You are on page 1of 24

Analytical representation of the Local Field Correction of the Uniform Electron Gas

within the Effective Static Approximation


Tobias Dornheim,1, 2, ∗ Zhandos A. Moldabekov,1, 2 and Panagiotis Tolias3
1
Center for Advanced Systems Understanding (CASUS), D-02826 Görlitz, Germany
2
Helmholtz-Zentrum Dresden-Rossendorf (HZDR), D-01328 Dresden, Germany
3
Space and Plasma Physics, Royal Institute of Technology, Stockholm, SE-100 44, Sweden
The description of electronic exchange–correlation effects is of paramount importance for
many applications in physics, chemistry, and beyond. In a recent Letter, Dornheim et al.
[Phys. Rev. Lett. 125, 235001 (2020)] have presented the effective static approximation (ESA) to the
arXiv:2101.05498v2 [physics.plasm-ph] 26 Feb 2021

local field correction (LFC), which allows for the highly accurate estimation of electronic properties
such as the interaction energy and the static structure factor. In the present work, we give an
analytical parametrization of the LFC within ESA that is valid for any wave number, and available
for the entire range of densities (0.7 ≤ rs ≤ 20) and temperatures (0 ≤ θ ≤ 4) that are relevant for
applications both in the ground state and in the warm dense matter regime. A short implementation
in Python is provided, which can easily be incorporated into existing codes.
In addition, we present an extensive analysis of the performance of ESA regarding the estimation
of various quantities like the dynamic structure factor, static dielectric function, the electronically
screened ion-potential, and also stopping power in electronic medium. In summary, we find that the
ESA gives an excellent description of all these quantities in the warm dense matter regime, and only
becomes inaccurate when the electrons start to form a strongly correlated electron liquid (rs ∼ 20).
Moreover, we note that the exact incorporation of exact asymptotic limits often leads to a superior
accuracy compared to the neural-net representation of the static LFC [J. Chem. Phys. 151, 194104
(2019)].

I. INTRODUCTION wards inertial confinement fusion [35], and is relevant for


the new field of hot-electron chemistry [36, 37].
The accurate description of many-electron systems is of Consequently, WDM is nowadays routinely realized in
paramount importance for many applications in physics, large research facilities around the globe; see Ref. [38] for
quantum chemistry, material science, and related disci- a recent review of different experimental techniques. Fur-
plines [1, 2]. In this regard, the uniform electron gas ther, we mention that there have been many remarkable
(UEG) [3, 4], which is comprised of correlated electrons experimental discoveries in this field, such as the obser-
in a homogeneous, neutralizing positive background (also vation of diamond formation by Kraus et al. [39, 40], or
known as ”jellium” or quantum one-component plasma), the measurement of plasmons in aluminum by Sperling
constitutes a fundamental model system. Indeed, our im- et al. [41].
proved understanding of the UEG has facilitated many At the same time, the theoretical description of WDM
key insights like the quasi-particle picture of collective is notoriously difficult [42, 43] due to the complicated in-
excitations [5] and the Bardeen-Cooper-Schrieffer theory terplay of i) Coulomb coupling, ii) quantum degeneracy
of superconductivity [6]. of the electrons, and iii) thermal excitations. Formally,
In the ground state, many properties of the UEG these conditions are conveniently expressed by two char-
have been accurately determined on the basis of quan- acteristic parameters that are of the order of one simul-
tum Monte Carlo (QMC) simulations [7–19], which have taneously: the density parameter (Wigner-Seitz radius)
subsequently been used as input for various parametriza- rs = r/aB , where r and aB are the average interparti-
tions [20–26]. These, in turn, have provided the basis cle distance and Bohr radius, and the degeneracy tem-
of the possibly unrivaled success of density functional perature θ = kB T /EF , with EF being the usual Fermi
theory (DFT) regarding the description of real materi- energy [1, 44]. In particular, the high temperature rules
als [27–29]. out ground state approaches and thermal DFT [45] sim-
Over the last decade or so, there has emerged a re- ulations, too, require as input an exchange–correlation
markable interest in warm dense matter (WDM)–an ex- (XC) functional that has been developed for finite tem-
otic state with high temperatures and extreme densities. perature [46–49].
In nature, these conditions occur in various astrophysi- This challenge has resulted in a substantial progress re-
cal objects such as giant planet interiors [30–32], brown garding the development of electronic QMC simulations
dwarfs [30, 33], and neutron star crusts [34]. On earth, at WDM conditions [50–68], which ultimately led to the
WDM has been predicted to occur on the pathway to- first parametrizations of the XC-free energy fxc of the
UEG [69, 70], allowing for thermal DFT calculations on
the level of the local density approximation (LDA). At
the same time, DFT approaches are being developed that
∗ t.dornheim@hzdr.de deal efficiently with the drastic increase in the basis size
2

