Accurate knowledge of the properties of hydrogen at high compression is crucial for astrophysics (e.g., planetary and stellar interiors, brown dwarfs, atmosphere of compact stars) and laboratory experiments, including inertial confinement fusion. There exists experimental data for the equation of state, conductivity, and Thomson scattering spectra. However, the analysis of the measurements at extreme pressures and temperatures typically involves additional model assumptions, which makes it difficult to assess the accuracy of the experimental data rigorously. On the other hand, theory and modeling have produced extensive collections of data. They originate from a very large variety of models and simulations including path integral Monte Carlo (PIMC) simulations, density functional theory (DFT), chemical models, machine-learned models, and combinations thereof. At the same time, each of these methods has fundamental limitations (fermion sign problem in PIMC, approximate exchange–correlation functionals of DFT, inconsistent interaction energy contributions in chemical models, etc.), so for some parameter ranges accurate predictions are difficult. Recently, a number of breakthroughs in first principles PIMC as well as in DFT simulations were achieved which are discussed in this review. Here we use these results to benchmark different simulation methods. We present an update of the hydrogen phase diagram at high pressures, the expected phase transitions, and thermodynamic properties including the equation of state and momentum distribution. Furthermore, we discuss available dynamic results for warm dense hydrogen, including the conductivity, dynamic structure factor, plasmon dispersion, imaginary-time structure, and density response functions. We conclude by outlining strategies to combine different simulations to achieve accurate theoretical predictions that are based on first principles.

Hydrogen was the first element that formed after the Big Bang and it has remained the most abundant species in the Universe. Its properties, therefore, determine the structure and evolution of astrophysical objects including stars, many planets, and interstellar gas clouds. These objects exist in a huge range of temperatures—spanning from a few Kelvin to billions of Kelvin—and pressures that extend from zero to trillions of atmospheres in neutron stars. Astrophysical observations have revealed extensive though indirect information about the properties of hydrogen. For example, observations of the oscillations of the sun's surface are employed to constrain its interior properties (“helioseismology”) (e.g., Refs. 1–3). More recently, a major source of knowledge has come from laboratory experiments that reach ever higher pressures. Examples include static compression using diamond anvil cells (DAC) (e.g., Ref. 4), dynamic compression using shock waves that are generated by impactors,5,6 explosives (e.g., Refs. 7 and 8), or high-intensity lasers.9 Extracting data from these measurements and translating them into static, dynamic, or optical properties of hydrogen can be challenging. The accuracy of the diagnostics is usually severely limited due to the short observation time and relaxation processes, unknown intrinsic properties of the equipment, and so forth. These gaps in knowledge are often bridged with simple models for the equation of state (EOS) or hydrodynamics, the validity of which may be questionable. All of this renders the accuracy of many measurements unclear. Moreover, high-pressure experiments are challenging and costly, and available only at a limited number of laboratories. Therefore, there is a very high demand for theoretical analysis and computer experiments that would allow for an improved interpretation of experimental data but also independent and reliable predictions.

Aside from experiments that focus on basic science aspects of the properties of hydrogen at high pressure, there is a rapidly increasing interest in technological applications of hydrogen. This includes inertial confinement fusion (ICF)10–12 and the use of hydrogen in green energy applications. In particular, the ignition campaign at the NIF has recently reported major breakthroughs. Remarkably, the fusion gain exceeded unity (not accounting for wasted energy).13–16 Technological advances, in particular ICF, rely heavily on theoretical modeling that allows one to understand the relevant physical processes, make reliable predictions, and enable the optimization of experiments. Moreover, the direct measurement even of basic parameters such as the temperature or density is often not possible; instead, they have to be inferred indirectly from other observations, which, in turn, requires theoretical results for different observables.17 

The theory and modeling of dense hydrogen has a long history. The hydrogen atom and molecule were the focus of the early developments of quantum mechanics, and hydrogen was the first many-body system for which the consequences of the Fermi statistics of the electrons and the Pauli principle were explored by Fowler.18 Pressure ionization of atoms was already predicted by Hund.19 At about the same time, Wigner and Huntington predicted that atomic hydrogen under high pressure would be metallic.20 Since then the questions about the liquid and solid phases of hydrogen, as well as of transitions between insulating, conducting and, possibly, a superconducting phase predicted by Ashcroft,21 have been a driver of both experimental and theoretical progress. There have been many reports of the experimental observation of hydrogen metallization at room temperature (see Refs. 22–24), but there is no consensus that metallization has been reproducibly observed.

On the other hand, it is well accepted that the reverberating shock experiments by Weir et al.6 achieved the metallization of hydrogen at high temperature and pressure. The conditions are broadly consistent with those in the interior of Jupiter. There is indeed strong evidence that a thick layer of metallic hydrogen exists in the planet's interior because Jupiter has a strong magnetic field, even though the precise location of the dynamo active layer is still being debated.25 

The equation of state of hydrogen (and helium) is of crucial importance26,27 when measurements of Jupiter's and Saturn's gravity fields by orbiting spacecraft like Juno and Cassini are interpreted.28,29 Interior models typically start from the isentropic pressure–density relationship for hydrogen (see Fig. 1) before helium and heavier elements are introduced. The gravity measurements have shown that Jupiter's interior cannot be perfectly homogeneous and almost all models have introduced a step in composition into the planet's interior, which requires a physical justification. Earlier models thus invoked a first-order plasma phase transition (PPT) in hydrogen30 until first-principles computer simulations could not confirm its existence and instead predicted a liquid–liquid phase transition (LLPT).31–33 However, the latter is believed to occur at too low temperatures to matter for Jupiter's interior. Modern Jupiter models now invoke the phase separation of hydrogen–helium mixtures34–36 or a dilute core37,38 to justify an inhomogeneous interior structure. Still, a major challenge in understanding Jupiter's interior is remaining. Measurements of the planet's atmosphere have revealed a heavy element abundance of 4%. If one assumes these measurements to represent the bulk of the planet and if one invokes common EOSs for hydrogen and helium,39–41 then gravity measurements of the Juno spacecraft cannot be reproduced.42–44 Various hypotheses have been proposed to reconcile this discrepancy, which include lowering the density of hydrogen,45 increasing the planet's interior temperature,46 or reducing the deep abundance of the heavy elements.47 In addition to the equation of state, further thermodynamic (e.g., compressibility, Grüneisen parameter) and transport properties (e.g., electrical and thermal conductivity, viscosity) are required to model the thermal evolution (cooling) and dynamo processes (magnetic field generation) in giant planets like Jupiter and Saturn; for recent work (see Refs. 48 and 49).

FIG. 1.

Regions of temperatures and pressures accessed by experiments. Solid blue lines: static Diamond anvil cell (DAC), dashed blue lines: dynamic DAC. The solid red line is the Hugoniot starting from cold solid hydrogen at ambient pressure. Dashed lines are the conditions at the interiors of Jupiter, Saturn, and the Sun.

FIG. 1.

Regions of temperatures and pressures accessed by experiments. Solid blue lines: static Diamond anvil cell (DAC), dashed blue lines: dynamic DAC. The solid red line is the Hugoniot starting from cold solid hydrogen at ambient pressure. Dashed lines are the conditions at the interiors of Jupiter, Saturn, and the Sun.

Close modal

The relevance to planetary science makes it important to characterize how hydrogen transitions from a molecular, insulating state to an atomic and conducting state. As noted above, an early conjecture was the plasma phase transition (PPT), predicted by Schramm50 and later by Norman and Starostin.51 Even though no longer favored today, this issue has led to considerable development in theory and simulations. The first models were based on the thermodynamics of chemically reacting gases applied to many-particle systems of electrons, protons, atoms, and molecules (“chemical models,” e.g., Ref. 51). They were systematically extended to take into account interactions between all constituents (“nonideality effects”) by Ebeling and co-workers,52,53 Mihalas et al.,54–56 Chabrier and coworkers,39,57,58 and the Rostock school around Kremp et al.59,60 and Röpke et al.61–63 This has led to important advances in the many-body theory of Coulomb systems in the framework of kinetic theory and the theory of fluctuations,64 statistical physics,65–67 nonequilibrium Green functions,68,69 density operator theory,70 and linear response theory.71,72

Chemical models rely on effective interaction potentials between the considered “species.” Their density and temperature dependence are usually treated within simple approximations such as static screening and perturbation theory. With the emergence of modern computers, many-body simulations such as quantum Monte Carlo73 (QMC) and density functional theory74 (DFT) became feasible. These methods are based on the fundamental properties of electrons and nuclei and work in the “physical picture.” They avoid any artificial classification into different chemical species. QMC simulations of hydrogen and helium, based on Feynman's imaginary-time path integral representation of quantum mechanics (PIMC) were pioneered by Fosdick and Jordan75,76 and Filinov et al.77,78 However, when applied to fermions, these simulations, even though being free of systematic errors, are severely hampered by the fermion sign problem79–81 (FSP) confining them to regions of low electron degeneracy. To address the FSP, Ceperley introduced the restricted PIMC (RPIMC) method,82 which is completely sign-problem free, as long as diagonal density matrix elements are evaluated, which is sufficient to compute the internal energy, pressure, pair correlation functions, and structure factor. However, some negative contributions need to be included when the momentum distribution is computed.83 While being formally exact, RPIMC requires information about the nodal structure of the thermal density matrix, which has to be approximated in practice. This has allowed Ceperley and coworkers to perform simulations for temperatures where hydrogen is dominated by atoms and molecules.84–88 

The second approach within the “physical picture” is Kohn–Sham (KS) DFT coupled to molecular dynamics for the ions (DFT-MD or Born–Oppenheimer MD, BOMD, also known as ab initio MD or AIMD) which have dramatically extended the scope of problems that can be studied while yielding a higher accuracy compared to simpler models that require the introduction of assumptions specific to hydrogen. While DFT-MD and RPIMC methods rely on some approximations, these do not depend on a particular material. Early DFT-MD simulations of hot, dense hydrogen were performed by Lenosky,86 Scandolo,31 Desjarlais,89 and Bonev et al.90 

The “first principles” character (formally exact quantum mechanical approach within the physical picture) of RPIMC and DFT simulations91 rests on two fundamental approximations: the choice of the nodal surface of the N-particle density matrix and the exchange–correlation (XC) functional, respectively. The exact versions of these quantities are, in general, not known and the accuracy of a particular choice is difficult to assess from within the methods, and it will depend on the phase of hydrogen. Of course, the ultimate accuracy arbitrator is given by experimental measurements; however, in the case of highly compressed matter, such comparisons are difficult and afflicted by large uncertainties and error bars. An instructive example provides the laser-driven shock wave measurements of liquid deuterium of Da Silva et al.9,92 They used the Nova laser and reported a very high compression ratio, ρ/ρ0, of up to 5.9 [cf. gray circles in Fig. 2(a)] significantly larger than predicted by the SESAME model.93 This discrepancy sparked an intense discussion and led to a series of additional experiments and many theoretical investigations. Already the first two articles that reported results from first-principles simulations with RPIMC94 and DFT-MD86 consistently predicted a much smaller compression ratio of  4.3 at 200 GPa pressure (right panel of Fig. 2). It was argued that the uncertainty in the simulation results was smaller than an energy increase in 3eV/atom that would be required to bring the simulation results in agreement with the Nova measurements. That conclusion was supported by many following articles that reported simulation results. Using chemical models, on the other hand, it is possible to find agreement with the Nova measurements. Examples are the linear mixing model of Ross95 and the equation of state of Saumon et al.,39 as shown in Fig. 2. Note that both models predated the experiments and were not fit to the results.

FIG. 2.

Pressure vs compression along the deuterium Hugoniot. Panel (a) Experimental results up to 250 GPa: NOVA laser (gray filled circles);92 Z-machine (red diamonds);96 Ω laser shocks (green and blue squares);97,98 explosions (cyan triangles);99 Sano laser shock (orange circles);100 Fernandez laser shock;101 Nellis gas gun;102 and Dick gas gun.103 Panel (b) Theory results up to 250 GPa: SESAME (solid black line);93 DFTMD (solid orange);104–106 DFT-MD including finite temperature XC functionals (dashed orange line);107 RPIMC (brown circles, solid line);94 CEIMC (red circles);108,109 tight binding MD (gray dashed line);110 linear mixing model (green dash-dotted line),95 S&C (Saumon, Chabrier, and van Horn, blue dashed);39 and WPMD (Wavepacket MD, blue dots).111 Panel (c) High pressure regime up to 600 GPa: curves as introduced in panels (a) and (b) (note the different scales). Notice that experiments do not measure temperature directly which gives rise to additional uncertainties when compared to simulations.

FIG. 2.

Pressure vs compression along the deuterium Hugoniot. Panel (a) Experimental results up to 250 GPa: NOVA laser (gray filled circles);92 Z-machine (red diamonds);96 Ω laser shocks (green and blue squares);97,98 explosions (cyan triangles);99 Sano laser shock (orange circles);100 Fernandez laser shock;101 Nellis gas gun;102 and Dick gas gun.103 Panel (b) Theory results up to 250 GPa: SESAME (solid black line);93 DFTMD (solid orange);104–106 DFT-MD including finite temperature XC functionals (dashed orange line);107 RPIMC (brown circles, solid line);94 CEIMC (red circles);108,109 tight binding MD (gray dashed line);110 linear mixing model (green dash-dotted line),95 S&C (Saumon, Chabrier, and van Horn, blue dashed);39 and WPMD (Wavepacket MD, blue dots).111 Panel (c) High pressure regime up to 600 GPa: curves as introduced in panels (a) and (b) (note the different scales). Notice that experiments do not measure temperature directly which gives rise to additional uncertainties when compared to simulations.

Close modal

This discrepancy and the resulting uncertainty in the EOS led to a number of new measurements, such as the Z-pinch experiments by Knudson et al.112 and shock experiments at the Omega laser97,98 [see Fig. 2(a)]. The difference in maximum compression (4- to 4.5-fold112 vs 5- to 6-fold9,92,100) is striking. Interestingly, the theoretical results may also be sorted into two groups: high compression, close to a maximum of 6 has been predicted by chemical models, such as Ebeling's Padé formulas,53 in addition to the two models mentioned above. On the other hand, there are the first-principle RPIMC and DFT-MD simulations, and also the SESAME tables,93 that predicted significantly lower compressibilities. It is one aim of the present article to resolve these discrepancies. Testing the accuracy of the different simulations for dense partially ionized hydrogen allows us to confirm and quantify the reliability of RPIMC and DFT-MD.

Still, more work will be needed to resolve the remaining disagreements between various experimental and theoretical predictions for the Hugoniot state in Fig. 2. For example, measurements at the Z-machine indicate a lower compression ratio at 50 GPa than is predicted by coupled electron–ion Monte Carlo (CEIMC) simulations which, on the other hand, are in good agreement with the laser measurements by Sano et al.100 Compared to the Nova results, the more recent laser experiments at the Ω facility have yielded much lower compression ratios, which brought them into better agreement with predictions from the Z-facility and from first principles simulations.

Finally, we note that a comparatively small discrepancy has remained between first principles simulations and the recent laser shock experiments by Fernandez-Panella et al.101 on cryogenic liquid deuterium at very high pressures up to 550GPa [see Fig. 2(c)]. Rygg et al.113 suggested that collective modes (plasmons) would be missing from existing first principles calculations and might be the reason for the discrepancy. They performed a Debye-type calculation of the specific heat of electron plasma waves113 and simply added their contributions to the RPIMC results of Hu et al.114 This brought the theoretical results within the one-sigma error bars of the measurements above 300 GPa but could not explain the deviations at lower pressures. The central question is whether such plasmon waves are excited under such conditions and whether they would introduce a sufficiently large correction to the computed PIMC energies. Note that PIMC simulations take into account the entire Hamiltonian, so that plasmons are automatically accounted for, as long as the simulation cell is sufficiently large. By comparing results featuring 32 and 64 atoms, Militzer et al.94 determined that the compression ratio shifts by less than 0.01 at 580GPa and renders it unlikely that finite size effects or plasmons can explain the deviation of 0.1 between the measurements by Fernandez-Panella et al. and the RPIMC predictions. In addition, in this paper, we present novel independent tests of the RPIMC data of Ref. 94. A comparison with the first-principle fermionic PIMC (FPIMC) results of Filinov and Bonitz115 shows that the deviations in energy and pressure between RPIMC and FPIMC, for the parameters of interest, are less than 1% (cf. Sec. IV A 3), which rules out the plasmon hypothesis. Note that the FPIMC data have undergone a finite size extrapolation, an issue that will be discussed in Sec. III A.

As an alternative to laboratory measurements, new insights into the quality of theoretical models might be obtained with novel simulation methods that involve fewer or no uncontrolled approximations, such as exact diagonalization for model systems. Another tool for benchmarks that was already mentioned is fermionic PIMC simulations (featuring the FSP) under conditions for which they are feasible. Indeed, progress has been made recently with first-principles PIMC simulations for jellium (the uniform electron gas model,116,117 UEG) by Bonitz and coworkers who were able to avoid the FSP by a combination of two complementary approaches—configuration PIMC (PIMC simulations in occupation number representation, CPIMC)118–121 and permutation blocking PIMC (an advanced coordinate space approach, PB-PIMC).122–124 Simulations turned out to be feasible for all densities and temperatures exceeding half of the Fermi energy.125–128 Both were employed to benchmark RPIMC simulations121,129 and popular chemical models,130 including the Padé formulas of Ebeling,53 the STLS parameterization of Ichimaru et al.,131–133 and the models of Vashishta and Singwi134 and Perrot and Dharma-wardana.135 Another result of these simulations is the highly accurate parameterization of the exchange–correlation free energy—the functionals of Groth et al. (GDSMFB)117,128 and of Karasiev et al. (KSDT)136—which constitute explicitly thermal exchange–correlation functionals for DFT simulations on the level of the local density approximation.137,138 This demonstrates the high potential of parameter-free computer experiments for theory and model development in the field of warm dense matter.

Similar benefits can be expected from recent first-principles fermionic PIMC simulations—further improvements of the PB-PIMC approach—for partially ionized dense hydrogen by Filinov and Bonitz115,139,140 that cover temperatures above 15 000 K and densities corresponding to values of the Wigner–Seitz (Brueckner) coupling parameter rs3 (see Sec. II A). Reference 115 presented tables of benchmark data and comparisons with several alternative simulations. Another step forward is fermionic PIMC simulations for the static density response of hydrogen for fixed ionic configurations by Böhme et al.,141,142 which extend previous first principles results for jellium.143–147 Finally, we mention very recent works by Dornheim et al.148,149 who have presented the first dependable PIMC results for a number of structural, imaginary-time, and density-response properties for hydrogen at the electronic Fermi temperature considering the cases of rs=1, rs=2, and rs=3.23.

When considering hydrogen at temperatures well below the electronic degeneracy temperature, TF=EF/kB, fermionic PIMC (FPIMC) becomes highly inefficient because of the FSP. RPIMC simulations have been applied to temperatures as low as 0.1TF, but nodal restriction makes sampling new paths and moving the nuclei inefficient below such temperatures (see below). On the other hand, electronic temperature effects become less relevant here, and one can resort to a finite-temperature treatment of the nuclei, combined with a ground state description of the electrons. This is the realm of coupled electron ion Monte Carlo (CEIMC). Usually, the electronic problem is solved by DFT methods (BOMD), but for hydrogen, even more controlled electronic ground state QMC methods have been applied: the CEIMC method uses electronic energies from ground state QMC to sample nuclear degrees of freedom using Monte Carlo (MC),150,151 while a similar method using QMC-derived forces in a Langevin dynamics has also been proposed.152 Despite being heavy in terms of computer resources, those methods are more controlled than DFT since their accuracy can be judged based on the variational principle.

The first goal of the present paper is to summarize the current knowledge of the phase diagram of hydrogen, in particular at low temperatures. Here, many new results have been obtained recently but many details still remain open. We update the review of Ref. 156. The analysis is also extended to higher temperatures of partially ionized hydrogen over a wide range of densities.

The second goal of this paper is to give an overview of the diverse arsenal of simulation methods that can be applied to dense hydrogen and range from first-principles methods, such as QMC, over Green functions and DFT, to simpler models, such as chemical models and hydrodynamic equations. We discuss which quantities are accessible by the different methods and compare them with respect to their accuracy and applicability range. To this end, we perform comparisons for the equation of state, pair distributions, and degree of ionization between recent first-principles QMC results and DFT simulations, semiclassical MD with quantum potentials, average atom codes, and chemical models.

The third goal of this article is to present an overview of recent developments in simulations of dense hydrogen. These include the influence of finite-temperature XC-functionals in DFT simulations for hydrogen107,136,138,157 and the development of wave packet molecular dynamics simulations. Today, ab initio approaches can be used to analyze x-ray Thomson scattering (XRTS) signals from warm dense matter experiments.158 It is, for instance, possible to infer the experimental plasma parameters by comparing DFT-MD data for the dynamic structure factor to the XRTS scattering signal.159 A novel first principles approach to accurately analyze XRTS experiments was developed by Dornheim et al.17,148,160–164 Thus, it has become possible to compare not just integrated quantities like the EOS but also dynamic (frequency dependent) quantities from first principles simulations to experimental measurements. PIMC results exist for jellium for the dynamic structure factor,165,166 the dynamic local field correction,167 the plasmon spectrum,168 the dielectric function, and the conductivity.167 These can be extended in a semi-quantitative way to hydrogen and predicted a similar roton-type feature in hydrogen as in strongly correlated jellium that might be observable.166,169

Given the high interest in ICF, we also reevaluate the open issues for the accurate theoretical modeling of the compression path of the deuterium–tritium fuel. The density temperature trajectory of the fuel and ablator materials traverses the warm dense matter into the high dense matter regimes as the ICF capsule implodes and then undergoes thermonuclear ignition, burns its fuel, and ultimately decompresses (cf. Fig. 3). Pressures range from around a Megabar, at a density of 1024 cm−3 and temperatures of 1 eV, to tens of Gigabars of pressure, during ignition and burn, where densities are on the order of 1026 cm−3 and temperatures are many kiloelectron volt. It is expected that the novel results obtained from first principles simulations presented in this article will also be valuable for ICF modeling.

FIG. 3.

Qualitative location of warm dense matter (WDM) in the plane of electron number density and temperature. We also introduce a number of relevant parameters that are defined in Sec. II A. In addition, we include examples of warm dense hydrogen in selected astrophysical objects such as giant planet interiors and ICF compression path from Refs. 114, 153, and 154. Adapted from Ref. 155.

FIG. 3.

Qualitative location of warm dense matter (WDM) in the plane of electron number density and temperature. We also introduce a number of relevant parameters that are defined in Sec. II A. In addition, we include examples of warm dense hydrogen in selected astrophysical objects such as giant planet interiors and ICF compression path from Refs. 114, 153, and 154. Adapted from Ref. 155.

Close modal

The structure of this article is as follows: In Sec. II, we summarize the current knowledge as well as open questions on the phase diagram of hydrogen, including a discussion on metallization, the plasma phase transition, and the liquid–liquid phase transition(s). In Sec. III, we present an overview of important simulation approaches for dense partially ionized hydrogen including the regimes of low and high temperatures. In Sec. IV, we present our numerical results and comparisons of methods, including thermodynamic properties of hydrogen, density response functions, and transport properties. We also make suggestions for the future developments of the different methods and conclude with a summary and outlook in Sec. V where we also outline strategies for accurately combining different simulation approaches.

In this section, we present an overview on the most important properties and phases of hydrogen at high pressures. This includes the hypothetical plasma phase transition (PPT) (Sec. II B 1), the insulator–metal (liquid–liquid) transition (Sec. II B 2), and the proton crystal phase (Sec. II B 3). First, we introduce the basic parameters that are being used to describe highly compressed hydrogen. Since we discuss a large variety of properties and quantities as well as simulation methods and approximations, we provide a list of common acronyms in  Appendix in Table IX.

TABLE I.

Results for the hypothetical plasma phase transition (top rows: chemical models, CM; middle rows: PIMC simulations; bottom rows: experiments, Exp.), including critical temperature, pressure, degree of ionization or density (ρ is given in g cm−3; n is given in cm−3). FPIMC: fermionic PIMC; RPIMC: restricted PIMC. f: free-particle nodes; v: variational nodes; 1: explosives-driven shock wave; 2: laser-driven shock wave. Predicted transitions in the liquid phase (LLPT) are listed separately, in Table II.

Reference, year Tcr(K) pcr(GPa) α,rs,n,ρ Method
50, 1961  15 920      CM 
51, 1968  0.1Tion    ncr=(0.1aB)3  CMa 
179, 1970   10 913    ncr=(0.096aB)3  CM 
52, 1973  12 600  95.0  ρcr=0.95  CM 
180, 1980  <9000    ρcr=1.0  CM 
181, 1983  19 000  24.0  αcr=0.50  CM 
53, 1985  16 500  22.8  αcr=0.32  CM 
182, 1989  15 000  64.6  αcr=0.2  CM 
58, 1992  15 300  61.4  αcr=0.18  CM 
39, 1995  15 311  61.4  αcr=0.075  CM 
59, 1995  14 900  72.3  αcr=0.4  CM 
61, 1995  15 000  23  ρcr=0.13  CM 
183, 1996  13 000    ncr=aB3  CM 
184, 1999  >10 000  100    CM 
77, 1975  >100 000      FPIMC 
85, 1996  11 000  48  rscr=2.2  RPIMCf 
87, 2000  ⋯  ⋯  no PPT  RPIMCv 
185, 2001  >10 000    ncr>1022  FPIMC 
186, 2003  >10 000    ncr>1022  FPIMC 
7, 1972    <280    Exp.1 
6, 1996  ⋯  ⋯  no PPT  Exp.2 
8, 2007        Exp.1 
Reference, year Tcr(K) pcr(GPa) α,rs,n,ρ Method
50, 1961  15 920      CM 
51, 1968  0.1Tion    ncr=(0.1aB)3  CMa 
179, 1970   10 913    ncr=(0.096aB)3  CM 
52, 1973  12 600  95.0  ρcr=0.95  CM 
180, 1980  <9000    ρcr=1.0  CM 
181, 1983  19 000  24.0  αcr=0.50  CM 
53, 1985  16 500  22.8  αcr=0.32  CM 
182, 1989  15 000  64.6  αcr=0.2  CM 
58, 1992  15 300  61.4  αcr=0.18  CM 
39, 1995  15 311  61.4  αcr=0.075  CM 
59, 1995  14 900  72.3  αcr=0.4  CM 
61, 1995  15 000  23  ρcr=0.13  CM 
183, 1996  13 000    ncr=aB3  CM 
184, 1999  >10 000  100    CM 
77, 1975  >100 000      FPIMC 
85, 1996  11 000  48  rscr=2.2  RPIMCf 
87, 2000  ⋯  ⋯  no PPT  RPIMCv 
185, 2001  >10 000    ncr>1022  FPIMC 
186, 2003  >10 000    ncr>1022  FPIMC 
7, 1972    <280    Exp.1 
6, 1996  ⋯  ⋯  no PPT  Exp.2 
8, 2007        Exp.1 
a

Data are from a preprint of Alekseev et al., Ref. 5 in Ref. 51.

Despite its chemical simplicity, warm dense hydrogen exists in a broad variety of phases, ranging from solid to liquid, gas, and plasma (for an overview, see Figs. 3 and 4). For systematics, it is useful to introduce dimensionless parameters. In the following, we will use CGS units, i.e., set 4πε0=1.

FIG. 4.

Hydrogen phase diagram. Solid black lines show the boundaries between the gas, liquid, and solid phases as measured in static experiments. The solid circles show the location of critical or triple points (black: observed, red: predicted). The black dashed lines are crossovers between the classical behavior of electrons and protons at high temperatures. TF(e) and TF(p+) are the Fermi temperatures of electrons and protons, respectively. “PPT” indicates the critical region of the hypothetical plasma phase transition (Sec. II B 1). “LPPT” indicates the liquid–liquid phase transition. Thermal ionization occurs at TTion=157000 K [Eq. (21)]. The red dotted line shows the estimated extent of solid atomic hydrogen.

FIG. 4.

Hydrogen phase diagram. Solid black lines show the boundaries between the gas, liquid, and solid phases as measured in static experiments. The solid circles show the location of critical or triple points (black: observed, red: predicted). The black dashed lines are crossovers between the classical behavior of electrons and protons at high temperatures. TF(e) and TF(p+) are the Fermi temperatures of electrons and protons, respectively. “PPT” indicates the critical region of the hypothetical plasma phase transition (Sec. II B 1). “LPPT” indicates the liquid–liquid phase transition. Thermal ionization occurs at TTion=157000 K [Eq. (21)]. The red dotted line shows the estimated extent of solid atomic hydrogen.

Close modal

1. Parameters of electrons and protons

The properties of the electrons and protons in thermal equilibrium are characterized by a combination of standard jellium parameters and Coulomb one-component plasma (OCP) parameters:170 

  • The Wigner–Seitz (Brueckner) coupling parameter of electrons
    (1)
    is the ratio of the mean interparticle distance (or Wigner–Seitz radius), d, given by
    (2)

    to the Bohr radius [Eq. (18)], where n denotes the number density.

  • The degeneracy parameter
    (3)
    is the ratio of thermal energy to the Fermi energy, EF=(qF)2/2m, where the Fermi wave number is qF=(3π2n)1/3. The parameter Θ is closely related to the degeneracy parameter χa, given in terms of the thermal de Broglie wavelength Λ of electrons and protons as
    (4)
    (5)

    This expression for Λ holds in thermal equilibrium for ideal particles, its modification due to interaction effects is discussed in Sec. III G 2.

  • The generalized coupling parameter
    (6)

    which takes into account the quantum kinetic energy via the Fermi-integral I3/2 and the spin degeneracy factor 2σ+1.171 It is shown as the red line in Fig. 3. For non-degenerate particles in the gas (plasma) phase, i.e., for χa1, this simplifies to the well-known classical Coulomb coupling parameter, Γ=e2/(dkBT). For high degeneracy, Γ becomes constant when the zero-temperature quantum coupling parameter, rs, is constant.

  • Screening and dynamic properties of the free electrons are characterized by the Debye length, rD and the plasma frequency, given as
    (7)
    (8)
    (9)
For strong degeneracy, the Debye length is replaced by the Thomas–Fermi screening length, rTF.

2. Parameters of partially ionized hydrogen

In the case of partial ionization, it is common to characterize the gas and plasma phases by the partial densities of free and bound electrons which add up to the total density, ne. Furthermore, the bound state density, in the case of hydrogen, consists of an atomic and a molecular component, nA and nM, respectively, which count the number of electrons (protons) per volume in an atom or a molecule,
(10)
(11)
  • The thermodynamic properties of partially ionized hydrogen are then characterized by the following dimensionless parameters: the finite fractions of free particles (degree of ionization), α, atoms, xA, and molecules, xM.
    (12)
    (13)
    (14)

    and α+xA+2xM=1. Such a subdivision into “free” and “bound” electrons is called “chemical picture” (cf. Sec. III B) and is somewhat artificial. The results may depend on the procedure; different criteria lead to different results. This issue will be discussed in Sec. IV A 6.

  • Often, instead of the particle density, the mass density is used, which we give for hydrogen (ρ) and deuterium (ρD) as
    (15)
    (16)

    where ρ is given in g/cm3.

  • The fractions of bound states are influenced by temperature and density in units of the binding energy,
    (17)
    (18)
    (19)
    (20)
    (21)

    where mr is the reduced mass, mr=meMM+1, with the mass ratio M=mp/me, and for hydrogen εr=1. In Eqs. (19) and (20), we indicated the values of the binding energy of hydrogen molecules and the associated equilibrium proton distance. Complete thermal ionization of atoms occurs for temperatures TTion [Eq. (21)], corresponding to the binding energy.

  • The free electron properties in a partially ionized hydrogen plasma can be estimated from the jellium parameters, by a simple rescaling of the density, nene*,169 the corresponding parameters will be denoted by an asterisk:
    (22)
    (23)
    (24)
    (25)
    (26)

    For a given total density n and temperature T, these quantities depend on the degree of ionization, α(n,T) [Eq. (12)]. This estimate of electronic properties applies the “chemical picture” (cf. Sec. III B) and assumes that the influence of the bound states on the free electron properties is negligible.

3. Pressure ionization. Mott effect. Insulator–metal transition (IMT)

In a many-particle system, ionization can occur also at zero temperature. In the case of a sustained overlap of neighboring atoms (molecules), Coulomb correlations, quantum, and exchange effects lead to a shift of the individual bound state energies and a lowering of the ionization (dissociation) energy which eventually vanishes at a critical density, which is sometimes called “Mott density” in the plasma physics literature.68,69 PIMC simulations predicted a critical coupling parameter
(27)
for pressure ionization of hydrogen atoms at low temperature,172 which agrees with the value of 1.19 reported by Rogers from the solution of the hydrogen Schrödinger equation with a Debye potential.173 At finite temperature, pressure ionization occurs at reduced densities. The topic of pressure ionization and ionization potential depression will be discussed in more detail in Sec. IV B. Note that the above picture of pressure ionization is essentially a single-atom picture, where an isolated atom is embedded into a surrounding medium which is often appropriate in the high-temperature plasma phase.

At low temperatures, in the condensed phase, the atoms or molecules of hydrogen form an ordered lattice. Then pressure ionization and dissociation is a many-particle effect. Compression of the crystal leads to a change of the band structure and, eventually, to a closure of the energy gap between the highest occupied and the lowest unoccupied band: an insulator to metal transition (IMT). This effect was first studied by Wigner and Huntington,20 predicting that molecular hydrogen (an insulator) would transition to metallic atomic hydrogen because of band filling. Mott174,175 further considered the effects of electron correlation and computed critical values for the electron density beyond which metallization occurs. These values strongly depend on the material and the crystal symmetry, and the corresponding values for the coupling parameter are significantly higher than the hydrogen plasma result (27), for a recent overview (see Ref. 176). The IMT in hydrogen will be discussed in detail in Sec. II B 2.

Hydrogen has a very rich phase diagram, many details of which are under active investigation. Figure 4 shows a simplified overview. At high temperatures and low pressures (top left corner) hydrogen behaves as a classical gas of charged particles—a two-component electron–ion plasma (TCP). Upon cooling below the binding energy, TTion, hydrogen atoms and molecules form, the fraction of which increases as the temperature decreases.

At higher pressures, cooling potentially gives rise to a second fluid–fluid phase transition in partially ionized hydrogen. There have been extensive discussions in the literature about whether there are two independent phase transitions. Landau and Zeldovich,177 Schramm,50 and Norman and Starostin51 predicted that ionization and recombination of the plasma proceeds via a “plasma phase transition” (PPT) (see Fig. 4). We discuss theoretical and experimental predictions in Sec. II B 1 and conclude that there is no evidence for the PPT. A second liquid–liquid phase transition (LLPT) at megabar pressures and temperatures around 1000K likely occurs between molecular hydrogen and metallic fluid hydrogen.6 

Lowering the temperature further, hydrogen freezes. At these lower temperatures, nuclear quantum and isotopic effects become important. We discuss the location of the melting line in Sec. II B 2. There, we also consider phenomena predicted to occur upon compression, in particular the insulator-to-metal transition in solid hydrogen.20 Whether this transition coincides with the transition from molecular hydrogen to atomic hydrogen is still an open question. Ashcroft et al.21,178 have suggested that the solid atomic hydrogen phase will be a room-temperature superconductor because of the large phonon energies and electron–proton coupling. Finally, in Sec. II B 3, we concentrate on the compression of hydrogen gas up to TPa and PPa at relatively high temperatures (see Fig. 4). These pressures give rise to the formation of an atomic (proton) crystal that is bounded by a classical proton fluid at lower pressures, and a quantum proton liquid at higher pressures.

1. Partially ionized hydrogen below the Mott density: Hypothetical plasma phase transition (PPT)

We start by considering partially ionized hydrogen at temperatures 10 000KTTion [Eq. (21)] and densities below the Mott density, rsrsMott [Eq. (27)]. The degree of ionization α increases monotonically with temperature (thermal ionization) and density (pressure ionization) which is accompanied by a strong increase in electrical conductivity. This parameter range attracted particular attention in the 1970s to 1990s, due to the prediction of the so-called plasma phase transition (PPT)—a hypothetical first-order phase transition that was thought to be analogous to the van der Waals gas–liquid transition. The idea was50,51 that the combination of overall attractive Coulomb forces in the plasma and short-range electron repulsion (due to quantum effects), below a critical temperature may lead to a fluid with two phases of different degree of ionization. Additional arguments for the existence of the PPT were similar effects in electrolytes, alkali metals, and electron–hole plasmas in semiconductors (e.g., Ref. 68). The critical point of the PPT was closely linked to the ionization potential of the hydrogen atom, and most predictions settled around Tcr0.1Tion 15 800 K; for a representative collection of results (see Table I). The majority of predictions of the PPT originated from chemical models (CM, the top part of Table I) (for details, cf. Sec. III B).

The PPT was also investigated with first principles fermionic PIMC simulations.77,185,186 However, these simulations are severely hampered by the sign problem (cf. Sec. III E 2) and could not achieve converged results.187 On the other hand, restricted PIMC simulations with free-particle nodes reported a PPT85 but were not confirmed by more accurate later simulations with variational nodes87 (see also Refs. 188 and 189, cf. bottom part of Table I).

Finally, on the experimental side (bottom of Table I) hydrogen metallization and a density jump at high pressure were reported in shock compression experiments using explosives.7 On the other hand, Weir et al.6 observed metallization of hydrogen and deuterium in shock experiments, around p=140GPa at temperatures where hydrogen is a fluid, but did not observe a phase transition. Fortov et al. also used explosives-driven shock compression of deuterium8 and claimed proof of the PPT, based on observation of a “density jump”' in the range of p = 127–150 GPa. However, the evidence is not convincing,190 and the studied temperature range of 2550–4000 K (as deduced from models) is confined to the fluid phase. Thus, the observations could rather be an indication of the liquid–liquid transition (see Sec. II B 2).

The topic of the PPT has been frequently discussed in reviews and textbooks (see, e.g., Refs. 156, 191, and 192). Also, Refs. 193 and 194 contain a recent overview of the history of the PPT, from the point of view of their Russian authors. Table I summarizes the extensive literature on the PPT. Note that chemical models become questionable if the ionization fraction changes significantly with density. This is the case, in particular, if the binding energies vanish near the Mott density, where atoms (or molecules) transform to short-lived transient states, as typical for WDM. The PPT may be an artifact that is “built into” these models.195,196 For more details, see Sec. III B. State-of-the-art first-principles simulations in the physical picture show no indications of a PPT. DFT-MD simulations of Lorenzen et al.33 and coupled electron–ion (CEIMC) simulations of Morales et al.32,197 have produced evidence of a phase transition at much lower temperatures in the fluid phase (LLPT), which differs qualitatively from the PPT (for more details, see Table II and Sec. II B 2).

TABLE II.

Experimental (first two lines) and simulation results for the first order liquid–liquid phase transition (LLPT) including temperature range (in 1000 K) and pressure range. PBE: PBE XC-functional. Predictions of phase transitions in the plasma phase (PPT) are listed separately in Table I. For details, see text and Fig. 6.

Reference, year Tcr(K) pcr(GPa) Method
6, 1996  <5200  No LLPT  Shock experiment 
8, 2007      Shock experiment 
31, 2003  2000  125  Car-Parrinello 
188, 2004  <4500    DFT, PBE 
198, 2006  ⋯  No LLPT  CEIMC 
88, 2007  ⋯  No LLPT  DFT 
104, 2008  ⋯  No LLPT  DFT 
33, 2010  1400  132  DFT 
197, 2010–16  2000  120  CEIMC 
  2000  97  DFT 
Reference, year Tcr(K) pcr(GPa) Method
6, 1996  <5200  No LLPT  Shock experiment 
8, 2007      Shock experiment 
31, 2003  2000  125  Car-Parrinello 
188, 2004  <4500    DFT, PBE 
198, 2006  ⋯  No LLPT  CEIMC 
88, 2007  ⋯  No LLPT  DFT 
104, 2008  ⋯  No LLPT  DFT 
33, 2010  1400  132  DFT 
197, 2010–16  2000  120  CEIMC 
  2000  97  DFT 

2. Hydrogen phase diagram at low temperature: Metallization and liquid–liquid phase transition (LLPT)

The phase diagram of hydrogen under extreme conditions is still uncertain, mainly because of experimental difficulties in producing and controlling the extreme physical conditions required199 and the limited information experiments obtained about the system. At low temperatures in the solid phase, the most important missing information is the crystal structure as a function of temperature and pressure. In Fig. 5, we report the present low-temperature phase diagram emerging from experimental information. Up to six different crystal structures have been detected but the transition lines are traced based on discontinuities and changes, respectively, of the vibrational frequencies of the infrared (IR) and Raman spectra, while measurements of Bragg peaks are mostly missing. A notable exception is along the 300 K isotherm where special x-ray spectroscopy has been developed and applied to confirm that the structure of phase I is m-HCP up to 250 GPa.200,201 Candidate structures have been inferred by the ab initio Random Structure Search (AIRSS) method202 based on DFT calculations with phonon corrections.

FIG. 5.

Experimentally inferred solid hydrogen phase diagram (solid and dashed black lines). Several nonmetallic crystal phases have been detected. The melting line (black line) is reentrant and has been measured up to about 300 GPa.205 The low-temperature phase III is where metallization has been measured. The semimetallic state is entered at 360 GPa204 and indicated by the blue shaded area. It persists up to 425 GPa where a sudden collapse of the direct gap is detected,24 which is represented by the red shaded area. At 495 GPa, a reflective sample has been reported203 (represented by the dark red vertical bar in the lower right corner). Two theoretical melting lines are also reported, one obtained by free energy methods for classical nuclei within BOMD with DFT-PBE (orange line)32 and one recently obtained by the two-phase method for a system of quantum nuclei using a machine learning force field (DeePMD) trained on QMC energies and forces (green continuous line).206 Green circles: experimental data for warm hydrogen in DAC where signatures of a phase transition were detected.207 Blue circles: CEIMC predictions for the LLPT line for quantum protons.197 

FIG. 5.

Experimentally inferred solid hydrogen phase diagram (solid and dashed black lines). Several nonmetallic crystal phases have been detected. The melting line (black line) is reentrant and has been measured up to about 300 GPa.205 The low-temperature phase III is where metallization has been measured. The semimetallic state is entered at 360 GPa204 and indicated by the blue shaded area. It persists up to 425 GPa where a sudden collapse of the direct gap is detected,24 which is represented by the red shaded area. At 495 GPa, a reflective sample has been reported203 (represented by the dark red vertical bar in the lower right corner). Two theoretical melting lines are also reported, one obtained by free energy methods for classical nuclei within BOMD with DFT-PBE (orange line)32 and one recently obtained by the two-phase method for a system of quantum nuclei using a machine learning force field (DeePMD) trained on QMC energies and forces (green continuous line).206 Green circles: experimental data for warm hydrogen in DAC where signatures of a phase transition were detected.207 Blue circles: CEIMC predictions for the LLPT line for quantum protons.197 

Close modal

A relevant and still partially unanswered question concerns the mechanism by which solid hydrogen metallizes upon increasing pressure.199 

Important advances toward the metallization of solid hydrogen at low temperature were announced by different groups in recent years. In 2017, the group at Harvard reported the observation of metallic hydrogen in diamond anvil cell (DAC) experiments at 495 GPa below 80 K.203 This conclusion was based on the sudden appearance of a reflective sample that was interpreted as evidence for the metallization of hydrogen. This interpretation has been criticized by others,199 and so far the results have not been reproduced. Almost simultaneously the group in Mainz reported evidence of the formation of a semi-metallic, still molecular, phase at around 350 GPa and below 100 K. They employed both optical probes and direct electric measurements in a DAC.204 In 2020, Loubeyre et al.24 reported results using a toroidal DAC with synchrotron radiation. They measured the IR absorption profile over a wide range of pressure and detected complete absorption at 427 GPa and 80 K, which was interpreted as a sudden closure of the direct energy gap, a strong indicator for the state of a “good metal.” Although sample visual inspection and reversibility of the transition upon pressure release suggest that this metal is still molecular, strong experimental evidence, such as the persistence of the vibron signal, is missing. Thus, it is conceivable that the observed collapse of the infrared gap signals a metallization through molecular dissociation.

In Fig. 5, we sketch the current low-temperature phase diagram up to the IMT region. Below 200 K in phase III, we report the transition to the semimetallic state at 350 GPa204 as a blue shaded area and the later transition to the “metallic” state at 425 GPa24 as the red shaded area. At 495 GPa, we also report the results of the Harvard group203 as the dark red-filled area. No experimental information is available for metallization at temperatures higher than ∼100 K. Candidate structures for the low temperature phase III, emerging from AIRSS202,216 and a subsequent QMC analysis,217 are C2/c-24 and Cmca-12. Gap closure of those structures has been detected with QMC methods218 finding that the C2/c-24 structure undergoes an indirect gap closure around 360 GPa, in quantitative agreement with the experimental transition to the semimetallic phase. Moreover, it was shown that the direct gap of the C2/c-24 phase follows the experimental absorption edge behavior with pressures from Ref. 24. We emphasize that this agreement can only be obtained by taking into account the zero point motion of the protons.218 Moreover, modeling of the correlated electrons requires going beyond common DFT approximations and using QMC methods.218–220 A recent study,221,222 based on the Stochastic Self-Consistent Harmonic Approximation (SSCHA) and QMC-corrected energies suggests that the observed collapse of the direct gap at 425 GPa24 is related to a structural transition from the C2/c-24 to the Cmca-12 metallic structure, in both hydrogen and deuterium, as verified by experiments.223 

Above 200 K, phase III transforms into a new phase IV when the molecular vibron line splits into two characteristic frequencies suggesting that different molecules may experience two different environments.224 For this phase, theory predicted a layered structure in which regular molecular layers are intercalated with planar layers where molecules form almost regular hexagons, precursors of a transition to a hexagonal monatomic layer at higher pressure.216,225 Not much is known about metallization in this higher temperature phase (see, however, Refs. 225 and 226).

The black lines in Fig. 5 are experimental phase boundaries. However, experimental information about the crystalline structures of those phases are missing, and the experimental determination of the melting line is not based on the evidence of the vanishing of long-range order (Bragg peaks) above the melting line. We also report the melting line from two different theories. The orange line is from BOMD simulations with the DFT-PBE approximation.32,90 The green line is a recent prediction from a Machine Learning force field trained on QMC energies and forces (ML-QMC).206 Within DFT, the location of the melting line and LLPT line strongly depends on the approximation for the exchange–correlation functional. The DFT-PBE melting line seems to agree better with the experiments, although it was observed that, considering quantum nuclei within this theory, leads to unphysical results.197,227 At pressures between 50 and 200 GPa, ML-QMC predicts a different solid structure that is much more stable than in PBE-PIMD calculations. This prediction has yet to be confirmed by experiments.

Above the melting line, fluid hydrogen can be either molecular or atomic. First-principle simulations, both by Born–Oppenheimer molecular dynamics (BOMD) and by coupled electron–ion Monte Carlo (CEIMC),31–33,197,214,228,229 predict the existence of a weakly first order LLPT between the molecular insulating fluid and the mostly monatomic metallic fluid, below a critical temperature estimated to be TLLPTcr 2000 K. This is signaled by a small discontinuity in the specific volume at a given pressure and temperature. The precise value of TLLPTcr is not known but, above TLLPTcr, molecular dissociation and metallization increase progressively with pressure. In Fig. 6, we report CEIMC predictions (cf. blue and red shaded lines) for the LLPT line of hydrogen and deuterium between 600 and 1500 K.197 Other lines from BOMD are in qualitative agreement, but their precise location depends on the specific XC approximation adopted.197,230 Unequivocal experimental evidence of the occurrence of a first-order transition is presently missing.

FIG. 6.

Phase diagram of fluid hydrogen and deuterium around the LLPT line. Shaded lines (blue for hydrogen and red for deuterium) are the LLPT predicted by CEIMC.197,208 Filled symbols: estimates of the LLPT from the reflectivity coefficient; open symbols indicate the inception of absorption. Squares correspond to deuterium, and circles to hydrogen. Shown are data from sp-DAC (green), Z-machine (orange), NIF (red), and lp-DAC methods (purple). DAC-p, data from sp-DAC corresponding to the temperature plateau from Refs. 207 and 209 (T 1700 K) and from Ref. 210 (T 1700 K); DAC-r, data from sp-DAC at R = 0.3; lp-DAC 212, filled purple points are conducting conditions, and open purple points are nonconducting conditions (for both hydrogen and deuterium); NIF-a, data from NIF when the absorption coefficient; NIF-r, data from NIF at R = 0.3;212 Z-a, data from Z-machine when the sample becomes dark; Z-r, data from Z-machine at the observed discontinuity in reflectivity.213 Blue points: theoretical predictions from Ref. 214 for quantum protons: Filled circles show when R = 0.3 for H/vacuum interface; open circles show when μ=1μm1. Brown shaded circles: recent predictions for a system of quantum protons from PIMD simulation with the DFT-SCAN-rvv10 approximation.215 The points correspond to a conductivity σ=2000S/cm. Melting lines are as in Fig. 5. Figure adapted from Ref. 214.

FIG. 6.

Phase diagram of fluid hydrogen and deuterium around the LLPT line. Shaded lines (blue for hydrogen and red for deuterium) are the LLPT predicted by CEIMC.197,208 Filled symbols: estimates of the LLPT from the reflectivity coefficient; open symbols indicate the inception of absorption. Squares correspond to deuterium, and circles to hydrogen. Shown are data from sp-DAC (green), Z-machine (orange), NIF (red), and lp-DAC methods (purple). DAC-p, data from sp-DAC corresponding to the temperature plateau from Refs. 207 and 209 (T 1700 K) and from Ref. 210 (T 1700 K); DAC-r, data from sp-DAC at R = 0.3; lp-DAC 212, filled purple points are conducting conditions, and open purple points are nonconducting conditions (for both hydrogen and deuterium); NIF-a, data from NIF when the absorption coefficient; NIF-r, data from NIF at R = 0.3;212 Z-a, data from Z-machine when the sample becomes dark; Z-r, data from Z-machine at the observed discontinuity in reflectivity.213 Blue points: theoretical predictions from Ref. 214 for quantum protons: Filled circles show when R = 0.3 for H/vacuum interface; open circles show when μ=1μm1. Brown shaded circles: recent predictions for a system of quantum protons from PIMD simulation with the DFT-SCAN-rvv10 approximation.215 The points correspond to a conductivity σ=2000S/cm. Melting lines are as in Fig. 5. Figure adapted from Ref. 214.

Close modal

Experiments on fluid hydrogen are performed either by static compression, with DAC,207 or by dynamic compression, with shock wave techniques6,8,212,213 (cf. Fig. 1). During the pressure increase, the sample first becomes opaque when the electronic gap (cf. Sec. IV C 3) becomes comparable to the energy of the probe laser (absorption coefficient μ1μm1), and later, at higher pressure, it turns reflective, signaling the occurrence of the metallic state (reflectivity R=0.3).

In Fig. 6, we show static compression data at R=0.3 (DAC-r) from Refs. 207 and 231 (short-pulse DAC, sp-DAC) and Ref. 211 (long pulse DAC, lp-DAC) on hydrogen and deuterium, together with shock wave data on deuterium from the Sandia group213 and from the Livermore group.212 The inception of absorption (μ1μm1) in those experiments is represented by open symbols while closed symbols represent the occurrence of reflective samples (R = 0.3). Also, data from sp-DAC show their temperature plateau (DAC-p). Together with experimental data, Fig. 6 also reports theoretical predictions (obtained within the Kubo–Greenwood approach, cf. Sec. IV E 2) for the inception of an opaque sample and inception of a metallic sample, from Ref. 214. While we see an essential agreement between different experiments and with the theory, for the absorption threshold, reflectivity measurements are not in agreement between the two different shock wave experiments. Note that in those experiments, temperature is not directly measured but it is inferred from a model EOS, and predictions are sensitive to the adopted model. Theoretical predictions for the reflectivity threshold are closer to the NIF-r data and are also in agreement with sp-DAC-r measurements.

In Ref. 214, it was shown that R=0.3 corresponds roughly to a conductivity value of σ= 2000 S/cm, although the correspondence depends on the refractive index of the considered interface (see Ref. 214 for details). This value of the conductivity of metallic hydrogen was observed as a saturation value in early shock wave experiments.6 In Fig. 6, we also report theoretical predictions of the occurrence of metallic hydrogen (σ=2000S/cm) from PIMD with a modern XC approximation (SCAN+rVV10).215 While below the critical point, those predictions are in agreement with other estimates, at higher temperatures this behavior deviates toward the absorption threshold.

Before closing this section, we should note that the new ML-QMC melting line has a potentially large impact on interpreting and understanding this part of the phase diagram. For instance, the thermodynamic path followed during the compression in the shock wave experiments might be influenced by the occurrence of crystallization and subsequent re-melting of the sample. We observe that, except for the μ=1μm1 data, all other data show a “kink” as the temperature is lowered including that of the sp-DAC-p data. The relative location of the new melting line and the LLPT line from CEIMC suggests that this kink might be the signature of a reentrant melting transition—a speculation that needs confirmation. Also, we note that a higher melting line restricts the LLPT region to a much smaller portion of the phase diagram.

3. Hydrogen phase diagram in the high-density plasma phase: Proton crystal (atomic solid hydrogen)

Finally, we consider temperatures exceeding 10 000 K and densities exceeding the Mott density, rsrsMott, where the hydrogen plasma is fully ionized. With increasing density, the electrons eventually approach an ideal Fermi gas; however, the protonic order can change. Due to the large proton to electron mass ratio, the proton degeneracy is smaller by almost five orders of magnitude, χe/χp78700. Consequently, there exists a density range given by [cf. the definitions (4) and (5)]
(28)
corresponding to a temperature range TFPTTF, where the electrons are strongly degenerate but the protons are still classical (TFP=EFP/kB denotes the Fermi temperature of the protons) (cf. Fig. 4). There, the system is reasonably well described by a classical one-component plasma (OCP) model of ions embedded into a uniform neutralizing electron background. Depending on the density and temperature, protons will exhibit gas-like, liquid-like, or solid-like behavior. The range of the proton crystal (solid atomic hydrogen in a b.c.c structure) is indicated in Fig. 4. Screening effects at high density reduce the melting temperature by a factor of about 2.232 The liquid–solid phase boundary at high temperature is approximately given by the classical proton coupling parameter [Eq. (6)], Γ175, whereas the transition to a quantum proton liquid occurs toward high density, around rsp=rsM100, as a result of proton quantum effects. Finally, the gas–liquid transition is observed in the range Γ1020 (e.g., Ref. 233).
This behavior is not restricted to hydrogen but is observed also for other dense electron–ion systems. For a given temperature and density, the liquid–solid boundary differs for different plasmas, as it depends on the ion to electron mass ratio M. The results of two-component PIMC simulations are shown in Fig. 7. There we plot the relative interparticle distance fluctuations
(29)
i.e., the mean fluctuations of the distance of neighboring particles in units of the mean interparticle distance, d, which provide a very approximate criterion for a solid–liquid transition (modified Lindemann criterion) (e.g., Ref. 234): whereas in the solid phase, particles are localized around their lattice positions and distance fluctuations are below a threshold value on the order of 0.050.15, R increases rapidly in the liquid phase. The figure shows that this transition occurs (at fixed temperature and density) when M is reduced, around a value of M=80.
FIG. 7.

Mean square relative distance fluctuations (in units of the mean interparticle distance) of the heavy particles (ions, holes) [Eq. (29)] as a function of the mass ratio M, showing the role of heavy particle quantum effects. Two-component PIMC simulations for kBT=23EB and rs=0.63, with the binding energy EB [Eq. (17)]. Reproduced from Ref. 241 with the permission of the authors. Copyright IOP Publishing. Reproduced with permission. All rights reserved.

FIG. 7.

Mean square relative distance fluctuations (in units of the mean interparticle distance) of the heavy particles (ions, holes) [Eq. (29)] as a function of the mass ratio M, showing the role of heavy particle quantum effects. Two-component PIMC simulations for kBT=23EB and rs=0.63, with the binding energy EB [Eq. (17)]. Reproduced from Ref. 241 with the permission of the authors. Copyright IOP Publishing. Reproduced with permission. All rights reserved.

Close modal
In Ref. 172, also analytical estimates for the existence of an ion crystal were provided. The density range, [n(1),n(2)], the maximum temperature T* as well as the critical mass ratio were found to be given by
(30)
(31)
(32)
(33)
where Z is the ion to electron charge ratio. For hydrogen, it follows n(1)=0.9×1024cm3,n(2)=1028cm3, and T*= 66 000 K. The approximate range of the proton crystal is indicated in Fig. 4 by the red dotted line and denoted as “solid (atomic) hydrogen.” Note that, for more accurate estimates of the solid–liquid transition, one has to take into account screening of the ion–ion interaction by the electrons,232,235 e.g., via an effective ion–ion potential, an effective screening dependent coupling parameter or based on the shape of the pair distribution function (e.g., Refs. 236–238).

As noted above, the crystallization of heavy charged particles is also relevant for other systems. Prominent examples are white dwarf stars where crystallization of carbon ions (C6+) and oxygen ions (O8+) is expected to occur in the core. Due to the increased ion mass, the inequality (28) leads to a much broader density range and higher maximum temperature. Equations (30)–(32) yield, for carbon (oxygen), n(1)=2×1026cm3(6.6×1026cm3),n(2)=3.7×1033cm3(2.7×1034cm3), and T*=109K(4.2×109K). As a final note, liquid or crystal formation is also relevant for certain semiconductors. In fact, hole crystallization was predicted, among others, by Abrikosov239,240 who predicted a critical hole to electron mass ratio of Mcr100 for CuCl. Fermionic PIMC simulations for electron–hole plasmas yielded172,241,242 Mcr83, in three dimensions (3D), and Mcr60, in 2D [cf. Eq. (33)]. Such mass ratios are feasible in intermediate valence semiconductors, such as Tm[Se, Te].243 

We conclude the analysis of the phase diagram of dense hydrogen by outlining open problems and indicating possible solutions. As we have seen before, the phase diagram of hydrogen is surprisingly complex and identification of phases and phase transitions or crossovers very sensitively depends on the analysis tools. As we have seen in the example of the shock Hugoniot (cf. Sec. I and Fig. 2), various experiments and simulation methods may lead to strongly differing predictions which requires a careful analysis.

  • The high-pressure phases of condensed hydrogen and the LLPT are particularly complex, with many competing possible orders. The simulation results presented in Sec. II B 2 and summarized in the phase diagrams of Figs. 4–6 include extensive predictions that still await experimental tests. For completeness, we mention that a different picture has been put forward by Fedorov et al. (see Ref. 244 and references therein). Based on restricted open shell Kohn–Sham simulations they predict the existence of excitons in the low-temperature fluid. The dissociation of excitons would represent an efficient energy transfer from nuclear to electronic degrees of freedom beyond the BO approximation which, however, should be clearly visible in the optical data. They hypothesize the existence of a plasma phase and PPT, at low temperatures.

  • Accurate experiments and first principle simulations are needed for definitive findings. Reliable simulation results are of great value for benchmarking other less fundamental methods. Examples are fermionic PIMC simulations for the jellium model,117,121,127 as well as for hydrogen115 as discussed in Sec. IV A.

  • The DFT and QMC simulations are currently limited to treating a few hundred atoms. While pressure and internal energy seem to converge fairly quickly with increasing system size, this limitation may lead to biases in the vicinity of phase transitions. Extrapolation to the thermodynamic limit (cf. Sec. III A 4) is necessary to establish an accurate long-range order and critical exponents. Methods that can treat larger system sizes are helpful. A promising technique is to combine DFT or QMC results with machine learning techniques (e.g., Refs. 206, 245, and 246) but the accuracy and reliability to reproduce in general the underlying DFT or QMC simulations still needs to be established.

  • The high-temperature and high-pressure range is of particular interest for ICF (cf. Fig. 3). This wide range of material conditions places a severe set of constraints on computational techniques used to compute equations of state and transport properties. The radiation-hydrodynamic codes (cf. Sec. III H 5) use physics models such as the equation of state, electrical and thermal conductivity, and plasma viscosity as input to the Navier–Stokes equations. It is important that the phenomena are well understood theoretically and modeled accurately so that the ICF design simulations are as accurate as possible.

A major challenge for experiments with matter at extreme densities, temperatures, and pressures, in general, and dense hydrogen, in particular, is given by accurate diagnostics. Indeed, very often, basic parameters such as density and temperature cannot be directly measured and have to be inferred from other observations and theories. The velocity interferometer for any reflector (VISAR) diagnostics is a standard tool to obtain pressure and density in a shocked state of the sample.247,248 If the pressure standard is known to have good accuracy and the shock is planar enough, it will deliver reliable results. Streaked optical pyrometry (SOP) is a diagnostic used to extract the temperature of the sample.248,249 However, it usually measures (near) surface temperatures and the results depend on the (model-dependent and often unknown) reflectivities of the warm dense matter state in a gray body model.

A potentially particularly powerful method of diagnostics that overcomes problems inherent in the standard diagnostics fielded in WDM experiments is given by x-ray Thomson scattering (XRTS).158,250 XRTS probes the electronic dynamic structure factor (DSF), See(q,ω), convolved with the combined source-and-instrument function R(ωs)161 
(34)
The momentum transfer q is determined by the scattering geometry, ω=ω0ωs denotes the energy loss of the scattered photon, and ω0 is the beam energy. In principle, the DSF gives one detailed insight into the microphysics of the probed sample and the measured scattering intensity can be used to infer key parameters, such as T, ρ, and the effective charge Z.163,248,251–254

In practice, however, XRTS experiments with hydrogen are notoriously difficult. Hydrogen has a small scattering cross section, leading to a low photon count and, consequently, a noisy intensity signal (e.g., Ref. 255). This is particularly problematic for XRTS experiments with fusion plasmas, which have a very low repetition rate (e.g., one to three shots per day at the NIF256). XRTS experiments with hydrogen jets257,258 at modern x-ray free electron laser (XFEL) facilities, such as LCLS in the USA or the European XFEL in Germany, on the other hand, allow one to average over thousands of shots. Here, a crucial problem is given by the uncertain degree of inhomogeneity and nonequilibrium of the sample.259 

Let us conclude by outlining two additional, general challenges of XRTS experiments with warm dense matter. On the experimental side, the accurate interpretation of the measured intensity requires detailed knowledge of the source-and-instrument function R(ωs), which is a nontrivial task for both backlighter sources260 and XFEL facilities.253 From the theoretical perspective, understanding the measured signal usually requires model or simulation results for See(q,ω), which is a difficult task. In our opinion, the most flexible tool for this purpose is given by time-dependent DFT (TDDFT), which is discussed in Secs. III H 4 and IV F 1. An alternative strategy has recently been suggested by Dornheim et al.,17,160,164 who have proposed to evaluate Eq. (34) in the imaginary-time domain (see Sec. IV D 2). Remarkably, this allows for model-free diagnostics of parameters such as the temperature160,161 and the static structure factor See(q),162 and allows for direct comparisons with quasi-exact PIMC simulations148,163 (see also Sec. IV D 3).

Given these recent developments, we are confident that future XRTS experiments with hydrogen will give valuable new insights into the equation of state components and other important properties.

1. Basic equations

The theoretical description of hydrogen starts with the Hamiltonian of Ne=Np=N electrons and protons with coordinates (r1,,rN) and (R1,,RN), respectively,
(35)
(36)
(37)
(38)
where w(r)=q2/r is the repulsive Coulomb potential of two electrons (or protons) with q=e0(e0). The total Hamiltonian Ĥ [Eq. (35)] can be decomposed in various ways, e.g., into contributions of electrons, Ĥe, protons, Ĥp, and electron–proton interactions, Ŵep, or into kinetic, K̂=K̂e+K̂p, and interaction energy, V̂=Ŵee+Ŵpp+Ŵep. For simulations of dense hydrogen, we will assume a macroscopic, spatially uniform system without external potentials. Exceptions (additional terms in the hamiltonian) are Sec. IV D, where we compute the density response of hydrogen by studying the reaction to a harmonic field, and, Sec. IV F 3, where the time evolution is studied that follows the impact of an energetic particle beam.

There exist a few methods that allow us to solve the many-body problem with the Hamiltonian (35) exactly (i.e., without approximations or statistical noise from sampling). These include exact diagonalization (configuration interaction, CI) and density matrix renormalization group (DMRG) approaches which are, however, limited to small particle numbers N and/or small number of basis functions and have not been used for dense hydrogen. They are valuable for benchmarking of other methods in model cases that might be relevant to dense hydrogen.

Before discussing simulation approaches that are being applied to dense hydrogen, we briefly summarize the main many-particle properties that are of interest, both for comparison with experiments and for basic physical understanding.

2. Observables of interest

The main quantities of interest can be classified into the following groups for which we also indicate the sections where they are discussed in this paper.

  • The first relevant set of observables is given by the thermodynamic properties of dense hydrogen. This includes the equation of state (EOS, pressure) (Sec. IV A 3), total energy, free energy etc. More detailed information on the structural properties and interaction effects is obtained from the species-resolved static structure factors Sab(q) and pair distribution functions gab(r) which will be studied in Sec. IV A 5. Complementary microscopic information is contained in both the electronic and ionic momentum distributions ne(q) and ni(q) (Sec. IV C), which are important, e.g., for the characterization of the effective charge state or, in the case of the ions, for the estimation of ion impact reaction or nuclear fusion rates (e.g., Refs. 261 and 262).

  • The second group of important observables describe electronic dynamic properties. These include the dynamic structure factor, See(q,ω), which is the key property in XRTS experiments [cf. Eq. (34)], and the imaginary-time correlation function (ITCF), Fee(q,τ), which is connected to the former via a two-sided Laplace transform [Eq. (121)]. Interestingly, the ITCF has emerged as an important observable for XRTS diagnostics in its own right,160–163,263 as it gives one straightforward and model-free access, e.g., to the temperature.

  • The third group of observables are electronic transport and optical properties, such as the conductivity σ(ω), the heat conductivity λ(ω) (cf. Sec. IV E 3), and the opacity κ(ω) (Sec. IV E 4). The latter is a key property in optical transmission experiments, e.g., at the NIF.264 

  • A related fourth group of properties includes spectral information, such as the density of states (DOS), as well as the single-particle spectral function, A(q,ω), and the Matsubara or nonequilibrium Green functions (Sec. III H 3). Note that the spectral function can be directly measured in photoemission experiments. Closely related properties are the energy gap that is important to characterize the low-temperature phases of hydrogen (cf. Sec. IV C 3) and the ionization potential depression (Sec. IV B 1), which characterizes the ionization of high density hydrogen in the gas phase.

  • The fifth set of relevant observables is given by the ion dynamic properties, that have traditionally been estimated on the basis of molecular dynamics simulation. In particular, we consider the ionic structure factor, Sii(q,ω), which is important for an analysis of the collective modes (ion acoustic modes) and sound speed of the heavy particles (Sec. IV D 6). Other important observables are the diffusivity, Dion, viscosity, ηion, and the ionic thermal conductivity λion.

  • The sixth group of observables are related to important non-equilibrium properties which characterize the equilibration of hydrogen after an external excitation. We list three examples: the relaxation time, trel, during which the electron momentum distribution, ne(q,t), thermalizes to a Fermi function (in general, to a correlated equilibrium distribution), the equilibration time between electrons and protons, teq, and the stopping power, Sx (Sec. IV F 3) (and related energy loss properties).

TABLE III.

First-principles simulation methods for dense hydrogen in thermodynamic equilibrium at finite temperature ordered by their expected accuracy. Numbers I.–V. in the column “Observables” refer to the list in Sec. III A 2. GM(q,τ): Matsubara Green function of “imaginary time” τ. The column “benchmarks” lists available relevant benchmarks of pressure (p), interaction energy (V), degree of ionization (α). For new benchmarks provided in this paper, the relevant figures are indicated, for details see footnotes and main text. For a more complete list of observables, see Table VI.

