We assess the accuracy of common hybrid exchange-correlation (XC) functionals (PBE0, PBE0-1/3, HSE06, HSE03, and B3LYP) within the Kohn–Sham density functional theory for the harmonically perturbed electron gas at parameters relevant for the challenging conditions of the warm dense matter. Generated by laser-induced compression and heating in the laboratory, the warm dense matter is a state of matter that also occurs in white dwarfs and planetary interiors. We consider both weak and strong degrees of density inhomogeneity induced by the external field at various wavenumbers. We perform an error analysis by comparing with the exact quantum Monte Carlo results. In the case of a weak perturbation, we report the static linear density response function and the static XC kernel at a metallic density for both the degenerate ground-state limit and for partial degeneracy at the electronic Fermi temperature. Overall, we observe an improvement in the density response when the PBE0, PBE0-1/3, HSE06, and HSE03 functionals are used, compared with the previously reported results for the PBE, PBEsol, local-density approximation, and AM05 functionals; B3LYP, on the other hand, does not perform well for the considered system. Additionally, the PBE0, PBE0-1/3, HSE06, and HSE03 functionals are more accurate for the density response properties than SCAN in the regime of partial degeneracy.

The Kohn–Sham density functional theory (KS-DFT) is the most popular first-principles electronic structure method both at ambient and extreme conditions. This includes simulations of the warm dense matter (WDM),1–4 where KS-DFT is widely used5–7 for the interpretation of X-Ray Thomson scattering experiments.8–10 WDM is characterized by high temperatures and densities and is generated using different techniques11 at large research facilities like the European X-ray Free-Electron Laser (XFEL),12 the National Ignition Facility (NIF),13,14 the Linac coherent light source at SLAC,15 or the Sandia Z-machine.16 Both experimental and theoretical investigations of WDM are of high current interest due to its relevance for astrophysics, material science, and inertial confinement fusion.14,17 Examples of astrophysical applications include white dwarfs18 and planetary interiors.19–21 In addition, WDM is relevant for the discovery of novel materials at extreme conditions19,22–24 and for hot-electron chemistry.25 

The rigorous understanding of the physics and chemistry of the WDM state is still in a stage of active development, since most widely used electronic structure calculations and diagnostic tools were developed for ambient conditions and require a consistent extension to high temperatures. For example, in the case of the KS-DFT, the key ingredient for an accurate simulation is the exchange-correlation (XC) functional. At ambient conditions, the quality of different XC functionals is tested by comparing them with the experimental data, e.g., with band gaps,26 lattice constants,27 and atomization energies.28 

In WDM simulations, it is a common practice to use ground-state XC functionals with thermal excitations included by the smearing of band occupation numbers according to the Fermi–Dirac distribution;29,30 this is often referred to as zero-temperature approximation in the literature. Although there has been a certain progress in the development of finite-temperature XC functionals,31–35 testing XC functionals at WDM conditions is challenging. First, there are significant uncertainties in the experimentally realized temperature and density.9,10,36 Second, besides the very recent path integral Monte Carlo (PIMC) simulation results for hydrogen at certain WDM parameters,37,38 the exact simulation data for complex materials like compressed solids at WDM conditions24,39,40 that would enable reliable benchmarks of XC functionals are sparse.

This situation is highly unsatisfactory because the KS-DFT results are often used for the interpretation of experimental observations6,24 and as an input to other methods and models, including machine-learning models,41 XC kernels for linear-response time-dependent density functional theory calculations,42 molecular dynamics with screened effective potentials,43–49 and quantum hydrodynamics.50–52 

Clearly, an extensive benchmark and analysis of the accuracy of KS-DFT calculations based on various XC functionals at high temperatures are needed. Recently, Moldabekov et al.42,53,54 have reported the analysis of the local-density approximation (LDA),55 PBE,56 PBEsol,57 AM05,58 and SCAN59 XC functionals at temperatures and densities relevant for WDM by considering an inhomogeneous electron gas in an external harmonic perturbation and comparing DFT results with exact PIMC reference data.60–62 This has provided important insights to the performance of these functionals at WDM conditions. For example, in contrast to the ground state calculations, it was found that at high temperatures, SCAN performs worse than LDA, PBE, and PBEsol. Therefore, it is important to test the accuracy of XC functionals at WDM conditions. Furthermore, for the examples of the warm dense electron gas and hydrogen, it has been pointed out that even finite-temperature versions of the LDA and generalized gradient approximation (GGA) level functionals—created to extend the corresponding ground state versions to high temperatures—do not always lead to improved results at WDM conditions.42 Recently, Moldabekov et al.42 have analyzed the static XC kernel and density response function of the uniform electron gas (UEG) and hydrogen at WDM conditions using a finite-temperature LDA by Groth et al.32 It was found that for the UEG and warm dense hydrogen, the finite-temperature LDA provides less accurate results compared with the ground-state LDA. This seemingly surprising result was explained by a reduced validity (regarding wave number) of the long-wavelength approximation for the static XC kernel with an increase in temperature. It is clear that the explicit inclusion of the temperature dependence in an XC functional does not guarantee an improved description compared with its counterpart in the ground state. This highlights the importance of accurately testing and benchmarking XC functionals at WDM conditions.

On the other hand, a generalized KS-DFT63 with orbital-dependent XC functionals has emerged as one of the most promising tools in the density-functional theory. The most often used orbital-dependent XC functionals63,64 belong to the class of the hybrid functionals, which entail a certain weighted combination of the exact Hartree–Fock exchange contribution and of a standard XC functional like the LDA or PBE.28,65 These orbital-dependent XC functionals naturally contain information about non-locality effects and reduce the self-interaction error.66 Moreover, they provide a superior description of various material properties like lattice constants27 and atomization energies.28 

Hybrid XC functionals—within generalized KS-DFT at finite temperature31,35,67—can potentially overcome some problems of finite-temperature XC functionals since, in addition to the reduction of the self-interaction error, they contain extra information about thermal effects through orbitals in the exact exchange part. Therefore, in this work, we extend the previous benchmark and analysis of the LDA, GGA, and meta-GGA functionals for the harmonically perturbed electron gas at WDM conditions53,54 to include a number of popular hybrid XC functionals. In our assessment, we consider the HSE06,68,69 HSE03,70 PBE0,71 PBE0-1/3,72 and B3LYP73 hybrid functionals.

First, we consider the weakly perturbed electron gas and present results for the static hybrid XC kernel Kxc(q) of the uniform electron gas (UEG) at both ambient and WDM conditions using the recent framework by Moldabekov et al.42 In particular, Kxc(q) constitutes a wave-vector resolved measure for electronic XC effects and, therefore, is uniquely suited to benchmark the performance of XC functionals. In addition, the kernel constitutes a key input for a gamut of applications such as quantum hydrodynamics50,51 or the interpretation of X-Ray Thomson-Scattering (XRTS) experiments.74–77 

Second, we consider a strongly inhomogeneous, warm dense electron gas. The analysis is carried out by comparing with results for the PBE functional and exact PIMC reference data. PIMC allows us to unambiguously quantify the accuracy of the considered hybrid XC functionals, whereas the comparison with the PBE functional gives us insight into the physics behind the observed behavior of the hybrid functionals at different conditions.

The paper is organized as follows: we begin with providing the theoretical background in Sec. II, the details of our KS-DFT and PIMC calculations are given in Sec. III, the results of the calculations are presented in Sec. IV, and the paper is concluded in Sec. V with a summary of the main findings and a brief outlook on future research.

Hybrid XC functionals are based on the adiabatic connection formula.28 At finite temperature, the thermal adiabatic connection formula for the exchange-correlation functional, in terms of the potential energy of exchange-correlation, Uxcλ,78 is given by31,
Fxc=01dλUxcλ,
(1)
where λ is the electron–electron coupling parameter (interaction strength). In the λ = 0 limit, Uxcλ=0 reduces to the finite temperature Hartree–Fock exchange, with Uxcλ=0=FXHF given by35,79
FXHF[ρσ]=12σ=,dr1dr2ρσr1,r22r12,
(2)
where r12=r1r2 and ρσr1,r2 are the exchange densities,
ρσr1,r2=ifiσϕiσr1ϕiσ*r2,
(3)
with fiσ being the orbital occupation number defined by the Fermi–Dirac distribution.
Within the hybrid XC functionals construction scheme, the λ = 1 limit Uxcλ=1 is approximated by a standard XC density functional, ExcDF[n], such as LDA or GGA28,65 (where DF is an abbreviation for density functional). In their essence, hybrid XC functionals are based on the approximation of the Uxcλ in the form of a certain (usually polynomial) interpolation between the limits Uxcλ=0=FxHF[ρσ] and Uxcλ=1=ExcDF[n].28,65,71–73 For example, employing65,
Uxcλ=ExcDF[n]+FxHF[ρσ]ExDF[n]1λn1,
(4)
with n = 4 and ExDF being the exchange part of ExcDF and implementing the PBE as the DF, one arrives at a finite temperature version of the PBE0 hybrid XC functional,71 
FxcPBE0[ρσ,n]=ExcPBE[n]+14FxHF[ρσ]ExPBE[n].
(5)

In Eq. (5), the implicit dependence on the electronic temperature is taken into account in FxHF[ρσ] due to the smearing of the occupation numbers. Similarly, the usage of orbitals in the Fermi–Dirac distribution of the occupation numbers in the PBE0-1/3, HSE06, HSE03, and B3LYP allows for incorporating the implicit portion of thermal effects. Clearly, this is not a fully consistent way of generating a finite-temperature hybrid XC functional. Nevertheless, e.g., the HSE03 has been successfully used for the description of the experimental X-ray spectra of aluminum80 and the reflectivity and dc electrical conductivity of liquid ammonia81 at high temperatures. Furthermore, it turns out that the straightforward use of the orbitals with smeared occupation numbers according to the Fermi–Dirac distribution in the PBE0, PBE0-1/3, HSE06, and HSE03 provides a significant improvement in the description of the electronic density response at high temperatures TTF compared with PBE as reported in Sec. IV. Note that in the case of the HSE06 and HSE03, the Coulomb potential in FxHF[ρσ] is substituted by a screened potential 1/r12 → erfc(Ωr12)/r1279 to avoid the computational difficulties due to the long-range character of the Coulomb potential. Consistently, the screened version for the exchange part of the PBE is also used. For example, Ω = 0.11 Bohr−1 is used for the HSE03.69 

We investigate the performance of various hybrid XC functionals at WDM conditions by analyzing the density response of the interacting electron gas governed by the following Hamiltonian:62,82,83
Ĥ=ĤUEG+k=1N2Acos(r̂kq),
(6)
where ĤUEG is the Hamiltonian of the unperturbed UEG3,30,84 and N is the number of electrons. Note that we restrict ourselves to the unpolarized case with an equal number of spin-up and spin-down electrons, N = N = N/2, throughout this work. In addition to ĤUEG, the Hamiltonian in Eq. (6) has an extra harmonic perturbation defined by the amplitude A and wavelength λ = 2π/|q|.