for high temperatures [71–75], and even gradient correc- neural net from Ref. [94]; a short Python implementation
tions to the LDA have become available [49, 76]. is freely available online [102]; ii) we further analyze the
Of particular relevance for the further development of performance of the ESA regarding the estimation of var-
WDM theory is the response of the electrons to an exter- ious electronic properties such as S(q, ω) and χ(q) over a
nal perturbation as it is described by the dynamic den- large range of densities and temperatures.
sity response function χ(q, ω), see Eq. (1) below, where The paper is organized as follows: In Sec. II, we in-
q and ω denote the wave vector and frequency. Such in- troduce the underlying theoretical background including
formation is vital for the interpretation of X-ray Thom- the density response function, its relation to the dynamic
son scattering experiments (XRTS)–a standard method structure factor, and the basic idea of the ESA scheme.
of diagnostics for WDM which gives access to plasma Sec. III is devoted to our new analytical parametrization
parameters such as the electronic temperature [77, 78]. of the LFC within ESA (see Sec. III C for the final result),
Furthermore, accurate knowledge of χ(q, ω) would al- which is analyzed in the subsequent Sec. IV regarding the
low for the construction of advanced XC-functionals for estimation of numerous electronic properties. The paper
DFT based on the adiabatic connection formula and the is concluded by a brief summary and outlook in Sec. V.
fluctuation-dissipation theorem, see Refs. [79–82] for de-
tails, or as the incorporation as the dynamic XC-kernel
in time-dependent DFT [83, 84]. Finally, we mention II. THEORY
the calculation of energy-loss properties like the stop-
ping power [85], the construction of effective ion-ion po- We assume Hartree atomic units throughout this work.
tentials [86–88], the description of electrical and thermal
conductivities [89], and the incorporation of electronic
exchange–correlation effects into other theories such as A. Density response and local field correction
quantum hydrodynamics [90, 91] or average atom mod-
els [92].
The density response of an electron gas to an exter-
Being motivated by these applications, Dornheim and nal harmonic perturbation [64] of wave-number q and
co-workers have recently presented a number of investiga- frequency ω is—within linear response theory—fully de-
tions of both the static and dynamic density response of scribed by the dynamic density response function χ(q, ω).
the warm dense electron gas based on ab initio path inte- The latter is conveniently expressed as [1, 103]
gral Monte Carlo (PIMC) [93] simulations [89, 94–99]. In
particular, they have reported that often a static treat- χ0 (q, ω)
ment of electronic XC-effects is sufficient for a highly ac- χ(q, ω) = 4π , (1)
1− q2 [1 − G(q, ω)] χ0 (q, ω)
curate description of dynamic properties such as χ(q, ω)
or the dynamic structure factor (DSF) S(q, ω). Unfor- where χ0 (q, ω) denotes the density response function
tunately, this static approximation (see Sec. II C below) of an ideal Fermi gas known from theory and the full
leads to a substantial bias in frequency-averaged proper- wave-number- and frequency-resolved information about
ties like the interaction energy v [100]. exchange–correlation effects is contained in the dynamic
To overcome this limitation, Dornheim et al. [100] local field correction G(q, ω). Hence, setting G(q, ω) = 0
have presented the effective static approximation (ESA), in Eq. (1) leads to the well known RPA which entails only
which entails a frequency-averaged description of elec- a mean-field description of the density response.
tronic XC-effects by combining the neural-net represen- Naturally, the computation of accurate data for G(q, ω)
tation of the static local field correction (LFC) from constitutes a most formidable challenge, although first ab
Ref. [94] with a consistent limit for large wave vectors initio results have become available recently at least for
based on QMC data for the pair distribution function parts of the WDM regime [89, 95–98].
evaluated at zero distance; see Ref. [101] for a recent Let us next consider the static limit, i.e.,
investigation of this quantity. In particular, the ESA
has been shown to give highly accurate results for differ- χ(q) = lim χ(q, ω) . (2)
ω→0
ent electronic properties such as the interaction energy
and the static structure factor (SSF) S(q) at the same In this limit, accurate data for Eq. (1) have been pre-
computational cost as the random phase approximation sented by Dornheim et al. [94, 99, 104] based on the re-
(RPA). Furthermore, the value of the ESA for the inter- lation [9]
pretation of XRTS experiments has been demonstrated Z β
by re-evaluating the study of aluminum by Sperling et
χ(q) = −n dτ F (q, τ ) , (3)
al. [41]. 0
The aim of the present work is two-fold: i) we introduce
an accurate analytical parametrization of the LFC within with the imaginary-time density–density correlation
ESA, which exactly reproduces the correct limits at both function being defined as
small and large wave numbers q = |q| and can be easily 1
incorporated into existing codes without relying on the F (q, τ ) = hρ(q, τ )ρ(−q, 0)i . (4)
N
3