Method Observables Approximations Main limitations Benchmarks Section
Fermionic PIMC (FPIMC)  I., II. GM(q,τ) nonlinear response  Number of particles N number of high-T factors P  Fermion sign problem Θ0.5 statistical errors  N/A  III E 1 
Restricted PIMC (RPIMC)  I.  Fixed node approximation  Θ0.1  p,α based on FPIMC (Figs. 14 and 23 III E 4 
CEIMC  I., IV. DOS, energy gap  BO approximation Electrons in ground state using VMC or RQMC  Θ0.1  N/A  III F 
Green functions  I., II., III., IV. A(q,ω) DOS  Selfenergy Σ  Moderate coupling strength  Interaction energy based on FPIMCa model systems  III H 3 
Kohn–Sham-DFT-MD  I., V. III.b  BO approx., XC functional    p for LDA, PBE and KDT16 (Fig. 15 III C 
Linear-response TDDFT  II.  BO approx., XC functional XC-kernel  No reliable dynamic XC-kernel  N/Ac  III H 4 
Orbital free DFT-MD  I., V.  BO approx., XC functional non-interacting free-energy functional  Accurate only for relatively high T due to free-energy functional  N/A  III C 
Method Observables Approximations Main limitations Benchmarks Section
Fermionic PIMC (FPIMC)  I., II. GM(q,τ) nonlinear response  Number of particles N number of high-T factors P  Fermion sign problem Θ0.5 statistical errors  N/A  III E 1 
Restricted PIMC (RPIMC)  I.  Fixed node approximation  Θ0.1  p,α based on FPIMC (Figs. 14 and 23 III E 4 
CEIMC  I., IV. DOS, energy gap  BO approximation Electrons in ground state using VMC or RQMC  Θ0.1  N/A  III F 
Green functions  I., II., III., IV. A(q,ω) DOS  Selfenergy Σ  Moderate coupling strength  Interaction energy based on FPIMCa model systems  III H 3 
Kohn–Sham-DFT-MD  I., V. III.b  BO approx., XC functional    p for LDA, PBE and KDT16 (Fig. 15 III C 
Linear-response TDDFT  II.  BO approx., XC functional XC-kernel  No reliable dynamic XC-kernel  N/Ac  III H 4 
Orbital free DFT-MD  I., V.  BO approx., XC functional non-interacting free-energy functional  Accurate only for relatively high T due to free-energy functional  N/A  III C 
a

Benchmarks of potential energy of Ref. 117 for jellium.

b

Optical properties can be computed from a set of KS-orbitals via the Kubo–Greenwood relation that does, however, not include Electron–electron collision (see Sec. IV E).

c

Note that LR-TDDFT results for different XC-functionals and XC-kernels can be benchmarked against PIMC results for the ITCF F(q, τ) [cf. Eq. (121)] in future studies.

The above list does not pretend to be complete but concentrates on quantities that have been in the focus recently. Other observables that are potentially interesting for future investigations, but which are not considered here, include various nonlinear response functions,265–269 effective forces,270 and potentials.271,272

Now, obvious questions arise: which simulation approaches are available to compute the above quantities? And, in case a computation is possible, how accurate and reliable are the results? How can the accuracy be estimated or verified? Finally, if several methods are available—which one is preferable? In the following, we attempt answers to these questions, with a focus on applications to dense hydrogen. A comparison of accuracy and limitations of various methods is presented in Tables III–V. A summary of the relation between methods and observables is presented in Table VI.

TABLE IV.

Further simulation methods for dense hydrogen in thermodynamic equilibrium at finite temperature, ordered by their expected accuracy. Numbers I.–VI. in the column “Observables” refer to the list in Sec. III A 2. α: degree of ionization, xA: degree of dissociation (atom fraction). For a more complete list of observables, see Table VI.

Method Observables Approximations Main limitations Benchmarks Section
Average atom model  I., II., III., IV., V.  XC functional, single ion center  Surrounding atoms not treated explicitly  PDF based on FPIMC (Fig. 18 III D 
Wavepacket MD  I., II. V., VI.  Classical dynamics pair potentials  Wave function modeled by one or few Gaussians approximations needed to prevent uncontrolled spreading  N/A  III G 1 
Semiclassical MD  I., II., V., VI.  Classical dynamics quantum pair potentials, e.g., improved Kelbg pot  kBT0.5 Ry only pair exchange  Pressure p, based on FPIMC (Fig. 14 III G 2 
Chemical models, e.g., FVT  I. part of thermo-dynamic functions  Equilibrium n(k) α,xA  Approximations for interaction spatial homogeneity  p,xA based on FPIMC (Figs. 14 and 22 III B 
Method Observables Approximations Main limitations Benchmarks Section
Average atom model  I., II., III., IV., V.  XC functional, single ion center  Surrounding atoms not treated explicitly  PDF based on FPIMC (Fig. 18 III D 
Wavepacket MD  I., II. V., VI.  Classical dynamics pair potentials  Wave function modeled by one or few Gaussians approximations needed to prevent uncontrolled spreading  N/A  III G 1 
Semiclassical MD  I., II., V., VI.  Classical dynamics quantum pair potentials, e.g., improved Kelbg pot  kBT0.5 Ry only pair exchange  Pressure p, based on FPIMC (Fig. 14 III G 2 
Chemical models, e.g., FVT  I. part of thermo-dynamic functions  Equilibrium n(k) α,xA  Approximations for interaction spatial homogeneity  p,xA based on FPIMC (Figs. 14 and 22 III B 
TABLE V.

Nonequilibrium simulation methods for dense hydrogen ordered by their expected accuracy. QKT: quantum kinetic theory. Numbers I.–VI. in the column “Observables” refer to the list in Sec. III A 2. n(r,t),u(r,t): density and velocity field. neq(k): equilibrium momentum distribution; ω(q): plasmon dispersion. nl excitations: nonlinear excitations. For a more complete list of observables, see Table VI.

Method Observables Approximations Main limitations Benchmarks Section
Nonequilibrium Green functions (NEGF, QKT)  I.–IV., VI. nonequilibrium DOS  Selfenergy Σ  Moderate coupling strength model systems  ni(t) based on DMRGa  III H 1 
Real-time TDDFT  II., III., VI.  XC potential adiabatic approximation  Accuracy of electron collisions unclear  N/A  III H 4 
Quantum Hydrodynamics  n(r,t),u(r,t) ω(q) nl excitations, shocks  n(k)neq(k) no exchange effects  Θ1, Γ1 limited length & time resolution  Forces and potentials based on KS-DFTb  III H 5 
Hydrodynamics  n(r,t),u(r,t) ω(q) nl excitations, shocks  n(k)neq(k) no quantum effects no exchange effects  Θ1, Γ1 limited length & time resolution  N/A  III H 5 
Method Observables Approximations Main limitations Benchmarks Section
Nonequilibrium Green functions (NEGF, QKT)  I.–IV., VI. nonequilibrium DOS  Selfenergy Σ  Moderate coupling strength model systems  ni(t) based on DMRGa  III H 1 
Real-time TDDFT  II., III., VI.  XC potential adiabatic approximation  Accuracy of electron collisions unclear  N/A  III H 4 
Quantum Hydrodynamics  n(r,t),u(r,t) ω(q) nl excitations, shocks  n(k)neq(k) no exchange effects  Θ1, Γ1 limited length & time resolution  Forces and potentials based on KS-DFTb  III H 5 
Hydrodynamics  n(r,t),u(r,t) ω(q) nl excitations, shocks  n(k)neq(k) no quantum effects no exchange effects  Θ1, Γ1 limited length & time resolution  N/A  III H 5 
a

Benchmarks of site occupations, ni(t), and doublon number of Ref. 274 for Hubbard model.

b

Benchmarks of Ref. 275.

3. Hierarchy of simulation approaches

For dense hydrogen in the ground state, the most accurate results, so far, have been obtained using diffusion or variational Monte Carlo (DMC, VMC). Ground state properties are not the focus of the present paper, for an overview (see Ref. 156). For finite temperature properties, one has to switch to a mixed ensemble description based on the N-particle density operator ρ̂. Here, the most accurate approach is provided by fermionic path integral Monte Carlo (FPIMC) which is free of systematic errors, and convergence to the exact thermodynamic limit can be achieved by proper extrapolation with respect to the particle number N and the number of high-temperature factors P (see Sec. III E 1). FPIMC is the only method which is capable of producing unbiased accurate predictions without requiring benchmarks by other simulations. The limiting factor of FPIMC is the fermion sign problem (FSP, cf. Sec. III E 2) manifesting itself by an exponentially small signal to noise ratio, which currently limits simulations to temperatures T0.5TF. There are various concepts to avoid this limitation by formulating PIMC in second quantization (CPIMC) (see, e.g., Sec. III E), but the accessible parameter range remains restricted by the FSP.

To extend the simulations of hydrogen to the parameters of interest (or to more complex systems), as well as to increase the computational efficiency, a large arsenal of approximate methods has been developed, both for equilibrium and nonequilibrium situations. An overview is presented in Tables IV–VI, and a more detailed discussion is given in Secs. III A 4 and III A 5. The purpose of these tables is to provide a compact comparison of important methods which includes their applicability range and the quantities that they are capable of providing. The methods are listed on the order of their accuracy that is generally expected, based on their construction and the involved basic approximation schemes. Note that the actual accuracy of different methods, when applied to dense hydrogen, may differ from these expectations. While the ultimate tests are, of course, experiments, they are extremely challenging and complex in the case of dense hydrogen and currently afflicted by large uncertainties or disagreements between different experimental groups and techniques. Therefore, we base our tests on computer experiments for which the accuracy has been convincingly established and which serve as benchmarks. These are fermionic PIMC (FPIMC) results for the uniform electron gas (e.g., Refs. 117 and 129) and hydrogen.115 Presently, only a few such benchmarks are available; therefore, in the present paper, we produce additional comparisons which are all listed in Tables IV and V.

Note that these tests refer to specific parameter ranges of dense plasmas and they are performed for a special choice of approximation (such as self-energy or exchange–correlation functional in the case of Green functions or DFT, respectively). Therefore, the results of the benchmarks cannot be directly generalized to other parameters or approximations. Nevertheless, they provide valuable guidelines for the behavior of different methods and approximations, when applied to dense hydrogen.

In the following, we summarize the results of the available benchmarks, starting with Table IV. (1) Among the approximate methods, RPIMC (even with free-particle nodes) is the most accurate one, achieving an accuracy of the equation of state of better than 2%, in a broad parameter range of partially ionized hydrogen (currently, no tests are possible for high densities, rs34). Only at the lowest temperature, T 15 000 K, deviations reach 6%115 (cf. Figs. 14 and 23 in Sec. IV A 3). (2) Equilibrium Green functions (EGF) with weak coupling self-energies achieve an accuracy better than 10%, for the thermodynamic properties of the model UEG, for rs1. With the cumulant approach, Kas and Rehr273 achieved similar accuracy for a significantly broader range of rs-values. (3) Kohn–Sham DFT with PBE exchange–correlation functionals exhibits deviations of up to 7% for the pressure of hydrogen in the density-temperature range corresponding to rs4 and T30000K. When finite-temperature functionals are being used, the accuracy improves substantially, to better than 2%, for T60000K, and larger deviations for T 30 000 K, for details (see Fig. 15 in Sec. IV A 4). Orbital-free (OF) DFT is generally expected to be significantly less accurate than KS-DFT, but no benchmarks are available so far. Linear-response TDDFT is expected to yield accurate dynamical quantities on the accuracy level of KSDFT (cf. Sec. III H 4), but no benchmarks are available yet.

A second group of methods that involve stronger approximations than the first are listed in Table V. They are expected to be of lower accuracy than the methods in Table IV. A few benchmarks are performed in this paper for semiclassical MD with quantum potentials (cf. Fig. 14 in Sec. IV A 3) and for chemical models (cf. Figs. 14 and 22 in Sec. IV A 3). For example, semiclassical MD with the improved Kelbg potential is able to reproduce the equation of state for temperatures above 60 000 K below a critical density (that varies with temperature) with an accuracy of the order of 1%–3% (see the discussion of Fig. 14 in Sec. IV A 3). On the other hand, standard fluid variational theory (FVT) that does not include ionization reproduces the equation of state at low temperatures within approximately 20% but fails at temperatures exceeding 20 000 K.

Important simulation methods that allow one to access nonequilibrium properties are summarized in Table VI. So far they have been tested much less and not for dense hydrogen, for which currently no benchmarks exist. The most accurate approach is nonequilibrium Green functions (NEGF) and the associated quantum kinetic equations (QKE) for which a few tests for lattice systems against DMRG exist. They indicate that, for weak to moderate coupling, time-dependent observables can be computed with errors below 10%, provided the proper high-level self-energies (such as third order approximation, GW or T-matrix) are being used. The remaining nonequilibrium approaches listed in Table VI involve significantly more restrictive approximations than NEGF, but presently no accuracy tests are available. Benchmarks could be produced in the future based on NEGF simulations by designing well-defined model cases for which all methods of interest are feasible. Examples include small systems or lattice models.

TABLE VI.

Summary of relevant observables and simulation methods available to compute them. EOS: equation of state (including energy, pressure and other thermodynamic variables); See(q), Sii(q), Sei(q): static structure factors; ne(q): electronic momentum distribution function; See(q,ω): electronic dynamic structure factor; Fee(q,τ): ITCF [Eq. (121)]; σ(ω): electric conductivity in the optical limit; κ(ω): opacity; DOS: density of states; IPD: ionization potential depression; Sii(q,ω): ion dynamic structure factor; Di: diffusion coefficient; ηi: viscosity; λi: ion thermal conductivity; trel: relaxation time of the electron momentum distribution; teq electron–ion equilibration time; Sx: stopping power. Yes (roman) indicates a demonstrated capability, Yes indicates the potential capability and/or some significant limitation that is explained in a footnote, and No indicates that estimating the observable in question is fundamentally not possible. Note that we do not compare the accuracy of different methods and do not list limitations of the parameter range that is accessible by different methods; this information is provided in Tables III–V.

FPIMC RPIMC CEIMC KS-DFT OF-DFT AA WP-MD SC MD CM RT-TDDFT GF, QKTa,b
EOS  Yes  Yes  Yes  Yes  Yes  Yes  Yes  Yes  Yes  No  Yes 
See(q)  Yes  Yes  Yes  Yesc  No  Yesd  Yes  Yes  No  Yese  Yes 
Sii(q)  Yes  Yes  Yes  Yes  Yes  Yes  Yes  Yes  No  No  No 
Sei(q)  Yes  Yes  Yes  Yes  Yes  Yes  Yes  Yes  No  No  No 
ne(q)  Yes  Yes  Yes  Yesf  No  No  Yesg  No  No  No  Yes 
See(q,ω)  Yesh  No  No  Yes  No  Yes  Yesi  Yes  No  Yes  Yes 
Fee(q,τ)  Yes  No  No  Yes  No  Yes  Yesi  No  No  Yes  Yes 
σ(ω)  No  No  No  Yes  No  Yes  No  No  No  Yes  Yes 
κ(ω)  No  No  No  Yes  No  Yes  No  No  No  Yes  Yes 
DOS  Yesj  No  No  Yes  No  Yes  No  No  No  No  Yes 
IPD  Yesk  No  No  Yes  No  Yes  No  No  Yes  Yes  Yes 
Sii(q,ω)  No  No  No  Yes  Yes  Yes  Yes  Yes  No  No  No 
Dion  No  No  No  Yes  Yes  Yes  Yes  Yes  No  No  No 
ηion  No  No  No  Yes  Yes  Yes  Yes  Yes  No  No  No 
λion  No  No  No  Yes  Yes  Yes  Yes  Yes  No  No  No 
trel  No  No  No  No  No  No  No  No  No  Yesl  Yes 
teq  No  No  No  Yesm  Yesn  Yeso  Yes  Yes  No  Yes  Yes 
Sx  No  No  No  Yesp  Yesq  Yesr  Yes  Yes  No  Yes  Yes 
FPIMC RPIMC CEIMC KS-DFT OF-DFT AA WP-MD SC MD CM RT-TDDFT GF, QKTa,b
EOS  Yes  Yes  Yes  Yes  Yes  Yes  Yes  Yes  Yes  No  Yes 
See(q)  Yes  Yes  Yes  Yesc  No  Yesd  Yes  Yes  No  Yese  Yes 
Sii(q)  Yes  Yes  Yes  Yes  Yes  Yes  Yes  Yes  No  No  No 
Sei(q)  Yes  Yes  Yes  Yes  Yes  Yes  Yes  Yes  No  No  No 
ne(q)  Yes  Yes  Yes  Yesf  No  No  Yesg  No  No  No  Yes 
See(q,ω)  Yesh  No  No  Yes  No  Yes  Yesi  Yes  No  Yes  Yes 
Fee(q,τ)  Yes  No  No  Yes  No  Yes  Yesi  No  No  Yes  Yes 
σ(ω)  No  No  No  Yes  No  Yes  No  No  No  Yes  Yes 
κ(ω)  No  No  No  Yes  No  Yes  No  No  No  Yes  Yes 
DOS  Yesj  No  No  Yes  No  Yes  No  No  No  No  Yes 
IPD  Yesk  No  No  Yes  No  Yes  No  No  Yes  Yes  Yes 
Sii(q,ω)  No  No  No  Yes  Yes  Yes  Yes  Yes  No  No  No 
Dion  No  No  No  Yes  Yes  Yes  Yes  Yes  No  No  No 
ηion  No  No  No  Yes  Yes  Yes  Yes  Yes  No  No  No 
λion  No  No  No  Yes  Yes  Yes  Yes  Yes  No  No  No 
trel  No  No  No  No  No  No  No  No  No  Yesl  Yes 
teq  No  No  No  Yesm  Yesn  Yeso  Yes  Yes  No  Yes  Yes 
Sx  No  No  No  Yesp  Yesq  Yesr  Yes  Yes  No  Yes  Yes 
a

Green functions (GF, including equilibrium and nonequilibrium Green functions) are the basis for standard and improved quantum kinetic equations (QKT). Simulations for dense hydrogen are currently limited to full ionization and moderate coupling.

b

Extension to partially ionized hydrogen appears realistic by using Kohn–Sham orbitals as input.

c

By integrating over linear-response TDDFT results for See(q,ω) involving additional uncontrolled approximations such as Kxc(q,ω).

d

By integrating over See(q,ω) computed from an approximate dynamic collision frequency in the optical limit.

e

Indirectly by integrating over See (q,ω).

f

In principle, one can compute ne(q) by integrating over the spectral function, e.g., in the GW approximation.276 

g

Using fully anti-symmetrized models non-classical momentum distributions are obtained277, however, the wave packet approximation will influence the result when q−1 is comparable to and smaller than the size of the wave packet.

h

Via a difficult analytic continuation of Fee(q, τ) [cf. Eq. (121)] (see, e.g., Ref. 166 for the UEG model).

i

The limited functional form is noticeable for q−1 comparable to packet size and the computations has previously been computed in a semi-classical manner278 but quantum formulations are under development.

j

While not having been demonstrated, it is in principle possible to compute the Matsubara Green function GM(q, τ),279 use it as the basis for a reconstruction of the single-particle spectral function A(q, τ)280, and then integrate the latter over the momentum q.

k

Despite the absence of orbitals in PIMC, information about IPD is encoded in Fee(q, τ), but this has not yet been explored for hydrogen. An alternative approach to the IPD is presented in Sec. IV B 2.

l

Computing trel is possible in principle, but would require a fully dynamic XC-potential, which is unknown in practice.

m

Within linear response, using KS-DFT obtained electron–phonon matrix elements or electron–ion friction coefficients.281,282

n

Using Sii(q) in approximate models for electron–ion collisions (e.g., see the discussions in Ref. 283) and combining it with the Mermin model for the dielectric function of electrons.284 

o

Using an effective ion potential in approximate models.285,286

p

Electronic stopping power is computed on the linear response level.

q

Via the Mermin dielectric function (131), by using Sii(q) in models for the calculation of the electron–ion collision frequency (e.g., in the Ziman formula287).

r

Within the linear response regime.288 

TABLE VII.

Pressure (in Mbar) for three isotherms, for fermionic (FP-PIMC) and restricted PIMC (RPIMC), KS-DFT with PBE XC functional and KDT16 functional, respectively, as well as for semiclassical MD with improved Kelbg (I. Kelbg) potential (for T=62500 K) and fluid variational theory (for T=15625 K and T=31250 K). Data correspond to Figs. 14 and 15. The FP-PIMC data for low densities, 14<rs100 can be found in Ref. 115.

T (K) rs FP-PIMC RPIMC DFT-MD (KDT16) DFT-MD (PBE) FVT
15 625  4.0  0.0452 (13)  0.047 90 (66)    0.0449 (07)  0.045 22 
5.0  0.024 09 (52)  0.025 37 (20)    0.0238 (55)   
6.0  0.014 84 (14)  0.014 85 (12)    0.0144 (37)  0.014 34 
8.0  0.6743 (50) × 10−2  0.659 (5) × 10−2    0.666 (2) × 10−2   
10.0  0.3572 (27) × 10−2  0.349 (2) × 10−2    0.346 (6) × 10−2  0.3346 × 10−2 
12.0  0.211 71 (98) × 10−2  0.205 (1) × 10−2       
14.0  0.134 37 (54) × 10−2  0.129 (1) × 10−2      0.1247 × 10−2 
31 250  4.0  0.1158 (31)  0.115 00 (53)  0.113 087 52  0.121 303 79  0.1002 
5.0  0.0610 (4)  0.059 80 (19)  0.060 772 701  0.063 940 243   
6.0  0.036 24 (23)  0.035 56 (11)  0.036 769 872  0.038 098 258  0.031 
8.0  0.015 877 (42)  0.015 80 (6)  0.016 679 863  0.016 922 071   
10.0  0.857 24 (71) × 10−2  0.845 (2) × 10−2      0.6884  ×102 
12.0  0.518 71 (36) × 10−2  0.511 (1) × 10−2       
14.0  0.339 20 (22) × 10−2  0.336 (1) ×102      0.2524  ×102 
62 500  2.6  1.044 (12)  1.048 00 (410)  1.043 151 9  1.097 589 5   
3.0  0.6901 (71)  0.673 00 (170)  0.680 517 94  0.712 121 22   
4.0  0.2966 (13)  0.295 00 (60)  0.298 404 59  0.308 110 06   
5.0  0.159 61 (18)  0.158 00 (16)  0.159 927 42  0.162 677 45  MD (I. Kelbg) 
6.0  0.096 196 (53)  0.095 50 (9)  0.096 265 651  0.096 967 732  0.0924 
8.0  0.043 430 (22)  0.043 25 (4)      0.0431 
10.0  0.023 37 (1)  0.023 25 (2)      0.0235 
12.0  0.014 045 2 (60)  0.013 98 (1)      0.0141 
  14.0  0.909 93 (40) ×102  0.905 (1) ×102      0.906  ×102 
T (K) rs FP-PIMC RPIMC DFT-MD (KDT16) DFT-MD (PBE) FVT
15 625  4.0  0.0452 (13)  0.047 90 (66)    0.0449 (07)  0.045 22 
5.0  0.024 09 (52)  0.025 37 (20)    0.0238 (55)   
6.0  0.014 84 (14)  0.014 85 (12)    0.0144 (37)  0.014 34 
8.0  0.6743 (50) × 10−2  0.659 (5) × 10−2    0.666 (2) × 10−2   
10.0  0.3572 (27) × 10−2  0.349 (2) × 10−2    0.346 (6) × 10−2  0.3346 × 10−2 
12.0  0.211 71 (98) × 10−2  0.205 (1) × 10−2       
14.0  0.134 37 (54) × 10−2  0.129 (1) × 10−2      0.1247 × 10−2 
31 250  4.0  0.1158 (31)  0.115 00 (53)  0.113 087 52  0.121 303 79  0.1002 
5.0  0.0610 (4)  0.059 80 (19)  0.060 772 701  0.063 940 243   
6.0  0.036 24 (23)  0.035 56 (11)  0.036 769 872  0.038 098 258  0.031 
8.0  0.015 877 (42)  0.015 80 (6)  0.016 679 863  0.016 922 071   
10.0  0.857 24 (71) × 10−2  0.845 (2) × 10−2      0.6884  ×102 
12.0  0.518 71 (36) × 10−2  0.511 (1) × 10−2       
14.0  0.339 20 (22) × 10−2  0.336 (1) ×102      0.2524  ×102 
62 500  2.6  1.044 (12)  1.048 00 (410)  1.043 151 9  1.097 589 5   
3.0  0.6901 (71)  0.673 00 (170)  0.680 517 94  0.712 121 22   
4.0  0.2966 (13)  0.295 00 (60)  0.298 404 59  0.308 110 06   
5.0  0.159 61 (18)  0.158 00 (16)  0.159 927 42  0.162 677 45  MD (I. Kelbg) 
6.0  0.096 196 (53)  0.095 50 (9)  0.096 265 651  0.096 967 732  0.0924 
8.0  0.043 430 (22)  0.043 25 (4)      0.0431 
10.0  0.023 37 (1)  0.023 25 (2)      0.0235 
12.0  0.014 045 2 (60)  0.013 98 (1)      0.0141 
  14.0  0.909 93 (40) ×102  0.905 (1) ×102      0.906  ×102 
TABLE VIII.

New results for the electrical conductivity σ given in 1/(Ω·mm) from DFT-MD-Kubo–Greenwood simulations with PBE XC functional, corresponding to the dashed lines in Fig. 42 and to the thermodynamic data shown in Table VII.

rs T (K) 15 625 31 250 62 500
25(3)  61(7)  208(20) 
7.2(7)  48(3)  183(4) 
5.7(6)  46(3)  172(4) 
4.8(5)  47(3)  168(7) 
rs T (K) 15 625 31 250 62 500
25(3)  61(7)  208(20) 
7.2(7)  48(3)  183(4) 
5.7(6)  46(3)  172(4) 
4.8(5)  47(3)  168(7) 

In the following, we give a brief overview of the methods presented in this article which can be loosely grouped into particle-based simulations and continuum models.

4. Particle-based simulations

Simulations using particles are well known from classical physics, the most important ones being thermodynamic (Metropolis or kinetic) Monte Carlo (MC) and molecular dynamics (MD). These simulations are potentially exact, i.e., they are free of approximations and systematic errors (based on “first principles”) if the pair interactions are accurately known. For quantum systems, such as dense hydrogen, a true “first principles” approach is fermionic PIMC (FPIMC), which was discussed above. To extend FPIMC to a broader parameter range and eliminate the fermion sign problem, several approximate methods have been proposed. The most important ones, based on the so-called fixed node approximation, are restricted PIMC (RPIMC, Sec. III E 4) and coupled electron–ion QMC (CEIMC, Sec. III F) that have been very successfully applied to dense hydrogen and other WDM systems. While the accuracy of the fixed node approximation is not known a priori, benchmarks against FPIMC simulations for dense hydrogen at Θ0.5 confirmed that the relative error of RPIMC in the equation of state and the total energy does not exceed a few percent (cf. Table III and Sec. IV A 3). Moreover, the fixed node approximation, both in the ground state and at finite temperature, is variational in nature which provides an internal consistency check among different nodal restrictions and the possibility of systematic improvement by more flexible reference density matrices (RPIMC) or electronic ground state wave functions (CEIMC).

The second successful particle-based quantum simulation approach is the combination of Kohn–Sham DFT with MD for the ions (cf. Sec. III C). It uses a much more severe set of approximations compared to RPIMC: the first is the Born–Oppenheimer approximation to decouple the electron and ion dynamics. The second is a many-body approximation to account for interaction effects which is formulated in terms of the exchange–correlation (XC) functional (see Table III). There exist a number of approximate functionals but their applicability range and accuracy are not known a priori; they have to be established by benchmarks for model systems or against FPIMC simulations, which will be done in Sec. IV A. A computationally less expensive alternative to Kohn–Sham DFT is given by the orbital-free DFT method, which is available at arbitrarily high temperatures and allows for the simulation of much larger systems. This advantage comes at the cost of an additional approximation to the kinetic energy of the noninteracting reference system, making it less accurate than other methods at low to moderate temperatures. While the computed pressure was found to be fairly accurate, significant errors in the internal energy were identified289 and deviations in the computed shock Hugoniot curves were reported.290 

Particle-based simulations use a finite simulation cell with a small to moderate particle number, N10103, and the results are repeated for an increasing number N. Scaling vs N1 is performed to establish convergence to the thermodynamic limit (TDL) for all quantities of interest (see, e.g., Refs. 115, 117, 127, and 291–294). If the scaling of certain observables with the particle number is known analytically, explicit finite size corrections can be derived and the transition to the TDL be performed for any finite value of N.

A prominent example for the latter case is the interaction energy of the UEG, which is readily expressed as an integral over the static structure factor S(q). This allows for two potential sources of finite-size errors: (a) the explicit dependence of SN(q) on the system size and (b) the approximation of the continuous q-integral by a discrete sum over lattice vectors in the finite simulation cell. In the ground state, Chiesa et al.291 found that it holds SN(q)S(q), to a remarkable degree. Moreover, they proposed to correct the discretization error in leading order based on the q0 limit of S(q), which is known for the UEG;295 the resulting finite-size correction scales as 1/N and is often sufficient for applications in the ground state. While the idea is easily extended to finite temperatures,296 this first-order correction becomes insufficient at high densities and temperatures.117,127,129 General strategies to correct finite size errors beyond leading order have been discussed293 and successfully applied to hydrogen systems at zero electronic temperatures. At finite temperature, Dornheim et al.127 introduced a beyond leading order correction based on a suitable trial function for S(q) that, in practice, can be computed on the level of the RPA. Finally, we note a recent idea by Dornheim and Vorberger,294 how to remove the small finite-size errors from SN(q) in the high-density regime.

5. Continuum models

Here we briefly discuss models that are formulated in the thermodynamic limit, i.e., the limits N and V have been performed, while the density, n=N/V, has been kept constant. These models, in one way or the other, use approximations for the interaction part, V̂, of the Hamiltonian (35), which are often weak coupling approximations (perturbation expansions). Alternatively, particular relevant physical effects can be taken into account selectively, such as bound states or dynamical screening. Examples of the latter approximation schemes are Feynman diagram expansions of many-body self-energies297,298 or decoupling approximations of the BBGKY hierarchy of reduced density operators.299 

The most accurate, but also the most complex of these approximation schemes, in the case of thermodynamic equilibrium, is Equilibrium (or imaginary time or Matsubara) Green functions. The method depends on a single input quantity—the self-energy Σ—and would be exact, would Σ be known. In practice, of course, one has to use approximations for Σ, and the accuracy of the results is not known a priori (cf. Table III). Here the central quantity is the Matsubara Green function, GM(iωn,k), which, besides statistical information (the k-dependence), also contains spectral information (spectral function and density of states) via the dependence on the Matsubara frequencies ωn. The accuracy of various self-energy approximations is unknown a priori, and only a few tests are available, for the UEG,300,301 for more information (see Sec. III H 3). They confirm good accuracy of EGF, if self-energies within their expected range of validity are being used (see Table III).

A less accurate scheme, compared to Green functions theory, is DFT. It uses a simpler quantity as the basis, compared to the Green function—the density, n(r), which depends only on the coordinate, but not the momentum (or frequency). There exist various approximation schemes for the interaction energy as a functional of n(r) which avoid Kohn–Sham orbitals (orbital-free DFT) (for details, see Sec. III C). Further simplified approaches are (cf. Table V) average atom models (Sec. III D) and chemical models (Sec. III B). The latter are used mainly to compute the mean chemical composition of partially ionized plasmas without any spatial or temporal resolution.

Finally, in Table V, we list nonequilibrium models that capture time-dependent properties and, in part, the relaxation behavior of dense hydrogen. Here the most accurate approach is Nonequilibrium Green functions (NEGF) theory which allows to derive standard quantum kinetic equations as well as improved equations (see Sec. III H 3). Other approaches, again in a sequence of decreasing accuracy, are Real-time TDDFT (Sec. III C 1), wavepacket MD (Sec. III G 1), semiclassical MD (Sec. III G 2), and Quantum and classical hydrodynamics. The latter solve equations for macroscopic space-dependent observables, such as densities, velocity fields, and energy density that are derived from kinetic equations via additional coarse graining (cf. Sec. III H 5).

Using simpler models with more approximations, in general, allows one to reduce the computational effort and access larger length scales, longer simulation times, or more complex systems. Finding the optimal compromise between accuracy and computational cost is an important issue to which we return in Sec. V B.

In the following we provide a detailed discussion of the different simulation approaches.

1. Physical vs chemical picture

First-principles many-particle approaches start with physical particles, such as electrons and protons, as “fundamental” ingredients. This can be regarded as the “physical picture” and it is at the heart of analytical approaches, such as the many-particle Schrödinger equation, density matrix methods, or nonequilibrium Green functions (cf. Sec. III H 3), as well as first-principle simulations, such as quantum Monte Carlo (cf. Sec. III E) or density functional theory (Sec. III C). Of course, the electron states in an isolated hydrogen atom can be rigorously subdivided into bound and scattering states, based on the sign of the electron energy (relative energy of the electron–proton pair). The situation changes immediately at finite temperatures, where thermal energy allows for the excitation of bound electrons into the continuum (ionization). Similarly, in a system of many atoms under high pressure, the electronic states are renormalized by the surrounding particles leading to the formation of bound states of several atoms (e.g., molecules) or the breakup of atoms or molecules, due to pressure ionization (Mott effect) (cf. Sec. II A). In these cases, a clear separation between bound and free particles does not exist anymore. Nevertheless, it is often, both intuitive and technically advantageous, to introduce such a subdivision, as it allows for mapping of some of the complex properties of the many-particle system onto a significantly simpler “chemical picture.”

There exist various criteria for how to “identify bound states” in a first principles approach. This includes the definition of bound state wave functions or of bound contributions of two-particle density matrices and Green functions (e.g., Refs. 68 and 69). In PIMC simulations, one can use criteria based on the electron–proton or proton-proton pair distribution function302 or on the extension of the electron path in the vicinity of a proton (e.g., Ref. 115, for QMC applications, see Sec. IV A 6). On the other hand, in DFT, one might use electronic density gradients,157 as in the electron localization functions (ELF),303 or maximally localized Wannier functions304 to obtain information about the nature of bound states. In DFT-AA models (cf. Sec. III D), there are several plausible definitions of the average ionization.305 Obviously, the results for the fractions of free and bound particles may depend on the criterion, which requires special care if such an analysis is performed. This is the case, in particular, for densities and temperatures around which the degree of ionization or dissociation changes rapidly (cf. Sec. IV A 6).

Once such a subdivision has been performed, a transition to chemical models can be made. The key to all “chemical picture” models is the assumption that electrons, protons, hydrogen atoms, and molecules can be regarded as distinct “chemical” species. This can be extended to treating additional bound states such as molecular ions or different excited atomic states as separate “species.” An extended discussion of why fundamental and composite particles “must be treated in a democratic way” (i.e., on equal footing) can be found in the text book of Ebeling et al.68 The main advantage of chemical models is that one can apply standard tools of thermodynamics or chemical kinetics to compute equilibrium states of dense plasmas including the fractions of free electrons, atoms, and molecules as well as the conditions for phase transitions (cf. Secs. II B 1 and II B 2).

2. Chemical equilibrium: Saha equation

Within the chemical picture, using standard thermodynamics of chemically reacting systems, the equilibrium composition of a hydrogen plasma follows either from minimizing the derivatives of the proper thermodynamic potential (e.g., free energy) with respect to the particle number or by equating the chemical potentials of the species for each “reaction” or balance equation. In the case of hydrogen, there are two main “reactions”: e+pAj and A+AM, where Aj denotes atoms in bound state “j” and M molecules, leading to the thermodynamic stability conditions171 
(39)
(40)
Note that each chemical potential is composed of an ideal part, μa=μaid, that is known, and an interaction contribution which we denote Δμ, hence μa=μaid+Δμa,a=e,p,A,M. Chemical models are formally exact, would all the Δμa be known.306 This can be exploited by mapping results from physical picture simulations, such as DFT or QMC, on a chemical model and extracting unknown parameters. We demonstrate this for the example of the ionization potential depression, using PIMC data, in Sec. IV B 2. In chemical models, however, for the interaction contributions, approximations are being used which we discuss below.

First, we recall the ideal chemical potentials entering the ionization-recombination reaction:

  • For free fermions (electrons, protons), the chemical potential is given by
    (41)

    where I1/21 is the inverse of the Fermi integral of order 1/2, we introduced the spin degeneracy factors, ge=gp=2, and the quantum degeneracy parameter was defined in Eq. (4) and involves the free electron density (fraction α of the total density).

  • For non-degenerate particles (χa1), expansion of the Fermi integral yields
    (42)
  • For hydrogen atoms (which are non-degenerate) in bound state |j=|nlm, the spin degeneracy factor is gA=gegp, and an additional contribution for the relative (bound state) energy appears
    (43)
    (44)

    where En0=E1s/n2, and Δnl(ne,T) is the plasma-induced modification to the bound state energy that adds to the ideal chemical potential. The energy shifts follow from the solution of the Bethe–Salpeter equation.171 

Using the results for classical particles (non-degenerate electrons), the ionization-recombination equilibria (39) become
(45)
where we took into account that Λp3ΛA3. Note that in Eq. (45) there is no explicit interaction contribution of the atoms, ΔμAj, because these effects are taken into account in the level shifts, Δnl. Summing this over all bound states, we obtain the total mass action law (Saha equation)
(46)
(47)
(48)
Here ΣAint denotes the partition function of atomic bound states that are renormalized by the surrounding plasma. In a similar way one derives the dissociation equilibrium (48) (see, e.g., Ref. 69).
Notice that the atomic partition function of Coulomb bound states in vacuum, nn2eβEn0, is diverging (e.g., Ref. 69). Therefore, regularizations that also take into account the spectrum of scattering states, are required. Applying higher-order Levinson's theorems, when performing integration by parts, the leading divergent terms can be projected into the contribution of scattering states so that the well-known Planck–Larkin partition function follows (here we apply this procedure using the renormalized bound state energies, instead of En0)
(49)
which is convergent and replaces ZAint (for details, see Ref. 61). For the molecular partition function, we refer to Refs. 59 and 61 and references therein.
Introducing the fractions of electrons that are unbound (α), bound in atoms (xA) and in molecules (xM) [cf. Eqs. (12)–(14)], and taking into account that ΛMΛA/2, the Saha equations are rewritten in dimensionless form as
(50)
(51)
Finally, we extend this result to the general case of electrons of arbitrary quantum degeneracy. From Eq. (41), it follows for the ionization equilibrium that
(52)
which generalizes Eq. (50) and the subscript “F” denotes Fermi statistics. For any given density and temperature, the two Saha equations can be solved for α and xA, whereas xM follows from Eq. (14). These equations still require input for the partition functions (bound state energy shifts) and for the interaction parts of the chemical potentials.
Before discussing these questions, we illustrate the physical meaning of the ionization equilibrium for the case of low densities and low temperatures where the atomic partition function can be approximated by the ground state, ZAPLZAinteβ(E1s0+Δ1s). Further, we divide Eq. (52) by α, to formally reproduce the classical (low-density) expression
(53)
(54)
(55)
where I=|E1s0| is the (positive) ionization potential of a free atom, which is modified by the remaining terms, ΔI and ΔIF. Here ΔI can be understood as an additional (negative) contribution to the binding energy, arising from all interactions in the system, where the ground state level shift is approximated by an interaction contribution to the atomic chemical potential, Δ1sΔμA that is discussed below. The additional term, ΔIF, is an energy shift due to the spin statistics of the unbound electrons, given by the difference of the chemical potentials of an ideal Fermi gas and an ideal classical gas. We will return to the issue of renormalization of the binding energy in Sec. IV B and to a discussion of the Saha equations in Sec. IV A 6.

To conclude this section, let us briefly discuss the approximations of the interaction contributions to the chemical potentials (for details, see Refs. 59, 61, and 191).

  • The contribution of the interaction between charged particles to the chemical potential is known analytically in the limiting cases of low (Debye–Hückel) and high densities (Thomas-Fermi). Efficient Padé formulas that are applicable for arbitrary densities and temperatures were constructed by Ebeling and coworkers,53 recently benchmarked129 and possibly improved by QMC results (see Sec. III E 1).

  • The (repulsive) interaction between neutrals is often treated within the hard-sphere model for which accurate formulas were derived by Carnahan and Starling307 and Mansoori et al.308 Improved approaches such as fluid variational theory (FVT, see Ref. 309) consider also attractive van-der-Waals-like contributions (see Refs. 62 and 63).

  • The interaction between charged and neutral particles is of particular importance in partially ionized plasmas. The respective contribution to the chemical potential can be described by calculating the second virial coefficient with respect to a screened polarization potential as derived by Redmer et al.310 

Detailed calculations for the thermodynamic properties of partially ionized hydrogen plasmas within the general scheme outlined above were performed by many groups using slightly different approximations for each of the three interaction parts. We would like to mention here the work of Däppen et al.56 on the solar atmosphere, of Saumon and Chabrier58,182 on a wide-range EOS of dense hydrogen plasma, and the FVT of Juranek et al.62,63 applied for the calculation of the EOS of hydrogen and, in particular, the Hugoniot curve.

The dissociation degree (fraction of atoms), as predicted by an advanced chemical model (FVT), is shown in Fig. 21 in in Sec. IV A 6).63 At low temperatures, good agreement with tight-binding molecular dynamics (TB-MD) simulations311 is achieved. With increasing temperature, deviations from QMC simulations are increasing, as is discussed in more detail in Sec. IV A 6. Also, the thermodynamic quantities of dense hydrogen calculated within FVT are in good agreement with quantum Monte Carlo results, at low temperature.63 Illustrations are shown for the equation of state in Fig. 14, a discussion is given in Sec. IV A 3.

3. Chemical models for dynamic properties: Chihara decomposition

The chemical picture concept to distinguish between bound and free electrons has also been extended to dynamical (frequency-dependent) quantities. In particular, this assumption leads to a decomposition of the total dynamic electron structure factor, See(q,ω), according to158,312–314
(56)
Here Z¯A is the average atomic number of the ion species α=β={} such that Ne=Z¯ANi. The first line describes the contribution of the bound electrons of various ion species represented by the form factor Fα and the contribution gα of the free electrons that screen the ionic charges in screening clouds. These electronic densities are convolved with the ion structure factors Sαβ and are weighted according to their concentrations xα=nα/βnβ. The second line contains the dynamic structure factor of the free electrons in the system, See0(q,ω), with Ne=ZfαNα+αZαNα. The third line describes transitions of electrons between free and bound states. Here, S̃αce denotes the correlations of the bound electrons of species α, whereas Sαs is the ionic self-structure factor.

Several methods exist to calculate the necessary input quantities for the Chihara formula (56). If the ion dynamics can be neglected, as is often the case [i.e., Sii(q,ω)=Sii(q)δ(ω)], then the ion structure can be obtained from HNC calculations,258,315 from DFT-MD,316 or even from PIMC simulations.148 The form factor F is from an isolated atom or ion and can either be approximated by hydrogenic wave functions or be calculated directly from DFT.317 The screening cloud, g, can be taken from linear response theory and may even include closed electron shell effects.318 The dynamic structure of the free electrons, See0, is described using the extended Mermin dielectric function featuring electron–electron local field corrections.319 The contributions of bound-free transitions and also free-bound transitions to the dynamic structure factor are usually obtained using the impulse approximation in an isolated atom picture in which the medium effects are inserted via effective ionization potentials, i.e., via atomic level shifts using ionization potential depression (IPD) models320–322 (see Sec. IV B).

Several improvements on the most basic models have been suggested and tested against experiments. DFT and DFT-MD methods can be applied to calculate, both the ion–ion and electron–ion structure, [F(q)+g(q)]=Sei(q)/Sii(q), so that the first line of Eq. (56) can be determined from quantum mechanical simulations alone.316 Similarly, calculations of the dynamic optical dielectric function, using the Kubo–Greenwood (KG) formula323,324 (cf. Sec. IV E 2), based on DFT Kohn–Sham orbitals can be used as improved collision frequency entering the Mermin dielectric function (131) for the free electron structure factor.159,325,326 The combination of all these concepts, including also the DFT determination of the bound-free part, allows one to almost overcome the Chihara picture. The full inelastic spectrum is then obtained using the KG approach based on DFT extended by a Mermin ansatz, whereas the elastic part can be directly sampled from DFT-MD.263,327,328 Recent efforts utilize the linear response time-dependent DFT (LR-TDDFT) for the inelastic part of the spectrum, together with the usual DFT-MD for the elastic part.263,328

DFT-based approaches such as Kohn–Sham (KS)-DFT and orbital-free (OF)-DFT have become the most commonly used first-principle simulation methods for WDM.137 This is due to a combination of often acceptable accuracy over the desired parameter space with a manageable computational cost.

1. Basic concepts

KS-DFT and OF-DFT treat the electronic component quantum mechanically while considering ions as classical particles. Both flavors of DFT rely on a one-to-one correspondence between the equilibrium electronic density n0(r) and the external potential, including the one of the ions, Vei(r). For the time-independent case, the proof of this correspondence is provided by the Hohenberg–Kohn theorems,329 for the ground state, which was extended to finite temperatures by Mermin.330,331 In the time-dependent case, Runge and Gross showed that a one-to-one mapping exists between the time-dependent single-particle density and the single-particle potential.332 Consequently, one can consider a system of non-interacting fermions in a local effective potential, vKS(r) (KS potential), that generates the same single-particle electronic density as the one of a real (interacting) system. Moreover, physical properties of electrons, such as free energy and pressure, can be computed as functionals of the electron density, n(r).

In the Kohn–Sham (KS) formulation of DFT, the electronic subsystem is described by the single-particle Schrödinger-type equations of auxiliary non-interacting fermions:
(57)
where the orbitals ψR¯j depend parametrically on the ionic positions R¯=(R1,,RNion) (Nion is the total number of ions). Equation (57) provides a framework for the simulation of the time evolution of the electronic density n(r,t)=jfj|ψR¯j(r,t)|2 in terms of the single-particle KS orbitals ψR¯j, with fj being the occupation numbers, according to the Fermi-Dirac distribution, and the sum is performed over all orbitals. In Eq. (57), the KS potential is defined as the sum of the classical electrostatic (Hartree) potential, vH(r), the exchange–correlation (XC) potential, vxc(r), and an external potential that contains the potential of the ions, vext(r), i.e., vKS(r)=vH(r)+vXC(r)+vext(r). The XC potential plays a key role in compensating for the differences between the auxiliary non-interacting fermions in the effective field and a real system. The exact form of vxc(r) is not known in advance and has to be approximated or provided by a higher-level quasi-exact method such as PIMC. In this way, as soon as the XC potential is defined, the KS approach substantially simplifies the computational task of describing quantum-many particle systems.
In practice, because of the large difference in masses of electrons and ions, an additional approximation that is often used for WDM simulations is the Born–Oppenheimer (BO) approximation,333 in which electrons are assumed to be in equilibrium in the field of a given configuration of ions. Therefore, the BO approximation reduces the KS-DFT to the time-independent situation in the absence of a time-dependent external potential. From Eq. (57) then follows
(58)
where ϕR¯j(r) results from the factorization ψR¯j(r,t)=ϕR¯j(r)exp(iεjt/), and for the density, we have n(r)=jfj|ϕR¯j(r)|2.
Since an equilibrium system is considered, the solution of Eq. (58) is combined with the minimization of the free energy of the system:
(59)
where Es,R¯kin[n] and Ss,R¯[n] are the kinetic energy and entropy of the auxiliary system of non-interacting fermions computed using KS orbitals and eigenenergies, respectively. Further, in Eq. (59), UR¯[n] is the Hartree mean-field potential energy, and VR¯ei[n] is the potential energy due to electron–ion interaction. The functional derivative of the XC functional, FR¯xc[n], with respect to density defines the XC potential, vxc(r), of the effective potential in KS-DFT. For WDM applications, we discuss the various approximations of the XC functional in Sec. III C 2.

Neglecting dynamic electronic polarization effects, the BO approximation allows one to drive the dynamics of ions using forces from the ground state (equilibrium) DFT density of electrons. The BO approximation is expected to be sufficiently accurate for most of the ensemble-averaged macroscopic properties, such as the EOS.

There are a number of packages available that solve Eq. (58) using a variety of numerical methods, libraries, and programming languages. For example, freely available open source codes are GPAW,334–337 Quantum-Espresso,338,339 ABINIT,340–345 ABACUS,346,347 CP2K,348,349 and SPARC.350,351 Commercial alternatives are for example given by VASP352–355 and CASTEP.356,357 Being massively parallel and rigorously tested, these codes with numerous postprocessing capabilities make KS-DFT a powerful tool for first principles simulation of quantum many-particle systems.

Within DFT, an alternative route is based on the direct minimization of the free energy [Eq. (59)] expressed as a functional of the density, without employing KS orbitals. This is the strategy of orbital-free DFT (OF-DFT). In addition to the approximation for the XC term, this requires an explicit formula for the non-interacting free energy, Fs[n]=Eskin[n]kBTSs[n], as a functional of the density. Since the exact form of this functional is not known, Fs[n] is approximated based on theoretical constraints and information about limiting cases, such as the UEG or an isolated hydrogen atom.358–362 While being less accurate than KS-DFT, on the other hand, OF-DFT has a much lower computational cost. This enables the simulation of tens of thousands of particles at elevated temperatures.363 For certain materials, the active development of various approximations for Fs[n] allows one to perform simulations with sufficient accuracy (e.g., for molecular dynamics) in both ambient364 and extreme conditions.365 Because of the high degree of ionization (i.e., weak electron–ion coupling) and reduced relevance of quantum effects (e.g., compared to the case with Θ1 for hydrogen), OF-DFT is particularly valuable in the limit of high temperatures with Θ1,363,366 e.g., for EOS calculations.367 In this way, OF-DFT is a valuable complement to KS-DFT. Note that there are several open source and freely available codes for OF-DFT-based MD simulations at high temperatures (see, e.g., Refs. 363, 365, and 368).

The main reason for the increase in the computational cost of KS-DFT is the need for a much larger number of orbitals as compared to the number of particles in the simulation. This is the result of the significant smearing of the occupation numbers beyond the Fermi energy when Θ1. To have accurate results at WDM conditions, in addition to other parameters, convergence with respect to the number of used orbitals (i.e., the smallest occupation number) is mandatory.369 For example, keeping the smallest occupation number at about 106, for 108 hydrogen atoms, it was shown using the highly efficient GPU implementation of VASP,370,371 that the computation cost scales as T2, at T10eV and rs=2.366 Such a quadratic scaling becomes intractable when the number of atoms increases to thousands. Therefore, new numerical methods are required to push the computational capabilities of KS-DFT to high temperatures. To meet the demands of WDM research, Zhang et al.372 introduced the so-called extended KS-DFT where high-energy orbitals are approximated using the analytical formula for the density of states of the free electron gas. The energy threshold Ec above which states are treated analytically is a convergence parameter in simulations using the extended KS-DFT. We note that the extended KS-DFT scheme is implemented in Abinit.373 Another similar hybrid approach was suggested by Hollebon and Sjostrom,374 where the density contribution due to orbitals above Ec is approximated using the Thomas-Fermi model. In addition, the most recent development in the adaptation of the KS-DFT to meet the needs of high-temperature applications is the stochastic formulation of the KS-DFT.375–379 

2. Exchange–correlation functionals

The accuracy of the KS-DFT results depends, to a large degree, on the form of the used XC functional. A local density approximation (LDA) is arguably the simplest form one can use for the XC functional FR¯xc[n] which, in the ground state, reduces to ER¯xc[n]. To simplify the notation, in the following, we omit the parametric dependence on the ionic positions and write FR¯xc[n]=Fxc[n].

For extended systems, the most successful XC-functionals on the LDA-level are based on the UEG model:
(60)
where fxcUEG[n(r)] is the XC-free energy per particle of the UEG at a given temperature and the local value of the density.

In practice, the quasi-exact QMC results for the UEG are parametrized (fitted) for KS-DFT applications. For example, widely used parametrizations of the ground state QMC results of Ceperley and Alder380 were provided by Perdew and Zunger381 and Vosko et al.382 

The inclusion of the corrections due to density inhomogeneity, in the first order, leads to the additional functional dependence on density gradients, n(r). The most often used and universal XC functionals on this second “rung” beyond the LDA are referred to as generalized gradient approximations (GGA). The GGA level functionals, such as PBE383 and PBEsol384 for T=0, are designed using various physical (e.g., the QMC data for the UEG limit) and mathematical constraints (e.g., Levy's uniform scaling condition385) The extension beyond the GGA leads to the meta-GGA rung of functionals, such as SCAN, where the functional dependencies on the Laplacian of the density, Δn(r), and on the kinetic energy density, τ=jfj|ϕj(r)|2, are added.386 Further advanced inclusion of the non-locality into the XC-functional results in higher rung classes, such as the hybrid-GGA HSE functional that features exact exchange.387 This hierarchy of the XC functionals is often referred to as Jacob's ladder, following Perdew and Schmidt.388 Numerous XC-functionals are available through the Libxc library of XC-functionals.389,390 Similar to the ground state applications, accurate QMC data for the UEG at finite temperatures allowed the development of the XC functionals on different rungs of Jacob's ladder for warm dense matter applications, as discussed in Sec. III C 3.

3. Finite-temperature effects in the functionals

Currently, the vast majority of DFT simulations of WDM use a ground state XC-functional without explicit temperature dependence, that is Fxc[n,T]Exc[n]—an approach known as the ground-state approximation (GSA). Ground-state functionals beyond the LDA rung take into account inhomogeneity effects, but completely miss thermal XC effects. The simplest LDA XC-free energy is based on the highly accurate QMC data for the uniform electron gas (UEG) at finite temperature of Dornheim et al.,128 and is presented by two equivalent parameterizations already mentioned above, the KSDT136 and GDSMFB128 functionals.

A ground-state GGA functional with an additive thermal correction at the LDA level was presented in Ref. 391. Kozlowski et al. exploited a similar idea presenting a ground-state PBE functional with a multiplicative LDA-level thermal correction.392 These two approximations take into account the inhomogeneity effects only at the ground-state level, i.e., depend on the ground-state reduced density gradient variables without explicit temperature dependence. Another drawback of these two functionals is that they are not consistent with the XC-finite-temperature gradient expansion—one of the known and important constraints for the XC-free energy.

Karasiev et al. developed a non-empirical fully thermal GGA-level XC-functional, KDT16,393 based on rigorous constraints with proper temperature-dependent reduced density gradient variables. The KDT16 XC functional reduces to the ground state PBE functional, in the zero-temperature limit, such that it can be used across the entire temperature range. Recently developed thermal functionals based on the de-orbitalized Laplacian-dependent SCAN and its regularized-restored (R2) version, TSCANL and TR2SCANL, respectively, are based on a universal thermal additive correction at the GGA level using a perturbative-like self-consistent approach.230 Semi-local functionals suffer from a fundamental drawback—underestimating the electronic bandgap. Hybrid XC functionals, such as the global PBE0394 and the range separated HSE387 are known to predict qualitatively correct bandgap values. The recently developed thermal KDT0 hybrid395 is based on a mixture of finite-temperature Hartree–Fock (HF) exchange and thermal KDT16 GGA XC. The KDT0 hybrid provides significant improvements to calculations of the entire band structure, at temperatures within the WDM regime, such that the accuracy of optical properties is improved as well upon accounting thermal XC effects at the hybrid (HF plus density functional approximation) level.396 In Sec. IV A 4, we test the relevance of temperature effects on the functionals for dense partially ionized hydrogen.

Like the DFT-MD models described in Secs. III A–III C, DFT-based average-atom models (DFT-AA) treat electrons quantum mechanically, determining an equilibrium electronic density that is self-consistent with an electron–ion potential. Like DFT-MD models, DFT-AA models can describe the electrons using either orbital-free approximations397,398 or Kohn–Sham orbitals399–401 and are sensitive to the choice of the exchange–correlation functional. Unlike DFT-MD models, however, DFT-AA models do not determine a full three-dimensional electronic density that varies with the positions of multiple ions; instead they model only a single ion, implicitly averaging both the electronic density ne(r) and the screened potential, Vei(r) into spherically symmetric quantities around a single ion center. The effects of changing ion density are incorporated through changes in the boundary conditions at the ionic Wigner–Seitz radius di [Eq. (2)]: typically the potential is forced to vanish at di (although other types of boundary conditions have been explored as well399,402).

DFT-AA models can also predict ionic properties, modeling correlations in the external environment by computing the self-consistent response of electrons in a cavity absent a central nuclear charge.403–405 This external potential can be used to define a neutral pseudo-atom (NPA) charge density and an ion–ion potential that determines the static ion–ion structure factor and radial ion distribution function.

Averaging over both electronic and ionic configurations makes all-electron DFT-AA models computationally highly efficient, over a very wide range of densities and up to very high temperatures. As a consequence, they provide electronic contributions to many modern equation of state tables, such as SESAME.93 However, averaging over ion ensembles and enforced spherical symmetry make DFT-AA models less reliable than DFT-MD, at low temperatures, where molecular and lattice structures become important.

Many extensions to basic DFT-AA models have extended their functionality beyond equations of state. The Kohn–Sham electronic orbitals can be used to compute transition matrix elements needed for static406–408 or dynamic409 conductivities and opacities, and for the collisions and dynamic structure factors relevant to x-ray Thomson scattering288,410,411 and stopping powers.288,406 Expansions of these orbitals into real, integer-occupied electronic configurations can provide detailed x-ray opacities.412,413 Examples of opacities from our DFT-AA model are shown in Fig. 44 in Sec. IV E 4.

Originally introduced for the simulation of low-temperature liquid 4He,75,76 the PIMC approach414–417 has emerged as one of the most powerful finite-temperature methods in statistical physics and quantum chemistry. PIMC simulations have been successfully applied to warm dense hydrogen, where a number of different variants have been developed which include restricted (fixed nodes) PIMC (RPIMC), fermionic (direct) PIMC (FPIMC), configuration PIMC (CPIMC), and coupled electron–ion Monte Carlo (CEIMC). In Fig. 8, we give an overview of the parameter range where the different finite temperature methods apply for dense hydrogen. In the limit of vanishing temperatures, variational and diffusion Monte Carlo methods directly address ground state properties.418 In this section, we focus on FPIMC (Secs. III E 1 and III E 5), RPIMC (Sec. III E 4), CPIMC (Sec. III E 3), and CEIMC (Sec. III F).

FIG. 8.

Applicability range of different simulation methods across the hydrogen phase diagram. RPIMC: restricted PIMC, extends to about 0.1TF (Sec. III E 4). CEIMC: coupled electron–ion Monte Carlo (Sec. III F), FPIMC: fermionic PIMC, extends to about 0.5TF and was applied to temperatures above 10 000 K (Sec. III E 5). ML denotes region fit by QMC-based machine learning force fields in Ref. 206 and also the region where CEIMC has been applied. Red pluses show FPIMC simulation points from Ref. 115 that mark the border of what is feasible today. Light black crosses show regions covered by the RPIMC database.114 The other lines are introduced in Fig. 4.

FIG. 8.

Applicability range of different simulation methods across the hydrogen phase diagram. RPIMC: restricted PIMC, extends to about 0.1TF (Sec. III E 4). CEIMC: coupled electron–ion Monte Carlo (Sec. III F), FPIMC: fermionic PIMC, extends to about 0.5TF and was applied to temperatures above 10 000 K (Sec. III E 5). ML denotes region fit by QMC-based machine learning force fields in Ref. 206 and also the region where CEIMC has been applied. Red pluses show FPIMC simulation points from Ref. 115 that mark the border of what is feasible today. Light black crosses show regions covered by the RPIMC database.114 The other lines are introduced in Fig. 4.

Close modal

Before discussing applications to hydrogen, we introduce the basic idea behind the PIMC method for the Uniform Electron Gas (UEG) model116,117—the archetypical model system of interacting quantum electrons419—in which the ions are replaced by a homogeneous neutralizing positive background. Subsequently, we will generalize the PIMC approach to actual electron–ion systems, such as hydrogen, and discuss how to avoid path collapse due to the singular Coulomb attraction in an efficient way.420 Finally, we will touch upon a variety of possible concepts to avoid the fermion sign problem,79,80 which otherwise renders direct PIMC simulations of hydrogen computationally unfeasible, for degenerate electrons.

1. PIMC simulation of the uniform electron gas

The UEG constitutes the quantum mechanical analog of the classical one-component plasma (OCP). Its Hamiltonian simplifies Eq. (35) and is given by (using Hartree atomic units)117 
(61)
where WE(r̂l,r̂k) is the Ewald pair potential that (a) takes into account the interaction also with the infinite periodic array of images and (b) with the neutralizing uniform positive background. The final term in Eq. (61) is given by the Madelung constant that accounts for the self-interaction between a charge and its own array of images. A detailed discussion of the Ewald potential has been presented by Fraser et al.421 
To derive the PIMC approach, we consider a system of N spin-unpolarized (i.e., N=N=N/2) electrons in the canonical ensemble where the volume Ω=L3, number density n=N/Ω and inverse temperature β=1/kBT are fixed. The partition function in coordinate space then reads
(62)
where R=(r1,,rN)T represents the coordinates of both spin-up and spin-down electrons. Note that we have to explicitly take into account all possible permutation elements σi of the respective permutation group SNi (with π̂σNi being the corresponding permutation operator, i=,). Evidently, the partition function is given by a sum over contributions that change their sign as a function of the number of pair permutations Npp. This is the root cause of the fermion sign problem80 that is discussed in more detail below.
At this point, the main practical problem is the evaluation of the matrix elements of the density operator ρ̂=eβĤ since the kinetic and potential contributions to the full Hamiltonian, Ĥ=K̂+V̂, do not commute. To overcome this obstacle, one can make use of the exact semi-group property of the density operator
(63)
where the reduced inverse temperature is defined as ε=β/P. Inserting P1 unity operators of the form 1̂=dRα|RαRα|, then leads to the familiar expression117,
(64)
with R0=RP. The original Eq. (62) has thus been re-cast as an integral over P sets of particle coordinates where the corresponding density matrix has to be evaluated at P times the original temperature. For a sufficiently large P, one may introduce a suitable high-temperature approximation, such as the primitive approximation eεĤeεK̂eεV̂; convergence of the latter is ensured by the well-known Trotter formula422,
(65)
It is important to note that Eq. (65) only holds for operators that are bounded from below.423 This is, however, not the case for hydrogen, due to the diverging Coulomb attraction between electrons and ions, which will require special care (cf. Sec. III E 5). In addition, we note that more accurate and efficient factorization schemes have been discussed in the literature,416,424,425 which is of particular interest for the PIMC simulation of fermions,115,122,124,426 as we elaborate on in Sec. III E 4.
Finally, the partition function can be expressed as
(66)
i.e., as an integral over the high-dimensional meta-variable X=(R0,,RP1)T where the symbolic notation dX contains all integrals as well as the summation over all permutations. Each particular X can be interpreted as a path configuration through the imaginary time, t=iτ, with τ[0,β], where each particle is represented on each of the P discrete imaginary “time slices”. Further, each path contributes proportionally to its weight, W(X), which is a function that can be readily evaluated in practice.

2. The fermion sign problem (FSP)

Equation (66) is a 3PN-dimensional integral, with NO(10102) and PO(102), and further contains the summation over N! permutations. Its numerical evaluation using standard quadrature methods is thus unfeasible in practice. However, a stochastic estimation based on the Metropolis algorithm427 overcomes this bottleneck, as its efficiency depends only weakly on the dimensionality.428 The task at hand is thus to generate a Markov chain of configurations X, which are distributed according to the probability P(X)=W(X)/ZN,Ω,β. Indeed, path sampling schemes279,414,417,429 allow for quasi-exact simulations of NO(103104) bosons (such as superfluid 4He414,430,431) or boltzmannons (i.e., hypothetical distinguishable quantum particles, e.g., Refs. 432 and 433), giving unprecedented insights into important phenomena such as superfluidity.414,431,434 For fermions, on the other hand, the (1)Npp term leads to both positive and negative weights W(X), thus preventing the interpretation of P(X) as a probability distribution.

Formally, this obstacle is easily avoided by switching to a modified configuration space that is defined by
(67)
The expectation value of an arbitrary observable  is then computed as
(68)
with S(X)=W(X)/W(X) and
(69)
where A(X) is the estimator of A in the path-integral representation. The denominator in Eq. (68) the average sign SS80 and measures the cancelation between positive and negative contributions to the fermionic partition function.
The relative error of a Monte Carlo estimate for Â, assuming NMC statistically independent samples, is then given by435,
(70)
Here σA is the intrinsic variance of the estimator A(X), whereas f and f are the free energy densities of ZN,Ω,β and ZN,Ω,β, with ff. Equation (70) thus directly implies that the statistical error of a fermionic PIMC simulation increases exponentially with increasing system size and decreasing temperature (or, equivalently, increasing β). This “exponential wall” can only be compensated by increasing the number of Monte Carlo samples as 1/NMC, which quickly becomes computationally unfeasible. This is the fermion sign problem,80 that prevents the direct application of the PIMC method over substantial parts of the relevant WDM parameter space (see also Ref. 129 for a recent overview article). This unsatisfactory situation has sparked a remarkable surge of developments in the field of fermionic QMC simulations over the last decade (e.g., Refs. 115, 120–123, 125, 126, 298, and 436–446); a selection of methods that are particularly relevant for the simulation of hydrogen is shown in Secs. III E 4 and III E 6. Another idea is to develop PIMC simulations in different quantum mechanical representations where the FSP may appear in different parameter regions. This concept is discussed in Sec. III E 3.

3. Configuration PIMC (CPIMC) simulation of the UEG

The physical origin of the fermion sign problem in FPIMC simulations are quantum exchange effects of fermions. The path integral concept maps the simulation of quantum systems at a temperature T on that of classical systems at a P times higher temperature [cf. Eq. (63)]. Thus, it is not surprising that fermionic PIMC simulations become increasingly difficult with increasing quantum degeneracy, i.e., for increasing χ or decreasing Θ, as indicated in Fig. 8. On the other hand, for the limit of strong degeneracy, i.e., low Θ, or that of a nearly ideal Fermi gas, i.e., rs1, well-known concepts exist. Here we will not consider the low-temperature case where ground state methods such as diffusion or variational Monte Carlo are appropriate (see Fig. 8), but concentrate on finite temperatures and low rs-values. In fact, for rs0, the ideal Fermi gas limit is recovered where N-particle states are given by Slater determinants of one-particle orbitals. The N-particle density operator ρ̂, as discussed in Sec. III E 1, can now be transformed into a new representation, in addition to the previously considered coordinate representation that we reproduce in the first line, we sketch the idea of configuration PIMC (CPIMC), developed by Bonitz et al., in the second quantization, as118,
(71)
In the coordinate representation, one uses N-particle coordinate states |R, which are (non-antisymmetrized) product states of one particle states, |ri, with i=1,,N, and the antisymmetrization being implemented via a sum over all different permutations, σ and σ, giving rise to sign-alternating terms that cause the FSP. In CPIMC (the second line), one proceeds differently: no antisymmetrization of the density operator is performed but, instead, the matrix elements are computed with antisymmetric N-particle states, |{n}=|n1n2, where nk are the occupation numbers of all single-particle orbitals (occupation number representation). While fermionic PIMC in coordinate representation becomes trivial in the classical limit (large rs), CPIMC becomes trivial for V̂0, i.e., for an ideal Fermi gas. In both cases, the density matrix is diagonal in the respective representation. Departure from these limits leads to an increasing fermion sign problem: for coordinate space simulations, if rs is lowered, and for CPIMC, if rs is increased. The behavior of CPIMC has been studied in detail for various systems (see Refs. 118, 120, 121, 125, and 126) where these general considerations are confirmed. Further results are available for the static structure factor143 and for the momentum distribution function447 (for overviews, see Refs. 117 and 119).

