties, such as the fractional quantum Hall effect [38] and ultrafast carrier mobility [39], due to its special band structure derived from the electronic structure confined in a plane. Graphene-based spintronics devices are also expected to be developed [39,40], although a band gap has to be generated for realizing these devices. GNRs were theoretically predicted to generate a band gap [41,42], and this has also been experimentally reported. Since the electronic properties of GNRs depend on geometrical factors such as width, length, and edge shape, it is essential to understand their relation with the electronic structure in order to design GNR-based nanodevices. To obtain the correct electronic structures of GNRs, known as strongly correlated systems, high-level ab initio multireference methods such as the density matrix renormalization group (DMRG) [4,43] and the second-order reduced density matrix-based variational method [44] are required. However, due to their large computational demands, their applications to larger GNRs are difficult. In addition, at the current stage, geometry optimization based on these high-level ab initio methods is not tractable.
Here, we performed geometry optimizations of GNRs with the DC-HFB energy gradient and investigated the polyradical character of GNRs from the HFB natural orbital occupation. After a (DC-)HFB calculation, natural orbitals can be obtained by diagonalizing the density matrix, D ¯ . The eigenvalues correspond to the natural orbital occupation numbers (NOONs), which have been discussed as indicators of diradical character in UHF calculations [45]. For single-reference systems, the NOON is either 0 (unoccupied) or 2 (two-electron-occupied), while for singlet diradical systems with N e electrons, the highest occupied (i.e., N e 2 -th) and lowest unoccupied (i.e., ( N e 2 + 1 ) -th) natural orbitals (HONO and LUNO) are degenerate, resulting in NOON being 1 [4]. Throughout this paper, the modified version of the Gamess program package [46,47] was used for computations of the standard and DC-HFB energy and gradient.

3.1. Determination of Parameter ζ

Before applying the HFB method to GNRs, the scaling factor, ζ , introduced in Equation (1) has to be fixed. Here, this parameter was determined by comparing the optimized geometries of phenalenyl cation, anthracene, and phenanthrene (Figure 1) with those by reference all- π CASSCF calculations, which were performed with the OpenMolcas 24.10 program [48]. No dynamical correlation correction was introduced to the CASSCF reference, since the HFB method basically cannot account for dynamical correlation. Throughout this paper, the 6-31G** basis set [49,50] was adopted unless otherwise noted. The obtained CASSCF geometries for phenalenyl cation, anthracene, and phenanthrene have D 3 h , D 2 h , and C 2 v symmetries, respectively, the stabilities of which are confirmed by frequency analysis.
Table 1 and Table 2 summarize the optimized C–C and C–H bond lengths, respectively, for these molecules obtained with the CASSCF and HFB methods. For HFB calculations, the parameter ζ is varied between 0.75 and 1.0. The mean absolute deviations (MADs) of the HFB results from the CASSCF reference results are also listed. In these systems, HFB calculations with ζ = 0.75 converge to the restricted HF (RHF) results, i.e., completely closed-shell solutions. If ζ increases, the obtained result gradually deviates from the RHF solution, leading to partially open-shell solutions. The deviations in the C–C bond lengths are one order (or more) larger than those of the C–H lengths. The HFB results with ζ = 1.0 include obviously larger deviations than the others. For both the C–C and C–H lengths, the HFB results with ζ = 0.85 gives the smallest MADs from the CASSCF ones. Hereafter, ζ = 0.85 was adopted. We also adopted the triple-zeta quality 6-311G(2df,2p) basis set, where the obtained geometries are given in Table S1 in the Supplementary Materials. The difference in C–C bond length between the 6-31G** and 6-311G(2df,2p) basis sets was sufficiently small, with a maximum of 0.007 Å and an average of 0.003 Å.