We note that Eq. (4) is the usual intermediate scattering the coupling parameter rs . This idea is at the heart of the
function [77], but evaluated at an imaginary-time argu- construction of advanced exchange–correlation function-
ment τ ∈ [0, β]. In addition, we note that it is straight- als for DFT calculations within the ACFDT formulation;
forward to then use χ(q) to solve Eq. (1) for the static see, e.g., Refs. [79–82] for more details.
local field correction
G(q) = lim G(q, ω) C. The static approximation
ω→0
q2
 
1 1
=1− − . (5) Since the full frequency-dependence of G(q, ω) remains
4π χ0 (q) χ(q)
to this date unknown for most parts of the WDM regime
Based on Eq. (5), Dornheim et al. [94] have obtained an (and also in the ground-state), one might neglect dy-
extensive data set for G(q) for Np ∼ 50 different density– namic effects and simply substitute G(q) in Eq. (1). This
temperature combinations. These data—together with leads to the dynamic density response function within the
the parametrization of G(q; rs ) at zero temperature by static approximation [89, 96],
Corradini et al. [25] based on ground-state QMC sim-
ulations [10, 11]—was then used to train a deep neu- χ0 (q, ω)
χstat (q, ω) = 4π , (10)
ral network that functions as an accurate representation 1− q2 [1 − G(q)] χ0 (q, ω)
G(q; rs , θ) for 0 ≤ q ≤ 5qF , 0.7 ≤ rs ≤ 20 and 0 ≤ θ ≤ 4.
which entails the frequency-dependence on an RPA level,
but exchange-correlation effects are incorporated stati-
B. Fluctuation–dissipation theorem cally. Indeed, it was recently shown that Eq. (10) allows
to obtain nearly exact results for χ(q, ω), S(q, ω), and
related quantities for rs . 5 and θ & 1.
The fluctuation–dissipation theorem [1]
Yet, while results for individual wave numbers are rel-
Imχ(q, ω) atively good, the static approximation is problematic for
S(q, ω) = − (6) quantities that require an integration over q, such as
πn(1 − e−βω )
the interaction energy v [100]. More specifically, it can
relates Eq. (1) to the dynamic structure factor S(q, ω) be shown that neglecting the frequency dependence in
and, thus, directly connects the LFC to different mate- the LFC (LFCs that are explicitly defined without a fre-
rial properties. First and foremost, we mention that the quency dependence are hereafter denoted as G(q)) leads
DSF can be directly measured, e.g. with the XRTS tech- to the relation [105]
nique [77], which means that the accurate prediction of
S(q, ω) from theory is of key importance for the diagnos- lim G(q) = 1 − g(0) , (11)
q→∞
tics of state-of-the-art WDM experiments [78].
The static structure factor is defined as the normaliza- where g(0) denotes the pair distribution function (PDF)
tion of the DSF g(r) evaluated at zero distance, sometimes also called
Z ∞ the on-top PDF or contact probability. Yet, is has been