Results for the thermodynamic properties of the uniform electron gas (the single-particle orbitals generating the Slater determinants are plane waves) were presented in Refs. 120 and 121, and we show the exchange–correlation energy of the UEG in Fig. 9. While coordinate space simulations (RPIMC, see Sec. III E 4, and PB-PIMC,122 see Sec. III E 6) are hampered by the FSP, when approaching rs=1 from above, CPIMC easily fills the gap at small rs-values. At the same time, CPIMC is afflicted by a FSP, when rs approaches 0.51, from below.

FIG. 9.

Exchange–correlation energy of N=33 spin-polarized electrons in jellium for two dimensionless temperatures: Θ=2 and Θ=0.5. First principle CPIMC and PB-PIMC data can be connected exactly. RPIMC data from Ref. 296, DMQMC from Ref. 439. In the limit rs0, the Hartree–Fock result is approached and the curves should be horizontal, as observed in CPIMC and DMQMC. Θ=0.5 is the lowest temperature for which CPIMC and PB-PIMC can be smoothly connected, as PB-PIMC is afflicted with increasing errors at rs2. For Θ=0.5, all data have been shifted by 0.05 Ha. Reprinted from Ref. 129 with permission of AIP Publishing.

FIG. 9.

Exchange–correlation energy of N=33 spin-polarized electrons in jellium for two dimensionless temperatures: Θ=2 and Θ=0.5. First principle CPIMC and PB-PIMC data can be connected exactly. RPIMC data from Ref. 296, DMQMC from Ref. 439. In the limit rs0, the Hartree–Fock result is approached and the curves should be horizontal, as observed in CPIMC and DMQMC. Θ=0.5 is the lowest temperature for which CPIMC and PB-PIMC can be smoothly connected, as PB-PIMC is afflicted with increasing errors at rs2. For Θ=0.5, all data have been shifted by 0.05 Ha. Reprinted from Ref. 129 with permission of AIP Publishing.

Close modal

Thus, an interesting complementarity of the two representations is observed, and a combination of the two allows one to achieve the first principles PIMC results, for all densities, thereby effectively avoiding the sign problem, as is demonstrated in Fig. 9. A combination of both approaches without gaps is possible for temperatures Θ0.5, for N=33 particles. The figure also contains a comparison to RPIMC simulations by E. W. Brown et al.296,448 which became increasingly inaccurate when rs was reduced toward unity. Finally, we also included in Fig. 9 data from density matrix QMC (DMQMC). This approach is similar to CPIMC and was developed by Blunt et al.,437 as a finite-temperature extension of full-CI-QMC (FCIQMC).449,450 DMQMC was subsequently applied to the UEG by Malone et al.,438,439 their data are shown by the green diamonds.

To conclude this section, we have shown that a combination of two fermionic PIMC methods in complementary representations is a very promising approach to avoid the FSP, not only for jellium but also for hydrogen. However, so far no CPIMC results for hydrogen are available yet. Other strategies to alleviate the fermion sign problem will be discussed in Secs. III E 4 and III E 6.

4. The restricted path integral Monte Carlo method (RPIMC)

As discussed above, the direct PIMC method requires sampling both paths and permutations of paths. For fermions, this requires a minus sign whenever an odd permutation is encountered as Feynman and Hibbs noted in 1965.451 As explained in Ref. 435, this leads to an exponential increase in the computational effort [cf. Eq. (70)].

In ground-state QMC methods, the fixed-node (FN) method has proved useful to get beyond this limitation.380,452 Here one posits a good many-body trial function (typically from a mean field calculations) and constructs a solution with the same sign. A stochastic process can be used to find the optimal magnitude. If the nodes are correct, one obtains the exact energy. Otherwise, the computed energy lies below the variational Monte Carlo energy but above the exact energy. This rigorous upper bound allows one to judge the relative accuracy of the nodal surfaces even in the absence of experimental data or other methods. If one could completely parameterize nodal surfaces, one could use this to arrive at the exact surface. The FN approximation is particularly accurate for hydrogenic systems since, in the molecular phase, the nodes are between the molecules and have less influence on the computed properties; in the atomic phase, the wave function is weakly correlated so nodes from DFT orbitals are accurate.

It is possible to generalize the FN method to non-zero temperature using the restricted path integral method82 (RPIMC). One posits a many-body fermion density matrix ρT(R,R;τ), for all 0τβ. Then one can show that the exact density matrix satisfies the identity
(72)
where the paths go from PR*Rβ without crossing a node. That is,
(73)
The many-body position that sets the nodes, R*, is called the reference point. The contribution of the paths W(X) is the same as in Eq. (66). This is an identity as long as the sign of the trial density matrix is correct. To obtain observables such as the internal energies, and the pair correlation functions or the momentum distributions, one samples the reference point R* and the permutation, P, over their allowed values. For observables diagonal in coordinate space (e.g., the partition function and the electron density), Rβ=R*, and only even permutations contribute, all contributions are non-negative and can be sampled without a sign problem. When off diagonal elements of the density matrix are sampled to compute, e.g., the momentum distribution, negative contributions still enter the computed averaged, even though all paths obey the nodal restriction in Eq. (73).83 
There remain two main challenges. First, the exact fermionic density matrix is only known in a few cases, e.g., for noninteracting particles. So one is required to employ a trial density matrix, which introduces an uncontrolled approximation into the calculations. For temperatures at or above the degeneracy temperature, one can show that the nodes approach exponentially (in T) those of the free particle density matrix,
(74)

On the other hand, in the low-temperature limit, if the ground state is unique (non-degenerate), the fermion density matrix will factor into the product ρ(R,R;β)Ψ*(R)Ψ(R), and RPIMC reduces to the ground state FN method. To determine the nodal structure between these two limits, two approximations have been derived. Variational nodes have been developed87 and then applied to dense hydrogen.94,189,453 Furthermore, free-particle nodes and bound states around the nuclei have been combined into one trial density matrix to extend the applicability range of RPIMC simulations to lower temperatures.454 

The second challenge is to sample the space of path coordinates and permutations efficiently with Monte Carlo methods in the presence of the nodal restriction. This poses a different challenge than in the bosonic case because the restriction singles out the reference point as being special and the nodal restriction places a non-local restriction on Monte Carlo updates. As the temperature is lowered the paths and the reference point become difficult to move. It is not known if this is a fundamental problem. See Refs. 82 and 435 for details of the RPIMC method.

Beginning with Ref. 84 in 1994, the RPIMC method has been applied to study hydrogen numerous times, which will be discussed in Sec. III E 5. Starting with helium,455,456 RPIMC with free-particle nodes has been applied to study many heavier elements up to neon.290,436,457–460 The inclusion of bound states into the nodal structure allowed one to simulate elements and compounds as heavy as silicon.461–464 For each material, the RPIMC results were combined with predictions from density functional molecular dynamics simulations to obtain one consistent equation of state (EOS) table that covers a wide range of density-temperature conditions so that shock Hugoniot curves can be inferred. These EOS tables were combined into one first principles EOS (FPEOS) database.465,466 RPIMC has also been used to construct an EOS for the uniform electron gas.136,296,448

5. PIMC simulations of hydrogen

The main difference between PIMC simulations of the UEG (cf. Sec. III E 1) and hydrogen is given by the Coulomb attraction between electrons and protons. Since the latter is not bounded from below, it prevents the direct utilization of the primitive approximation in the Trotter formula [Eq. (65)]. To overcome this obstacle, one can replace the primitive approximation with the pair action as described in414,420,467,
(75)
where the kinetic part has been omitted for simplicity. Here ϕ is the logarithm of the exact density matrix for two charged particles: electron–electron or electron–proton. A further simplification has been proposed by Kelbg,468,469 based on a first-order perturbation expansion of Eq. (75) with respect to the coupling strength. This result was extended to strong coupling in Ref. 470, leading to the “improved Kelbg potential” (IKP) (see Sec. III G 2). Although the resulting effective quantum pair potential471 is easy to implement and avoids the path collapse in PIMC simulations, due to the divergence of the bare Coulomb attraction, its convergence is substantially slower compared to the full pair approximation,140–142,471 making the latter the preferred approach for PIMC simulations. For completeness, we note that such effective potentials are still being used in approximate dynamic simulations of electron–ion systems (see, e.g., Refs. 472 and 473 and Sec. IV D 6).

A second difference between PIMC simulations of the UEG and hydrogen is the impact of fermionic exchange effects and their manifestation with respect to the fermion sign problem. In Fig. 10, we show the average sign from direct PIMC simulations at the electronic Fermi temperature Θ=1 for rs=2 (ρ= 0.34 g/cm3, T= 12.53 eV) and rs=3.23ρ= 0.08g/cm3, T= 4.80 eV) as a function of the number of electrons N. For all depicted datasets, we find an exponential decrease in S with increasing system size [cf. Eq. (70)]. For the UEG model (blue diamonds and yellow squares), decreasing the density (at constant degeneracy) leads to a less severe sign problem as the formation of permutation cycles474 is effectively suppressed by the stronger Coulomb repulsion between the electrons. Interestingly, the opposite trend has been reported for hydrogen.149 First, the sign problem is overall more severe compared to the UEG under the same conditions. This is a direct consequence of the increased degree of inhomogeneity due to the presence of the protons around which the electrons tend to cluster even, if they are effectively unbound. Second, we find a reduced sign for hydrogen at the lower density, as the inhomogeneity induced formation of permutation cycles of electrons located around the protons predominates over the separation of electrons due to the Coulomb repulsion observed for the UEG. We note that the Pauli principle ensures the stability of a system of protons and electrons when the temperature is lowered, whereas a corresponding system of hypothetical distinguishable quantum particles would collapse.475 

FIG. 10.

Average sign S as a function of the number of electrons N at the electronic Fermi temperature Θ=1, for rs=2 (ρ= 0.34 g/cm3, T= 12.53 eV) and rs=3.23ρ= 0.08g/cm3, T= 4.80 eV). The dashed black lines show exponential fits to the PIMC data [cf. Eq. (70)]. Reproduced from Ref. 149 with the permission of the authors.

FIG. 10.

Average sign S as a function of the number of electrons N at the electronic Fermi temperature Θ=1, for rs=2 (ρ= 0.34 g/cm3, T= 12.53 eV) and rs=3.23ρ= 0.08g/cm3, T= 4.80 eV). The dashed black lines show exponential fits to the PIMC data [cf. Eq. (70)]. Reproduced from Ref. 149 with the permission of the authors.

Close modal

To deal with the more pronounced sign problem in PIMC simulations of hydrogen, three strategies have been successfully employed: (i) Militzer, Ceperley and others87 have used the restricted PIMC method (see Sec. III E 4); (ii) Filinov and Bonitz115 have employed antisymmetric imaginary-time propagators (i.e., determinants). By grouping together positive and negative terms in the determinants, the sign problem is substantially alleviated,122,123 allowing for simulations at lower temperature compared to direct PIMC; and (iii) the ξ-extrapolation method introduced by Xiong and Xiong,476 which removes the exponential scaling with respect to the system size at moderate temperatures.148,149,163,445,446 These concepts are discussed in more detail in Sec. III E 6.

An additional difference between hydrogen and the UEG arises for either the CEIMC method, described below (Sec. III F), or in the RPIMC method (Sec. III E 4), because the trial density matrix or electron trial function, has an important dependence on the protonic configuration. Accurate antisymmetric trial functions may give rise to additional computational effort.

6. Concepts to alleviate the Fermion sign problem

Over the past few years, a number of concepts to alleviate the fermion sign problem have been suggested. Here, we focus on two ideas that have recently been used for the PIMC simulation of warm dense hydrogen, as discussed in Sec. III E 5: (i) the utilization of antisymmetrized imaginary-time propagators115 and (ii) the controlled extrapolation over the continuous spin-variable ξ.148,149

As mentioned in Sec. III E 2, the sign problem comes from the cancelation of positive and negative terms due to the (1)Npp term in Eq. (64). Using the idempotency property of the antisymmetrization operator and the definition of the determinant, one can rewrite the partition function as78,117,122,416,426,477
(76)
where V(Rα) includes all potential energy contributions from time slice α and ρ0() is the noninteracting kinetic density matrix. Comparing Eqs. (64) and (76), it becomes evident that a large number of positive and negative terms have been grouped together in the determinants. While the determinants can still be both positive and negative, this can lead to a considerable reduction of the sign problem since at least parts of the cancelation are carried out analytically.

The basic idea behind this strategy is illustrated in Fig. 11, where we show a configuration of N=2 identical fermions (e.g., electrons) in the τ-x-plane. Within the direct PIMC approach, one would either consider the diagonal (red) connections that are associated with a positive sign of the configuration weight, or the off diagonal (blue) connections corresponding to a pair exchange with a negative sign. By using the determinants in Eq. (76), one can evaluate both terms at the same time, which reduces the amount of cancelations in the average sign S. In practice, this strategy only works when the diagonal and off diagonal contributions have a comparable weight; this is the case when the distance between two beads on the same time slice is comparable to the associated thermal wavelength λε=2πε (see the black arrow in Fig. 11). Since λε decreases with increasing number of time slices P, the effect of the determinants vanishes, and one recovers the full sign problem of the direct PIMC method for P.124 

FIG. 11.

Schematic illustration of the grouping of diagonal (positive) and off diagonal (negative) terms in an antisymmetrized imaginary-time propagator. Using the fourth-order propagator from Ref. 478, every imaginary-time step ε is further divided into three sub-intervals of unequal length. Reprinted from Dornheim, Groth, and Bonitz, Phys. Rep. 744, 1–86 (2018), Copyright 2018, with permission from Elsevier.

FIG. 11.

Schematic illustration of the grouping of diagonal (positive) and off diagonal (negative) terms in an antisymmetrized imaginary-time propagator. Using the fourth-order propagator from Ref. 478, every imaginary-time step ε is further divided into three sub-intervals of unequal length. Reprinted from Dornheim, Groth, and Bonitz, Phys. Rep. 744, 1–86 (2018), Copyright 2018, with permission from Elsevier.

Close modal

To avoid this conundrum, it has been suggested122,426 to combine the determinants with a higher-order factorization of the density operator that allows for sufficient accuracy with a small number of high-temperature factors P. This is the basic idea of the permutation blocking PIMC (PB-PIMC) method,117,122–124,129,144 which has recently been extended to the grand canonical ensemble139 and applied to the simulation of warm dense hydrogen by Filinov and Bonitz.115 

A different approach has recently been introduced by Xiong and Xiong,476 who have suggested to replace the factor (1)Npp in Eq. (64) by ξNpp, where ξ[1,1] is a continuous variable. The physically meaningful cases of Bose–Einstein, Maxwell–Boltzmann, and Fermi–Dirac statistics are recovered for ξ=1, ξ=0, and ξ=1, respectively. The basic strategy to avoid the sign problem is then to carry out simulations in the sign-problem free domain of ξ[0,1] and subsequently extrapolate to the fermionic limit based on the empirical relation476,
(77)
where a0, a1, and a2 are the free fit parameters. Having originally been introduced for path integral MD, this idea was applied to PIMC simulations of electrons in a harmonic trap and the warm dense UEG in Ref. 445. In practice, Eq. (77) works remarkably well for weak to moderate levels of quantum degeneracy, where it is capable of delivering quasi-exact results with a typical relative error bar of 0.1%. However, the idea breaks down for low temperatures, when the bosonic and fermionic systems become too dissimilar.445 In practice, the applicability of the extrapolation can be checked for small systems with N10 electrons where exact fermionic PIMC simulations (i.e., for ξ=1) are still feasible. Its reliability for larger systems is then ensured by the local nature of fermionic exchange effects at moderate to high temperatures.479 As a result, one gets a non-empirical PIMC approach for the simulation of fermions without the exponential bottleneck with respect to the system size. This has recently allowed for simulations of the UEG with up to N=1000 electrons.446 

Subsequently, the ξ-extrapolation method has been successfully applied to study the static density response,148 imaginary-time structure, and structural properties149 of warm dense hydrogen; for completeness, there are further results for strongly compressed beryllium in Ref. 163. As an example, we show results for the Electron–electron static structure factor, See(q), for N=14 hydrogen atoms at rs=2 and Θ=1 in Fig. 12. The green crosses show exact fermionic PIMC results, and the red circles have been obtained by extrapolating from the sign-problem free sector 0ξ1 (shaded gray area). The two datasets are in excellent agreement over the entire q-range, which demonstrates the reliability of the ξ-extrapolation for these parameters.

FIG. 12.

Electron–electron static structure factor of warm dense hydrogen for N=14 atoms at rs=2 (ρ= 0.34g/cm3) and Θ=1(T=12.53eV). Green crosses: exact direct fermionic PIMC results (i.e., ξ=1); red circles: extrapolation from the sign-problem free domain of ξ[0,1] via Eq. (77); gray area: PIMC results for 0ξ1.Reproduced from Ref. 149 with the permission of the authors.

FIG. 12.

Electron–electron static structure factor of warm dense hydrogen for N=14 atoms at rs=2 (ρ= 0.34g/cm3) and Θ=1(T=12.53eV). Green crosses: exact direct fermionic PIMC results (i.e., ξ=1); red circles: extrapolation from the sign-problem free domain of ξ[0,1] via Eq. (77); gray area: PIMC results for 0ξ1.Reproduced from Ref. 149 with the permission of the authors.

Close modal

Let us conclude by comparing a few strengths and weaknesses of the RPIMC, permutation blocking PIMC, and ξ-extrapolation approaches. The RPIMC method is available over the broadest range of parameters (in particular with respect to low temperatures) (cf. Fig. 8). Moreover, it works for a variety of materials,465 including second-row elements.454 These strengths come at the cost of an approximation that is very difficult to check in practice. Additionally, the straightforward access of direct PIMC to different ITCFs is lost. The PB-PIMC method, on the other hand, provides a higher degree of certainty, as the convergence with the number of high-temperature factors P can, in principle, be checked, even though this can be difficult in practice. Moreover, PB-PIMC does provide access to all ITCFs, but the imaginary-time grid on which these can be resolved is often very small, due to the requirement of a small P for the alleviation of the sign problem. Finally, the ξ-extrapolation method appears to be as limited with respect to the temperature as the direct PIMC method (although new concepts are continually being developed480) i.e., to a region of the phase space where the impact of Fermi statistics is not too profound. Instead, its key strength is the removal of the exponential scaling with the number of electrons at parameters where PIMC simulations are generally possible, at least for small systems. At the same time, it does not require any external input such as the nodal structure in RPIMC, and it gives one full access to all ITCFs on an arbitrarily dense τ-grid.

Finally, we mention recent attempts of Filinov et al. to develop alternative representations of the density matrix that would reduce the FSP (see Ref. 481 and references therein), however; accuracy tests against available benchmark data are missing so far.

Within the adiabatic approximation, solutions to the full Schrödinger equation of the coupled electron–ion system are expanded in exact eigenfunctions of the electronic Hamiltonian for fixed (given) nuclear positions. The coupling of different adiabatic electronic eigenstates only occurs via the ionic kinetic energy operator and is suppressed by the large mass ratio, M=mI/me1. Within the Born–Oppenheimer approximation (BOA),333 the wave function of the coupled electron–ion system factorizes into an electronic and nuclear part which can be regarded as the starting point of an expansion in powers of (1/M)1/4 for the total energy.482 Far below the electronic Fermi temperature, electrons are frozen in the adiabatic ground state, and the nuclear distribution can be obtained from its Born–Oppenheimer energy surface, e.g., the electronic ground state energy for a static nuclear configuration. By employing the BOA, we have completely decoupled the electronic problem. Calculations of the BO potential of the nuclei can now be done by any suitable method, independently from sampling of the nuclear degrees of freedom. A frequent choice is the use of DFT to determine electronic energies and forces within classical molecular dynamics simulations of the nuclei (BOMD).483 

Within coupled electron–ion Monte Carlo (CEIMC)150 the electronic BO energy is determined via ground state Monte Carlo methods, e.g., variational (VMC) or reptation (RMC) Monte Carlo, whereas the nuclear configurations at finite temperature are sampled either classically, according to their Boltzmann weight, or quantum mechanically, using path-integral Monte Carlo methods. This decoupling of electronic and nuclear degrees of freedom enables QMC simulation for temperatures below those applicable to RPIMC or FPIMC calculations (as shown in the figure). However, since VMC or RMC calculations of the electronic energies have stochastic errors, to guarantee unbiased Monte Carlo sampling of the nuclear configurations the nuclear Monte Carlo process uses the penalty method.484 A detailed description of CEIMC can be found in Refs. 151 and 156.

The accuracy of CEIMC depends on the choice of the trial wave function underlying the electronic QMC calculations. Typically one uses a Slater–Jastrow wave function augmented by backflow and three-body correlations485 with orbitals in the Slater determinant taken from DFT.486 Any residual dependence of QMC energies on parameters of the trial wave functions as well as the choice of the DFT functional used for the orbitals can be quantified and controlled by minimizing the QMC energies. The possibility to quantify the accuracy of different QMC calculations based on the variational principle represents one of the main advantages compared to using DFT energies where one usually relies on experimental input to justify the choice of the functional.

Diffusion (DMC)73 or reptation Monte Carlo (RMC)487 stochastically improves any trial wave function via imaginary time propagation. Similar to PIMC described above, direct sampling suffers from a strong sign problem which can be circumvented by the use of the fixed-node (FN) approximation. Assuming the BOA, the quality of the trial wave function and its residual fixed node error can be studied separately on snapshots of nuclear configurations; this greatly simplifies the search for accurate electronic wave functions, e.g., compared to improving the nodal surface underlying RPIMC. In practice, the influence of the nodal surface can be studied by the influence of the underlying DFT orbitals,152 optimizing localized atomic orbitals,197 multi-determinant wave functions,488 or backflow transformations.485 

However, the simplifications occurring in the BOA, due to the electronic description at zero temperature, come with one major drawback, since electronic correlations are no longer limited by the thermal wavelength; electronic coherence may extend over the whole simulation cell. The resulting sensitivity to boundary conditions can result in an important sensitivity to the size of the supercell; in other words to the number of particles in the simulation. Those finite-size errors can be drastically reduced by employing twist averaged boundary conditions293,489 in the QMC calculations of the BO energy. Leading order finite-size effects can be estimated directly from the properties of the underlying trial wave function291,293 without the need for extrapolations based on simulating different system sizes. This is particularly important for CEIMC where numerical extrapolations are computationally expensive or impossible.

Since finite size considerations, based on the analytical structure of the wave functions,293 guarantee size consistency, more recently developed iterative and neural backflow wave functions490–496 as well as FCIQMC449,450 and coupled cluster methods497 offer the possibility to systematically study and reduce the fixed-node error for systems with a small number of electrons. So far, most of these studies have been done on the homogeneous electron gas.

Beyond structural properties of the nuclei, including quantum and thermal motion beyond the harmonic approximation, CEIMC also provides important insights into electronic properties. The many-body BO wave function of the electrons yields direct access to off diagonal matrix elements of the reduced (single-electron) density matrix which encodes the localized (insulating) or extended (Fermi-liquid) behavior of the electrons197,498,499 (cf. Sec. IV C 2). Further, electronic excitation gaps can be determined within QMC accuracy, by variations of the electronic chemical potential within the grand-canonical ensemble500 (cf. Sec. IV C 3).

Within CEIMC, excitation gaps including nuclear quantum and thermal effects can be consistently determined.218,501 The closure of the fundamental gap pinpoints the transition to metallic hydrogen. In solid hydrogen, the character of the electronic wavefunction at different chemical potential contains information on the band structure of added or removed electrons502 and indicates the closure of the indirect gap at 330–380 GPa into a bad metal phase (blue shaded area in Fig. 5) with a direct gap closing at higher pressures ∼450–500 GPa vertical dark red shaded area in Fig. 5).218 Concerning the LLPT, CEIMC found an abrupt closure of the gap at the molecular-atomic transition (see Figs. 5 and 6) whereas a crossover between insulating and metallic liquid is observed at around 3000 K, above the critical temperature.499 

Due to the BO decoupling of electronic and nuclear degrees of freedom, CEIMC naturally connects with many electronic structure methods used in materials science. On one hand, it provides QMC based results of fundamental quantities which can be used to validate DFT based methods as well as GW-Bethe–Salpeter equation (BSE) approaches where experimental results are missing or difficult to interpret, as is the case in high pressure hydrogen. On the other hand, it allows one to investigate approximations such as BO, by comparison with PIMC or zero temperature DMC for conditions of temperature and pressure where several methods are applicable.

Sampling classical nuclear degrees of freedom by Langevin dynamics based on QMC forces within the BO approximation has been developed in Refs. 503–506. Comparable results to CEIMC are found,152 once finite size and basis set errors of previous results504–506 are accounted for.

There exist a variety of concepts to extend classical molecular dynamics to quantum systems. This includes Wigner function QMD,507,508 path integral molecular dynamics (PIMD),509,510 or wave packet molecular dynamics (WPMD).511 In addition, many attempts were made to employ classical MD and account for quantum and spin effects by proper modification of the pair interaction (see Sec. III G 2).

1. Wave packet molecular dynamics (WPMD)

Wave packet molecular dynamics was originally proposed by Heller511 after the observation that a Gaussian wave function remains a Gaussian in a quadratic potential, and its time evolution follows classical equations. More generally, equations of motion can be derived via the time-dependent variational principle by considering variations of the action,277 
(78)
where the state |Q=|Q(Qμ) is parametrized by a complex set of parameters Qμ. The equations of motion for the parameters readily follow
(79)
with the hamiltonian function H=Q|Ĥ|Q/Q|Q and the norm matrix
(80)
where we have used the convention |Q=|Q(Qμ) and Q|=Q(Qμ*)|.

Commonly, a restricted state |Q that does not span the full Hilbert space is used to aid computational performance and model electron and ion dynamics simultaneously. However, such an approximation limits the dynamics to a sub-manifold, and the approximation is closely related to the McLachlan and Dirac–Frenkel variational principles.512 The actual choice of the trial state is, therefore, important for the model, and a variety of suggestions has been put forward (see Ref. 513 and references therein). For the study of hydrogen systems, a Gaussian with a time-dependent width has been used most frequently for single-electron states, even if recent extensions have been suggested.514,515 This limited functional has a ground-state binding energy of the hydrogen atom E011.5eV, but some attempts have been made to explicitly include the 1s state in the model.516 

Klakow et al. suggested an approximation to exchange effects, by considering a pairwise antisymmetrization of the kinetic energy,517,518 an idea that was extended by the electron force field (eFF) model by the introduction of fitting parameters to better match electron–ions bounding in low-Z elements.519–521 These types of pairwise Pauli potentials have been primarily used to study dynamic processes in hydrogen, e.g., proton transport,522,523 electron–proton temperature equilibration,524 plasma oscillations,278 and electron stopping power525 but also constructed a variational approximation to the imaginary time density matrix87 

For lower temperature conditions, a complete Slater determinant has been used as a trial state, producing a correctly antisymmetrized state, including effects on energies and the norm matrix.526–528 A low-temperature insulator–metal phase transition has been predicted for temperatures T< 4000 K, below 150 GPa.527,528

Lavrinenko et al. have suggested including the effect of exchange and correlations in the wave packet model by evaluating DFT functionals based on the density profiles from the wave packets,111,529,530 instead of manipulating the state directly. This has the greatest effect at high densities, ne> 1023 cm−3 and is seen to reduce the maximum compression along the Hugoniot compared to the explicitly anti-symmetrized models, more in accordance with experimental evidence111 (see also Fig. 2).

It is well documented526,531–533 that, at sufficiently high temperatures, the wave packets expand indefinitely, something that needs to be regularized. This expansion is arguably due to an inability to localize an electron on multiple ions.534 Multiple different suggestions have been made to limit the expansion,524,531,535,536 but the effect of this regularization must be tested.533 

2. Effective quantum pair potentials

The idea to “correct” the pair interaction for quantum diffraction and exchange effects was advanced by many authors, including Kelbg and coworkers468,469,537 and Deutsch and coworkers.538–540 Kelbg considered the canonical two-particle density matrix at a finite temperature kBT=1/β, which obeys the Bloch equation
(81)
For weak coupling, the result for the off diagonal density matrix is
and involves kinetic energy parts (the Gaussian factors) and an effective “quantum potential” Φ. The complicated coordinate dependence is often simplified to a diagonal potential, UIK, that depends only on the distance of the pair,
(82)
where erf denotes the error function, and we introduced the de Broglie wavelength with the reduced mass of the pair, λij=h(2πmijkBT)1/2, and γij is a free parameter. Consider first the case γij1, for which one recovers the potential UK derived originally by Kelbg468,469 (for a recent overview, see Ref. 140). In the limit T, this potential approaches the classical Coulomb potential whereas, for decreasing T, the potential exhibits systematically increasing deviations from the Coulomb potential, due to quantum diffraction effects and approaches a finite value at zero separation, UijK(0)=qiqjλij. At the same time, it is known that, while the derivative UijK(0) is correct, the absolute value is not, if the weak coupling approximation is violated.470 The correct value can be restored by fitting to the exact solution of the pair problem, with a single additional parameter, γij(T), and the corresponding results have been termed “Improved Kelbg potential (IKP).”470,471 Its value at zero pair separation is
which gives the fit parameter a simple interpretation: as a result of pair interaction effects the “extension” of a quantum particle becomes γij(T)λij(T) compared to the “ideal extension” λij(T). While e-e repulsion increases the extension, electron–proton attraction reduces it, ultimately to the in Bohr radius. The parameters γee(T) and γep(T) have a simple Padé representation given in Ref. 471 and are shown in Fig. 13.
FIG. 13.

Fit parameters of the improved Kelbg potential (82) for e-p and e-e (without exchange) interaction. γee and γep can be understood as modifications of the de Broglie wavelength due to e-e and e-p interaction, respectively, see text. Reprinted from Ref. 471 with permission of the authors. Copyright IOP Publishing. Reproduced with permission. All rights reserved.

FIG. 13.

Fit parameters of the improved Kelbg potential (82) for e-p and e-e (without exchange) interaction. γee and γep can be understood as modifications of the de Broglie wavelength due to e-e and e-p interaction, respectively, see text. Reprinted from Ref. 471 with permission of the authors. Copyright IOP Publishing. Reproduced with permission. All rights reserved.

Close modal

The Kelbg potential was used in PIMC simulations of hydrogen and electron–hole plasmas by Filinov et al.185,186,541,542 and also in simulations of quark-gluon plasmas.543–545 Further, the improved Kelbg potential was employed by Filinov et al. in recent PIMC simulations of hydrogen.115,140 There it was found that, at low temperatures, the convergence (with respect to the number of high-temperature factors P) with the Kelbg potential is very slow and faster with the IKP. Thus, the use of the exact pair density matrix is advantageous for such conditions. With increasing temperature the convergence improves rapidly for the IKP (see also Sec. III E 5).

Aside from PIMC, the Kelbg potential was also used in semiclassical MD simulations of partially ionized hydrogen for the computation of the density correlation function and the plasmon spectrum.546–548 MD simulations with the IKP of the thermodynamic properties of partially ionized hydrogen were reported in Ref. 471 and allowed one to accurately reproduce the equation of state for T 50 000 K. They break down, however, when molecules form because the IKP does not prevent cluster formation of more than two hydrogen atoms, due to the missing exchange of more than two electrons. Novel MD results with the IKP will be presented below, in Sec. IV D 6, where we compute the ion–acoustic plasmon mode in dense hydrogen.

The use of the IKP in classical MD simulations can be understood as a simple case of force fields. Further improvements of this concept have recently been achieved via machine learning approaches which we discuss in Sec. III G 3.

Note that the quantum potentials discussed above are important in order to reproduce short-distance phenomena in semiclassical simulations. On the other hand, to capture screening and collective effects of quantum particles inside a plasma, the Coulomb potential has to be renormalized at large distances. Examples are the Debye potential or the dynamically screened Coulomb potential [Eq. (87)].

Finally, the appropriate potential to compute the interaction between two electrons in a nonideal plasma is the Kukkonen-Overhauser potential.271 It has been tested in PIMC simulations in Ref. 270 and was recently used to explain the roton minimum observed in the plasmon dispersion of the correlated UEG549 (see Sec. IV D 4).

3. Effective ion–ion potentials: Machine-learning concepts

Methods to extract effective ion–ion potentials from first principles simulations have been studied for a number of years. They are expected to close the gap between cheap but heuristic pair potentials and computationally expensive ab initio calculations, needed to better explore all parts of the phase diagram of hydrogen. Forces or potentials from DFT-MD can be matched to a functional form of the potential or can be fit by a freely varying function.550–553 At the same time, model potentials have been fit to structural or thermodynamic data.554 

In this way, machine-learned (ML) effective potentials between molecules, atoms, and ions are a new class of potentials that are promising to be more flexible and transferable than effective potentials that are obtained in the more traditional ways (cf. Sec. III G 2).

In particular, the part of the phase diagram where the Born–Oppenheimer approximation can be applied seems to be predisposed to replacing ab initio DFT or QMC based BO energies by approximate ML models developed over recent years.245,246,555–558 The red box in Fig. 8 indicates the phase-space region where ML effective potentials have been created and applied to liquid and solid hydrogen.206,559–561

A ML model for the BO potential surface is trained on energies and forces of 103105 static nuclear configurations calculated by DFT or QMC methods, e.g., from ab initio MD, PIMD, or CEIMC runs. ML-based methods seem to perform equally well for QMC-based energies or forces despite the intrinsic stochastic errors.206,562

By running MD or PIMD simulations with effective ion potentials, one can then afford calculations on much bigger time and length scales. Although ML model potentials in particular, are built to ensure transferability, there is limited practical experience for hydrogen so far. The smooth crossover between the molecular and atomic limit found in Refs. 560 and 563 from a ML model of PBE-BO energies has not been confirmed later by direct ab initio PBE-MD.228 The absence of the liquid–liquid phase transition in the ML simulations might be due to the too coarse grid of the initial BO-DFT-MD simulations.

A ML potential study206 including nuclear quantum effects and QMC BO-energies (as well as PBE and vdW-DF1) predicted melting of the solid at considerably higher temperatures, for the ML QMC potential energy surfaces (see green line in Fig. 5), than the one obtained within PBE, with ML vdW lying in between, as well as a possible structural phase transition to a Fmmm-4 phase at higher temperature. The higher melting line prediction from ML models is in qualitative agreement with direct simulation by DFT-MD with vdW-DF functional and by CEIMC, as reported in the Supplementary material in Ref. 206.

An alternative, though closely related strategy to achieve size transferability based on machine-learned DFT results has recently been proposed by Ellis et al.564 Instead of constructing an effective potential, their strategy is to learn the local density of states (LDOS) and to parameterize it based on local SNAP descriptors.565 The LDOS then gives one access to standard DFT observables such as the energy or the electronic density, which, in principle, can be evaluated for very large numbers of ions; results for 131 072 Be atoms have been presented in Ref. 246. The application of this method to dense hydrogen constitutes an interesting topic for future work.

Dense plasmas exposed to external fields may be driven far from equilibrium and undergo a cascade of fast to slow relaxation processes, extending from femtosecond to nanosecond scales, before they return to equilibrium. Knowledge of the nonequilibrium behavior is important for understanding many experiments with dense hydrogen and deuterium, including inertial confinement fusion, for a recent discussion (see Ref. 259).

While truly first principles methods such as QMC (Sec. III E 3) exist for the ground state and thermodynamic properties of dense quantum plasmas, no approaches of similar accuracy and capability have been developed for nonequilibrium situations yet. Even though there exist real-time QMC methods, such as continuous time QMC that are being used in condensed matter and cold atom physics to study impurity models (e.g., Ref. 566) they are afflicted with an additional phase problem. This limits their application to very short simulation times and, to our knowledge, has ruled out their application to dense plasmas so far. Similarly, extensions of path integral methods to nonequilibrium are available; various versions of path integral molecular dynamics (PIMD) have been proposed (e.g., Refs. 507, 509, and 510). Yet applications to dense hydrogen have concentrated so far on equilibrium situations to capture nuclear quantum effects.215 On the other hand, the accuracy and applicability range of these methods for dense plasmas out of equilibrium remains to be fully explored.

Thus, the main method to simulate dense plasmas in nonequilibrium are quantum kinetic equations that can be derived using nonequilibrium Green functions or density operator methods (cf. Secs. III H 1 and III H 3). Other nonequilibrium approaches are extensions of DFT to nonequilibrium (cf. Sec. III H 4) and hydrodynamics (cf. Sec. III H 5). For an overview, see also Table VI.

1. Quantum kinetic equations (QKE)

Even though the equations of motion of quantum many-particle systems are known—the time-dependent N-particle Schrödinger equation, for the N-particle wave function, |ΨN(t), in the case of pure states, and the von Neumann equation for the density operator ρ̂N(t), for a mixed state—their solution for plasmas is neither possible nor necessary. To compute time-dependent macroscopic observables, such as transport, optical or dielectric properties, collision rates, or reaction cross sections, knowledge of far simpler quantities is sufficient. The proper link between the microscopic quantum-mechanical equations and the macroscopic quantities and their equations of motion is provided by quantum kinetic theory70—the proper generalization of Boltzmann's classical kinetic equation,
(83)
(84)
Here fa(r,p,t) is the phase space distribution function of particle species “a” which evolves in time due to spatial inhomogeneities, all forces, Fa [terms on the l.h.s. of Eq. (83)], as well as scattering processes described by the collision integral Ia. For the case of a plasma, Fa is the Lorentz force, Eq. (84), which involves the electromagnetic field that obeys Maxwell's equations. Since the charged particles induce an electromagnetic field themselves, the associated “mean field” (Vlasov or Hartree field) is fully accounted for by this force. The distribution function is normalized to the total particle number, Na, and gives access to all single-particle observable in general, nonequilibrium situations, including the particle density, na and the mean velocity, ua,
(85)
which provides the starting point for deriving hydrodynamic equations (cf. Sec. III H 5). Furthermore, the direct relation between distribution function and density [Eq. (85)] allows one to establish connections between quantum kinetic equations and time-dependent DFT (cf. Sec. III H 4).
Let us now turn to the r.h.s. of Eq. (83), which describes all scattering processes involving two or more particles, i.e.,
The three-particle integrals describe a large variety of scattering processes including inelastic processes, such as excitation/de-excitation or ionization of atoms and molecules (e.g., Ref. 171). On the other hand, the integrals Iab describe binary elastic scattering processes and are crucial for correctly capturing correlation and thermalization effects in the plasma. Here we concentrate on the two-particle integrals and, for illustration, present the Balescu–Lenard integral,567,568 which contains other collision integrals as limiting cases (see below) and plays a central role in plasma kinetic theory70 
(86)
(87)
a. Discussion of the collision integral: Special cases

The integral Ia describes the temporal change of the distribution fa(r,p,t) due to binary (or more complex) collisions with particles of all types “b” in all possible momentum states p which are occupied at time t with probability fb(p,t). The collision integral (86) is the lowest order perturbation theory result, it is proportional to the square of the interaction potential (second order in the coupling parameter Γ, second Born approximation, SOA) and contains a number of limiting cases which we briefly recall.

  • For ε(k,ω)1, Ia reduces to the Landau collision integral which contains the square of the Fourier transform of the Coulomb potential. In that case the k-integral diverges logarithmically which is “repaired” by introducing finite minimum and maximum values for the k-integration, resulting in the so-called “Coulomb logarithm” (for a discussion of its use in classical plasma kinetic theory, see Ref. 569),
    (88)

    The existence of a finite maximum wave number, kmax (or minimum distance, bmin), is usually justified by the finite extension of particles which is of the order of the De Broglie wavelength Λa (cf. Sec. III G 2), leading to the choice kmaxΛa1. On the other hand, the common choice of kmin=1/bmax is 1/rD—the inverse of the Debye screening length, Eq. (7). This is motivated by the screening of the long-range Coulomb interaction in a plasma, giving rise to the effective replacement of the Coulomb potential by the Debye potential, qaqbrqaqbrer/rD. In degenerate quantum systems, the Debye length is replaced by the Thomas-Fermi length, rTF [Eq. (9)], or by a proper interpolation between the two limits.

  • Use of the Debye potential leads to ii) the statically screened second Born approximation where, in Eq. (87), ε(k)=1+(krD)2.

  • This result is further improved by taking into account the dynamics of the screening cloud via the dynamic Vlasov dielectric function, ε(k,ω), of an ideal plasma giving rise to the dynamically screened pair potential, Vs(k,ω) [Eq. (87)]. This corresponds to the full Balescu–Lenard kinetic equation which selfconsistently includes screening effects (no phenomenological cutoff kmin is needed) and collective excitations (plasmons) in the two-particle scattering process (for a derivation, see Ref. 570). In the many-body NEGF theory, this approximation is directly related to the GW approximation (see below).

  • In nonequilibrium plasmas the dielectric function, in general, becomes time dependent as it involves the distribution function
    (89)

    and has to be updated during the solution of the Balescu–Lenard equation. This has been done in a number of works but leads to problems that will be discussed at the end of this section. Similarly, the static dielectric function becomes time-dependent in nonequilibrium, where the time-dependence of fa gives rise to a time-dependent screening length ε(k)ε(k,[rD(t)]) (for details, see Ref. 70).

Note further that, for dense plasmas, quantum generalizations of the kinetic equation (83) and of the collision integral are necessary (in particular, Pauli blocking and exchange contributions) which are straightforward and will be discussed below in the context of the G1–G2 scheme.

b. Linear response: Relaxation time approximation (RTA)
An important and popular special case of the collision integral (86) is obtained by linearization with respect to small deviations from the asymptotic equilibrium distribution, faeq,571,572
(90)
where, for faeq(p;n,T), usually a Maxwell, Bose or Fermi distribution is being used. Furthermore, τ denotes a characteristic time during which fa(p,t) approaches faeq which is computed for a plasma in thermodynamic equilibrium. The explicit expression for τ depends on the choice of the collision integral which is then linearized around fa=faeq. The integral (90) is sometimes called BGK (Bhatnagar, Gross, Krook) collision integral and is the starting point for the derivation of linear response quantities from kinetic equations (e.g., Ref. 573). An example is dielectric and optical properties that can be derived from the kinetic equation (83). While setting Ia0 leads to the mean field (Vlasov or RPA) results for the dielectric function, using Ia(p,t)|RTA allows one to include collision effects in a sum-rule preserving way, see the discussion of the Mermin dielectric function [Eq. (131) in Sec. IV D 5).

Aside from linear response theory applications, both, the Landau and Balescu–Lenard equation have been frequently solved for electron–hole plasmas and dense quantum plasmas that are driven far away from equilibrium (e.g., Refs. 574–577) to study the relaxation toward thermodynamic equilibrium.578 However, the proper choice of the collision integral is often not clear, and, therefore, the resulting time evolution of fa(r,p,t) is only qualitatively correct.

c. Properties and failures of the kinetic equation (83) with the collision integral (86)

It easily verified70 that the kinetic equations (83) and 86)

  1. obey conservation laws of particle number and mean momentum;

  2. conserve the mean kinetic energy, p2/2m where averaging is with fa(t). This, however, violates the correct conservation law of an interacting system where total energy, i.e., the sum of kinetic and interaction energy, is conserved.

  3. have an asymptotic solution, faeq(p), which is given by the Fermi or Bose distribution (if quantum generalizations for fermions or bosons have been performed, e.g., Ref. 70). However, in an interacting quantum system, the equilibrium distribution is different, as we will show explicitly in Sec. IV C.

  4. Numerical solutions of the quantum Balescu–Lenard equation [(83) and (86)] with the time-dependent dielectric function (89) revealed an unphysically rapid thermalization,574 indicating that the collision integral (86) is not applicable in situations far from equilibrium, at short times. In particular, the use of an equilibrium dielectric function in combination with a nonequilibrium distribution function is not justified.70 

Before discussing how to overcome these limitations of Markovian kinetic equations we briefly comment on the status of equilibrium Green functions (EGF) theory.

2. Equilibrium Green functions (EGF)

Nonequilibrium Green functions, Gij(t,t), which will be discussed in Sec. III H 3, depend on two time arguments and are defined on the round trip Keldysh contour.579 The NEGF contains thermodynamic equilibrium as a limiting case where the Keldysh contour shrinks to its imaginary branch (e.g., Refs. 297 and 580) and Gij(t,t)GijM(ω), where the frequency ω is the Fourier adjoint of the time difference, tt. This limit coincides with the Matsubara Green functions that have been actively investigated since the late 1950s (e.g., Refs. 581 and 582). This technique has been extensively applied to dense partially ionized plasmas in the 1970s and 1980s by the Rostock group (e.g., Refs. 68 and 69). The strength of equilibrium and nonequilibrium Green functions theory is the systematic approach (via Feynman diagrams) to incorporate correlation effects into the theory. At the same time, the set of available self-energies is limited and their validity range and accuracy are not known a priori. A few tests have been performed against FPIMC simulations for the uniform electron gas using the dynamic second order Born approximation (Montroll-Ward and e4 self-energies).117,121 Within their expected range of validity, rs1, the relative error of the interaction energy V reaches 10%117 which is the value shown in Table IV. Kas et al. presented improved EGF results that are based on a cumulant expansion and were also applied to the strong coupling regime, rs40.273,583 They reported overall good agreement with FPIMC-based fits (errors of the order of 10% for the interaction parts of the thermodynamic functions), however, significantly larger errors and unphysical temperature dependence were observed at intermediate densities.117 Note that these errors translate into much smaller errors for the total thermodynamic quantities, which is remarkable, considering the large density and temperature range.

A major advantage of Green functions methods is the direct access—via the frequency dependence of GM—to spectral properties such as the spectral function or density of states. This is achieved via an iterative solution of the Dyson equation,
(91)
This equation would be exact, would the self-energy be known. In practice, however, approximations for the functional dependence, ΣM[GM], have to be used. For a discussion of the solution of the Dyson equation (see Ref. 584). The second approach to the spectral properties consists of solving the Bethe–Salpeter equation (BSE)—the equation of motion of the two-particle Green function (e.g., Refs. 171 and 297).

Ab initio EGF simulations. The large computational effort of Green functions theory arises from the dependence of the Green function on two basis indices, i.e., the quantities in Eq. (91) have to be understood as matrices, GM(ω)GijM(ω). In the case of spatially uniform systems, such as dense plasmas or jellium, a plane wave basis in terms of momentum eigenstates is appropriate, leading to diagonal matrices, Gp,pMδp,p and Σp,pMδp,p. However, even in that case, the dimension of the matrices can be very large, since p are vectors (for details, see Sec. III H 3).

A very promising concept to reduce the computational effort and increase the accuracy of the results is to use, instead of plane waves, a basis of Kohn–Sham orbitals obtained from an independent KS-DFT simulation for the same system. This leads to a powerful combination of DFT with Green functions known as ab initio BSE or ab initio GW simulations and was successfully applied to various systems in the ground state (e.g., Ref. 276). This has also been extended to excited state and optical properties, e.g., with the yambo code.585 

3. Nonequilibrium Green functions and the G1–G2 scheme

The deficiencies of standard kinetic equations that were listed above led in the 1990s to the development of generalized quantum kinetic equations by Kremp and coworkers, who used nonequilibrium Green functions (NEGF)171,579,580,586,587 and reduced density operators.70 The resulting quantum kinetic equations are free of the aforementioned problems (e.g., Refs. 588–590) and were applied, among others, to quantum plasmas in strong laser fields591–593 and to the correlated dielectric function of the uniform electron gas.594 In recent years more advanced collision integrals could be implemented that include strong coupling effects (T-matrix approximation). Moreover, extensive tests of the accuracy were performed by comparison to cold atom experiments and exact diagonalization and density matrix renormalization group (DMRG) calculations.274,298,595 Even though these tests were possible only for lattice models, they have allowed one, for the first time, to rigorously benchmark the accuracy of quantum kinetic theory simulations with different collision integrals (self-energies). This has made quantum kinetic theory a predictive tool for nonequilibrium applications and a possible benchmark for other time-dependent approaches for dense quantum plasmas such as time-dependent DFT and quantum hydrodynamics.

a. The G1-G2 scheme

The main limitation of NEGF simulations is their high computational load: nonequilibrium simulations scale cubically with the number of time steps, Nt, which is due to the two-time dependence of the NEGF and the time nonlocal structure of the collision integrals (memory integration). This has changed dramatically when Schlünzen et al. were able to reformulate the NEGF equations as two coupled time-local equations for the one- and two-particle Green functions by invoking the generalized Kadanoff–Baym ansatz (GKBA):596 this allowed them to reduce the scaling to first order in Nt.597,598 Remarkably, this scaling is achieved not only for simple perturbation theory (SOA, Landau collision integral) but also for advanced self-energies, including T-matrix and GW [generalization of the Balescu–Lenard approximation, Eq. (86)] and combinations thereof resulting in the dynamically screened ladder (DSL) approximation.599 This has already triggered a large number of applications for lattice systems and 2D quantum materials. Below, we present the equations in a form suitable for dense quantum plasmas including fully ionized hydrogen.

Assuming a spatially uniform system consisting of multiple components (labeled α and β with charge numbers Zα and Zβ), the equations of motion, in the G1–G2 scheme, are conveniently rewritten in the momentum representation (to simplify the notation, we omit the time dependencies of G and Vq in some places)600,
(92)
(93)
(94)
where
(95)
To shorten the notation, we write the momentum argument as a subscript, i.e., Gpα<Gα<(p), and so on. In these equations, the single-particle Green function is related to the distribution function by iGα<(p,t)=fα(p,t), whereas iGα>(p,t)=1fα(p,t). The two-particle function entering the collision integral in the kinetic equation (92) denotes the correlated part of the two-particle Green function, G=G(2)GHF(2), where G(2) is related to the two-particle distribution function by (i)2Gαβ(2)=fαβ(2), and the Hartree–Fock hamiltonian is given by hp,αHF(t)=hp,α(t)+iqVpqGα<(q,t). Note that the G1–G2 scheme is complemented with initial conditions (93) for the one- and two-particle Green function at time t=0. In contrast to the Boltzmann equation, the initial state may be correlated,601,602 i.e., G(0)0, which is achieved by slowly turning on the pair interaction (“adiabatic switching”) during the solution of the equations, for a finite period of time t<0 (e.g., Refs. 298, 603, and 604).
Equation (92) is the generalized quantum kinetic equation which, due to the assumed homogeneity and absence of external fields does not contain additional terms on the l.h.s. (for generalizations to plasmas in strong fields, see Refs. 591–593). The collision integral I generalizes the Balescu–Lenard integral (86) to the case of arbitrary correlations the properties of which are defined by the two-particle correlation function G. It obeys a separate equation of motion, Eq. (94) which is the nonequilibrium generalization of the time-diagonal Bethe–Salpeter equation (BSE) when the GKBA has been applied.605 The terms on the r.h.s. specify the many-body approximation (self-energy): if Π is neglected, Eq. (94) corresponds to the static second Born approximation (generalization of the Landau equation). The Ψ term describes the two-particles scattering process including exchange and Pauli blocking. On the other hand, if the polarization term Π is included, the treatment of correlations is significantly improved to the dynamically screened second Born approximation, i.e., to the GW approximation (non-Markovian quantum Lenard–Balescu equation and Dyson equation for W). It is well known that the polarization term Π gives rise to a Dyson equation for the inverse dielectric function or the dynamically screened potential Vs(k,t,t)70—the nonequilibrium generalization of the potential (87),
(96)
where P is the retarded longitudinal polarization function. Equation (87) yields the nonequilibrium inverse dielectric function
(97)
and the complete plasmon spectrum, including its time evolution in nonequilibrium. This is the correct nonequilibrium generalization of the (inverse) dielectric function and eliminates the problems of the dielectric function (89) that were mentioned above.
For a stationary system, the dependence on the center of mass time, T=(t+t)/2 vanishes, and Fourier transformation with respect to τ=tt yields, applying the convolution theorem, the frequency dependence of all quantities
(98)
This leads the dielectric function of an ideal system as well as the inverse dielectric function
(99)
(100)
where the latter is expressed in terms of the density response function χ [for the definition, see Eq. (105)]. Using Eqs. (99) and (100), one finds that the density response function obeys the equation
(101)
a relation that is also being used in time-dependent DFT (see Sec. III H 4). The RPA result for χ follows when the polarization function of an ideal Fermi gas (Lindhard polarization, Pχ0) is being used. One way to go beyond the RPA result for χ is to introduce the dynamic local field correction, G(k,ω), leading to
(102)
The function G is discussed in more detail in Sec. IV D.

The second way to incorporate correlation effects is to further upgrade the BSE (94), which leads to the dynamically screened ladder approximation which takes into account plasmon effects beyond the RPA as well as bound states in the plasma environment (ionization potential depression) (cf. Sec. IV B).

So far, the G1–G2 equations have been solved for dense quasi-1D plasmas,606 and new results will be presented in Sec. IV F 3. Extensions of the G1–G2 scheme to 2D and 3D plasmas are straightforward, in principle, but suffer from a new bottleneck: the large memory consumption when storing the correlation function Gkpqαβ. This problem is expected to be solved in the near future by applying a novel quantum fluctuations approach developed by Schroedter and coworkers.605,607,608

4. Time-dependent DFT (TDDFT)

Real-time time-dependent Kohn–Sham density functional theory (RT-TDDFT) is based on Eqs. (57), which provides the time-dependent one-particle density n(r,t)=jfj|ψR¯j(r,t)|2. Although RT-TDDFT can be formulated to be formally exact, in practice several approximations are used. Most importantly for WDM applications, the XC functional is often utilized in a static (adiabatic) approximation, meaning that vxc depends only on the density value at a given time moment and does not have an explicit dependence on time; memory effects in the XC potential are thus neglected. This is not critical for systems such as WDM and dense plasmas if the excitation spectrum is qualitatively similar to that of the ideal electron gas. Examples of such features are plasmons at small wavenumbers (where collective oscillations are not strongly damped) and features of single-particle oscillations at large wavenumbers (where the kinetic energy of an electron, 2q2/2me, dominates over interaction energy terms). In addition, in the same way, for the same reasons, and with the same quality as in the equilibrium KS-DFT, RT-TDDFT with an adiabatic XC functional provides information about the energy levels of the orbitals localized around ions.

The drawback of using the static XC functional in RT-TDDFT is that it leads to time-independent occupation numbers.609 Second, for extended systems in general, and WDM simulations in particular, the initial state of the system is usually prepared using an equilibrium KS-DFT calculation with corresponding occupation numbers according to the Fermi–Dirac distribution. This in combination with the static approximation for vxc prevents the legitimate application of RT-TDDFT for the simulation of non-equilibrium effects on a timescale shorter than the relaxation time (Sec. III H 1), where the occupation numbers change significantly. Furthermore, the adiabatic approximation in vxc fails at perturbation frequencies (energies) that are outside the spectrum of the equilibrium KS-DFT state.610 

RT-TDDFT does not require an explicit condition of weak coupling between the perturbing external field and the electrons. This allows one to use it for the simulation of processes that are beyond the linear response approximation, e.g., in the simulation of the stopping power.611 In the context of ICF, the latter application is of particular importance for the alpha particle energy loss in warm dense hydrogen.612,613

An alternative and formally equivalent formulation of TDDFT in the linear response regime is referred to as linear-response TDDFT (LR-TDDFT). LR-TDDFT does not require an explicit time-dependent propagation of the wave functions since all needed information about the dynamic properties of the equilibrium system is already contained in the equilibrium state wavefunctions and eigenenergies. Therefore, the derivation of LR-TDDFT is based on a standard perturbative approach starting from the ideal (non-interacting) lowest-order approximation to the dynamic density response function. Further, the addition of density inhomogeneity effects and corrections due to exchange and correlations results in a generalization of Eq. (100) for the inverse dielectric function to non-uniform systems,
(103)
where k is a wave vector restricted to the first Brillouin zone, and G and G are reciprocal lattice vectors. The dielectric function of a homogeneous system follows from G=G components of the matrix εG,G1(k,ω).
Similar as in Sec. III H 3, this equation can be transformed into a Dyson equation for the microscopic electronic density response function χG,G(k,ω):614,615
(104)
where χG,G0(k,ω) is an ideal (non-interacting) density response function computed using Kohn–Sham orbitals, vG1(k)=4π/|k+G1|2 is the Coulomb potential in reciprocal space, and KG1,G2xc(k,ω) is the exchange–correlation (XC) kernel defined as functional derivative of the exchange–correlation potential.616 Equation (104) represents a microscopic density response to an external perturbation with the frequency ω and the wavenumber q=G+k, where G is a reciprocal lattice vector, and thus constitutes a generalization of Eq. (117) below.
The dependence on two different wavenumbers of the microscopic electronic density response function is due to the inhomogeneity of the system. To see this, it is illustrative to write the connection between the perturbing field, δv(r,t), and the resulting density perturbation, δn(r,t), in real space:
(105)
For an equilibrium system perturbed by a weak field, we have χ(r,t,r,t)=χ(r,t,r,t)=χ(r,r,tt). The introduction of periodic boundary conditions allows one to rewrite χ(r,t,r,t)=χ(r,r,tt) in Fourier space in terms of G and G as χG,G(k,ω).419 

For disordered (melted) extended systems such as warm dense matter, the system properties are homogeneous on average (over time or configurations). Therefore, to connect LR-TDDFT results with an experimental observable, such as XRTS, one needs to compute the averaged macroscopic density response function. This has been explained in detail for the example of warm dense hydrogen in Refs. 617 and 618.

We also point out that the Dyson equation (104) is the equilibrium limit of the two-time Dyson equation (96) that is studied in quantum kinetic theory (cf. Sec. III H 1). This also provides the opportunity to establish direct links between approximations for Kxc and for the self-energy, as well as between the results of the two approaches.

5. Classical and quantum hydrodynamics for ICF modeling

Radiation-hydrodynamics (RH) codes such as Hydra619,620 and FLASH,621,622 are the main computational tools used to design Inertial Confinement Fusion (ICF) experiments. The codes are multi-dimensional, multi-physics (coupled radiation, hydrodynamics, and thermonuclear burn), and run on large parallel computing machines. Behind every ICF experiment are thousands and thousands of runs using the RH codes. This design element is critical to an experiment's success, due to the great expense of the targets and diagnostics. Hence, the accuracy of RH codes is paramount. The process that each RH code goes through to ensure this accuracy is called verification and validation (VandV)623 and has to answer two questions: (1) Is the RH code solving the correct equations (validation)? (2) Is the RH code solving the equations correctly (verification)? The former question is typically addressed through a comparison of the RH code with experimental results of varying complexity. This might involve comparisons of Rayleigh-Taylor (RT) instability growth rates with experimental data coming from linear electric motor experiments.624 The latter question is typically addressed by a comparison of the RH code with analytic and semi-analytic results, including the Sod shock tube problem and the Marshak wave.625 

RH codes are tested on a large suite of validation and verification test problems before a code is used to simulate an ICF capsule. The trajectory (cf. Fig. 3) of the fuel burning region in an ICF capsule traverses the warm dense to hot dense matter regime. This wide range of physical conditions places a severe demand on the accuracy of RH codes. The hydrodynamics usually involves some form of the Navier–Stokes equations, and the algorithms are designed to capture shock formation and propagation. A description of turbulent mixing is also needed in ICF simulations due to the prevalence of Rayleigh-Taylor, Richtmyer-Meshkov, and Kelvin-Helmholtz instabilities. Radiation is frequently treated with gray or multi-group diffusion.625 RH codes are made up of hundreds of input parameters if not thousands. The equations require as input numerous physics quantities, including EOS, fusion reaction rates, opacities, and electronic and ionic transport coefficients, such as thermal and electrical conductivity and viscosity. EOS models, in particular, are a focus of study in ICF implosions since they can affect shock timing and material compressibility and thereby determine instability growth rates and hydrodynamic coupling to the DT fuel.

The challenge users and code developers of the RH codes face is: what choices should they should make for an ICF simulation? For example, what is the best choice for the EOS for the fuel or ablator material? In many cases, high-quality experimental physics data that could inform users of the RH codes about what choices to make does not exist. The reason is that it is difficult to obtain data for a single physics model in regimes where the temperatures are in excess of 100 eV and pressures are far in excess of 1 Mbar. Most experiments in this regime are integrated, involving many different physics models. Computational physicists, therefore, must rely on the results for EOS, opacities, transport coefficients, etc., coming from fundamental physics codes like MD, atomic kinetics, kinetic theory, QMC, and TDDFT.

Recently, a comparison of EOS models626 and transport coefficients627,628 in regimes relevant to ICF and based on fundamental physics codes was documented from three code comparison workshops to which the high energy density physics community were invited to participate. A brief discussion of the results and open questions will be given in Secs. IV A 1 and IV E 1. One conclusion is that there is a strong demand for accurate simulation data for hydrogen that have predictive capability. In this paper, we present novel data for the equation of state and transport properties of dense hydrogen in the relevant parameter range that are based on first principles simulations. We also perform accuracy tests of various models for dense hydrogen (cf. Secs. IV A 3 and IV A 4). This should help to reduce uncertainties of existing model predictions for parameters relevant to ICF.

At the same time, first principles-quality modeling of the entire ICF explosion is still out of reach. QMC simulations, so far, only describe thermodynamic equilibrium situations, including static and dynamic properties (see Secs. III E and IV D 2). On the other hand, first principles time-dependent approaches, such as quantum kinetic methods (Sec. III H 1) and time-dependent DFT (TD-DFT) (Sec. III H 4) only capture relatively short time periods, on the femtosecond to picosecond scale. At the same time, phenomena such as the hydrodynamic implosion of an ICF capsule or the shock propagation and various instabilities (such as the Rayleigh-Taylor instability) take place at significantly larger length and time scales that are currently inaccessible to the aforementioned ab initio methods. Nevertheless, these first principles approaches could be very useful for benchmarks of the hydrodynamic models in well-defined test cases.

A promising compromise between first principles simulations and hydrodynamics could be Quantum hydrodynamics (QHD)—an approach that captures the dynamics of the quantum many-body system in terms of hydrodynamic field variables, such as density and velocity, directly extending classical hydrodynamics. A version of QHD that extends the picture of Madelung and Bohm629,630 of one-electron quantum mechanics to many-particle systems was proposed by Manfredi and Haas631 and became very popular in quantum plasma simulations leading, however, also to poorly controlled applications (for discussions, see Refs. 632–635). At the same time, this model does not reproduce the correct plasmon dispersion. It was demonstrated by Zh. Moldabekov et al. how this problem can be fixed and, moreover, how one can correctly account for exchange and correlation effects (missing in the original formulation), e.g., by using Local field Corrections from QMC simulations,635,636 or how one can use input from DFT simulations.275 In fact, the importance of quantum effects for shock wave propagation was demonstrated recently637 which means that QHD could become a valuable tool also for ICF modeling. We will return to these questions in Sec. V B.

In this section, we present state-of-the-art simulation results for dense hydrogen. This includes, in Sec. IV A, the equation of state, pair distribution functions, degree of ionization, and ionization potential depression (IPD). In Sec. IV C, we consider the momentum distribution function and, in Sec. IV D, static and dynamic density response properties. In Sec. IV E, we summarize transport and optical properties, such as the electrical and thermal conductivity and the opacity. The results are a combination of existing data and novel simulations. At the beginning of each section, the origin of the data are explained including necessary details to allow for reproducibility. Finally, in Sec. IV F, we discuss time-dependent simulations.

1. Previous comparisons

Comparisons of the equation of state results for dense hydrogen from different models have been performed at various places. A recent comparison by Gaffney et al.626 focused on ICF parameters and included extensive deuterium data that resulted from a code comparison workshop to which the high energy density physics community was invited. The models present included CEIMC, KS-DFT, OF-DFT, AA-DFT, and various combined EOS tables. The conclusions can be summarized as follows: (1) 510% model-model variations exist throughout the relevant parameter space and can be much larger in regions where ionization and dissociation are occurring, (2) the deuterium EOS is particularly uncertain, with no single model able to match the available experimental data, and this drives similar uncertainties in the CH EOS, and (3) new experimental capabilities such as Hugoniot measurements around 100 Mbar and high-quality temperature measurements are essential to reducing EOS uncertainty.

The reported large variations between models regarding the deuterium EOS is one of the motivations of the present paper. In Secs. IV A 2 and IV A 3, we reevaluate the hydrogen EOS in the difficult region of coexistence of atoms, molecules, and free charges for T15000 K and rs3. As we will show, in this parameter range, our benchmark comparisons allow for the appropriate choice of approximations in the first-principles approaches and significantly reduce the uncertainties in the EOS.

2. Origin of simulation data

The quantum Monte Carlo data presented in this section, which serve as benchmark data have been published recently: the RPIMC data are due to Militzer et al., published in Ref. 46. The fermionic PIMC data are due to Filinov and Bonitz, published in Ref. 115.

These PIMC data are compared to extensive new DFT simulation results. They have been obtained using the code VASP.352–355 Standard issue PAW pseudopotentials were used.638,639 The Mermin formalism of DFT was used to include temperature effects with the appropriate Fermi smearing of the occupation of the bands.330 The XC functionals used include ground state LDA,381 PBE-GGA,383 and the temperature-dependent KDT16 functional393 (for details, see Sec. III C). The number of protons in the simulation box varies from N=512, for the highest densities, to N=64, for the lowest densities shown in the figures. We also ran DFT-MD simulations with different particle numbers and confirmed that finite size effects are negligible. The MD time step size was generally chosen to be Δt=0.10.2 fs. We monitored the temperature, energy, and pressure fluctuations in the DFT-MD runs within the NVT ensemble for proper sampling and adjusted the thermostat parameter Nosé mass accordingly.640,641 The number of bands needed to converge the DFT calculations (to within 0.2%) was adjusted such that the highest energy eigenvalue has an occupation not exceeding 5×105. This requires several thousand bands. Checks with a different number of k-points confirmed that most of the time the Γ-point was sufficient, and the Baldereschi mean value point was chosen for higher densities. The plane wave cutoff needed (for convergence to within 0.2%) was between Ecut= 600 eV for the standard PAW potential (used at lower densities) up to Ecut=1200eV for the hard GW PAW potentials used for the highest densities. These measures ensure that the pressure as obtained from DFT-MD is converged to better than 1%.

3. Comparison of RPIMC, semiclassical MD, and FVT with fermionic PIMC simulations for the pressure

We start the analysis of the thermodynamic properties of dense hydrogen by a comparison of the equation of state between available QMC results in Fig. 14. Using the recent fermionic PIMC results of Ref. 115 as benchmarks allows us to judge the accuracy of the RPIMC simulations and the validity range of the used free-particle nodes. We also include new results from semiclassical MD simulations (Sec. III G 2) and from a chemical model (Fluid Variational Theory, FVT) of Juranek et al.63 (for details, see Sec. III B and Table VII). It is quite obvious that, over a very broad range of densities and temperatures, FP-PIMC and RPIMC, which are completely independent simulations, agree to a remarkable degree. That validates the use of free particle nodes in RPIMC. For the two higher isotherms, the deviation is below 2%. On the other hand, at T= 15 625 K we find a maximum deviation of 6% in the region with a substantial amount of molecules. These differences are observed at the remarkably low value of Θ0.5 which corresponds to the smallest rs-value accessible in FP-PIMC due to the FSP, where simulations are difficult to converge. We expect that these differences are due to the deteriorating quality of the nodal surfaces input in RPIMC. FVT, on the other hand, can serve as a quick estimation of the EOS, up to about 16 000 K/with up to 20% deviation in the pressure. The molecular dissociation from rs=3 to rs=14 is described very well at T= 15 625 K using FVT, which is also confirmed by Fig. 22 in in Sec. IV A 6.