It allows us to investigate both the weakly and strongly inhomogeneous electron gas by tuning the external perturbation. As a measure of the density inhomogeneity, one can use the maximum value of the density perturbation δnmax/n0=maxδn(r)/n0, where δn(r) = n(r) − n0 is the density deviation from the UEG with density n0.

When δnmax/n0 ≳ 1 and sufficiently small perturbation wavelength, we access the regime where standard XC functionals fail due to the strong degree of density inhomogeneity and self-interaction.53 In the limit δnmax/n0 ≪ 1, the static density response function is defined only by the properties of the UEG and is independent of A. In this case, for the static density response function, χ(q), we have
δn(r)=2Acos(qr)χ(q).
(7)
Within linear response theory, the density response function χ(q, ω) can be represented in terms of the ideal response function χ0(q, ω) and the XC kernel,
χ(q,ω)=χ0(q,ω)1v(q)+Kxc(q,ω)χ0(q,ω),
(8)
where, for the UEG, χ0(q, ω) is the temperature-dependent Lindhard function.
Following Ref. 42, one can invert Eq. (8) to compute the static XC kernel using χ(q) = χ(q, ω = 0), and χ0(q) = χ0(q, ω = 0),
Kxc(q)=v(q)+1χ(q)1χ0(q),=1χRPA(q)1χ(q),
(9)
where χRPA(q) is the density response function in the random phase approximation (RPA) defined by setting Kxc(q) ≡ 0. In the static case, ω = 0, from Eq. (8) we get
χRPA(q)=χ0(q)1v(q)χ0(q),
(10)
which is the temperature-dependent analytical RPA result for the UEG.

The XC kernel constitutes the key input to linear-response time-dependent density functional theory. Clearly, the XC kernels from the hybrid XC functionals are also interesting in their own right since they provide insight into the XC features at different wavenumbers. We stress that the static XC kernels obtained from hybrid XC functionals for the UEG had not been reported prior to our work—neither for the ground state nor for WDM.

For the UEG, the XC kernel is equivalent to the local field correction (LFC) commonly used in the quantum theory of the electron liquid.30,85 The relation between the static XC kernel and the static LFC reads
G(q,ω)=1v(q)Kxc(q,ω).
(11)
For a correct description of the UEG and related systems like metals at ambient conditions and at WDM parameters, the LFC must obey the long wave-length expansion based on the compressibility sum-rule,86 
limq0G(q)=q24π2n02n0Fxc,
(12)
with n0 = N/V denoting the average number density of the electrons in the UEG or the mean value of the conduction electron density in metals.

We use the available exact quantum Monte Carlo data for G(q) of the UEG to analyze various XC kernels.60,76,82,87–89

Beyond the linear response theory, accurate data for the XC kernel (LFC) in combination with ideal density response functions allows one to compute higher order quadratic and cubic density response functions.62,90,91

Commonly used parameters for the description of the electrons in the WDM regime are the coupling parameter rs and the reduced temperature θ = T/TF, where the former is the mean inter-particle distance in Hartree atomic units and the latter is the ratio of the temperature to the Fermi temperature.92 The WDM state is defined by the regime where electrons are non-ideal rs ≳ 1, and thermal excitations and quantum degeneracy are simultaneously important, i.e., θ ∼ 1. Often in experiments, the WDM state is generated by shock-compression and (isochoric) heating of solid-state samples.24,39,40 In this regard, understanding the WDM state around metallic densities is particularly important. Thus, we consider the characteristic metallic density of rs = 2 and the partially degenerate case with θ = 1. This corresponds to an electron temperature T ≃ 12.5 eV. In the case of a weak perturbation δnmax/n0 ≪ 1, we additionally consider the ground state by setting θ = 0.01. For this case, we analyze the quality of the considered hybrid XC functionals in the UEG limit in terms of quantum Monte Carlo data for the density response function by Moroni et al.82 Furthermore, the analysis of the ground state allows us to gauge the impact of thermal excitation effects on the performance of the hybrid XC functionals at TTF.

For the perturbation in Eq. (6), we consider A = 0.01, A = 0.1, A = 0.5, and A = 1.0 (in Hartree). These parameters correspond to various degrees of the density inhomogeneity from δnmax/n0 ≲ 10−2 (at A = 0.01) to δnmax/n0 ≳ 1 (at A = 0.5 and A = 1.0). Thus, A = 0.01 allows one to probe the properties of the system in the UEG limit. We particularly extract the static density response function and the static XC kernel following the approach introduced in Ref. 42 and compare them with the PIMC data.

The wavenumber of the perturbation q is directed along the z-axis. Note that we drop the vector notation in the following. The values of the perturbation wavenumber are defined by the length of the simulation cell as q = 2πj/L, where j ≥ 1 is a positive integer number. The smallest value of q corresponding to j = 1 is denoted as qmin. We consider q in the range from qmin to 4qmin and up to 5qmin at A = 0.01 for the static density response and XC kernel calculations.

The KS-DFT calculations were performed using the ABINIT package.93–98 The calculations at θ = 1 were performed with the cubic main cell length L = 7.7703 Bohr, Nb = 200 orbitals, and the maximal kinetic energy cutoff 13 Ha. The used cell length corresponds to N = L3 × n0 = 14 electrons, where at rs = 2, we have n0 ≃ 0.02984 Bohr−3. The smallest perturbation wavenumber for 14 particles is qmin[N = 14] ≃ 0.842 68 qF, where qF is the Fermi wavenumber. Periodic boundary conditions and a k-point grid of size 8 × 8 × 8 were used. Self-consistent-field cycles for the solution of the Kohn–Sham equations were converged with absolute differences of the total energy δE < 10−7 Ha. A large number of bands NbN are needed due to significant thermal smearing of the occupation numbers. At the considered parameters, Nb = 200 corresponds to the smallest occupation number fmin < 10−4, which is sufficiently small to acquire convergence.53 In prior works, it was shown that, in the case of the harmonically perturbed electron gas, the results for the response functions with N = 14 particles are not affected by finite size effects.42,53,99–102 In this work, keeping other parameters equivalent, the convergence is cross-checked by comparing the data for N = 14 with the results computed with N = 20 and Nb = 300 at q = 2qmin (with A = 0.01). Additionally, we used three different particle numbers in the simulation cell for the ground state calculations: N = 38, 54, and 66; with a k-point grid of 8 × 8 × 8. The agreement and consistency of the results for all considered hybrid XC functionals computed with different particle numbers reassure the convergent character of the calculations. Finally, for both θ = 1 and θ = 0.01, we performed KS-DFT calculations with the XC functional set to zero yielding the static density response function of the UEG in the random phase approximation (RPA) (see Sec. II B) and compared it with the analytical RPA (in N limit). The perfect agreement of our KS-DFT data with the analytical RPA [shown in Figs. 1(a) and 2(a)] substantiates our confidence in the convergence of our calculations.

FIG. 1.

The linear static electron density response function (top) and the static XC kernel (bottom) at θ = 0.01 and rs = 2. Solid black line: exact UEG results based on the neural-net representation of Ref. 60. Black squares: exact quantum Monte Carlo results by Moroni et al.82 The RPA result for the density response function in the top panel is given by the dashed line. The gray circles in the top panel are the KS-DFT results for the density response function with the XC functional set to zero (denoted as NullXC). The other symbols distinguish the KS-DFT results computed using different XC-functionals.

FIG. 1.

The linear static electron density response function (top) and the static XC kernel (bottom) at θ = 0.01 and rs = 2. Solid black line: exact UEG results based on the neural-net representation of Ref. 60. Black squares: exact quantum Monte Carlo results by Moroni et al.82 The RPA result for the density response function in the top panel is given by the dashed line. The gray circles in the top panel are the KS-DFT results for the density response function with the XC functional set to zero (denoted as NullXC). The other symbols distinguish the KS-DFT results computed using different XC-functionals.

Close modal
FIG. 2.

The linear static electron density response function (top) and the static XC kernel (bottom) at θ = 1 and rs = 2. Solid black line: exact UEG results based on the neural-net representation of Ref. 60. The RPA result for the density response function in the top panel is given by the dashed line. The gray circles in the top panel are the KS-DFT results for the density response function with the XC functional set to zero (denoted as NullXC). The other symbols distinguish KS-DFT results computed using different XC-functionals.

FIG. 2.

The linear static electron density response function (top) and the static XC kernel (bottom) at θ = 1 and rs = 2. Solid black line: exact UEG results based on the neural-net representation of Ref. 60. The RPA result for the density response function in the top panel is given by the dashed line. The gray circles in the top panel are the KS-DFT results for the density response function with the XC functional set to zero (denoted as NullXC). The other symbols distinguish KS-DFT results computed using different XC-functionals.

Close modal

We report results for a total of 155 KS-DFT calculations for the HSE06, HSE03, PBE0, PBE0-1/3, and B3LYP hybrid XC functionals with different choices of parameters (amplitude A and wavenumber q) in the density perturbation.

A fraction of the PIMC data (with q = qmin, 2qmin, and 3qmin) used in this work has been reported in the analysis of the ground state LDA, PBE, PBEsol, AM05, and SCAN functionals in Ref. 53. It was shown there that the quality of the KS-DFT calculations based on these XC functionals has the tendency to deteriorate with an increasing wavenumber of the perturbation. Here, we test the hybrid XC functionals in the regime where the functionals in the previous analysis have failed. We have, therefore, generated new PIMC data for q = 4qmin at A = 0.1, 0.5, and 1.0.

Since a detailed introduction to the direct PIMC method103 and the associated fermion sign problem104 have been presented elsewhere, we here restrict ourselves to a brief summary of the main simulation details. We have used the extended ensemble approach introduced in Ref. 105, which is a canonical adaption of the worm algorithm by Boninsegni et al.106,107 Moreover, we use P = 200 imaginary-time propagators from a primitive factorization scheme, which is sufficient to ensure convergence; see the supplementary material of Ref. 61 for a corresponding analysis.

We start with the weakly perturbed electron gas where |δn(r)|/n0 ≪ 1 and consider both the ground state and the parameters relevant for WDM. We analyze the results both in terms of the static density response function χ(q) and the static XC kernel Kxc(q). In the next step, we consider strongly perturbed systems with |δn(r)|/n0 ≳ 1 and provide the analysis of the accuracy of the electron density calculations based on the considered hybrid XC functionals.

1. Ground state results

First, the analysis of the T → 0 limit helps to establish a baseline for further analyzing the performance of the hybrid XC kernels at high temperatures. Second, for the construction of the LDA and GGA (e.g., PBE) XC functionals, the reproduction of the LFC of the UEG at q < 2qF was one of the main criteria for an adequate description of metals.56 Thus, it is interesting to see to what degree the considered hybrid XC functionals can reproduce the LFC of the UEG.