S(q) = dω S(q, ω) , (7) shown both in the ground state [106, 107] and at finite
−∞ temperature [94, 104] that the exact static limit of the
and thus entails an averaging over the full frequency dynamic LFC diverges towards either positive or nega-
range. We stress that this is in contrast to the static tive infinity in the q → ∞ limit. Eq. (11) thus implies
density response function χ(q) introduced in the previ- that using G(q) as G(q) in Eq. (10) leads to a diverging
ous section, which is defined as the limit of ω → 0. The on-top PDF, which is, of course, unphysical. This, too,
SSF, in turn, gives direct access to the interaction energy is the reason for spurious contributions to wave-number
of the system, and for a uniform system it holds [4] integrated quantities like v at large q.

1 ∞
Z
v= dq [S(q) − 1] . (8) D. The Effective Static Approximation
π 0
Finally, we mention the adiabatic connection formula [4,
To overcome these limitation of the static approxima-
69, 70]
tion, Dornheim et al. [100] have proposed to define an
Z rs
1 effectively frequency-averaged theory that combines the
fxc (rs , θ) = 2 drs v(rs , θ)rs , (9) good performance of Eq. (10) for q . 3qF with the con-
rs 0
sistent limit of G(q) from Eq. (11).
which implies that the free energy (and, equivalently the More specifically, this so-called effective static approx-
partition function Z) can be inferred if the dynamic den- imation is constructed as [100]
sity response function—the only unknown part of which
is the dynamic LFC G(q, ω)—of a system is known for all GESA (q; rs , θ) = Gnn (q; rs , θ) (1 − A(x)) (12)
wave numbers and frequencies, and for different values of + (1 − g(0; rs , θ)) A(x) ,
4

with x = q/qF , and where Gnn (q; ts , θ) is the neural-


net representation of PIMC data for the exact static 1.6
limit G(q) = G(q, 0) of the UEG [94], and g(0; rs , θ) rs=20, θ=1 1
1.4
denotes the on-top pair distribution function that was
parametrized in Ref. [100] on the basis of restricted PIMC 1.2
data by Brown et al. [53]. Further, A(x) denotes the ac-
tivation function 1

G(q)

A(x)
1 0.8 0.5
A(x) = A(x, xm , η) = [1 + tanh (η(x − xm ))] (13)
2 PIMC
0.6 STLS
resulting in a smooth transition between Gnn and ESA
Eq. (11) for large q. Here the parameters xm and η can 0.4
1-g(0)
be used to tune the position and width of the activation. 0.2 Static
In practice, the performance of the ESA only weakly de- A(x,xm,η)
pends on η and we always use η = 3 throughout this 0 0
work. The appropriate choice of the position xm is less 0 1 2 3 4 5
trivial and is discussed below. 1.2 q/qF
An example for the construction of the ESA is shown
in Fig. 1 for the UEG at rs = 20 and θ = 1. In the top 1
panel, we show the wave-number dependence of the static
LFC G(q), with the green squares depicting exact PIMC 0.8
data for N = 66 taken from Ref. [104] and the black rs=20, θ=1