3.2. Polyradicality of Graphene Nanoribbons

HFB calculations were performed for GNRs of different widths (W) and lengths (L) to investigate the effect of edge structure on polyradicality. The width W and length L are defined as shown in Figure 2. For this sample molecule, we denote { L , W } = { 4 , 2 } . NOONs for (HONO 2 ) to (LUNO + 2 ) were examined as a measure of polyradicality. The geometries were also optimized at the HFB/6-31G** level of theory.
First, we fixed the width W = 2 and varied the zigzag edge length L = 2 8 . The NOONs for six frontier natural orbitals are summarized in Figure 3. The occupation number difference between HONO and LUNO rapidly decreases as L increases, indicating that the diradical character of the system is increasing along L. In addition, the differences between (HONO 1 )/(HONO 2 ) and (LUNO + 1 )/(LUNO + 2 ) also decrease gradually, indicating the increase in tetra- and hexaradical characters in these systems.
Next, we fixed the length L = 2 and varied the armchair edge width W = 1 7 , the results of which are depicted in Figure 4. Again, the occupation number difference between HONO and LUNO decreases as W increases, but the rate of decrease is slower than the W = 2 result in Figure 3. The reason can be deduced from the shapes of HONO and LUNO. The HONOs and LUNOs of the GNRs with { L , W } = { 8 , 2 } and { L , W } = { 3 , 4 } are illustrated in Figure 5 and Figure 6, respectively. Since the occupation numbers for these NOs are ∼1, the electron distribution represented by these NOs corresponds to the distribution of the singlet diradical. It can be confirmed that radical electrons are mainly localized on zigzag edges. Therefore, diradical character is enhanced when the zigzag edge is elongated compared to when the armchair edge is elongated.
Figure 7a shows the HONO-LUNO occupation number difference of GNRs plotted against the number of carbon atoms. For the same number of carbon atoms, the W = 1 series, in which the zigzag edge increases with the smallest armchair edge fraction, has the highest diradical character. Conversely, the L = 2 series, in which the armchair edge increases with the smallest zigzag edge fraction, has the lowest diradical character. The difference in the occupation numbers between HONO 1 and LUNO + 1 , which is an indicator of tetraradicality, is shown in Figure 7b, and a similar trend is seen especially in W-constant series. In the L-constant series, if the number of carbon atoms is taken as the horizontal axis, the L = 2 and L = 3 series are reversed in the middle. This is probably due to the fact that radicals tend to localize at the armchair edge as opposed to the center of the GNR.

3.3. DC-HFB Optimization of GNR Structures

In this subsection, the geometry optimization of the GNR with { L , W } = { 15 , 1 } was performed by the standard and DC-HFB method. In the DC-HFB calculations, two different gradient formalisms, namely KKASN and YL formulas, as well as two different fragmentation schemes, namely type A and B, as shown in Figure 8, were investigated.
First, the fragmentation scheme was set to type A. Figure 9 shows the optimized C–C bond lengths at the zigzag edge with respect to the sequential position. In the DC calculations, up to n b -th nearest fragments were included in the buffer region. Although an alternating single- and double-bond structure can be found around the end points, the bond alternation rapidly decreases around the center. For DC with n b = 3 , the structure obtained with the YL formula significantly differs from the standard one. The difference significantly diminishes when adopting the KKASN formula. The mean absolute deviations (MADs) from the standard HFB result are 0.0246  Å and 0.0043  Å for the YL and KKASN formulas, respectively. As the buffer size increases to n b = 4 , the structure deviation decreases by about one order of magnitude both for the KKASN and YL formulas. The MADs are 0.0040  Å and 0.0004  Å, respectively.
Figure 10 shows the energy convergence process in the geometry optimization of GNR with { L , W } = { 15 , 1 } . The initial structure was set with uniform bond lengths and angles, namely, 1.4  Å, 1.1  Å, and 120° for C–C bond lengths, C–H bond lengths, and bond angles, respectively. For the DC-HFB results with n b = 4 , the energy convergence process is almost equivalent to the standard HFB one. The deviations in energy at the optimized structure are −1.12 mEh and −0.07 mEh for the KKASN and YL formulas, respectively. At first glance, the YL formula seems to produce better results than the KKASN one, but the energy difference between the DC and standard HFB results for the initial structure is −1.14 mEh, so this is due to the error cancellation caused by the slightly poorer structure by the YL formula. For the combination of a smaller buffer size of n b = 3 and the YL formula, the energy gradually increased from the third iteration and converged to a higher energy structure after more than 20 iterations. However, when combined with the KKASN formula, the energy converged as smoothly as the n b = 4 case, while the energy is 28.77 mEh lower than the standard HFB result.
Finally, dependence on the fragmentation scheme was investigated. Figure 11 shows the optimized C–C bond lengths at the zigzag edge with respect to the sequential position. The type A results are equivalent to those given in Figure 9. Especially for a small buffer size of n b = 3 , the trend is different between different fragmentation schemes; namely, type A fragmentation emphasizes the bond alternation compared to type B fragmentation. However, for a large buffer size of n b = 4 , the geometry is almost identical. For type B fragmentation, the MADs from the standard HFB result are 0.0026  Å and 0.0001  Å for n b = 3 and 4, respectively, which are slightly smaller than those for type A fragmentation.
For reference, the computational times for the single-point gradient calculation of GNR with { L , W } = { 15 , 1 } are summarized in Table 3. The calculations were performed on a single node equiped with two Intel Xeon Gold 5118 processors. The standard (non-DC) HFB took more than 10 times longer time than the standard UHF. The first reason for the lengthening of the calculation is that the dimension of the Hamiltonian matrix is doubled, as is seen in Equation (6). In addition, the degradation of the SCF convergence significantly increased the computational time. However, by applying the DC method, we were able to reduce the calculation time to about one-third. Furthermore, the acceleration ratio will be increased for larger systems.