The results for the static density response function and the static XC kernel are shown in Figs. 1(a) and 1(b), respectively. In Fig. 1, we compare the KS-DFT results with the diffusion quantum Monte Carlo (DMC) data by Moroni et al. (black squares)82 computed for T → 0. In addition, we show the results from the machine-learning representation (ML-PIMC) by Dornheim et al. of extensive PIMC data at finite temperatures60 and the DMC data82 for T = 0. In addition, the results based on the hybrid XC functionals, in Fig. 1, we also show results computed using PBE. The latter comparison is meaningful because PBE0, PBE0-1/3, HSE06, and HSE03 are based on PBE.

Before discussing our results on the hybrid XC functionals, we validate the absence of any significant finite size effects in Fig. 1(a) by comparing the analytical RPA with the static density response from the KS-DFT calculations where the XC functional is set to zero (labeled NullXC42). The latter is equivalent to the RPA and does indeed agree with the analytical RPA as expected.

In Fig. 1(a), we observe that the results for χ(q) based on PBE0, PBE0-1/3, HSE06, HSE03, and PBE are closely grouped and are in rather good agreement with the reference DMC data. In contrast, the B3LYP results for χ(q) significantly deviate from the reference ML-PIMC data at q > qF. This is expected since B3LYP does not attain the UEG limit by design. This was elucidated by Paier et al. for the unperturbed UEG.27 Here, we further substantiate it for finite perturbation wavenumbers in the range 0.6 ≤ q ≤ 3qF as it can be seen in Fig. 1(b) for the static XC kernel, where a strong deviation of the B3LYP results from the DMC and ML-PIMC data is clearly visible. Furthermore, from Fig. 1(b), we see that the B3LYP data do not obey the long wave-length expansion of the LFC Eq. (12) with quadratic q dependence of the LFC at qqF; this can be seen particularly well in the inset that shows a magnified segment.

In contrast to B3LYP, the LFCs of PBE0, PBE0-1/3, HSE06, and HSE03 reproduce the quadratic dependence on the wavenumber at q ≲ 1.5qF in accordance with Eq. (12), as shown in Fig. 1(b). We also observe in Fig. 1(b) that the PBE results have a quadratic dependence on q at all considered wave numbers since it was designed in a way that the gradient corrections to the exchange and to the correlation parts exactly cancel each other in the limit of the UEG.56 

At large wavenumbers, we observe from Fig. 1(b) that the PBE0, PBE0-1/3, HSE06, and HSE03 results deviate from the DMC data at q ≳ 2qF. Nevertheless, at 2qFq ≤ 3qF, the results from the considered hybrid XC functionals exhibit smaller absolute deviations from the DMC data compared with the PBE results. An interesting feature of the LFC based on the PBE0, PBE0-1/3, HSE06, and HSE03 is that it has a maximum somewhere around 2qF and a monotonically decreasing trend with increasing q at q > 2qF. A similar behavior—with a somewhat worse agreement with the DMC data—was reported by Armiento and Mattsson for the AM0558 functional, which is a semi-local GGA. Another related observation is that AM05 provides as accurate a description as the PBE0 and HSE06 hybrid XC functionals for lattice constants, bulk moduli, and surfaces of solids.108 The latter property, i.e., the electronic behavior at surfaces, corresponds to a domain of strongly inhomogeneous electron densities characterized by large wave numbers qqF.109 Thus, it is an interesting observation that the AM05 XC kernel has a similar behavior as the XC kernel computed using the HSE06, HSE03, PBE0, and PBE0-1/3 functionals at q > 2qF. We do not further extend the analysis of the observed similarities between the AM05 XC kernel and the hybrid XC kernels since it would deviate from the main focus of this work, but we leave such investigation as a possible direction for future research.

In Fig. 1(b), one can see that at q ≲ 1.5qF, the PBE0, PBE0-1/3, HSE06, HSE03, and PBE functionals provide static XC kernels in excellent agreement with the DMC data. Additionally, the PBE0 and PBE0-1/3 results are much closer to the DMC data at q ≃ 2qF compared with HSE06, HSE06, and PBE. At q ≳ 2qF, PBE0 provides a much better description of the static XC kernel of the UEG compared with the other hybrid XC functionals.

In addition, we find that increasing the contribution of HF exchange from 1/4 (as in PBE0) to 1/3 (as in PBE0-1/3) lowers the resulting values of Kxc(q) at q > 2qF. Keeping the coefficient 1/4 in front of the HF exchange fixed in Eq. (5), the same effect is achieved by using a screened potential instead of the Coulomb potential as illustrated by the static XC kernel based on the HSE06 and the HSE03. On the other hand, the behavior of the Kxc(q) at q < 2qF is not sensitive to the screening of the Coulomb potential or the variation of the HF exchange contribution. One could adjust the screening parameter or the weight of the HF exchange contribution in order to accurately reproduce the exact data for the Kxc(q) in a wider range of wavenumbers.

2. Results at a high temperature TTF

Our new results for the static density response function and the static XC kernel of the UEG at θ = 1 are presented in Figs. 2(a) and 2(b), respectively. Similar to the ground state, we observe that the B3LYP results for the static density response function [Fig. 2(a)] and the static XC kernel [Fig. 2(b)] disagree significantly with the reference ML-PIMC data at qqF. Also, the B3LYP data for the static XC kernel do not obey a quadratic dependence on the wavenumber at qqF and approaches the PBE result from above with a decrease in q and crosses the PBE data at q ≃ 0.84qF. The PBE data show a quadratic dependence on q in the entire considered range of wavenumbers and agree well with the reference data (ML-PIMC) at q ≲ 1.5qF. The PBE0, PBE0-1/3, HSE06, and HSE03 results are in close agreement with the reference data (ML-PIMC) at q ≲ 2.5qF (with the PBE0 being the most accurate). Therefore, these functionals lead to a better description of the density response compared with PBE. At q ≳ 2.5qF, the PBE0, PBE0-1/3, HSE06, and HSE03 functionals produce static XC kernels that decrease with increasing q and eventually become negative at q ≳ 3.5qF. A similar behavior has been reported for the AM05 functional at TTF.42 In general, the trends in the q-dependence and main features of χ(q) and Kxc(q) of the considered hybrid XC functionals are similar to those in the ground state discussed above. This indicates that the partial occupation of the orbitals at high temperatures does not lead to fundamentally new features in the hybrid XC functional results.

With that, we conclude our consideration of the weak perturbation limit. We continue further with the case of the strong density inhomogeneity.

To generate a strong density perturbation, we apply the external harmonic perturbation as defined in Eq. (6) with amplitudes A = 0.1, 0.5, and 1.0 and wavenumbers in the range from q = qmin ≃ 0.84 qF up to q = 4qmin. The amplitude of the density inhomogeneity is qualitatively illustrated in Fig. 3, where we plot the largest values of the density accumulation (δn) and the density depletion (δn) in the simulation cell due to the external harmonic field with different wavenumbers. The gray area indicates density perturbations less than 0.1n0. We see that at the considered amplitudes of the perturbation, we always have |δn|≥ 0.1n0 and |δn|≥ 0.1n0. These density perturbation magnitudes are well beyond the linear response regime considered in Sec. IV A. At A = 0.5, we have |δn|∼ n0 and |δn|∼ 0.5n0. Furthermore, at A = 1.0, the electron density values almost vanish in the density depletion regions with δn → −n0 at q = qmin and q = 2qmin, where a physical lower bound δn > − n0 is in place (indicated by the dashed horizontal line in Fig. 3). Correspondingly, most of the electrons are localized in the density accumulation regions with δn ≈ 2n0.

FIG. 3.

The largest value of the density perturbation at different external field amplitudes and wavenumbers in the density accumulation region is denoted as ⊕ (positive values) and in the density depletion region is denoted as ⊖ (negative values). The smallest considered value of the perturbation wavenumber is q ≃ 0.84qF. The shaded area indicates the density perturbations δn ≤ 0.1n0. The dashed horizontal line represents lower bound for the density perturbation in the depletion region since n = n0 + δn cannot be negative. The data are from the PIMC calculations at rs = 2 and θ = 1.

FIG. 3.

The largest value of the density perturbation at different external field amplitudes and wavenumbers in the density accumulation region is denoted as ⊕ (positive values) and in the density depletion region is denoted as ⊖ (negative values). The smallest considered value of the perturbation wavenumber is q ≃ 0.84qF. The shaded area indicates the density perturbations δn ≤ 0.1n0. The dashed horizontal line represents lower bound for the density perturbation in the depletion region since n = n0 + δn cannot be negative. The data are from the PIMC calculations at rs = 2 and θ = 1.

Close modal

An important point for our analysis is that at fixed values of A, the density perturbation magnitude can have a non-monotonic dependence on q due to a stronger response of electrons around q = 2qmin ≈ 1.5qF as it is clearly illustrated in Fig. 3 for A = 0.1 and A = 0.5. A stronger density response at the wavelength corresponding to the mean inter-particle distance d ≃ 2rs, i.e., at q ≃ 2π/d ≃ 2qF, is a manifestation of the fact that it is energetically more favorable for particles to be localized at the distance rd from each other. This was demonstrated and used for the explanation of the roton feature in the excitation spectrum of the UEG by Dornheim et al.,.110 Contrarily, at fixed values of q, the increase in A always leads to a monotonic increase in the density perturbation amplitude. Therefore, to analyze the error of the hybrid XC functionals in the KS-DFT calculations, it is more transparent to compare the deviations from the exact PIMC data for various A values at a fixed q, rather than vice versa.

We perform an error analysis following Refs. 53 and 54, where we use the relative density deviation between the KS-DFT data and the reference quantum Monte Carlo (QMC) data defined as
Δñ[%]=δnDFTδnQMCmax{δnQMC}×100,
(13)
where max{δnQMC} is the maximum deviation of the QMC data from n0. From a physical perspective, Eq. (13) defines an error in the density response computed using KS-DFT in comparison with the exact density response from the PIMC calculations. It becomes obvious by considering a weak perturbation where δn(r) = 2 A cos(qr)χ(q) and Eq. (13) reduces to the error evaluation of the linear density response function.111 

Before we dive into the details of the error analysis of the hybrid XC functionals, in Fig. 4, we illustrate the density perturbation δn(r) = n(r) − n0 and the density error Δñ profiles when the harmonic perturbation with A = 0.1 and q = 3qmin is applied. From Fig. 4(a), we observe that the PBE data exhibit a certain deviation from the PIMC data at the maximum and minimum values of δn(r). The B3LYP results show a significant disagreement with the PIMC data in both the density accumulation (with δn(r) > 1) and density depletion regions (with δn(r) < 1). The results for the other considered hybrid XC functionals are in a close agreement with the exact PIMC data. This is further reaffirmed by Δñ in Fig. 4(b), where |Δñ|1% for the PBE0, PBE0-1/3, HSE06, and HSE03 results. Additionally, from Fig. 4(b), we clearly see that the PBE (the B3LYP) data have an error of up to ±5% (±15%) in the relative density deviation Δñ. Obviously, in this case, the PBE0, PBE0-1/3, HSE06, and HSE03 hybrid XC functionals provide a significant improvement due to mixing in a fraction of HF exchange.

