Also at ]Department of Mathematics and Statistics, University of Massachusetts Amherst.
FEAST nonlinear eigenvalue algorithm for quasiparticle equations
Abstract
The use of Green’s function in quantum many-body theory often leads to nonlinear eigenvalue problems, as Green’s function needs to be defined in energy domain. The approximation method is one of the typical examples. In this article, we introduce a method based on the FEAST eigenvalue algorithm for accurately solving the nonlinear eigenvalue quasiparticle equation, eliminating the need for the Kohn-Sham wavefunction approximation. Based on the contour integral method for nonlinear eigenvalue problem, the energy (eigenvalue) domain is extended to complex plane. Hypercomplex number is introduced to the contour deformation calculation of self-energy to carry imaginary parts of both Green’s functions and FEAST quadrature nodes. Calculation results for various molecules are presented and compared with a more conventional graphical solution approximation method. It is confirmed that the Highest Occupied Molecular Orbital (HOMO) from the Kohn-Sham equation is very close to that of , while the Least Unoccupied Molecular Orbital (LUMO) shows noticeable differences.
The approximation has established itself as a cornerstone in the realm of many-body quantum theory, providing a robust framework for the calculation of quasiparticle (QP) properties in a wide range of materials. Originating from the pioneering work of Lars Hedin in 1965 Hedin (1965), the method has undergone extensive development, culminating in a versatile tool for investigating electronic excitations. The method derives its name from the product of the Green’s function () and the screened Coulomb interaction (), which collectively account for the exchange-correlation effects beyond the Hartree-Fock approximation.
The approximation emerged as a response to the limitations of simpler models like Hartree-Fock and density functional theory (DFT) in describing excited states. Hedin’s formalism provided a systematic way to improve upon these methods by incorporating dynamic screening effects. Initially, practical applications of the method were limited by computational constraints. However, advancements in numerical techniques and computational power have facilitated its application to increasingly complex systems. Early applications of were focused on simple semiconductors and insulators Hybertsen and Louie (1986); Godby et al. (1986), demonstrating significant improvements over DFT in predicting band gaps and electronic spectra. The advent of iterative and self-consistent schemes in the 1990s von Barth and Holm (1996); Holm and von Barth (1998); Schöne and Eguiluz (1998), 2000s García-González and Godby (2001); Shishkin and Kresse (2007); Stan et al. (2009) and 2010s Rostgaard et al. (2010); Blase et al. (2011); Caruso et al. (2012, 2013); Marom et al. (2012); Wilhelm et al. (2016); Grumet et al. (2018), notably the ev and sc approaches, marked a significant leap in accuracy and reliability.
In recent years, the approximation has seen substantial advancements in both theoretical formulations and computational implementations. Hybrid methods that combine with other techniques, such as +BSE (Bethe-Salpeter Equation) Salpeter and Bethe (1951); Strinati (1982); Rohlfing and Louie (2000); Hanke and Sham (1980); Komsa and Krasheninnikov (2012); Ramasubramaniam (2012); Hüser et al. (2013); Shi et al. (2013); Qiu et al. (2016); Dvorak and Wu (2015); K/"orbel et al. (2014); Bruneval et al. (2015); Jacquemin et al. (2015) and +DMFT (Dynamical Mean-Field Theory) Georges and Kotliar (1992); Biermann et al. (2003); Kotliar et al. (2006); Biermann (2014); Choi et al. (2016); Nilsson et al. (2017); Zhu and Chan (2021), have expanded the applicability of to more complex systems, including strongly correlated materials. The development of efficient and high-performance computing algorithms has further enhanced the feasibility of calculations for large-scale systems. Gao et al. (2016, 2024); Kim et al. (2020); Shao et al. (2016, 2018)
Despite these advancements, implementations often rely on linearized eigenvalue equations and DFT wavefunction approximations, which may limit their accuracy in certain scenarios. It is claimed that the DFT wavefunctions are often very close to the wavefunctions, but we cannot yet say that it is correct for all cases. Nonlinear eigenvalue algorithm, by contrast, offer a more precise framework for capturing the full extent of QP interactions. The eigenvalue-dependent nonlinear eigenvalue approach directly tackles the energy term inside the self-energy in the QP equations, avoiding approximations that can lead to inaccuracies in QP energies. The formulation of the QP equation as a nonlinear eigenvalue problem involves solving the Dyson equation, where the self-energy depends on the QP energies themselves.