S(q)
dashed curve the neural-net representation from Ref. [94]. 0.6
Observe the positively increasing tail at large q from both
data sets, which is consistent to the positive value of the
0.4 1.05
exchange-correlation contribution to the kinetic energy
at these conditions [106, 109].
The solid red line corresponds to the ESA and is indis- 0.2
1
tinguishable from the neural net for q . 2qF . Further, it
smoothly goes over into Eq. (11) for larger q and attains 0
this limit for q & 3.5qF . The purple dash-dotted curve 0 1 2 3 4 5
shows the corresponding activation function A(x) [using q/qF
xm = 3] on the right y-axis and illustrates the shape of
the switchover between the two limits. As a reference,
we have also included G(q) computed within the finite- FIG. 1. Illustration of the effective static approximation
temperature version [105, 108] of the STLS approxima- (ESA) [100] for rs = 20 and θ = 1. Top panel: Static LFC.
tion [110], see the dotted blue curve. First and foremost, Green squares are exact PIMC data for G(q; rs , θ) taken from
Ref. [104], and dashed black line the neural net representa-
we note that STLS constitutes a purely static theory for
tion from Ref. [94]. The solid red curve shows the frequency-
the LFC and, thus, exactly fulfills Eq. (11), i.e., it attains averaged LFC G(q; rs , θ) within ESA [Eq. (12)] and the dot-
a constant value in the limit of large wave numbers, al- ted blue curve the same quantity within STLS [105, 108].
though for significantly larger values of q. In addition, The purple dash-dotted line shows the activation function
STLS is well known to violate the exact compressibility A(x, xm , η) [for xm = η = 3, see Eq. (13)] and corresponds to
sum-rule [108] (see Eq. (15) below) and deviates from the the right y-axis. Top panel: Static structure factor S(q) from
other curves even in the small-q limit. Finally, we note the same methods, and in RPA (dash-dotted yellow).
that it does not reproduce the peak of both the neural
net and ESA around q = 2.5qF .
The bottom panel of Fig. 1 shows the correspond- green squares do correspond to the results one would ob-
ing results for the static structure factor S(q), with the tain if the correct, dynamic LFC G(q, ω) was inserted
green crosses again being the exact PIMC results from into Eq. (1).
Ref. [104]. At this point, we feel that a note of caution This is in contrast to the black dashed curve, that
is pertinent. On the one hand, the PIMC method is lim- has been obtained on the basis of the static approxima-
ited to simulations in the static limit, as dynamic sim- tion, Eq. (10), using as input the neural-net representa-
ulations are afflicted with an exponentially hard phase tion [94] of the exact static limit G(q). Evidently, the
problem [111] in addition to the usual fermion sign prob- static treatment of exchange–correlation effects is well
lem [112]. Therefore, PIMC results for both χ(q, ω) and justified for q . 2qF , but there appear systematic devia-
G(q, ω) are only available for ω = 0. Yet, the PIMC tions for larger q; see also the inset showing a magnified
method is also capable to give exact results for frequency- segment around the maximum of S(q). In particular,
averaged quantities like S(q), as the frequency integra- S(q) does not decay to 1, and, while being small for each
tion is carried out in the imaginary time [93]. Thus, the individual q, the error accumulates under the integral in
5

Eq. (8). 5.5


The solid red curve has been obtained by inserting G(q) fit
within the ESA into Eq. (10). Plainly, the inclusion of 5
the on-top PDF via Eq. (12) removes the spurious ef-
fects from the static approximation, and the ESA curve 4.5
is strikingly accurate over the entire q-range.