FIG. 4.

(a) The density perturbation δn(r) and (b) the density error Δñ[%] profiles along the perturbation direction at A = 0.1, q = 3qmin ≃ 2.528 qF, rs = 2, and θ = 1.

FIG. 4.

(a) The density perturbation δn(r) and (b) the density error Δñ[%] profiles along the perturbation direction at A = 0.1, q = 3qmin ≃ 2.528 qF, rs = 2, and θ = 1.

Close modal

To perform a general analysis of the errors in the density response, we consider the largest value of the Δñ in the density accumulation region, Δñ, and the largest magnitude of the Δñ in the density depletion region, Δñ, for different A and q.

In Fig. 5, we show the histogram representation of the largest value of the error in the relative density deviation as defined by Δñ and Δñ at q = qmin ≃ 0.84 qF and different values of A. We observe that the B3LYP data have the smallest error with Δñ<1% and Δñ<1% compared with other KS-DFT data. However, it appears to be the crossing point between the B3LYP density response and the exact density response as the function of the parameters A and q, where one can say that the good performance of B3LYP is due to chance. We have already observed such a crossing point when considering the linear density response function in the case of the weak perturbation in Sec. IV A. Indeed, the B3LYP data significantly deviate from the reference PIMC data at other considered q > qmin values at A = 0.1, 0.5, and 1.0 (see Figs. 68). In fact, at A = 0.1 and q = 2qmin, 3qmin, and 4qmin, the B3LYP data have the largest errors in both density accumulation and depletion regions among all considered XC functionals. This is consistent with the conclusions drawn from the linear density response function in Sec. IV A. The reason is that at A = 0.1, with |δn|∼ 0.1n0 and |δn|∼ 0.1n0 (see Fig. 3), the linear density response function has the largest contribution to the total density response within non-linear response theory, which includes quadratic and cubic density response functions.62,91 This reasoning is not applicable at A = 0.5 and A = 1.0 since density deviation from the UEG case is well beyond of the regular perturbation expansion, which is valid for |δn(r)|/n0 ≪ 1.

FIG. 5.

Histogram representation of the largest value of the error in the relative density deviation of the KS-DFT data from the exact PIMC data, Eq. (13), in the density accumulation region (the top panel denoted as ⊕) and in the density depletion region (the bottom panel denoted as ⊖) for different A at q = qmin ≃ 0.84 qF. The results are computed for rs = 2 and θ = 1.

FIG. 5.

Histogram representation of the largest value of the error in the relative density deviation of the KS-DFT data from the exact PIMC data, Eq. (13), in the density accumulation region (the top panel denoted as ⊕) and in the density depletion region (the bottom panel denoted as ⊖) for different A at q = qmin ≃ 0.84 qF. The results are computed for rs = 2 and θ = 1.

Close modal
FIG. 6.

Histogram representation of the largest value of the error in the relative density deviation of the KS-DFT data from the exact PIMC data, Eq. (13), in the density accumulation region (the top panel denoted as ⊕) and in the density depletion region (the bottom panel denoted as ⊖) for different A at q = 2qmin ≃ 1.685 qF. The gray area corresponds Δñ±1% and Δñ±1%. The results are computed for rs = 2 and θ = 1.

FIG. 6.

Histogram representation of the largest value of the error in the relative density deviation of the KS-DFT data from the exact PIMC data, Eq. (13), in the density accumulation region (the top panel denoted as ⊕) and in the density depletion region (the bottom panel denoted as ⊖) for different A at q = 2qmin ≃ 1.685 qF. The gray area corresponds Δñ±1% and Δñ±1%. The results are computed for rs = 2 and θ = 1.

Close modal
FIG. 7.

Histogram representation of the largest value of the error in the relative density deviation of the KS-DFT data from the exact PIMC data, Eq. (13), in the density accumulation (the top panel denoted as ⊕) and density depletion regions (the bottom panel denoted as ⊖) for different A at q = 3qmin ≃ 2.528 qF. The gray area corresponds Δñ±1% and Δñ±1%. The results are computed for rs = 2 and θ = 1.

FIG. 7.

Histogram representation of the largest value of the error in the relative density deviation of the KS-DFT data from the exact PIMC data, Eq. (13), in the density accumulation (the top panel denoted as ⊕) and density depletion regions (the bottom panel denoted as ⊖) for different A at q = 3qmin ≃ 2.528 qF. The gray area corresponds Δñ±1% and Δñ±1%. The results are computed for rs = 2 and θ = 1.

Close modal
FIG. 8.

Histogram representation of the largest value of the error in the relative density deviation of the KS-DFT data from the exact PIMC data, Eq. (13), in the density accumulation (the top panel denoted as ⊕) and density depletion regions (the bottom panel denoted as ⊖) for different A at q = 4qmin ≃ 3.371 qF. The gray area corresponds Δñ±1% and Δñ±1%. The results are computed for rs = 2 and θ = 1.

FIG. 8.

Histogram representation of the largest value of the error in the relative density deviation of the KS-DFT data from the exact PIMC data, Eq. (13), in the density accumulation (the top panel denoted as ⊕) and density depletion regions (the bottom panel denoted as ⊖) for different A at q = 4qmin ≃ 3.371 qF. The gray area corresponds Δñ±1% and Δñ±1%. The results are computed for rs = 2 and θ = 1.

Close modal

From Fig. 5, one can see that at q = qmin ≃ 0.84 qF and A = 0.1, 0.5, and 1.0, PBE0, PBE0-1/3, HSE06, and HSE03 have a similar performance as PBE, with Δñ2% and Δñ1%. As shown in Fig. 6, at q = 2qmin ≃ 1.685 qF, the PBE0, PBE0-1/3, HSE06, and HSE03 results are a bit more accurate than PBE with the improvement of the accuracy approximately between 1% and 2% at A = 0.5 and A = 1.0 in the density accumulation region.

At q = 3qmin ≃ 2.528 qF and A = 0.1, we observe the largest improvement in the accuracy of the density response using PBE0, PBE0-1/3, HSE06, and HSE03 compared with PBE (see Fig. 7). At these parameters, the PBE data deviate significantly with about 5% in Δñ and Δñ. In contrast, the PBE0, PBE0-1/3, HSE06, and HSE03 results are highly accurate with deviations of |Δñ|1% and |Δñ|1%. From Fig. 7, one can see that at q = 3qmin and A = 0.5, PBE0, PBE0-1/3, HSE06, and HSE03 provide an improvement of a few percent in accuracy compared with PBE. At A = 1.0 and q = 3qmin, all considered XC functionals show similar accuracy with |Δñ|2% and |Δñ|2%.

Finally, in Fig. 8, we show the results for q = 4qmin ≃ 3.371 qF. Similar to the case with q = 3qmin ≃ 2.528 qF, PBE0, PBE0-1/3, HSE06, and HSE03 yield significantly higher accuracy than PBE for the density response at A = 0.1 and A = 0.5, with up to two times smaller errors in |Δñ| and/or |Δñ|. At last, for A = 1.0 and q = 4qmin, we observe similar accuracy for all considered XC functionals with |Δñ|4% and |Δñ|4%.

Before drawing general conclusions, we recall that in previous work, Moldabekov et al.53 had assessed the accuracy of both the LDA and GGA functionals for the density response in the warm dense electron gas. They had shown that the PBEsol and LDA functionals were approximately as accurate as PBE for the density response in the partially degenerate case with TTF for both weak and strong perturbations. The SCAN functional—being highly accurate in the ground state—had deviated significantly from the reference PIMC data at q > qF.42,53

In this work, the benchmark against the exact PIMC data at q ≲ 3qF for both weak and strong perturbations demonstrates that the PBE0, PBE0-1/3, HSE06, and HSE03 hybrid XC functionals yield a more accurate description of the density response compared with PBE. In contrast, B3LYP is significantly less accurate for the perturbed electron gas compared with the PBE or the other hybrid functionals.

Furthermore, we presented new results for the static XC kernel computed with the PBE0, PBE0-1/3, HSE06, HSE03, and B3LYP functionals for both the ground state and partially degenerate case with TTF. Among the considered XC functionals, it has been revealed that PBE0 yields the closest agreement with the exact PIMC and DMC reference data. Particularly, at q < 3qF, the PBE0 static XC kernel is practically exact. For the partially degenerate regime, this is a significant improvement over LDA, PBEsol, AM05, and SCAN static XC kernels.42 At q > 3qF and TTF, none of the aforementioned LDA, GGA, meta-GGA, and hybrid XC functionals are able to accurately describe the static LFC of the UEG, G(q) = (q2/4π)Kxc(q). This is not a critical point for the density response function since Kxc(q) ≪ 1 and Kxc(q) do not introduce a significant correction into χ(q) as evident from Eq. (8). Experimentally, large wave numbers are probed in WDM and dense plasmas using the XRTS signal collected at large scattering angles.75,112,113

The static XC kernel, being a crucial quantity in linear-response time-dependent density functional theory, enables a unique insight into the performance of the various XC functionals. Thus, in combination with the exact quantum Monte Carlo data for the UEG,60,76,82,87–89 it also can help with the construction of improved XC functionals for WDM applications, where the XRTS signal from large q values is crucial for the diagnostics of system parameters in experiments.

We are convinced that our present findings will be useful for future DFT calculations of WDM, in general, and for the many applications that utilize the XC kernel,42 in particular.

This work was funded by the Center for Advanced Systems Understanding (CASUS), which was financed by Germany’s Federal Ministry of Education and Research (BMBF) and by the Saxon state government out of the State budget approved by the Saxon State Parliament. We gratefully acknowledge computation time at the Norddeutscher Verbund für Hoch-und Höchstleistungsrechnen (HLRN) under Grant No. shp00026, and on the Bull Cluster at the Center for Information Services and High Performance Computing (ZIH) at Technische Universität Dresden. The authors also thank Henrik Schulz and Jens Lasch for providing very helpful support on high performance computing at HZDR.

The authors have no conflicts to disclose.