4. Concluding Remarks

This study presented the divide-and-conquer Hartree–Fock–Bogoliubov (DC-HFB) method as an effective approach for the treatment of static electron correlation in large molecular systems. The analysis of natural orbitals obtained from the HFB denstiy matrix revealed that GNRs exhibit strong polyradical character, particularly in structures with extended zigzag edges, which is consistent with previous computational reports. The two proposed approximate energy gradient formulations, specifically the YL and KKASN methods, demonstrated utility in the geometry optimization of graphene nanoribbons (GNRs). In general, the KKASN method gives a more accurate gradient and geometry than the YL method. These results highlight the DC-HFB method’s scalability and accuracy for large-scale systems, making it a promising tool for quantum chemical calculations in strongly correlated systems. As mentioned in the introduction, incorporating dynamical electron correlation is required to further enhance the quantitative accuracy of the method. A simple scheme to incorporate dynamical electron correlation is to combine the HFB method with the Kohn–Sham density functional theory (DFT), which has already been accomplished by Tsuchimochi et al. [51]. Its DC-based linear-scaling implementation as well as its energy gradient formulation will be straightforward, since the DC method was originally proposed in DFT.
Recently, we have proposed a time-dependent (TD) HFB method for the excited-state calculation of molecular systems [52]. Since the real-time propagation scheme is adopted in this excited-state calculation method, a DC-TD-HFB formulation will be possible, opening up the excited-state calculation of large-scale statically correlated systems. In addition, the HFB wavefunction (and of course the DC-HFB method) suffers from the contamination of different-particle-number wavefunctions, which is related to the spin contamination in the UHF wavefunction. The projected HFB or UHF method, which extracts the desired particle-number or spin states, respectively, from the contaminated wavefunction, has been proposed [26,53]. The application of the DC scheme to these projected HF theories is currently underway and will be presented in the near future.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/chemistry7020046/s1, Table S1: Optimized C–C bond lengths (in Å) of phenalenyl cation, anthracene, and phenanthrene obtained by HFB ( ζ = 0.85 ) calculations with 6-31G** and 6-311G(2df,2p) basis sets.

Author Contributions

Conceptualization, M.K.; methodology, M.K., R.K. and T.A.; software, M.K. and R.K.; investigation, M.K.; writing—original draft preparation, M.K.; writing—review and editing, M.K., T.A. and T.T.; project administration, M.K.; funding acquisition, M.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by JSPS KAKENHI Grant Numbers JP25810011, JP15H00908, JP17K05738, and JP23H04093, by MEXT as “Program for Promoting Research on the Supercomputer Fugaku” Grant Number JPMXP1020230325.

Data Availability Statement

The original data presented in the study are openly available in github at https://github.com/k-masato/DC-HFB_GNR/, accessed on 28 January 2025. Further inquiries can be directed to the corresponding author.