FIG. 14.

Panel (a) Three isotherms of the hydrogen pressure (in units of the ideal Fermi pressure)—comparison of FP-PIMC simulations (fermionic propagator PIMC of Ref. 115, RPIMC simulations 488, classical MD using the improved Kelbg potential (Sec. III G 2), and from FVT.63 Panel (b) Ratio of the pressures from RPIMC and FP-PIMC for three isotherms as a function of rs, for which results from both QMC simulations are available. The data of subfigure (a) are also presented in Table VII.

FIG. 14.

Panel (a) Three isotherms of the hydrogen pressure (in units of the ideal Fermi pressure)—comparison of FP-PIMC simulations (fermionic propagator PIMC of Ref. 115, RPIMC simulations 488, classical MD using the improved Kelbg potential (Sec. III G 2), and from FVT.63 Panel (b) Ratio of the pressures from RPIMC and FP-PIMC for three isotherms as a function of rs, for which results from both QMC simulations are available. The data of subfigure (a) are also presented in Table VII.

Close modal

Finally, the comparison with semiclassical MD simulations that use the improved Kelbg potential (Sec. III G 2) is presented for one isotherm of T=62500 K (red crosses). The results are within 13% of the FPIMC data, for rs7, which is remarkable since the plasma contains a significant fraction of atoms. A similar agreement is observed for higher temperatures where the accessible density range increases with T. For example, for T=95000 K (T=125000 K) the density range grows to rs5 (rs3). For temperatures below the shown isotherm, the results are not reliable anymore because the current version of SC-MD does not describe molecules sufficiently accurately.

4. Comparison of KS-DFT with fermionic PIMC

An in-depth comparison of state of the art DFT-MD simulations with quantum Monte Carlo results for the EOS of warm dense hydrogen is performed in Fig. 15. We show the same isotherms as in Fig. 14 but, in addition, also an isochore, for rs=3. The comparison between the two different flavors of PIMC (at conditions where both RPIMC and FP-PIMC provide data) has already been established above. Here, we see clearly that RPIMC can provide data for high densities, rs3, or low temperatures where the vanishing average sign makes this prohibitively expensive for FP-PIMC, even though the accuracy of RPIMC cannot be quantified.

FIG. 15.

Benchmarks of KS-DFT-MD results for the EOS of hydrogen (normalized to the pressure of an ideal Fermi gas). Panels (a)–(c) show three isotherms, and panel (d) an isochore. We compare DFT-MD results with the zero-temperature LDA functional (violet circles), the PBE XC-functional (black stars), and using the temperature-dependent KDT16 functional (red diamonds),393 respectively. RPIMC simulations465 are shown by the cyan squares and FP-PIMC data115 by the blue squares. Also shown are average atom data (green circles), for a discussion, see text. The statistical errors of the data are within the size of the symbols. Panels (b)–(d) show, in addition, the ratio of various results to FP-PIMC (lines with arrows, right hand side y-axis). Part of the data are also presented in Table VII.

FIG. 15.

Benchmarks of KS-DFT-MD results for the EOS of hydrogen (normalized to the pressure of an ideal Fermi gas). Panels (a)–(c) show three isotherms, and panel (d) an isochore. We compare DFT-MD results with the zero-temperature LDA functional (violet circles), the PBE XC-functional (black stars), and using the temperature-dependent KDT16 functional (red diamonds),393 respectively. RPIMC simulations465 are shown by the cyan squares and FP-PIMC data115 by the blue squares. Also shown are average atom data (green circles), for a discussion, see text. The statistical errors of the data are within the size of the symbols. Panels (b)–(d) show, in addition, the ratio of various results to FP-PIMC (lines with arrows, right hand side y-axis). Part of the data are also presented in Table VII.

Close modal

Let us now compare Kohn–Sham DFT-MD simulations to the FPIMC results for densities larger than rs=4. At T= 15 625 K, taken as an example for situations when a ground state XC-functional should be a good approximation, due to the still high electron degeneracy, we find excellent agreement for rs4 with FP-PIMC. For lower rs (higher densities), the RPIMC results consistently give higher pressures than DFT-MD indicating differences in the description of this system within which a substantial amount of molecules is present. There are two uncontrolled approximations here. First, the XC-functional used in DFT seems to be sufficiently well approximated using PBE, for rs4. On the other hand, for rs<4, no conclusion on the appropriate functional can be drawn yet, and it remains to investigate how other functionals, such as vdW, HSE, or SCAN behave. Second, the free-particle nodes used in RPIMC may lose validity. This is supported by the systematically too high pressures of RPIMC, already at rs=5 and rs=4.

The situation is different for elevated temperatures of T= 31 250 K and T= 62 500 K [see panels (b) and (c) of Fig. 15]. Again, for rs4, PIMC shows a consistent picture whereas DFT-MD using a ground state PBE functional (but the temperature-dependent Mermin formulation of DFT) overestimates the pressure by up to 7%. On the other hand, taking into account temperature-dependent XC effects with the KDT16 GGA functional lowers the pressure such that the deviation from FP-PIMC is below 2%, for all conditions. It is interesting to observe that the main improvements, when using temperature-dependent XC-functionals, are found at intermediate temperatures of a few eV (up to 20 eV) such that the degeneracy is lifted but the ideal kinetic pressure is not dominant yet.

The general conclusion is that, for the range of conditions presented here, that includes the molecular, atomic, and partially ionized fluid/plasma state of hydrogen and covers densities slightly higher than that of the solid to very low densities, and temperatures T 15 000 K, the (temperature-dependent) GGA functional is accurate, and no meta, hybrid, exact exchange, or van der Waals functionals appear to be required. At the same time, larger differences are observed for T = 31 250 K, and the best choice of XC-functional remains open.

For completeness, in Fig. 15 we also include data from a DFT-AA model using LDA exchange. Here, the ion pressure contribution is taken to be the classical ideal pressure at the given ion density and temperature, which tends to overestimate ion pressures at lower temperatures and densities where molecular bonding can occur. Electron pressures from DFT-AA models can be calculated using the virial theorem642 or free-energy derivatives;401,407 however, here the electron pressure has been calculated using the simple prescription of Ref. 643.

5. Pair distribution functions and static structure factor

The pair distribution functions (PDF or radial distributions) are a sensitive indicator of quantum and correlation effects in dense hydrogen. They also directly reflect the structural properties. Here, we illustrate this for several phases of hydrogen that were discussed in Sec. II B, starting from low temperatures. Figure 16 shows results for fluid hydrogen at T= 1200 K, for two densities corresponding to rs=1.34 and rs=1.44. Those results have been obtained by CEIMC for classical protons and illustrate how correlations change across the LLPT. At the lower density (rs=1.44) the system is molecular, as seen from the pronounced peak around 1.4aB in gpp, followed by a nearly vanishing signal around 1.8aB. Instead, at the higher density (rs=1.34), the molecular peak disappears and the correlation function has a much smoother behavior signaling the absence of stable structures. In panel (b) of the figure we report the proton-electron correlation =4πρr2gep(r), where ρ is the electron density, and we compare with the electronic density corresponding to the isolated atom ground state (red line). Instead of a peak at r=1aB, which would be due to hydrogen atoms, the simulations exhibit only a weak shoulder indicating that the vast majority of the electrons are delocalized, which is also confirmed by the results of Fig. 30 in Sec. IV C 2.

FIG. 16.

Radial distribution functions for hydrogen at T=1200 K for two densities across the LLPT. Panel (a): proton-proton gpp(r), panel (b): proton–electron r2gep(r), panel (c): spin-unlike electron gee(r), and panel (d): spin-like electrons gee(r)=gee(r). Distances are in units of aB. Red dashed-dotted line: radial probability density of the hydrogen ground state. rs=1.44 corresponds to the molecular fluid phase, whereas rs=1.34 corresponds to a metallic phase with mostly delocalized electrons (see also Fig. 30).

FIG. 16.

Radial distribution functions for hydrogen at T=1200 K for two densities across the LLPT. Panel (a): proton-proton gpp(r), panel (b): proton–electron r2gep(r), panel (c): spin-unlike electron gee(r), and panel (d): spin-like electrons gee(r)=gee(r). Distances are in units of aB. Red dashed-dotted line: radial probability density of the hydrogen ground state. rs=1.44 corresponds to the molecular fluid phase, whereas rs=1.34 corresponds to a metallic phase with mostly delocalized electrons (see also Fig. 30).

Close modal

In panels (c) and (d) we report the electron–electron pair distribution functions for spin-unlike and spin-like electrons, respectively. While little change with density is observed in the spin-like function, besides the trivial change in average distance, in the spin-unlike functions, at the lower density, we can see the presence of molecules as a characteristic structure with a first maximum around r0.7aB induced by the charge accumulation between the protons to form the molecular bond, and a second maximum at about the distance between molecules. A more complete characterization of correlation in hydrogen around the LLPT, including a discussion of nuclear quantum effects obtained from CEIMC with nuclear path integrals is reported in Refs. 197, 208, and 498.

Next, we show results for the partially ionized plasma phase. Data from fermionic PIMC simulations of Ref. 115 are shown in Fig. 17 for T= 15 640 K and rs=7. As in the previous low-temperature case, the presence of hydrogen molecules can be identified from the peak in gii(r) for ion–ion distances, 0.7År1.3Å, as well as from g(r), which indicates accumulation of electron pairs with different spins between two protons. The peak position of gii(r) agrees well with the ground state atom separation in a hydrogen molecule. The simultaneous presence of atoms is reflected by the peak of the electron–ion function multiplied by 4πr2 which is located close to 1aB. It is interesting to again compare the FPIMC data to KS-DFT results with the PBE functional (dotted lines). Both the e-i and i-i PDF are in reasonable agreement, indicating a similar fraction of molecules in the plasma.

FIG. 17.

Pair distribution functions of partially ionized hydrogen for T= 15 640 K and rs=7, from FP-PIMC simulations of Ref. 115 using P=96 fourth order high-temperature factors and N=34. Green line: electron–ion PDF multiplied by 4πr2. Dotted lines: KS-DFT results with PBE-XC functional.

FIG. 17.

Pair distribution functions of partially ionized hydrogen for T= 15 640 K and rs=7, from FP-PIMC simulations of Ref. 115 using P=96 fourth order high-temperature factors and N=34. Green line: electron–ion PDF multiplied by 4πr2. Dotted lines: KS-DFT results with PBE-XC functional.

Close modal

The pair distributions for a four times higher temperature and two densities, corresponding to rs=5 and rs=3, are shown in Fig. 18. At this temperature, no molecules exist, as is confirmed by the monotonic ion–ion PDF. Also, there is only a small fraction, xA, of atoms which is indicated by the broad peak (shoulder) of r2gei, around one in bohr radius for rs=5 (rs=3), where xA increases when the density is lowered. The comparison between FPIMC and KS-DFT shows reasonable agreement for gii, where the latter predicts a slightly too strong proton-proton repulsion (too broad minimum at small distances). The electron–proton PDFs (bottom right figure) show excellent agreement with FPIMC. The figure also allows for a comparison with average atom models (AA, dashed lines in the lower left figure). The results underestimate the i-i-repulsion leading to a significantly narrower minimum.

FIG. 18.

Pair distribution functions of partially ionized hydrogen at T= 62 500 K and rs=3 and rs=5. Top panels: e-e PDF with the indicated spin combinations. Bottom panels: i-i PDF (left) and 4πner2gei(r) (right). Full lines: FP-PIMC results; dotted lines: KS-DFT with the PBE functional. Dashes (bottom left figure): average atom model (AA).

FIG. 18.

Pair distribution functions of partially ionized hydrogen at T= 62 500 K and rs=3 and rs=5. Top panels: e-e PDF with the indicated spin combinations. Bottom panels: i-i PDF (left) and 4πner2gei(r) (right). Full lines: FP-PIMC results; dotted lines: KS-DFT with the PBE functional. Dashes (bottom left figure): average atom model (AA).

Close modal

We continue the analysis of the AA model, but now in Fourier space. Figure 19 shows a comparison of the electronic and ionic static structure factor, between PIMC and DFT-AA, at a relatively high temperature (145000K, or about 12.5eV) and rs=2, where the plasma is fully ionized. There is excellent agreement in the static ion–ion structure factor Sii(q) predicted by the two approaches. For Sei(q), DFT-AA has two possible representations of the static electron–ion structure: one based on the self-consistent electron density within the Wigner–Seitz/ion-sphere cell about a single ion (labeled DFT-AA), and another one based on the neutral pseudo-atom electronic density (DFT-AA-NPA), which includes the effect of ion correlations and allows the electron density belonging to a single ion to extend beyond the ion-sphere radius.403,405 The PIMC results are in much better agreement with the NPA results.

FIG. 19.

Static ion–ion and electron–ion structure factors for hydrogen at T= 145 000 K and rs=2. PIMC results from Ref. 149 are compared with two versions of the DFT-AA model, see text for details.

FIG. 19.

Static ion–ion and electron–ion structure factors for hydrogen at T= 145 000 K and rs=2. PIMC results from Ref. 149 are compared with two versions of the DFT-AA model, see text for details.

Close modal

6. Degree of ionization and dissociation

For chemical models, the degree of ionization, α(n,T), and the degree of dissociation, β(n,T) (or, equivalently, the fraction of atoms xA), follow from the solution of the coupled Saha equations [Eqs. (51) and (52)]. The accuracy of the results sensitively depends on the quality of the interaction contributions to the chemical potentials, as discussed in Sec. III B and is not known a priori. A typical example will be discussed in Fig. 20.

FIG. 20.

Cluster analysis for classical protons at T=1200 K. (a) Distribution of the bond distance in the pairs found by the cluster algorithm with a cutoff distance of 2.6aB. (b) Distribution of the number of pairs with a cutoff distance of 1.8aB. (c) Distribution of the number of different partners, for a given proton, in the pairs found with the large cutoff value. (d) Molecular fraction from different estimators: Nav from the average number of molecules with the short cutoff value, from the coordination number at r=1.4aB, and Pp from the values of the distribution in panel (c) for np=1.

FIG. 20.

Cluster analysis for classical protons at T=1200 K. (a) Distribution of the bond distance in the pairs found by the cluster algorithm with a cutoff distance of 2.6aB. (b) Distribution of the number of pairs with a cutoff distance of 1.8aB. (c) Distribution of the number of different partners, for a given proton, in the pairs found with the large cutoff value. (d) Molecular fraction from different estimators: Nav from the average number of molecules with the short cutoff value, from the coordination number at r=1.4aB, and Pp from the values of the distribution in panel (c) for np=1.

Close modal

In contrast, in physical approaches, such as quantum Monte Carlo or DFT, there is no strict subdivision of free and bound electrons. Only for an isolated atom such a subdivision could be introduced based on the sign of the energy eigenvalue—negative (positive) for bound (free) electrons. However, for finite temperatures this boundary is being “washed out” and, due to the additional thermal energy, highly excited atomic bound states will statistically contribute to the scattering spectrum. Moreover, in a plasma at high density, the interaction between atoms and the overlap of electronic orbitals from neighboring atoms may significantly modify the energy spectrum (shift of energy eigenvalues, lowering of ionization energy, Mott effect, etc.) which will be discussed in Sec. IV B. In this section, we present first principles results for the degree of ionization and dissociation of hydrogen in different phases, that were discussed before. We start with the low-temperature case and the LLPT and then consider partially ionized hydrogen in the gas phase.

The characterization of hydrogen across the LLPT in terms of the molecular fraction and the possibility of stable ionic composites like H2+ and H3+, based on CEIMC results, have been discussed in detail in Ref. 208. There, different possible definitions of the molecular fraction were discussed and compared. As an illustration, in Fig. 21 we show the analysis for classical hydrogen along the T= 1200 Kisotherm across the LLPT. Results for quantum protons, detailed in Ref. 208, are qualitatively similar and pushes the LLPT to slightly lower pressures (5–10% depending on temperature). As explained in Ref. 208, we assigned molecules for each configuration along the Monte Carlo trajectory by a cluster analysis and by a pair distance criterion. The “bond length” distribution at six different densities is shown in panel (a) of Fig. 21. A discontinuous behavior occurs at the LLPT: in the molecular phase (continuous lines) the distribution is narrow, strongly peaked around the molecular bond length (1.4aB), and independent of density. In the dissociated phase (dot-dashed lines) the distribution is wider, peaked at a distance larger than the molecular bond (1.6aB), more asymmetric, with a detectable tail at large distances, and with a larger sensitivity to density.

FIG. 21.

Results for the fraction of atoms α of fluid hydrogen as a function of density for three temperatures. A comparison of the FVT version of Ref. 62 with tight-binding molecular dynamics simulations311 is performed. These models neglect free electrons and protons and use the definition α=nA/(nA+nM). The pressure-induced breakup of molecules occurs around ρ= 1 g/cm3, corresponding to rs1.392, which is consistent with the low-temperature behavior shown in Fig. 21(d). Adapted from Juranek et al.63 with the permission of the authors. Reprinted with the permission of AIP Publishing.

FIG. 21.

Results for the fraction of atoms α of fluid hydrogen as a function of density for three temperatures. A comparison of the FVT version of Ref. 62 with tight-binding molecular dynamics simulations311 is performed. These models neglect free electrons and protons and use the definition α=nA/(nA+nM). The pressure-induced breakup of molecules occurs around ρ= 1 g/cm3, corresponding to rs1.392, which is consistent with the low-temperature behavior shown in Fig. 21(d). Adapted from Juranek et al.63 with the permission of the authors. Reprinted with the permission of AIP Publishing.

Close modal

We also look at the distribution of the number of different neighbors a single proton experiences during the simulation, as shown in panel c) of Fig. 21. This quantity is useful when nuclear dynamics is not available, as in CEIMC and PIMD, to emulate the persistence time criterion employed in ab initio molecular dynamics simulations of classical protons88 where a pair of neighboring protons was defined as a molecule if it remained bound for a time interval of at least 10 characteristic molecular vibrational periods. Again we observe a striking change at the LLPT: for densities up to rs=1.4, very stable molecules are observed since the protons experience basically the same neighbor along the entire trajectory. For higher densities (rs<1.4), the distribution exhibits a long tail, indicating a strong attitude to being paired with many different neighbors during the sampling of the configurational space. One possible estimator of the molecular fraction is given by the maximum of this distribution and is reported in panel (d) of the figure and denoted as Pp. Another possible estimator is obtained by computing the distribution of the number of distinct pairs within a cutoff distance of 1.8aB corresponding to the first minimum of gpp(r), as shown in panel (b) of the figure. Again an abrupt change of the distribution is observed at the LLPT: in the dissociated phase (dot-dashed lines) the distribution is rather broad and peaked around the value of 20, while in the molecular phase (continuous lines) we observe a strong peak at 27 which is the maximum number of pairs in our simulated system. The molecular fraction can then be defined as the average over those distributions, and it is denoted by Nav in the plot in panel (d).

Finally, a third definition of the molecular fraction has been proposed in Ref. 104 as twice the coordination number obtained from gpp(r) at r=1.4aB, i.e., at the distance of the molecular peak. This definition is also shown in panel d) and denoted as Holst. For the present case, we see that the Holst estimator and Pp are in rather good agreement, whereas Nav is strongly overestimating the molecular fraction in the dissociated phase. This is the general trend also for other isotherms and for systems of quantum nuclei below the critical temperature of the LLPT.208 Above the critical temperature, when molecular dissociation with pressure becomes a continuous process, the agreement between the Holst estimator and Pp is lost, and it is not clear which estimator is more reliable (see Ref. 208 for discussion).

We now turn to higher temperatures which are in between the liquid phase studied in Fig. 21 and the partially ionized plasma phase above T= 15 000 K that is considered below. This range is difficult to access by QMC simulations so we resort to an advanced chemical model (FVT) (cf. Sec. III B).

The dissociation degree, as predicted for hydrogen in the temperature range between 3000 and 10 000 K, is shown in Fig. 20.63 Very good agreement with tight-binding molecular dynamics (TB-MD) simulations311 is evident. The comparison with Fig. 21(d) shows that, for the lowest temperature, the breakup of molecules occurs nearly at the same density, rs1.39, corresponding to the density ρ 1 g/cm3. At the same time, the density interval of the pressure-induced molecule dissociation is much larger than in the fluid phase where it occurred in the interval rs[1.37,1.40] corresponding to ρ[0.982,1.049] gcm−3. The comparison with Fig. 20 indicates that FVT predicts a much softer pressure dissociation. The comparison with TB-MD indicates that this could be due to limitations of the chemical model, so additional first principles simulations are needed to resolve this question.

We now turn to lower densities, ρ 0.5 g/cm3 where the fraction of atoms again increases. This is a statistical effect related to a decrease in the probability of two atoms to approach each other sufficiently closely in order to form a bond. This behavior is confirmed by fermionic PIMC simulations as well as by RPIMC simulations for T= 15 625 K115 (cf. left part of Fig. 22). There the two QMC simulations are compared to the results of FVT and another chemical model. Interestingly, we observe very good agreement between FP-PIMC and FVT, whereas the RPIMC results for the fraction of atoms differ significantly.

FIG. 22.

Fraction of free electrons, atoms, and molecules, for two isotherms, T= 15 640 K (left) and T = 312 50 K (right). FP-PIMC results from Ref. 115 are plotted for dHcr=1.9aB (brown solid dots) (dHcr=2.25aB, open gray circles), for details, see text. Blue lines: RPIMC data for dHcr=1.9aB. Red lines: chemical model. Reproduced from Ref. 115 with permission of the authors. Copyright (2024) by the American Physical Society.

FIG. 22.

Fraction of free electrons, atoms, and molecules, for two isotherms, T= 15 640 K (left) and T = 312 50 K (right). FP-PIMC results from Ref. 115 are plotted for dHcr=1.9aB (brown solid dots) (dHcr=2.25aB, open gray circles), for details, see text. Blue lines: RPIMC data for dHcr=1.9aB. Red lines: chemical model. Reproduced from Ref. 115 with permission of the authors. Copyright (2024) by the American Physical Society.

Close modal

The behavior of FVT changes dramatically when the simulations move further into the gas phase, at T= 31 250 K. Here FVT yields good results for the fraction of molecules, however, it strongly overestimates the fraction of atoms. This is due to the ionization of atoms which is important for this temperature already at rs=3, but missing in the FVT model indicating that it is restricted to lower temperatures. Note that a generalization of FVT to the case of partial ionization is possible (see, e.g., Ref. 644).

In contrast to FVT, first-principles FPIMC simulations are possible for temperatures above T= 15 000 K for a very broad range of densities (cf. Fig. 22). The fermion sign problem sets a lower limit for FPIMC of about rs4.115 The criterion for the identification of molecules in FPIMC and RPIMC is based on a critical proton-proton distance, dii that is indicated in the figure. By using two different values, the impact of the threshold can be quantified. The present criterion is similar to the one used in part (b) of Fig. 20 for the analysis of hydrogen in the vicinity of the LLPT. On the other hand, the procedure to identify atoms in a fermionic PIMC simulation is based on a cluster analysis that counts the number of electrons inside a sphere around each proton, for details we refer to Ref. 115. The FPIMC simulations demonstrate that, for low densities corresponding to rs15 (rs7), at T= 15 625 K (at T= 31 250 K) the atomic fraction decreases rapidly in favor of unbound electrons and protons. As in the case of the molecule breakup discussed above, this is a statistics effect.

Results for the degree of ionization from FPIMC and RPIMC for four isotherms are shown separately in Fig. 23. The behavior that was discussed for the two temperatures shown in Fig. 22 is here confirmed also for two higher temperatures. With increasing temperature, the fraction of atoms is, obviously, significantly lower, and their breakup sets in already at higher densities. Of particular interest is the minimum of the atom fraction which is observed around rs4.8 (rs3.9), for T= 62 500 K (T= 125 000 K). To the right of the minimum, i.e., for higher densities, the atom fraction, xA, decreases due to pressure ionization, reaching zero at around rs=1.2.

FIG. 23.

Four isotherms of the degree of ionization α of hydrogen vs density or rs. Comparison of RPIMC results of Ref. 189 and fermionic PIMC results of Ref. 115. Thin lines indicate the expected connection of the FPIMC data to the Mott density, rs=1.2, where the FSP prohibits FPIMC simulations, for the two higher temperatures (for the lower temperatures, this extrapolation is not possible).

FIG. 23.

Four isotherms of the degree of ionization α of hydrogen vs density or rs. Comparison of RPIMC results of Ref. 189 and fermionic PIMC results of Ref. 115. Thin lines indicate the expected connection of the FPIMC data to the Mott density, rs=1.2, where the FSP prohibits FPIMC simulations, for the two higher temperatures (for the lower temperatures, this extrapolation is not possible).

Close modal

In conclusion, we stress again that the degree of ionization and the fractions of atoms and molecules are not physical observables. They cannot be rigorously computed, even by a first principles simulation. Any result depends on the used criterion, as discussed above, even though the sensitivity to the chosen criterion can be verified and minimized. For completeness, we mention that there exist other approaches use different criteria, in particular, dynamical quantities to estimate the degree of ionization. For example, it was suggested to use the plasma frequency as an observable.645 Furthermore, in Ref. 327 it was demonstrated for carbon, how the degree of ionization can be derived from the dynamic conductivity in a DFT simulation, using the Kubo–Greenwood formula (Sec. IV E 2) and the Thomas-Reiche-Kuhn sum rule.

1. Overview

The problem of a Coulomb bound state in a plasma medium has been studied for a long time. Physically, one expects that the ionization energy of a bound state will be reduced by the surrounding medium, giving rise to an effective ionization potential, Iioneff(n,T), that depends on density and temperature. Early works in that field are due to Rompe and Steenbeck,646 as well as Ecker and Kröll, who considered non-degenerate electrons647 and computed the “ionization potential depression” (IPD), i.e., the medium-induced lowering of the ionization potential. An improved model of IPD is due to Stewart and Pyatt.322 We also mention the computational analysis of Rogers et al.173 who solved the Schrödinger equation with a screened Coulomb potential and computed the reduction of the binding energy in dependence on the screening length. However, this does not fully describe the situation in a dense hydrogen plasma. A more comprehensive description is achieved within the Bethe–Salpeter equation (cf. Sec. III H 3), which allows one to take into account dynamical screening, quantum, and Pauli blocking effects (e.g., Refs. 68, 648, and 649). Recent work by Massacrier et al.650 developed a new approach to describe the IPD that treats bound and free electronic states under a consistent set of assumptions and thereby removes the discrepancies in dense plasmas between first principles calculations and average atom models that were reported in Ref. 462. We also mention KS-DFT simulations of Hu651 for the shift of the K-edge in a strongly coupled fully degenerate carbon plasma that indicates significant differences from existing IPD models.

Before proceeding with a discussion of IPD, let us comment on the analogies with the insulator–metal transition in the fluid phase. While, in the gas phase, the pressure-induced transition from a non-conducting atomic (or molecular) phase to a conducting (fully ionized) plasma proceeds via the vanishing of the atomic binding energy (IPD), in the condensed phase, compression leads to bandgap closure, resulting in a metallic state with delocalized electrons, for results (see Sec. IV C 3).

2. First principles QMC results for the IPD

Based on the recently obtained first-principles QMC results for hydrogen,115 here we present a novel approach to the ionization potential depression. The idea is to use, as input, the fractions of free electrons, α, and atoms, xA, in a partially ionized hydrogen plasma which depend on the effective binding energy of the electrons. However, what is not known is how different atomic (and molecular) orbitals are affected by the interactions between the particles and by quantum effects upon compression, i.e., spectral information is missing which is not available from the QMC simulations.

To obtain this information, it is necessary to resort to alternative many-body theories. As was discussed in Sec. III B 2, it is known from the solution of the Bethe–Salpeter equation that all atomic levels are down-shifted in a plasma,68,648,649,652 i.e., En0Enl(ne,T)=E1s0n2+Δnl(ne,T), with Δnl(ne,T)0, and, simultaneously, the continuum edge (the zero energy level) shifts down as well, 0ΔI0(n,T)0. Thus, the ionization potentials are renormalized according to
(106)
(107)
Upon density increase, the ionization potentials vanish one by one, starting with the highest excited states, i.e., the bound states merge into the continuum and disappear, i.e., electrons become unbound, as we demonstrate in the figures below. Vanishing of the ground state happens at the Mott density, I1eff(neMott,T)0 corresponding to rsMott1.2, Eq. (27).
Our first principles approach to the renormalization of the binding energies is then to solve the Saha equation (52) for a single unknown quantity—the continuum lowering, ΔI0, for input values of α and xA that are provided by a QMC simulation. Our starting point is the Saha equation for ionization and recombination, Eq. (52) which we bring to the classical form and rewrite by introducing the Planck–Larkin mass action constant, KAPL, details are given in Ref. 653,
(108)
(109)
(110)
(111)
(112)
where ΔIF denotes an additional energy shift, relative to the low-density limit, that is due to the Fermi statistics, defined by Eq. (55). Also, with the definition (110), we enforce that bound states vanish from the bound state partition function when their binding energy vanishes.

We solve the Saha equation (108) iteratively for ΔI0 until the r.h.s. matches the given ratio xA/α2, on the l.h.s.

Fermionic PIMC results for the continuum edge are shown in Figs. 24 and 25, for two approximations for the level renormalization: the first (shown by the triangles and denoted “FPIMC1”) entirely neglects the shift of the bound states, ΔEnl0, and only retains the continuum shift, ΔI0. The second takes into account the renormalization of the bound states in a consistent way, using approximate results from the solution of the Schrödinger equation with a Yukawa potential given in Ref. 654. Even though this neglects, among others, dynamic screening and spin statistics effect, this approximation should be appropriate for almost the entire density range for which the FPIMC results of Ref. 115 are available. For the screening parameter, we take into account quantum effects. The results for several effective bound state energies, Enl(ne,T), used in approach 2 are also shown in Figs. 24 and 25 by the red dotted line, for the ground state, and the lowest excited states, in the insets. Vertical gray dashed lines indicate the free electron densities where the (renormalized) 4s, 3s, and 2s levels vanish.

FIG. 24.

Shift of the continuum edge, ΔI0, vs free electron density ne*=αne (lower x-axis) and rs* [Eq. (22), upper x-axis], for temperatures T= 62 500 and 125 000 K. Lines with symbols: FPIMC results without (FPIMC1) and with (FPIMC2) renormalization of the bound state energies.654 SP and EK denote the models of Stewart and Pyatt, and Ecker and Kröll, respectively. μex denotes the interaction part of the chemical potentials computed within the chemical model of Ref. 115. Insets show the renormalization of the 2s, 2p, 3s, 3p, 4s, and 4p states. Vertical lines indicate the density where the 4s, 3s, and 2s levels merge into the continuum, according to the FPIMC2 simulation data. This happens around rs*=13.59,8.96, and 4.86, for T= 62 500 K, and rs*=10.93,7.26, and 4.0, for T= 125 000 K, respectively.

FIG. 24.

Shift of the continuum edge, ΔI0, vs free electron density ne*=αne (lower x-axis) and rs* [Eq. (22), upper x-axis], for temperatures T= 62 500 and 125 000 K. Lines with symbols: FPIMC results without (FPIMC1) and with (FPIMC2) renormalization of the bound state energies.654 SP and EK denote the models of Stewart and Pyatt, and Ecker and Kröll, respectively. μex denotes the interaction part of the chemical potentials computed within the chemical model of Ref. 115. Insets show the renormalization of the 2s, 2p, 3s, 3p, 4s, and 4p states. Vertical lines indicate the density where the 4s, 3s, and 2s levels merge into the continuum, according to the FPIMC2 simulation data. This happens around rs*=13.59,8.96, and 4.86, for T= 62 500 K, and rs*=10.93,7.26, and 4.0, for T= 125 000 K, respectively.

Close modal
FIG. 25.

Same as Fig. 24, but for T= 15 625 and 31 250 K, respectively. The 4s, 3s, and 2s levels vanish at rs*=20.71,13.48, and 5.93, for 15 625 K, and at rs*=16.81,11.01, and 5.61, for 31 250 K, respectively. These estimates are based on the FPIMC2 data, where available, and on the μex data, otherwise.

FIG. 25.

Same as Fig. 24, but for T= 15 625 and 31 250 K, respectively. The 4s, 3s, and 2s levels vanish at rs*=20.71,13.48, and 5.93, for 15 625 K, and at rs*=16.81,11.01, and 5.61, for 31 250 K, respectively. These estimates are based on the FPIMC2 data, where available, and on the μex data, otherwise.

Close modal

Comparing the two approximations, we observe that both are in good agreement at low free electron densities, however, for ne* 1021 cm−3 differences increase rapidly. The renormalization of the bound state energies (FPIMC2) yields a substantially larger lowering of the continuum edge. Our first-principles results also allow us to benchmark other models. While the Ecker-Kröll model (EK)647 is in reasonable agreement for low temperatures and up to moderate densities, the Stewart-Pyatt model (SP)322 significantly underestimates the continuum lowering for most of the studied parameters.

Finally, we present in Fig. 26 the effective ionization potential of the hydrogen ground state, i.e., the energy distance from the renormalized 1s-state energy to the lowered continuum, as a function of the total density parameter rs. Since the level shifts (Figs. 24 and 25) depend on the free electron density, the transition from rs* to rs invokes the FPIMC results for the degree of ionization, α, which is reproduced, for reference, in the top part of the figure.

FIG. 26.

The effective ionization energy of the ground state, I1eff (bottom) and a fraction of free electrons (α, dots) and atoms (full line), cf. top figure, vs total density parameter, rs=rs*α1/3, for 31 250 and 15 625 K (left) and for 62 500 K and 125 000 (right). Squares (triangles): FPIMC2 (chemical model) results using shifts of the continuum and ground state, as shown in Figs. 24 and 25. Dotted line: extrapolation of the FPIMC2 results to the Mott density.

FIG. 26.

The effective ionization energy of the ground state, I1eff (bottom) and a fraction of free electrons (α, dots) and atoms (full line), cf. top figure, vs total density parameter, rs=rs*α1/3, for 31 250 and 15 625 K (left) and for 62 500 K and 125 000 (right). Squares (triangles): FPIMC2 (chemical model) results using shifts of the continuum and ground state, as shown in Figs. 24 and 25. Dotted line: extrapolation of the FPIMC2 results to the Mott density.

Close modal

The ionization potential I1eff remains close to 1ER for low densities, up to approximately rs20. In this parameter range (which is only of minor interest) the solution procedure sometimes does not converge (for FPIMC1) or it yields small fluctuations of the FPIMC2 results for the continuum edge (see the curves for T= 125 000 K in Fig. 24), which leads to reduced accuracy of I1eff. This behavior is caused by the very low fraction xA of atoms [cf. Fig. 26(c)] where statistical errors become important.

For higher densities rs20, I1eff starts to decrease monotonically, where the decrease becomes faster when the temperature increases. Comparing the two theoretical concepts where the bound state level shifts are neglected (FPIMC1) or taken into account (FPIMC2), we observe only small differences for the ground state ionization potential, in the studied parameter range. Therefore, we only show FPIMC2 data and compare them to the results for the interaction part of the chemical potential, μex, from the chemical model that is shown in Figs. 24 and 25. Interestingly, even though the FPIMC simulations are strongly hampered by the FSP and reach pressure ionization only for the two highest temperatures [cf. part (c) of Fig. 26], the smooth behavior of I1eff [cf. parts (b) and (d) of the figure] allow us to extrapolate the curves to zero (cf. the dotted lines), i.e., to the Mott density, except for the lowest temperature.

To summarize this section, we presented a novel first principles approach to plasma-induced renormalization of hydrogen bound states that is quite general and equally applies to other materials and other theoretical methods, including RPIMC and DFT. The approach uses first principles PIMC data for the degree of ionization, α, as an input. Even though the definition of α depends on the chosen criterion, the influence of the chosen procedure can be easily tested and quantified. The results indicated that it is important to accurately and consistently include the bound state level shifts (FPIMC2 results) which have to be provided as external input. Improved solutions to the two-particle problem in a medium, in particular at high density, will allow for further improvement of the IPD results.

The momentum distribution of electrons and protons in hydrogen is of crucial importance for many properties of the system. This includes impact excitation and ionization rates, chemical reaction rates, and fusion rates. In particular, it has been argued that the interplay of Coulomb correlations and quantum effects gives rise to an enhancement of fusion rates in dense plasmas, as compared to classical plasmas (e.g., Refs. 261, 262, and 655). In fact, in classical plasmas, the momentum distribution, n(k), always exhibits an exponential tail, in the high momentum limit, independently of the interaction strength, which is a consequence of the factorization of coordinate and momentum dependent terms in the classical partition function. In contrast, in quantum systems, an exponential tail is only observed in the ideal gas limit (Bose or Fermi function). In the presence of interactions however, non-commutation of kinetic and interaction energy gives rise to a non-exponential power law decay of n(k). This means that an analysis of the momentum tail allows for a direct access to correlation effects in the plasma, in particular, to the so-called “on top pair distribution”, g(0)=limr0g(r)—the pair distribution of spin up and spin down particles at zero separation.

Recently accurate results have been obtained for n(k) and g(0) of jellium which we discuss in Sec. IV C. After this, we briefly discuss the situation in hydrogen.

1. Jellium

The existence of a non-exponential tail of n(k)k8 for jellium in the ground state was demonstrated by Daniel and Vosko.657 Similar analytical predictions were due to Galitskii et al.658 and Kimball,659 whereas Yasuhara et al. established a direct relation between n(k) and g(0), for T=0,660 
(113)
that was extended to finite temperatures by Hofmann et al.661 These are important results which, however, refer only to the occupation of the very high momentum states but do not contain information on the overall shape of n(k) in a correlated Fermi gas, which requires simulations.

For the ground state, and moderate to strong coupling, at 1rs10, the entire momentum distribution has been computed by Holzmann et al.662 by reptation quantum Monte Carlo simulations. They particularly investigated the region around the Fermi momentum and determined the quasiparticle renormalization factor.

For finite temperatures, Θ1, exact CPIMC results were presented recently by Hunger et al.,447 whereas Dornheim et al. presented direct PIMC results.269,429 In Fig. 27, we show CPIMC results for the momentum distribution, n(k), for a weakly coupled moderately degenerate electron gas and compare to the ground state and to an ideal Fermi gas. The figure clearly demonstrates the algebraic decay of the correlated momentum distribution (full lines) and its deviation from the exponential tail of the ideal system (dashed lines). The k8 asymptotic is clearly confirmed in Fig. 28 and, in addition, the temperature and density dependence of the pre-factor [i.e., of g(0)] has been established in Ref. 447. Note that, for rs1 and temperatures above 0.5TF, CPIMC proves to be highly efficient, being able to resolve n(k) with an accuracy of ten decimal digits. Figure 28 also indicates that the power law asymptotic is reached only for large momenta where the occupation is already very low.

FIG. 27.

Nonexponential decay of the momentum distribution of moderately correlated and weakly degenerate electrons, rs=0.5 and T/TF=1,5,2.0,4.0. CPIMC results with N=54 particles are compared to the ground state (solid black, data of Ref. 656). For comparison, the ideal Fermi distribution is shown by dashed lines of the same color as the interacting result. Reproduced from Ref. 447 with the permission of the authors.

FIG. 27.

Nonexponential decay of the momentum distribution of moderately correlated and weakly degenerate electrons, rs=0.5 and T/TF=1,5,2.0,4.0. CPIMC results with N=54 particles are compared to the ground state (solid black, data of Ref. 656). For comparison, the ideal Fermi distribution is shown by dashed lines of the same color as the interacting result. Reproduced from Ref. 447 with the permission of the authors.

Close modal
FIG. 28.

Tail region of the momentum distribution for rs=0.5 and Θ=2. Bottom figure shows the deviation from the asymptotic, δ(k)=n(k)n(k)1. The k8 asymptotic is reached for k10kF with a prefactor that is provided by the CPIMC simulations. Reproduced from Ref. 447 with the permission of the authors.

FIG. 28.

Tail region of the momentum distribution for rs=0.5 and Θ=2. Bottom figure shows the deviation from the asymptotic, δ(k)=n(k)n(k)1. The k8 asymptotic is reached for k10kF with a prefactor that is provided by the CPIMC simulations. Reproduced from Ref. 447 with the permission of the authors.

Close modal

In Fig. 29, we show complementary direct PIMC results (red circles) for n(k) of a spin-polarized UEG with N=33 electrons at rs=4 and Θ=1. We note the larger occupation for kkF of the PIMC results compared to the ideal Fermi distribution (dashed black). This effect has first been reported by Militzer and Pollock663 and is connected to a lowering of the kinetic energy due to XC-effects. The green crosses show restricted PIMC results for the same conditions; they are in qualitative agreement with the exact PIMC dataset but are systematically too high for all k. Rescaling the RPIMC dataset with a constant factor results in the yellow crosses, which are in good agreement with the direct PIMC data over the entire momentum range. The observed discrepancy between PIMC and RPIMC has thus been attributed to a systematic error in the determination of the proper normalization in the latter implementation, whereas the direct PIMC results have been normalized automatically based on the extended ensemble approach introduced in Ref. 429. At the same time, we note that no nodal error was found at these conditions.269,429 A similar comparison for warm dense hydrogen constitutes an important topic for future works. Finally, we include ground-state QMC results for the unpolarized UEG662,664 at the same density by the dotted blue curve. We note the substantially rounded edges; these features would be absent for an ideal Fermi gas where the ground-state momentum distribution is given by a simple step function and, therefore, originates from correlation effects due to the interacting electrons.

FIG. 29.

Momentum distribution n(k) of the spin-polarized UEG at rs=4 and Θ=1 (T= 3.13 eV) for N = 33 electrons. Red circles: direct PIMC; green crosses: RPIMC;83 yellow crosses: RPIMC with a modified normalization constant; dotted blue: ground-state QMC results662 for the unpolarized UEG at the same density. Note that the Fermi momentum for both polarizations is different. Adapted from Ref. 269.

FIG. 29.

Momentum distribution n(k) of the spin-polarized UEG at rs=4 and Θ=1 (T= 3.13 eV) for N = 33 electrons. Red circles: direct PIMC; green crosses: RPIMC;83 yellow crosses: RPIMC with a modified normalization constant; dotted blue: ground-state QMC results662 for the unpolarized UEG at the same density. Note that the Fermi momentum for both polarizations is different. Adapted from Ref. 269.

Close modal

2. Hydrogen

By lowering the temperature below the electronic degeneracy temperature, TF, the momentum distribution of jellium becomes steeper at the Fermi momentum, developing a discontinuity in the limit of Θ0. Thermal broadening of the jump of the Fermi–Dirac occupation function is of order Θ, typically beyond the experimental resolutions of synchrotron light-scattering experiments at ambient temperatures,664–666 in the regime where the Born–Oppenheimer approximation is expected to be applicable.

Coulomb interactions between electrons lead to the deviation from the ideal gas step-function at zero temperature, reducing the magnitude of the discontinuity at kF, illustrated in Fig. 29 for rs=4, corresponding to the density of valence electrons in crystalline metallic sodium at ambient pressure and temperature. However, anisotropies in the Fermi surface, e.g., due to band structure effects, smear out the jump in n(k) when spherically averaged.664 For an insulator, the electronic momentum distribution is expected to remain continuous in any direction even at zero temperature.

The electronic momentum distribution, n(k), of hydrogen has been computed by Pierleoni et al. using CEIMC simulations,197,498 for details on the method (cf. Sec. III F). In Fig. 30 we show results for n(k) at T= 1200 K, for two densities in the range of the LLPT (cf. Sec. II B 2). For rs=1.44 the system is in the molecular phase, whereas molecules are mostly dissociated at rs=1.34 (see Fig. 20) and the corresponding discussion in Sec. IV A 6. We, therefore, expect a noticeable change in the momentum distribution reflecting metallic behavior with a sharp Fermi surface on the atomic side. Indeed, the width of the distribution around kF is reduced by about a factor of two. This change occurs abruptly at the LLPT transition, whereas further density changes on the atomic or molecular side are much smoother.498 The broadening of n(k) around kF, in the atomic phase, is due to anisotropies of the Fermi surface of the snapshots of different nuclear configurations in the Born–Oppenheimer approximation.

FIG. 30.

Electronic momentum distribution of hydrogen at two densities across the LLPT at T= 1200 K, computed using a Slater–Jastrow backflow trial wavefunction with configurations from CEIMC. In the molecular phase (blue) the single-electron density matrix, n(r), decreases exponentially with r, while in the dissociated phase (red) it decreases as r3 because the electrons are delocalized.

FIG. 30.

Electronic momentum distribution of hydrogen at two densities across the LLPT at T= 1200 K, computed using a Slater–Jastrow backflow trial wavefunction with configurations from CEIMC. In the molecular phase (blue) the single-electron density matrix, n(r), decreases exponentially with r, while in the dissociated phase (red) it decreases as r3 because the electrons are delocalized.

Close modal
That these changes of the momentum distribution from rs=1.44 to rs=1.34 indicate a phase transition from insulating to metallic behavior is further supported by analyzing the long distance behavior of the reduced single-particle density matrix n(r), the Fourier transform of n(k), shown in Fig. 30. Within the Born–Oppenheimer approximation, we have n(r)=n(r|R¯)R¯, with R¯ indicating the averaging over the nuclear configurations of
(114)
where Ψ0()Ψ0(|R¯) is the Born–Oppenheimer electronic wave function which parametrically depends on the nuclear configuration R¯.

From Fig. 30 we see that the envelope of the asymptotic long-range behavior changes from an exponential behavior to a power law, at the LLPT, confirming the change from the localized character of the molecular liquid to a metallic state with a Fermi-liquid-type r3 decay, in the atomic phase.

3. CEIMC results for the gap

Within CEIMC, it is further possible to calculate the fundamental electronic gap in the insulating phase,218,499,500 the energy gap corresponding to electron/hole doping. Assuming a strictly uniform background charge to assure charge neutrality, the change of the free energy with respect to addition/removal of electrons reduces to different electronic Born–Oppenheimer excitation energies, averaged over nuclear configurations. Since the additional electronic charge is necessarily smeared out over the entire system, due to the charge-compensating homogeneous background, modifications of the nuclear states can be neglected in the thermodynamic limit.218 In the crystalline state, electron addition and removal energies can be characterized by the symmetries of the crystal structure, defined by the average charge density, even in the presence of zero point and thermal nuclear motion502 within the region of validity of the BO approximation.

Experimentally, the electronic excitation spectrum is probed by optical absorption corresponding to charge-neutral particle-hole excitations which can also be addressed with QMC-based methods.667 Electron–hole binding and excitonic effects may lower the gap compared to the fundamental gap. Those effects are particularly important at lower pressures where excitons are strongly bound.501 In addition, the localization of e-h excitations may also affect the BO energy surface of the nuclei in the vicinity of the excitons. Neutral electronic excitations averaged over nuclear trajectories of the ground state BO energy surface remains an upper bound of the neutral excitation gap.

CEIMC calculations of the closing of the fundamental gap in solid hydrogen218 predicted that for ∼370– 380GPa an indirect gap closure occurs at 200K, for C2/c-24, and at 340GPa, for Cmca-12, whereas the direct gap remains open until ∼450 and 500GPa, respectively, suggesting a bad metal (or semimetal) scenario, qualitatively similar to experimental observations.204,668–670 Predictions for the direct gap are in agreement with absorption measurements24,671 with a strong reduction of 2eV due to nuclear quantum effects, slightly larger than the experimental extrapolation223 of 1.5eV from D2 to H2 assuming a m1/2 isotope effect. Experiments have been performed at 80 K so that the experimental value might be slightly biased by residual finite temperature shifts.

The gap closure of the molecular liquid has been studied at three different temperatures around the LLPT.499  Figure 31 shows that, below the critical temperature, the gap closure seems to occur abruptly, together with the molecular to atomic transition, whereas a smoother closure is observed at T= 3000 K, supporting the existence of a crossover between the two phases above the critical temperature. This picture is also supported by the electronic density of states, an example of which is reported in Fig. 32. Experimental determinations of the gap are also reported in Fig. 31 and are in reasonable agreement with QMC results, suggesting that the temperatures reached in the shock-wave experiments are in the same range as in the simulations.

FIG. 31.

The fundamental energy gap of liquid hydrogen along the isotherms: T = 900, 1500, and 3000 K, as a function of pressure. Inset: the same gap as a function of rs, a measure of density [Eq. (1)]. The lines connect the gap data only up to the molecular-atomic transition region. The colored rectangles show the coexistence region of the LLPT according to Ref. 208. The dotted lines are the gaps reported by Cellier et al.212 The brown and green squares are the results of Nellis et al. for temperatures of 2000–3000 K672 reanalyzed in Ref. 673. The red dot is the gap reported by McWilliams et al. at 2400 K.674 Adapted from Ref. 499.

FIG. 31.

The fundamental energy gap of liquid hydrogen along the isotherms: T = 900, 1500, and 3000 K, as a function of pressure. Inset: the same gap as a function of rs, a measure of density [Eq. (1)]. The lines connect the gap data only up to the molecular-atomic transition region. The colored rectangles show the coexistence region of the LLPT according to Ref. 208. The dotted lines are the gaps reported by Cellier et al.212 The brown and green squares are the results of Nellis et al. for temperatures of 2000–3000 K672 reanalyzed in Ref. 673. The red dot is the gap reported by McWilliams et al. at 2400 K.674 Adapted from Ref. 499.

Close modal
FIG. 32.

Density of states (DOS) of liquid hydrogen near the band edge at densities near the gap closure for the isotherm T= 1500 K. The inset shows the equation of state, as reported in Ref. 208. The dashed and solid red lines indicate the atomic and molecular regions, respectively. The colors of the DOS match the colors of points in the insets. Adapted from Ref. 499.

FIG. 32.

Density of states (DOS) of liquid hydrogen near the band edge at densities near the gap closure for the isotherm T= 1500 K. The inset shows the equation of state, as reported in Ref. 208. The dashed and solid red lines indicate the atomic and molecular regions, respectively. The colors of the DOS match the colors of points in the insets. Adapted from Ref. 499.

Close modal

To characterize the Fermi liquid behavior of metallic hydrogen, determination of the effective mass of single-particle excitations should be carried out. However, since the effective mass of jellium is very close to the bare electron mass,301,675 one expects the excitation spectrum to be determined by (single-particle) band structure effects to a large extent.

In this section, we will give an overview of a number of density response properties of warm dense hydrogen. As a starting point, we consider the modified Hamiltonian17,
(115)
where Ĥ is the unperturbed Hamiltonian, Eq. (35), to which we add an external monochromatic perturbation of wave vector q, frequency ω, and perturbation amplitude A. In the limit of infinitesimally weak perturbations, the induced density change with respect to the unperturbed system is described by [cf. Eq. (105)]
(116)
it is thus a linear function of the perturbation strength, which is fully described by the dynamic linear density response function χ(q,ω). We note that the latter is an equilibrium property of the unperturbed system. From a theoretical perspective, it is very convenient to express the density response as [cf. Eq. (102)]
(117)
where χ0(q,ω) denotes a known reference function. For the UEG, the latter is typically given by the temperature-dependent Lindhard function that describes the density response of an ideal Fermi gas at the same conditions.419 In that case, the complete information about electronic XC effects is included in the dynamic exchange–correlation kernel Kxc(q,ω), which is related to the dynamic local field correction (LFC)676 by
(118)
Indeed, setting Kxc(q,ω)0 in Eq. (117) corresponds to the well-known random phase approximation (RPA) describing the electronic density response on a mean-field level. Unsurprisingly, G(q,ω) of the UEG constitutes key input for a host of practical applications such as the construction of nonlocal XC-functionals for DFT simulations677–679 or the interpretation of XRTS experiments.146,319,680 Therefore, the development of approximate expressions for G(q, ω) [or its static limit G(q, 0)] has been a highly active topic over the past decades.66,131,133,681–689 Very recently, highly accurate results have become available both in the ground state300 and at finite temperatures.143,147,155,165,166,688,690–692 Among other things, these results have revealed a roton-type feature in the spectrum of density fluctuations166,270,549 for intermediate wavenumbers, which has subsequently been predicted to be even more pronounced in warm dense hydrogen169 (see Sec. IV D 4).

For electron–ion systems such as hydrogen, the reference function in Eq. (117) is typically replaced by the Kohn–Sham response function χKS(q,ω), which depends on the utilized XC-functional, eventually leading to a Dyson's type equation [cf. Eq. (104)]. Therefore, the clear-cut interpretation of Kxc(q,ω)0 as a mean-field description is no longer valid, as both, the kernel and the reference function contain some information about electronic XC-effects. For hydrogen, first dependable results for Kxc(q,0) have been presented only recently17,141,142,693 and are discussed in Sec. IV D 1.

We also note that χ is directly related to the inverse dielectric function [cf. Eq. (100)], which describes the dynamical screening of the Coulomb interaction and also enters collision integrals of kinetic equations (cf. Sec. III H 1).

A particularly important practical application of Eq. (117) is given by the fluctuation–dissipation theorem419,
(119)
which relates the dynamic structure factor to the dynamic density response function. For example, Eq. (119) constitutes the basis of LR-TDDFT simulations of XRTS measurements.17,263,617,694–700 In practice, XRTS measurements of hydrogen are generally challenging due to its comparably small scattering cross section.257,258

1. Static density response: Snapshots

In the static limit of ω0, the electronic density response of hydrogen can be studied by carrying out a set of harmonically perturbed calculations that are governed by the Hamiltonian (115) for a set of fixed proton coordinates. Fully integrated PIMC simulations where both the electrons and protons are treated dynamically on the same level are discussed in detail in Sec. IV D 2. Still, using PIMC to solve the electronic problem in the fixed external proton potential allows for direct comparisons with DFT simulations, which is interesting in its own right.

Following this idea, Böhme et al.141 have presented the first accurate results for the static density response function, χ(q)=χ(q,0) [and, in this way, also the corresponding static XC-kernel Kxc(q)]. The results are shown in Fig. 33 as the black squares for Θ=1, with the top and bottom panels corresponding to rs=2 and rs=4, respectively. In addition, we also show the exact static density response of the UEG145 (solid black), as well as the RPA result for the latter (dashed black) at the same conditions. For the higher density, most electrons are strongly delocalized throughout the system, and the density response of hydrogen closely follows the density response of the UEG over the entire wavenumber range. In stark contrast, the electrons are strongly localized around the protons for rs=4, and the overall density response of hydrogen is substantially reduced compared to the UEG model at the same conditions. Assuming a decomposition into effectively free and bound populations of electrons where only the former respond to the external perturbation, Böhme et al.141 have reported a free-electron fraction (degree of ionization) of α=0.6. This is in close agreement with the value of α=0.54 reported in the RPIMC simulations of Militzer and Ceperley,189 whereas more recent simulations by Filinov and Bonitz115 find a substantially lower value, α0.3 (cf. Fig. 23).

FIG. 33.

Comparing PIMC and DFT results for the static linear density response function χ(q). For details, see text. Taken from Moldabekov et al.693 with the permisson of the authors.

FIG. 33.

Comparing PIMC and DFT results for the static linear density response function χ(q). For details, see text. Taken from Moldabekov et al.693 with the permisson of the authors.

Close modal

Subsequently, Modabekov et al.693 have used the same setup to perform KS-DFT simulations of hydrogen, and the results for different XC-functionals are included as the colored symbols in Fig. 33, for both densities. For rs=2 all depicted functionals work very well, over the entire depicted q-range, although, in particular, the SCAN functional overestimates the true density response for q>2qF. In addition, the gray circles that have been computed without any XC-functional (“NullXC”) are in close agreement with the RPA result for the UEG. For rs=4, the situation is more complex. First and foremost, we find that here, too, all DFT results are in qualitative agreement with the exact PIMC reference data for all wavenumbers. Interestingly, DFT overestimates the magnitude of the true density response for q2qF independent of the utilized functional. This is likely a consequence of self-interaction effects,701 which lead to an underestimation of the localization of the electrons around the ions. Since the latter depletes the ability of effectively bound electrons to react to the external perturbation, DFT does not capture the full reduction of χ(q) compared to the free UEG model. Second, we mention the increase in the density response compared to the UEG for the largest depicted value of q. This has been explained in terms of isotropy breaking at certain length scales in the original Ref. 141, and is again captured qualitatively by DFT.

In essence, by implementing the direct perturbation approach into DFT, Moldabekov et al.693 have demonstrated an alternative way to obtain the static XC-kernel KXC(q) within DFT itself without the need for numerical derivatives of the XC-functional. In particular, this method is straightforward even for hybrid functionals on higher rungs on Jacob's ladder of functional approximations,388 for which numerical derivatives are known to be unstable. A particularly interesting application of Kxc(q) is given by LR-TDDFT simulations to compute the dynamic structure factor S(q,ω), which is discussed briefly in Sec. IV F 1.

2. Imaginary-time correlation functions (ITCF)

A well-known bottleneck of the first principles PIMC method is its restriction to the imaginary-time domain (see Sec. III E). While dynamic PIMC simulations are, in principle, possible,423 the complex exponent of the time-evolution operator would lead to an additional phase problem, limiting its application to ultra-short time scales.

A possible way to circumvent this limitation is to consider imaginary-time correlation functions, which can be easily computed within PIMC. A particularly important example is the imaginary-time density–density correlation function (ITCF)164,702–705
(120)
corresponding to the usual intermediate scattering function F(q,t),158 but evaluated at the imaginary-time argument t=iτ. For example, F(q,τ) is directly connected to the dynamic structure factor via a two-sided Laplace transform,
(121)
The numerical inversion of Eq. (121) to compute See(q,ω) is often denoted as analytic continuation in the literature,280 and constitutes a difficult, ill-conditioned inverse problem. Although several methods exist (e.g., Refs. 430 and 706–711), the reliability of these methods is difficult to judge in the absence of additional constraints. A benign exception is given by the UEG, where it has been possible to sufficiently constrain the space of possible trial spectra by imposing a number of exactly known conditions.165,166 The generalization of these ideas to warm dense hydrogen is a topic of ongoing efforts and will be covered in dedicated future works.
Very recently, Dornheim et al.164 have argued that, due to the formally unique mapping inherent to Eq. (121), often no analytic continuation is necessary to understand the physical mechanisms in a given system of interest. In fact, both See(q,ω) and Fee(q,τ) contain the same information, but in different representations. Surprisingly, considering Fee(q,τ) might even be more practical for the interpretation of experimental XRTS measurements due to the well-known convolution theorem,160,
(122)
where R(ω) denotes the experimental source function. In other words, it is straightforward—and very robust with respect to experimental noise161,162—to obtain the ITCF from an XRTS measurement, given accurate knowledge for R(ω), whereas the corresponding deconvolution of Eq. (122) to get See(q,ω) is notoriously unstable. The task at hand is thus to understand the physical manifestation of effects such as temperature, density, or Coulomb coupling in the imaginary-time domain.164 
For example, the detailed balance relation
(123)
provides a direct relation between the DSF at positive and negative frequencies in thermal equilibrium, which exclusively depends on the temperature.326,712 Unfortunately, its direct application for temperature diagnostics is generally prevented by the convolution with the source function. Yet, it is straightforward to translate [Eq. (123)] into the τ%-domain, resulting in the symmetry relation17,160,161
(124)
In other words, F(q,τ) is symmetric around τ=β/2=1/(2kBT), where it attains a minimum. This allows one to directly extract the temperature of an XRTS measurement without the need for simulations and approximations.160 It can thus be expected that Eq. (124) will give direct insights into upcoming experiments with hydrogen jets,257,258 and potential ambitious spherical implosion experiments at the NIF.256 
In Fig. 34, we show first principles PIMC results for the ITCF of hydrogen for N=32 atoms at rs=2 and Θ=1 taken from Ref. 148. Clearly, Fee(q,τ) displays a rich structure, and its main features can be summarized as follows. For τ=0, the ITCF is equal to the static structure factor, Fee(q,0)=See(q); the latter is, by definition, the normalization of See(q,ω). The decay of Fee(q,τ) with τ becomes increasingly steep for large wavenumbers. It is a direct consequence of the quantum delocalization of the electrons, which is more pronounced on shorter length scales λ=2π/q; it would be entirely absent in a classical system and is substantially less pronounced for the heavier protons (cf. Fig. 35). Interestingly, the main trends of Fee(q,τ) can be explained by a simplified single-particle imaginary-time Gaussian diffusion model.178 To get additional insights into the behavior of Fee(q,τ), we consider the frequency moments of See(q,ω), which are defined as17 
(125)
FIG. 34.

First principles PIMC results for the imaginary-time density–density correlation function Fee(q,τ) of hydrogen with N=32 atoms at rs=2 and Θ=1 shown in the q- τ-plane. Taken from Ref. 148 with the permission of the authors.

FIG. 34.

First principles PIMC results for the imaginary-time density–density correlation function Fee(q,τ) of hydrogen with N=32 atoms at rs=2 and Θ=1 shown in the q- τ-plane. Taken from Ref. 148 with the permission of the authors.

Close modal
FIG. 35.

First principles PIMC results for various ITCFs of hydrogen at rs=2 and Θ=1 for q= 1.53  Å1 (top) and q= 7.65  Å1 (bottom); solid red: Fee(q,τ), dashed green: Fpp(q,τ), dotted blue: Fep(q,τ), double-dashed yellow: UEG.164 The shaded intervals correspond to 1σ error bars. The insets in the lower panel show magnified segments around zero of Fep(q,τ) and Fpp(q,τ). Taken from Ref. 148 with the permission of the authors.

FIG. 35.

First principles PIMC results for various ITCFs of hydrogen at rs=2 and Θ=1 for q= 1.53  Å1 (top) and q= 7.65  Å1 (bottom); solid red: Fee(q,τ), dashed green: Fpp(q,τ), dotted blue: Fep(q,τ), double-dashed yellow: UEG.164 The shaded intervals correspond to 1σ error bars. The insets in the lower panel show magnified segments around zero of Fep(q,τ) and Fpp(q,τ). Taken from Ref. 148 with the permission of the authors.

Close modal
It is easy to see that the MαS can alternatively be computed from the ITCF via17,164,713
(126)
i.e., as its derivatives with respect to τ around τ=0. This relation has recently been used to extract the first five moments of the DSF from ab initio PIMC results for the warm dense UEG over a broad range of wave numbers.713 In addition to being interesting in their own right, accurate knowledge of various MαS is important to constrain the aforementioned analytic continuation [cf. Eq. (121)];703,714 the application of such methods to warm dense hydrogen based on highly accurate PIMC input data will be explored in dedicated future works. A particularly important example is given by the universal f-sum rule148,419
(127)
It implies that the slope of F(q,τ) around the origin increases quadratically in magnitude with the wave number q. In practice, Eq. (127) is also very useful to obtain the proper normalization of a measured XRTS signal, which is a-priory unknown,162 but strictly required to use the ITCF for other purpose, such as the study of Electron–electron correlations based on the static structure factor See(q)=Fee(q,0), or for the evaluation of the imaginary–time version of the fluctuation–dissipation theorem [Eq. (129)].
In Fig. 35, we show various species-resolved ITCFs as a function of τ for two selected wave numbers. The double-dashed yellow lines correspond to the UEG146,147 and have been included as a reference. For q= 1.53 Å−1, Fee(q,τ) of hydrogen (solid red) is substantially shifted upwards compared to the UEG at the same conditions, whereas this shift is nearly entirely absent for q=7.65 Å1. To understand this observation, we decompose the full dynamic structure factor into (quasi-)elastic and inelastic contributions,159,316
(128)
The former is conventionally attributed to bound electrons and an additional screening cloud of free electrons that follows the protons, whereas the latter contains bound-free (and its reverse, free-bound715) and free-free transitions within the widely employed Chihara picture.158,312,716 The spectral weight of the elastic feature is given by the Rayleigh weight WR(q)=Sep2(q)/Spp(q), which constitutes a direct measure for the degree of electronic localization around the ions. Clearly, the elastic contribution to the full DSF manifests in the ITCF as a constant shift, which explains the observed differences between the UEG and hydrogen in Fig. 35.

The dashed green and dotted blue curves correspond to Fpp(q,τ) and Fep(q,τ), which appear to be constant on the depicted scale. The insets in the bottom panel show magnified plots for the larger wave number. Evidently, no τ-dependence can be observed for Fep(q,τ) within the given uncertainty interval. This is unsurprising as Eq. (127) implies that Fep(q,τ) is constant in linear order. In contrast, we find a small yet significant τ-decay for Fpp(q,τ); its reduced magnitude compared to the electrons directly follows from the heavier mass.

3. Static density response of hydrogen

A key application of the ITCF is given by the imaginary-time version of the fluctuation–dissipation theorem,164,717
(129)
In practice, it is thus possible to obtain the full q-dependence of the static density response from a single PIMC simulation of the unperturbed system. This is a considerable advantage over the direct perturbation approach governed by the modified Hamiltonian Eq. (115). For completeness, we note that similar relations exist for various nonlinear response functions.17,704,718

In Fig. 36, we show the species-resolved static density response of hydrogen, for rs=2 and Θ=1, i.e., for the same conditions that were explored in Sec. IV D 2. Let us first consider the blue symbols that show PIMC results for χep(q), i.e., the response of the electrons to a perturbation of the protons (or vice versa). It monotonically decays with q, as the electronic localization around the ions becomes less important on small length scales. The green symbols show the static proton response function, χpp(q). It exhibits the opposite trend compared to Fep(q,τ), which can be understood as follows: in the limit of large q, Fpp(q,τ)Spp(q) approaches one; its small τ-decay shown in Fig. 35 is negligible. Equation (129) then implies the classical short wavelength limit, limqχppcl(q)=nβ, which explains the apparent convergence of the green symbols in Fig. 36. For smaller q, electronic screening of the effective proton–proton interaction leads to a reduction of the density response, which attains a finite value in the limit of q0.

FIG. 36.

First principles PIMC results for the species-resolved static density response of hydrogen at rs=2 and Θ=1. Red, green, and blue symbols: χee(q), χpp(q), and χep(q), for full hydrogen evaluated from the ITCF via Eq. (129); black circles: electronic density response of fixed proton snapshot (see Ref. 141); solid yellow line: machine-learning representation of χee(q) of the UEG model145 at the same conditions. Adapted from Ref. 148 with the permission of the authors.

FIG. 36.

First principles PIMC results for the species-resolved static density response of hydrogen at rs=2 and Θ=1. Red, green, and blue symbols: χee(q), χpp(q), and χep(q), for full hydrogen evaluated from the ITCF via Eq. (129); black circles: electronic density response of fixed proton snapshot (see Ref. 141); solid yellow line: machine-learning representation of χee(q) of the UEG model145 at the same conditions. Adapted from Ref. 148 with the permission of the authors.

Close modal
Let us next consider the red symbols that display the static electronic density response, χee(q), of full two-component hydrogen. In stark contrast to χpp(q), it monotonically decays for q2 Å1. This is exclusively due to the quantum delocalization of the electrons, as Fee(q,τ) vanishes increasingly fast with τ for large wave numbers. In practice, the response of the electrons vanishes when the wavelength of the external perturbation becomes comparable to their thermal wavelength λβ. In fact, the protons are subject to the same effect, but they manifest only for substantially larger q due to their heavier mass. In the limit of small q, χee(q) plateaus; a hypothetical minimum at a finite value of q is possible but cannot be resolved with the considered N=32 hydrogen atoms. To understand the rich physics that guides the electronic density response in this regime, it is convenient to compare it with other models to isolate a particular effect. The solid yellow curve shows χee(q) of the UEG145 at the same conditions. It agrees well with the red symbols, for q4 Å1, but increasingly deviates for long wavelengths. Indeed, it completely vanishes for q0, due to the perfect screening of the UEG.295 The latter breaks down for full two-component hydrogen as the unperturbed protons, on average, follow the harmonically perturbed electrons; this leads to an effective screening of the Electron–electron interaction that facilitates the larger response of hydrogen compared to the UEG on large-length scales. An alternative, though fully equivalent, way to understand this is to recall the direct relation of χee(q) to the inverse frequency moment of the DSF,706 
(130)
The static density response is thus highly sensitive to the low-frequency behavior of See(q,ω), which, for an electron–ion system, is dominated by the elastic feature [cf. Eq. (128)]. The larger density response of hydrogen is thus a direct consequence of the electronic localization around the ions described by the Rayleigh weight, WR(q). Finally, the black dots in Fig. 36 correspond to the linear density response of the electrons in a fixed proton snapshot that has been discussed in Sec. IV D 1. They are in close qualitative agreement with the UEG at these conditions, but, unsurprisingly, miss the rich interplay between the electrons and protons of the full electronic density response, i.e., the red symbols.