Zhandos A. Moldabekov: Conceptualization (lead); Data curation (lead); Formal analysis (lead); Investigation (lead); Methodology (equal); Visualization (lead); Writing – original draft (lead). Mani Lokamani: Data curation (equal); Writing – review & editing (equal). Jan Vorberger: Methodology (equal); Writing – review & editing (equal). Attila Cangi: Conceptualization (equal); Methodology (equal); Writing – review & editing (equal). Tobias Dornheim: Conceptualization (equal); Data curation (equal); Funding acquisition (equal); Methodology (equal); Project administration (equal); Writing – review & editing (equal).

The data that support the findings of this study are available from the Rossendorf Data Repository (RODARE).114 

1.
Frontiers and Challenges in Warm Dense Matter
, edited by
F.
Graziani
,
M. P.
Desjarlais
,
R.
Redmer
, and
S. B.
Trickey
(
Springer, International Publishing
,
2014
).
2.
M.
Bonitz
,
T.
Dornheim
,
Z. A.
Moldabekov
,
S.
Zhang
,
P.
Hamann
,
H.
Kählert
,
A.
Filinov
,
K.
Ramakrishna
, and
J.
Vorberger
, “
Ab initio simulation of warm dense matter
,”
Phys. Plasmas
27
,
042710
(
2020
).
3.
T.
Dornheim
,
S.
Groth
, and
M.
Bonitz
, “
The uniform electron gas at warm dense matter conditions
,”
Phys. Rep.
744
,
1
86
(
2018
).
4.
V. E.
Fortov
, “
Extreme states of matter on Earth and in space
,”
Phys.-Usp
52
,
615
647
(
2009
).
5.
A. D.
Baczewski
,
L.
Shulenburger
,
M. P.
Desjarlais
,
S. B.
Hansen
, and
R. J.
Magyar
, “
X-ray Thomson scattering in warm dense matter without the Chihara decomposition
,”
Phys. Rev. Lett.
116
,
115004
(
2016
).
6.
C.
Mo
,
Z.
Fu
,
W.
Kang
,
P.
Zhang
, and
X. T.
He
, “
First-principles estimation of electronic temperature from x-ray Thomson scattering spectrum of isochorically heated warm dense matter
,”
Phys. Rev. Lett.
120
,
205002
(
2018
).
7.
K.
Ramakrishna
,
A.
Cangi
,
T.
Dornheim
,
A.
Baczewski
, and
J.
Vorberger
, “
First-principles modeling of plasmons in aluminum under ambient and extreme conditions
,”
Phys. Rev. B
103
,
125118
(
2021
).
8.
S. H.
Glenzer
and
R.
Redmer
, “
X-ray Thomson scattering in high energy density plasmas
,”
Rev. Mod. Phys.
81
,
1625
(
2009
).
9.
D.
Kraus
,
B.
Bachmann
,
B.
Barbrel
,
R. W.
Falcone
,
L. B.
Fletcher
,
S.
Frydrych
,
E. J.
Gamboa
,
M.
Gauthier
,
D. O.
Gericke
,
S. H.
Glenzer
,
S.
Göde
,
E.
Granados
,
N. J.
Hartley
,
J.
Helfrich
,
H. J.
Lee
,
B.
Nagler
,
A.
Ravasio
,
W.
Schumaker
,
J.
Vorberger
, and
T.
Döppner
, “
Characterizing the ionization potential depression in dense carbon plasmas with high-precision spectrally resolved x-ray scattering
,”
Plasma Phys. Controlled Fusion
61
,
014015
(
2019
).
10.
T.
Dornheim
,
M.
Böhme
,
D.
Kraus
,
T.
Döppner
,
T. R.
Preston
,
Z. A.
Moldabekov
, and
J.
Vorberger
, “
Accurate temperature diagnostics for matter under extreme conditions
,”
Nat. Commun.
13
,
7911
(
2022
).
11.
K.
Falk
, “
Experimental methods for warm dense matter research
,”
High Power Laser Sci. Eng
6
,
e59
(
2018
).
12.
T.
Tschentscher
,
C.
Bressler
,
J.
Grünert
,
A.
Madsen
,
A.
Mancuso
,
M.
Meyer
,
A.
Scherz
,
H.
Sinn
, and
U.
Zastrau
, “
Photon beam transport and scientific instruments at the european XFEL
,”
Appl. Sci.
7
,
592
(
2017
).
13.
R.
Betti
and
O. A.
Hurricane
, “
Inertial-confinement fusion with lasers
,”
Nat. Phys.
12
,
435
448
(
2016
).
14.
S. X.
Hu
,
B.
Militzer
,
V. N.
Goncharov
, and
S.
Skupsky
, “
Strong coupling and degeneracy effects in inertial confinement fusion implosions
,”
Phys. Rev. Lett.
104
,
235003
(
2010
).
15.
B. K.
Ofori-Okai
,
M. C.
Hoffmann
,
A. H.
Reid
,
S.
Edstrom
,
R. K.
Jobe
,
R. K.
Li
,
E. M.
Mannebach
,
S. J.
Park
,
W.
Polzin
,
X.
Shen
,
S. P.
Weathersby
,
J.
Yang
,
Q.
Zheng
,
M.
Zajac
,
A. M.
Lindenberg
,
S. H.
Glenzer
, and
X. J.
Wang
, “
A terahertz pump mega-electron-volt ultrafast electron diffraction probe apparatus at the SLAC Accelerator Structure Test area facility
,”
J. Instrum
13
,
P06014
(
2018
).
16.
M. K.
Matzen
,
M. A.
Sweeney
,
R. G.
Adams
,
J. R.
Asay
,
J. E.
Bailey
,
G. R.
Bennett
,
D. E.
Bliss
,
D. D.
Bloomquist
,
T. A.
Brunner
,
R. B.
Campbell
,
G. A.
Chandler
,
C. A.
Coverdale
,
M. E.
Cuneo
,
J.-P.
Davis
,
C.
Deeney
,
M. P.
Desjarlais
,
G. L.
Donovan
,
C. J.
Garasi
,
T. A.
Haill
,
C. A.
Hall
,
D. L.
Hanson
,
M. J.
Hurst
,
B.
Jones
,
M. D.
Knudson
,
R. J.
Leeper
,
R. W.
Lemke
,
M. G.
Mazarakis
,
D. H.
McDaniel
,
T. A.
Mehlhorn
,
T. J.
Nash
,
C. L.
Olson
,
J. L.
Porter
,
P. K.
Rambo
,
S. E.
Rosenthal
,
G. A.
Rochau
,
L. E.
Ruggles
,
C. L.
Ruiz
,
T. W. L.
Sanford
,
J. F.
Seamen
,
D. B.
Sinars
,
S. A.
Slutz
,
I. C.
Smith
,
K. W.
Struve
,
W. A.
Stygar
,
R. A.
Vesey
,
E. A.
Weinbrecht
,
D. F.
Wenger
, and
E. P.
Yu
, “
Pulsed-power-driven high energy density physics and inertial confinement fusion research
,”
Phys. Plasmas
12
,
055503
(
2005
).
17.
S. X.
Hu
,
B.
Militzer
,
V. N.
Goncharov
, and
S.
Skupsky
, “
First-principles equation-of-state table of deuterium for inertial confinement fusion applications
,”
Phys. Rev. B
84
,
224109
(
2011
).
18.
A. L.
Kritcher
,
D. C.
Swift
,
T.
Döppner
,
B.
Bachmann
,
L. X.
Benedict
,
G. W.
Collins
,
J. L.
DuBois
,
F.
Elsner
,
G.
Fontaine
,
J. A.
Gaffney
,
S.
Hamel
,
A.
Lazicki
,
W. R.
Johnson
,
N.
Kostinski
,
D.
Kraus
,
M. J.
MacDonald
,
B.
Maddox
,
M. E.
Martin
,
P.
Neumayer
,
A.
Nikroo
,
J.
Nilsen
,
B. A.
Remington
,
D.
Saumon
,
P. A.
Sterne
,
W.
Sweet
,
A. A.
Correa
,
H. D.
Whitley
,
R. W.
Falcone
, and
S. H.
Glenzer
, “
A measurement of the equation of state of carbon envelopes of white dwarfs
,”
Nature
584
,
51
54
(
2020
).
19.
D.
Kraus
,
J.
Vorberger
,
A.
Pak
,
N. J.
Hartley
,
L. B.
Fletcher
,
S.
Frydrych
,
E.
Galtier
,
E. J.
Gamboa
,
D. O.
Gericke
,
S. H.
Glenzer
,
E.
Granados
,
M. J.
MacDonald
,
A. J.
MacKinnon
,
E. E.
McBride
,
I.
Nam
,
P.
Neumayer
,
M.
Roth
,
A. M.
Saunders
,
A. K.
Schuster
,
P.
Sun
,
T.
van Driel
,
T.
Döppner
, and
R. W.
Falcone
, “
Formation of diamonds in laser-compressed hydrocarbons at planetary interior conditions
,”
Nat. Astron.
1
,
606
611
(
2017
).
20.
A. K.
Schuster
,
N. J.
Hartley
,
J.
Vorberger
,
T.
Döppner
,
T.
van Driel
,
R. W.
Falcone
,
L. B.
Fletcher
,
S.
Frydrych
,
E.
Galtier
,
E. J.
Gamboa
,
D. O.
Gericke
,
S. H.
Glenzer
,
E.
Granados
,
M. J.
MacDonald
,
A. J.
MacKinnon
,
E. E.
McBride
,
I.
Nam
,
P.
Neumayer
,
A.
Pak
,
I.
Prencipe
,
K.
Voigt
,
A. M.
Saunders
,
P.
Sun
, and
D.
Kraus
, “
Measurement of diamond nucleation rates from hydrocarbons at conditions comparable to the interiors of icy giant planets
,”
Phys. Rev. B
101
,
054301
(
2020
).
21.
R. F.
Smith
,
J. H.
Eggert
,
R.
Jeanloz
,
T. S.
Duffy
,
D. G.
Braun
,
J. R.
Patterson
,
R. E.
Rudd
,
J.
Biener
,
A. E.
Lazicki
,
A. V.
Hamza
,
J.
Wang
,
T.
Braun
,
L. X.
Benedict
,
P. M.
Celliers
, and
G. W.
Collins
, “
Ramp compression of diamond to five terapascals
,”
Nature
511
,
330
333
(
2014
).
22.
A.
Lazicki
,
D.
McGonegle
,
J. R.
Rygg
,
D. G.
Braun
,
D. C.
Swift
,
M. G.
Gorman
,
R. F.
Smith
,
P. G.
Heighway
,
A.
Higginbotham
,
M. J.
Suggit
,
D. E.
Fratanduono
,
F.
Coppari
,
C. E.
Wehrenberg
,
R. G.
Kraus
,
D.
Erskine
,
J. V.
Bernier
,
J. M.
McNaney
,
R. E.
Rudd
,
G. W.
Collins
,
J. H.
Eggert
, and
J. S.
Wark
, “
Metastability of diamond ramp-compressed to 2 terapascals
,”
Nature
589
,
532
535
(
2021
).
23.
J.
Lütgert
,
J.
Vorberger
,
N. J.