In this work, we provide accurate solutions to the (one-shot ) QP equation, which is the most widely used scheme and the starting point of the self-consistent calculations. Our approach leverages recent advancements in the FEAST algorithm Polizzi (2009); Gavin et al. (2018); Brenneck and Polizzi (2020) to solve the nonlinear eigenvalue QP equation without resorting to linear eigenvalue and DFT wavefunction approximations.
We begin by outlining the theoretical framework of the approximation and its formulation as a nonlinear eigenvalue problem.
non-linear formulation of
The approximation quasiparticle equation in atomic units is given as
(1) |
where , correspond to the external and the Hartree potentials, respectively. The term is the exchange-correlation self-energy which contains the many-body quantum effects information. Eq. (1) is clearly a non-local and non-linear eigenvalue equation since the self-energy term contains all of these two difficult-to-handle features. The self-energy in one-shot approximation is given by the convolution integral of the Green’s function () and the screened Coulomb interaction () in energy domain.
(2) |
An exponential term with a positive infinitesimal is inserted into the integral to make sure that the integral is convergent in the complex plane and physically meaningful. The means mean-field one-body Green’s function that could be obtained by a mean-field Hamiltonian, such as DFT, and is the screened Coulomb interaction which is determined by the mean-field one-body Green’s function again and given by
(3) |
where is bare Coulomb interaction, and the dielectric function is calculated as follow.
(4) |
Using the Randon Phase Approximation (RPA), the irreducible polarizability, , in Eq. (4), is the convolution integral of two mean-field one-body Green’s functions usually with a factor 2 to represent the spin - for the closed shell systems:
(5) |
It can be seen from Eq. (2) to (5) that the self-energy is a energy domain integral that is solely determined by two physical quantities, the bare Coulomb interaction and the Green’s function . Therefore, the construction of the Green’s function and the evaluation of the energy integral are essential for calculations. For Green’s function, typically, the most commonly used method is the eigenfunction decomposition, another way is to construct the mean-field Hamiltonian – , compute the resolvent (inverse) directly and add the remainder imaginary term.
(6) |
where and are the eigenvalues and eigenfunctions of , is the Fermi energy level of the system and could be computed using the occupied orbitals only.
Using eigenfunction decomposition, Eq. (5) can also be equivalently rewritten as the Lehmann representation. Alternatively, Eq. (5) can be solved analytically and expressed by the Green’s function and the occupied orbitals.
(7) |
In practice, exchange and correlation parts of the self-energy are treated separately, , using the following two expressions for the exchange part , and the correlation part :
(8) |
(9) |
It could be noted that the exchange part (8) has an analytical solution and becomes energy independent, leaving only the correlation part (9) energy dependent (for simplicity, the exponential term and the infinity notations will be omitted from now on).
Several different approaches are available to carefully evalaute the correlation component, including: Analytical Continuation (AC) Rojas et al. (1995); Ren et al. (2012a, b); Wilhelm et al. (2016), Contour Deformation (CD) Godby et al. (1988); Gonze et al. (2009); Blase et al. (2011); Govoni and Galli (2015); Golze et al. (2018, 2020), Plasmon-Pole Models (PPM) Hybertsen and Louie (1986); Soininen et al. (2003, 2005); Kas et al. (2007), and fully analytic method using the Casida equation Casida (1995); Tiago and Chelikowsky (2006); Bruneval (2012); Gao and Chelikowsky (2019). In this work, the CD method is utilized as the main way to calculate the energy integral. The CD method converts the integral on real axis into a contour integral minus an integral on imaginary axis, thus avoiding instability on the real axis due to the poles of Green’s functions:
(10) |
The summation notation in the first term of Eq. (10) is sum over all the residues of the poles that lie in the contour, and for the second term, the Gauss quadrature could be used to evaluate the imaginary axis integral.