Acknowledgments

Some of the present calculations were performed using the computer facilities at Research Center for Computational Science, Okazaki (Project: 24-IMS-C017), and at Research Institute for Information Technology, Kyushu University, Japan. During the preparation of this manuscript, the authors used ChatGPT for writing the draft of abstract and DeepL for proofreading the English. The authors have reviewed and edited the output and take full responsibility for the content of this publication. The Institute for Chemical Reaction Design and Discovery (ICReDD) was established by the World Premier International Research Initiative (WPI), MEXT, Japan.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Houk, K.N.; Lee, P.S.; Nendel, M. Polyacene and Cyclacene Geometries and Electronic Structures: Bond Equalization, Vanishing Band Gaps, and Triplet Ground States Contrast with Polyacetylene. J. Org. Chem. 2001, 66, 5517–5521. [Google Scholar] [CrossRef]
  2. Tian, C.; Miao, W.; Zhao, L.; Wang, J. Graphene nanoribbons: Current status and challenges as quasi-one-dimensional nanomaterials. Rev. Phys. 2023, 10, 100082. [Google Scholar] [CrossRef]
  3. Caetano, E.W.S.; Freire, V.N.; dos Santos, S.G.; Galvão, D.S.; Sato, F. Möbius and twisted graphene nanoribbons: Stability, geometry, and electronic properties. J. Chem. Phys. 2008, 128, 164719. [Google Scholar] [CrossRef]
  4. Mizukami, W.; Kurashige, Y.; Yanai, T. More π Electrons Make a Difference: Emergence of Many Radicals on Graphene Nanoribbons Studied by Ab Initio DMRG Theory. J. Chem. Theory Comput. 2012, 9, 401–407. [Google Scholar] [CrossRef]
  5. Hayes, E.F.; Siu, A.K.Q. Electronic structure of the open forms of three-membered rings. J. Am. Chem. Soc. 1971, 93, 2090–2091. [Google Scholar] [CrossRef]
  6. Head-Gordon, M. Characterizing unpaired electrons from the one-particle density matrix. Chem. Phys. Lett. 2003, 372, 508–511. [Google Scholar] [CrossRef]
  7. Nakano, M.; Fukui, H.; Minami, T.; Yoneda, K.; Shigeta, Y.; Kishi, R.; Champagne, B.; Botek, E.; Kubo, T.; Ohta, K.; et al. (Hyper)polarizability density analysis for open-shell molecular systems based on natural orbitals and occupation numbers. Theor. Chem. Accounts 2011, 130, 711–724. [Google Scholar] [CrossRef]
  8. Nakano, M. Excitation Energies and Properties of Open-Shell Singlet Molecules: Applications to a New Class of Molecules for Nonlinear Optics and Singlet Fission; Springer International Publishing: Cham, Switzerland, 2014. [Google Scholar] [CrossRef]
  9. Nakano, M. Open-Shell-Character-Based Molecular Design Principles: Applications to Nonlinear Optics and Singlet Fission. Chem. Rec. 2016, 17, 27–62. [Google Scholar] [CrossRef]
  10. Nakano, M.; Champagne, B. Nonlinear optical properties in open-shell molecular systems. WIREs Comput. Mol. Sci. 2016, 6, 198–210. [Google Scholar] [CrossRef]
  11. Roos, B.O. The Complete Active Space Self-Consistent Field Method and Its Applications in Electronic Structure Calculations; Wiley Blackwell (John Wiley & Sons): Hoboken, NJ, USA, 1987; pp. 399–445. [Google Scholar] [CrossRef]
  12. Zgid, D.; Nooijen, M. The density matrix renormalization group self-consistent field method: Orbital optimization with the density matrix renormalization group method in the active space. J. Chem. Phys. 2008, 128, 144116. [Google Scholar] [CrossRef]
  13. Nakatani, N.; Guo, S. Density matrix renormalization group (DMRG) method as a common tool for large active-space CASSCF/CASPT2 calculations. J. Chem. Phys. 2017, 146, 094102. [Google Scholar] [CrossRef]
  14. 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. [Google Scholar] [CrossRef] [PubMed]
  15. Baiardi, A.; Reiher, M. The density matrix renormalization group in chemistry and molecular physics: Recent developments and new challenges. J. Chem. Phys. 2020, 152, 040903. [Google Scholar] [CrossRef] [PubMed]
  16. Surján, P.R. An Introduction to the Theory of Geminals. In Correlation and Localization; Topics in Current Chemistry; Surján, P.R., Ed.; Springer: Berlin/Heidelberg, Germany, 1999; Volume 203, pp. 63–88. [Google Scholar] [CrossRef]
  17. Limacher, P.A.; Ayers, P.W.; Johnson, P.A.; De Baerdemacker, S.; Van Neck, D.; Bultinck, P. A New Mean-Field Method Suitable for Strongly Correlated Electrons: Computationally Facile Antisymmetric Products of Nonorthogonal Geminals. J. Chem. Theory Comput. 2013, 9, 1394–1401. [Google Scholar] [CrossRef]