Hartley
,
K.
Voigt
,
M.
Rödel
,
A. K.
Schuster
,
A.
Benuzzi-Mounaix
,
S.
Brown
,
T. E.
Cowan
,
E.
Cunningham
,
T.
Döppner
,
R. W.
Falcone
,
L. B.
Fletcher
,
E.
Galtier
,
S. H.
Glenzer
,
A.
Laso Garcia
,
D. O.
Gericke
,
P. A.
Heimann
,
H. J.
Lee
,
E. E.
McBride
,
A.
Pelka
,
I.
Prencipe
,
A. M.
Saunders
,
M.
Schölmerich
,
M.
Schörner
,
P.
Sun
,
T.
Vinci
,
A.
Ravasio
, and
D.
Kraus
, “
Measuring the structure and equation of state of polyethylene terephthalate at megabar pressures
,”
Sci. Rep.
11
,
12883
(
2021
).
24.
S.
White
,
B.
Kettle
,
J.
Vorberger
,
C. L. S.
Lewis
,
S. H.
Glenzer
,
E.
Gamboa
,
B.
Nagler
,
F.
Tavella
,
H. J.
Lee
,
C. D.
Murphy
,
D. O.
Gericke
, and
D.
Riley
, “
Time-dependent effects in melting and phase change for laser-shocked iron
,”
Phys. Rev. Res.
2
,
033366
(
2020
).
25.
M. L.
Brongersma
,
N. J.
Halas
, and
P.
Nordlander
, “
Plasmon-induced hot carrier science and technology
,”
Nat. Nanotechnol.
10
,
25
34
(
2015
).
26.
P.
Borlido
,
T.
Aull
,
A. W.
Huran
,
F.
Tran
,
M. A. L.
Marques
, and
S.
Botti
, “
Large-scale benchmark of exchange–correlation functionals for the determination of electronic band gaps of solids
,”
J. Chem. Theory Comput.
15
,
5069
5079
(
2019
).
27.
J.
Paier
,
M.
Marsman
, and
G.
Kresse
, “
Why does the B3LYP hybrid functional fail for metals?
,”
J. Chem. Phys.
127
,
024103
(
2007
).
28.
A. D.
Becke
, “
Density-functional thermochemistry. III. The role of exact exchange
,”
J. Chem. Phys.
98
,
5648
5652
(
1993
).
29.
N. D.
Mermin
, “
Thermal properties of the inhomogeneous electron gas
,”
Phys. Rev.
137
,
A1441
A1443
(
1965
).
30.
G.
Giuliani
and
G.
Vignale
,
Quantum Theory of the Electron Liquid, Masters Series in Physics and Astronomy
(
Cambridge University Press
,
2005
).
31.
S.
Pittalis
,
C. R.
Proetto
,
A.
Floris
,
A.
Sanna
,
C.
Bersier
,
K.
Burke
, and
E. K. U.
Gross
, “
Exact conditions in finite-temperature density-functional theory
,”
Phys. Rev. Lett.
107
,
163001
(
2011
).
32.
S.
Groth
,
T.
Dornheim
,
T.
Sjostrom
,
F. D.
Malone
,
W. M. C.
Foulkes
, and
M.
Bonitz
, “
Ab initio exchange–correlation free energy of the uniform electron gas at warm dense matter conditions
,”
Phys. Rev. Lett.
119
,
135001
(
2017
).
33.
V. V.
Karasiev
,
J. W.
Dufty
, and
S. B.
Trickey
, “
Nonempirical semilocal free-energy density functional for matter under extreme conditions
,”
Phys. Rev. Lett.
120
,
076401
(
2018
).
34.
V. V.
Karasiev
,
T.
Sjostrom
, and
S. B.
Trickey
, “
Generalized-gradient-approximation noninteracting free-energy functionals for orbital-free density functional calculations
,”
Phys. Rev. B
86
,
115101
(
2012
).
35.
D. I.
Mihaylov
,
V. V.
Karasiev
, and
S. X.
Hu
, “
Thermal hybrid exchange-correlation density functional for improving the description of warm dense matter
,”
Phys. Rev. B
101
,
245141
(
2020
).
36.
T.
Dornheim
,
Z.
Moldabekov
,
P.
Tolias
,
M.
Böhme
, and
J.
Vorberger
, “
Physical insights from imaginary-time density–density correlation functions
,” arXiv:2209.02254 (
2022
).
37.
M.
Böhme
,
Z. A.
Moldabekov
,
J.
Vorberger
, and
T.
Dornheim
, “
Static electronic density response of warm dense hydrogen: Ab initio path integral Monte Carlo simulations
,”
Phys. Rev. Lett.
129
,
066402
(
2022
).
38.
M.
Böhme
,
Z. A.
Moldabekov
,
J.
Vorberger
, and
T.
Dornheim
, “
Ab initio path integral Monte Carlo simulations of hydrogen snapshots at warm dense matter conditions
,”
Phys. Rev. E
107
,
015206
(
2023
).
39.
A.
Denoeud
,
N.
Ozaki
,
A.
Benuzzi-Mounaix
,
H.
Uranishi
,
Y.
Kondo
,
R.
Kodama
,
E.
Brambrink
,
A.
Ravasio
,
M.
Bocoum
,
J.-M.
Boudenne
,
M.
Harmand
,
F.
Guyot
,
S.
Mazevet
,
D.
Riley
,
M.
Makita
,
T.
Sano
,
Y.
Sakawa
,
Y.
Inubushi
,
G.
Gregori
,
M.
Koenig
, and
G.
Morard
, “
Dynamic x-ray diffraction observation of shocked solid iron up to 170 GPa
,”
Proc. Natl. Acad. Sci.
113
,
7745
7749
(
2016
).
40.
N.
Booth
,
A. P. L.
Robinson
,
P.
Hakel
,
R. J.
Clarke
,
R. J.
Dance
,
D.
Doria
,
L. A.
Gizzi
,
G.
Gregori
,
P.
Koester
,
L.
Labate
,
T.
Levato
,
B.
Li
,
M.
Makita
,
R. C.
Mancini
,
J.
Pasley
,
P. P.
Rajeev
,
D.
Riley
,
E.
Wagenaars
,
J. N.
Waugh
, and
N. C.
Woolsey
, “
Laboratory measurements of resistivity in warm dense plasmas relevant to the microphysics of brown dwarfs
,”
Nat. Commun.
6
,
8742
(
2015
).
41.
L.
Fiedler
,
K.
Shah
,
M.
Bussmann
, and
A.
Cangi
, “
Deep dive into machine learning density functional theory for materials science and chemistry
,”
Phys. Rev. Mater.
6
,
040301
(
2022
).
42.
Z. A.
Moldabekov
,
M.
Böhme
,
J.
Vorberger
,
D.
Blaschke
, and
T.
Dornheim
, “
Ab initio static exchange–correlation kernel across Jacob’s ladder without functional derivatives
,”
J. Chem. Theory Comp.
(published online
2022
).
43.
G.
Senatore
,
S.
Moroni
, and
D. M.
Ceperley
, “
Local field factor and effective potentials in liquid metals
,”
J. Non-Cryst. Solids
205–207
,
851
854
(
1996
).
44.
Z.
Moldabekov
,
H.
Kählert
,
T.
Dornheim
,
S.
Groth
,
M.
Bonitz
, and
T. S.
Ramazanov
, “
Dynamical structure factor of strongly coupled ions in a dense quantum plasma
,”
Phys. Rev. E: Stat. Phys., Plasmas, Fluids, Relat. Interdiscip. Top.
99
,
053203
(
2019
).
45.
Z. A.
Moldabekov
,
S.
Groth
,
T.
Dornheim
,
H.
Kählert
,
M.
Bonitz
, and
T. S.
Ramazanov
, “
Structural characteristics of strongly coupled ions in a dense quantum plasma
,”
Phys. Rev. E
98
,
023207
(
2018
).
46.
J.
Vorberger
,
Z.
Donko
,
I. M.
Tkachenko
, and
D. O.
Gericke
, “
Dynamic ion structure factor of warm dense matter
,”
Phys. Rev. Lett.
109
,
225001
(
2012
).
47.
T.
Dornheim
,
P.
Tolias
,
Z. A.
Moldabekov
,
A.
Cangi
, and
J.
Vorberger
, “
Effective electronic forces and potentials from ab initio path integral Monte Carlo simulations
,”
J. Chem. Phys.
156
,
244113
(
2022
).
48.
Z. A.
Moldabekov
,
T.
Dornheim
, and
M.
Bonitz
, “
Screening of a test charge in a free-electron gas at warm dense matter and dense non-ideal plasma conditions
,”
Contrib. Plasma Phys.
62
,
e202000176
(
2022
).
49.
Z. A.
Moldabekov
,
S.
Groth
,
T.
Dornheim
,
M.
Bonitz
, and
T. S.
Ramazanov
, “
Ion potential in non-ideal dense quantum plasmas
,”
Contrib. Plasma Phys.
57
,
532
538
(
2017
).
50.
A.
Diaw
and
M. S.
Murillo
, “
A viscous quantum hydrodynamics model based on dynamic density functional theory
,”
Sci. Rep.
7
,
15352
(
2017
).
51.
Z. A.
Moldabekov
,
M.
Bonitz
, and
T. S.
Ramazanov
, “
Theoretical foundations of quantum hydrodynamics for plasmas
,”
Phys. Plasmas
25
,
031903
(
2018
).
52.
Z.
Moldabekov
,
T.
Dornheim
,
G.
Gregori
,
F.
Graziani
,
M.
Bonitz
, and
A.
Cangi
, “
Towards a quantum fluid theory of correlated many-fermion systems from first principles
,”
SciPost Phys.
12
,
62
(
2022
).
53.
Z.
Moldabekov
,
T.
Dornheim
,
M.
Böhme
,
J.
Vorberger
, and
A.
Cangi
, “
The relevance of electronic perturbations in the warm dense electron gas
,”
J. Chem. Phys.
155
,
124116
(
2021
).
54.
Z.
Moldabekov
,
T.
Dornheim
,
J.
Vorberger
, and
A.
Cangi
, “
Benchmarking exchange-correlation functionals in the spin-polarized inhomogeneous electron gas under warm dense conditions
,”
Phys. Rev. B
105
,
035134
(
2022
).
55.
J. P.
Perdew
and
Y.
Wang
, “
Accurate and simple analytic representation of the electron-gas correlation energy
,”
Phys. Rev. B
45
,
13244
13249
(
1992
).
56.
J. P.
Perdew
,
K.
Burke
, and
M.
Ernzerhof
, “