Now, substitute equation into Eq. (1) , we get
(11) |
Then, after discretization, the nonlinear eigenvalue equation (1) can be written in the following matrix form:
(12) |
There are several ways to discretize and construct the system of matrices Eq. (12), real space finite element method (FEM) is used in this work NES (2022); Kestyn and Polizzi (2020); Li et al. (2024); Li and Polizzi (2024). represents then the FEM symmetric positive definite mass matrix.
One way to approximately solve this nonlinear eigenvalue equation is to use the DFT wavefunctions as the eigenvector solutions, . The eigenvalues can then be found graphically by plotting the two expressions on the left and right hand sides of the equal sign of the following expression:
(13) |
where the intersections are the graphical solutions defined by the set .
solution of the non-linear problem using the FEAST algorithm
Eigenvalue problems in which the coefficient matrices depend nonlinearly on the eigenvalues arise in a variety of applications Betcke et al. (2013); Güttel and Tisseur (2017). Eigenvalues associated with eigenvectors are then solutions of the following general form:
(14) |
which includes the common linear eigenvalue problem as a special case, letting , as well as the broader polynomial case letting . For the QP equation (12), we note that
(15) |
The FEAST algorithm is a contour integral method initially developed to address linear eigenvalue problems Polizzi (2009); FEA (2020). It has since been adapted for polynomial eigenvalue problems Gavin et al. (2018) and, more recently, for general nonlinear eigenvalue problems (14) Brenneck and Polizzi (2020). The latter can be considered an advancement over Beyn’s contour integration method Beyn (2012), as it incorporates several enhancements from the FEAST algorithm, such as residual inverse iterations for improved convergence and mixed-precision arithmetic. Essentially, the FEAST algorithm and other similar contour integral methods involve converting an eigenvalue problem into solving a series of independent linear systems (i.e. ) at particular points in the complex plane. When applied to solving the genereral non-linear system (14), the FEAST algorithm involve performing the following contour integrations along for and :
(16) |
where represents the eigenvector subspace, and regroups the residuals of the eigenpairs located within . In practice, a numerical quadrature is needed to compute the contour integral (16) and the resulting FEAST algorithm is presented in Figure 1. It is important to note the following: (i) while the quality of numerical integration, the number of quadrature nodes, and the size of the search subspace can impact the convergence rate, they do not affect the accuracy of the final solutions; (ii) in practice, using a relatively small number of quadrature nodes (around 10) is generally sufficient to achieve convergence to machine precision with very few iterations; (iii) the linear systems that need to be solved at the complex points are independent of one another and can be solved in parallel; (iv) importantly, these linear systems can be solved with lower accuracy (using lower arithmetic precision or modest relative residuals, such as with iterative methods) without affecting the convergence rate of the eigenpairs to machine precision.
Input: |
Contour containing wanted eigenvalues |
Set of quadrature nodes and weights |
Subspace (random) with |
While not converged for all inside |
Solve for all contour nodes |
Compute |
Compute and |
Perform the QR decomposition |
Compute |
Solve reduced eigenvalue problem |
Update X=qW, noting that |
Form with |
Output: all eigenpairs inside |
Finally, when applying FEAST for solving the QP Eq. (12), it requires replacing the in Eq. (10) with the complex number () at the quadrature nodes. However, due to the nonlinear properties, the process of incorporating FEAST contour integral projection (15) and (16) to (also ) involves the calculations of residues of the poles with the complex energy deviated from the real axis, which could lead to ill-defined of Eq. 10 (also of Eq. (8)). To avoid the complication caused by the confusion between the imaginary parts of FEAST nodes and the imaginary numbers present in (namely the terms and ), we propose handling these two types of imaginary components separately. As a result, hypercomplex number (more precisely, quaternion) should be introduced in the calculations. Moreover, to make the basis of quaternion complete, another dummy imaginary part (, where ) needs to be added. The energy domain of the self-energy is then a hypercomplex space defined by the quaternion:
(17) |
In practice, the domain could be reduced to complex plane for some cases. In the context of considering only real part of the self-energy in complex plane, by taking advantage of the symmetry, only the Green’s function in the second term of Eq. (10) needs the quaternion operations, the first and third imaginary parts will be cancelled out after the integration in the end, only left with the imaginary part of . Stated otherwise, in the context of considering real part of the self-energy, for Eq. (10), the first term is the normal complex number calculation in complex plane smearing out the term, and the real part of is used to determine whether if the poles lie in the contour without the need of considering the imaginary part of . For the second term, after computing the Green’s function in the hypercomplex space , only the real and second imaginary part are taken from the results to turn it back to the complex plane, and multiply it with the real part of . Consequently, Eq. (10) can be rewritten in FEAST algorithm as:
(18) |
where and stand for the real and the imaginary part.