  18. Tarumi, M.; Kobayashi, M.; Nakai, H. Accelerating convergence in the antisymmetric product of strongly orthogonal geminals method. Int. J. Quantum Chem. 2013, 113, 239–244. [Google Scholar] [CrossRef]
  19. 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, NY, USA, 1977; pp. 79–127. [Google Scholar] [CrossRef]
  20. Kobayashi, M.; Szabados, Á.; Nakai, H.; Surján, P.R. Generalized Møller-Plesset Partitioning in Multiconfiguration Perturbation Theory. J. Chem. Theory Comput. 2010, 6, 2024–2033. [Google Scholar] [CrossRef]
  21. Tarumi, M.; Kobayashi, M.; Nakai, H. Generalized Møller-Plesset Multiconfiguration Perturbation Theory Applied to an Open-Shell Antisymmetric Product of Strongly Orthogonal Geminals Reference Wave Function. J. Chem. Theory Comput. 2012, 8, 4330–4335. [Google Scholar] [CrossRef]
  22. Wang, Q.; Duan, M.; Xu, E.; Zou, J.; Li, S. Describing Strong Correlation with Block-Correlated Coupled Cluster Theory. J. Phys. Chem. Lett. 2020, 11, 7536–7543. [Google Scholar] [CrossRef]
  23. Henderson, T.M.; Bulik, I.W.; Scuseria, G.E. Pair extended coupled cluster doubles. J. Chem. Phys. 2015, 142, 214116. [Google Scholar] [CrossRef]
  24. Lehtola, S.; Head-Gordon, M. Coupled-cluster pairing models for radicals with strong correlations. arXiv 2025. [Google Scholar] [CrossRef]
  25. Staroverov, V.N.; Scuseria, G.E. Optimization of density matrix functionals by the Hartree-Fock-Bogoliubov method. J. Chem. Phys. 2002, 117, 11107–11112. [Google Scholar] [CrossRef]
  26. 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. [Google Scholar] [CrossRef] [PubMed]
  27. Yang, W.; Lee, T.S. A density-matrix divide-and-conquer approach for electronic structure calculations of large molecules. J. Chem. Phys. 1995, 103, 5674–5678. [Google Scholar] [CrossRef]
  28. Nakai, H.; Kobayashi, M.; Yoshikawa, T.; Seino, J.; Ikabata, Y.; Nishimura, Y. Divide-and-Conquer Linear-Scaling Quantum Chemical Computations. J. Phys. Chem. A 2023, 127, 589–618. [Google Scholar] [CrossRef]
  29. Kobayashi, M.; Nakai, H. Divide-and-conquer approaches to quantum chemistry: Theory and implementation. In Linear-Scaling Techniques in Computational Chemistry and Physics: Methods and Applications; Challenges and Advances in Computational Chemistry and Physics; Chapter 5; Papadopoulos, M.G., Zalesny, R., Mezey, P.G., Leszczynski, J., Eds.; Springer: Dordrecht, The Netherlands, 2011; Volume 13, pp. 97–127. [Google Scholar] [CrossRef]
  30. Kobayashi, M.; Taketsugu, T. Divide-and-Conquer Hartree-Focko-Bogoliubov Method and Its Application to Conjugated Diradical Systems. Chem. Lett. 2016, 45, 1268–1270. [Google Scholar] [CrossRef]
  31. Pulay, P. Analytical derivatives, forces, force constants, molecular geometries, and related response properties in electronic structure theory. WIREs Comput. Mol. Sci. 2014, 4, 169–181. [Google Scholar] [CrossRef]
  32. Kobayashi, M. Gradient of molecular Hartree–Fock–Bogoliubov energy with a linear combination of atomic orbital quasiparticle wave functions. J. Chem. Phys. 2014, 140, 084115. [Google Scholar] [CrossRef]
  33. Chai, J.D. Density functional theory with fractional orbital occupations. J. Chem. Phys. 2012, 136, 154104. [Google Scholar] [CrossRef]
  34. Wu, C.S.; Chai, J.D. Electronic Properties of Zigzag Graphene Nanoribbons Studied by TAO-DFT. J. Chem. Theory Comput. 2015, 11, 2003–2011. [Google Scholar] [CrossRef]
  35. Chen, B.J.; Chai, J.D. TAO-DFT fictitious temperature made simple. RSC Adv. 2022, 12, 12193–12210. [Google Scholar] [CrossRef]
  36. Kobayashi, M.; Kunisada, T.; Akama, T.; Sakura, D.; Nakai, H. Reconsidering an analytical gradient expression within a divide-and-conquer self-consistent field approach: Exact formula and its approximate treatment. J. Chem. Phys. 2011, 134, 034105. [Google Scholar] [CrossRef] [PubMed]
  37. Pulay, P. Ab initio calculation of force constants and equilibrium geometries in polyatomic molecules. I. Theory. Mol. Phys. 1969, 17, 197–204. [Google Scholar] [CrossRef]
  38. Bolotin, K.I.; Ghahari, F.; Shulman, M.D.; Stormer, H.L.; Kim, P. Observation of the fractional quantum Hall effect in graphene. Nature 2009, 462, 196–199. [Google Scholar] [CrossRef] [PubMed]
  39. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Katsnelson, M.I.; Grigorieva, I.V.; Dubonos, S.V.; Firsov, A.A. Two-dimensional gas of massless Dirac fermions in graphene. Nature 2005, 438, 197–200. [Google Scholar] [CrossRef]