Generalized gradient approximation made simple
,”
Phys. Rev. Lett.
77
,
3865
3868
(
1996
).
57.
J. P.
Perdew
,
A.
Ruzsinszky
,
G. I.
Csonka
,
O. A.
Vydrov
,
G. E.
Scuseria
,
L. A.
Constantin
,
X.
Zhou
, and
K.
Burke
, “
Restoring the density-gradient expansion for exchange in solids and surfaces
,”
Phys. Rev. Lett.
100
,
136406
(
2008
).
58.
R.
Armiento
and
A. E.
Mattsson
, “
Functional designed to include surface effects in self-consistent density functional theory
,”
Phys. Rev. B
72
,
085108
(
2005
).
59.
J.
Sun
,
A.
Ruzsinszky
, and
J. P.
Perdew
, “
Strongly constrained and appropriately normed semilocal density functional
,”
Phys. Rev. Lett.
115
,
036402
(
2015
).
60.
T.
Dornheim
,
J.
Vorberger
,
S.
Groth
,
N.
Hoffmann
,
Z. A.
Moldabekov
, and
M.
Bonitz
, “
The static local field correction of the warm dense electron gas: An ab initio path integral Monte Carlo study and machine learning representation
,”
J. Chem. Phys.
151
,
194104
(
2019
).
61.
T.
Dornheim
,
J.
Vorberger
, and
M.
Bonitz
, “
Nonlinear electronic density response in warm dense matter
,”
Phys. Rev. Lett.
125
,
085001
(
2020
).
62.
T.
Dornheim
,
M.
Böhme
,
Z. A.
Moldabekov
,
J.
Vorberger
, and
M.
Bonitz
, “
Density response of the warm dense electron gas beyond linear response theory: Excitation of harmonics
,”
Phys. Rev. Res.
3
,
033231
(
2021
).
63.
S.
Kümmel
and
L.
Kronik
, “
Orbital-dependent density functionals: Theory and applications
,”
Rev. Mod. Phys.
80
,
3
60
(
2008
).
64.
R.
Garrick
,
A.
Natan
,
T.
Gould
, and
L.
Kronik
, “
Exact generalized Kohn-Sham theory for hybrid functionals
,”
Phys. Rev. X
10
,
021040
(
2020
).
65.
J. P.
Perdew
,
M.
Ernzerhof
, and
K.
Burke
, “
Rationale for mixing exact exchange with density functional approximations
,”
J. Chem. Phys.
105
,
9982
9985
(
1996
).
66.
D.
Zhang
and
D. G.
Truhlar
, “
Unmasking static correlation error in hybrid Kohn–Sham density functional theory
,”
J. Chem. Theory Comput.
16
,
5432
5440
(
2020
).
67.
M.
Greiner
,
P.
Carrier
, and
A.
Görling
, “
Extension of exact-exchange density functional theory of solids to finite temperatures
,”
Phys. Rev. B
81
,
155119
(
2010
).
68.
J.
Heyd
,
G. E.
Scuseria
, and
M.
Ernzerhof
, “
Erratum: Hybrid functionals based on a screened Coulomb potential” [J. Chem. Phys. 118, 8207 (2003)]
,”
J. Chem. Phys.
124
,
219906
(
2006
).
69.
A. V.
Krukau
,
O. A.
Vydrov
,
A. F.
Izmaylov
, and
G. E.
Scuseria
, “
Influence of the exchange screening parameter on the performance of screened hybrid functionals
,”
J. Chem. Phys.
125
,
224106
(
2006
).
70.
J.
Heyd
,
G. E.
Scuseria
, and
M.
Ernzerhof
, “
Hybrid functionals based on a screened Coulomb potential
,”
J. Chem. Phys.
118
,
8207
8215
(
2003
).
71.
C.
Adamo
and
V.
Barone
, “
Toward reliable density functional methods without adjustable parameters: The PBE0 model
,”
J. Chem. Phys.
110
,
6158
6170
(
1999
).
72.
P.
Cortona
, “
Note: Theoretical mixing coefficients for hybrid functionals
,”
J. Chem. Phys.
136
,
086101
(
2012
).
73.
P. J.
Stephens
,
F. J.
Devlin
,
C. F.
Chabalowski
, and
M. J.
Frisch
, “
Ab initio calculation of vibrational absorption and circular dichroism spectra using density functional force fields
,”
J. Phys. Chem.
98
,
11623
11627
(
1994
).
74.
C.
Fortmann
,
A.
Wierling
, and
G.
Röpke
, “
Influence of local-field corrections on Thomson scattering in collision-dominated two-component plasmas
,”
Phys. Rev. E
81
,
026405
(
2010
).
75.
T. R.
Preston
,
K.
Appel
,
E.
Brambrink
,
B.
Chen
,
L. B.
Fletcher
,
C.
Fortmann-Grote
,
S. H.
Glenzer
,
E.
Granados
,
S.
Göde
,
Z.
Konôpková
,
H. J.
Lee
,
H.
Marquardt
,
E. E.
McBride
,
B.
Nagler
,
M.
Nakatsutsumi
,
P.
Sperling
,
B. B. L.
Witte
, and
U.
Zastrau
, “
Measurements of the momentum-dependence of plasmonic excitations in matter around 1 Mbar using an X-ray free electron laser
,”
Appl. Phys. Lett.
114
,
014101
(
2019
).
76.
T.
Dornheim
,
A.
Cangi
,
K.
Ramakrishna
,
M.
Böhme
,
S.
Tanaka
, and
J.
Vorberger
, “
Effective static approximation: A fast and reliable tool for warm-dense matter theory
,”
Phys. Rev. Lett.
125
,
235001
(
2020
).
77.
X.
Zan
,
C.
Lin
,
Y.
Hou
, and
J.
Yuan
, “
Local field correction to ionization potential depression of ions in warm or hot dense matter
,”
Phys. Rev. E
104
,
025203
(
2021
).
78.
O.
Gunnarsson
and
B. I.
Lundqvist
, “
Exchange and correlation in atoms, molecules, and solids by the spin-density-functional formalism
,”
Phys. Rev. B
13
,
4274
4298
(
1976
).
79.
B. G.
Janesko
,
T. M.
Henderson
, and
G. E.
Scuseria
, “
Screened hybrid density functionals for solid-state chemistry and physics
,”
Phys. Chem. Chem. Phys.
11
,
443
454
(
2009
).
80.
B. B. L.
Witte
,
L. B.
Fletcher
,
E.
Galtier
,
E.
Gamboa
,
H. J.
Lee
,
U.
Zastrau
,
R.
Redmer
,
S. H.
Glenzer
, and
P.
Sperling
, “
Warm dense matter demonstrating non-Drude conductivity from observations of nonlinear plasmon damping
,”
Phys. Rev. Lett.
118
,
225001
(
2017
).
81.
A.
Ravasio
,
M.
Bethkenhagen
,
J.-A.
Hernandez
,
A.
Benuzzi-Mounaix
,
F.
Datchi
,
M.
French
,
M.
Guarguaglini
,
F.
Lefevre
,
S.
Ninet
,
R.
Redmer
, and
T.
Vinci
, “
Metallization of shock-compressed liquid ammonia
,”
Phys. Rev. Lett.
126
,
025003
(
2021
).
82.
S.
Moroni
,
D. M.
Ceperley
, and
G.
Senatore
, “
Static response and local field factor of the electron gas
,”
Phys. Rev. Lett.
75
,
689
692
(
1995
).
83.
J. C.
Smith
,
A.
Pribram-Jones
, and
K.
Burke
, “
Exact thermal density functional theory for a model system: Correlation components and accuracy of the zero-temperature exchange-correlation approximation
,”
Phys. Rev. B
93
,
245131
(
2016
).
84.
P. F.
Loos
and
P. M. W.
Gill
, “
The uniform electron gas
,”
Comput. Mol. Sci.
6
,
410
429
(
2016
).
85.
A. A.
Kugler
, “
Theory of the local field correction in an electron gas
,”
J. Stat. Phys.
12
,
35
(
1975
).
86.
S.
Ichimaru
,
H.
Iyetomi
, and
S.
Tanaka
, “
Statistical physics of dense plasmas: Thermodynamics, transport coefficients and dynamic correlations
,”
Phys. Rep.
149
,
91
205
(
1987
).
87.
K.
Chen
and
K.
Haule
, “
A combined variational and diagrammatic quantum Monte Carlo approach to the many-electron problem
,”
Nat. Commun.
10
,
3725
(
2019
).
88.
P.-C.
Hou
,
B.-Z.
Wang
,
K.
Haule
,
Y.
Deng
, and
K.
Chen
, “
Exchange-correlation effect in the charge response of a warm dense electron gas
,”
Phys. Rev. B
106
,
L081126
(
2022
).
89.
T.
Dornheim
,
Z. A.
Moldabekov
, and
P.
Tolias
, “
Analytical representation of the local field correction of the uniform electron gas within the effective static approximation
,”
Phys. Rev. B
103
,
165102
(
2021
).
90.
T.
Dornheim
,
J.
Vorberger
, and
Z. A.
Moldabekov
, “
Nonlinear density response and higher order correlation functions in warm dense matter
,”
J. Phys. Soc. Jpn.
90
,
104002
(
2021
).
91.
Z.
Moldabekov
,
J.
Vorberger
, and
T.
Dornheim
, “
Density functional theory perspective on the nonlinear response of correlated electrons across temperature regimes
,”
J. Chem. Theory Comput.
18
,
2900
2912
(
2022
).
92.
T.
Ott
,
H.
Thomsen
,
J. W.
Abraham
,
T.
Dornheim
, and
M.
Bonitz
, “
Recent progress in the theory and simulation of strongly correlated plasmas: Phase transitions, transport, quantum, and magnetic field effects
,”
Eur. Phys. J. D
72
,
84
(
2018
).
93.
X.
Gonze
,
B.
Amadon
,
G.
Antonius
,
F.
Arnardi
,
L.
Baguet