Results and Discussion
Combining the methods mentioned above, we performed calculations on a few molecules. In principle, nonlinear FEAST can solve all the occupied states and the LUMO state at the same time, as long as the contour is chosen appropriately. However, it has to be noted that due to the pronounced multisolution behavior of the core electrons in Golze et al. (2018, 2020), contouring energy states that are lower than HOMO can sometimes cause the calculation results to not converge. This could be fixed by using different mean-field Hamiltonian other than DFT, such as Hartree-Fock. In this article, we take the advantage of FEAST to choose the contour appropriately and solve the HOMO and LUMO states using DFT-PBE as the starting point.
The following two expressions are used to evaluate the solutions of the nonlinear eigenvalue equation:
(19) |
where means energy obtained by graphical solution method using DFT wavefunction approximations, which is equavalent to (13), means energy obtained by nonlinear FEAST algorithm. The only difference is the use of different wavefunctions, while the Hamiltonians keep unchanged. As a result, if the energies are different, it would be due to the DFT wavefunction approximation made in the GW calculations.
@PBE | @PBE | |||
---|---|---|---|---|
HOMO | DFT-PBE | graphical | nonlinear | Deviation |
(eV) | solution | FEAST | ||
He | -15.96 | -23.66 | -23.69 | 0.03 |
H2 | -10.42 | -15.91 | -15.92 | 0.01 |
Ne | -13.44 | -20.75 | -20.78 | 0.03 |
LiH | -4.36 | -6.94 | -6.97 | 0.03 |
Li2 | -2.89 | -5.05 | -5.02 | -0.03 |
CO | -9.54 | -14.08 | -14.09 | 0.01 |
@PBE | @PBE | |||
---|---|---|---|---|
LUMO | DFT-PBE | graphical | nonlinear | Deviation |
(eV) | solution | FEAST | ||
He | 1.18 | 2.06 | 1.91 | 0.15 |
H2 | 0.60 | 2.57 | 1.95 | 0.62 |
Ne | 0.63 | 2.45 | 2.16 | 0.29 |
LiH | -1.12 | 0.47 | 0.32 | 0.15 |
Li2 | -1.24 | -0.03 | -0.09 | 0.05 |
CO | -2.99 | 0.35 | 0.21 | 0.14 |
From the Table 1 and 2, it can be seen that the wavefunctions of DFT and are almost identical at the HOMO level, but the difference is very pronounced at the LUMO level. The deviations in LUMO level are roughly one order larger than that in HOMO level. It is also found that LUMO energy states obtained by nonlinear eigenvalue algorithm are lower than the graphical solutions using DFT wavefunction approximations, which means DFT wavefunction approximations would slightly overestimate the HOMO-LUMO gap. From these results, it could be carefully extrapolated that all the DFT orbitals of occupied states are very close to ’s, but the unoccupied states are different.
To summarize, a nonlinear eigenvalue algorithm for solving QP equations is presented. We utilized the FEAST algorithm, which is capable of handling both linear and nonlinear eigenvalue problems, to solve the QP equation and compute the eigenfunctions. By introducing hypercomplex numbers into the FEAST spectral projection process, we effectively circumvent the complications associated with two imaginary numbers. Additionally, to simplify quaternion calculations, we reduce the eigenvalue domain back to the complex plane, making calculations more accessible and affordable. Our method was applied to several molecules, revealing that the LUMO wavefunctions differ from DFT wavefunctions, which challenges the previously accepted consensus. This discrepancy highlights the importance of considering nonlinear eigenvalue problems in approximations, an area that may often be overlooked within the community but could greatly benefit from the recent progress made in numerical nonlinear algorithms, such as FEAST. We hope our findings will stimulate interest and lead to the calculation of nonlinear eigenvalue equation in future quantum many-body studies, paving the way for more accurate and comprehensive condensed matter physics analyses.