In addition to being interesting in their own right, highly accurate PIMC results for hydrogen, and potentially other light elements, are important for several reasons: (i) as a basis to compute the exchange–correlation kernel Kxc(q)—a key input for TDDFT,617,720 ionization potential depression models,680 and the construction of advanced XC-functionals for thermal DFT simulations;679 (ii) to benchmark dynamic simulations and widely used approximate models such as the Chihara decomposition715,716 (cf. Sec. III B 3); and (iii) as an unambiguous prediction of upcoming XRTS measurements with fusion plasmas and hydrogen jets.257,258

4. Prediction of a roton-type feature for jellium at low density

In Sec. IV D 2, we have discussed the dynamic density response of hydrogen based on PIMC results in the imaginary-time domain. For jellium, it is possible to carry out a reliable analytic continuation (cf. Refs. 165–167 and 721) to obtain highly accurate results for the DSF over a broad range of parameters. This study has revealed two important points: (1) the static approximation [Eq. (143)] is remarkably accurate in the case of the UEG for high to moderate densities and (2) the position of the maximum in See(q,ω)—here simply denoted as ω(q)—exhibits a non-monotonous dependence on the wave number q in the low-density regime, with a pronounced minimum around q2qF. Both points are illustrated in Fig. 37, where we show ω(q) at Θ=1, for rs=2 (left) and rs=10 (right). More specifically, the dotted red, dashed black, and solid green curves have been obtained based on the RPA [ Kxc(q,ω)0], static approximation [ Kxc(q,ω)Kxc(q,0)], and the full, reconstructed results for Kxc(q,ω), respectively. For both conditions, all curves exactly reproduce the collective plasmon excitation in the limit of q0. Similarly, we find the same parabolic dependence of ω(q) on q in the single-particle limit of qqF.

FIG. 37.

Position of the maximum of S(q,ω) for jellium at Θ=1 computed within the RPA (dotted red line), the static approximation [cf. Eq. (143), Ref. 166, dashed black line], and from the full analytic continuation of PIMC results165,166 (solid green line) for rs=2 (left) and rs=10 (right). The shaded gray area indicates the pair continuum.719 In all cases, the DSF exhibits the well-defined collective plasmon excitation for q0 and a parabolic single-particle dispersion for qqF. The roton feature occurs at intermediate q where the wavelength λ=2π/q is comparable to the average interparticle distance d. It is a direct consequence of the exchange–correlation induced alignment of pairs of electrons (cf. Fig. 38) and is, therefore, not captured by the RPA. Taken from Hamann et al.169 with the permission of the authors.

FIG. 37.

Position of the maximum of S(q,ω) for jellium at Θ=1 computed within the RPA (dotted red line), the static approximation [cf. Eq. (143), Ref. 166, dashed black line], and from the full analytic continuation of PIMC results165,166 (solid green line) for rs=2 (left) and rs=10 (right). The shaded gray area indicates the pair continuum.719 In all cases, the DSF exhibits the well-defined collective plasmon excitation for q0 and a parabolic single-particle dispersion for qqF. The roton feature occurs at intermediate q where the wavelength λ=2π/q is comparable to the average interparticle distance d. It is a direct consequence of the exchange–correlation induced alignment of pairs of electrons (cf. Fig. 38) and is, therefore, not captured by the RPA. Taken from Hamann et al.169 with the permission of the authors.

Close modal

Due to the role of rs as the quantum coupling parameter, electronic exchange–correlation effects are of moderate importance in the regime of metallic densities, and the systematic deviation of RPA to the other curves is comparably small; the static approximation can, in fact, hardly be distinguished from the exact results in this regime. This situation changes dramatically at rs=10, which is located at the margins of the strongly coupled electron liquid regime.688 Indeed, the exact PIMC-based results for ω(q) exhibit a pronounced minimum for intermediate wave numbers, which phenomenologically resembles the well-known roton feature of quantum liquids such as ultracold helium.414,430,722–724 The static approximation qualitatively captures this nontrivial feature, whereas it is completely absent from the mean-field based RPA results.

To understand the physical origin of the roton minimum in the UEG, Dornheim et al.549 have proposed to consider the fluctuation–dissipation theorem [cf. Eq. (119)], which implies that the DSF is fully determined by the linear density response of the system to an infinitesimal external harmonic perturbation. In Fig. 38, the latter situation is schematically illustrated, with the green bead corresponding to an arbitrarily chosen reference particle. In addition, the blue beads correspond to other particles in the system, which are, on average, disordered at rs=10 and Θ=1 [note the absence of a pronounced peak in See(q) discussed, e.g., in Ref. 549]. The dark gray cosinusoidal curve shows an external harmonic perturbation along the x-direction, where the wavelength is of the same order as the average interparticle distance, i.e., λ=2π/qd. For any value of q, the electrons are energetically incentivized to align themselves to the minima of the perturbation. For small q, where λd, this trend is suppressed by the perfect screening in the UEG.295 Conversely, the alignment is negligible for qqF, as it happens on increasingly small length scales, λd. For the situation that is depicted in Fig. 38, the alignment of the leftmost blue bead to the cosine perturbation coincides with a minimization of the interaction energy W, i.e., a negative change ΔW. It has been shown in Refs. 270 and 549 that this alignment of pairs of electrons explains the roton minimum shown in Fig. 37. Further, this shift in the interaction energy is an electronic XC-effect and, therefore, fundamentally not captured by the RPA that entails a mean-field description of the density response. We note that these findings are consistent with similar observations for jellium in the ground-state.725–727 

FIG. 38.

Schematic illustration of electronic pair alignment causing the roton feature shown in Fig. 37. The green bead shows an arbitrary reference particle, and the blue bead other electrons which are, on average, disordered. Applying an external harmonic perturbation [cf. Eq. (115), gray curve] with a wavelength λ that is commensurate to the average interparticle distance d changes the free energy landscape and reinforces the minimization of the interaction energy W. The corresponding change ΔW of the latter qualitatively explains the negative roton shift. Taken from Dornheim et al.549 

FIG. 38.

Schematic illustration of electronic pair alignment causing the roton feature shown in Fig. 37. The green bead shows an arbitrary reference particle, and the blue bead other electrons which are, on average, disordered. Applying an external harmonic perturbation [cf. Eq. (115), gray curve] with a wavelength λ that is commensurate to the average interparticle distance d changes the free energy landscape and reinforces the minimization of the interaction energy W. The corresponding change ΔW of the latter qualitatively explains the negative roton shift. Taken from Dornheim et al.549 

Close modal

While being interesting in their own right, the results discussed in this section have been obtained for the UEG model (cf. Sec. III E 1). It is thus important to ask if the roton minimum can be observed in a real system, with warm dense hydrogen being a promising candidate.169 

5. Prediction of a roton-type feature in warm dense hydrogen

Roton excitations are well-known from superfluid helium but, as discussed in Sec. IV D 4, a similar “roton feature” was recently also observed in simulations of the electron liquid.549 Interestingly, it was predicted by Hamann et al. that this feature should be observable also in the plasmon dispersion of hydrogen—in the metallic (ionized plasma) phase.169 In the following, we briefly outline this result.

As a first step to investigate how the dispersion behavior changes when considering a two-component partially ionized plasma, as compared to jellium, we treat electron–ion interactions in the relaxation time approximation (RTA). This corresponds to introducing a BGK collision operator, as discussed in Sec. III H 1, which contains a characteristic relaxation time τ for the electron momentum distribution which is the inverse of the electron–ion collision frequency, ν=τ1.

Solving the quantum kinetic equation of the electrons [Eq. (83)], with the collision integral (90) in linear response, for a weak monochromatic field, one derives the following expression for the dielectric function,573 which is due to Mermin728 and is referred to as Mermin dielectric function,
(131)
were, ε(q,ω) is the dielectric function of jellium, which is recovered in the limit of vanishing electron–ion collisions, ν0. The original derivation uses the Lindhard function (random-phase approximation), further research has shown this approach to be compatible with dielectric functions that include electronic correlations, e.g., are described using local field corrections G(q,ω).319 There exist various extensions of the Mermin dielectric function that introduce complex and frequency-dependent collision frequencies, ν(ω), i.e., collision rates that depend on the energy scale.284,729 Expressions for the collision frequency can be derived using approximations of quantum kinetic theory—either from reduced density operators (BBGKY-hierarchy) or nonequilibrium Green functions70,594 (cf. Sec. III H 1). Several approximations were discussed (e.g., in Ref. 283). The analysis of Ref. 169 used the dynamically screened Born approximation but neglected the plasmon peak in the integrand, for numerical reasons,
(132)
Here the ion–ion static structure factor, Sii(q), enters in the calculation of the potential at which electrons are scattered.
Another approach interprets the Mermin dielectric function as a generalized version of the Drude formula, which matches in the optical limit
This is particularly useful in the context of DFT-MD simulations where it is possible to obtain results for the dynamic conductivity via the Kubo–Greenwood (KG) formula730 (cf. Sec. IV 5 2), from which collision frequencies can be extracted and used in Eq. (131), thereby extrapolating such results to finite wave vectors.325 However, in the absence of true first-principles calculations providing electron–ion local field corrections, such an extrapolation is of unknown quality and has, in some cases, been shown to be not reliable.731 Further research into transport properties of hydrogen under high pressure and, in particular, comparisons between KG approaches and linear response TDDFT will shed light on this important issue (see Sec. IV E).

The dispersion of the peak position of the dynamic structure factor is shown in Fig. 39. There we compare results for jellium, using the static local field correction (dashed line, labeled LFC), and calculations including, in addition, collisions with the ionic background based on the Mermin dielectric function, Eq. (131) (solid line). Collision frequencies were calculated using Eq. (132) while setting Sii(q)=1.732 

FIG. 39.

Emergence of the negative dispersion of the peak of S(q,ω) for partially ionized hydrogen at T=50000 K. Results for jellium using the static LFC (dashed) are compared to results which additionally account for collisions with ions using the Mermin dielectric function [Eq. (131)] (solid). In contrast to Fig. 37, here absolute units are used on the wave number and frequency axes. At this temperature, a local minimum would start to emerge for jellium at significantly lower densities only, corresponding to rs*>6.

FIG. 39.

Emergence of the negative dispersion of the peak of S(q,ω) for partially ionized hydrogen at T=50000 K. Results for jellium using the static LFC (dashed) are compared to results which additionally account for collisions with ions using the Mermin dielectric function [Eq. (131)] (solid). In contrast to Fig. 37, here absolute units are used on the wave number and frequency axes. At this temperature, a local minimum would start to emerge for jellium at significantly lower densities only, corresponding to rs*>6.

Close modal

The expected location of the roton feature in jellium and dense hydrogen is shown in Fig. 40; the location is to the left of the black (for jellium) or red dashed (for hydrogen) curves, respectively. As the roton is expected to occur in the unbound part of the electrons that form the free electron component within the hydrogen sample, the partial ionization was taken into account, based on PIMC simulations.115 The red data points for several isotherms indicate the maximum total electron densities to be around 610×1022 cm−3 (0.1 0.17 g/cm3). The necessary densities can be achieved in experiments for instance using a hydrogen jet.257,258

FIG. 40.

Change of the line indicating the onset of roton behavior (to its left), from jellium to hydrogen. The arrows illustrate the effect of partial ionization at a given temperature. Updated version of figure from Hamann et al.,169 using degree of ionization data from Ref. 115 (cf. Fig. 23).

FIG. 40.

Change of the line indicating the onset of roton behavior (to its left), from jellium to hydrogen. The arrows illustrate the effect of partial ionization at a given temperature. Updated version of figure from Hamann et al.,169 using degree of ionization data from Ref. 115 (cf. Fig. 23).

Close modal

The resolution in angular and energy space that is necessary to observe the predicted roton feature in XRTS experiments169 will only be available at modern XFEL facilities that combine a brilliant, narrow-band x-ray laser with fast detectors and a high repetition optical laser and target delivery system.733,734 In addition, advanced spectral techniques700,735 might be necessary and, naturally, a theory-free determination of basic parameters such as temperature and density.160–163 

6. Ion–acoustic mode in partially ionized hydrogen

After analyzing correlation effects in the electron plasmon dispersion in dense hydrogen, we now turn to the low-frequency collective oscillations of the protons.

We have used the LAMMPS code736 to perform semiclassical molecular dynamics simulations of a hydrogen plasma with electrons and ions interacting via the improved Kelbg potential (Sec. III G 2), similar to previous simulations in Refs. 471 and 737. Compared to DFT-MD simulations (Sec. III C), they are computationally much cheaper and allow one to treat much larger system sizes with thousands of protons and electrons. In addition, they do not rely on the Born–Oppenheimer approximation and treat electrons dynamically (non-adiabatic). On the other hand, they are applicable only at sufficiently high temperatures and low densities. A comparison with PIMC simulations471 indicates that the improved Kelbg potential (IKP) is applicable for temperatures well above 60 000 K, at rs=4 and rs=6. New tests of the IKP were presented in Sec. IV A 3 and provided more details on the accessible density-temperature range. They also confirm that the MD simulations should be accurate for the parameters of Fig. 41.

FIG. 41.

Ion–ion dynamic structure factor Sii(q,ω) at rs=3 and T= 125 000 K for various wave vectors q where d is the Wigner–Seitz radius [Eq. (2)]. Frequencies are given in units of the ion plasma frequency, ωpi.

FIG. 41.

Ion–ion dynamic structure factor Sii(q,ω) at rs=3 and T= 125 000 K for various wave vectors q where d is the Wigner–Seitz radius [Eq. (2)]. Frequencies are given in units of the ion plasma frequency, ωpi.

Close modal

The MD simulations give direct access to various time correlation functions and their spectra, including the ion–ion dynamic structure factor. When calculating dynamic properties, one must keep in mind, however, that the (improved) Kelbg potential was designed to reproduce thermodynamic properties, and its use for the calculation of dynamic quantities remains to be tested.

In Fig. 41, we show preliminary results from a simulation with 40 000 protons and electrons for the ion–ion DSF of a hydrogen plasma at rs=3 and T= 125 000 K. At the smallest wave number, the data indicate the formation of an ion–acoustic mode well below the ion plasma frequency, which appears to vanish at larger q. Longer simulations are still required to reduce the noise and to verify the peak formation. This would provide access to the ion–acoustic speed (see also Ref. 737). Experimentally, acoustic modes in methane under WDM conditions have recently been resolved in experiments using inelastic x-ray scattering.738 

We now turn to the transport (e.g., electrical and thermal conductivity) and optical (e.g., opacity) properties of dense hydrogen in thermodynamic equilibrium. Here, a variety of simulation methods is available. For a summary of the relevant properties and simulation approaches (see Sec. III A 3 and Table III).

1. Previous comparisons of transport coefficients

At a transport coefficients comparison workshop that took place in 2016,627 participants were invited to compute electrical conductivity, thermal conductivity, viscosity, diffusion, and stopping power for three plasma compositions (H, C, and CH) with densities varying from 0.1 to 100g/cm3 and temperatures from 0.2 eV to 2 keV. The models included KS-DFT, OF-DFT, TDDFT, average atom models, and various analytical models. All models tested showed significant variation (factors of three or more) for a portion of the density–temperature space. Agreement among models was generally higher in the classical weakly coupled regime. At low temperatures and densities, uncertainties in the ionization state were the major source of disagreement. For most transport coefficients, typical best-case variations among codes were 20%, in the weakly coupled regime, factors of two in warm dense matter, and worsening to factors of ten or more, at low temperatures. A follow up transport workshop took place in 2023 and significantly refined the analysis for a broad group of materials (cf. Ref. 628). One conclusion was that the difference in the DC electrical conductivity was at worst a factor of seven between all models and a factor of two between similar models.

The workshops mentioned above did not attempt to rank the different simulations or to identify the most accurate ones. In contrast, in the present paper, we have attempted to evaluate the accuracy of different approaches—for hydrogen and a limited range of parameters—and presented comparisons of the associated thermodynamics results (which are also the basis for transport applications) in Sec. IV A. A clear preference has been observed for first-principles methods, i.e., QMC and DFT. In Secs. IV E 2 and IV E 3, we focus on DFT results and theoretical issues associated with the method since QMC does not give access to transport and optical properties (see Tables III and IV). In Sec. IV E 2, we discuss the Kubo–Greenwood formula that is the basis of the KS-DFT approach to transport quantities. This is followed by state of the art results for the electrical and thermal conductivity in Sec. IV E 3 and opacity, in Sec. IV E 4. We conclude with a discussion of the important issue of scattering effects in DFT simulations (Sec. IV E 5).

2. KS-DFT approach. Kubo–Greenwood formula

The electronic transport properties, such as electrical and thermal conductivity, can be derived from linear response theory.323,324 The resulting frequency-dependent Onsager coefficients can be expressed as739 
(133)
where ω is the frequency and V denotes the volume of the simulation box. Ekμ and fkμ describe the energy eigenvalue and the occupation number of the respective Bloch state |kμ. The states are calculated within the Kohn–Sham framework of DFT, together with the transition matrix elements with the velocity operator, kν|v̂|kμ. The enthalpy per electron, h, can be directly obtained from the chemical potential and the entropy, which are self-consistently calculated within the DFT as the free-energy is minimized. The real parts of the electrical conductivity, σ, and the thermal conductivity, λ, are subsequently calculated via
(134)
(135)
When calculating the DC limit, ω0, of these quantities, great care has to be taken of the representation of the δ function in Eq. (133), which is usually done by a Gaussian whose width has to be chosen to the typical distance between bands in the eigenvalue spectrum. Additionally, Eqs. (134) and (135) are generally tensor quantities, so components need to be combined, which is only trivial for cubic systems. Also, note that most DFT codes are only calculating transitions at the same k point (kk) and might not be applicable in their standard version for non-local potentials. Many of these technical aspects are carefully described in recent literature.730,740–742

The above Kubo–Greenwood formalism is the standard approach to obtain electronic transport properties from DFT and has led to many studies of the electrical and thermal conductivity of hydrogen. Most of the pressure–temperature range covered in these studies are typical for the interior of giant planets,33,104,739,743 where Kohn–Sham DFT is very efficient. Other studies pushed the calculations to extremely high temperatures in the ICF domain,740,744–746 which is only possible when the density is sufficiently high so that the number of Bloch states remains computationally tractable or by combining Kohn–Sham DFT with orbital-free DFT. On the low density end of the warm dense matter region, data are scarce, because the calculations are computationally very demanding due to the large simulation box sizes.740 

3. DFT results for the electrical conductivity

The electrical conductivity of dense hydrogen plasma is shown in Fig. 42, as computed by Holst et al.739 using the PBE XC functional when evaluating the Kubo–Greenwood formula (cf. Sec. IV E 2). Since these results terminate at densities ρ=0.1 g cm−3, here we present new independent DFT-PBE results that extend to one order of magnitude lower densities, see also Table VIII, and are found to be consistent with the earlier results. The corresponding EOS data were tested against FP-PIMC in Sec. IV A 4.

FIG. 42.

Results for the electrical conductivity, Eq. (134), of dense hydrogen, as a function of density for different temperatures. Average atom data (AA, circles) and DFT-KG using the PBE functional (lines) are compared. Some DFT data (full lines) were taken from Holst et al.,739 new results of this paper, that correspond to the EOS results of Fig. 14, are shown by the dashed lines (note the slightly different temperature values in both datasets). The new data for low densities, rs3, are reproduced in Table VIII.

FIG. 42.

Results for the electrical conductivity, Eq. (134), of dense hydrogen, as a function of density for different temperatures. Average atom data (AA, circles) and DFT-KG using the PBE functional (lines) are compared. Some DFT data (full lines) were taken from Holst et al.,739 new results of this paper, that correspond to the EOS results of Fig. 14, are shown by the dashed lines (note the slightly different temperature values in both datasets). The new data for low densities, rs3, are reproduced in Table VIII.

Close modal

Let us summarize the observed results reported in Fig. 42. The electrical conductivity increases with density, which is a result of pressure ionization (cf. Sec. II A). For temperatures below 1500 K, i.e., in the dense fluid phase, the electrical conductivity shows jumps which are typical of a first-order phase transition—the LLPT (cf. Sec. II B 2). Above the critical point of the LLPT, the increase is steep but continuous and becomes less pronounced with increasing temperature. Here, thermal ionization leads to a broadening of the nonmetal-to-metal transition. Note that the temperature trend of the electrical conductivity is inverted above the transition density which is a typical metal-like behavior. Here, increasing temperature broadens the Fermi function which allows for more scattering processes of the electrons at ions so that their mobility is being reduced.

For the three temperatures considered at the lower densities shown in Fig. 42, we can observe satisfactory agreement between DFT-KG results and AA calculations using the Ziman approach.287,407 For these conditions, the conductivity never drops to zero as there always is some remnant, fluctuating partial ionization, in agreement with the PIMC results (cf. Fig. 23).

The electronic thermal conductivity shows a very similar behavior (see Fig. 43). In the dense fluid below 1500 K, a sharp rise over several orders of magnitude appears at about 0.9g/cm3. This is a result of an abrupt pressure-driven ionization. Above the critical temperature, the nonmetal-to-metal transition is thermally broadened as in the case of the electrical conductivity. Contrary to the electrical conductivity, the isotherms of the thermal conductivity increase systematically with temperature for all densities.

FIG. 43.

Results for the electronic thermal conductivity [Eq. (135)] of dense hydrogen plasma as a function of the density for different temperatures. Reproduced from Holst et al.739 with the permission of the authors. Copyright (2011) by the American Physical Society.

FIG. 43.

Results for the electronic thermal conductivity [Eq. (135)] of dense hydrogen plasma as a function of the density for different temperatures. Reproduced from Holst et al.739 with the permission of the authors. Copyright (2011) by the American Physical Society.

Close modal

4. DFT results for the opacity

The radiative properties of dense hydrogen (deuterium) can be derived from the frequency-dependent electrical conductivity calculated by DFT-based molecular-dynamics (quantum molecular dynamics, QMD) simulations combined with the linear-response theory of Kubo and Greenwood (cf. Sec. IV E 2). Starting from the real part of the electric conductivity, σ1(ω), Eq. (134), we can use the Kramers–Kronig relation to obtain its imaginary part, σ2(ω), by principle-value integration. The obtained complex conductivity is directly related to the frequency-dependent dielectric function [ q0 limit of the full dielectric function, Eq. (99)], ε(ω)=ε1(ω)+iε2(ω):
(136)
(137)
which allows one to calculate the refractive index, n(ω)+ik(ω):
(138)
(139)
Finally, the frequency-grouped Rosseland mean opacity of dense hydrogen can be computed, for photon energies in between ω1 and ω2, by using the following integration formula:
(140)
with the mass absorption coefficient being defined as αm(ω)=4πσ1(ω)/[n(ω)cρ], where c and ρ are the speed of light and mass density, respectively. Here, the Planck function
(141)
gives the radiation spectrum of a black-body system at temperature T.

This method of combining DFT with Kubo–Greenwood linear response theory has been applied to build first-principles opacity tables (FPOT) of warm-dense hydrogen (deuterium)747 and polystyrene748 in a wide range of densities and temperatures. One example is shown in Fig. 44 for the opacity of deuterium where we compare FPOT (marked as “QMD” as it was derived from QMD+Kubo–Greenwood calculations where the PBE XC functional was used) and the astrophysics opacity table (AOT) created by Los Alamos National Lab.749  Figure 44 displays the 48-group Rosseland mean opacity of deuterium in the warm dense matter range for two different conditions: (a) ρ= 5.388 g/cm3 and kBT= 10.8 eV and (b) ρ= 199.561 g/cm3 and kBT= 43.1 eV. It indicates that the opacity is sensitive to strong coupling and degeneracy effects and can differ significantly from traditional atomic code calculations in the WDM regime. The difficulty of extending first-principles DFT-MD/QMD+KG calculations of the opacity to high temperature (beyond the Fermi temperature) lies in the number of bands needed for convergence. It is noted that the DFT-based average-atom (DFT-AA, Sec. III D) model calculations (green diamonds in Fig. 44), which are more readily extended to high temperatures, give good agreement with QMD calculations, in particular in the low and mid frequency ranges. High-frequency opacity calculations with the QMD+KG method invoke extrapolations, which can result in some uncertainties.

FIG. 44.

Comparison of the deuterium 48-group Rosseland mean opacity among QMD, AOT, and DFT-AA calculations as a function of the group photon energy. (a) ρ= 5.388 g/cm3 and kBT= 10.8 eV. (b) ρ=199.561g/cm3 and kBT= 43.1 eV.

FIG. 44.

Comparison of the deuterium 48-group Rosseland mean opacity among QMD, AOT, and DFT-AA calculations as a function of the group photon energy. (a) ρ= 5.388 g/cm3 and kBT= 10.8 eV. (b) ρ=199.561g/cm3 and kBT= 43.1 eV.

Close modal

In conclusion, we note that the opacity of hydrogen has recently been measured at the NIF for conditions that are relevant for the interior of red dwarf stars. The conceptual paper264 shows also DFT-MD predictions for a (25%/75%) hydrogen–tritium mixture at ρ= 200 g/cm3 and temperatures of kBT= 100 eV and kBT= 150 eV, in comparison with an AA model (cf. Sec. III D), very similar to those displayed in Fig. 44.

5. Correlation and collision effects in DFT and TDDFT

While e-e correlations are treated on the level of the chosen XC functional, it has long been debated whether or not e-e collisions are properly included in calculations of electronic transport using KS-DFT for the evaluation of the Kubo–Greenwood formula (Sec. IV E 2; see, e.g., Refs. 746, 750, and 751). The Spitzer theory752 gives exact values for the transport coefficients in the non-degenerate (Θ1) and weakly coupled (Γ1) limit, both for considering and neglecting e-e collisions, besides e-i collisions. This information can be used to determine the coefficients of a virial expansion of the electrical conductivity, σ(n,T) [Eq. (134)] and to benchmark DFT results (see Ref. 753). In this limit, the electrical (σ) and thermal conductivity (λ), as well as the thermopower (α) can simply be expressed by prefactors f, a, and L as follows:
(142)

Here, lnΛc is the well-known Coulomb logarithm [cf. Eq. (88)]. For a fully ionized hydrogen plasma with Z=1, the Spitzer values for the prefactors with and without considering e-e collisions are fei=1.0159, fei+ee=0.5908, aei=1.5, aei+ee=0.7003, Lei=4.0, and Lei+ee=1.5966.752 Overall, e-e collisions reduce the electronic transport coefficients considerably in the non-degenerate limit.

Recently, French et al.740 performed extensive calculations of the transport properties of hydrogen across the plasma parameter plane, along the line Γ8Θ7=1, in order to check if the correct Spitzer values are reproduced in the non-degenerate (Θ1) and weakly coupled (Γ1) limit. Evaluating the Kubo–Greenwood formula in that limit, using KS-DFT, is numerically demanding since huge simulation cells and a large number of bands are required in order to converge the calculations. This has so far prevented obtaining conclusive results. French et al.740 could show explicitly that the KS-DFT results approach the Lorentz plasma values without e-e collisions, but not the Spitzer values, so this approach does not capture e-e scattering processes (see Fig. 45 for a and L).

FIG. 45.

Thermopower coefficient a (red) and Lorenz number L (black) from DFT (full) and the relaxation time approximation (RTA, dashed) along the Γ8Θ7=1 line. Reproduced from French et al.740 with the permission of the authors. Copyright (2022) by the American Physical Society.

FIG. 45.

Thermopower coefficient a (red) and Lorenz number L (black) from DFT (full) and the relaxation time approximation (RTA, dashed) along the Γ8Θ7=1 line. Reproduced from French et al.740 with the permission of the authors. Copyright (2022) by the American Physical Society.

Close modal

In the degenerate limit (Θ1), the influence of e-e collisions vanishes, due to strong screening of the Coulomb interactions. More importantly, Pauli blocking prevents e-e scattering processes, except for a small energy range near the Fermi energy. Inspecting Fig. 45, the correct values for L=π2/3 (Wiedemann–Franz law) and a=0 indeed do follow from KS-DFT. For a more detailed discussion of the influence of e-e collisions on the transport coefficients (see Refs. 191 and 753). Note that Reinholz et al.750 have derived a correction factor Ree(T,Θ,Z) for the influence of e-e collisions on the electrical conductivity which can be applied in a wide temperature and density range.

A potential alternative route toward optical properties from KS-DFT simulations is given by LR-TDDFT693,720 and RTTDDFT699,754,755 (cf. Sec. III H 4). In principle, both methods allow one to consistently include e-e correlation effects, although both, the dynamic XC-kernel and the dynamic XC-potential are usually treated in an adiabatic (i.e., static) approximation in practice.

After discussing simulations of thermodynamic, dynamic, transport, and optical properties of dense hydrogen in thermal equilibrium, we now briefly outline the prospects for time-dependent simulations. We consider two examples of quantities: in Sec. IV F 1 we discuss the application of linear-response TDDFT to the dynamic structure factor. Then, in Sec. IV F 3 we discuss the application of quantum kinetic equations to the electronic stopping power.

1. Dynamic structure factor from linear-response TDDFT simulations

A promising route toward ab initio results for dynamic properties of hydrogen is given by linear-response TDDFT. As explained above, here the central quantity is the dynamic electronic density response function, χee(q,ω), which is expressed in terms of the dynamic KS response, χ0(q,ω) and the, in principle, also dynamic XC-kernel Kxc(q,ω). In practice, both the XC-functional and the XC-kernel have to be approximated, with the latter being particularly difficult. Very recently, Moldabekov et al.617,693 have suggested a new way that allows one to compute the static XC-kernel Kxc(q)=Kxc(q,0) for arbitrary XC functionals on any rung of Jacob's ladder of functional approximations388 without the need to explicitly evaluate functional derivatives. This, in turn, makes it possible to compute the dynamic density response function, Eq. (102), within the static approximation147,166,617
(143)
which can be generalized to a Dyson's type equation [cf. Eq. (104)] for inhomogeneous systems. The fluctuation–dissipation theorem [cf. Eq. (119)] then opens the way to estimate See(q,ω) from Eq. (143). The static approximation is expected to work well for moderate densities rs4 and moderate temperatures Θ1, based on previous investigations of the UEG.147,165,166

In Fig. 46, we show corresponding linear-response TDDFT results for the inelastic component of the DSF [cf. Eq. (128)], for hydrogen at rs=2 and Θ=1, for q=3.06 Å1. The dotted green and blue curves have been obtained for the UEG, where the inclusion of the local field correction leads to a well-known red shift.166 For hydrogen, the RPA curve that has been obtained by setting Kxc(q)0, in Eq. (143), is qualitatively close to the RPA result for the UEG and exhibits a peak at the same frequency. Similarly, using the static XC-kernel from Ref. 141 (solid black), or the adiabatic LDA (ALDA) kernel (solid yellow) leads to the same red shift as for the UEG.

FIG. 46.

Linear response TDDFT results for the inelastic component of the dynamic structure factor [cf. Eq. (128)] of hydrogen at rs=2 and Θ=1 for q=3.06 Å1. Taken from Böhme et al.141 with the permission of the authors.

FIG. 46.

Linear response TDDFT results for the inelastic component of the dynamic structure factor [cf. Eq. (128)] of hydrogen at rs=2 and Θ=1 for q=3.06 Å1. Taken from Böhme et al.141 with the permission of the authors.

Close modal

The further exploration of this framework to study the dynamic density response and dynamic structure factor of hydrogen (and potentially other elements) constitutes an important topic that will be pursued in dedicated future works.

2. Capabilities of real-time-dependent DFT simulations

Real-time TDDFT simulations will be of much use for improved descriptions, in particular of conductivities and stopping power.627,628 The direct time-dependent simulation of the reaction of the system to an external perturbation, be it via radiation or particle impact, constitutes an advantage over linear response methods such as LR-TDDFT. Thus, the dynamic structure of warm dense matter at finite wavenumbers, i.e., the XRTS signal, can be computed.699 In combination with Ehrenfest dynamics, the stopping of fast beam particles in warm dense hydrogen can be investigated.288,611,612,756,757 A special advantage of a direct simulation is the capability to include non-linear effects and, in principle, also Electron–electron collisions. In addition, the optical limit of the conductivity can be accessed more straightforwardly in comparison to the DFT-Kubo–Greenwood approach.

3. Time-dependent charged particle stopping from quantum kinetic theory simulations

The stopping power, Sx=dEdx—the energy loss of energetic particles per unit length in a plasma—is of crucial importance for many applications. A large arsenal of simulation techniques have been used to compute Sx, including molecular dynamics (e.g., Ref. 758) or time-dependent DFT (for recent results, see Refs. 611 and 612).

On the other hand, the recent advances in the description of correlated quantum plasmas, as described in this manuscript, have made it possible to obtain the density fluctuations, δn(r), in a WDM system from first principles. This naturally includes non-linear fluctuations.265,759 On this basis, it is possible to determine the stopping power from the consideration of the friction experienced by a beam particle due to the nonlinear (electronic) density fluctuations760 
(144)
where Ze is the charge of the beam ions moving at velocity v. The density fluctuation can then be obtained in linear response theory via the inverse dielectric function, which leads to
(145)
where nB is the Bose function and ω±=k2/2m±kv.

It follows from Eq. (102) that the inverse dielectric function allows to systematically account for correlation effects via the dynamic local field correction, G(q,ω), that is accurately available from QMC simulations.613 The effect of electronic correlations is demonstrated in Fig. 47 where we compare RPA results (G0) to QMC results and the analytical STLS model.761 While for large projectile velocities, vvth, correlation effects are not important, for velocities, v3vth, they cause a significant increase in the electronic stopping power.

FIG. 47.

Stopping power as a function of projectile velocity (in units of the thermal velocity and the Fermi velocity, respectively) for rs=4 and Θ=0.5. Uncorrelated results (RPA) are compared to QMC results and the STLS model. Reproduced from Ref. 613 with the permission of the authors. Copyright (2020) by the American Physical Society.

FIG. 47.

Stopping power as a function of projectile velocity (in units of the thermal velocity and the Fermi velocity, respectively) for rs=4 and Θ=0.5. Uncorrelated results (RPA) are compared to QMC results and the STLS model. Reproduced from Ref. 613 with the permission of the authors. Copyright (2020) by the American Physical Society.

Close modal
Independently from linear response theory, the stopping power can be obtained from quantum kinetic theory, which directly benefits from the first-principles simulation results shown in this paper. The equation for the stopping power of the beam particles interacting with a warm dense matter system reads
(146)
The time derivative of the beam Wigner function, fb, is given by a kinetic equation, which reads for the homogeneous case,762,
(147)
This equation has to be solved in tandem with a similar one for the plasma particles. The collision integral can be of, e.g., Boltzmann type, Lenard–Balescu like, or have any other shape763–765 (cf. Sec. III H 3). Nonlinear and strong coupling effects can be accounted for via higher order (beyond second order Born) contributions to the selfenergies (scattering rates) Σ.148,149

Here we present new results for the application of quantum kinetic equations. They have been obtained within the G1–G2 scheme that was introduced in Sec. III H 3. An example of the solution of the G1–G2 equations for a dense quasi-1D plasma, which extends previous simulations (Ref. 606), is shown in Fig. 48. We consider a fully ionized electron–ion plasma at high density and moderate temperature corresponding to rs=0.5 and Θ=0.81. For technical reasons, we consider a quasi-one-dimensional plasma (for details of the model, cf. Ref. 606) and reduce the ion mass to 5me. We start with an electron plasma in thermal equilibrium which is impacted by a nearly monoenergetic ion beam (dashed blue curve in the upper left part of Fig. 48). The simulations demonstrate how the electron distribution changes: it shifts to the right (electrons gain momentum) and broadens (electrons are being heated). At the same time, ions lose energy, i.e., their distribution broadens.

FIG. 48.

Time evolution of electronic (full lines) and ionic (dashed lines) distribution functions and observables during ion impact. The electronic parameters are rs=0.5 and Θ=0.81. The ion impact starts at t=0.3 fs, cf. vertical gray line. Top figure: Electron and ion distribution functions. Second graph: total, kinetic, mean field (Fock), and correlation energy per particle. Third figure: momentum per particle. Bottom: Electronic plasmon number for five wave numbers. For details, see text.

FIG. 48.

Time evolution of electronic (full lines) and ionic (dashed lines) distribution functions and observables during ion impact. The electronic parameters are rs=0.5 and Θ=0.81. The ion impact starts at t=0.3 fs, cf. vertical gray line. Top figure: Electron and ion distribution functions. Second graph: total, kinetic, mean field (Fock), and correlation energy per particle. Third figure: momentum per particle. Bottom: Electronic plasmon number for five wave numbers. For details, see text.

Close modal

From the time-dependent distributions, we also compute macroscopic observables, the time evolution of which is shown in Fig. 48. Note that the G1–G2 scheme has the same conservation properties as the exact many-particle system: it conserves particle number, total momentum, and total energy (including interaction energy). The time evolution of the mean momentum (momentum per particle) of electrons and ions is depicted in the middle plot. When multiplied with the respective particle numbers, the momentum loss of ions is exactly compensated by the momentum gain of the electrons. In the second plot from the top of Fig. 48, we plot the different energy contributions of electrons (full lines) and ions (dashed lines), where the total energy consists of kinetic, exchange, and correlation energy, Etota=Ekina+EFocka+Ecora, with a=e,i. Let us have a closer look at the dynamics. At t=0, we start with uncorrelated electrons (ideal Fermi gas). To obtain the correct initial state, which includes correlation effects due to the Coulomb interaction of the electrons, we apply the adiabatic switching procedure (e.g., Refs. 298 and 604) during which correlation energy forms which is negative, see the green curve in the inset (due to the weak coupling conditions the correlation energy is substantially smaller than the kinetic energy).

At time t=0.3 fs the ion beam impacts the correlated electron plasma, giving rise to a rapid increase in kinetic, mean field, and (negative) correlation energy. Correlation and Fock energy saturate after about 1 fs, which can be identified with the correlation time in the system (see Refs. 589 and 766). After this time, the energy exchange between ions and electrons continues but at a significantly slower rate. Also, the distribution functions continue to change their shape up to about 8 fs (cf. top part of Fig. 48), which can be identified with the relaxation time. Subsequently, the dynamics are dominated by a kinetic energy exchange between the two components, i.e., by electron–ion temperature equilibration. Another interesting question is that on the mechanism of the energy exchange between electrons and ions—are these two-particle collisions or are collective excitations involved? This question is answered in the bottom panel of Fig. 48 where we plot the plasmon occupation number for five wave numbers. First, we observe a rapid build up of the plasmon occupations, during the adiabatic switching interval, t0.3 fs. With the impact of the ions, the plasmon population increases and exhibits oscillations with the q-dependent plasmon frequency, ω(q), which are damped out when the correlation time is reached. The direct access to the time-dependent plasmon properties is provided by the G1–G2 scheme if it is solved with GW self-energies (non-Markovian quantum generalization of the Balescu–Lenard approximation) [cf. Eq. (94)], as in the present case.

This example demonstrates the capabilities of NEGF simulations and the G1–G2 scheme, in particular, for dense quantum plasmas that are far from equilibrium. These simulations fully include electronic correlations and their dynamics, as well as electron–electron scattering effects which are missing in KS-DFT (cf. Sec. IV E 5). For a fully ionized plasma, all relevant observables can be computed accurately, including their time evolution. This includes the time-dependent stopping power and dielectric properties. Moreover, it is possible to compute the time evolution of the electronic density of states. Overcoming the current main limitation—the memory bottleneck arising from the storage of the two-particle function and, with this, the extension of the G1–G2 scheme to 2D and 3D plasmas should become possible in the near future by applying the quantum fluctuations approach605,607,608 (see Sec. III H 3).

In conclusion, we note that the stopping power is not only of interest for dense plasmas but has recently been investigated also for other targets, such as correlated quantum materials for which NEGF-Ehrenfest dynamics simulations767–769 as well as TDDFT-Ehrenfest simulations770 were reported. There not only the energy loss of the projectile is important but also the charge transfer from the target to the projectile, in particular in the case of highly charged ions.771,772

Hydrogen at high pressure continues to be a focus of research in many fields, including astrophysics, high pressure physics, basic science, and technology, in particular, for inertial confinement fusion. Our understanding of this simplest of all chemical elements remains far from complete. This may seem surprising since the basic equation governing its properties—the Schrödinger equation—has been known for nearly one century. However, strong compression gives rise to a complicated interplay of many particles—a combination of quantum, spin, correlation, and thermal effects, which poses challenges both for experiments and simulations.

In this review article, we focused on important simulation approaches that are capable of substantially advancing our knowledge about dense hydrogen in the near future. Of particular interest was the question about the accuracy and predictive capability of commonly used simulation techniques. In many fields of physics, including atomic physics, semiconductor optics, or cold atoms in traps or optical lattices, this question is routinely answered by high-precision experiments. In the field of dense plasmas, this is not yet possible. Therefore, we outlined strategies to use first-principle simulations, i.e., “computer experiments”, that can be used to benchmark models and simulations. In the following, we summarize the main results and discuss open questions.

The first part of this review (Sec. II) was devoted to the phase diagram of hydrogen at high pressure. We discussed the hypothetical plasma phase transition (PPT) (Sec. II B 1) that has been predicted to occur in partially ionized hydrogen in the gas phase and concluded that there is neither experimental nor reliable theoretical evidence. We then turned to hydrogen at low temperature, in the liquid and solid phases, and reviewed the knowledge about the solid–liquid and insulator–metal transition (LLPT).

A related topic is given by the additional metal-to-superconductor phase transition that was proposed by Ashcroft,21 and further explored in subsequent works (e.g., Refs. 773–776). In this regard, accurate QMC simulations might help by giving insights into effective electronic interactions in the medium,270,272 which are important for the estimation of the critical temperature.777 In the case of deuterium, it is further feasible to carry out either PIMD or PIMC simulations based on an effective ionic potential and, in this way, to directly estimate the superfluid fraction.414 This has recently allowed Myung et al.778 to report a hypothetical supersolid phase779 at high pressure and low temperature, although further verification is required.

In Sec. III, we reviewed the most important simulation methods. A particular emphasis was given to first principles approaches, such as quantum Monte Carlo (QMC) and density functional theory (DFT) simulations, but we also discussed average atom and chemical models, classical and quantum hydrodynamics, and semiclassical simulations with quantum potentials. The methods differ strongly in terms of computational effort, applicable parameter range, and resolution. Most importantly, there are fundamental differences between the methods with respect to their theoretical level and approximations and, hence, their expected accuracy. A detailed comparison of the relevant methods was summarized in Tables III–VI.

The expectations regarding the accuracy are derived from the general (qualitative) validity range of the approximations (including self-energies, in the case of Green functions, or XC functionals, in the case of DFT). At the same time, the quantitative accuracy when an approximation is applied to dense hydrogen can only be established using strict benchmarks against known results.

Therefore, in Sec. IV we presented extensive tests of various methods using fermionic PIMC results of Ref. 115 as reference. The results are summarized as follows:

  • RPIMC with free particle nodes is the most accurate of the approximate methods, in the considered parameter range, with relative errors of the pressure not exceeding 2%, except for the lowest temperature, T=15625 K. The accuracy of the results (and the quality of the nodes) at high density, rs3, remains to be tested.

  • KS-DFT-MD with PBE functionals exhibits pressure deviations from FPIMC of up to 7%. The agreement improves substantially with the KDT16 finite-temperature PBE functional, especially for T60000 K, to about 2%, whereas for T30000 K the deviations are larger.

  • Semiclassical MD simulations with the improved Kelbg potential achieve an accuracy of 1%–3%, for the pressure, for T60000 K and rs7, where the density range quickly widens with increasing temperature.

  • DFT-AA models show good agreement with FPIMC, for pair distributions and static structure factors, and with KS-DFT, for the conductivity. For quantitative benchmarks additional, tests will be required.

Further, we demonstrated that, aside from benchmarking other methods, first-principles simulations can also be combined with simpler models, such as chemical models, to compute quantities of interest that are difficult to access otherwise. An example is the ionization potential depression (IPD) for which various models exist but only a few first principles results have been reported.651 In Sec. IV B 2 we showed that, by using the fractions of free particles and atoms from FPIMC simulations as input, allows one to directly obtain the IPD. Similar approaches can also be developed for other quantities, on the basis of DFT simulations, and for more complex systems.

An interesting property of correlated quantum many-body systems is the non-exponential decay of the momentum distribution, n(k), at large values of the argument, which we demonstrated for jellium in Sec. IV C. It is expected that similar behavior exists in dense hydrogen, both for electrons and ions. Quantification of this effect could have implication for nuclear fusion rates in dense plasmas.

1. Upcoming experiments and suggested model developments

The importance of dense hydrogen is reflected in a large number of ambitious upcoming experiments at top facilities around the world. This includes experiments with liquid hydrogen jets at the European XFEL and LCLS, equation of state measurements at FAIR (GSI Darmstadt), compression experiments at the Omega laser facility in Rochester, and compression experiments at Sandia's Z machine, via the Fundamental Science Program using pulsed power (see, e.g., Ref. 213). Furthermore, at the National Ignition Facility extensive, implosion experiments are planned based on proposals that are submitted via the NIF Discovery Science Program (e.g., Ref. 212) (see also the newly developed colliding planar shock platform at the NIF).780 

These experiments pose tremendous challenges for theory and simulations and require a long-term strategy in the field. The experience with hydrogen simulations shows that there is no “silver bullet,”781 i.e., no single method that would allow one to compute all quantities precisely and to solve all problems. Instead, a smart combination of methods is required. For example, for ICF modeling, large length and time scales and spatial inhomogeneities have to be resolved which is currently only possible with a hydrodynamic description of the dense plasma. At the same time, hydrodynamic models have an unknown accuracy, in particular, in the warm and hot dense matter regime. Substantial improvements are possible with input for the equation of state and transport and optical properties from more accurate approaches, such as average atom models or semiclassical MD with quantum potentials. These methods, in general, also have an unknown accuracy and can be benchmarked and improved with input from more accurate simulations such as Kohn–Sham DFT. However, KS-DFT requires the choice of an exchange–correlation functional the accuracy of which is generally unknown. The use of FPIMC results as benchmarks—even if they are available only for a limited set of parameters—allows one to significantly improve KS-DFT results for dense hydrogen (cf. Sec. V A). Such a combination of methods promises to be both accurate and efficient, not only for ICF modeling.

The approach we are putting forward is sketched in Fig. 49. It involves the development of first principles fermionic PIMC simulations which provide the highest accuracy results for a limited range of parameters, which can be used to test and improve (indicated by the vertical red arrows with “Benchmarks”) more approximate methods, including RPIMC, Green functions, and KS-DFT. The latter, in turn, are capable of simulations for a significantly larger length scales, larger range of parameters and materials. Similarly, having accurate KS-DFT results available may substantially improve (vertical blue arrow) other less accurate simulations, including average atom models or classical and quantum hydrodynamics.

FIG. 49.

Schematic overview of important simulation methods for dense hydrogen, ordered by accuracy (and resolution) vs maximum length and time scales that can be reached by the simulation. First principles QMC methods can be used to benchmark other methods and to improve their accuracy. They also allow for the construction of accurate force fields using machine learning approaches (ML-FF). DFT, on the other hand, can be used to benchmark more approximate methods, including AA models. Horizontal arrows: input for fluid simulations (e.g., for ICF modeling) which include EOS: equation of state; σ (λ): electrical (thermal) conductivity; Sx: stopping power [Eq. (146)]; KR: opacity [Eq. (140)]. More details of the listed methods and quantities can be found in Sec. III and Tables III–VI.

FIG. 49.

Schematic overview of important simulation methods for dense hydrogen, ordered by accuracy (and resolution) vs maximum length and time scales that can be reached by the simulation. First principles QMC methods can be used to benchmark other methods and to improve their accuracy. They also allow for the construction of accurate force fields using machine learning approaches (ML-FF). DFT, on the other hand, can be used to benchmark more approximate methods, including AA models. Horizontal arrows: input for fluid simulations (e.g., for ICF modeling) which include EOS: equation of state; σ (λ): electrical (thermal) conductivity; Sx: stopping power [Eq. (146)]; KR: opacity [Eq. (140)]. More details of the listed methods and quantities can be found in Sec. III and Tables III–VI.

Close modal

For such a hybrid approach to be successful, it is necessary to continuously develop and improve all key simulation methods, including FPIMC, RPIMC, CEIMC, Green functions, DFT, average atom models, and hydrodynamics which is motivated by their complementary strengths and limitations, that were illustrated in Table III. There are many important developments going on in these fields, among which we highlight

  • Further improvement of fermionic PIMC (and potentially PIMD442,782–784) simulations for dense hydrogen and extension to stronger degeneracy. Application of the ξ-extrapolation method148,163,445,446,476,480,784 (cf. Sec. III E 6) to hydrogen over a broader range of parameters, and more observables, in particular static properties related to the EOS.

  • Improving RPIMC simulations of hydrogen (and beyond) (cf. Sec. III E 4), by using variationally optimized nodal surfaces.87 We note that a somewhat related optimization of the thermal density matrix has recently been pursued by Xie et al.,494 and was successfully benchmarked against FPIMC results for electrons in a 2D harmonic trap.785 

  • Strongly degenerate fully ionized hydrogen plasmas can be simulated with high accuracy using configuration PIMC (CPIMC), as was demonstrated for jellium120,121,125 (cf. Sec. III E 3). Developing CPIMC for hydrogen appears to be promising to obtain benchmark data for high densities, rs1.

  • Using highly accurate QMC simulations on the microscale to train machine-learning representations is likely to constitute a viable route toward larger systems.

  • Further development of finite temperature XC functionals for DFT128,136–138,157,230,363,392,393 on higher rungs of Jacob's ladder of functional approximations,388 for example, based on FPIMC results for the XC-kernel either of the UEG167,786,787 or warm dense hydrogen148 (see, e.g., Ref. 679). This might be complemented by the further exploration of high-T DFT methods such as extended KS-DFT372,373 or the spectral quadrature approach.788 

  • Alternative DFT schemes including ab initio molecular simulations with numeric atom-centered orbitals (FHI-aims),789 and stochastic DFT, which scales linearly with N and is, therefore, suited for performing simulations with large particle numbers in order to answer “large system questions.”375–377 Recently, a mixed stochastic-deterministic approach has been proposed,378 which combines the best aspects of both schemes.

  • Nonlinear effects can be analyzed efficiently either based on the LFC/the XC-kernel,759 the ITCF,704 or by direct simulation of the full density response.265 Relevant observables include the stopping power790,791 and effective potentials and forces.270 

  • The combination of Green functions methods (e.g., Dyson and Bethe–Salpeter equation) with Kohn–Sham orbitals from a DFT simulation as input792 should allow for high-quality spectral information (spectral function, DOS, cf. Table III) also for partially ionized hydrogen.

2. First principles input for ICF modeling

Let us return to the discussion of ICF modeling of Sec. III H 5 where we pointed out the importance of accurate data for the equation of state, transport, and optical properties. As we demonstrated in this paper, very accurate data are available from first principles simulations that are suitable as input into hydrodynamic simulations. A general scheme illustrating this “hand over” of results is sketched in Fig. 49.

For improved ICF modeling, moreover, it is important to understand the validity of the hydrodynamic or diffusion approximations used in ICF simulations. Codes such as those relying on kinetic theory (e.g., Vlasov–Fokker–Planck) can provide useful information as to the accuracy of the hydrodynamics description during ICF burn.793,794 This approach is valuable to better understand the micro-physics of diffusive and mixing processes at material interfaces795 which in turn contribute to Rayleigh–Taylor and Richtmyer–Meshkov instabilities. However, in the initial compression phase where quantum degeneracy effects of the electrons are important, classical kinetic theory should be replaced by more accurate kinetic theory, as was discussed in this paper (cf. Sec. III H 1).

An example of uncertain input in the radiation matter equations that has gotten some attention472,796 is the Coulomb logarithm [Eq. (88)], which is used to determine the rate at which electrons and ions couple in a burning plasma. Several different model forms have been proposed to improve the Coulomb logarithm285 and account for the different plasma length scales (e.g., Landau length, thermal de Broglie wavelength, Debye length, see Sec. II A). At the same time, in the quantum degenerate regime, the concept of the Coulomb logarithm breaks down and has to be replaced by scattering rates that follow from improved QKE (cf. Sec. III H 1). Moreover, time-dependent simulations of the initial relaxation phase that include the nonlinear stopping power appear to be possible by using combinations of TDDFT, semiclassical MD, and quantum kinetic theory (cf. Sec. IV F 3).

3. Toward high precision diagnostics of warm and hot dense hydrogen

The upcoming experimental developments that were mentioned in Sec. V B 1 crucially depend on accurate diagnostics and interpretation of the measurements. Let us start with the XRTS technique, which is arguably one of the most valuable methods of diagnostics for WDM but, in the case of hydrogen, is severely hampered due to its comparably small scattering cross section. In this regard, we note exciting new capabilities to probe optically heated hydrogen jet targets257,258 with high-repetition rate XFEL x-ray sources. First, the high repetition rate (pump-probe experiments using the DiPOLE797 and ReLaX798 lasers can be performed with a repetition rate of 10 Hz at the European XFEL) allows one to average over 105 individual shots to accumulate a sufficiently strong scattering signal. This is further facilitated by the high brightness of XFELs (see, e.g., Ref. 799). Second, it is now routinely possible to produce x-ray sources with a bandwidth of ΔE 0.5 eV by monochromating seeded XFELs. In combination with ultrahigh resolution Si (533) Diced Crystal Analyzers at XFEL facilities,735,800 this allows for extremely precise and high resolution measurements of the inelastic scattering. In fact, McBride et al.735 have presented a specialized set-up with a resolution of ΔE0.05eV, which has allowed Descamps et al.801 to resolve the dynamic ion feature of single-crystal diamond. More recently, Gawne et al.700 demonstrated a set-up for measuring the XRTS spectrum with a resolution of ΔE0.1eV over a spectral window of tens of eelectron volt. In combination, these developments will allow for unprecedented XRTS measurements of hydrogen at solid state density over a broad range of temperatures, kBT(1102) eV.

In particular, a narrow and well-characterized source-and-instrument function is key to deconvolve the XRTS intensity in the Laplace domain161 [cf. Eq. (122) in Sec. IV D 2]. This will allow for model-free temperature diagnostics,160 as well as for a host of other applications, such as extracting the normalization162,163 and, in this way, the electron static structure factor See(q). Moreover, probing the hydrogen jet at multiple scattering angles allows for simultaneous measurements, in the collective and non-collective regimes. This is crucial to check the consistency of theoretical models and simulations, and to detect possible non-equilibrium and inhomogeneity effects.261 Finally, the high resolution might allow one to detect the theoretically predicted roton-type feature in warm dense hydrogen169 in a dedicated experiment (see Sec. IV D 5). For completeness, we note that hydrogen jets are not limited to XRTS experiments, but can be combined with a multitude of other diagnostics. For example, Yang et al.802 have recently proposed to use optical shadowgraphy measurements to benchmark hydrodynamics and particle-in-cell (PIC) simulations.803 

At the same time, XRTS measurements on capsule implosion experiments (e.g., at the NIF or the OMEGA laser facility) remain challenging and, in fact, unlikely. An alternative has recently been suggested by Lütgert et al.,264 who proposed a new platform to measure the opacity of hydrogen at strong compression and high temperatures (T 200 eV), resembling the conditions encountered in red dwarfs. In addition to their utility for diagnostics, dependable results for the opacity will help to further advance our understanding of the optical properties of dense hydrogen (cf. Table III). Moreover, it will allow for the benchmarking of theoretical methodologies such as the Kubo–Greenwood formula (Sec. IV E 2) and corresponding improvements that take into account electronic collisions,750 LR-TDDFT with different static and, potentially, dynamic XC-kernels (Sec. IV F 1), RT-TDDFT simulations (Sec. III H 4), and potentially NEGF (e.g., Sec. III H 3).

4. Toward future benchmarks of ab initio simulations and models

The present work has clearly shown the importance of a critical comparison of different methods and demonstrated practical approaches. We presented extensive comparisons of static properties, such as the pressure, pair correlation functions, and related observables using, as a reference, fermionic QMC results.

It is highly desirable to extend the comparisons to dynamic observables, in order to assess the accuracy and expected utility of different methods for the description and interpretation of different types of experiments. Naturally, this is less straightforward than for static properties, since QMC methods generally do not give one direct access to dynamic properties (cf. Table III in Sec. III A 3). Similarly, DFT-based results, e.g., for the dynamic structure factor, See(q,ω), are usually based on one of the following three approximations: (i) Linear-response TDDFT, with an adiabatic approximation to the XC-kernel693 (Sec. IV F 1); even though empirical non-adiabatic kernels do exist, their accuracy is generally unclear, in particular at WDM conditions. (ii) Combination of a dynamic collision frequency, computed in the optical (i.e., q0) limit, with the Born–Mermin approach.159 (iii) Real-time TDDFT with an adiabatic approximation for the dynamic XC-potential699 (cf. Sec. III H 4).

In lieu of unassailable frequency-resolved benchmark data, we are confident that the ITCF, Fee(q,τ), will play an important role in such a project. From a theoretical perspective, it contains the same information as the dynamic structure factor, See(q,ω), due to the uniqueness of the two-sided Laplace transform [Eq. (121)]. In practice, it is straightforward to compute the ITCF either from FPIMC simulations,704 or from any other approach that gives one access to See(q,ω). Moreover, being a wavenumber-resolved property, it is known to exhibit a negligible dependence on the system size,721 which was recently substantiated for the case of warm dense hydrogen.148 We thus propose to benchmark DFT simulations, average atom models, and other theoretical approaches, such as the usual Chihara decomposition (Sec. III B 3), against highly accurate FPIMC results for Fee(q,τ), both for collective and non-collective scattering regimes, over a broad range of densities for Θ1.

In addition, one might consider other properties such as the static density response, static structure factor, and the positive frequency moments of See(q,ω) that can be extracted from the derivatives of Fee(q,τ) with respect to τ around τ=0.713 Finally, we mention the enticing possibility to develop a sufficiently constrained analytic continuation (possibly by extending the stochastic sampling of the dynamic local field correction introduced in Ref. 166) to invert Eq. (121), and, in this way, obtain FPIMC results for See(q,ω) and a number of other dynamic observables.167,168

These efforts might be complemented by considering known analytical limits (cf. Sec. IV E 5), and to derive new relations that extend previous results for the more simple UEG model (e.g., Refs. 166 and 419). Finally, we note that, while unassailable benchmarks for real WDM systems in non-equilibrium are currently lacking, one can benchmark time-dependent methods such as NEGF or RT-TDDFT against each other for well controlled test cases, such as lattice models (e.g., Refs. 274, 298, and 804) or small systems. This might be a viable way to derive and test improved non-adiabatic XC correlation potentials and to obtain insights into dynamic effects beyond the linear response regime.

To conclude this article, we are confident that the field of dense hydrogen is entering a new era—the era of high-precision simulations. We have discussed the arsenal of simulation methods that are available to compute relevant observables, based on first principles, and achieve reliable results that have predictive power. We outlined possible strategies to combine the strengths of different approaches to develop high precision simulations for experimentally relevant situations.

MB acknowledges fruitful discussions with P.R. Levashov and V.S. Filinov on PIMC simulations for hydrogen. This work was partially supported by the Center for Advanced Systems Understanding (CASUS) which is 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. MB acknowledges funding by the Deutsche Forschungsgemeinschaft via project Nos. BO1366-13/2 and BO1366-16. TD acknowledges funding from the European Research Council (ERC) under the European Union's Horizon 2022 research and innovation programme (Grant agreement No. 101076233, “PREXTREME”). Views and opinions expressed are however those of the authors only and do not necessarily reflect those of the European Union or the European Research Council Executive Agency. Neither the European Union nor the granting authority can be held responsible for them. TD acknowledges funding from the European Union's Just Transition Fund (JTF) within the project Röntgenlaser-Optimierung der Laserfusion (ROLF), Contract No. 5086999001, co-financed by the Saxon state government out of the State budget approved by the Saxon State Parliament. DC was supported by DOE No. DE-SC0020177 and BM by DOE No. DE-NA0004147. CP was supported by the European Union – NextGenerationEU under the Italian Ministry of University and Research (MUR) project Nos. PRIN2022-2022NRBLPT CUP E53D23001790006 and PRIN2022-P2022MC742PNRR, CUP E53D23018440001. PS acknowledges funding from OXPEG and AWE UK. RR acknowledges funding by the Deutsche Forschungsgemeinschaft via the Research Unit FOR 2440. SBH was supported by Sandia National Laboratories, a multimission laboratory managed and operated by National Technology and Engineering Solutions of Sandia, LLC, a wholly owned subsidiary of Honeywell International Inc., for the U.S. Department of Energy's National Nuclear Security Administration under Contract No. DE-NA0003525. This paper describes objective technical results and analysis. Any subjective views or opinions that might be expressed in the paper do not necessarily represent the views of the U.S. Department of Energy or the United States Government. Some of the authors would like to thank the Institut Henri Poincaré (No. UAR 839 CNRS-Sorbonne Université) and the LabEx CARMIN (No. ANR-10-LABX-59-01) for their support. Portions of this work were performed under the auspices of the U.S. Department of Energy by Lawrence Livermore National Laboratory under Contract No. DE-AC52-07NA27344 Lawrence Livermore National Security, LLC. SXH and VVK acknowledge the support by the Department of Energy [National Nuclear Security Administration] University of Rochester “National Inertial Confinement Program” under Award No. DE-NA0004144 and U.S. National Science Foundation PHY Grant No. 2205521.

The PIMC and DFT calculations were partly carried out at the Norddeutscher Verbund für Hoch- und Höchstleistungsrechnen (HLRN) under Grant Nos. shp00026, mvp00018, and mvp00024, on a Bull Cluster at the Center for Information Services and High Performance Computing (ZIH) at Technische Universität Dresden, and on the HoreKa supercomputer funded by the Ministry of Science, Research and the Arts Baden-Württemberg and by the Federal Ministry of Education and Research.

The authors have no conflicts to disclose.

Michael Bonitz: Conceptualization (lead); Project administration (lead); Writing – original draft (equal). Paul Hamann: Investigation (equal); Writing – original draft (equal). Stephanie B. Hansen: Investigation (equal); Writing – original draft (equal). Markus Holzmann: Investigation (equal); Writing – original draft (equal). Suxing Hu: Investigation (equal); Writing – original draft (equal). Hanno Kählert: Investigation (equal); Writing – original draft (equal). Valentin V. Karasiev: Investigation (equal); Writing – original draft (equal). Uwe Kleinschmidt: Investigation (equal). Linda Kordts: Investigation (equal); Writing – original draft (equal). Christopher Makait: Investigation (equal); Writing – original draft (equal). Burkhard Militzer: Investigation (equal); Writing – original draft (equal). Jan Vorberger: Conceptualization (lead); Investigation (equal); Writing – original draft (lead). Zhandos Moldabekov: Investigation (equal); Writing – original draft (equal). Carlo Pierleoni: Investigation (equal); Writing – original draft (equal). Martin Preising: Investigation (equal). Kushal Ramakrishna: Investigation (equal). Ronald Redmer: Investigation (equal); Writing – original draft (equal). Sebastian Schwalbe: Investigation (equal). Pontus Erik Martin Svensson: Investigation (equal). Tobias Dornheim: Conceptualization (lead); Investigation (equal); Writing – original draft (lead). Mandy Bethkenhagen: Investigation (equal); Writing – original draft (equal). Maximilian Böhme: Investigation (equal). David M. Ceperley: Conceptualization (equal); Writing – original draft (equal). Alexey Filinov: Investigation (equal). Thomas Gawne: Writing – original draft (equal). Frank Graziani: Writing – original draft (equal). Gianluca Gregori: Writing – original draft (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

TABLE IX.

Acronyms used in this article and place of their first/detailed reference.