  40. Zhang, Y.; Tan, Y.W.; Stormer, H.L.; Kim, P. Experimental observation of the quantum Hall effect and Berry’s phase in graphene. Nature 2005, 438, 201–204. [Google Scholar] [CrossRef]
  41. Son, Y.W.; Cohen, M.L.; Louie, S.G. Energy Gaps in Graphene Nanoribbons. Phys. Rev. Lett. 2006, 97, 216803. [Google Scholar] [CrossRef]
  42. Barone, V.; Hod, O.; Scuseria, G.E. Electronic Structure and Stability of Semiconducting Graphene Nanoribbons. Nano Lett. 2006, 6, 2748–2754. [Google Scholar] [CrossRef]
  43. Hachmann, J.; Dorando, J.J.; Avilés, M.; Chan, G.K.L. The radical character of the acenes: A density matrix renormalization group study. J. Chem. Phys. 2007, 127, 134309. [Google Scholar] [CrossRef]
  44. Pelzer, K.; Greenman, L.; Gidofalvi, G.; Mazziotti, D.A. Strong Correlation in Acene Sheets from the Active-Space Variational Two-Electron Reduced Density Matrix Method: Effects of Symmetry and Size. J. Phys. Chem. A 2011, 115, 5632–5640. [Google Scholar] [CrossRef]
  45. Pulay, P.; Hamilton, T.P. UHF natural orbitals for defining and starting MC-SCF calculations. J. Chem. Phys. 1988, 88, 4926–4933. [Google Scholar] [CrossRef]
  46. Schmidt, M.W.; Baldridge, K.K.; Boatz, J.A.; Elbert, S.T.; Gordon, M.S.; Jensen, J.H.; Koseki, S.; Matsunaga, N.; Nguyen, K.A.; Su, S.; et al. General atomic and molecular electronic structure system. J. Comput. Chem. 1993, 14, 1347–1363. [Google Scholar] [CrossRef]
  47. Gordon, M.S.; Schmidt, M.W. Advances in electronic structure theory: GAMESS a decade later. In Theory and Applications of Computational Chemistry: The First Forty Years; Chapter 41; Dykstra, C.E., Frenking, G., Kim, K.S., Scuseria, G.E., Eds.; Elsevier: Amsterdam, The Netherlands, 2005; pp. 1167–1189. [Google Scholar]
  48. Li Manni, G.; Galván, I.F.; Alavi, A.; Aleotti, F.; Aquilante, F.; Autschbach, J.; Avagliano, D.; Baiardi, A.; Bao, J.J.; Battaglia, S.; et al. The OpenMolcas Web: A Community-Driven Approach to Advancing Computational Chemistry. J. Chem. Theory Comput. 2023, 19, 6933–6991. [Google Scholar] [CrossRef] [PubMed]
  49. Hehre, W.J.; Ditchfield, R.; Pople, J.A. Self-consistent molecular orbital methods. XII. Further extensions of Gaussian-type basis sets for use in molecular orbital studies of organic molecules. J. Chem. Phys. 1972, 56, 2257. [Google Scholar] [CrossRef]
  50. Hariharan, P.C.; Pople, J.A. The influence of polarization functions on molecular orbital hydrogenation energies. Theor. Chim. Acta 1973, 28, 213–222. [Google Scholar] [CrossRef]
  51. Tsuchimochi, T.; Scuseria, G.E.; Savin, A. Constrained-pairing mean-field theory. III. Inclusion of density functional exchange and correlation effects via alternative densities. J. Chem. Phys. 2010, 132, 024111. [Google Scholar] [CrossRef]
  52. Nishida, M.; Akama, T.; Kobayashi, M.; Taketsugu, T. Time-dependent Hartree–Fock–Bogoliubov method for molecular systems: An alternative excited-state methodology including static electron correlation. Chem. Phys. Lett. 2023, 816, 140386. [Google Scholar] [CrossRef]
  53. Jiménez-Hoyos, C.A.; Henderson, T.M.; Tsuchimochi, T.; Scuseria, G.E. Projected Hartree–Fock theory. J. Chem. Phys. 2012, 136, 164109. [Google Scholar] [CrossRef]
Figure 1. Structures of phenalenyl cation, anthracene, and phenanthrene.
Figure 1. Structures of phenalenyl cation, anthracene, and phenanthrene.
Chemistry 07 00046 g001
Figure 2. Structure of GNRs and definitions of length L and width W.
Figure 2. Structure of GNRs and definitions of length L and width W.
Chemistry 07 00046 g002
Figure 3. NOONs for (HONO 2 ) to (LUNO + 2 ) of GNRs with { L , W } = { 2 8 , 2 } .
Figure 3. NOONs for (HONO 2 ) to (LUNO + 2 ) of GNRs with { L , W } = { 2 8 , 2 } .
Chemistry 07 00046 g003
Figure 4. NOONs for (HONO 2 ) to (LUNO + 2 ) of GNRs with { L , W } = { 2 , 1 7 } .
Figure 4. NOONs for (HONO 2 ) to (LUNO + 2 ) of GNRs with { L , W } = { 2 , 1 7 } .
Chemistry 07 00046 g004
Figure 5. (a) HONO and (b) LUNO obtained from the HFB calculation of GNR with { L , W } = { 8 , 2 } . The value in parentheses is the occupation number.
Figure 5. (a) HONO and (b) LUNO obtained from the HFB calculation of GNR with { L , W } = { 8 , 2 } . The value in parentheses is the occupation number.
Chemistry 07 00046 g005