References
- Hedin (1965) L. Hedin, Phys. Rev. 139, A796 (1965).
- Hybertsen and Louie (1986) M. S. Hybertsen and S. G. Louie, Phys. Rev. B 34, 5390 (1986).
- Godby et al. (1986) R. W. Godby, M. Schlüter, and L. J. Sham, Phys. Rev. Lett. 56, 2415 (1986).
- von Barth and Holm (1996) U. von Barth and B. Holm, Phys. Rev. B 54, 8411 (1996).
- Holm and von Barth (1998) B. Holm and U. von Barth, Phys. Rev. B 57, 2108 (1998).
- Schöne and Eguiluz (1998) W.-D. Schöne and A. G. Eguiluz, Phys. Rev. Lett. 81, 1662 (1998).
- García-González and Godby (2001) P. García-González and R. W. Godby, Phys. Rev. B 63, 075112 (2001).
- Shishkin and Kresse (2007) M. Shishkin and G. Kresse, Phys. Rev. B 75, 235102 (2007).
- Stan et al. (2009) A. Stan, N. E. Dahlen, and R. van Leeuwen, The Journal of Chemical Physics 130, 114105 (2009), https://pubs.aip.org/aip/jcp/article-pdf/doi/10.1063/1.3089567/15424617/114105_1_online.pdf .
- Rostgaard et al. (2010) C. Rostgaard, K. W. Jacobsen, and K. S. Thygesen, Phys. Rev. B 81, 085103 (2010).
- Blase et al. (2011) X. Blase, C. Attaccalite, and V. Olevano, Phys. Rev. B 83, 115103 (2011).
- Caruso et al. (2012) F. Caruso, P. Rinke, X. Ren, M. Scheffler, and A. Rubio, Phys. Rev. B 86, 081102 (2012).
- Caruso et al. (2013) F. Caruso, P. Rinke, X. Ren, A. Rubio, and M. Scheffler, Phys. Rev. B 88, 075105 (2013).
- Marom et al. (2012) N. Marom, F. Caruso, X. Ren, O. T. Hofmann, T. Körzdörfer, J. R. Chelikowsky, A. Rubio, M. Scheffler, and P. Rinke, Phys. Rev. B 86, 245127 (2012).
- Wilhelm et al. (2016) J. Wilhelm, M. Del Ben, and J. Hutter, Journal of Chemical Theory and Computation 12, 3623 (2016), pMID: 27348184, https://doi.org/10.1021/acs.jctc.6b00380 .
- Grumet et al. (2018) M. Grumet, P. Liu, M. Kaltak, J. c. v. Klimeš, and G. Kresse, Phys. Rev. B 98, 155143 (2018).
- Salpeter and Bethe (1951) E. E. Salpeter and H. A. Bethe, Phys. Rev. 84, 1232 (1951).
- Strinati (1982) G. Strinati, Phys. Rev. Lett. 49, 1519 (1982).
- Rohlfing and Louie (2000) M. Rohlfing and S. G. Louie, Phys. Rev. B 62, 4927 (2000).
- Hanke and Sham (1980) W. Hanke and L. J. Sham, Phys. Rev. B 21, 4656 (1980).
- Komsa and Krasheninnikov (2012) H.-P. Komsa and A. V. Krasheninnikov, Phys. Rev. B 86, 241201 (2012).
- Ramasubramaniam (2012) A. Ramasubramaniam, Phys. Rev. B 86, 115409 (2012).
- Hüser et al. (2013) F. Hüser, T. Olsen, and K. S. Thygesen, Phys. Rev. B 88, 245309 (2013).
- Shi et al. (2013) H. Shi, H. Pan, Y.-W. Zhang, and B. I. Yakobson, Phys. Rev. B 87, 155304 (2013).