Acronym Explanation Section
AA  Average atom model  III D 
AIMD  Ab initio MD (= BOMD)  III C 
AIRSS  Ab initio random structure search  II B 2 
ALDA  Adiabatic LDA  III C 
BOA  Born–Oppenheimer approximation  III C 
BOMD  Born–Oppenheimer MD  III C 
CEIMC  Coupled electron–ion MC  III F 
CPIMC  Configuration PIMC  III E 3 
DAC  Diamond anvil cell  I 
DFT  Density functional theory  III C 
DFT-AA  DFT-based average atom model  III D 
DFT-MD  DFT with MD for ions (=BOMD)  III C 
DMQMC  Density matrix QMC  III E 3 
DSF  Dynamic structure factor  IV D 
EOS  Equation of state  IV A 
FCIQMC  Full-CI QMC  III F 
FSP  Fermion sign problem  III E 2 
FP-PIMC  Fermionic propagator PIMC  115  
FVT  Fluid variational theory  III B 2 
GDSMFB  Finite-T LDA by Groth et al.  128  
GGA  Generalized gradient approximation  III C 
ICF  Inertial confinement fusion  I 
IPD  Ionization potential depression  IV B 2 
IMT  Insulator to metal transition  II A 
ITCF  Imaginary time correlation function  IV D 2 
KG  Kubo–Greenwood  IV E 
KS-DFT  Kohn–Sham DFT  III C 
KSDT  Finite-T LDA by Karasiev et al.  136  
LDA  Local density approximation  III C 
LFC  Local field correction  IV D 
LLPT  Liquid–liquid phase transition  II B 2 
LR-TDDFT  Linear response TDDFT  III H 4 
MC  Monte Carlo  III E 
MD  Molecular dynamics  III C 
ML-QMC  Machine learning QMC  III G 3 
NEGF  Noequilibrium Green functions  III H 3 
NIF  National Ignition Facility  I 
OCP  One-component plasma model  II A 
PBE  Perdew–Burke–Ernzerhof  III C 
OF-DFT  Orbital-free DFT  III C 
PB-PIMC  Permutation blocking PIMC  III E 6 
PDF  Pair distribution function  IV A 5 
PIMC  Path integral MC  III E 
PIMD  Path integral MD  III G 
PPT  Plasma phase transition  II B 1 
RPA  Random phase approximation  IV D 
QMC  Quantum Monte Carlo  III E 
RPIMC  Restricted PIMC  III E 4 
RT-TDDFT  Real-time TDDFT  III H 4 
TB-MD  tight-binding molecular dynamics  III B 2 
TDDFT  Time-dependent DFT  III H 4 
UEG  Uniform electron gas model (jellium)  III E 1 
VMC  Variational MC  III F 
XC  Exchange–correlation  III C 1 
XFEL  X-ray free electron laser  II C 
XRTS  X-ray Thomson scattering  I 
Acronym Explanation Section
AA  Average atom model  III D 
AIMD  Ab initio MD (= BOMD)  III C 
AIRSS  Ab initio random structure search  II B 2 
ALDA  Adiabatic LDA  III C 
BOA  Born–Oppenheimer approximation  III C 
BOMD  Born–Oppenheimer MD  III C 
CEIMC  Coupled electron–ion MC  III F 
CPIMC  Configuration PIMC  III E 3 
DAC  Diamond anvil cell  I 
DFT  Density functional theory  III C 
DFT-AA  DFT-based average atom model  III D 
DFT-MD  DFT with MD for ions (=BOMD)  III C 
DMQMC  Density matrix QMC  III E 3 
DSF  Dynamic structure factor  IV D 
EOS  Equation of state  IV A 
FCIQMC  Full-CI QMC  III F 
FSP  Fermion sign problem  III E 2 
FP-PIMC  Fermionic propagator PIMC  115  
FVT  Fluid variational theory  III B 2 
GDSMFB  Finite-T LDA by Groth et al.  128  
GGA  Generalized gradient approximation  III C 
ICF  Inertial confinement fusion  I 
IPD  Ionization potential depression  IV B 2 
IMT  Insulator to metal transition  II A 
ITCF  Imaginary time correlation function  IV D 2 
KG  Kubo–Greenwood  IV E 
KS-DFT  Kohn–Sham DFT  III C 
KSDT  Finite-T LDA by Karasiev et al.  136  
LDA  Local density approximation  III C 
LFC  Local field correction  IV D 
LLPT  Liquid–liquid phase transition  II B 2 
LR-TDDFT  Linear response TDDFT  III H 4 
MC  Monte Carlo  III E 
MD  Molecular dynamics  III C 
ML-QMC  Machine learning QMC  III G 3 
NEGF  Noequilibrium Green functions  III H 3 
NIF  National Ignition Facility  I 
OCP  One-component plasma model  II A 
PBE  Perdew–Burke–Ernzerhof  III C 
OF-DFT  Orbital-free DFT  III C 
PB-PIMC  Permutation blocking PIMC  III E 6 
PDF  Pair distribution function  IV A 5 
PIMC  Path integral MC  III E 
PIMD  Path integral MD  III G 
PPT  Plasma phase transition  II B 1 
RPA  Random phase approximation  IV D 
QMC  Quantum Monte Carlo  III E 
RPIMC  Restricted PIMC  III E 4 
RT-TDDFT  Real-time TDDFT  III H 4 
TB-MD  tight-binding molecular dynamics  III B 2 
TDDFT  Time-dependent DFT  III H 4 
UEG  Uniform electron gas model (jellium)  III E 1 
VMC  Variational MC  III F 
XC  Exchange–correlation  III C 1 
XFEL  X-ray free electron laser  II C 
XRTS  X-ray Thomson scattering  I 
1.
D.
Gough
, “
Inverting helioseismic data
,”
Sol. Phys.
100
,
65
99
(
1985
).
2.
J.
Christensen-Dalsgaard
,
T. L.
Duvall
,
D. O.
Gough
,
J. W.
Harvey
, and
E. J.
Rhodes
, “
Speed of sound in the solar interior
,”
Nature
315
,
378
382
(
1985
).
3.
D. O.
Gough
,
A. G.
Kosovichev
,
J.
Toomre
,
E.
Anderson
,
H. M.
Antia
,
S.
Basu
,
B.
Chaboyer
,
S. M.
Chitre
,
J.
Christensen-Dalsgaard
,
W. A.
Dziembowski
et al, “
The seismic structure of the sun
,”
Science
272
,
1296
1300
(
1996
).
4.
R. J.
Hemley
and
H. K.
Mao
, “
Phase transition in solid molecular hydrogen at ultrahigh pressures
,”
Phys. Rev. Lett.
61
,
857
860
(
1988
).
5.
W. J.
Nellis
,
A. C.
Mitchell
,
P. C.
McCandless
,
D. J.
Erskine
, and
S. T.
Weir
, “
Electronic energy gap of molecular hydrogen from electrical conductivity measurements at high shock pressures
,”
Phys. Rev. Lett.
68
,
2937
2940
(
1992
).
6.
S. T.
Weir
,
A. C.
Mitchell
, and
W. J.
Nellis
, “
Metallization of fluid molecular hydrogen at 140 GPa (1.4 Mbar)
,”
Phys. Rev. Lett.
76
,
1860
1863
(
1996
).
7.
F. V.
Grigorev
,
S. B.
Kormer
,
O. L.
Mikhailova
,
A. P.
Tolochko
, and
V. D.
Urlin
, “
Experimental determination of the compressibility of hydrogen at densities 0.5–2 g/cm3. Metallization of hydrogen
,”
Sov. J. Exp. Theor. Phys. Lett.
16
,
201
(
1972
).
8.
V. E.
Fortov
,
R. I.
Ilkaev
,
V. A.
Arinin
,
V. V.
Burtzev
,
V. A.
Golubev
,
I. L.
Iosilevskiy
,
V. V.
Khrustalev
,
A. L.
Mikhailov
,
M. A.
Mochalov
,
V. Y.
Ternovoi
et al, “
Phase transition in a strongly nonideal deuterium plasma generated by quasi-isentropical compression at megabar pressures
,”
Phys. Rev. Lett.
99
,
185001
(
2007
).
9.
L. B.
Da Silva
,
P.
Celliers
,
G. W.
Collins
,
K. S.
Budil
,
N. C.
Holmes
,
T. W.
Barbee
, Jr.
,
B. A.
Hammel
,
J. D.
Kilkenny
,
R. J.
Wallace
,
M.
Ross
et al, “
Absolute equation of state measurements on shocked liquid deuterium up to 200 GPa (2 Mbar)
,”
Phys. Rev. Lett.
78
,
483
486
(
1997
).
10.
J.
Nuckols
,
L.
Wood
,
A.
Thiessen
, and
G.
Zimmermann
, “
Laser compression of matter to super-high densities: Thermonuclear (CTR) applications
,”
Nature
239
,
139
142
(
1972
).
11.
J.
Lindl
, “
Development of the indirect‐drive approach to inertial confinement fusion and the target physics basis for ignition and gain
,”
Phys. Plasmas
2
,
3933
4024
(
1995
).
12.
R. S.
Craxton
,
K. S.
Anderson
,
T. R.
Boehly
,
V. N.
Goncharov
,
D. R.
Harding
,
J. P.
Knauer
,
R. L.
McCrory
,
P. W.
McKenty
,
D. D.
Meyerhofer
,
J. F.
Myatt
et al, “
Direct-drive inertial confinement fusion: A review
,”
Phys. Plasmas
22
,
110501
(
2015
).
13.
H.
Abu-Shawareb
,
R.
Acree
,
P.
Adams
,
J.
Adams
,
B.
Addis
,
R.
Aden
,
P.
Adrian
,
B.
Afeyan
,
M.
Aggleton
,
L.
Aghaian
et al, “
Lawson criterion for ignition exceeded in an inertial fusion experiment
,”
Phys. Rev. Lett.
129
,
075001
(
2022
).
14.
H.
Abu-Shawareb
,
R.
Acree
,
P.
Adams
,
B.
Addis
, and
R.
Aden
, “
Achievement of target gain larger than unity in an inertial fusion experiment
,”
Phys. Rev. Lett.
132
,
065102
(
2024
).
15.
O. A.
Hurricane
,
D. A.
Callahan
,
D. T.
Casey
,
A. R.
Christopherson
,
A. L.
Kritcher
,
O. L.
Landen
,
S. A.
Maclaren
,
R.
Nora
,
P. K.
Patel
,
J.
Ralph
et al, “
Energy principles of scientific breakeven in an inertial fusion experiment
,”
Phys. Rev. Lett.
132
,
065103
(
2024
).
16.
A.
Pak
,
A. B.
Zylstra
,
K. L.
Baker
,
D. T.
Casey
,
E.
Dewald
,
L.
Divol
,
M.
Hohenberger
,
A. S.
Moore
,
J. E.
Ralph
,
D. J.
Schlossberg
et al, “
Observations and properties of the first laboratory fusion experiment to exceed a target gain of unity
,”
Phys. Rev. E
109
,
025203
(
2024
).
17.
T.
Dornheim
,
Z. A.
Moldabekov
,
K.
Ramakrishna
,
P.
Tolias
,
A. D.
Baczewski
,
D.
Kraus
,
T. R.
Preston
,
D. A.
Chapman
,
M. P.
Böhme
,
T.
Döppner
et al, “
Electronic density response of warm dense matter
,”
Phys. Plasmas
30
,
032705
(
2023
).
18.
R.
Fowler
, “
On dense matter
,”
Mon. Not. R. Astron. Soc.
87
,
114
122
(
1926
).
19.
F.
Hund
,
Matter under Very High Presseure and Temperature (in German)
, Ergebnisse der Exakten Naturwissenschaften, Vol.
15
(
Springer
,
Berlin
,
1936
).
20.
E.
Wigner
and
H. B.
Huntington
, “
On the possibility of a metallic modification of hydrogen
,”
J. Chem. Phys.
3
,
764
770
(
1935
).
21.
N. W.
Ashcroft
, “
Metallic hydrogen: A high-temperature superconductor?
,”
Phys. Rev. Lett.
21
,
1748
1749
(
1968
).
22.
R. J.
Hemley
and
H. K.
Mao
, “
Optical studies of hydrogen above 200 gigapascals: Evidence for metallization by band overlap
,”
Science
244
,
1462
1465
(
1989
).
23.
R. J.
Hemley
and
H. K.
Mao
, “
Critical behavior in the hydrogen insulator-metal transition
,”
Science
249
,
391
393
(
1990
).
24.
P.
Loubeyre
,
F.
Occelli
, and
P.
Dumas
, “
Synchrotron infrared spectroscopic evidence of the probable transition to metal hydrogen
,”
Nature
577
,
631
635
(
2020
).
25.
Y.-K.
Tsang
and
C. A.
Jones
, “
Characterising jupiter's dynamo radius using its magnetic energy spectrum
,”
Earth Planet. Sci. Lett.
530
,
115879
(
2020
).
26.
B.
Militzer
,
F.
Soubiran
,
S. M.
Wahl
, and
W.
Hubbard
, “
Understanding Jupiter's Interior
,”
J. Geophys. Res. Planets
121
,
1552
, https://doi.org/10.1002/2016JE005080 (
2016
).
27.
R.
Helled
,
G.
Mazzola
, and
R.
Redmer
, “
Understanding dense hydrogen at planetary conditions
,”
Nat. Rev. Phys.
2
,
562
574
(
2020
).
28.
D.
Durante
,
D. R.
Buccino
,
G.
Tommei
,
M.
Parisi
,
D.
Serra
,
M.
Zannoni
,
V.
Notaro
,
P.
Racioppa
,
G.
Lari
,
L. G.
Casajus
,
L.
Iess
,
W. M.
Folkner
,
P.
Tortora
, and
S. J.
Bolton
,
Geophys. Res. Lett.
47
,
e2019
, https://doi.org/10.1029/2019GL086572 (
2020
).
29.
L.
Iess
,
B.
Militzer
,
Y.
Kaspi
,
P.
Nicholson
,
D.
Durante
,
P.
Racioppa
,
A.
Anabtawi
,
E.
Galanti
,
W.
Hubbard
,
M. J.
Mariani
et al, “
Measurement and implications of Saturn's gravity field and ring mass
,”
Science
364
,
eaat2965
(
2019
).
30.
D.
Saumon
and
T.
Guillot
, “
Shock compression of deuterium and the interiors of jupiter and saturn
,”
Astrophys. J.
609
,
1170
(
2004
).
31.
S.
Scandolo
, “
Liquid–liquid phase transition in compressed hydrogen from first-principles simulations
,”
Proc. Natl. Acad. Sci. U. S. A.
100
,
3051
3053
(
2003
).
32.
M. A.
Morales
,
C.
Pierleoni
,
E.
Schwegler
, and
D. M.
Ceperley
, “
Evidence for a first-order liquid-liquid transition in high-pressure hydrogen from ab initio simulations
,”
Proc. Natl. Acad. Sci. U. S. A.
107
,
12799
12803
(
2010
).
33.
W.
Lorenzen
,
B.
Holst
, and
R.
Redmer
, “
First-order liquid-liquid phase transition in dense hydrogen
,”
Phys. Rev. B
82
,
195107
(
2010
).
34.
M. A.
Morales
,
E.
Schwegler
,
D.
Ceperley
,
C.
Pierleoni
,
S.
Hamel
, and
K.
Caspersen
, “
Phase separation in hydrogen–helium mixtures at Mbar pressures
,”
Proc. Natl. Acad. Sci. U. S. A.
106
,
1324
1329
(
2009
).
35.
W.
Lorenzen
,
B.
Holst
, and
R.
Redmer
, “
Demixing of hydrogen and helium at megabar pressures
,”
Phys. Rev. Lett.
102
,
115701
(
2009
).
36.
M.
Schöttler
and
R.
Redmer
, “
Ab initio calculation of the miscibility diagram for hydrogen-helium mixtures
,”
Phys. Rev. Lett.
120
,
115703
(
2018
).
37.
S. M.
Wahl
,
W. B.
Hubbard
,
B.
Militzer
,
T.
Guillot
,
Y.
Miguel
,
N.
Movshovitz
,
Y.
Kaspi
,
R.
Helled
,
D.
Reese
,
E.
Galanti
et al, “
Comparing Jupiter interior structure models to Juno gravity measurements and the role of a dilute core
,”
Geophys. Res. Lett.
44
,
4649
4659
, https://doi.org/10.1002/2017GL073160 (
2017
).
38.
B.
Militzer
,
W. B.
Hubbard
,
S.
Wahl
,
J. I.
Lunine
,
E.
Galanti
,
Y.
Kaspi
,
Y.
Miguel
,
T.
Guillot
,
K. M.
Moore
,
M.
Parisi
et al, “
Juno spacecraft measurements of Jupiter's gravity imply a dilute core
,”
Planet. Sci. J.
3
,
185
(
2022
).
39.
D.
Saumon
,
G.
Chabrier
, and
H. M.
van Horn
, “
An equation of state for low-mass stars and giant planets
,”
Astrophys. J. Sci.
99
,
713
(
1995
).
40.
B.
Militzer
and
W. B.
Hubbard
, “
Ab initio equation of state for hydrogen-helium mixtures with recalibration of the giant-planet mass-radius relation
,”
Astrophys. J.
774
,
148
(
2013
).
41.
G.
Chabrier
,
S.
Mazevet
, and
F.
Soubiran
, “
A new equation of state for dense hydrogen–helium mixtures
,”
Astrophys. J.
872
,
51
(
2019
).
42.
W. B.
Hubbard
and
B.
Militzer
, “
A preliminary Jupiter model
,”
Astrophys. J.
820
,
80
(
2016
).
43.
R.
Helled
,
D. J.
Stevenson
,
J. I.
Lunine
,
S. J.
Bolton
,
N.
Nettelmann
,
S.
Atreya
,
T.
Guillot
,
B.
Militzer
,
Y.
Miguel
, and
W. B.
Hubbard
, “
Revelations on jupiter's formation, evolution and interior: Challenges from juno results
,”
Icarus
378
,
114937
(
2022
).
44.
B.
Militzer
, “
Study of jupiter's interior with quadratic monte carlo simulations
,”
Astrophys. J.
953
,
111
(
2023
).
45.
S.
Howard
,
T.
Guillot
,
M.
Bazot
,
Y.
Miguel
,
D. J.
Stevenson
,
E.
Galanti
,
Y.
Kaspi
,
W. B.
Hubbard
,
B.
Militzer
,
R.
Helled
et al, “
Jupiter's interior from juno: Equation-of-state uncertainties and dilute core extent
,”
Astron. Astrophys.
672
,
A33
(
2023
).
46.
Y.
Miguel
,
M.
Bazot
,
T.
Guillot
,
S.
Howard
,
E.
Galanti
,
Y.
Kaspi
,
W. B.
Hubbard
,
B.
Militzer
,
R.
Helled
,
S. K.
Atreya
et al, “
Jupiter's inhomogeneous envelope
,”
Astron. Astrophys.
662
,
A18
(
2022
).
47.
F.
Debras
,
G.
Chabrier
, and
D. J.
Stevenson
, “
Superadiabaticity in Jupiter and giant planet interiors
,”
Astrophys. J. Lett.
913
,
21
(
2021
).
48.
M.
French
,
A.
Becker
,
W.
Lorenzen
,
N.
Nettelmann
,
M.
Bethkenhagen
,
J.
Wicht
, and
R.
Redmer
, “
Ab initio simulations for material properties along the Jupiter adiabat
,”
Astrophys. J. Sci.
202
,
5
(
2012
).
49.
M.
Preising
,
M.
French
,
C.
Mankovich
,
F.
Soubiran
, and
R.
Redmer
, “
Material properties of Saturn's interior from ab initio simulations
,”
Astrophys. J. Sci.
269
,
47
(
2023
).
50.
K. H.
Schramm
, “
Ionisation und Zustandsgleichung eines Plasmas unter hohen Drucken und Temperaturen
,”
Z. Phys.
165
,
336
355
(
1961
).
51.
G.
Norman
and
A.
Starostin
, “
Insufficiency of the classical description of a nondegenerate dense plasma
,”
High Temp.
6
,
394
(
1968
).
52.
W.
Ebeling
and
R.
Sändig
, “
Theory of the lonization equilibrium in dense plasmas
,”
Ann. Phys.
483
,
289
305
(
1973
).
53.
W.
Ebeling
and
W.
Richert
, “
Plasma phase transition in hydrogen
,”
Phys. Lett. A
108
,
80
82
(
1985
).
54.
D. G.
Hummer
and
D.
Mihalas
, “
The equation of state for stellar envelopes. i. an occupation probability formalism for the truncation of internal partition functions
,”
Astrophys. J.
331
,
794
(
1988
).
55.
D.
Mihalas
,
W.
Dappen
, and
D. G.
Hummer
, “
The equation of state for stellar envelopes. II. Algorithm and selected results
,”
Astrophys. J.
331
,
815
(
1988
).
56.
W.
Daeppen
,
D.
Mihalas
,
D. G.
Hummer
, and
B. W.
Mihalas
, “
The equation of state for stellar envelopes. III. Thermodynamic quantities
,”
Astrophys. J.
332
,
261
(
1988
).
57.
D.
Saumon
,
W. B.
Hubbard
,
G.
Chabrier
, and
H. M.
van Horn
, “
The role of the molecular-metallic transition of hydrogen in the evolution of Jupiter, Saturn, and brown dwarfs
,”
Astrophys. J.
391
,
827
831
(
1992
).
58.
D.
Saumon
and
G.
Chabrier
, “
Fluid hydrogen at high density: Pressure ionization
,”
Phys. Rev. A
46
,
2084
2100
(
1992
).
59.
M.
Schlanges
,
M.
Bonitz
, and
A.
Tschttschjan
, “
Plasma phase transition in fluid hydrogen–helium mixtures
,”
Contrib. Plasma Phys.
35
,
109
(
1995
).
60.
V.
Bezkrovniy
,
V. S.
Filinov
,
D.
Kremp
,
M.
Bonitz
,
M.
Schlanges
,
W. D.
Kraeft
,
P. R.
Levashov
, and
V. E.
Fortov
, “
Monte Carlo results for the hydrogen Hugoniot
,”
Phys. Rev. E
70
,
057401
(
2004
).
61.
H.
Reinholz
,
R.
Redmer
, and
S.
Nagel
, “
Thermodynamic and transport properties of dense hydrogen plasma
,”
Phys. Rev. E
52
,
5368
(
1995
).
62.
H.
Juranek
and
R.
Redmer
, “
Self-consistent fluid variational theory for pressure dissociation in dense hydrogen
,”
J. Chem. Phys.
112
,
3780
3786
(
2000
).
63.
H.
Juranek
,
R.
Redmer
, and
Y.
Rosenfeld
, “
Fluid variational theory for pressure dissociation in dense hydrogen: Multicomponent reference system and nonadditivity effects
,”
J. Chem. Phys.
117
,
1768
1774
(
2002
).
64.
Y. L.
Klimontovich
,
Kinetic Theory of Nonideal Gases and Nonideal Plasmas
(
Pergamon Press
,
1982
).
65.
Y. L.
Klimontovich
,
Statistical Physics
(
Harwood Acad. Publ.
,
1986
).
66.
S.
Ichimaru
,
H.
Iyetomi
, and
S.
Tanaka
, “
Statistical physics of dense plasmas: Thermodynamics, transport coefficients and dynamic correlations
,”
Phys. Rep.
149
,
91
205
(
1987
).
67.
W.
Ebeling
,
V.
Fortov
, and
V.
Filinov
,
Quantum Statistics of Dense Gases and Nonideal Plasmas
(
Springer
,
Heidelberg
,
2017
).
68.
W.
Ebeling
,
W.
Kraeft
, and
D.
Kremp
,
Theory of Bound States and Ionization Equilibrium in Plasmas and Solids
(
Akademie-Verlag
,
Berlin
,
1976
).
69.
W.-D.
Kraeft
,
D.
Kremp
,
W.
Ebeling
, and
G.
Röpke
,
Quantum Statistics of Charged Particle Systems
(
Akademie-Verlag
,
Berlin
,
1986
).
70.
M.
Bonitz
,
Quantum Kinetic Theory
, 2nd ed. (
Springer
,
Cham
,
2016
).
71.
D.
Zubarev
,
V.
Morozov
, and
G.
Röpke
,
Statistical Mechanics of Nonequilibrium Processes, Statistical Mechanics of Nonequilibrium Processes
. Basic Concepts, Kinetic Theory, Statistical Mechanics of Nonequilibrium Processes, Vol.
1
(
Wiley
,
1996
).
72.
D.
Zubarev
,
V.
Morozov
, and
G.
Röpke
,
Statistical Mechanics of Nonequilibrium Processes, Statistical Mechanics of Nonequilibrium Processes
. Relaxation and Hydrodynamic Processes, Statistical Mechanics of Nonequilibrium Processes, Vol.
2
(
Wiley
,
1997
).
73.
W. M. C.
Foulkes
,
L.
Mitas
,
R. J.
Needs
, and
G.
Rajagopal
, “
Quantum Monte Carlo simulations of solids
,”
Rev. Mod. Phys.
73
,
33
83
(
2001
).
74.
R. O.
Jones
, “
Density functional theory: Its origins, rise to prominence, and future
,”
Rev. Mod. Phys.
87
,
897
923
(
2015
).
75.
L. D.
Fosdick
and
H. F.
Jordan
, “
Path-integral calculation of the two-particle slater sum for He4
,”
Phys. Rev.
143
,
58
66
(
1966
).
76.
H. F.
Jordan
and
L. D.
Fosdick
, “
Three-particle effects in the pair distribution function for He4 gas
,”
Phys. Rev.
171
,
128
149
(
1968
).
77.
V.
Filinov
and
G.
Norman
, “
On phase transition in a non-ideal plasma
,”
Phys. Lett. A
55
,
219
(
1975
).
78.
V. M.
Zamalin
,
G. E.
Norman
, and
V. S.
Filinov
,
The Monte-Carlo Method in Statistical Thermodynamics
(
Nauka
,
Moscow
,
1977
).
79.
M.
Troyer
and
U. J.
Wiese
, “
Computational complexity and fundamental limitations to fermionic quantum Monte Carlo simulations
,”
Phys. Rev. Lett.
94
,
170201
(
2005
).
80.
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
).
81.
E. Y.
Loh
,
J. E.
Gubernatis
,
R. T.
Scalettar
,
S. R.
White
,
D. J.
Scalapino
, and
R. L.
Sugar
, “
Sign problem in the numerical simulation of many-electron systems
,”
Phys. Rev. B
41
,
9301
9307
(
1990
).
82.
D. M.
Ceperley
, “
Fermion nodes
,”
J. Stat. Phys.
63
,
1237
1267
(
1991
).
83.
B.
Militzer
,
E.
Pollock
, and
D.
Ceperley
, “
Path integral Monte Carlo calculation of the momentum distribution of the homogeneous electron gas at finite temperature
,”
High Energy Density Phys.
30
,
13
20
(
2019
).
84.
C.
Pierleoni
,
D.
Ceperley
,
B.
Bernu
, and
W.
Magro
, “
Equation of state of the hydrogen plasma by path integral Monte Carlo simulation
,”
Phys. Rev. Lett.
73
,
2145
2149
(
1994
).
85.
W. R.
Magro
,
D. M.
Ceperley
,
C.
Pierleoni
, and
B.
Bernu
, “
Molecular dissociation in hot, dense hydrogen
,”
Phys. Rev. Lett.
76
,
1240
1243
(
1996
).
86.
T. J.
Lenosky
,
J. D.
Kress
,
L. A.
Collins
, and
I.
Kwon
, “
Molecular-dynamics modeling of shock-compressed liquid hydrogen
,”
Phys. Rev. B
55
,
R11907
R11910
(
1997
).
87.
B.
Militzer
and
E. L.
Pollock
, “
Variational density matrix method for warm, condensed matter: Application to dense hydrogen
,”
Phys. Rev. E
61
,
3470
3482
(
2000
).
88.
J.
Vorberger
,
I.
Tamblyn
,
B.
Militzer
, and
S. A.
Bonev
, “
Hydrogen-helium mixtures in the interiors of giant planets
,”
Phys. Rev. B
75
,
024206
(
2007
).
89.
M. P.
Desjarlais
, “
Density-functional calculations of the liquid deuterium hugoniot, reshock, and reverberation timing
,”
Phys. Rev. B
68
,
064204
(
2003
).
90.
S. A.
Bonev
,
E.
Schwegler
,
T.
Ogitsu
, and
G.
Galli
, “
A quantum fluid of metallic hydrogen suggested by first-principles calculations
,”
Nature
431
,
669
672
(
2004
).
91.
We note that, while DFT is formally exact given the correct exchange–correlation functional, DFT-MD simulations are based on the Born–Oppenheimer approximation and, strictly speaking, not a first-principles approach, according to our definition.
92.
G. W.
Collins
,
L. B. D.
Silva
,
P.
Celliers
,
D. M.
Gold
,
M. E.
Foord
,
R. J.
Wallace
,
A.
Ng
,
S. V.
Weber
,
K. S.
Budil
, and
R.
Cauble
, “
Measurements of the equation of state of deuterium at the fluid insulator-metal transition
,”
Science
281
,
1178
1181
(
1998
).
93.
G.
Kerley
,
Molcular Based Study of fluids
(
ACS
,
Washington, D.C
.,
1983
).
94.
B.
Militzer
and
D. M.
Ceperley
, “
Path integral monte carlo calculation of the deuterium hugoniot
,”
Phys. Rev. Lett.
85
,
1890
1893
(
2000
).
95.
M.
Ross
, “
Linear-mixing model for shock-compressed liquid deuterium
,”
Phys. Rev. B
58
,
669
677
(
1998
).
96.
M. D.
Knudson
,
D. L.
Hanson
,
J. E.
Bailey
,
C. A.
Hall
,
J. R.
Asay
, and
C.
Deeney
, “
Principal hugoniot, reverberating wave, and mechanical reshock measurements of liquid deuterium to 400 gpa using plate impact techniques
,”
Phys. Rev. B
69
,
144209
(
2004
).
97.
D. G.
Hicks
,
T. R.
Boehly
,
P. M.
Celliers
,
J. H.
Eggert
,
S. J.
Moon
,
D. D.
Meyerhofer
, and
G. W.
Collins
, “
Laser-driven single shock compression of fluid deuterium from 45 to 220 GPa
,”
Phys. Rev. B
79
,
014112
(
2009
).
98.
M. D.
Knudson
and
M. P.
Desjarlais
, “
Shock compression of quartz to 1.6 TPa: Redefining a pressure standard
,”
Phys. Rev. Lett.
103
,
225501
(
2009
).
99.
G. V.
Boriskov
,
A. I.
Bykov
,
R. I.
Il'kaev
,
V. D.
Selemir
,
G. V.
Simakov
,
R. F.
Trunin
,
V. D.
Urlin
,
A. N.
Shuikin
, and
W. J.
Nellis
, “
Shock compression of liquid deuterium up to 109 GPa
,”
Phys. Rev. B
71
,
092104
(
2005
).
100.
T.
Sano
,
N.
Ozaki
,
T.
Sakaiya
,
K.
Shigemori
,
M.
Ikoma
,
T.
Kimura
,
K.
Miyanishi
,
T.
Endo
,
A.
Shiroshita
,
H.
Takahashi
et al, “
Laser-shock compression and hugoniot measurements of liquid hydrogen to 55 GPa
,”
Phys. Rev. B
83
,
054117
(
2011
).
101.
A.
Fernandez-Pañella
,
M.
Millot
,
D. E.
Fratanduono
,
M. P.
Desjarlais
,
S.
Hamel
,
M. C.
Marshall
,
D. J.
Erskine
,
P. A.
Sterne
,
S.
Haan
,
T. R.
Boehly
et al, “
Shock compression of liquid deuterium up to 1 TPa
,”
Phys. Rev. Lett.
122
,
255702
(
2019
).
102.
W. J.
Nellis
,
A. C.
Mitchell
,
M.
van Thiel
,
G. J.
Devine
,
R. J.
Trainor
, and
N.
Brown
, “
Equation‐of‐state data for molecular hydrogen and deuterium at shock pressures in the range 2–76 GPa (20–760 kbar)
,”
J. Chem. Phys.
79
,
1480
1486
(
1983
).
103.
R. D.
Dick
and
G. I.
Kerley
, “
Shock compression data for liquids. II. Condensed hydrogen and deuterium
,”
J. Chem. Phys.
73
,
5264
5271
(
1980
).
104.
B.
Holst
,
R.
Redmer
, and
M. P.
Desjarlais
, “
Thermophysical properties of warm dense hydrogen using quantum molecular dynamics simulations
,”
Phys. Rev. B
77
,
184201
(
2008
).
105.
L.
Caillabet
,
S.
Mazevet
, and
P.
Loubeyre
, “
Multiphase equation of state of hydrogen from ab initio calculations in the range 0.2 to 5 g/cc up to 10 eV
,”
Phys. Rev. B
83
,
094101
(
2011
).
106.
C.
Wang
and
P.
Zhang
, “
Wide range equation of state for fluid hydrogen from density functional theory
,”
Phys. Plasmas
20
,
092703
(
2013
).
107.
V. V.
Karasiev
,
S. X.
Hu
,
M.
Zaghoo
, and
T. R.
Boehly
, “
Exchange-correlation thermal effects in shocked deuterium: Softening the principal hugoniot and thermophysical properties
,”
Phys. Rev. B
99
,
214110
(
2019
).
108.
N. M.
Tubman
,
E.
Liberatore
,
C.
Pierleoni
,
M.
Holzmann
, and
D. M.
Ceperley
, “
Molecular-atomic transition along the deuterium hugoniot curve with coupled electron-ion monte carlo simulations
,”
Phys. Rev. Lett.
115
,
045301
(
2015
).
109.
M.
Ruggeri
,
M.
Holzmann
,
D. M.
Ceperley
, and
C.
Pierleoni
, “
Quantum Monte Carlo determination of the principal Hugoniot of deuterium
,”
Phys. Rev. B
102
,
144108
(
2020
).
110.
T. J.
Lenosky
,
J. D.
Kress
, and
L. A.
Collins
, “
Molecular-dynamics modeling of the hugoniot of shocked liquid deuterium
,”
Phys. Rev. B
56
,
5164
5169
(
1997
).
111.
Y.
Lavrinenko
,
P. R.
Levashov
,
D. V.
Minakov
,
I. V.
Morozov
, and
I. A.
Valuev
, “
Equilibrium properties of warm dense deuterium calculated by the wave packet molecular dynamics and density functional theory method
,”
Phys. Rev. E
104
,
045304
(
2021
).
112.
M. D.
Knudson
,
D. L.
Hanson
,
J. E.
Bailey
,
C. A.
Hall
,
J. R.
Asay
, and
W. W.
Anderson
, “
Equation of state measurements in liquid deuterium to 70 GPa
,”
Phys. Rev. Lett.
87
,
225501
(
2001
).
113.
J. R.
Rygg
,
P. M.
Celliers
, and
G. W.
Collins
, “
Specific heat of electron plasma waves
,”
Phys. Rev. Lett.
130
,
225101
(
2023
).
114.
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
).
115.
A.
Filinov
and
M.
Bonitz
, “
The equation of state of partially ionized hydrogen and deuterium plasma revisited
,”
Phys. Rev. E
108
,
055212
(
2023
).
116.
P.-F.
Loos
and
P. M. W.
Gill
, “
The uniform electron gas
,”
WIREs Comput. Mol. Sci.
6
,
410
429
(
2016
).
117.
T.
Dornheim
,
S.
Groth
, and
M.
Bonitz
, “
The uniform electron gas at warm dense matter conditions
,”
Phys. Rep.
744
,
1
86
(
2018
).
118.
T.
Schoof
,
M.
Bonitz
,
A.
Filinov
,
D.
Hochstuhl
, and
J.
Dufty
, “
Configuration path integral Monte Carlo
,”
Contrib. Plasma Phys.
51
,
687
697
(
2011
).
119.
T.
Schoof
,
S.
Groth
, and
M.
Bonitz
, “
Introduction to configuration path integral monte carlo
,” in
Complex Plasmas
, edited by
M.
Bonitz
,
J.
Lopez
,
K.
Becker
, and
H.
Thomsen
(
Springer International Publishing
,
2014
), Vol.
82
, pp.
153
194
.
120.
T.
Schoof
,
S.
Groth
, and
M.
Bonitz
, “
Towards ab initio thermodynamics of the electron gas at strong degeneracy
,”
Contrib. Plasma Phys.
55
,
136
143
(
2015
).
121.
T.
Schoof
,
S.
Groth
,
J.
Vorberger
, and
M.
Bonitz
, “
Ab initio thermodynamic results for the degenerate electron gas at finite temperature
,”
Phys. Rev. Lett.
115
,
130402
(
2015
).
122.
T.
Dornheim
,
S.
Groth
,
A.
Filinov
, and
M.
Bonitz
, “
Permutation blocking path integral Monte Carlo: A highly efficient approach to the simulation of strongly degenerate non-ideal fermions
,”
New J. Phys.
17
,
073017
(
2015
).
123.
T.
Dornheim
,
T.
Schoof
,
S.
Groth
,
A.
Filinov
, and
M.
Bonitz
, “
Permutation blocking path integral Monte Carlo approach to the uniform electron gas at finite temperature
,”
J. Chem. Phys.
143
,
204101
(
2015
).
124.
T.
Dornheim
,
S.
Groth
, and
M.
Bonitz
, “
Permutation blocking path integral Monte Carlo simulations of degenerate electrons at finite temperature
,”
Contrib. Plasma Phys.
59
,
e201800157
(
2019
).
125.
S.
Groth
,
T.
Schoof
,
T.
Dornheim
, and
M.
Bonitz
, “
Ab initio quantum Monte Carlo simulations of the uniform electron gas without fixed nodes
,”
Phys. Rev. B
93
,
085102
(
2016
).
126.
T.
Dornheim
,
S.
Groth
,
T.
Schoof
,
C.
Hann
, and
M.
Bonitz
, “
Ab initio quantum Monte Carlo simulations of the uniform electron gas without fixed nodes: The unpolarized case
,”
Phys. Rev. B
93
,
205134
(
2016
).
127.
T.
Dornheim
,
S.
Groth
,
T.
Sjostrom
,
F. D.
Malone
,
W. M. C.
Foulkes
, and
M.
Bonitz
, “
Ab initio quantum Monte Carlo simulation of the warm dense electron gas in the thermodynamic limit
,”
Phys. Rev. Lett.
117
,
156403
(
2016
).
128.
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
).
129.
T.
Dornheim
,
S.
Groth
,
F. D.
Malone
,
T.
Schoof
,
T.
Sjostrom
,
W. M. C.
Foulkes
, and
M.
Bonitz
, “
Ab initio quantum Monte Carlo simulation of the warm dense electron gas
,”
Phys. Plasmas
24
,
056303
(
2017
).
130.
S.
Groth
,
T.
Dornheim
, and
M.
Bonitz
, “
Free energy of the uniform electron gas: Testing analytical models against first‐principles results
,”
Contrib. Plasma Phys.
57
,
137
146
(
2017
).
131.
S.
Tanaka
and
S.
Ichimaru
, “
Thermodynamics and correlational properties of finite-temperature electron liquids in the Singwi-Tosi-Land-Sjölander approximation
,”
J. Phys. Soc. Jpn.
55
,
2278
2289
(
1986
).
132.
S.
Tanaka
,
S.
Mitake
, and
S.
Ichimaru
, “
Parametrized equation of state for electron liquids in the Singwi-Tosi-Land-Sjölander approximation
,”
Phys. Rev. A
32
,
1896
1899
(
1985
).
133.
T.
Sjostrom
and
J.
Dufty
, “
Uniform electron gas at finite temperatures
,”
Phys. Rev. B
88
,
115123
(
2013
).
134.
S.
Ichimaru
, “
Nuclear fusion in dense plasmas
,”
Rev. Mod. Phys.
65
,
255
299
(
1993
).
135.
F.
Perrot
and
M. W. C.
Dharma-wardana
, “
Spin-polarized electron liquid at arbitrary temperatures: Exchange-correlation energies, electron-distribution functions, and the static response functions
,”
Phys. Rev. B
62
,
16536
(
2000
).
136.
V. V.
Karasiev
,
T.
Sjostrom
,
J.
Dufty
, and
S. B.
Trickey
, “
Accurate homogeneous electron gas exchange-correlation free energy for local spin-density calculations
,”
Phys. Rev. Lett.
112
,
076403
(
2014
).
137.
V. V.
Karasiev
,
L.
Calderin
, and
S. B.
Trickey
, “
Importance of finite-temperature exchange correlation for warm dense matter calculations
,”
Phys. Rev. E
93
,
063207
(
2016
).
138.
K.
Ramakrishna
,
T.
Dornheim
, and
J.
Vorberger
, “
Influence of finite temperature exchange-correlation effects in hydrogen
,”
Phys. Rev. B
101
,
195129
(
2020
).
139.
A.
Filinov
,
P. R.
Levashov
, and
M.
Bonitz
, “
Thermodynamics of the uniform electron gas: Fermionic path integral Monte Carlo simulations in the restricted grand canonical ensemble
,”
Contrib. Plasma Phys.
61
,
e202100112
(
2021
).
140.
M.
Bonitz
,
W.
Ebeling
,
A.
Filinov
,
W.
Kraeft
,
R.
Redmer
, and
G.
Röpke
, “
Günter Kelbg, the Kelbg potential and its impact on quantum plasma theory
,”
Contrib. Plasma Phys.
63
,
e202300029
(
2023
).
141.
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
).
142.
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
).
143.
S.
Groth
,
T.
Dornheim
, and
M.
Bonitz
, “
Configuration path integral Monte Carlo approach to the static density response of the warm dense electron gas
,”
J. Chem. Phys.
147
,
164108
(
2017
).
144.
T.
Dornheim
,
S.
Groth
,
J.
Vorberger
, and
M.
Bonitz
, “
Permutation-blocking path-integral Monte Carlo approach to the static density response of the warm dense electron gas
,”
Phys. Rev. E
96
,
023203
(
2017
).
145.
T.
Dornheim
,
J.
Vorberger
,
S.
Groth
,
N.
Hoffmann
,
Z.
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
).
146.
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
).
147.
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
).
148.
T.
Dornheim
,
S.
Schwalbe
,
P.
Tolias
,
M. P.
Böhme
,
Z. A.
Moldabekov
, and
J.
Vorberger
, “
Ab initio density response and local field factor of warm dense hydrogen
,”
Matter Radiat. Extremes
9
,
057401
(
2024
).
149.
T.
Dornheim
,
S.
Schwalbe
,
M. P.
Böhme
,
Z. A.
Moldabekov
,
J.
Vorberger
, and
P.
Tolias
, “
Ab initio path integral Monte Carlo simulations of warm dense two-component systems without fixed nodes: Structural properties
,”
J. Chem. Phys.
160
,
164111
(
2024
).
150.
C.
Pierleoni
,
D.
Ceperley
, and
M.
Holzmann
, “
Coupled electron-ion Monte Carlo calculations of dense metallic hydrogen
,”
Phys. Rev. Lett.
93
,
146402
(
2004
).
151.
C.
Pierleoni
and
D.
Ceperley
, “
The coupled electron-ion Monte Carlo method
,” in
Computer Simulations in Condensed Matter Systems: From Materials to Chemical Biology
, edited by
M.
Ferrario
,
G.
Ciccotti
, and
K.
Binder
(
Springer Berlin Heidelberg
,
Berlin, Heidelberg
,
2006
), pp.
641
683
.
152.
G.
Mazzola
,
R.
Helled
, and
S.
Sorella
, “
Phase diagram of hydrogen and a hydrogen-helium mixture at planetary conditions by quantum Monte Carlo simulations
,”
Phys. Rev. Lett.
120
,
25701
(
2018
).
153.
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
).
154.
S. X.
Hu
,
V. N.
Goncharov
,
T. R.
Boehly
,
R. L.
McCrory
,
S.
Skupsky
,
L. A.
Collins
,
J. D.
Kress
, and
B.
Militzer
, “
Impact of first-principles properties of deuterium–tritium on inertial confinement fusion target designsa)
,”
Phys. Plasmas
22
,
056304
(
2015
).
155.
T.
Dornheim
,
J.
Vorberger
,
Z.
Moldabekov
,
G.
Röpke
, and
W.-D.
Kraeft
, “
The uniform electron gas at high temperatures: Ab initio path integral Monte Carlo simulations and analytical theory
,”
High Energy Density Phys.
45
,
101015
(
2022
).
156.
J. M.
McMahon
,
M. A.
Morales
,
C.
Pierleoni
, and
D. M.
Ceperley
, “
The properties of hydrogen and helium under extreme conditions
,”
Rev. Mod. Phys.
84
,
1607
1653
(
2012
).
157.
Z.
Moldabekov
,
S.
Schwalbe
,
M.
Böhme
,
J.
Vorberger
,
X.
Shao
,
M.
Pavanello
,
F.
Graziani
, and
T.
Dornheim
, “
Bound state breaking and the importance of thermal exchange-correlation effects in warm dense hydrogen
,”
J Chem Theory Comput.
20
(
1
),
68
78
(
2024
).
158.
S. H.
Glenzer
and
R.
Redmer
, “
X-ray Thomson scattering in high energy density plasmas
,”
Rev. Mod. Phys.
81
,
1625
(
2009
).
159.
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
).
160.
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
).
161.
T.
Dornheim
,
M. P.
Böhme
,
D. A.
Chapman
,
D.
Kraus
,
T. R.
Preston
,
Z. A.
Moldabekov
,
N.
Schlünzen
,
A.
Cangi
,
T.
Döppner
, and
J.
Vorberger
, “
Imaginary-time correlation function thermometry: A new, high-accuracy and model-free temperature analysis technique for x-ray Thomson scattering data
,”
Phys. Plasmas
30
,
042707
(
2023
).
162.
T.
Dornheim
,
T.
Döppner
,
A. D.
Baczewski
,
P.
Tolias
,
M. P.
Böhme
,
Z. A.
Moldabekov
,
T.
Gawne
,
D.
Ranjan
,
D. A.
Chapman
,
M. J.
MacDonald
,
T. R.
Preston
,
D.
Kraus
, and
J.
Vorberger
, “
X-ray Thomson scattering absolute intensity from the f-sum rule in the imaginary-time domain
,”
Sci. Rep.
14
,
14377
(
2024
).
163.
T.
Dornheim
,
T.
Döppner
,
P.
Tolias
,
M.
Böhme
,
L.
Fletcher
,
T.
Gawne
,
F.
Graziani
,
D.
Kraus
,
M.
MacDonald
,
Z.
Moldabekov
et al, “
Unraveling electronic correlations in warm dense quantum plasmas
,” arXiv:2402.19113 (
2024
).
164.
T.
Dornheim
,
Z.
Moldabekov
,
P.
Tolias
,
M.
Böhme
, and
J.
Vorberger
, “
Physical insights from imaginary-time density–density correlation functions
,”
Matter Radiat. Extremes
8
,
056601
(
2023
).
165.
S.
Groth
,
T.
Dornheim
, and
J.
Vorberger
, “
Ab initio path integral Monte Carlo approach to the static and dynamic density response of the uniform electron gas
,”
Phys. Rev. B
99
,
235122
(
2019
).
166.
T.
Dornheim
,
S.
Groth
,
J.
Vorberger
, and
M.
Bonitz
, “
Ab initio path integral Monte Carlo results for the dynamic structure factor of correlated electrons: From the electron liquid to warm dense matter
,”
Phys. Rev. Lett.
121
,
255001
(
2018
).
167.
P.
Hamann
,
T.
Dornheim
,
J.
Vorberger
,
Z.
Moldabekov
, and
M.
Bonitz
, “
Dynamic Properties of the warm dense Electron gas: An ab initio path integral Monte Carlo approach
,”
Phys. Rev. B
102
,
125150
(
2020a
).
168.
P.
Hamann
,
T.
Dornheim
,
J.
Vorberger
,
Z.
Moldabekov
, and
M.
Bonitz
, “
Ab initio results for the plasmon dispersion and damping of the warm dense electron gas
,”
Contrib. Plasma Phys.
60
,
e202000147
(
2020
).
169.
P.
Hamann
,
L.
Kordts
,
A.
Filinov
,
M.
Bonitz
,
T.
Dornheim
, and
J.
Vorberger
, “
Prediction of a roton-type feature in warm dense hydrogen
,”
Phys. Rev. Res.
5
,
033039
(
2023
).
170.
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
).
171.
D.
Kremp
,
M.
Schlanges
, and
W.-D.
Kraeft
,
Quantum Statistics of Nonideal Plasmas
(
Springer
,
Heidelberg
,
2005
).
172.
M.
Bonitz
,
V. S.
Filinov
,
V. E.
Fortov
,
P. R.
Levashov
, and
H.
Fehske
, “
Crystallization in two-component Coulomb systems
,”
Phys. Rev. Lett.
95
,
235006
(
2005
).
173.
F. J.
Rogers
,
H. C.
Graboske
, and
D. J.
Harwood
, “
Bound eigenstates of the static screened coulomb potential
,”
Phys. Rev. A
1
,
1577
1586
(
1970
).
174.
N. F.
Mott
, “
The basis of the electron theory of metals, with special reference to the transition metals
,”
Proc. Phys. Soc. A
62
,
416
(
1949
).
175.
N. F.
Mott
, “
Metal-insulator transition
,”
Rev. Mod. Phys.
40
,
677
683
(
1968
).
176.
M.
Imada
,
A.
Fujimori
, and
Y.
Tokura
, “
Metal-insulator transitions
,”
Rev. Mod. Phys.
70
,
1039
1263
(
1998
).
177.
Y.
Zel'dovich
and
L.
Landau
, “
On the relationship between the liquid and gaseous state in metals
,”
Zh. Eksp. Teor. Fiz.
14
,
32
34
(
1944
).
178.
D. M.
Straus
and
N. W.
Ashcroft
, “
Self-consistent structure of metallic hydrogen
,”
Phys. Rev. Lett.
38
,
415
418
(
1977
).
179.
G.
Norman
and
A.
Starostin
, “
Thermodynamics of a strongly nonideal plasma
,”
High Temp.
8
,
381
408
(
1970
).
180.
S.
Franck
, “
On the dielectric-metal transition in hydrogen
,”
Ann. Phys.
492
,
349
356
(
1980
).
181.
M.
Robnik
and
W.
Kundt
, “
Hydrogen at high pressures and temperatures
,”
Astron. Astrophys.
120
,
227
(
1983
).
182.
D.
Saumon
and
G.
Chabrier
, “
Fluid hydrogen at high density: The plasma phase transition
,”
Phys. Rev. Lett.
62
,
2397
2400
(
1989
).
183.
W.
Ebeling
,
A.
Förster
,
H.
Hess
, and
M. Y.
Romanovsky
, “
Thermodynamic and kinetic properties of hot nonideal plasmas
,”
Plasma Phys. Controlled Fusion
38
,
A31
(
1996
).
184.
D.
Beule
,
W.
Ebeling
,
A.
Förster
,
H.
Juranek
,
S.
Nagel
,
R.
Redmer
, and
G.
Röpke
, “
Equation of state for hydrogen below 10000 K: From the fluid to the plasma
,”
Phys. Rev. B
59
,
14177
14181
(
1999
).
185.
V.
Filinov
,
V.
Fortov
,
M.
Bonitz
, and
P.
Levashov
, “
Phase transition in strongly degenerate hydrogen plasma
,”
Jetp. Lett.
74
,
384
(
2001
).
186.
V.
Filinov
,
M.
Bonitz
,
P.
Levashov
,
V.
Fortov
,
W.
Ebeling
, and
M.
Schlanges
, “
Plasma phase transition in hydrogen and electron-hole plasmas
,”
Contrib. Plasma Phys.
43
,
290
294
(
2003
).
187.
In Ref. 77, the authors published “preliminary estimates” for the two-phase region that extends beyond T= 100 000 K, based on PIMC results with just P=2 high-temperature factors. In Refs. 185 and 186, the authors used P=20 high-temperature factors, N=50 electrons and ions. At T= 10 000 K, in the density range of 0.1g/cm3 1.5g/cm3, they reported strong fluctuations and the formation of “metallic clusters” that they interpreted as evidence for the PPT, without providing evidence for convergence of the simulations.
188.
S. A.
Bonev
,
B.
Militzer
, and
G.
Galli
, “
Ab initio simulations of dense liquid deuterium: Comparison with gas-gun shock-wave experiments
,”
Phys. Rev. B
69
,
014101
(
2004
).
189.
B.
Militzer
and
D. M.
Ceperley
, “
Path integral Monte Carlo simulation of the low-density hydrogen plasma
,”
Phys. Rev. E
63
,
066404
(
2001
).
190.
The analysis is based on the comparison of simple chemical models and to the PIMC simulations of Ref. 185. Also, the reported strong increase of the conductivity in the same density range is not proof of a phase transition.
191.
R.
Redmer
, “
Physical properties of dense, low-temperature plasmas
,”
Phys. Rep.
282
,
35
157
(
1997
).
192.
W.
Ebeling
and
G.
Norman
, “
Coulombic phase transitions in dense plasmas
,”
J. Stat. Phys.
110
,
861
(
2003
).
193.
G. E.
Norman
and
I. M.
Saitov
, “
Plasma phase transition (by the fiftieth anniversary of the prediction)
,”
Contrib. Plasma Phys.
59
,
e201800182
(
2019
).
194.
G.
Norman
and
I. M.
Saitov
, “
Plasma phase transition
,”
Phys-Usp.
64
,
1094
1124
(
2021
).
195.
G.
Chabrier
,
D.
Saumon
, and
A. Y.
Potekhin
, “
Dense plasmas in astrophysics: From giant planets to neutron stars
,”
J. Phys. A: Math. Gen.
39
,
4411
(
2006
).
196.
G.
Chabrier
,
D.
Saumon
, and
C.
Winisdoerffer
, “
Hydrogen and helium at high density and astrophysical implications
,”
Astrophys. Space Sci.
307
,
263
267
(
2007
).
197.
C.
Pierleoni
,
M. a.
Morales
,
G.
Rillo
,
M.
Holzmann
, and
D. M.
Ceperley
, “
Liquid–liquid phase transition in hydrogen by coupled electron–ion Monte Carlo simulations
,”
Proc. Natl. Acad. Sci. U. S. A.
113
,
4953
4957
(
2016
).
198.
K. T.
Delaney
,
C.
Pierleoni
, and
D. M.
Ceperley
, “
Quantum Monte Carlo simulation of the high-pressure molecular-atomic crossover in fluid hydrogen
,”
Phys. Rev. Lett.
97
,
235702
(
2006
).
199.
E.
Gregoryanz
,
C.
Ji
,
P.
Dalladay-Simpson
,
B.
Li
,
R. T.
Howie
, and
H.-K.
Mao
, “
Everything you always wanted to know about metallic hydrogen but were afraid to ask
,”
Matter Radiat. Extremes
5
,
038101
(
2020
).
200.
C.
Ji
,
B.
Li
,
W.
Liu
,
J. S.
Smith
,
A.
Majumdar
,
W.
Luo
,
R.
Ahuja
,
J.
Shu
,
J.
Wang
,
S.
Sinogeikin
et al, “
Ultrahigh-pressure isostructural electronic transitions in hydrogen
,”
Nature
573
,
558
(
2019
).
201.
C.
Ji
,
B.
Li
,
W.
Liu
,
J. S.
Smith
,
A.
Björling
,
A.
Majumdar
,
W.
Luo
,
R.
Ahuja
,
J.
Shu
,
J.
Wang
et al, “
Crystallography of low Z material at ultrahigh pressure: Case study on solid hydrogen
,”
Matter Radiat. Extremes
5
,
038401
(
2020
).
202.
C. J.
Pickard
and
R. J.
Needs
, “
Structure of phase III of solid hydrogen
,”
Nat. Phys.
3
,
473
476
(
2007
).
203.
R. P.
Dias
and
I. F.
Silvera
, “
Observation of the Wigner-Huntington transition to metallic hydrogen
,”
Science
355
,
715
718
(
2017
).
204.
M. I.
Eremets
,
A. P.
Drozdov
,
P. P.
Kong
, and
H.
Wang
, “
Semimetallic molecular hydrogen at pressure above 350 GPa
,”
Nat. Phys.
15
,
1246
1249
(
2019
).
205.
C.-S.
Zha
,
H.
Liu
,
J. S.
Tse
, and
R. J.
Hemley
, “
Melting and high pt transitions of hydrogen up to 300 GPa
,”
Phys. Rev. Lett.
119
,
075302
(
2017
).
206.
H.
Niu
,
Y.
Yang
,
S.
Jensen
,
M.
Holzmann
,
C.
Pierleoni
, and
D. M.
Ceperley
, “
Stable solid molecular hydrogen above 900k from a machine-learned potential trained with diffusion quantum monte carlo
,”
Phys. Rev. Lett.
130
,
76102
(
2023
).
207.
M.
Zaghoo
,
A.
Salamat
, and
I. F.
Silvera
, “
Evidence of a first-order phase transition to metallic hydrogen
,”
Phys. Rev. B
93
,
155128
(
2016
).
208.
C.
Pierleoni
,
M.
Holzmann
, and
D. M.
Ceperley
, “
Local structure in dense hydrogen at the liquid-liquid phase transition by coupled electron-ion Monte Carlo
,”
Contrib. Plasma Phys.
58
,
99
106
(
2018
).
209.
V.
Dzyabura
,
M.
Zaghoo
, and
I. F.
Silvera
, “
Evidence of a liquid–liquid phase transition in hot dense hydrogen
,”
Proc. Natl. Acad. Sci. U. S. A.
110
,
8040
8044
(
2013
).
210.
K.
Ohta
,
K.
Ichimaru
,
M.
Einaga
,
S.
Kawaguchi
,
K.
Shimizu
,
T.
Matsuoka
,
N.
Hirao
, and
Y.
Ohishi
, “
Phase boundary of hot dense fluid hydrogen
,”
Sci. Rep.
5
,
16560
(
2015
).
211.
S.
Jiang
,
N.
Holtgrewe
,
Z. M.
Geballe
,
S. S.
Lobanov
,
M. F.
Mahmood
,
R. S.
McWilliams
, and
A. F.
Goncharov
, “
A spectroscopic study of the insulator–metal transition in liquid hydrogen and deuterium
,”
Adv. Sci.
7
,
1901668
(
2020
).
212.
P. M.
Celliers
,
M.
Millot
,
S.
Brygoo
,
R. S.
McWilliams
,
D. E.
Fratanduono
,
J. R.
Rygg
,
A. F.
Goncharov
,
P.
Loubeyre
,
J. H.
Eggert
,
J. L.
Peterson
et al, “
Insulator-metal transition in dense fluid deuterium
,”
Science
361
,
677
682
(
2018
).
213.
M. D.
Knudson
,
M. P.
Desjarlais
,
A.
Becker
,
R. W.
Lemke
,
K. R.
Cochrane
,
M. E.
Savage
,
D. E.
Bliss
,
T. R.
Mattsson
, and
R.
Redmer
, “
Direct observation of an abrupt insulator-to-metal transition in dense liquid deuterium
,”
Science
348
,
1455
1460
(
2015
).
214.
G.
Rillo
,
M. A.
Morales
,
D. M.
Ceperley
, and
C.
Pierleoni
, “
Optical properties of high-pressure fluid hydrogen across molecular dissociation
,”
Proc. Natl. Acad. Sci. U. S. A.
116
,
9770
9774
(
2019
).
215.
J.
Hinz
,
V. V.
Karasiev
,
S. X.
Hu
,
M.
Zaghoo
,
D.
Mejiá-Rodríguez
,
S. B.
Trickey
, and
L.
Calderín
, “
Fully consistent density functional theory determination of the insulator-metal transition boundary in warm dense hydrogen
,”
Phys. Rev. Res.
2
,
032065(R)
(
2020
).
216.
C. J.
Pickard
,
M.
Martinez-Canales
, and
R. J.
Needs
, “
Density functional theory study of phase IV of solid hydrogen
,”
Phys. Rev. B
85
,
214114
(
2012
).
217.
J.
McMinis
,
R. C.
Clay
,
D.
Lee
, and
M. A.
Morales
, “
Molecular to atomic phase transition in hydrogen under high pressure
,”
Phys. Rev. Lett.
114
,
105305
(
2015
).
218.
V.
Gorelov
,
M.
Holzmann
,
D. M.
Ceperley
, and
C.
Pierleoni
, “
Energy gap closure of crystalline molecular hydrogen with pressure
,”
Phys. Rev. Lett.
124
,
116401
(
2020
).
219.
G.
Rillo
,
M. A.
Morales
,
D. M.
Ceperley
, and
C.
Pierleoni
, “
Coupled electron-ion Monte Carlo simulation of hydrogen molecular crystals
,”
J. Chem Phys.
148
,
102314
(
2018
).
220.
S.
Azadi
and
T. D.
Kühne
, “
Unconventional phase iii of high-pressure solid hydrogen
,”
Phys. Rev. B
100
,
155103
(
2019
).
221.
L.
Monacelli
,
I.
Errea
,
M.
Calandra
, and
F.
Mauri
, “
Black metal hydrogen above 360 GPa driven by proton quantum fluctuations
,”
Nat. Phys.
17
,
63
(
2020
).
222.
L.
Monacelli
,
M.
Casula
,
K.
Nakano
,
S.
Sorella
, and
F.
Mauri
, “
Quantum phase diagram of high-pressure hydrogen
,”
Nat. Phys.
19
,
845
850
(
2023
).
223.
P.
Loubeyre
,
F.
Occelli
, and
P.
Dumas
, “
Compression of D2 to 460 GPa and isotopic effects in the path to metal hydrogen
,”
Phys. Rev. Lett.
129
,
35501
(
2022
).
224.
R. T.
Howie
,
T.
Scheler
,
C. L.
Guillaume
, and
E.
Gregoryanz
, “
Proton tunneling in phase IV of hydrogen and deuterium
,”
Phys. Rev. B
86
,
214104
(
2012
).
225.
B.
Monserrat
,
N. D.
Drummond
,
P.
Dalladay-simpson
,
R. T.
Howie
,
P.
Lopez Rios
,
E.
Gregoryanz
,
C. J.
Pickard
, and
R. J.
Needs
, “
Structure and metallicity of phase V of hydrogen
,”
Phys. Rev. Lett.
120
,
255701
(
2018
).
226.
P.
Dalladay-Simpson
,
R. T.
Howie
, and
E.
Gregoryanz
, “
Evidence for a new phase of dense hydrogen above 325 gigapascals
,”
Nature
529
,
63
67
(
2016
).
227.
M.
Morales
,
R.
Clay
,
C.
Pierleoni
, and
D.
Ceperley
, “
First principles methods: A perspective from quantum Monte Carlo
,”
Entropy
16
,
287
321
(
2013
).
228.
V. V.
Karasiev
,
J.
Hinz
,
S. X.
Hu
, and
S. B.
Trickey
, “
On the liquid–liquid phase transition of dense hydrogen
,”
Nature
600
,
E12
E16
(
2021
).
229.
T.
Bryk
,
C.
Pierleoni
,
G.
Ruocco
, and
A. P.
Seitsonen
, “
Characterization of molecular-atomic transformation in fluid hydrogen under pressure via long-wavelength asymptote of charge density fluctuations
,”
J. Mol. Liquids
312
,
113274
(
2020
).
230.
V. V.
Karasiev
,
D. I.
Mihaylov
, and
S. X.
Hu
, “
Meta-gga exchange-correlation free energy density functional to increase the accuracy of warm dense matter simulations
,”
Phys. Rev. B
105
,
L081109
(
2022
).
231.
M.
Zaghoo
,
R. J.
Husband
, and
I. F.
Silvera
, “
Striking isotope effect on the metallization phase lines of liquid hydrogen and deuterium
,”
Phys. Rev. B
98
,
104102
(
2018
).
232.
E.
Liberatore
,
C.
Pierleoni
, and
D. M.
Ceperley
, “
Liquid-solid transition in fully ionized hydrogen at ultra-high pressures
,”
J. Chem. Phys.
134
,
184505
(
2011
).
233.
M.
Bonitz
,
C.
Henning
, and
D.
Block
, “
Complex plasmas: A laboratory for strong correlations
,”
Rep. Prog. Phys.
73
,
066501
(
2010
).
234.
J.
Böning
,
A.
Filinov
,
P.
Ludwig
,
H.
Baumgartner
,
M.
Bonitz
, and
Y. E.
Lozovik
, “
Melting of trapped few-particle systems
,”
Phys. Rev. Lett.
100
,
113401
(
2008
).
235.
B.
Militzer
and
R. L.
Graham
, “
Simulations of dense atomic hydrogen in the Wigner crystal phase
,”
J. Phys. Chem. Solids
67
,
2136
2143
(
2006
).
236.
S.
Hamaguchi
,
R. T.
Farouki
, and
D. H. E.
Dubin
, “
Phase diagram of Yukawa systems near the one‐component‐plasma limit revisited
,”
J. Chem. Phys.
105
,
7641
7647
(
1996
).
237.
T.
Ott
,
M.
Stanley
, and
M.
Bonitz
, “
Non-invasive determination of the parameters of strongly coupled 2D Yukawa liquids
,”
Phys. Plasmas
18
,
063701
(
2011
).
238.
T.
Ott
,
M.
Bonitz
,
L. G.
Stanton
, and
M. S.
Murillo
, “
Coupling strength in Coulomb and Yukawa one-component plasmas
,”
Phys. Plasmas
21
,
113704
(
2014
).
239.
A.
Abrikosov
, “
A possible mechanism of high temperature superconductivity
,”
J. Less Common Met.
62
,
451
455
(
1978
).
240.
A.
Abrikosov
,
Fundamentals of the Theory of Metals
(
Dover Publications Inc
.,
North-Holland
,
1988
).
241.
M.
Bonitz
,
V. S.
Filinov
,
V. E.
Fortov
,
P. R.
Levashov
, and
H.
Fehske
, “
Hole crystallization in semiconductors
,”
J. Phys. A: Math. Gen.
39
,
4717
(
2006
).
242.
V. S.
Filinov
,
M.
Bonitz
,
H.
Fehske
,
V. E.
Fortov
, and
P. R.
Levashov
, “
Proton crystallization in a dense hydrogen plasma
,”
Contrib. Plasma Phys.
52
,
224
228
(
2012
).
243.
P.
Wachter
,
B.
Bucher
, and
J.
Malar
, “
Possibility of a superfluid phase in a bose condensed excitonic state
,”
Phys. Rev. B
69
,
094502
(
2004
).
244.
D.
Fedorov
and
V. V.
Stegailov
, “
Exciton nature of plasma phase transition in warm dense fluid hydrogen: Roks simulation
,”
ChemPhysChem
24
,
e202200730
(
2023
).
245.
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
).
246.
L.
Fiedler
,
N. A.
Modine
,
S.
Schmerler
,
D. J.
Vogel
,
G. A.
Popoola
,
A. P.
Thompson
,
S.
Rajamanickam
, and
A.
Cangi
, “
Predicting electronic structures at any length scale with machine learning
,”
npj Comput. Mater.
9
,
115
(
2023
).
247.
L. M.
Barker
and
R. E.
Hollenbach
, “
Laser interferometer for measuring high velocities of any reflecting surface
,”
J. Appl. Phys.
43
,
4669
4675
(
1972
).
248.
K.
Falk
,
S.
Regan
,
J.
Vorberger
,
M.
Barrios
,
T.
Boehly
,
D.
Fratanduono
,
S.
Glenzer
,
D.
Hicks
,
S.
Hu
,
C.
Murphy
et al, “
Self-consistent measurement of the equation of state of liquid deuterium
,”
High Energy Density Phys.
8
,
76
80
(
2012
).
249.
E.
Miller
,
T. R.
Boehly
,
A.
Melchior
,
D. D.
Meyerhofer
,
P. M.
Celliers
,
J. H.
Eggert
,
D. G.
Hicks
,
C. M.
Sorce
,
J. A.
Oertel
, and
P. M.
Emmel
, “
Streaked optical pyrometer system for laser-driven shock-wave experiments on OMEGA
,”
Rev. Sci. Instrum.
78
,
034903
(
2007
).
250.
J.
Sheffield
,
D.
Froula
,
S.
Glenzer
, and
N.
Luhmann
,
Plasma Scattering of Electromagnetic Radiation: Theory and Measurement Techniques
(
Elsevier Science
,
2010
).
251.
S. H.
Glenzer
,
O. L.
Landen
,
P.
Neumayer
,
R. W.
Lee
,
K.
Widmann
,
S. W.
Pollaine
,
R. J.
Wallace
,
G.
Gregori
,
A.
Höll
,
T.
Bornath
et al, “
Observations of plasmons in warm dense matter
,”
Phys. Rev. Lett.
98
,
065002
(
2007
).
252.
K.
Falk
,
E. J.
Gamboa
,
G.
Kagan
,
D. S.
Montgomery
,
B.
Srinivasan
,
P.
Tzeferacos
, and
J. F.
Benage
, “
Equation of state measurements of warm dense carbon using laser-driven shock and release technique
,”
Phys. Rev. Lett.
112
,
155003
(
2014
).
253.
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
et al, “
Characterizing the ionization potential depression in dense carbon plasmas with high-precision spectrally resolved x-ray scattering
,”
Plasma Phys. Controlled Fusion
61
,
014015
(
2019
).
254.
T.
Döppner
,
M.
Bethkenhagen
,
D.
Kraus
,
P.
Neumayer
,
D. A.
Chapman
,
B.
Bachmann
,
R. A.
Baggott
,
M. P.
Böhme
,
L.
Divol
,
R. W.
Falcone
et al, “
Observing the onset of pressure-driven k-shell delocalization
,”
Nature
618
,
270
(
2023
).
255.
B.
Fletcher
,
U.
Zastrau
,
E.
Galtier
,
E. J.
Gamboa
,
S.
Goede
,
W.
Schumaker
,
A.
Ravasio
,
M.
Gauthier
,
M. J.
MacDonald
,
Z.
Chen
et al, “
High resolution x-ray Thomson scattering measurements from cryogenic hydrogen jets using the linac coherent light source