,
J.-M.
Beuken
,
J.
Bieder
,
F.
Bottin
,
J.
Bouchet
,
E.
Bousquet
,
N.
Brouwer
,
F.
Bruneval
,
G.
Brunin
,
T.
Cavignac
,
J.-B.
Charraud
,
W.
Chen
,
M.
Côté
,
S.
Cottenier
,
J.
Denier
,
G.
Geneste
,
P.
Ghosez
,
M.
Giantomassi
,
Y.
Gillet
,
O.
Gingras
,
D. R.
Hamann
,
G.
Hautier
,
X.
He
,
N.
Helbig
,
N.
Holzwarth
,
Y.
Jia
,
F.
Jollet
,
W.
Lafargue-Dit-Hauret
,
K.
Lejaeghere
,
M. A. L.
Marques
,
A.
Martin
,
C.
Martins
,
H. P. C.
Miranda
,
F.
Naccarato
,
K.
Persson
,
G.
Petretto
,
V.
Planes
,
Y.
Pouillon
,
S.
Prokhorenko
,
F.
Ricci
,
G.-M.
Rignanese
,
A. H.
Romero
,
M. M.
Schmitt
,
M.
Torrent
,
M. J.
van Setten
,
B. V.
Troeye
,
M. J.
Verstraete
,
G.
Zérah
, and
J. W.
Zwanziger
, “
The ABINIT project: Impact, environment and recent developments
,”
Comput. Phys. Commun.
248
,
107042
(
2020
).
94.
A. H.
Romero
,
D. C.
Allan
,
B.
Amadon
,
G.
Antonius
,
T.
Applencourt
,
L.
Baguet
,
J.
Bieder
,
F.
Bottin
,
J.
Bouchet
,
E.
Bousquet
,
F.
Bruneval
,
G.
Brunin
,
D.
Caliste
,
M.
Côté
,
J.
Denier
,
C.
Dreyer
,
P.
Ghosez
,
M.
Giantomassi
,
Y.
Gillet
,
O.
Gingras
,
D. R.
Hamann
,
G.
Hautier
,
F.
Jollet
,
G.
Jomard
,
A.
Martin
,
H. P. C.
Miranda
,
F.
Naccarato
,
G.
Petretto
,
N. A.
Pike
,
V.
Planes
,
S.
Prokhorenko
,
T.
Rangel
,
F.
Ricci
,
G.-M.
Rignanese
,
M.
Royo
,
M.
Stengel
,
M.
Torrent
,
M. J.
van Setten
,
B. V.
Troeye
,
M. J.
Verstraete
,
J.
Wiktor
,
J. W.
Zwanziger
, and
X.
Gonze
, “
ABINIT: Overview and focus on selected capabilities
,”
J. Chem. Phys.
152
,
124102
(
2020
).
95.
X.
Gonze
,
F.
Jollet
,
F.
Abreu Araujo
,
D.
Adams
,
B.
Amadon
,
T.
Applencourt
,
C.
Audouze
,
J.-M.
Beuken
,
J.
Bieder
,
A.
Bokhanchuk
,
E.
Bousquet
,
F.
Bruneval
,
D.
Caliste
,
M.
Côté
,
F.
Dahm
,
F.
Da Pieve
,
M.
Delaveau
,
M.
Di Gennaro
,
B.
Dorado
,
C.
Espejo
,
G.
Geneste
,
L.
Genovese
,
A.
Gerossier
,
M.
Giantomassi
,
Y.
Gillet
,
D. R.
Hamann
,
L.
He
,
G.
Jomard
,
J.
Laflamme Janssen
,
S.
Le Roux
,
A.
Levitt
,
A.
Lherbier
,
F.
Liu
,
I.
Lukačević
,
A.
Martin
,
C.
Martins
,
M. J. T.
Oliveira
,
S.
Poncé
,
Y.
Pouillon
,
T.
Rangel
,
G.-M.
Rignanese
,
A. H.
Romero
,
B.
Rousseau
,
O.
Rubel
,
A. A.
Shukri
,
M.
Stankovski
,
M.
Torrent
,
M. J.
Van Setten
,
B.
Van Troeye
,
M. J.
Verstraete
,
D.
Waroquiers
,
J.
Wiktor
,
B.
Xu
,
A.
Zhou
, and
J. W.
Zwanziger
, “
Recent developments in the ABINIT software package
,”
Comput. Phys. Commun.
205
,
106
131
(
2016
).
96.
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. R.
Hamann
,
P.
Hermet
,
F.
Jollet
,
G.
Jomard
,
S.
Leroux
,
M.
Mancini
,
S.
Mazevet
,
M. J. T.
Oliveira
,
G.
Onida
,
Y.
Pouillon
,
T.
Rangel
,
G.-M.
Rignanese
,
D.
Sangalli
,
R.
Shaltaf
,
M.
Torrent
,
M. J.
Verstraete
,
G.
Zerah
, and
J. W.
Zwanziger
, “
ABINIT: First-principles approach to material and nanosystem properties
,”
Comput. Phys. Commun.
180
,
2582
2615
(
2009
).
97.
X.
Gonze
,
G.-M.
Rignanese
,
M.
Verstraete
,
J.-M.
Beuken
,
Y.
Pouillon
,
R.
Caracas
,
F.
Jollet
,
M.
Torrent
,
G.
Zerah
,
M.
Mikami
,
P.
Ghosez
,
M.
Veithen
,
J.-Y.
Raty
,
V.
Olevano
,
F.
Bruneval
,
L.
Reining
,
R.
Godby
,
G.
Onida
, and
D. H. D. C.
Allan
, “
A brief introduction to the ABINIT software package
,”
Z. Kristallogr. - Cryst. Mater.
220
,
558
562
(
2005
).
98.
X.
Gonze
,
J.-M.
Beuken
,
R.
Caracas
,
F.
Detraux
,
M.
Fuchs
,
G.-M.
Rignanese
,
L.
Sindic
,
M.
Verstraete
,
G.
Zerah
,
F.
Jollet
,
M.
Torrent
,
A.
Roy
,
M.
Mikami
,
P.
Ghosez
,
J.-Y.
Raty
, and
D. C.
Allan
, “
First-principles computation of material properties: The ABINIT software project
,”
Comput. Mater. Sci.
25
,
478
492
(
2002
).
99.
Z. A.
Moldabekov
,
T.
Dornheim
, and
A.
Cangi
, “
Thermal excitation signals in the inhomogeneous warm dense electron gas
,”
Sci. Rep.
12
,
1093
(
2022
).
100.
Z. A.
Moldabekov
,
M.
Pavanello
,
M. P.
Boehme
,
J.
Vorberger
, and
T.
Dornheim
, “
Linear-response time-dependent density functional theory approach to warm dense matter with adiabatic exchange-correlation kernels
,” arXiv:2302.04822.
101.
Z. A.
Moldabekov
,
M.
Lokamani
,
J.
Vorberger
,
A.
Cangi
, and
T.
Dornheim
,
J. Phys. Chem. Lett.
14
(
5
),
1326
1333
(
2023
).
102.
T.
Dornheim
,
J.
Vorberger
, and
M.
Bonitz
, “
Nonlinear electronic density response in warm dense matter
,” arXiv:2004.03229 [cond-mat, physics:physics] (
2020
).
103.
D. M.
Ceperley
, “
Path integrals in the theory of condensed helium
,”
Rev. Mod. Phys.
67
,
279
(
1995
).
104.
T.
Dornheim
, “
Fermion sign problem in path integral Monte Carlo simulations: Quantum dots, ultracold atoms, and warm dense matter
,”
Phys. Rev. E
100
,
023307
(
2019
).
105.
T.
Dornheim
,
M.
Böhme
,
B.
Militzer
, and
J.
Vorberger
, “
Ab initio path integral Monte Carlo approach to the momentum distribution of the uniform electron gas at finite temperature without fixed nodes
,”
Phys. Rev. B
103
,
205142
(
2021
).
106.
M.
Boninsegni
,
N. V.
Prokofev
, and
B. V.
Svistunov
, “
Worm algorithm and diagrammatic Monte Carlo: A new approach to continuous-space path integral Monte Carlo simulations
,”
Phys. Rev. E
74
,
036701
(
2006
).
107.
M.
Boninsegni
,
N.
Prokof’ev
, and
B.
Svistunov
, “
Worm algorithm for continuous-space path integral Monte Carlo simulations
,”
Phys. Rev. Lett.
96
,
070601
(
2006
).
108.
A. E.
Mattsson
,
R.
Armiento
,
J.
Paier
,
G.
Kresse
,
J. M.
Wills
, and
T. R.
Mattsson
, “
The AM05 density functional applied to solids
,”
J. Chem. Phys.
128
,
084714
(
2008
).
109.
X.
Shao
,
W.
Mi
, and
M.
Pavanello
, “
Revised Huang-Carter nonlocal kinetic energy functional for semiconductors and their surfaces
,”
Phys. Rev. B
104
,
045118
(
2021
).
110.
T.
Dornheim
,
Z.
Moldabekov
,
J.
Vorberger
,
H.
Kählert
, and
M.
Bonitz
, “
Electronic pair alignment and roton feature in the warm dense electron gas
,”
Commun. Phys.
5
,
304
(
2022
).
111.
T.
Dornheim
,
Z. A.
Moldabekov
,
J.
Vorberger
, and
S.
Groth
, “
Ab initio path integral Monte Carlo simulation of the uniform electron gas in the high energy density regime
,”
Plasma Phys. Controlled Fusion
62
,
075003
(
2020
).
112.
S. H.
Glenzer
,
L. B.
Fletcher
,
E.
Galtier
,
B.
Nagler
,
R.
Alonso-Mori
,
B.
Barbrel
,
S. B.
Brown
,
D. A.
Chapman
,
Z.
Chen
,
C. B.
Curry
,
F.
Fiuza
,
E.
Gamboa
,
M.
Gauthier
,
D. O.
Gericke
,
A.
Gleason
,
S.
Goede
,
E.
Granados
,
P.
Heimann
,
J.
Kim
,
D.
Kraus
,
M. J.
MacDonald
,
A. J.
Mackinnon
,
R.
Mishra
,
A.
Ravasio
,
C.
Roedel
,
P.
Sperling
,
W.
Schumaker
,
Y. Y.
Tsui
,
J.
Vorberger
,
U.
Zastrau
,
A.
Fry
,
W. E.
White
,
J. B.
Hasting
, and
H. J.
Lee
, “
Matter under extreme conditions experiments at the Linac Coherent Light Source
,”
J. Phys. B: At., Mol. Opt. Phys.
49
,
092001
(
2016
).
113.
S.
Frydrych
,
J.
Vorberger
,
N. J.
Hartley
,
A. K.
Schuster
,
K.
Ramakrishna
,
A. M.
Saunders
,
T.
van Driel
,
R. W.
Falcone
,
L. B.
Fletcher
,
E.
Galtier
,
E. J.
Gamboa
,
S. H.
Glenzer
,
E.
Granados
,
M. J.
MacDonald
,
A. J.
MacKinnon
,
E. E.
McBride
,
I.
Nam
,
P.
Neumayer
,
A.
Pak
,
K.
Voigt
,
M.
Roth
,
P.
Sun
,
D. O.
Gericke
,
T.
Döppner
, and
D.
Kraus
, “
Demonstration of X-ray Thomson scattering as diagnostics for miscibility in warm dense matter
,”
Nat. Commun.
11
,
2620
(
2020
).
114.
Seehttp://doi.org/10.14278/rodare.2164 for the data available according to the FAIR principles on the Rossendorf Data Repository.
Published open access through an agreement with Helmholtz-Zentrum Dresden-Rossendorf