- Qiu et al. (2016) D. Y. Qiu, F. H. da Jornada, and S. G. Louie, Phys. Rev. B 93, 235435 (2016).
- Dvorak and Wu (2015) M. Dvorak and Z. Wu, Phys. Rev. B 92, 035422 (2015).
- K/"orbel et al. (2014) S. K/"orbel, P. Boulanger, I. Duchemin, X. Blase, M. A. L. Marques, and S. Botti, Journal of Chemical Theory and Computation 10, 3934 (2014), pMID: 26588537, https://doi.org/10.1021/ct5003658 .
- Bruneval et al. (2015) F. Bruneval, S. M. Hamed, and J. B. Neaton, The Journal of Chemical Physics 142, 244101 (2015), https://pubs.aip.org/aip/jcp/article-pdf/doi/10.1063/1.4922489/15496098/244101_1_online.pdf .
- Jacquemin et al. (2015) D. Jacquemin, I. Duchemin, and X. Blase, Journal of Chemical Theory and Computation 11, 3290 (2015), pMID: 26207104, https://doi.org/10.1021/acs.jctc.5b00304 .
- Georges and Kotliar (1992) A. Georges and G. Kotliar, Phys. Rev. B 45, 6479 (1992).
- Biermann et al. (2003) S. Biermann, F. Aryasetiawan, and A. Georges, Phys. Rev. Lett. 90, 086402 (2003).
- Kotliar et al. (2006) G. Kotliar, S. Y. Savrasov, K. Haule, V. S. Oudovenko, O. Parcollet, and C. A. Marianetti, Rev. Mod. Phys. 78, 865 (2006).
- Biermann (2014) S. Biermann, J. Phys.: Condens. Matter 26, 173202 (2014).
- Choi et al. (2016) S. Choi, A. Kutepov, K. Haule, M. v. Schilfgaarde, and G. Kotliar, npj Quantum Materials 1, 16001 (2016).
- Nilsson et al. (2017) F. Nilsson, L. Boehnke, P. Werner, and F. Aryasetiawan, Phys. Rev. Mater. 1, 043803 (2017).
- Zhu and Chan (2021) T. Zhu and G. K.-L. Chan, Phys. Rev. X 11, 021006 (2021).
- Gao et al. (2016) W. Gao, W. Xia, X. Gao, and P. Zhang, Scientific Reports 6, 36849 (2016).
- Gao et al. (2024) W. Gao, Z. Tang, J. Zhao, and J. R. Chelikowsky, Phys. Rev. Lett. 132, 126402 (2024).
- Kim et al. (2020) M. Kim, G. J. Martyna, and S. Ismail-Beigi, Phys. Rev. B 101, 035139 (2020).
- Shao et al. (2016) M. Shao, L. Lin, C. Yang, F. Liu, F. H. d. Jornada, J. Deslippe, and S. Louie, Sci. China Math. 59, 1593–1612 (2016).
- Shao et al. (2018) M. Shao, F. H. da Jornada, L. Lin, C. Yang, J. Deslippe, and S. G. Louie, SIAM Journal on Matrix Analysis and Applications 39, 683 (2018), https://doi.org/10.1137/16M1102641 .
- Polizzi (2009) E. Polizzi, Phys. Rev. B 79, 115112 (2009).
- Gavin et al. (2018) B. Gavin, A. Międlar, and E. Polizzi, Journal of Computational Science 27, 107 (2018).
- Brenneck and Polizzi (2020) J. Brenneck and E. Polizzi, “An iterative method for contour-based nonlinear eigensolvers,” (2020), arXiv:2007.03000 [math.NA] .
- Rojas et al. (1995) H. N. Rojas, R. W. Godby, and R. J. Needs, Phys. Rev. Lett. 74, 1827 (1995).
- Ren et al. (2012a) X. Ren, P. Rinke, V. Blum, J. Wieferink, A. Tkatchenko, A. Sanfilippo, K. Reuter, and M. Scheffler, New Journal of Physics 14, 053020 (2012a).
- Ren et al. (2012b) X. Ren, P. Rinke, C. Joas, and M. Scheffler, Journal of Materials Science 47, 7447–7471 (2012b).