,”
Rev. Sci. Instrum.
87
,
11E524
(
2016
).
256.
E. I.
Moses
,
R. N.
Boyd
,
B. A.
Remington
,
C. J.
Keane
, and
R.
Al-Ayat
, “
The national ignition facility: Ushering in a new age for high energy density science
,”
Phys. Plasmas
16
,
041006
(
2009
).
257.
U.
Zastrau
,
P.
Sperling
,
M.
Harmand
,
A.
Becker
,
T.
Bornath
,
R.
Bredow
,
S.
Dziarzhytski
,
T.
Fennel
,
L. B.
Fletcher
,
E.
Förster
et al, “
Resolving ultrafast heating of dense cryogenic hydrogen
,”
Phys. Rev. Lett.
112
,
105002
(
2014
).
258.
L. B.
Fletcher
,
J.
Vorberger
,
W.
Schumaker
,
C.
Ruyer
,
S.
Goede
,
E.
Galtier
,
U.
Zastrau
,
E. P.
Alves
,
S. D.
Baalrud
,
R. A.
Baggott
et al, “
Electron-ion temperature relaxation in warm dense hydrogen observed with picosecond resolved x-ray scattering
,”
Front. Phys.
10
,
838524
(
2022
).
259.
J.
Vorberger
,
T. R.
Preston
,
N.
Medvedev
,
M. P.
Böhme
,
Z. A.
Moldabekov
,
D.
Kraus
, and
T.
Dornheim
, “
Revealing non-equilibrium and relaxation in laser heated matter
,”
Phys. Lett. A
499
,
129362
(
2024
).
260.
J.
MacDonald
,
A. M.
Saunders
,
B.
Bachmann
,
M.
Bethkenhagen
,
L.
Divol
,
M. D.
Doyle
,
L. B.
Fletcher
,
S. H.
Glenzer
,
D.
Kraus
,
O. L.
Landen
et al, “
Demonstration of a laser-driven, narrow spectral bandwidth x-ray source for collective x-ray scattering experiments
,”
Phys. Plasmas
28
,
032708
(
2021
).
261.
A. N.
Starostin
,
A. G.
Leonov
,
Y. V.
Petrushevich
, and
V. K.
Rerikh
, “
Quantum corrections to the particle distribution function and reaction rates in dense media
,”
Plasma Phys. Rep.
31
,
123
132
(
2005
).
262.
A. N.
Starostin
,
V. K.
Gryaznov
, and
Y. V.
Petrushevich
, “
Development of the theory of momentum distribution of particles with regard to quantum phenomena
,”
J. Exp. Theor. Phys.
125
,
940
947
(
2017
).
263.
M.
Schörner
,
M.
Bethkenhagen
,
T.
Döppner
,
D.
Kraus
,
L. B.
Fletcher
,
S. H.
Glenzer
, and
R.
Redmer
, “
X-ray Thomson scattering spectra from density functional theory molecular dynamics simulations based on a modified chihara formula
,”
Phys. Rev. E
107
,
065207
(
2023
).
264.
J.
Lütgert
,
M.
Bethkenhagen
,
B.
Bachmann
,
L.
Divol
,
D. O.
Gericke
,
S. H.
Glenzer
,
G. N.
Hall
,
N.
Izumi
,
S. F.
Khan
,
O. L.
Landen
et al, “
Platform for probing radiation transport properties of hydrogen at conditions found in the deep interiors of red dwarfs
,”
Phys. Plasmas
29
,
083301
(
2022
).
265.
T.
Dornheim
,
J.
Vorberger
, and
M.
Bonitz
, “
Nonlinear electronic density response in warm dense matter
,”
Phys. Rev. Lett.
125
,
085001
(
2020
).
266.
T.
Dornheim
,
J.
Vorberger
,
Z. A.
Moldabekov
, and
P.
Tolias
, “
Spin-resolved density response of the warm dense electron gas
,”
Phys. Rev. Res.
4
,
033018
(
2022
).
267.
P.
Tolias
,
T.
Dornheim
,
Z. A.
Moldabekov
, and
J.
Vorberger
, “
Unravelling the nonlinear ideal density response of many-body systems
,”
EPL.
142
,
44001
(
2023
).
268.
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
).
269.
T.
Dornheim
,
J.
Vorberger
,
B.
Militzer
, and
Z. A.
Moldabekov
, “
Momentum distribution of the uniform electron gas at finite temperature: Effects of spin polarization
,”
Phys. Rev. E
104
,
055206
(
2021
).
270.
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
).
271.
C. A.
Kukkonen
and
A. W.
Overhauser
, “
Electron-electron interaction in simple metals
,”
Phys. Rev. B
20
,
550
557
(
1979
).
272.
C. A.
Kukkonen
and
K.
Chen
, “
Quantitative electron-electron interaction using local field factors from quantum Monte Carlo calculations
,”
Phys. Rev. B
104
,
195142
(
2021
).
273.
J. J.
Kas
and
J. J.
Rehr
, “
Finite temperature green's function approach for excited state and thermodynamic properties of cool to warm dense matter
,”
Phys. Rev. Lett.
119
,
176403
(
2017
).
274.
N.
Schlünzen
,
J.-P.
Joost
,
F.
Heidrich-Meisner
, and
M.
Bonitz
, “
Nonequilibrium dynamics in the one-dimensional Fermi-Hubbard model: Comparison of the nonequilibrium Green-functions approach and the density matrix renormalization group method
,”
Phys. Rev. B
95
,
165139
(
2017
).
275.
Z. A.
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
).
276.
V.
Olevano
,
A.
Titov
,
M.
Ladisa
,
K.
Hämäläinen
,
S.
Huotari
, and
M.
Holzmann
, “
Momentum distribution and Compton profile by the ab initio gw approximation
,”
Phys. Rev. B
86
,
195123
(
2012
).
277.
H.
Feldmeier
and
J.
Schnack
, “
Molecular dynamics for fermions
,”
Rev. Mod. Phys.
72
,
655
(
2000
).
278.
G.
Zwicknagel
and
T.
Pschiwul
, “
WPMD simulations of a two-component plasma
,”
J. Phys. A: Math. Gen.
39
,
4359
(
2006
).
279.
M.
Boninsegni
,
N. V.
Prokofev
, and
B. V.
Svistunov
, “
Worm algorithm for continuous-space path integral Monte Carlo simulations
,”
Phys. Rev. Lett.
96
,
070601
(
2006
).
280.
M.
Jarrell
and
J.
Gubernatis
, “
Bayesian inference and the analytic continuation of imaginary-time quantum Monte Carlo data
,”
Phys. Rep.
269
,
133
195
(
1996
).
281.
F.
Akhmetov
,
I.
Milov
,
I. A.
Makhotkin
,
M.
Ackermann
, and
J.
Vorberger
, “
Electron-phonon coupling in transition metals beyond Wang's approximation
,”
Phys. Rev. B
108
,
214301
(
2023
).
282.
J.
Simoni
and
J.
Daligault
, “
First-principles determination of electron-ion couplings in the warm dense matter regime
,”
Phys. Rev. Lett.
122
,
205001
(
2019
).
283.
Z.
Moldabekov
,
S.
Amirov
,
P.
Ludwig
,
M.
Bonitz
, and
T.
Ramazanov
, “
Effect of the dynamical collision frequency on quantum wakefields
,”
Contrib. Plasma Phys.
59
,
e201800161
(
2019
).
284.
H.
Reinholz
,
R.
Redmer
,
G.
Röpke
, and
A.
Wierling
, “
Long-wavelength limit of the dynamical local-field factor and dynamical conductivity of a two-component plasma
,”
Phys. Rev. E
62
,
5648
5666
(
2000
).
285.
O.
Gericke
,
M. S.
Murillo
, and
M.
Schlanges
, “
Dense plasma temperature equilibration in the binary collision approximation
,”
Phys. Rev. E
65
,
036418
(
2002
).
286.
S.
Kodanova
,
M.
Issanova
,
S.
Amirov
,
T.
Ramazanov
,
A.
Tikhonov
, and
Z.
Moldabekov
, “
Relaxation of non-isothermal hot dense plasma parameters
,”
Matter Radiat. Extremes
3
,
40
49
(
2017
).
287.
J. M.
Ziman
, “
A theory of the electrical properties of liquid metals. I: The monovalent metals
,”
Philos. Mag.: J. Theor. Exp. Appl. Phys.
6
,
1013
1034
(
1961
).
288.
T. W.
Hentschel
,
A.
Kononov
,
A.
Olmstead
,
A.
Cangi
,
A. D.
Baczewski
, and
S. B.
Hansen
, “
Improving dynamic collision frequencies: Impacts on dynamic structure factors and stopping powers in warm dense matter
,”
Phys. Plasmas
30
,
062703
(
2023
).
289.
S. X.
Hu
,
B.
Militzer
,
L. A.
Collins
,
K. P.
Driver
, and
J. D.
Kress
, “
First-principles prediction of the softening of the silicon shock hugoniot curve
,”
Phys. Rev. B
94
,
094109
(
2016
).
290.
S.
Zhang
,
K. P.
Driver
,
F. M. C.
Soubiran
, and
B.
Militzer
, “
First-principles equation of state and shock compression predictions of warm dense hydrocarbons
,”
Phys. Rev. E
96
,
013204
(
2017
).
291.
S.
Chiesa
,
D. M.
Ceperley
,
R. M.
Martin
, and
M.
Holzmann
, “
Finite-size error in many-body simulations with long-range interactions
,”
Phys. Rev. Lett.
97
,
076404
(
2006
).
292.
N. D.
Drummond
,
R. J.
Needs
,
A.
Sorouri
, and
W. M. C.
Foulkes
, “
Finite-size errors in continuum quantum Monte Carlo calculations
,”
Phys. Rev. B
78
,
125106
(
2008
).
293.
M.
Holzmann
,
R. C.
Clay
,
M. A.
Morales
,
N. M.
Tubman
,
D. M.
Ceperley
, and
C.
Pierleoni
, “
Theory of finite size effects for electronic quantum Monte Carlo calculations of liquids and solids
,”
Phys. Rev. B
94
,
035126
(
2016
).
294.
T.
Dornheim
and
J.
Vorberger
, “
Overcoming finite-size effects in electronic structure simulations at extreme conditions
,”
J. Chem. Phys.
154
,
144103
(
2021
).
295.
A. A.
Kugler
, “
Bounds for some equilibrium properties of an electron gas
,”
Phys. Rev. A
1
,
1688
(
1970
).
296.
E. W.
Brown
,
B. K.
Clark
,
J. L.
DuBois
, and
D. M.
Ceperley
, “
Path-integral Monte Carlo simulation of the warm dense homogeneous electron gas
,”
Phys. Rev. Lett.
110
,
146405
(
2013
).
297.
G.
Stefanucci
and
R.
van Leeuwen
,
Nonequilibrium Many-Body Theory of Quantum Systems: A Modern Introduction
(
Cambridge University Press
,
2013
).
298.
N.
Schlünzen
,
S.
Hermanns
,
M.
Scharnke
, and
M.
Bonitz
, “
Ultrafast dynamics of strongly correlated fermions—Nonequilibrium Green functions and selfenergy approximations
,”
J. Phys.: Condens. Matter
32
,
103001
(
2020
).
299.
M.
Bonitz
,
Quantum Kinetic Theory
, Teubner-Texte zur Physik, Vol.
33
(
B.G. Teubner
,
1998
).
300.
K.
Chen
and
K.
Haule
, “
A combined variational and diagrammatic quantum Monte Carlo approach to the many-electron problem
,”
Nat. Commun.
10
,
3725
(
2019
).
301.
K.
Haule
and
K.
Chen
, “
Static self-energy and effective mass of the homogeneous electron gas from quantum Monte Carlo calculations
,”
Sci. Rep.
12
,
2294
(
2022
).
302.
V.
Filinov
,
V.
Fortov
,
M.
Bonitz
, and
D.
Kremp
, “
Pair distribution functions of dense partially ionized hydrogen
,”
Phys. Lett. A
274
,
228
235
(
2000
).
303.
A. D.
Becke
and
K. E.
Edgecombe
, “
A simple measure of electron localization in atomic and molecular systems
,”
J. Chem. Phys.
92
,
5397
5403
(
1990
).
304.
N.
Marzari
,
A. A.
Mostofi
,
J. R.
Yates
,
I.
Souza
, and
D.
Vanderbilt
, “
Maximally localized Wannier functions: Theory and applications
,”
Rev. Mod. Phys.
84
,
1419
1475
(
2012
).
305.
M. S.
Murillo
,
J.
Weisheit
,
S. B.
Hansen
, and
M.
Dharma-Wardana
, “
Partial ionization in dense plasmas: Comparisons among average-atom density functional models
,”
Phys. Rev. E
87
,
063113
(
2013
).
306.
This implies, in particular, that the atomic contribution, ΔμA, includes contributions from all bound states and that all interactions appearing in the Hamiltonian are mapped uniquely on the interaction parts of the chemical potentials.
307.
N.
Carnahan
and
K.
Starling
, “
Equation of state for nonattracting rigid spheres
,”
J. Chem. Phys.
51
,
635
(
1969
).
308.
G.
Mansoori
,
N.
Carnahan
,
K.
Starling
, and
T.
Leland
, “
Equilibrium thermodynamic properties of the mixture of hard spheres
,”
J. Chem. Phys.
54
,
1523
(
1971
).
309.
M.
Ross
,
F. H.
Ree
, and
D. A.
Young
, “
The equation of state of molecular hydrogen at very high densitya)
,”
J. Chem. Phys.
79
,
1487
1494
(
1983
).
310.
R.
Redmer
,
G.
Röpke
, and
R.
Zimmermann
, “
Effective electron-atom interactions and virial coefficients in alkali plasmas
,”
J. Phys. B: At. Mol. Phys.
20
,
4069
(
1987
).
311.
T.
Lenosky
,
J.
Kress
,
L.
Collins
,
R.
Redmer
, and
H.
Juranek
, “
Simulations of fluid hydrogen: Comparison of a dissociation model with tight-binding molecular dynamics
,”
Phys. Rev. E
60
,
1665
(
1999
).
312.
J.
Chihara
, “
Difference in x-ray scattering between metallic and non-metallic liquids due to conduction electrons
,”
J. Phys. F: Met. Phys.
17
,
295
304
(
1987
).
313.
J.
Chihara
, “
Interaction of photons with plasmas and liquid metals—Photoabsorption and scattering
,”
J. Phys.: Condens. Matter
12
,
231
(
2000
).
314.
K.
Wünsch
,
J.
Vorberger
,
G.
Gregori
, and
D. O.
Gericke
, “
X-ray scattering as a probe for warm dense mixtures and high-pressure miscibility
,”
Europhys. Lett.
94
,
25001
(
2011
).
315.
K.
Wünsch
,
J.
Vorberger
, and
D. O.
Gericke
, “
Ion structure in warm dense matter: Benchmarking solutions of hypernetted-chain equations by first-principle simulations
,”
Phys. Rev. E
79
,
010201
(
2009
).
316.
J.
Vorberger
and
D. O.
Gericke
, “
Ab initio approach to model x-ray diffraction in warm dense matter
,”
Phys. Rev. E
91
,
033112
(
2015
).
317.
K.
Wünsch
,
J.
Vorberger
,
M.
Schlanges
, and
D. O.
Gericke
, “
Ion structure for x-ray Thomson scattering in dense fusion plasmas
,”
J. Phys.: Conf. Ser.
112
,
032077
(
2008
).
318.
D. O.
Gericke
,
J.
Vorberger
,
K.
Wünsch
, and
G.
Gregori
, “
Screening of ionic cores in partially ionized plasmas within linear response
,”
Phys. Rev. E
81
,
065401
(
2010
).
319.
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
).
320.
B. A.
Mattern
,
G. T.
Seidler
,
J. J.
Kas
,
J. I.
Pacold
, and
J. J.
Rehr
, “
Real-space green's function calculations of Compton profiles
,”
Phys. Rev. B
85
,
115135
(
2012
).
321.
B. A.
Mattern
and
G. T.
Seidler
, “
Theoretical treatments of the bound-free contribution and experimental best practice in X-ray Thomson scattering from warm dense matter
,”
Phys. Plasmas
20
,
022706
(
2013
).
322.
J. C.
Stewart
and
K. D.
Pyatt
, Jr.
, “
Lowering of ionization potentials in plasmas
,”
Astrophys. J.
144
,
1203
(
1966
).
323.
R.
Kubo
, “
Statistical-mechanical theory of irreversible processes. I. General theory and simple applications to magnetic and conduction problems
,”
J. Phys. Soc. Jpn.
12
,
570
586
(
1957
).
324.
D.
Greenwood
, “
The Boltzmann equation in the theory of electrical conduction in metals
,”
Proc. Phys. Soc.
71
,
585
(
1958
).
325.
K.
Plagemann
,
P.
Sperling
,
R.
Thiele
,
M.
Desjarlais
,
C.
Fortmann
,
T.
Döppner
,
H.
Lee
,
S. H.
Glenzer
, and
R.
Redmer
, “
Dynamic structure factor in warm dense beryllium
,”
New J. Phys.
14
,
055020
(
2012
).
326.
P.
Sperling
,
E. J.
Gamboa
,
H. J.
Lee
,
H. K.
Chung
,
E.
Galtier
,
Y.
Omarbakiyeva
,
H.
Reinholz
,
G.
Röpke
,
U.
Zastrau
,
J.
Hastings
,
L. B.
Fletcher
, and
S. H.
Glenzer
, “
Free-electron x-ray laser measurements of collisional-damped plasmons in isochorically heated warm dense matter
,”
Phys. Rev. Lett.
115
,
115001
(
2015
).
327.
M.
Bethkenhagen
,
B. B. L.
Witte
,
M.
Schörner
,
G.
Röpke
,
T.
Döppner
,
D.
Kraus
,
S. H.
Glenzer
,
P. A.
Sterne
, and
R.
Redmer
, “
Carbon ionization at gigabar pressures: An ab initio perspective on astrophysical high-density plasmas
,”
Phys. Rev. Res.
2
,
023260
(
2020
).
328.
C.
Mo
,
Z.-G.
Fu
,
P.
Zhang
,
W.
Kang
,
W.
Zhang
, and
X. T.
He
, “
First-principles method for x-ray Thomson scattering including both elastic and inelastic features in warm dense matter
,”
Phys. Rev. B
102
,
195127
(
2020
).
329.
P.
Hohenberg
and
W.
Kohn
, “
Inhomogeneous electron gas
,”
Phys. Rev.
136
,
B864
B871
(
1964
).
330.
N. D.
Mermin
, “
Thermal properties of the inhomogeneous electron gas
,”
Phys. Rev.
137
,
A1441
A1443
(
1965
).
331.
H.
Eschrig
, “
T ensemble-state density functional theory via Legendre transform
,”
Phys. Rev. B
82
,
205120
(
2010
).
332.
E.
Runge
and
E. K. U.
Gross
, “
Density-functional theory for time-dependent systems
,”
Phys. Rev. Lett.
52
,
997
1000
(
1984
).
333.
M.
Born
and
R.
Oppenheimer
, “
Zur Quantentheorie der Molekeln
,”
Ann. Phys.
389
,
457
484
(
1927
).
334.
J.
Mortensen
,
L. B.
Hansen
, and
K. W.
Jacobsen
, “
Real-space grid implementation of the projector augmented wave method
,”
Phys. Rev. B
71
,
035109
(
2005
).
335.
J.
Enkovaara
,
C.
Rostgaard
,
J. J.
Mortensen
,
J.
Chen
,
M.
Dułak
,
L.
Ferrighi
,
J.
Gavnholt
,
C.
Glinsvad
,
V.
Haikola
,
H. A.
Hansen
et al, “
Electronic structure calculations with GPAW: A real-space implementation of the projector augmented-wave method
,”
J. Phys.: Condens. Matter
22
,
253202
(
2010
).
336.
A. H.
Larsen
,
J. J.
Mortensen
,
J.
Blomqvist
,
I. E.
Castelli
,
R.
Christensen
,
M.
Dułak
,
J.
Friis
,
M. N.
Groves
,
B.
Hammer
,
C.
Hargus
et al, “
The atomic simulation environment—A Python library for working with atoms
,”
J. Phys.: Condens. Matter
29
,
273002
(
2017
).
337.
S. R.
Bahn
and
K. W.
Jacobsen
, “
An object-oriented scripting interface to a legacy electronic structure code
,”
Comput. Sci. Eng.
4
,
56
66
(
2002
).
338.
P.
Giannozzi
,
S.
Baroni
,
N.
Bonini
,
M.
Calandra
,
R.
Car
,
C.
Cavazzoni
,
D.
Ceresoli
,
G. L.
Chiarotti
,
M.
Cococcioni
et al, “
QUANTUM ESPRESSO: A modular and open-source software project for quantum simulations of materials
,”
J. Phys.: Condens. Matter
21
,
395502
(
2009
).
339.
P.
Giannozzi
,
O.
Andreussi
,
T.
Brumme
,
O.
Bunau
,
M. B.
Nardelli
,
M.
Calandra
,
R.
Car
,
C.
Cavazzoni
,
D.
Ceresoli
et al, “
Advanced capabilities for materials modelling with quantum espresso
,”
J. Phys.: Condens. Matter
29
,
465901
(
2017
).
340.
X.
Gonze
,
B.
Amadon
,
G.
Antonius
,
F.
Arnardi
,
L.
Baguet
,
J.-M.
Beuken
,
J.
Bieder
,
F.
Bottin
,
J.
Bouchet
, and
E. o
Bousquet
, “
The abinit project: Impact, environment and recent developments
,”
Comput. Phys. Commun.
248
,
107042
(
2020
).
341.
A. H.
Romero
,
D. C.
Allan
,
B.
Amadon
,
G.
Antonius
,
T.
Applencourt
,
L.
Baguet
,
J.
Bieder
,
F.
Bottin
,
J.
Bouchet
, and
E. o
Bousquet
, “
Abinit: Overview, and focus on selected capabilities
,”
J. Chem. Phys.
152
,
124102
(
2020
).
342.
X.
Gonze
,
F.
Jollet
,
F.
Abreu Araujo
,
D.
Adams
,
B.
Amadon
,
T.
Applencourt
,
C.
Audouze
,
J.-M.
Beuken
,
J.
Bieder
,
A.
Bokhanchuk
et al, “
Recent developments in the ABINIT software package
,”
Comput. Phys. Commun.
205
,
106
131
(
2016
).
343.
X.
Gonze
,
B.
Amadon
,
P.-M.
Anglade
,
J.-M.
Beuken
,
F.
Bottin
,
P.
Boulanger
,
F.
Bruneval
,
D.
Caliste
,
R.
Caracas
,
M.
Côté
et al, “
ABINIT: First-principles approach to material and nanosystem properties
,”
Comput. Phys. Commun.
180
,
2582
2615
(
2009
).
344.
X.
Gonze
,
G.-M.
Rignanese
,
M.
Verstraete
,
J.-M.
Beuken
,
Y.
Pouillon
,
R.
Caracas
,
F.
Jollet
,
M.
Torrent
,
G.
Zerah
,
M.
Mikami
et al, “
A brief introduction to the ABINIT software package
,”
Z. Kristallogr. Cryst. Mater.
220
,
558
562
(
2005
).
345.
X.
Gonze
,
J.-M.
Beuken
,
R.
Caracas
,
F.
Detraux
,
M.
Fuchs
,
G.-M.
Rignanese
,
L.
Sindic
,
M.
Verstraete
,
G.
Zerah
,
F.
Jollet
et al, “
First-principles computation of material properties: The ABINIT software project
,”
Comput. Mater. Sci.
25
,
478
492
(
2002
).
346.
M.
Chen
,
G.-C.
Guo
, and
L.
He
, “
Systematically improvable optimized atomic basis sets for ab initio calculations
,”
J. Phys.: Condens. Matter
22
,
445501
(
2010
).
347.
P.
Li
,
X.
Liu
,
M.
Chen
,
P.
Lin
,
X.
Ren
,
L.
Lin
,
C.
Yang
, and
L.
He
, “
Large-scale ab initio simulations based on systematically improvable atomic basis
,”
Comput. Mater. Sci.
112
,
503
517
(
2016
).
348.
J.
Hutter
,
M.
Iannuzzi
,
F.
Schiffmann
, and
J.
VandeVondele
, “
cp2k: Atomistic simulations of condensed matter systems
,”
WIREs Comput. Mol. Sci.
4
,
15
25
(
2014
).
349.
T. D.
Kühne
,
M.
Iannuzzi
,
M.
Del Ben
,
V. V.
Rybkin
,
P.
Seewald
,
F.
Stein
,
T.
Laino
,
R. Z.
Khaliullin
,
O.
Schütt
,
F.
Schiffmann
,
D.
Golze
,
J.
Wilhelm
,
S.
Chulkov
,
M. H.
Bani-Hashemian
,
V.
Weber
,
U.
Borštnik
,
M.
Taillefumier
,
A. S.
Jakobovits
,
A.
Lazzaro
,
H.
Pabst
,
T.
Müller
,
R.
Schade
,
M.
Guidon
,
S.
Andermatt
,
N.
Holmberg
,
G. K.
Schenter
,
A.
Hehn
,
A.
Bussy
,
F.
Belleflamme
,
G.
Tabacchi
,
A.
Glöß
,
M.
Lass
,
I.
Bethune
,
C. J.
Mundy
,
C.
Plessl
,
M.
Watkins
,
J.
VandeVondele
,
M.
Krack
, and
J.
Hutter
, “
CP2K: An electronic structure and molecular dynamics software package—Quickstep: Efficient and accurate electronic structure calculations
,”
J. Chem. Phys.
152
,
194103
(
2020
).
350.
B.
Zhang
,
X.
Jing
,
Q.
Xu
,
S.
Kumar
,
A.
Sharma
,
L.
Erlandson
,
S. J.
Sahoo
,
E.
Chow
,
A. J.
Medford
,
J. E.
Pask
, and
P.
Suryanarayana
, “
Sparc v2.0.0: Spin-orbit coupling, dispersion interactions, and advanced exchange–correlation functionals
,”
Software Impacts
20
,
100649
(
2024
).
351.
S.
Ghosh
and
P.
Suryanarayana
, “
Sparc: Accurate and efficient finite-difference formulation and parallel implementation of density functional theory: Extended systems
,”
Comput. Phys. Commun.
216
,
109
125
(
2017
).
352.
G.
Kresse
and
J.
Hafner
, “
Ab initio molecular dynamics for liquid metals
,”
Phys. Rev. B
47
,
558
561
(
1993
).
353.
G.
Kresse
and
J.
Hafner
, “
Ab initio molecular-dynamics simulation of the liquid-metal–amorphous-semiconductor transition in germanium
,”
Phys. Rev. B
49
,
14251
14269
(
1994
).
354.
G.
Kresse
and
J.
Furthmüller
, “
Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set
,”
Phys. Rev. B
54
,
11169
(
1996
).
355.
G.
Kresse
and
J.
Furthmüller
, “
Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set
,”
Comput. Mater. Sci.
6
,
15
50
(
1996
).
356.
S. J.
Clark
,
M. D.
Segall
,
C. J.
Pickard
,
P. J.
Hasnip
,
M. I. J.
Probert
,
K.
Refson
, and
M. C.
Payne
, “
First principles methods using castep
,”
Z. Kristallogr. Cryst. Mater.
220
,
567
570
(
2005
).
357.
M. D.
Segall
,
P. J. D.
Lindan
,
M. J.
Probert
,
C. J.
Pickard
,
P. J.
Hasnip
,
S. J.
Clark
, and
M. C.
Payne
, “
First-principles simulation: Ideas, illustrations and the castep code
,”
J. Phys.: Condens. Matter
14
,
2717
(
2002
).
358.
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
).
359.
V.
Karasiev
,
D.
Chakraborty
,
O. A.
Shukruto
, and
S. B.
Trickey
, “
Nonempirical generalized gradient approximation free-energy functional for orbital-free simulations
,”
Phys. Rev. B
88
,
161108
(
2013
).
360.
T.
Sjostrom
and
J.
Daligault
, “
Nonlocal orbital-free noninteracting free-energy functional for warm dense matter
,”
Phys. Rev. B
88
,
195103
(
2013
).
361.
K.
Luo
,
V. V.
Karasiev
, and
S. B.
Trickey
, “
Towards accurate orbital-free simulations: A generalized gradient approximation for the noninteracting free energy density functional
,”
Phys. Rev. B
101
,
075116
(
2020
).
362.
Z. A.
Moldabekov
,
X.
Shao
,
M.
Pavanello
,
J.
Vorberger
,
F.
Graziani
, and
T.
Dornheim
, “
Imposing correct jellium response is key to predict the density response by orbital-free dft
,”
Phys. Rev. B
108
,
235168
(
2023b
).
363.
D. I.
Mihaylov
,
S.
Hu
, and
V. V.
Karasiev
, “
Dragon: A multi-GPU orbital-free density functional theory molecular dynamics simulation package for modeling of warm dense matter
,”
Comput. Phys. Commun.
294
,
108931
(
2024
).
364.
W.
Mi
,
K.
Luo
,
S. B.
Trickey
, and
M.
Pavanello
, “
Orbital-free density functional theory: An attractive electronic structure method for large-scale first-principles simulations
,”
Chem. Rev.
123
,
12039
12104
(
2023
).
365.
V. V.
Karasiev
,
T.
Sjostrom
, and
S.
Trickey
, “
Finite-temperature orbital-free dft molecular dynamics: Coupling profess and quantum espresso
,”
Comput. Phys. Commun.
185
,
3240
3249
(
2014
).
366.
L.
Fiedler
,
Z. A.
Moldabekov
,
X.
Shao
,
K.
Jiang
,
T.
Dornheim
,
M.
Pavanello
, and
A.
Cangi
, “
Accelerating equilibration in first-principles molecular dynamics with orbital-free density functional theory
,”
Phys. Rev. Res.
4
,
043033
(
2022
).
367.
D. I.
Mihaylov
,
V. V.
Karasiev
,
S. X.
Hu
,
J. R.
Rygg
,
V. N.
Goncharov
, and
G. W.
Collins
, “
Improved first-principles equation-of-state table of deuterium for high-energy-density applications
,”
Phys. Rev. B
104
,
144104
(
2021
).
368.
X.
Shao
,
K.
Jiang
,
W.
Mi
,
A.
Genova
, and
M.
Pavanello
, “
DFTpy: An efficient and object-oriented platform for orbital-free DFT simulations
,”
Wiley Interdiscip. Rev. Comput. Mol. Sci.
11
,
e1482
(
2021
).
369.
A.
Blanchet
,
M.
Torrent
, and
J.
Clérouin
, “
Requirements for very high temperature Kohn–Sham DFT simulations and how to bypass them
,”
Phys. Plasmas
27
,
122706
(
2020
).
370.
M.
Hacene
,
A.
Anciaux-Sedrakian
,
X.
Rozanska
,
D.
Klahr
,
T.
Guignon
, and
P.
Fleurat-Lessard
, “
Accelerating VASP electronic structure calculations using graphic processing units
,”
J. Comput. Chem.
33
,
2581
2589
(
2012
).
371.
M.
Hutchinson
and
M.
Widom
, “
VASP on a GPU: Application to exact-exchange calculations of the stability of elemental boron
,”
Comput. Phys. Commun.
183
,
1422
1426
(
2012
).
372.
S.
Zhang
,
H.
Wang
,
W.
Kang
,
P.
Zhang
, and
X. T.
He
, “
Extended application of Kohn-Sham first-principles molecular dynamics method with plane wave approximation at high energy—From cold materials to hot dense plasmas
,”
Phys. Plasmas
23
,
042707
(
2016
).
373.
A.
Blanchet
,
J.
Clérouin
,
M.
Torrent
, and
F.
Soubiran
, “
Extended first-principles molecular dynamics model for high temperature simulations in the Abinit code: Application to warm dense aluminum
,”
Comput. Phys. Commun.
271
,
108215
(
2022
).
374.
P.
Hollebon
and
T.
Sjostrom
, “
Hybrid KohnSham+ThomasFermi scheme for high-temperature density functional theory
,”
Phys. Rev. B
105
,
235114
(
2022
).
375.
Y.
Cytter
,
E.
Rabani
,
D.
Neuhauser
, and
R.
Baer
, “
Stochastic density functional theory at finite temperatures
,”
Phys. Rev. B
97
,
115207
(
2018
).
376.
R.
Baer
,
D.
Neuhauser
, and
E.
Rabani
, “
Self-averaging stochastic Kohn-Sham density-functional theory
,”
Phys. Rev. Lett.
111
,
106402
(
2013
).
377.
M. D.
Fabian
,
B.
Shpiro
,
E.
Rabani
,
D.
Neuhauser
, and
R.
Baer
, “
Stochastic density functional theory
,”
WIREs Comput. Mol. Sci.
9
,
e1412
(
2019
).
378.
J.
White
and
L. A.
Collins
, “
Fast and universal Kohn-Sham density functional theory algorithm for warm dense matter to hot dense plasma
,”
Phys. Rev. Lett.
125
,
055002
(
2020
).
379.
T.
Chen
,
Q.
Liu
,
Y.
Liu
,
L.
Sun
, and
M.
Chen
, “
Combining stochastic density functional theory with deep potential molecular dynamics to study warm dense matter
,”
Matter Radiat. Extremes
9
,
015604
(
2024
).
380.
D. M.
Ceperley
and
B. J.
Alder
, “
Ground state of the electron gas by a stochastic method
,”
Phys. Rev. Lett.
45
,
566
569
(
1980
).
381.
J. P.
Perdew
and
A.
Zunger
, “
Self-interaction correction to density-functional approximations for many-electron systems
,”
Phys. Rev. B
23
,
5048
5079
(
1981
).
382.
S. H.
Vosko
,
L.
Wilk
, and
M.
Nusair
, “
Accurate spin-dependent electron liquid correlation energies for local spin density calculations: A critical analysis
,”
Can. J. Phys.
58
,
1200
1211
(
1980
).
383.
J. P.
Perdew
,
K.
Burke
, and
M.
Ernzerhof
, “
Generalized gradient approximation made simple
,”
Phys. Rev. Lett.
77
,
3865
3868
(
1996
).
384.
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
).
385.
M.
Levy
, “
Asymptotic coordinate scaling bound for exchange-correlation energy in density-functional theory
,”
Int. J. Quantum Chem.
36
,
617
619
(
2009
).
386.
J.
Sun
,
A.
Ruzsinszky
, and
J. P.
Perdew
, “
Strongly constrained and appropriately normed semilocal density functional
,”
Phys. Rev. Lett.
115
,
036402
(
2015
).
387.
J.
Heyd
,
G. E.
Scuseria
, and
M.
Ernzerhof
, “
Hybrid functionals based on a screened Coulomb potential
,”
J. Chem. Phys.
118
,
8207
8215
(
2003
).
388.
J. P.
Perdew
and
K.
Schmidt
, “
Jacob's ladder of density functional approximations for the exchange-correlation energy
,”
AIP Conf. Proc.
577
,
1
20
(
2001
).
389.
M. A.
Marques
,
M. J.
Oliveira
, and
T.
Burnus
, “
Libxc: A library of exchange and correlation functionals for density functional theory
,”
Comput. Phys. Commun.
183
,
2272
2281
(
2012
).
390.
S.
Lehtola
,
C.
Steigemann
,
M. J.
Oliveira
, and
M. A.
Marques
, “
Recent developments in libxc—A comprehensive library of functionals for density functional theory
,”
SoftwareX
7
,
1
5
(
2018
).
391.
T.
Sjostrom
and
J.
Daligault
, “
Gradient corrections to the exchange-correlation free energy
,”
Phys. Rev. B
90
,
155109
(
2014
).
392.
J.
Kozlowski
,
D.
Perchak
, and
K.
Burke
, “
Generalized gradient approximation made thermal
,” arXiv:2308.03319 (
2023
).
393.
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
).
394.
J. P.
Perdew
,
M.
Ernzerhof
, and
K.
Burke
, “
Rationale for mixing exact exchange with density functional approximations
,”
J. Chem. Phys.
105
,
9982
9985
(
1996
).
395.
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
).
396.
R. M. N.
Goshadze
,
V. V.
Karasiev
,
D. I.
Mihaylov
, and
S. X.
Hu
, “
Shock-induced metallization of polystyrene along the principal hugoniot investigated by advanced thermal density functionals
,”
Phys. Rev. B
107
,
155116
(
2023
).
397.
L. H.
Thomas
, “
The calculation of atomic fields
,”
Math. Proc. Camb. Philos. Soc.
23
,
542
548
(
1927
).
398.
E.
Fermi
, “
Eine statistische Methode zur Bestimmung einiger Eigenschaften des Atoms und ihre Anwendung auf die Theorie des periodischen Systems der Elemente
,”
Z. Phys.
48
,
73
79
(
1928
).
399.
D. A.
Liberman
, “
Self-consistent field model for condensed matter
,”
Phys. Rev. B
20
,
4981
4989
(
1979
).
400.
W. C.
Dharma-wardana
and
F. m c
Perrot
, “
Density-functional theory of hydrogen plasmas
,”
Phys. Rev. A
26
,
2096
2104
(
1982
).
401.
B.
Wilson
,
V.
Sonnad
,
P.
Sterne
, and
W.
Isaacs
, “
Purgatorio: A new implementation of the inferno algorithm
,”
J. Quant. Spectrosc. Radiative Transfer
99
,
658
679
(
2006
).
402.
T. J.
Callow
,
S. B.
Hansen
,
E.
Kraisler
, and
A.
Cangi
, “
First-principles derivation and properties of density-functional average-atom models
,”
Phys. Rev. Res.
4
,
023055
(
2022
).
403.
M. W. C.
Dharma-wardana
, “
Electron-ion and ion-ion potentials for modeling warm dense matter: Applications to laser-heated or shock-compressed Al and Si
,”
Phys. Rev. E
86
,
036407
(
2012
).
404.
C.
Starrett
and
D.
Saumon
, “
Fully variational average atom model with ion-ion correlations
,”
Phys. Rev. E
85
,
026403
(
2012
).
405.
C.
Starrett
and
D.
Saumon
, “
A simple method for determining the ionic structure of warm dense matter
,”
High Energy Density Phys.
10
,
35
(
2014
).
406.
G.
Faussurier
,
C.
Blancard
,
P.
Cossé
, and
P.
Renaudin
, “
Equation of state, transport coefficients, and stopping power of dense plasmas from the average-atom model self-consistent approach for astrophysical and laboratory plasmas
,”
Phys. Plasmas
17
,
052707
(
2010
).
407.
P.
Sterne
,
S.
Hansen
,
B.
Wilson
, and
W.
Isaacs
, “
Equation of state, occupation probabilities and conductivities in the average atom purgatorio code
,”
High Energy Density Phys.
3
,
278
282
(
2007
).
408.
D.
Burrill
,
D.
Feinblum
,
M.
Charest
, and
C.
Starrett
, “
Comparison of electron transport calculations in warm dense matter using the ziman formula
,”
High Energy Density Phys.
19
,
1
10
(
2016
).
409.
W.
Johnson
,
C.
Guet
, and
G.
Bertsch
, “
Optical properties of plasmas based on an average-atom model
,”
J. Quant. Spectrosc. Radiative Transfer
99
,
327
340
(
2006
).
410.
W. R.
Johnson
,
J.
Nilsen
, and
K. T.
Cheng
, “
Thomson scattering in the average-atom approximation
,”
Phys. Rev. E
86
,
036410
(
2012
).
411.
A. N.
Souza
,
D. J.
Perkins
,
C. E.
Starrett
,
D.
Saumon
, and
S. B.
Hansen
, “
Predictions of x-ray scattering spectra for warm dense matter
,”
Phys. Rev. E
89
,
023108
(
2014
).
412.
G.
Faussurier
and
C.
Blancard
, “
Density effects on electronic configurations in dense plasmas
,”
Phys. Rev. E
97
,
023206
(
2018
).
413.
S. B.
Hansen
, “
Self-consistent and detailed opacities from a non-equilibrium average-atom model
,”
Phil. Trans. R Soc. A
381
,
20220212
(
2023
).
414.
D. M.
Ceperley
, “
Path integrals in the theory of condensed helium
,”
Rev. Mod. Phys.
67
,
279
(
1995
).
415.
M. F.
Herman
,
E. J.
Bruskin
, and
B. J.
Berne
, “
On path integral monte carlo simulations
,”
J. Chem. Phys.
76
,
5150
5155
(
1982
).
416.
M.
Takahashi
and
M.
Imada
, “
Monte carlo calculation of quantum systems
,”
J. Phys. Soc. Jpn.
53
,
963
974
(
1984
).
417.
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
).
418.
D. M.
Ceperley
and
B. J.
Alder
, “
Ground state of solid hydrogen at high pressures
,”
Phys. Rev. B
36
,
2092
2106
(
1987
).
419.
G.
Giuliani
and
G.
Vignale
,
Quantum Theory of the Electron Liquid
(
Cambridge University Press
,
Cambridge
,
2008
).
420.
B.
Militzer
, “
Computation of the high temperature coulomb density matrix in periodic boundary conditions
,”
Comput. Phys. Commun.
204
,
88
96
(
2016
).
421.
L. M.
Fraser
,
W. M. C.
Foulkes
,
G.
Rajagopal
,
R. J.
Needs
,
S. D.
Kenny
, and
A. J.
Williamson
, “
Finite-size effects and coulomb interactions in quantum monte carlo calculations for homogeneous systems with periodic boundary conditions
,”
Phys. Rev. B
53
,
1814
1832
(
1996
).
422.
H.
De Raedt
and
B.
De Raedt
, “
Applications of the generalized trotter formula
,”
Phys. Rev. A
28
,
3575
3580
(
1983
).
423.
H.
Kleinert
,
Path Integrals in Quantum Mechanics, Statistics, Polymer Physics, and Financial Markets
(
World Scientific
,
2009
).
424.
K.
Sakkos
,
J.
Casulleras
, and
J.
Boronat
, “
High order chin actions in path integral monte carlo
,”
J. Chem. Phys.
130
,
204109
(
2009
).
425.
R. E.
Zillich
,
J. M.
Mayrhofer
, and
S. A.
Chin
, “
Extrapolated high-order propagators for path integral monte carlo simulations
,”
J. Chem. Phys.
132
,
044103
(
2010
).
426.
S. A.
Chin
, “
High-order path-integral monte carlo methods for solving quantum dot problems
,”
Phys. Rev. E
91
,
031301
(
2015
).
427.
N.
Metropolis
,
A. W.
Rosenbluth
,
M. N.
Rosenbluth
,
A. H.
Teller
, and
E.
Teller
, “
Equation of state calculations by fast computing machines
,”
J. Chem. Phys.
21
,
1087
1092
(
1953
).
428.
M.
Bonitz
and
D.
Semkat
, “
Generalized quantum kinetic equations
,” in
Introduction to Computational Methods in Many Body Physics
, edited by
M.
Bonitz
and
D.
Semkat
(
Rinton Press
,
Princeton
,
2006
), pp.
94
102
.
429.
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
).
430.
G.
Ferré
and
J.
Boronat
, “
Dynamic structure factor of liquid 4He across the normal-superfluid transition
,”
Phys. Rev. B
93
,
104510
(
2016
).
431.
E. L.
Pollock
and
D. M.
Ceperley
, “
Path-integral computation of superfluid densities
,”
Phys. Rev. B
36
,
8343
(
1987
).
432.
B. K.
Clark
,
M.
Casula
, and
D. M.
Ceperley
, “
Hexatic and mesoscopic phases in a 2D quantum coulomb system
,”
Phys. Rev. Lett.
103
,
055701
(
2009
).
433.
T.
Dornheim
,
H.
Thomsen
,
P.
Ludwig
,
A.
Filinov
, and
M.
Bonitz
, “
Analyzing quantum correlations made simple
,”
Contrib. Plasma Phys.
56
,
371
379
(
2016
).
434.
T.
Dornheim
, “
Path-integral Monte Carlo simulations of quantum dipole systems in traps: Superfluidity, quantum statistics, and structural properties
,”
Phys. Rev. A
102
,
023307
(
2020
).
435.
D.
Ceperley
, “
Path integral monte carlo methods for fermions
,” in
Monte Carlo and Molecular Dynamics of Condensed Matter Systems
(
Compositori
,
1996
), pp.
1
23
.
436.
P.
Driver
and
B.
Militzer
, “
All-electron path integral Monte Carlo simulations of warm dense matter: Application to water and carbon plasmas
,”
Phys. Rev. Lett.
108
,
115502
(
2012
).
437.
N. S.
Blunt
,
T. W.
Rogers
,
J. S.
Spencer
, and
W. M. C.
Foulkes
, “
Density-matrix quantum Monte Carlo method
,”
Phys. Rev. B
89
,
245124
(
2014
).
438.
F. D.
Malone
,
N. S.
Blunt
,
J. J.
Shepherd
,
D. K. K.
Lee
,
J. S.
Spencer
, and
W. M. C.
Foulkes
, “
Interaction picture density matrix quantum Monte Carlo
,”
J. Chem. Phys.
143
,
044116
(
2015
).
439.
F. D.
Malone
,
N. S.
Blunt
,
E. W.
Brown
,
D. K. K.
Lee
,
J. S.
Spencer
,
W. M. C.
Foulkes
, and
J. J.
Shepherd
, “
Accurate exchange-correlation energies for the warm dense electron gas
,”
Phys. Rev. Lett.
117
,
115701
(
2016
).
440.
A.
Yilmaz
,
K.
Hunger
,
T.
Dornheim
,
S.
Groth
, and
M.
Bonitz
, “
Restricted configuration path integral Monte Carlo
,”
J. Chem. Phys.
153
,
124114
(
2020
).
441.
T.
Shen
,
Y.
Liu
,
Y.
Yu
, and
B. M.
Rubenstein
, “
Finite temperature auxiliary field quantum Monte Carlo in the canonical ensemble
,”
J. Chem. Phys.
153
,
204108
(
2020
).
442.
B.
Hirshberg
,
M.
Invernizzi
, and
M.
Parrinello
, “
Path integral molecular dynamics for fermions: Alleviating the sign problem with the Bogoliubov inequality
,”
J. Chem. Phys.
152
,
171102
(
2020
).
443.
T.
Dornheim
,
M.
Invernizzi
,
J.
Vorberger
, and
B.
Hirshberg
, “
Attenuating the fermion sign problem in path integral Monte Carlo simulations using the Bogoliubov inequality and thermodynamic integration
,”
J. Chem. Phys.
153
,
234104
(
2020
).
444.
J.
Lee
,
M. A.
Morales
, and
F. D.
Malone
, “
A phaseless auxiliary-field quantum Monte Carlo perspective on the uniform electron gas at finite temperatures: Issues, observations, and benchmark study
,”
J. Chem. Phys.
154
,
064109
(
2021
).
445.
T.
Dornheim
,
P.
Tolias
,
S.
Groth
,
Z. A.
Moldabekov
,
J.
Vorberger
, and
B.
Hirshberg
, “
Fermionic physics from ab initio path integral Monte Carlo simulations of fictitious identical particles
,”
J. Chem. Phys.
159
,
164113
(
2023
).
446.
T.
Dornheim
,
S.
Schwalbe
,
Z. A.
Moldabekov
,
J.
Vorberger
, and
P.
Tolias
, “
Ab initio path integral Monte Carlo simulations of the uniform electron gas on large length scales
,”
J. Phys. Chem. Lett.
15
,
1305
1313
(
2024
).
447.
K.
Hunger
,
T.
Schoof
,
T.
Dornheim
,
M.
Bonitz
, and
A.
Filinov
, “
Momentum distribution function and short-range correlations of the warm dense electron gas: Ab initio quantum Monte Carlo results
,”
Phys. Rev. E
103
,
053204
(
2021
).
448.
E. W.
Brown
,
J. L.
DuBois
,
M.
Holzmann
, and
D. M.
Ceperley
, “
Exchange-correlation energy for the three-dimensional homogeneous electron gas at arbitrary temperature
,”
Phys. Rev. B
88
,
081102
(
2013
).
449.
J. J.
Shepherd
,
G. H.
Booth
, and
A.
Alavi
, “
Investigation of the full configuration interaction quantum Monte Carlo method using homogeneous electron gas models
,”
J. Chem. Phys.
136
,
244101
(
2012
).
450.
M.
Ruggeri
,
P. L.
Ríos
, and
A.
Alavi
, “
Correlation energies of the high-density spin-polarized electron gas to mev accuracy
,”
Phys. Rev. B
98
,
161105
(
2018
).
451.
R. P.
Feynman
and
A. R.
Hibbs
,
Quantum Mechanics and Path Integrals
(
McGraw Hill
,
1965
).
452.
J.
Anderson
,
Quantum Monte Carlo: Origins, Development, Applications
(
Oxford University Press
,
2007
).
453.
B.
Militzer
,
D. M.
Ceperley
,
J. D.
Kress
,
J. D.
Johnson
,
L. A.
Collins
, and
S.
Mazevet
,
Phys. Rev. Lett.
87
,
275502
(
2001
).
454.
B.
Militzer
and
K. P.
Driver
, “
Development of path integral Monte Carlo simulations with localized nodal surfaces for second-row elements
,”
Phys. Rev. Lett.
115
,
176403
(
2015
).
455.
B.
Militzer
, “
First principles calculations of shock compressed fluid helium
,”
Phys. Rev. Lett.
97
,
175501
(
2006
).
456.
B.
Militzer
, “
Path integral Monte Carlo and density functional molecular dynamics simulations of hot, dense helium
,”
Phys. Rev. B
79
,
155105
(
2009
).
457.
S.
Zhang
,
B.
Militzer
,
M. C.
Gregor
,
K.
Caspersen
,
L. H.
Yang
,
J.
Gaffney
,
T.
Ogitsu
,
D.
Swift
,
A.
Lazicki
,
D.
Erskine
et al, “
Theoretical and experimental investigation of the equation of state of boron plasmas
,”
Phys. Rev. E
98
,
023205
(
2018
).
458.
K. P.
Driver
and
B.
Militzer
, “
First-principles equation of state calculations of warm dense nitrogen
,”
Phys. Rev. B
93
,
064101
(
2016
).
459.
K. P.
Driver
,
F.
Soubiran
,
S.
Zhang
, and
B.
Militzer
, “
First-principles equation of state and electronic properties of warm dense oxygen
,”
J. Chem. Phys.
143
,
164507
(
2015
).
460.
K. P.
Driver
and
B.
Militzer
, “
First-principles simulations and shock Hugoniot calculations of warm dense neon
,”
Phys. Rev. B
91
,
045103
(
2015
).
461.
S.
Zhang
,
K. P.
Driver
,
F.
Soubiran
, and
B.
Militzer
, “
Equation of state and shock compression of warm dense sodium—A first-principles study
,”
J. Chem. Phys.
146
,
074505
(
2017
).
462.
K. P.
Driver
,
F.
Soubiran
, and
B.
Militzer
, “
Path integral Monte Carlo simulations of warm dense aluminum
,”
Phys. Rev. E
97
,
063207
(
2018
).
463.
F.
González-Cataldo
,
F.
Soubiran
,
H.
Peterson
, and
B.
Militzer
, “
Path integral Monte Carlo and density functional molecular dynamics simulations of warm dense MgSiO3
,”
Phys. Rev. B
101
,
024107
(
2020
).
464.
F.
González-Cataldo
,
F.
Soubiran
, and
B.
Militzer
, “
Equation of state of hot, dense magnesium derived with first-principles computer simulations
,”
Phys. Plasmas
27
,
092706
(
2020
).
465.
B.
Militzer
,
F.
Gonzalez-Cataldo
,
S.
Zhang
,
K. P.
Driver
, and
F.
Soubiran
,
Phys. Rev. E
103
,
013203
(
2021
).
466.
B.
Militzer
,
F.
González-Cataldo
,
S.
Zhang
,
H. D.
Whitley
,
D. C.
Swift
, and
M.
Millot
, “
Nonideal mixing effects in warm dense matter studied with first-principles computer simulations
,”
J. Chem. Phys.
153
,
184101
(
2020
).
467.
E. L.
Pollock
,
Comp. Phys. Commun.
52
,
49
(
1988
).
468.
G.
Kelbg
, “
Theorie des Quanten-Plasmas
,”
Annalen der Phys.
467
,
219
224
(
1963
).
469.
G.
Kelbg
, “
Quantenstatistik der Gase mit Coulomb-Wechselwirkung
,”
Ann. Phys.
467
,
354
360
(
1963
).
470.
A.
Filinov
,
M.
Bonitz
, and
W.
Ebeling
, “
Improved Kelbg potential for correlated Coulomb systems
,”
J. Phys. A: Math. Gen.
36
,
5957
5962
(
2003
).
471.
A. V.
Filinov
,
V. O.
Golubnychiy
,
M.
Bonitz
,
W.
Ebeling
, and
J. W.
Dufty
, “
Temperature-dependent quantum pair potentials and their application to dense partially ionized hydrogen plasmas
,”
Phys. Rev. E
70
,
046411
(
2004
).
472.
J.
Glosli
,
F.
Graziani
,
R.
More
,
M.
Murillo
,
F.
Streitz
,
M.
Surh
,
L.
Benedict
,
S.
Hau-Riege
,
A.
Langdon
, and
R.
London
, “
Molecular dynamics simulations of temperature equilibration in dense hydrogen
,”
Phys. Rev. E
78
,
025401
(
2008
).
473.
F. R.
Graziani
,
V. S.
Batista
,
L. X.
Benedict
,
J. I.
Castor
,
H.
Chen
,
S. N.
Chen
,
C. A.
Fichtl
,
J. N.
Glosli
,
P. E.
Grabowski
,
A. T.
Graf
et al, “
Large-scale molecular dynamics simulations of dense plasmas: The Cimarron Project
,”
High Energy Density Phys.
8
,
105
131
(
2012
).
474.
T.
Dornheim
,
S.
Groth
,
A. V.
Filinov
, and
M.
Bonitz
, “
Path integral Monte Carlo simulation of degenerate electrons: Permutation-cycle properties
,”
J. Chem. Phys.
151
,
014108
(
2019
).
475.
E. H.
Lieb
, “
The stability of matter
,”
Rev. Mod. Phys.
48
,
553
569
(
1976
).
476.
Y.
Xiong
and
H.
Xiong
, “
On the thermodynamic properties of fictitious identical particles and the application to fermion sign problem
,”
J. Chem. Phys.
157
,
094112
(
2022
).
477.
A. P.
Lyubartsev
, “
Simulation of excited states and the sign problem in the path integral Monte Carlo method
,”
J. Phys. A: Math. Gen.
38
,
6659
(
2005
).
478.
S. A.
Chin
and
C. R.
Chen
, “
Gradient symplectic algorithms for solving the Schrödinger equation with time-dependent potentials
,”
J. Chem. Phys.
117
,
1409
1415
(
2002
).
479.
E.
Prodan
and
W.
Kohn
, “
Nearsightedness of electronic matter
,”
Proc. Natl. Acad. Sci. U. S. A.
102
,
11635
11638
(
2005
).
480.
Y.
Xiong
and
H.
Xiong
, “
Thermodynamics of fermions at any temperature based on parametrized partition function
,”
Phys. Rev. E
107
,
055308
(
2023
).
481.
V.
Filinov
,
P.
Levashov
, and
A.
Larkin
, “
Monte Carlo simulations of the electron short-range quantum ordering in coulomb systems and the ‘fermionic sign problem’
,”
J. Phys. A: Math. Theor.
55
,
035001
(
2021
).
482.
M.
Born
and
K.
Huang
,
Dynamical Theory of Crystal Lattices, International Series of Monographs on Physics
(
Clarendon Press
,
1988
).
483.
D.
Marx
and
J.
Hutter
,
Ab Initio Molecular Dynamics
(
Cambridge University Press
,
Cambridge
,
2009
).
484.
M.
Ceperley
and
M.
Dewing
, “
The penalty method for random walks with uncertain energies
,”
J. Chem. Phys.
110
,
9812
9820
(
1999
).
485.
M.
Holzmann
,
D. M.
Ceperley
,
C.
Pierleoni
, and
K.
Esler
, “
Backflow correlations for the electron gas and metallic hydrogen
,”
Phys. Rev. E
68
,
046707
(
2003
).
486.
C.
Pierleoni
,
K. T.
Delaney
,
M. A.
Morales
,
D. M.
Ceperley
, and
M.
Holzmann
, “
Trial wave functions for high-pressure metallic hydrogen
,”
Comput. Phys. Commun.
179
,
89
97
(
2008
).
487.
S.
Baroni
and
S.
Moroni
, “
Reptation quantum Monte Carlo: A method for unbiased ground-state averages and imaginary-time correlations
,”
Phys. Rev. Lett.
82
,
4745
4748
(
1999
).
488.
R. C.
Clay
,
M. P.
Desjarlais
, and
L.
Shulenburger
, “
Deuterium Hugoniot: Pitfalls of thermodynamic sampling beyond density functional theory
,”
Phys. Rev. B
100
,
075103
(
2019
).
489.
C.
Lin
,
F. H.
Zong
, and
D. M.
Ceperley
, “
Twist-averaged boundary conditions in continuum quantum Monte Carlo algorithms
,”
Phys. Rev. E
64
,
016702
(
2001
).
490.
M.
Taddei
,
M.
Ruggeri
,
S.
Moroni
, and
M.
Holzmann
, “
Iterative backflow renormalization procedure for many-body ground-state wave functions of strongly interacting normal fermi liquids
,”
Phys. Rev. B
91
,
115106
(
2015
).
491.
M.
Holzmann
and
S.
Moroni
, “
Orbital-dependent backflow wave functions for real-space quantum Monte Carlo
,”
Phys. Rev. B
99
,
085121
(
2019
).
492.
M.
Holzmann
and
S.
Moroni
, “
Itinerant-electron magnetism: The importance of many-body correlations
,”
Phys. Rev. Lett.
124
,
206404
(
2020
).
493.
M.
Wilson
,
S.
Moroni
,
M.
Holzmann
,
N.
Gao
,
F.
Wudarski
,
T.
Vegge
, and
A.
Bhowmik
, “
Neural network ansatz for periodic wave functions and the homogeneous electron gas
,”
Phys. Rev. B
107
,
235139
(
2023
).
494.
H.
Xie
,
Z.-H.
Li
,
H.
Wang
,
L.
Zhang
, and
L.
Wang
, “
Deep variational free energy approach to dense hydrogen
,”
Phys. Rev. Lett.
131
,
126501
(
2023
).
495.
G.
Cassella
,
H.
Sutterud
,
S.
Azadi
,
N. D.
Drummond
,
D.
Pfau
,
J. S.
Spencer
, and
W. M. C.
Foulkes
, “
Discovering quantum phase transitions with fermionic neural networks
,”
Phys. Rev. Lett.
130
,
036401
(
2023
).
496.
G.
Pescia
,
J.
Nys
,
J.
Kim
,
A.
Lovato
, and
G.
Carleo
, “
Message-passing neural quantum states for the homogeneous electron gas
,” arXiv:2305.07240 (
2023
).
497.
J. J.
Shepherd
and
A.
Grüneis
, “
Many-body quantum chemistry for the electron gas: Convergent perturbative theories
,”
Phys. Rev. Lett.
110
,
226401
(
2013
).
498.
C.
Pierleoni
,
G.
Rillo
,
D. M.
Ceperley
, and
M.
Holzmann
, “
Electron localization properties in high pressure hydrogen at the liquid-liquid phase transition by coupled electron-ion Monte Carlo
,”
J. Phys.: Conf. Ser.
1136
,
012005
(
2018
).
499.
V.
Gorelov
,
D. M.
Ceperley
,
M.
Holzmann
, and
C.
Pierleoni
, “
Electronic energy gap closure and metal-insulator transition in dense liquid hydrogen
,”
Phys. Rev. B
102
,
195133
(
2020
).
500.
Y.
Yang
,
V.
Gorelov
,
C.
Pierleoni
,
D. M.
Ceperley
, and
M.
Holzmann
, “
Electronic band gaps from Quantum Monte Carlo methods
,”
Phys. Rev. B
101
,
85115
(
2020
).
501.
V.
Gorelov
,
M.
Holzmann
,
D. M.
Ceperley
, and
C.
Pierleoni
, “
Electronic excitation spectra of molecular hydrogen in phase I from quantum Monte Carlo and many-body perturbation methods
,”
Phys. Rev. B
109
,
L241111
(
2024
).
502.
V.
Gorelov
,
D. M.
Ceperley
,
M.
Holzmann
, and
C.
Pierleoni
, “
Electronic structure and optical properties of quantum crystals from first principles calculations in the Born-Oppenheimer approximation
,”
J. Chem. Phys.
153
,
234117
(
2020d
).
503.
C.
Attaccalite
and
S.
Sorella
, “
Stable liquid hydrogen at high pressure by a novel ab initio molecular-dynamics calculation
,”
Phys. Rev. Lett.
100
,
114501
(
2008
).
504.
G.
Mazzola
,
S.
Yunoki
, and
S.
Sorella
, “
Unexpectedly high pressure for molecular dissociation in liquid hydrogen by electronic simulation
,”
Nat. Commun.
5
,
1
6
(
2014
).
505.
G.
Mazzola
and
S.
Sorella
, “
Distinct metallization and atomization transitions in dense liquid hydrogen
,”
Phys. Rev. Lett.
114
,
1
5
(
2015
).
506.
G.
Mazzola
and
S.
Sorella
, “
Accelerating ab initio molecular dynamics and probing the weak dispersive forces in dense liquid hydrogen
,”
Phys. Rev. Lett.
118
,
015703
(
2017
).
507.
V.
Filinov
,
P.
Thomas
,
I.
Varga
,
T.
Meier
,
M.
Bonitz
,
V.
Fortov
, and
S. W.
Koch
, “
Interacting electrons in a one-dimensional random array of scatterers: A quantum dynamics and Monte Carlo study
,”
Phys. Rev. B
65
,
165124
(
2002
).
508.
V. S.
Filinov
,
P.
Thomas
,
I.
Varga
,
T.
Meier
,
M.
Bonitz
,
V. E.
Fortov
, and
S. W.
Koch
, “
Electronic transport in a one-dimensional random array of scatterers
,”
J. Phys. A: Math. Gen.
36
,
5905
(
2003
).
509.
J.
Cao
and
G. A.
Voth
, “
The formulation of quantum statistical mechanics based on the Feynman path centroid density. I. Equilibrium properties
,”
J. Chem. Phys.
100
,
5093
5105
(
1994
).
510.
E. A.
Polyakov
,
A. P.
Lyubartsev
, and
P. N.
Vorontsov-Velyaminov
, “
Centroid molecular dynamics: Comparison with exact results for model systems
,”
J. Chem. Phys.
133
,
194103
(
2010
).
511.
J.
Heller
, “
Time-dependent approach to semiclassical dynamics
,”
J. Chem. Phys.
62
,
1544
1555
(
1975
).
512.
C.
Lasser
and
C.
Su
, “
Various variational approximations of quantum dynamics
,”
J. Math. Phys.
63
,
072107
(
2022
).
513.
E.
Grabowski
, “
A review of wave packet molecular dynamics
,” in
Frontiers and Challenges in Warm Dense Matter
, edited by
F.
Graziani
,
M. P.
Desjarlais
,
R.
Redmer
, and
S. B.
Trickey
(
Springer International Publishing
,
Cham
,
2014
), pp.
265
282
.
514.
W. A.
Angermeier
and
T. G.
White
, “
An investigation into the approximations used in wave packet molecular dynamics for the study of warm dense matter
,”
Plasma
4
,
294
308
(
2021
).
515.
P.
Svensson
,
T.
Campbell
,
F.
Graziani
,
Z.
Moldabekov
,
N.
Lyu
,
V. S.
Batista
,
S.
Richardson
,
S. M.
Vinko
, and
G.
Gregori
, “
Development of a new quantum trajectory molecular dynamics framework
,”
Philos. Trans. R. Soc. A
381
,
20220325
(
2023
).
516.
W.
Ebeling
and
B.
Militzer
, “
Quantum molecular dynamics of partially ionized plasmas
,”
Phys. Lett. A
226
,
298
304
(
1997
).
517.
D.
Klakow
,
C.
Toepffer
, and
P.-G.
Reinhard
, “
Hydrogen under extreme conditions
,”
Phys. Lett. A
192
,
55
59
(
1994
).
518.
D.
Klakow
,
C.
Toepffer
, and
P.-G.
Reinhard
, “
Semiclassical molecular dynamics for strongly coupled coulomb systems
,”
J. Chem. Phys.
101
,
10766
10774
(
1994
).
519.
J. T.
Su
and
W. A.
Goddard
III
, “
Excited electron dynamics modeling of warm dense matter
,”
Phys. Rev. Lett.
99
,
185003
(
2007
).
520.
J. T.
Su
and
W. A.
Goddard
III
, “
The dynamics of highly excited electronic systems: Applications of the electron force field
,”
J. Chem. Phys.
131
,
244501
(
2009
).
521.
A.
Jaramillo-Botero
,
J.
Su
,
A.
Qi
, and
W. A.
Goddard
III
, “
Large-scale, long-term nonadiabatic electron molecular dynamics for describing material properties and phenomena in extreme environments
,”
J. Comput. Chem.
32
,
497
512
(
2011
).
522.
Y.
Yao
,
Q.
Zeng
,
K.
Chen
,
D.
Kang
,
Y.
Hou
,
Q.
Ma
, and
J.
Dai
, “
Reduced ionic diffusion by the dynamic electron–ion collisions in warm dense hydrogen
,”
Phys. Plasmas
28
,
012704
(
2021
).
523.
W. A.
Angermeier
,
B. S.
Scheiner
,
N. R.
Shaffer
, and
T. G.
White
, “
Disentangling the effects of non-adiabatic interactions upon ion self-diffusion within warm dense hydrogen
,”
Philos. Trans. R. Soc. A
381
,
20230034
(
2023
).
524.
Q.
Ma
,
J.
Dai
,
D.
Kang
,
M.
Murillo
,
Y.
Hou
,
Z.
Zhao
, and
J.
Yuan
, “
Extremely low electron-ion temperature relaxation rates in warm dense hydrogen: Interplay between quantum electrons and coupled ions
,”
Phys. Rev. Lett.
122
,
015001
(
2019
).
525.
Y.
Liu
,
X.
Liu
,
S.
Zhang
,
H.
Liu
,
C.
Mo
,
Z.
Fu
, and
J.
Dai
, “
Molecular dynamics investigation of the stopping power of warm dense hydrogen for electrons
,”
Phys. Rev. E
103
,
063215
(
2021
).
526.
M.
Knaup
,
P.
Reinhard
,
C.
Toepffer
, and
G.
Zwicknagel
, “
Wave packet molecular dynamics simulations of warm dense hydrogen
,”
J. Phys. A: Math. Gen.
36
,
6165
(
2003
).
527.
B.
Jakob
,
P.-G.
Reinhard
,
C.
Toepffer
, and
G.
Zwicknagel
, “
Wave packet simulation of dense hydrogen
,”
Phys. Rev. E
76
,
036406
(
2007
).
528.
B.
Jakob
,
P.
Reinhard
,
C.
Toepffer
, and
G.
Zwicknagel
, “
Wave packet simulations for the insulator–metal transition in dense hydrogen
,”
J. Phys. A: Math. Theor.
42
,
214055
(
2009
).
529.
Y. S.
Lavrinenko
,
I. V.
Morozov
, and
I. A.
Valuev
, “
Wave packet molecular dynamics–density functional theory method for non-ideal plasma and warm dense matter simulations
,”
Contrib. Plasma Phys.
59
,
e201800179
(
2019
).
530.
Y. S.
Lavrinenko
,
I.
Morozov
, and
I.
Valuev
, “
High performance wave packet molecular dynamics with density functional exchange-correlation term for non-ideal plasma simulations
,”
J. Phys.: Conf. Ser.
1787
,
012043
(
2021
).
531.
M.
Knaup
,
P.-G.
Reinhard
, and
C.
Toepffer
, “