Figure 6. (a) HONO and (b) LUNO obtained from the HFB calculation of GNR with { L , W } = { 3 , 4 } . The value in parentheses is the occupation number.
Figure 6. (a) HONO and (b) LUNO obtained from the HFB calculation of GNR with { L , W } = { 3 , 4 } . The value in parentheses is the occupation number.
Chemistry 07 00046 g006
Figure 7. Occupation number differences of GNRs between (a) HONO and LUNO and (b) (HONO 1 ) and (LUNO + 1 ) against the number of carbon atoms, N C .
Figure 7. Occupation number differences of GNRs between (a) HONO and LUNO and (b) (HONO 1 ) and (LUNO + 1 ) against the number of carbon atoms, N C .
Chemistry 07 00046 g007
Figure 8. Two fragmentation schemes adopted in the DC-HFB calculations of GNR with { L , W } = { 15 , 1 } .
Figure 8. Two fragmentation schemes adopted in the DC-HFB calculations of GNR with { L , W } = { 15 , 1 } .
Chemistry 07 00046 g008
Figure 9. C–C bond lengths of GNR with { L , W } = { 15 , 1 } optimized with DC and standard HFB methods.
Figure 9. C–C bond lengths of GNR with { L , W } = { 15 , 1 } optimized with DC and standard HFB methods.
Chemistry 07 00046 g009
Figure 10. Energy convergence process in the geometry optimization of GNR with { L , W } = { 15 , 1 } in DC and standard HFB calculations.
Figure 10. Energy convergence process in the geometry optimization of GNR with { L , W } = { 15 , 1 } in DC and standard HFB calculations.
Chemistry 07 00046 g010
Figure 11. Dependence of the optimized C–C bond lengths on the fragmentation scheme in the DC-HFB calculations of GNR with { L , W } = { 15 , 1 } .
Figure 11. Dependence of the optimized C–C bond lengths on the fragmentation scheme in the DC-HFB calculations of GNR with { L , W } = { 15 , 1 } .
Chemistry 07 00046 g011
Table 1. Optimized C–C bond lengths (in Å) of phenalenyl cation, anthracene, and phenanthrene obtained with the CASSCF and HFB methods.
Table 1. Optimized C–C bond lengths (in Å) of phenalenyl cation, anthracene, and phenanthrene obtained with the CASSCF and HFB methods.
HFB
MoleculePositionCASSCFζ = 0.750.800.850.901.00
Phenalenyl cationC1–C21.41781.41071.41641.42281.42971.4487
C2–C31.41411.40871.41211.41571.41981.4344
C3–C41.39211.38341.38731.39211.39821.4173
AnthraceneC1–C21.40081.38941.38931.40091.41061.4299
C2–C31.43851.43591.43591.42521.42181.4343
C3–C41.36491.34731.34731.36711.38631.4156
C4–C51.43371.43241.43251.41831.41141.4213
C2–C61.42871.42461.42451.43301.44091.4557
PhenanthreneC1–C21.35571.33861.33861.34791.37301.4080
C2–C31.44241.44051.44051.43561.42691.4329
C3–C41.41661.40841.40841.41001.41631.4341
C4–C51.37801.36561.36561.37071.38561.4134
C5–C61.40901.40171.40171.40221.40481.4188
C6–C71.38021.36771.36771.37301.38861.4165
C3–C81.41271.40431.40431.41271.43241.4570
C7–C81.41921.41061.41061.41191.41771.4370
C8–C91.46341.46121.46121.45761.45201.4589
MAD 0.00800.00730.00570.01130.0251
Table 2. Optimized C–H bond lengths (in Å) of phenalenyl cation, anthracene, and phenanthrene obtained with the CASSCF and HFB methods.
Table 2. Optimized C–H bond lengths (in Å) of phenalenyl cation, anthracene, and phenanthrene obtained with the CASSCF and HFB methods.
HFB
MoleculePositionCASSCFζ = 0.750.800.850.901.00
Phenalenyl cationC3–H11.07521.07581.07531.07521.07551.0791
C4–H21.07341.07311.07361.07411.07471.0784
AnthraceneC1–H11.07691.07681.07681.07681.07731.0818
C3–H21.07631.07621.07621.07631.07671.0806
C4–H31.07561.07561.07561.07571.07601.0796
PhenanthreneC2–H11.07611.07611.07611.07611.07651.0806
C4–H21.07631.07631.07631.07631.07671.0806
C5–H31.07551.07551.07551.07551.07591.0794
C6–H41.07561.07571.07571.07571.07601.0797
C7–H51.07271.07271.07271.07281.07321.0776
MAD 0.00010.00010.00010.00050.0044
Table 3. Wall-clock computational times (in s) for the single-point gradient calculation of GNR with { L , W } = { 15 , 1 } .
Table 3. Wall-clock computational times (in s) for the single-point gradient calculation of GNR with { L , W } = { 15 , 1 } .
MethodTime
Standard UHF232.1
Standrad HFB ( ζ = 0.85 )3765.5
DC-HFB ( ζ = 0.85 , n b = 3 )1396.0
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kobayashi, M.; Kodama, R.; Akama, T.; Taketsugu, T. Fragmentation-Based Linear-Scaling Method for Strongly Correlated Systems: Divide-and-Conquer Hartree–Fock–Bogoliubov Method, Its Energy Gradient, and Applications to Graphene Nano-Ribbon Systems. Chemistry 2025, 7, 46. https://doi.org/10.3390/chemistry7020046