xm
4
The dotted blue curve has been computed using G(q)
within the STLS approximation. For small q, it, too 3.5
obeys the correct parabolic limit [59, 113], which is the
consequence of perfect screening in the UEG [1]. For 3
larger q, there appear systematic deviations, and the
correlation-induced peak of S(q) around q ∼ 2.2qF is 2.5
not reproduced by this theory; see also Ref. [104] for an 0 1 2 3 4
extensive analysis including even stronger values of the θ
coupling strength rs .
Finally, the dash-dotted yellow curve has been com- FIG. 2. Temperature dependence of the transition wave-
puted within the RPA. Clearly, neglecting exchange– number xm from Eq. (14).
correlation effects in Eq. (1) leads to an insufficient de-
scription of the SSF, and we find systematic deviations
of up to ∼ 30%. the decreased slope of the positive tail at large wave num-
bers. The red curve shows the ESA results for G(q) using
the transition wave-number obtained from Eq. (14), i.e.,
III. ANALYTICAL REPRESENTATION OF THE xm ≈ 3.58. In particular, the red curve reproduces the
ESA peak structure of the exact static limit G(q), and subse-
quently approaches the large-q limit from Eq. (11) [light
A. Choice of the activation function dotted grey line]. In contrast, the dash-dotted yellow
and dashed-double-dotted purple lines are ESA results
The ESA as it has been defined in Eq. (12) has, in for xm = 3 and xm = 2.5, respectively, and start to sig-
principle, two free parameters, which have to be de- nificantly deviate from G(q) before the peak. Finally, the
fined/parametrized before an analytical representation of dotted blue curve shows G(q) from STLS, and has been
G(q; rs , θ) can be introduced. More specifically, these are included as a reference.
the transition wave number xm and scaling parameter η Regarding S(q), the solid red curve shows the best
from the activation function A(x; xm , η); see Eq. (13). agreement to the PIMC data, whereas the static approx-
Scaling parameter η: We choose η(rs , θ) = 3 = imation again exhibits the spurious behaviour for large
const, as GESA (q; rs , θ) only weakly depends on this pa- q, albeit less pronounced than for θ = 1 shown above.
rameter; see Ref. [100] for an example. The ESA results for xm = 3, too, is in good agreement
Transition wave-number xm : The choice of a to the PIMC data, although there appears an unphysical
reasonable wave-number of the transition between the minimum around q = 3qF . The ESA curve for xm = 2.5,
neural-net and Eq. (11) is less trivial. What we need is on the other hand, does not reproduce the maximum in
a transition around xm ∼ 2.5qF for θ . 1, whereas it S(q) from the other data sets. Finally, the STLS curve
should move to larger wave-number for higher tempera- does not provide an accurate description of the physi-
tures. The dependence on the density parameter rs , on cal behaviour and systematically deviates from the exact
the other hand, is less pronounced and can be neglected. results except in the limits of large and small q.
We thus construct the function

xm (θ) = Ax + Bx θ + Cx θ2 , (14) B. Analytical representation

with Ax , Bx , and Cx being free parameters that we de- Let us start this discussion by introducing a suitable
termine empirically. In particular, we find Ax = 2.64, functional form for the q-dependence of GESA when rs
Bx = 0.31, and Cx = 0.08. A graphical depiction of and θ are fixed. First and foremost, we note that our
Eq. (14) is shown in Fig. 2 parametrization is always constructed from Eq. (12),
An example for the impact of xm on both G(q) and which means that the task at hand is to find an appro-
the corresponding SSF is shown in Fig. 3. The top panel priate representation of Gnn (q; rs , θ) that is sufficiently
shows the LFC, and we observe an overall similar trend accurate in the wave-number regime where the neural
as for θ = 1 depicted in Fig. 1. The main differences both net contributes to the ESA. The correct limit for large q,
in the PIMC data and the neural net results for G(q) are on the other hand, is built in automatically.
i) the comparably reduced height of the maximum, ii) In addition, we would like to incorporate the exact
the increased width of the maximum regarding q, and iii) long-wavelength limit of the static LFC that is given by
6