- Godby et al. (1988) R. W. Godby, M. Schlüter, and L. J. Sham, Phys. Rev. B 37, 10159 (1988).
- Gonze et al. (2009) X. Gonze, B. Amadon, P.-M. Anglade, J.-M. Beuken, F. Bottin, P. Boulanger, F. Bruneval, D. Caliste, R. Caracas, M. Côté, T. Deutsch, L. Genovese, P. Ghosez, M. Giantomassi, S. Goedecker, D. Hamann, P. Hermet, F. Jollet, G. Jomard, S. Leroux, M. Mancini, S. Mazevet, M. Oliveira, G. Onida, Y. Pouillon, T. Rangel, G.-M. Rignanese, D. Sangalli, R. Shaltaf, M. Torrent, M. Verstraete, G. Zerah, and J. Zwanziger, Computer Physics Communications 180, 2582 (2009), 40 YEARS OF CPC: A celebratory issue focused on quality software for high performance, grid and novel computing architectures.
- Govoni and Galli (2015) M. Govoni and G. Galli, Journal of Chemical Theory and Computation 11, 2680 (2015), pMID: 26575564, https://doi.org/10.1021/ct500958p .
- Golze et al. (2018) D. Golze, J. Wilhelm, M. J. van Setten, and P. Rinke, Journal of Chemical Theory and Computation 14, 4856 (2018), pMID: 30092140, https://doi.org/10.1021/acs.jctc.8b00458 .
- Golze et al. (2020) D. Golze, L. Keller, and P. Rinke, The Journal of Physical Chemistry Letters 11, 1840 (2020), pMID: 32043890, https://doi.org/10.1021/acs.jpclett.9b03423 .
- Soininen et al. (2003) J. A. Soininen, J. J. Rehr, and E. L. Shirley, Journal of Physics: Condensed Matter 15, 2573 (2003).
- Soininen et al. (2005) J. A. Soininen, J. J. Rehr, and E. L. Shirley, Physica Scripta 2005, 243 (2005).
- Kas et al. (2007) J. J. Kas, A. P. Sorini, M. P. Prange, L. W. Cambell, J. A. Soininen, and J. J. Rehr, Phys. Rev. B 76, 195116 (2007).
- Casida (1995) M. E. Casida, “Time-dependent density functional response theory for molecules,” in Recent Advances in Density Functional Methods (1995) pp. 155–192.
- Tiago and Chelikowsky (2006) M. L. Tiago and J. R. Chelikowsky, Phys. Rev. B 73, 205334 (2006).
- Bruneval (2012) F. Bruneval, The Journal of Chemical Physics 136, 194107 (2012), https://pubs.aip.org/aip/jcp/article-pdf/doi/10.1063/1.4718428/15450022/194107_1_online.pdf .
- Gao and Chelikowsky (2019) W. Gao and J. R. Chelikowsky, Journal of Chemical Theory and Computation 15, 5299 (2019), pMID: 31424933, https://doi.org/10.1021/acs.jctc.9b00520 .
- NES (2022) “Nessie- high-performance first-principle simulations,” http://www.nessie-code.org/ (2022).
- Kestyn and Polizzi (2020) J. Kestyn and E. Polizzi, IEEE Nanotechnology Magazine 14, 52 (2020).
- Li et al. (2024) D. Li, J. Kestyn, and E. Polizzi, Computer Physics Communications 295, 109014 (2024).
- Li and Polizzi (2024) D. Li and E. Polizzi, unpublished (2024).
- Betcke et al. (2013) T. Betcke, N. J. Higham, V. Mehrmann, C. Schröder, and F. Tisseur, ACM Transactions on Mathematical Software 39, 1 (2013).
- Güttel and Tisseur (2017) S. Güttel and F. Tisseur, Acta Numerica 26, 1 (2017).
- FEA (2020) “Feast eigenvalue solver,” http://www.feast-solver.org/ (2009-2020).
- Beyn (2012) W.-J. Beyn, BIT Numerical Mathematics 52, 109 (2012), arXiv:1003.1580 .