Wave packet molecular dynamics simulations of hydrogen near the transition to a metallic fluid
,”
Contrib. Plasma Phys.
39
,
57
60
(
1999
).
532.
M.
Knaup
,
P.-G.
Reinhard
, and
C.
Toepffer
, “
Wave packet molecular dynamics simulations of deuterium in the region of laser shock-wave experiments
,”
Contrib. Plasma Phys.
41
,
159
162
(
2001
).
533.
I.
Morozov
and
I.
Valuev
, “
Localization constraints in Gaussian wave packet molecular dynamics of nonideal plasmas
,”
J. Phys. A: Math. Theor.
42
,
214044
(
2009
).
534.
E.
Grabowski
,
A.
Markmann
,
I. V.
Morozov
,
I. A.
Valuev
,
C. A.
Fichtl
,
D. F.
Richards
,
V. S.
Batista
,
F. R.
Graziani
, and
M. S.
Murillo
, “
Wave packet spreading and localization in electron-nuclear scattering
,”
Phys. Rev. E
87
,
063104
(
2013
).
535.
W.
Ebeling
,
A.
Filinov
,
M.
Bonitz
,
V.
Filinov
, and
T.
Pohl
, “
The method of effective potentials in the quantum-statistical theory of plasmas
,”
J. Phys. A: Math. Gen.
39
,
4309
(
2006
).
536.
Y. S.
Lavrinenko
,
I.
Morozov
, and
I.
Valuev
, “
Reflecting boundary conditions for classical and quantum molecular dynamics simulations of nonideal plasmas
,”
Contrib. Plasma Phys.
56
,
448
458
(
2016
).
537.
J.
Hoffmann
and
G.
Kelbg
, “
Density matrix and slater sums of interacting many-particle systems
,”
Ann. Phys.
17
,
356
367
(
1966
).
538.
C.
Deutsch
and
M.
Lavaud
, “
The partition function of a two-dimensional plasma
,”
Phys. Lett. A
39
,
253
254
(
1972
).
539.
C.
Deutsch
,
M.
Gombert
, and
H.
Minoo
, “
Classical modelization of symmetry effects in the dense high-temperature electron gas
,”
Phys. Lett. A
66
,
381
382
(
1978
).
540.
H.
Minoo
,
M. M.
Gombert
, and
C.
Deutsch
, “
Temperature-dependent coulomb interactions in hydrogenic systems
,”
Phys. Rev. A
23
,
924
943
(
1981
).
541.
V.
Filinov
,
M.
Bonitz
, and
V.
Fortov
, “
High density phenomena in hydrogen plasma
,”
J. Exp. Theor. Phys. Lett.
72
,
245
(
2000
).
542.
V. S.
Filinov
,
M.
Bonitz
,
W.
Ebeling
, and
V. E.
Fortov
, “
Thermodynamics of hot dense H-plasmas: Path integral Monte Carlo simulations and analytical approximations
,”
Plasma Phys. Controlled Fusion
43
,
743
(
2001
).
543.
V. S.
Filinov
,
Y. B.
Ivanov
,
V. E.
Fortov
,
M.
Bonitz
, and
P. R.
Levashov
, “
Color path-integral Monte-Carlo simulations of quark-gluon plasma: Thermodynamic and transport properties
,”
Phys. Rev. C
87
,
035207
(
2013
).
544.
V.
Filinov
,
M.
Bonitz
,
Y.
Ivanov
,
V.
Skokov
,
P.
Levashov
, and
V.
Fortov
, “
Equation of state of strongly coupled quark–gluon plasma—Path integral Monte Carlo results
,”
Contrib. Plasma Phys.
49
,
536
543
(
2009
).
545.
V. S.
Filinov
,
Y. B.
Ivanov
,
M.
Bonitz
,
P. R.
Levashov
, and
V. E.
Fortov
, “
Quantum Monte Carlo simulations of strongly coupled quark‐gluon plasma
,”
Contrib. Plasma Phys.
52
,
135
139
(
2012
).
546.
V.
Golubnychiy
,
M.
Bonitz
,
D.
Kremp
, and
M.
Schlanges
, “
Plasmon dispersion of a weakly degenerate nonideal one-component plasma
,”
Contrib. Plasma Phys.
42
,
37
41
(
2002
).
547.
V.
Golubnychiy
,
M.
Bonitz
,
D.
Kremp
, and
M.
Schlanges
, “
Dynamical properties and plasmon dispersion of a weakly degenerate correlated one-component plasma
,”
Phys. Rev. E
64
,
016409
(
2001
).
548.
M.
Bonitz
,
D.
Semkat
,
A.
Filinov
,
V.
Golubnychyi
,
D.
Kremp
,
D. O.
Gericke
,
M. S.
Murillo
,
V.
Filinov
,
V.
Fortov
et al, “
Theory and simulation of strong correlations in quantum Coulomb systems
,”
J. Phys. A: Math. Gen.
36
,
5921
(
2003
).
549.
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
(
2022e
).
550.
S.
Izvekov
,
M.
Parrinello
,
C. J.
Burnham
, and
G. A.
Voth
, “
Effective force fields for condensed phase systems from ab initio molecular dynamics simulation: A new method for force-matching
,”
J. Chem. Phys.
120
,
10896
10913
(
2004
).
551.
G.
Tóth
, “
Effective potentials from complex simulations: A potential-matching algorithm and remarks on coarse-grained potentials
,”
J. Phys.: Condens. Matter
19
,
335222
(
2007
).
552.
O.
Akin-Ojo
,
Y.
Song
, and
F.
Wang
, “
Developing ab initio quality force fields from condensed phase quantum-mechanics/molecular-mechanics calculations through the adaptive force matching method
,”
J. Chem. Phys.
129
,
064108
(
2008
).
553.
J.
Vorberger
and
D.
Gericke
, “
Effective ion–ion potentials in warm dense matter
,”
High Energy Density Phys.
9
,
178
186
(
2013
).
554.
H.
Los
,
L. M.
Ghiringhelli
,
E. J.
Meijer
, and
A.
Fasolino
, “
Improved long-range reactive bond-order potential for carbon. I. Construction
,”
Phys. Rev. B
72
,
214102
(
2005
).
555.
J.
Behler
and
M.
Parrinello
, “
Generalized neural-network representation of high-dimensional potential-energy surfaces
,”
Phys. Rev. Lett.
98
,
146401
(
2007
).
556.
A. P.
Bartók
,
M. C.
Payne
,
R.
Kondor
, and
G.
Csányi
, “
Gaussian approximation potentials: The accuracy of quantum mechanics, without the electrons
,”
Phys. Rev. Lett.
104
,
136403
(
2010
).
557.
S.
Chmiela
,
A.
Tkatchenko
,
H. E.
Sauceda
,
I.
Poltavsky
,
K. T.
Schütt
, and
K.-R.
Müller
, “
Machine learning of accurate energy-conserving molecular force fields
,”
Sci. Adv.
3
,
e1603015
(
2017
).
558.
R.
Jinnouchi
,
J.
Lahnsteiner
,
F.
Karsai
,
G.
Kresse
, and
M.
Bokdam
, “
Phase transitions of hybrid perovskites simulated by machine-learning force fields trained on the fly with Bayesian inference
,”
Phys. Rev. Lett.
122
,
225701
(
2019
).
559.
H.
Zong
,
H.
Wiebe
, and
G.
Ackland
, “
Understanding high pressure hydrogen with a hierarchical machine-learned potential
,”
Nat. Commun.
11
,
5014
(
2020
).
560.
B.
Cheng
,
G.
Mazzola
,
C. J.
Pickard
, and
M.
Ceriotti
, “
Evidence for supercritical behaviour of high-pressure liquid hydrogen
,”
Nature
585
,
217
220
(
2020
).
561.
A.
Tirelli
,
G.
Tenti
,
K.
Nakano
, and
S.
Sorella
, “
High-pressure hydrogen by machine learning and quantum Monte Carlo
,”
Phys. Rev. B
106
,
L041105
(
2022
).
562.
D. M.
Ceperley
,
S.
Jensen
,
Y.
Yang
,
H.
Niu
,
C.
Pierleoni
, and
M.
Holzmann
, “
Training models using forces computed by stochastic electronic structure methods
,”
Electron. Struct.
6
,
015011
(
2024
).
563.
B.
Cheng
,
G.
Mazzola
,
C. J.
Pickard
, and
M.
Ceriotti
, “
Reply to: On the liquid–liquid phase transition of dense hydrogen
,”
Nature
600
,
E15
E16
(
2021
).
564.
J. A.
Ellis
,
L.
Fiedler
,
G. A.
Popoola
,
N. A.
Modine
,
J. A.
Stephens
,
A. P.
Thompson
,
A.
Cangi
, and
S.
Rajamanickam
, “
Accelerating finite-temperature Kohn-Sham density functional theory with deep neural networks
,”
Phys. Rev. B
104
,
035120
(
2021
).
565.
A.
Thompson
,
L.
Swiler
,
C.
Trott
,
S.
Foiles
, and
G.
Tucker
, “
Spectral neighbor analysis method for automated generation of quantum-accurate interatomic potentials
,”
J. Comput. Phys.
285
,
316
330
(
2015
).
566.
E.
Gull
,
A. J.
Millis
,
A. I.
Lichtenstein
,
A. N.
Rubtsov
,
M.
Troyer
, and
P.
Werner
, “
Continuous-time Monte Carlo methods for quantum impurity models
,”
Rev. Mod. Phys.
83
,
349
404
(
1960
).
567.
R.
Balescu
, “
Irreversible processes in ionized gases
,”
Phys. Fluids
3
,
52
(
1960
).
568.
A.
Lenard
, “
On Bogoliubov's kinetic equation for a spatially homogeneous plasma
,”
Phys. Plasmas
8
,
2275
(
2001
).
569.
S. I.
Braginskii
, “
Transport processes in a plasma
,”
Rev. Plasma Phys.
1
,
205
(
1965
).
570.
E.
Schroedter
and
M.
Bonitz
, “
Classical and quantum theory of fluctuations for many-particle systems out of equilibrium
,”
Contrib. Plasma Phys.
64
,
e202400015
(
2024
).
571.
L.
Bhatnagar
,
E. P.
Gross
, and
M.
Krook
, “
A model for collision processes in gases. I. Small amplitude processes in charged and neutral one-component systems
,”
Phys. Rev.
94
,
511
525
(
1954
).
572.
N.
Rostoker
and
M.
Rosenbluth
, “
Test particles in a completely ionized plasma
,”
Phys. Fluids
3
,
1
(
1960
).
573.
P.
Ludwig
,
M.
Bonitz
,
H.
Kählert
, and
J. W.
Dufty
, “
Dynamics of strongly correlated ions in a partially ionized quantum plasma
,”
J. Phys.: Conf. Ser.
220
,
012003
(
2010
).
574.
D. C.
Scott
,
R.
Binder
, and
S. W.
Koch
, “
Ultrafast dephasing through acoustic plasmon undamping in nonequilibrium electron-hole plasmas
,”
Phys. Rev. Lett.
69
,
347
350
(
1992
).
575.
R.
Binder
,
D.
Scott
,
A. E.
Paul
,
M.
Lindberg
,
K.
Henneberger
, and
S. W.
Koch
, “
Carrier-carrier scattering and optical dephasing in highly excited semiconductors
,”
Phys. Rev. B
45
,
1107
1115
(
1992
).
576.
S.
Kosse
,
M.
Bonitz
,
M.
Schlanges
, and
W.
Kraeft
, “
Evaluation of the quantum Landau collision integral
,”
Contrib. Plasma Phys.
37
,
499
(
1997
).
577.
R.
Scullard
,
A. P.
Belt
,
S. C.
Fennell
,
M. R.
Janković
,
N.
Ng
,
S.
Serna
, and
F. R.
Graziani
, “
Numerical solution of the quantum Lenard-Balescu equation for a non-degenerate one-component plasma
,”
Phys. Plasmas
23
,
092119
(
2016
).
578.
J.
Vorberger
and
D. O.
Gericke
, “
Coupled mode effects on energy transfer in weakly coupled, two-temperature plasmas
,”
Phys. Plasmas
16
,
082702
(
2009
).
579.
V.
Keldysh
, “
Diagram technique for nonequilibrium processes
,”
J. Exp. Theor. Phys.
20
,
1018
(
1965
).
580.
K.
Balzer
and
M.
Bonitz
,
Nonequilibrium Green's Functions Approach to Inhomogeneous Systems
(
Springer
,
Berlin Heidelberg
,
2013
).
581.
T.
Matsubara
, “
A new approach to quantum-statistical mechanics
,”
Prog. Theor. Phys.
14
,
351
(
1955
).
582.
A.
Abrikosov
,
L.
Gor'kov
, and
I.
Dzialoshinskii
,
Methods of Quantum Field Theory in Statistical Physics
(
Nauka
,
Moscow
,
1962
) [English transl.: Dover, New York, 1975].
583.
J. J.
Kas
,
T. D.
Blanton
, and
J. J.
Rehr
, “
Exchange-correlation contributions to thermodynamic properties of the homogeneous electron gas from a cumulant Green's function approach
,”
Phys. Rev. B
100
,
195144
(
2019
).
584.
J.-P.
Joost
,
N.
Schlünzen
,
S.
Hese
,
M.
Bonitz
,
C.
Verdozzi
,
P.
Schmitteckert
, and
M.
Hopjan
, “
Löwdin's symmetry dilemma within Green functions theory for the one-dimensional Hubbard model
,”
Contributions to Plasma Phys.
61
,
e202000220
(
2021
).
585.
A.
Marini
,
C.
Hogan
,
M.
Grüning
, and
D.
Varsano
, “
yambo: An ab initio tool for excited state calculations
,”
Comput. Phys. Commun.
180
,
1392
1403
(
2009
).
586.
M.
Bonitz
,
A.-P.
Jauho
,
M.
Sadovskii
, and
S.
Tikhodeev
, “
In memoriam Leonid V. Keldysh
,”
Phys. Status Solidi (b)
256
,
1800600
(
2019
).
587.
L.
Kadanoff
and
G.
Baym
,
Quantum Statistical Mechanics: Green's Function Methods in Equilibrium and Nonequilibrium Problems
, Frontiers in Physics. A Lecture Note and Reprint Series (
W.A. Benjamin
,
1962
).
588.
M.
Bonitz
,
D.
Kremp
,
D. C.
Scott
,
R.
Binder
,
W. D.
Kraeft
, and
H. S.
Köhler
, “
Numerical analysis of non-Markovian effects in charge-carrier scattering: One-time versus two-time kinetic equations
,”
J. Phys.: Cond. Matter
8
,
6057
(
1996
).
589.
M.
Bonitz
and
D.
Kremp
, “
Kinetic energy relaxation and correlation time of nonequilibrium many-particle systems
,”
Phys. Lett. A
212
,
83
90
(
1996
).
590.
R.
Binder
,
S.
Köhler
, and
M.
Bonitz
, “
Memory effects in the momentum orientation relaxation of electron hole plasmas in semiconductors
,”
Phys. Rev. B
55
,
5110
(
1997
).
591.
D.
Kremp
,
T.
Bornath
,
M.
Bonitz
, and
M.
Schlanges
, “
Quantum kinetic theory of plasmas in strong laser fields
,”
Phys. Rev. E
60
,
4725
4732
(
1999
).
592.
M.
Bonitz
,
T.
Bornath
,
D.
Kremp
,
M.
Schlanges
, and
W. D.
Kraeft
, “
Quantum kinetic theory for laser plasmas. Dynamical screening in strong fields
,”
Contrib. Plasma Phys.
39
,
329
347
(
1999
).
593.
H.
Haberland
,
M.
Bonitz
, and
D.
Kremp
, “
Harmonics generation in electron-ion collisions in a short laser pulse
,”
Phys. Rev. E
64
,
026405
(
2001
).
594.
N.-H.
Kwong
and
M.
Bonitz
, “
Real-time Kadanoff-Baym approach to plasma oscillations in a correlated electron gas
,”
Phys. Rev. Lett.
84
,
1768
1771
(
2000
).
595.
N.
Schlünzen
,
S.
Hermanns
,
M.
Bonitz
, and
C.
Verdozzi
, “
Dynamics of strongly correlated fermions: Ab initio results for two and three dimensions
,”
Phys. Rev. B
93
,
035107
(
2016
).
596.
P.
Lipavský
,
V.
Špička
, and
B.
Velický
, “
Generalized Kadanoff-Baym ansatz for deriving quantum transport equations
,”
Phys. Rev. B
34
,
6933
6942
(
1986
).
597.
N.
Schlünzen
,
J.-P.
Joost
, and
M.
Bonitz
, “
Achieving the scaling limit for nonequilibrium green functions simulations
,”
Phys. Rev. Lett.
124
,
076601
(
2020
).
598.
J.-P.
Joost
,
N.
Schlünzen
, and
M.
Bonitz
, “
G1-G2 scheme: Dramatic acceleration of nonequilibrium Green functions simulations within the Hartree-Fock generalized Kadanoff-Baym ansatz
,”
Phys. Rev. B
101
,
245101
(
2020
).
599.
J.-P.
Joost
,
N.
Schlünzen
,
H.
Ohldag
,
M.
Bonitz
,
F.
Lackner
, and
I.
Brezinova
, “
The dynamically screened ladder approximation: Simultaneous treatment of strong electronic correlations and dynamical screening out of equilibrium
,”
Phys. Rev. B
105
,
165155
(
2022
).
600.
M.
Bonitz
,
J.
Jan-Philip
,
C.
Makait
,
E.
Schroedter
,
T.
Kalsberger
, and
K.
Balzer
, “
Accelerating nonequilibrium Green function simulations: The G1-G2 scheme and beyond
,”
Phys. Stat. Sol. (b)
261
,
e2300578
(
2024
).
601.
D.
Semkat
,
D.
Kremp
, and
M.
Bonitz
, “
Kadanoff–Baym equations and non-Markovian Boltzmann equation in generalized T-matrix approximation
,”
J. Math. Phys.
41
,
7458
7467
(
2000
).
602.
D.
Semkat
,
D.
Kremp
, and
M.
Bonitz
, “
Kadanoff-Baym equations with initial correlations
,”
Phys. Rev. E
59
,
1557
(
1999
).
603.
M.
Bonitz
,
S.
Hermanns
, and
K.
Balzer
, “
Dynamics of Hubbard nano-clusters following strong excitation
,”
Contrib. Plasma Phys.
53
,
778
787
(
2013
).
604.
S.
Hermanns
,
N.
Schlünzen
, and
M.
Bonitz
, “
Hubbard nanoclusters far from equilibrium
,”
Phys. Rev. B
90
,
125111
(
2014
).
605.
E.
Schroedter
and
M.
Bonitz
, “
Two-time quantum fluctuations approach and its relation to the Bethe-Salpeter equation
,”
Phys. Stat. Sol. (b)
261
,
e2300564
(
2024
).
606.
C.
Makait
,
F. B.
Fajardo
, and
M.
Bonitz
, “
Time-dependent charged particle stopping in quantum plasmas: Testing the G1–G2 scheme for quasi-one-dimensional systems
,”
Contrib. Plasma Phys.
63
,
e202300008
(
2023
).
607.
E.
Schroedter
,
J.-P.
Joost
, and
M.
Bonitz
, “
Quantum fluctuations approach to the nonequilibrium GW-approximation
,”
Condense. Matter Phys.
25
,
23401
(
2022
).
608.
E.
Schroedter
,
B.
Wurst
,
J.-P.
Joost
, and
M.
Bonitz
, “
Quantum fluctuations approach to the nonequilibrium GW-approximation: Density correlations and dynamic structure factor
,”
Phys. Rev. B
108
,
205109
(
2023
).
609.
H.
Appel
and
E. K. U.
Gross
, “
Time-dependent natural orbitals and occupation numbers
,”
Europhys. Lett.
92
,
23001
(
2010
).
610.
M.
Thiele
,
E. K. U.
Gross
, and
S.
Kümmel
, “
Adiabatic approximation in nonperturbative time-dependent density-functional theory
,”
Phys. Rev. Lett.
100
,
153004
(
2008
).
611.
A.
Kononov
,
T. W.
Hentschel
,
S. B.
Hansen
, and
A. D.
Baczewski
, “
Trajectory sampling and finite-size effects in first-principles stopping power calculations
,”
npj Comput. Mater.
9
,
205
(
2023
).
612.
A.
Kononov
,
A. J.
White
,
K. A.
Nichols
,
S. X.
Hu
, and
A. D.
Baczewski
, “
Reproducibility of real-time time-dependent density functional theory calculations of electronic stopping power in warm dense matter
,” arXiv:2401.08793 (
2024
).
613.
Z. A.
Moldabekov
,
T.
Dornheim
,
M.
Bonitz
, and
T. S.
Ramazanov
, “
Ion energy-loss characteristics and friction in a free-electron gas at warm dense matter and nonideal dense plasma conditions
,”
Phys. Rev. E
101
,
053203
(
2020
).
614.
Y.-M.
Byun
,
J.
Sun
, and
C. A.
Ullrich
, “
Time-dependent density-functional theory for periodic solids: Assessment of excitonic exchange–correlation kernels
,”
Electron. Struct.
2
,
023002
(
2020
).
615.
M.
Martin
,
L.
Reining
, and
D. M.
Ceperley
,
Interacting Electrons: Theory and Computational Approaches
(
Cambridge University Press
,
2016
).
616.
E. K. U.
Gross
and
W.
Kohn
, “
Local density-functional theory of frequency-dependent linear response
,”
Phys. Rev. Lett.
55
,
2850
(
1985
).
617.
Z. A.
Moldabekov
,
M.
Pavanello
,
M. P.
Böhme
,
J.
Vorberger
, and
T.
Dornheim
, “
Linear-response time-dependent density functional theory approach to warm dense matter with adiabatic exchange-correlation kernels
,”
Phys. Rev. Res.
5
,
023089
(
2023
).
618.
Z. A.
Moldabekov
,
J.
Vorberger
,
M.
Lokamani
, and
T.
Dornheim
, “
Averaging over atom snapshots in linear-response TDDFT of disordered systems: A case study of warm dense hydrogen
,”
J. Chem. Phys.
159
,
014107
(
2023
).
619.
M. M.
Marinak
,
G. D.
Kerbel
,
N. A.
Gentile
,
O.
Jones
,
D.
Munro
,
S.
Pollaine
,
T. R.
Dittrich
, and
S. W.
Haan
, “
Three-dimensional hydra simulations of national ignition facility targets
,”
Phys. Plasmas
8
,
2275
2280
(
2001
).
620.
A. L.
Kritcher
et al, “
Design of inertial fusion implosions reaching the burning plasma regime
,”
Nat. Phys.
18
,
251
(
2022
).
621.
B.
Fryxell
,
K.
Olson
,
P.
Ricker
,
F. X.
Timmes
,
M.
Zingale
,
D. Q.
Lamb
,
P.
MacNeice
,
R.
Rosner
,
J. W.
Truran
, and
H.
Tufo
, “
Flash: An adaptive mesh hydrodynamics code for modeling astrophysical thermonuclear flashes
,”
Astrophys. J. Suppl. S
131
,
273
(
2000
).
622.
J. P.
Sauppe
,
Y.
Lu
,
P.
Tzeferacos
,
A. C.
Reyes
,
S.
Palaniyappan
,
K. A.
Flippo
,
S.
Li
, and
J. L.
Kline
, “
On the importance of three-dimensional modeling for high-energy-density physics experiments
,”
Phys. Plasmas
30
,
062707
(
2023
).
623.
R.
McLarren
,
Uncertainty Quantification and Predictive Computational Science
(
Springer
,
Heidelberg
,
2018
).
624.
A.
Banerjee
, “
Rayleigh-Taylor instability: a status review of experimental designs and measurement diagnostics
,”
J. Fluids Eng.
142
,
120801
(
2020
).
625.
J. I.
Castor
,
Radiation Hydrodynamics
, Vol.
15
(
Cambridge University Press
,
New York
,
2004
).
626.
J.
Gaffney
et al, “
A review of equation-of-state models for inertial confinement fusion materials
,”
High Energy Dens. Phys.
28
,
7
24
(
2018
).
627.
P.
Grabowski
et al, “
Review of the first charged-particle transport coeffcient comparison workshop
,”
High Energy Dens. Phys.
37
,
100905
(
2020
).
628.
L. J.
Stanek
,
A.
Kononov
,
S. B.
Hansen
,
B. M.
Haines
,
S. X.
Hu
,
P. F.
Knapp
,
M. S.
Murillo
,
L. G.
Stanton
,
H. D.
Whitley
,
S. D.
Baalrud
et al, “
Review of the second charged-particle transport coefficient code comparison workshop
,”
Phys. Plasmas
31
,
052104
(
2024
).
629.
E.
Madelung
, “
Quantentheorie in hydrodynamischer form
,”
Z. Phys.
40
,
322
326
(
1927
).
630.
D.
Bohm
, “
A suggested interpretation of the quantum theory in terms of “hidden” variables. I
,”
Phys. Rev.
85
,
166
179
(
1952
).
631.
G.
Manfredi
and
F.
Haas
, “
Self-consistent fluid model for a quantum electron gas
,”
Phys. Rev. B
64
,
075316
(
2001
).
632.
J.
Vranjes
,
B. P.
Pandey
, and
S.
Poedts
, “
On quantum plasma: A plea for a common sense
,”
Europhys. Lett.
99
,
25001
(
2012
).
633.
M.
Bonitz
,
E.
Pehlke
, and
T.
Schoof
, “
Attractive forces between ions in quantum plasmas: Failure of linearized quantum hydrodynamics
,”
Phys. Rev. E
87
,
033105
(
2013
).
634.
M.
Bonitz
,
E.
Pehlke
, and
T.
Schoof
, “
Reply to ‘Comment on ‘Attractive forces between ions in quantum plasmas: Failure of linearized quantum hydrodynamics’ ’
,”
Phys. Rev. E
87
,
037102
(
2013
).
635.
M.
Bonitz
,
Z. A.
Moldabekov
, and
T. S.
Ramazanov
, “
Quantum hydrodynamics for plasmas—Quo vadis?
,”
Phys. Plasmas
26
,
090601
(
2019
).
636.
Z. A.
Moldabekov
,
M.
Bonitz
, and
T. S.
Ramazanov
, “
Theoretical foundations of quantum hydrodynamics for plasmas
,”
Phys. Plasmas
25
,
031903
(
2018
).
637.
F.
Graziani
,
Z.
Moldabekov
,
B.
Olson
, and
M.
Bonitz
, “
Shock physics in warm dense matter—A quantum hydrodynamics perspective
,”
Contrib. Plasma Phys.
62
,
e202100170
(
2022
).
638.
E.
Blöchl
, “
Projector augmented-wave method
,”
Phys. Rev. B
50
,
17953
(
1994
).
639.
G.
Kresse
and
D.
Joubert
, “
From ultrasoft pseudopotentials to the projector augmented-wave method
,”
Phys. Rev. B
59
,
1758
(
1999
).
640.
S.
Nosé
, “
A unified formulation of the constant temperature molecular dynamics methods
,”
J. Chem. Phys.
81
,
511
519
(
1984
).
641.
W. G.
Hoover
, “
Canonical dynamics: Equilibrium phase-space distributions
,”
Phys. Rev. A
31
,
1695
1697
(
1985
).
642.
R.
Piron
and
T.
Blenski
, “
Variational-average-atom-in-quantum-plasmas (VAAQP) code and virial theorem: Equation-of-state and shock-Hugoniot calculations for warm dense Al, Fe, Cu, and Pb
,”
Phys. Rev. E
83
,
026403
(
2011
).
643.
W. R.
Johnson
, “
Pressure in the average-atom model
,” https://www3.nd.edu/∼johnson/.
644.
H.
Juranek
,
N.
Nettelmann
,
S.
Kuhlbrodt
,
V.
Schwarz
,
B.
Holst
, and
R.
Redmer
, “
Metal-nonmetal transition in dense plasmas
,”
Contrib. Plasma Phys.
45
,
432
440
(
2005
).
645.
G. E.
Norman
,
I. M.
Saitov
, and
V. V.
Stegailov
, “
Plasma-plasma and liquid-liquid first-order phase transitions
,”
Contrib. Plasma Phys.
55
,
215
221
(
2015
).
646.
R.
Rompe
and
M.
Steenbeck
,
The Plasma State of Gases (in German)
, Ergebnisse der Exakten Naturwissenschaften, Vol.
18
(
Springer
,
Berlin
,
1939
).
647.
G.
Ecker
and
W.
Kröll
, “
Lowering of the ionization energy for a plasma in thermodynamic equilibrium
,”
Phys. Fluids
6
,
62
69
(
1963
).
648.
R.
Zimmermann
,
K.
Kilimann
,
W. D.
Kraeft
,
D.
Kremp
, and
G.
Röpke
, “
Dynamical screening and self-energy of excitons in the electron–hole plasma
,”
Phys. Sta. Sol. (b)
90
,
175
187
(
1978
).
649.
H.
Haug
and
S.
Schmitt-Rink
, “
Electron theory of the optical properties of laser-excited semiconductors
,”
Prog. Quantum Electron.
9
,
3
100
(
1984
).
650.
G.
Massacrier
,
M.
Böhme
,
J.
Vorberger
,
F.
Soubiran
, and
B.
Militzer
, “
Reconciling ionization energies and band gaps of warm dense matter derived with ab initio simulations and average atom models
,”
Phys. Rev. Res.
3
,
023026
(
2021
).
651.
X.
Hu
, “
Continuum lowering and fermi-surface rising in strongly coupled and degenerate plasmas
,”
Phys. Rev. Lett.
119
,
065001
(
2017
).
652.
J.
Seidel
,
S.
Arndt
, and
W. D.
Kraeft
, “
Energy spectrum of hydrogen atoms in dense plasmas
,”
Phys. Rev. E
52
,
5387
5400
(
1995
).
653.
M.
Bonitz
and
L.
Kordts
, “First principles approach to the ionization potential depression and continuum lowering in warm dense matter,” Contrib. Plasma Phys. (to be published).
654.
C. A.
Onate
and
J. O.
Ojonubah
, “
Eigensolutions of the Schrödinger equation with a class of Yukawa potentials via supersymmetric approach
,”
J. Theor. Appl. Phys.
10
,
21
26
(
2016
).
655.
A.
Starostin
,
A.
Mironov
,
N.
Aleksandrov
,
N.
Fisch
, and
R.
Kulsrud
, “
Quantum corrections to the distribution function of particles over momentum in dense media
,”
Phys. A: Stat. Mech. Appl.
305
,
287
296
(
2002
).
656.
P.
Gori-Giorgi
and
J. P.
Perdew
, “
Short-range correlation in the uniform electron gas: Extended Overhauser model
,”
Phys. Rev. B
64
,
155102
(
2001
).
657.
E.
Daniel
and
S. H.
Vosko
, “
Momentum distribution of an interacting electron gas
,”
Phys. Rev.
120
,
2041
2044
(
1960
).
658.
V. M.
Galitskii
and
V. V.
Yakimets
, “
Particle relaxation in a maxwell gas
,”
Sov. Phys. JETP
24
,
1071
(
1967
).
659.
J. C.
Kimball
, “
Short-range correlations and the structure factor and momentum distribution of electrons
,”
J. Phys. A: Math. Gen.
8
,
1513
(
1975
).
660.
H.
Yasuhara
and
Y.
Kawazoe
, “
A note on the momentum distribution function for an electron gas
,”
Phys. A: Stat. Mech. Appl.
85
,
416
424
(
1976
).
661.
J.
Hofmann
,
M.
Barth
, and
W.
Zwerger
, “
Short-distance properties of Coulomb systems
,”
Phys. Rev. B
87
,
235125
(
2013
).
662.
M.
Holzmann
,
B.
Bernu
,
C.
Pierleoni
,
J.
Mcminis
,
D. M.
Ceperley
,
V.
Olevano
, and
L.
Delle Site
, “
Momentum distribution of the homogeneous electron gas
,”
Phys. Rev. Lett.
107
,
110402
(
2011
).
663.
B.
Militzer
and
E. L.
Pollock
, “
Lowering of the kinetic energy in interacting quantum systems
,”
Phys. Rev. Lett.
89
,
280401
(
2002
).
664.
S.
Huotari
,
J. A.
Soininen
,
T.
Pylkkänen
,
K.
Hämäläinen
,
A.
Issolah
,
A.
Titov
,
J.
McMinis
,
J.
Kim
,
K.
Esler
,
D. M.
Ceperley
et al, “
Momentum distribution and renormalization factor in sodium and the electron gas
,”
Phys. Rev. Lett.
105
,
086403
(
2010
).
665.
N.
Hiraoka
,
Y.
Yang
,
T.
Hagiya
,
A.
Niozu
,
K.
Matsuda
,
S.
Huotari
,
M.
Holzmann
, and
D. M.
Ceperley
, “
Direct observation of the momentum distribution and renormalization factor in lithium
,”
Phys. Rev. B
101
,
165124
(
2020
).
666.
Y.
Yang
,
N.
Hiraoka
,
K.
Matsuda
,
M.
Holzmann
, and
D. M.
Ceperley
, “
Quantum Monte Carlo Compton profiles of solid and liquid lithium
,”
Phys. Rev. B
101
,
165125
(
2020b
).
667.
V.
Gorelov
,
Y.
Yang
,
M.
Ruggeri
,
D. M.
Ceperley
,
C.
Pierleoni
, and
M.
Holzmann
, “
Neutral band gap of carbon by quantum Monte Carlo methods
,”
Condens. Matter Phys.
26
,
33701
(
2023
).
668.
M.
Eremets
,
A. P.
Drozdov
,
P.
Kong
, and
H.
Wang
, “
Molecular semimetallic hydrogen
,” arXiv:1708.05217 (
2017
).
669.
R. P.
Dias
,
O.
Noked
, and
I. F.
Silvera
, “
New insulating low temperature phase in dense hydrogen: The phase diagram to 421 GPa
,” arXiv:1603.02162 (
2016
).
670.
R. P.
Dias
,
O.
Noked
, and
I. F.
Silvera
, “
Quantum phase transition in solid hydrogen at high pressure
,”
Phys. Rev. B
100
,
184112
(
2019
).
671.
P.
Loubeyre
,
F.
Occelli
, and
R.
LeToullec
, “
Optical studies of solid hydrogen to 320 GPa and evidence for black hydrogen
,”
Nature
416
,
613
617
(
2002
).
672.
W. J.
Nellis
,
S. T.
Weir
, and
A. C.
Mitchell
, “
Minimum metallic conductivity of fluid hydrogen at 140 GPa (1.4 Mbar)
,”
Phys. Rev. B
59
,
3434
3449
(
1999
).
673.
M. D.
Knudson
,
M. P.
Desjarlais
,
M.
Preising
, and
R.
Redmer
, “
Evaluation of exchange-correlation functionals with multiple-shock conductivity measurements in hydrogen and deuterium at the molecular-to-atomic transition
,”
Phys. Rev. B
98
,
174110
(
2018
).
674.
R. S.
McWilliams
,
D. A.
Dalton
,
M. F.
Mahmood
, and
A. F.
Goncharov
, “
Optical properties of fluid hydrogen at the transition to a conducting state
,”
Phys. Rev. Lett.
116
,
1
6
(
2016
).
675.
M.
Holzmann
,
F.
Calcavecchia
,
D. M.
Ceperley
, and
V.
Olevano
, “
Static self-energy and effective mass of the homogeneous electron gas from quantum Monte Carlo calculations
,”
Phys. Rev. Lett.
131
,
186501
(
2023
).
676.
A. A.
Kugler
, “
Theory of the local field correction in an electron gas
,”
J. Stat. Phys.
12
,
35
(
1975
).
677.
D.
Lu
, “
Evaluation of model exchange-correlation kernels in the adiabatic connection fluctuation-dissipation theorem for inhomogeneous systems
,”
J. Chem. Phys.
140
,
18A520
(
2014
).
678.
C. E.
Patrick
and
K. S.
Thygesen
, “
Adiabatic-connection fluctuation-dissipation DFT for the structural properties of solids—The renormalized ALDA and electron gas kernels
,”
J. Chem. Phys.
143
,
102802
(
2015
).
679.
A.
Pribram-Jones
,
P. E.
Grabowski
, and
K.
Burke
, “
Thermal density functional theory: Time-dependent linear response and approximate functionals from the fluctuation-dissipation theorem
,”
Phys. Rev. Lett.
116
,
233001
(
2016
).
680.
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
).
681.
W.
Stolzmann
and
M.
Rösler
, “
Static local-field corrected dielectric and thermodynamic functions
,”
Contrib. Plasma Phys.
41
,
203
(
2001
).
682.
K. S.
Singwi
,
M. P.
Tosi
,
R. H.
Land
, and
A.
Sjölander
, “
Electron correlations at metallic densities
,”
Phys. Rev.
176
,
589
(
1968
).
683.
P.
Vashishta
and
K. S.
Singwi
, “
Electron correlations at metallic densities. V
,”
Phys. Rev. B
6
,
875
(
1972
).
684.
S.
Tanaka
, “
Correlational and thermodynamic properties of finite-temperature electron liquids in the hypernetted-chain approximation
,”
J. Chem. Phys.
145
,
214104
(
2016
).
685.
S.
Tanaka
, “
Improved equation of state for finite-temperature spin-polarized electron liquids on the basis of Singwi–Tosi–Land–Sjölander approximation
,”
Contrib. Plasma Phys.
57
,
126
136
(
2017
).
686.
P.
Tolias
,
F.
Lucco Castello
, and
T.
Dornheim
, “
Integral equation theory based dielectric scheme for strongly coupled electron liquids
,”
J. Chem. Phys.
155
,
134115
(
2021
).
687.
L.
Castello
,
P.
Tolias
, and
T.
Dornheim
, “
Classical bridge functions in classical and quantum plasma liquids
,”
Europhys. Lett.
138
,
44003
(
2022
).
688.
T.
Dornheim
,
T.
Sjostrom
,
S.
Tanaka
, and
J.
Vorberger
, “
Strongly coupled electron liquid: Ab initio path integral Monte Carlo simulations and dielectric theories
,”
Phys. Rev. B
101
,
045129
(
2020
).
689.
P.
Tolias
,
F.
Lucco Castello
, and
T.
Dornheim
, “
Quantum version of the integral equation theory-based dielectric scheme for strongly coupled electron liquids
,”
J. Chem. Phys.
158
,
141102
(
2023
).
690.
J. P. F.
LeBlanc
,
K.
Chen
,
K.
Haule
,
N. V.
Prokof'ev
, and
I. S.
Tupitsyn
, “
Dynamic response of an electron gas: Towards the exact exchange-correlation kernel
,”
Phys. Rev. Lett.
129
,
246401
(
2022
).
691.
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
).
692.
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
).
693.
Z.
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 Comput.
19
,
1286
1299
(
2023
).
694.
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
).
695.
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
).
696.
S.
Frydrych
,
J.
Vorberger
,
N.
Hartley
,
A.
Schuster
,
K.
Ramakrishna
,
A.
Saunders
,
T.
van Driel
,
R.
Falcone
,
L.
Fletcher
,
E.
Galtier
et al, “
Demonstration of X-ray Thomson scattering as diagnostics for miscibility in warm dense matter
,”
Nat. Commun.
11
,
2620
(
2020
).
697.
K.
Voigt
,
M.
Zhang
,
K.
Ramakrishna
,
A.
Amouretti
,
K.
Appel
,
E.
Brambrink
,
V.
Cerantola
,
D.
Chekrygina
,
T.
Döppner
,
R. W.
Falcone
et al, “
Demonstration of an x-ray Raman spectroscopy setup to study warm dense carbon at the high energy density instrument of European XFEL
,”
Phys. Plasmas
28
,
082701
(
2021
).
698.
D.
Ranjan
,
K.
Ramakrishna
,
K.
Voigt
,
O. S.
Humphries
,
B.
Heuser
,
M. G.
Stevenson
,
J.
Lütgert
,
Z.
He
,
C.
Qu
,
S.
Schumacher
et al, “
Toward using collective x-ray Thomson scattering to study C–H demixing and hydrogen metallization in warm dense matter conditions
,”
Phys. Plasmas
30
,
052702
(
2023
).
699.
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
).
700.
T.
Gawne
,
Z. A.
Moldabekov
,
O. S.
Humphries
,
K.
Appel
,
C.
Baehtz
,
V.
Bouffetier
,
E.
Brambrink
,
A.
Cangi
,
S.
Göde
,
Z.
Konôpková
,
M.
Makita
,
M.
Mishchenko
,
M.
Nakatsutsumi
,
K.
Ramakrishna
,
L.
Randolph
,
S.
Schwalbe
,
J.
Vorberger
,
L.
Wollenweber
,
U.
Zastrau
,
T.
Dornheim
, and
T. R.
Preston
, “
Ultrahigh resolution x-ray Thomson scattering measurements at the European x-ray free electron laser
,”
Phys. Rev. B
109
,
L241112
(
2024
).
701.
P.
Mori-Sánchez
and
A. J.
Cohen
, “
The derivative discontinuity of the exchange–correlation functional
,”
Phys. Chem. Chem. Phys.
16
,
14378
14387
(
2014
).
702.
M.
Boninsegni
and
D. M.
Ceperley
, “
Density fluctuations in liquid 4He. Path integrals and maximum entropy
,”
J. Low Temp. Phys.
104
,
339
357
(
1996
).
703.
A.
Filinov
and
M.
Bonitz
, “
Collective and single-particle excitations in two-dimensional dipolar Bose gases
,”
Phys. Rev. A
86
,
043628
(
2012
).
704.
T.
Dornheim
,
Z. A.
Moldabekov
, and
J.
Vorberger
, “
Nonlinear density response from imaginary-time correlation functions: Ab initio path integral Monte Carlo simulations of the warm dense electron gas
,”
J. Chem. Phys.
155
,
054110
(
2021
).
705.
T.
Dornheim
,
J.
Vorberger
,
Z. A.
Moldabekov
, and
M.
Böhme
, “
Analysing the dynamic structure of warm dense matter in the imaginary-time domain: Theoretical models and simulations
,”
Phil. Trans. R Soc. A
381
,
20220217
(
2023
).
706.
E.
Vitali
,
M.
Rossi
,
L.
Reatto
, and
D. E.
Galli
, “
Ab initio low-energy dynamics of superfluid and solid 4He
,”
Phys. Rev. B
82
,
174510
(
2010
).
707.
A. S.
Mishchenko
,
N. V.
Prokof'ev
,
A.
Sakamoto
, and
B. V.
Svistunov
, “
Diagrammatic quantum Monte Carlo study of the Fröhlich polaron
,”
Phys. Rev. B
62
,
6317
6336
(
2000
).
708.
Y.
Kora
and
M.
Boninsegni
, “
Dynamic structure factor of superfluid 4He from quantum Monte Carlo: Maximum entropy revisited
,”
Phys. Rev. B
98
,
134509
(
2018
).
709.
J.
Schött
,
E. G. C. P.
van Loon
,
I. L. M.
Locht
,
M. I.
Katsnelson
, and
I.
Di Marco
, “
Comparison between methods of analytical continuation for bosonic functions
,”
Phys. Rev. B
94
,
245140
(
2016
).
710.
J.
Otsuki
,
M.
Ohzeki
,
H.
Shinaoka
, and
K.
Yoshimi
, “
Sparse modeling approach to analytical continuation of imaginary-time quantum Monte Carlo data
,”
Phys. Rev. E
95
,
061302
(
2017
).
711.
O.
Goulko
,
A. S.
Mishchenko
,
L.
Pollet
,
N.
Prokof'ev
, and
B.
Svistunov
, “
Numerical analytic continuation: Answers to well-posed questions
,”
Phys. Rev. B
95
,
014102
(
2017
).
712.
T.
Döppner
,
O.
Landen
,
H.
Lee
,
P.
Neumayer
,
S.
Regan
, and
S.
Glenzer
, “
Temperature measurement through detailed balance in x-ray Thomson scattering
,”
High Energy Density Phys.
5
,
182
186
(
2009
).
713.
T.
Dornheim
,
D. C.
Wicaksono
,
J. E.
Suarez-Cardona
,
P.
Tolias
,
M. P.
Böhme
,
Z. A.
Moldabekov
,
M.
Hecht
, and
J.
Vorberger
, “
Extraction of the frequency moments of spectral densities from imaginary-time correlation function data
,”
Phys. Rev. B
107
,
155148
(
2023
).
714.
M.
Tkachenko
,
Y. V.
Arkhipov
, and
A.
Askaruly
,
The Method of Moments and Its Applications in Plasma Physics
(
Akademikerverlag
,
Saarbrücken
,
2012
).
715.
M. P.
Böhme
,
L. B.
Fletcher
,
T.
Döppner
,
D.
Kraus
,
A. D.
Baczewski
,
T. R.
Preston
,
M. J.
MacDonald
,
F. R.
Graziani
,
Z. A.
Moldabekov
et al, “
Evidence of free-bound transitions in warm dense matter and their impact on equation-of-state measurements
,” arXiv:2306.17653 (
2023
).
716.
G.
Gregori
,
S. H.
Glenzer
,
W.
Rozmus
,
R. W.
Lee
, and
O. L.
Landen
, “
Theoretical model of x-ray scattering as a dense matter probe
,”
Phys. Rev. E
67
,
026412
(
2003
).
717.
C.
Bowen
,
G.
Sugiyama
, and
B. J.
Alder
, “
Static dielectric response of the electron gas
,”
Phys. Rev. B
50
,
14838
(
1994
).
718.
T.
Dornheim
,
J.
Vorberger
,
Z. A.
Moldabekov
, and
M.
Bonitz
, “
Nonlinear interaction of external perturbations in warm dense matter
,”
Contrib. Plasma Phys.
62
,
e202100247
(
2022
).
719.
G.
Mahan
,
Many-Particle Physics
, Physics of Solids and Liquids (
Springer US
,
1990
).
720.
C.
Ullrich
,
Time-Dependent Density-Functional Theory: Concepts and Applications
(
OUP Oxford
,
2012
).
721.
T.
Dornheim
and
J.
Vorberger
, “
Finite-size effects in the reconstruction of dynamic properties from ab initio path integral Monte Carlo simulations
,”
Phys. Rev. E
102
,
063301
(
2020
).
722.
V.
Bobrov
,
S.
Trigger
, and
D.
Litinski
, “
Universality of the phonon–roton spectrum in liquids and superfluidity of 4he
,”
Z. Naturforsch. A
71
,
565
575
(
2016
).
723.
H.
Godfrin
,
M.
Meschke
,
H.-J.
Lauter
,
A.
Sultan
,
H. M.
Böhm
,
E.
Krotscheck
, and
M.
Panholzer
, “
Observation of a roton collective mode in a two-dimensional fermi liquid
,”
Nature
483
,
576
579
(
2012
).
724.
T.
Dornheim
,
Z. A.
Moldabekov
,
J.
Vorberger
, and
B.
Militzer
, “
Path integral Monte Carlo approach to the structural properties and collective excitations of liquid 3He without fixed nodes
,”
Sci. Rep.
12
,
708
(
2022
).
725.
Y.
Takada
and
H.
Yasuhara
, “
Dynamical structure factor of the homogeneous electron liquid: Its accurate shape and the interpretation of experiments on aluminum
,”
Phys. Rev. Lett.
89
,
216402
(
2002
).
726.
Y.
Takada
, “
Emergence of an excitonic collective mode in the dilute electron gas
,”
Phys. Rev. B
94
,
245106
(
2016
).
727.
J.
Koskelo
,
L.
Reining
, and
M.
Gatti
, “
Short-range excitonic phenomena in low-density metals
,” arXiv:2301.00474 (
2023
).
728.
N. D.
Mermin
, “
Lindhard dielectric function in the relaxation-time approximation
,”
Phys. Rev. B
1
,
2362
2363
(
1970
).
729.
A.
Selchow
,
G.
Röpke
,
A.
Wierling
,
H.
Reinholz
,
T.
Pschiwul
, and
G.
Zwicknagel
, “
Dynamic structure factor for a two-component model plasma
,”
Phys. Rev. E
64
,
056410
(
2001
).
730.
M.
Gajdoš
,
K.
Hummer
,
G.
Kresse
,
J.
Furthmüller
, and
F.
Bechstedt
, “
Linear optical properties in the projector-augmented wave methodology
,”
Phys. Rev. B
73
,
045112
(
2006
).
731.
K.
Ramakrishna
and
J.
Vorberger
, “
Ab initio dielectric response function of diamond and other relevant high pressure phases of carbon
,”
J. Phys.: Condens. Matter
32
,
095401
(
2019
).
732.
Changes in the collision frequency due to the ion-ion structure factor cause only minor changes in the dynamic structure factor, in general, and no visible changes for the peak position, in particular, for the considered densities and temperatures.
733.
T.
Tschentscher
,
C.
Bressler
,
J.
Grünert
,
A.
Madsen
,
A. P.
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
).
734.
C.
Bostedt
,
S.
Boutet
,
D. M.
Fritz
,
Z.
Huang
,
H. J.
Lee
,
H. T.
Lemke
,
A.
Robert
,
W. F.
Schlotter
,
J. J.
Turner
, and
G. J.
Williams
, “
Linac coherent light source: The first five years
,”
Rev. Mod. Phys.
88
,
015007
(
2016
).
735.
E. E.
McBride
,
T. G.
White
,
A.
Descamps
,
L. B.
Fletcher
,
K.
Appel
,
F. P.
Condamine
,
C. B.
Curry
,
F.
Dallari
,
S.
Funk
,
E.
Galtier
et al, “
Setup for meV-resolution inelastic X-ray scattering measurements and X-ray diffraction at the matter in extreme conditions endstation at the linac coherent light source
,”
Rev. Sci. Instrum.
89
,
10F104
(
2018
).
736.
A. P.
Thompson
,
H. M.
Aktulga
,
R.
Berger
,
D. S.
Bolintineanu
,
W. M.
Brown
,
P. S.
Crozier
,
P. J.
in 't Veld
,
A.
Kohlmeyer
,
S. G.
Moore
,
T. D.
Nguyen
et al, “
LAMMPS—A flexible simulation tool for particle-based materials modeling at the atomic, meso, and continuum scales
,”
Comput. Phys. Commun.
271
,
108171
(
2022
).
737.
M.
Bonitz
,
A.
Filinov
,
V. O.
Golubnychiy
,
T.
Bornath
, and
W. D.
Kraeft
, “
First principle thermodynamic and dynamic simulations for dense quantum plasmas
,”
Contrib. Plasma Phys.
45
,
450
458
(
2005
).
738.
T. G.
White
,
H.
Poole
,
E. E.
McBride
,
M.
Oliver
,
A.
Descamps
,
L. B.
Fletcher
,
W. A.
Angermeier
,
C. H.
Allen
,
K.
Appel
,
F. P.
Condamine
et al, “
Speed of sound in methane under conditions of planetary interiors
,”
Phys. Rev. Res.
6
,
L022029
(
2024
).
739.
B.
Holst
,
M.
French
, and
R.
Redmer
, “
Electronic transport coefficients from ab initio simulations and application to dense liquid hydrogen
,”
Phys. Rev. B
83
,
235120
(
2011
).
740.
M.
French
,
G.
Röpke
,
M.
Schörner
,
M.
Bethkenhagen
,
M. P.
Desjarlais
, and
R.
Redmer
, “
Electronic transport coefficients from density functional theory across the plasma plane
,”
Phys. Rev. E
105
,
065204
(
2022
).
741.
S.
Demyanov
,
D. V.
Knyazev
, and
P. R.
Levashov
, “
Continuous Kubo-Greenwood formula: Theory and numerical implementation
,”
Phys. Rev. E
105
,
035307
(
2022
).
742.
C. A.
Melton
,
R. C.
Clay
III
,
K. R.
Cochrane
,
A.
Dumi
,
T. A.
Gardiner
,
M. K.
Lentz
, and
J. P.
Townsend
, “
Transport coefficients of warm dense matter from Kohn-Sham density functional theory
,”
Phys. Plasmas
31
,
043903
(
2024
).
743.
S.
van de Bund
,
H.
Wiebe
, and
G. J.
Ackland
, “
Isotope quantum effects in the metallization transition in liquid hydrogen
,”
Phys. Rev. Lett.
126
,
225701
(
2021
).
744.
V.
Recoules
,
F.
Lambert
,
A.
Decoster
,
B.
Canaud
, and
J.
Clérouin
, “
Ab initio determination of thermal conductivity of dense hydrogen plasmas
,”
Phys. Rev. Lett.
102
,
075002
(
2009
).
745.
F.
Lambert
,
V.
Recoules
,
A.
Decoster
,
J.
Clérouin
, and
M.
Desjarlais
, “
On the transport coefficients of hydrogen in the inertial confinement fusion regime
,”
Phys. Plasmas
18
,
056306
(
2011
).
746.
M. P.
Desjarlais
,
C. R.
Scullard
,
L. X.
Benedict
,
H. D.
Whitley
, and
R.
Redmer
, “
Density-functional calculations of transport properties in the nondegenerate limit and the role of electron-electron scattering
,”
Phys. Rev. E
95
,
033203
(
2017
).
747.
S. X.
Hu
,
L. A.
Collins
,
V. N.
Goncharov
,
T. R.
Boehly
,
R.
Epstein
,
R. L.
McCrory
, and
S.
Skupsky
, “
First-principles opacity table of warm dense deuterium for inertial-confinement-fusion applications
,”
Phys. Rev. E
90
,
033111
(
2014
).
748.
S. X.
Hu
,
L. A.
Collins
,
J. P.
Colgan
,
V. N.
Goncharov
, and
D. P.
Kilcrease
, “
Optical properties of highly compressed polystyrene: An ab initio study
,”
Phys. Rev. B
96
,
144203
(
2017
).
749.
W.
Huebner
,
A.
Merts
,
N.
Magee
, Jr.
, and
M.
Argo
, “
Astrophysical opacity library
,”
Tech. Rep. No. LA-6760
(
1977
).
750.
H.
Reinholz
,
G.
Röpke
,
S.
Rosmej
, and
R.
Redmer
, “
Conductivity of warm dense matter including electron-electron collisions
,”
Phys. Rev. E
91
,
043105
(
2015
).
751.
N. R.
Shaffer
and
C. E.
Starrett
, “
Dense plasma opacity via the multiple-scattering method
,”
Phys. Rev. E
105
,
015203
(
2022
).
752.
L.
Spitzer
and
R.
Härm
, “
Transport phenomena in a completely ionized gas
,”
Phys. Rev.
89
,
977
(
1953
).
753.
G.
Röpke
,
M.
Schörner
,
R.
Redmer
, and
M.
Bethkenhagen
, “
Virial expansion of the electrical conductivity of hydrogen plasmas
,”
Phys. Rev. E
104
,
045204
(
2021
).
754.
K.
Ramakrishna
,
M.
Lokamani
,
A.
Baczewski
,
J.
Vorberger
, and
A.
Cangi
, “
Electrical conductivity of iron in earth's core from microscopic ohm's law
,”
Phys. Rev. B
107
,
115131
(
2023
).
755.
K.
Ramakrishna
,
M.
Lokamani
,
A.
Baczewski
,
J.
Vorberger
, and
A.
Cangi
, “
Impact of electronic correlations on high-pressure iron: Insights from time-dependent density functional theory
,”
Electron. Struct.
5
,
045002
(
2023
).
756.
R. J.
Magyar
,
L.
Shulenburger
, and
A. D.
Baczewski
, “
Stopping of deuterium in warm dense deuterium from Ehrenfest time-dependent density functional theory
,”
Contrib. Plasma Phys.
56
,
459
466
(
2016
).
757.
Y. H.
Ding
,
A. J.
White
,
S. X.
Hu
,
O.
Certik
, and
L. A.
Collins
, “
Ab initio studies on the stopping power of warm dense matter with time-dependent orbital-free density functional theory
,”
Phys. Rev. Lett.
121
,
145001
(
2018
).
758.
H. B.
Nersisyan
,
M.
Walter
, and
G.
Zwicknagel
, “
Stopping power of ions in a magnetized two-temperature plasma
,”
Phys. Rev. E
61
,
7022
7033
(
2000
).
759.
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
).
760.
C. D.
Hu
and
E.
Zaremba
, “
Z3 correction to the stopping power of ions in an electron gas
,”
Phys. Rev. B
37
,
9268
9277
(
1988
).
761.
S.
Singwi
and
M. P.
Tosi
, “
Correlations in electron liquids
,”
Solid State Phys.
36
,
177
266
(
1981
).
762.
D.
Kremp
,
M.
Schlanges
, and
T.
Bornath
,
Quantum Statistics of Nonideal Plasmas
(
Springer
,
2005
).
763.
D. O.
Gericke
,
S.
Kosse
,
M.
Schlanges
, and
M.
Bonitz
, “
T-matrix approach to equilibium and nonequilibrium carrier-carrier scattering in semiconductors
,”
Phys. Rev. B
59
,
10639
10650
(
1999
).
764.
D. O.
Gericke
and
M.
Schlanges
, “
Beam-plasma coupling effects on the stopping power of dense plasmas
,”
Phys. Rev. E
60
,
904
910
(
1999
).
765.
D. O.
Gericke
,
M.
Schlanges
, and
T.
Bornath
, “
Stopping power of nonideal, partially ionized plasmas
,”
Phys. Rev. E
65
,
036406
(
2002
).
766.
M.
Bonitz
, “
Correlation time approximation in non-Markovian kinetics
,”
Phys. Lett. A
221
,
85
93
(
1996
).
767.
K.
Balzer
,
N.
Schlünzen
, and
M.
Bonitz
, “
Stopping dynamics of ions passing through correlated honeycomb clusters
,”
Phys. Rev. B
94
,
245118
(
2016
).
768.
N.
Schlünzen
,
K.
Balzer
,
M.
Bonitz
,
L.
Deuchler
, and
E.
Pehlke
, “
Time-dependent simulation of ion stopping: Charge transfer and electronic excitations
,”
Contrib. Plasma Phys.
59
,
e201800184
(
2019
).
769.
L.
Borkowski
,
N.
Schlünzen
,
J.-P.
Joost
,
F.
Reiser
, and
M.
Bonitz
, “
Doublon production in correlated materials by multiple ion impacts
,”
Phys. Sta. Sol. (b)
259
,
2100511
(
2022
).
770.
A.
Kononov
,
A.
Olmstead
,
A. D.
Baczewski
, and
A.
Schleife
, “
First-principles simulation of light-ion microscopy of graphene
,”
2D Mater.
9
,
045023
(
2022
).
771.
K.
Balzer
and
M.
Bonitz
, “
Neutralization dynamics of slow highly charged ions passing through graphene nanoflakes - an embedding self-energy approach
,”
Contrib. Plasma Phys.
62
,
e202100040
(
2021
).
772.
A.
Niggas
,
J.
Schwestka
,
K.
Balzer
,
D.
Weichselbaum
,
N.
Schlünzen
,
R.
Heller
,
S.
Creutzburg
,
H.
Inani
,
M.
Tripathi
,
C.
Speckmann
et al, “
Ion-induced surface charge dynamics in freestanding monolayers of graphene and MoS2 probed by the emission of electrons
,”
Phys. Rev. Lett.
129
,
086802
(
2022
).
773.
E.
Babaev
,
A.
Sudbø
, and
N. W.
Ashcroft
, “
A superconductor to superfluid phase transition in liquid metallic hydrogen
,”
Nature
431
,
666
668
(
2004
).
774.
C. F.
Richardson
and
N. W.
Ashcroft
, “
High temperature superconductivity in metallic hydrogen: Electron-electron enhancements
,”
Phys. Rev. Lett.
78
,
118
121
(
1997
).
775.
J. M.
McMahon
and
D. M.
Ceperley
, “
High-temperature superconductivity in atomic metallic hydrogen
,”
Phys. Rev. B
84
,
144515
(
2011
).
776.
C. F.
Richardson
and
N. W.
Ashcroft
, “
Effective electron-electron interactions and the theory of superconductivity
,”
Phys. Rev. B
55
,
15130
15145
(
1997
).
777.
C.
Pellegrini
,
C.
Kukkonen
, and
A.
Sanna
, “
Ab initio calculations of superconducting transition temperatures: When going beyond RPA is essential
,”
Phys. Rev. B
108
,
064511
(
2023
).
778.
C. W.
Myung
,
B.
Hirshberg
, and
M.
Parrinello
, “
Prediction of a supersolid phase in high-pressure deuterium
,”
Phys. Rev. Lett.
128
,
045301
(
2022
).
779.
M.
Boninsegni
and
N. V.
Prokof'ev
, “
Colloquium: Supersolids: What and where are they?
,”
Rev. Mod. Phys.
84
,
759
776
(
2012
).
780.
J.
MacDonald
,
C. A.
Di Stefano
,
T.
Döppner
,
L. B.
Fletcher
,
K. A.
Flippo
,
D.
Kalantar
,
E. C.
Merritt
,
S. J.
Ali
,
P. M.
Celliers
,
R.
Heredia
et al, “
The colliding planar shocks platform to study warm dense matter at the National Ignition Facility
,”
Phys. Plasmas
30
,
062701
(
2023
).
781.
Quoting
P. C.
Martin
, in Ref. 820, p. 2.
782.
B.
Hirshberg
,
V.
Rizzi
, and
M.
Parrinello
, “
Path integral molecular dynamics for bosons
,”
Proc. Natl. Acad. Sci. U. S. A.
116
,
21445
21449
(
2019
).
783.
M. Y.
Feldman
and
B.
Hirshberg
, “
Quadratic scaling bosonic path integral molecular dynamics
,”
J. Chem. Phys.
159
,
154107
(
2023
).
784.
Y.
Xiong
, “
GPU acceleration of ab initio simulations of large-scale identical particles based on path integral molecular dynamics
,” arXiv:2404.02628 (
2024
).
785.
H.
Xie
,
L.
Zhang
, and
L.
Wang
, “
Ab-initio study of interacting fermions at finite temperature with neural canonical transformation
,”
J. Mach. Learn.
1
,
38
59
(
2022
).
786.
P.
Tolias
,
F.
Kalkavouras
, and
T.
Dornheim
, “
Fourier–Matsubara series expansion for imaginary–time correlation functions
,”
J. Chem. Phys.
160
,
181102
(
2024
).
787.
T.
Dornheim
,
P.
Tolias
,
F.
Kalkavouras
,
Z.
Moldabekov
, and
J.
Vorberger
, “
Dynamic exchange-correlation effects in the strongly coupled electron liquid
,” arXiv:2405.08480 (
2024
).
788.
M.
Bethkenhagen
,
A.
Sharma
,
P.
Suryanarayana
,
J. E.
Pask
,
B.
Sadigh
, and
S.
Hamel
, “
Properties of carbon up to 10 million kelvin from Kohn-Sham density functional theory molecular dynamics
,”
Phys. Rev. E
107
,
015306
(
2023
).
789.
V.
Blum
,
R.
Gehrke
,
F.
Hanke
,
P.
Havu
,
V.
Havu
,
X.
Ren
,
K.
Reuter
, and
M.
Scheffler
, “
Ab initio molecular simulations with numeric atom-centered orbitals
,”
Comput. Phys. Commun.
180
,
2175
2196
(
2009
).
790.
M.
Echenique
,
R. M.
Nieminen
,
J. C.
Ashley
, and
R. H.
Ritchie
, “
Nonlinear stopping power of an electron gas for slow ions
,”
Phys. Rev. A
33
,
897
904
(
1986
).
791.
I.
Nagy
,
A.
Arnau
, and
P. M.
Echenique
, “
Nonlinear stopping power and energy-loss straggling of an interacting electron gas for slow ions
,”
Phys. Rev. A
40
,
987
994
(
1989
).
792.
M. J.
van Setten
,
F.
Weigend
, and
F.
Evers
, “
The GW-method for quantum chemistry applications: Theory and implementation
,”
J. Chem. Theory Comput.
9
,
232
246
(
2013
).
793.
C.-K.
Huang
,
B.
Molvig
,
B.
Albright
,
E.
Dodd
,
E.
Vold
,
G.
Kagan
, and
N. M.
Hoffman
, “
Study of the ion kinetic effects in ICF run-away burn using a quasi-1D hybrid model
,”
Phys. Plasmas
24
,
022704
(
2017
).
794.
W.
Taitano
,
B.
Keenan
,
L.
Chacon
,
A. S.
E
,
H.
Hammer
, and
A.
Simakov
, “
An Eulerian Vlasov-Fokker–Planck algorithm for spherical implosion simulations of inertial confinement fusion capsules
,”
Comp. Phys. Comm.
263
,
107861
(
2021
).
795.
L.
Stanton
,
J.
Glosli
, and
M.
Murillo
, “
Multiscale molecular dynamics model for heterogeneous charged systems
,”
Phys. Rev. X
8
,
021044
(
2018
).
796.
G.
Dimonte
and
J.
Daligault
, “
Molecular-dynamics simulations of electron-ion temperature relaxation in a classical coulomb plasma
,”
Phys. Rev. Lett.
101
,
135001
(
2008
).
797.
P.
Mason
,
S.
Banerjee
,
J.
Smith
,
T.
Butcher
,
J.
Phillips
,
H.
Höppner
,
D.
Möller
,
K.
Ertel
,
M.
De Vido
,
I.
Hollingham
et al, “
Development of a 100 J, 10 Hz laser for compression experiments at the High Energy Density instrument at the European XFEL
,”
High Pow. Laser Sci. Eng.
6
,
e65
(
2018
).
798.
A.
Laso Garcia
,
H.
Höppner
,
A.
Pelka
,
C.
Bähtz
,
E.
Brambrink
,
S.
Di Dio Cafiso
,
J.
Dreyer
,
S.
Göde
,
M.
Hassan
,
T.
Kluge
et al, “
ReLaX: The Helmholtz International Beamline for Extreme Fields high-intensity short-pulse laser driver for relativistic laser–matter interaction and strong-field science using the high energy density instrument at the European X-ray free electron laser facility
,”
High Power Laser Sci. Eng.
9
,
e59
(
2021
).
799.
L. B.
Fletcher
,
H. J.
Lee
,
T.
Döppner
,
E.
Galtier
,
B.
Nagler
,
P.
Heimann
,
C.
Fortmann
,
S.
LePape
,
T.
Ma
,
M.
Millot
et al, “
Ultrabright X-ray laser scattering for dynamic warm dense matter physics
,”
Nat. Photonics
9
,
274
279
(
2015
).
800.
L.
Wollenweber
,
T. R.
Preston
,
A.
Descamps
,
V.
Cerantola
,
A.
Comley
,
J. H.
Eggert
,
L. B.
Fletcher
,
G.
Geloni
,
D. O.
Gericke
,
S. H.
Glenzer
et al, “
High-resolution inelastic x-ray scattering at the high energy density scientific instrument at the European X-Ray Free-Electron Laser
,”
Rev. Sci. Instrum.
92
,
013101
(
2021
).
801.
A.
Descamps
,
B. K.
Ofori-Okai
,
K.
Appel
,
V.
Cerantola
,
A.
Comley
,
J. H.
Eggert
,
L. B.
Fletcher
,
D. O.
Gericke
,
S.
Göde
,
O.
Humphries
et al, “
An approach for the measurement of the bulk temperature of single crystal diamond using an x-ray free electron laser
,”
Sci. Rep.
10
,
14564
(
2020
).
802.
L.
Yang
,
L.
Huang
,
S.
Assenbaum
,
T. E.
Cowan
,
I.
Goethel
,
S.
Göde
,
T.
Kluge
,
M.
Rehwald
,
X.
Pan
,
U.
Schramm
et al, “
Time-resolved optical shadowgraphy of solid hydrogen jets as a testbed to benchmark particle-in-cell simulations
,”
Commun. Phys.
6
,
368
(
2023
).
803.
L.
Obst
,
S.
Göde
,
M.
Rehwald
,
F.-E.
Brack
,
J.
Branco
,
S.
Bock
,
M.
Bussmann
,
T. E.
Cowan
,
C. B.
Curry
,
F.
Fiuza
et al, “
Efficient laser-driven proton acceleration from cylindrical and planar cryogenic hydrogen jets
,”
Sci. Rep.
7
,
10248
(
2017
).
804.
N.
Schlünzen
and
M.
Bonitz
, “
Nonequilibrium Green functions approach to strongly correlated fermions in lattice systems
,”
Contrib. Plasma Phys.
56
,
5
91
(
2016
).
805.
M.
Bonitz
,
R.
Nareyka
, and
D.
Semkat
, “
Progress in nonequilibrium Green's functions
,” in
Proceedings of the Conference on Kadanoff-Baym Equations: Progress and Perspectives for Many-Body Physics
(
World Scientific
,
2000
).