1.4 2

1.2
rs=20, θ=2 rs=5, θ=0
1.5
1

0.8
G(q)

G(q)
PIMC 1
0.6 STLS QMC
ESA CSR
0.4 1-g(0) STLS
Static 0.5 ESA
xm=3 fit
0.2
xm=2.5 1-g(0)
Static
0 0
0 1 2 3 4 5 0 1 2 3 4 5
1.2 1.4
q/qF q/qF
1.2
rs=5, θ=3
1
1

0.8 0.8

G(q)
rs=20, θ=2 PIMC
CSR
S(q)

0.6
0.6 1.05 STLS
0.4 ESA
fit
0.4 1-g(0)
1 0.2
Static

0.2 0
0 1 2 3 4 5
0.95 q/qF
0
0 1 2 3 4 5
q/qF FIG. 4. Static local field correction for rs = 5 and θ = 0 (top)
and θ = 3 (bottom). Green squares are ground-state QMC
data from Ref. [11] (PIMC data for G(q) from Ref. [94]) for
FIG. 3. Top panel: Static local field correction for rs = 20 θ = 0 (θ = 3), and dashed black lines the neural-net represen-
and θ = 2. Green squares are PIMC data for G(q) from tation from Ref. [94]. The solid red line shows G(q) within the
Ref. [104], and dashed black line the neural-net representation ESA using Eq. (14), and the light blue dash-dotted curve the
from Ref. [94]. The solid red line shows G(q) within the ESA corresponding fit from Eq. (16). The dotted blue line shows
using Eq. (14) [i.e., xm = 3.58], and the dash-dotted yellow G(q) from STLS, and the light grey line the analytical limit
and dash-double-dotted purple line show the ESA for xm = 3 from Eq. (11).
and xm = 2.5. The dotted blue line shows G(q) from STLS,
and the light grey line the analytical limit from Eq. (11).
Bottom panel: Corresponding results for the static structure in the denominator of the square brackets compensates
factor S(q).
the equal pre-factor for large q.
Two examples for the application of Eq. (16) are shown
the compressibility sum-rule [94, 108] (CSR) in Fig. 4, where the local field correction is shown for
rs = 5 and θ = 0 (top) and θ = 3 (bottom). The red
lim G(q; rs , θ) = GCSR (q; rs , θ) (15) curve shows G(q) within the ESA, and the light blue
q→0
dash-dotted curve a fit to these data using Eq. (16) as a
q2 ∂ 2 functional form for θ and rs being constant. First and
=−
(nfxc ) .
4π ∂n2 foremost, we note that the fit perfectly reproduces the
This is achieved by the ansatz ESA, and no fitting error can be resolved with the naked
eye.
Grnn,fit
s ,θ
(q) = GCSR (q; rs , θ) (16) The dash-dotted yellow curves show the CSR
√ [Eq. (15)], which has been included into Eq. (16). In the
1 + αrs ,θ x + β rs ,θ x
 
× , ground state, we indeed find good agreement between
1 + γ rs ,θ x + δ rs ,θ x1.25 + GCSR (q; rs , θ)
the CSR, the QMC data, the neural net, and also the
where x = q/qF is the reduced wave-number and the ESA for q . 2qF . This is somewhat changed for θ = 3,
super-scripts in the four free parameters αrs ,θ , β rs ,θ , where the yellow curve exhibits more pronounced devia-
γ rs ,θ , and δ rs ,θ indicate that they are obtained for fixed tions from the PIMC data and all other curves. Still, we
values of θ and rs . We note that the GCSR (q; rs , θ) term note that the functional form from Eq. (16) is capable
7