AMA Style

Kobayashi M, Kodama R, Akama T, Taketsugu T. Fragmentation-Based Linear-Scaling Method for Strongly Correlated Systems: Divide-and-Conquer Hartree–Fock–Bogoliubov Method, Its Energy Gradient, and Applications to Graphene Nano-Ribbon Systems. Chemistry. 2025; 7(2):46. https://doi.org/10.3390/chemistry7020046

Chicago/Turabian Style

Kobayashi, Masato, Ryosuke Kodama, Tomoko Akama, and Tetsuya Taketsugu. 2025. "Fragmentation-Based Linear-Scaling Method for Strongly Correlated Systems: Divide-and-Conquer Hartree–Fock–Bogoliubov Method, Its Energy Gradient, and Applications to Graphene Nano-Ribbon Systems" Chemistry 7, no. 2: 46. https://doi.org/10.3390/chemistry7020046

APA Style

Kobayashi, M., Kodama, R., Akama, T., & Taketsugu, T. (2025). Fragmentation-Based Linear-Scaling Method for Strongly Correlated Systems: Divide-and-Conquer Hartree–Fock–Bogoliubov Method, Its Energy Gradient, and Applications to Graphene Nano-Ribbon Systems. Chemistry, 7(2), 46. https://doi.org/10.3390/chemistry7020046

Article Metrics

Back to TopTop