to accommodate this finding, and attains the small-wave the functional form of Eq. (18) constitutes an adequate
number limit only for small q in this case. θ
representation of GESA (q; rs ).
We thus conclude that Eq. (16) constitutes a suit-
able basis for the desired analytical representation
GESA (q; rs , θ). As a next step, we make Eq. (16) de- C. Final representation of GESA (q; rs , θ)
pendent on the density parameter rs . To achieve this
goal, we parametrize the free parameters as:
The final step is then given by the construction of an
aθ + bθ rs analytical representation of the full rs -θ-q-dependence by
κ (rs ) = κ θκ ,
θ
(17)
1 + cκ rs expressing the parameters aθκ , bθκ , and cθκ in Eq. (17) as
a function of θ,
with κ ∈ {α, β, γ, δ}. Thus, the characterization of the
rs -dependence for a single isotherm requires the determi- fκ (θ) = af + bf θ + cf θ1.5 . (19)
nation of 12 free parameters. This results in the isother-
mic representation of the LFC of the form This results in three free parameters for each of the 12
coefficients required for the characterization of the rs -
Gθnn,fit ( q ; rs ) = GCSR (q; rs , θ) (18) dependence, i.e., a total of 36 parameters that have to be
 θ θ
√ 
determined by the fitting procedure.
1 + α (rs )x + β (rs ) x
× , The full three-dimensional fit-function is then given by
1 + γ θ (rs )x + δ θ (rs )x1.25 + GCSR (q; rs , θ)
This isothermic representation is illustrated in Fig. 5, Gnn,fit ( q ; rs , θ) = GCSR (q; rs , θ) (20)

where we show the full rs -dependence of the four free pa-
 
1 + α(rs , θ)x + β(rs , θ) x
rameter α − δ (clockwise) for θ = 0 (green), θ = 1 (red), × ,
1 + γ(rs , θ)x + δ(rs , θ)x1.25 + GCSR (q; rs , θ)
and θ = 4 (black). The symbols have been obtained by
fitting Eq. (16) to ESA data for G(q) for constant values where the functions κ(rs , θ) [with κ ∈ {α, β, γ, δ}] are
of rs and θ. The solid lines have been subsequently ob- given by
tained by fitting the representation of Eq. (17) to these
data over the entire rs -range. The resulting curves are aκ (θ) + bκ (θ)rs
κ(rs , θ) = , (21)
indeed smooth and qualitatively capture the main trends 1 + cκ (θ)rs
from the data points. Finally, the dashed curves have
been computed by fitting Eq. (18) to ESA data over the and the θ-dependent coefficients follow Eq. (19).
entire rs -range, but for constant values of θ. Interest- Our final analytical representation of the LFC within
ingly, this final optimization step results in qualitative the effective static approximation immediately follows
change of the description of all four parameters for θ = 4, from plugging Eq. (20) into Eq. (12),
but only mildly changes the results for both θ = 1 and GESA,fit (q; rs , θ) = Gnn,fit (q; rs , θ) (1 − A(x)) (22)
θ = 0.
Let us for now postpone the discussion of the dotted + (1 − g(0; rs , θ)) A(x) .
curve in Fig. 5, and consider Fig. 6 instead. In particular,
The thus fitted coefficients are given in Tab. I, and a
we show the results of the isothermic fitting procedure for
corresponding python implementation is freely available
θ = 1 (top) and θ = 4 (bottom), with the red, green, and
online [102].
black curves corresponding to different data sets for rs =
The resulting analytical representation GESA (q; rs , θ)
0.7, rs = 2, and rs = 5, respectively. More specifically,
is illustrated in Fig. 7, where we compare it (dashed lines)
the solid lines show the ESA reference data for G(q),
to the original ESA data at rs = 5 (top) and rs = 2,
and the dashed curves have been obtained by fitting the
i.e., two metallic densities that are of high interest in the
data points for α − δ shown in Fig. 5 via Eq. (17). For
context of WDM research.
θ = 1, this simple procedure alone leads to an excellent
θ More specifically, rs = 5 corresponds to a strongly cou-
representation of G (q; rs ). The dotted curve has been pled system, where an accurate treatment of electronic
obtained by performing the full isothermic fits, i.e., by exchange–correlation effects is paramount [114]. These
fitting Eq. (18) to ESA data over the entire rs -range, conditions can be realized experimentally in hydrogen
but with θ being constant. Indeed, we find only minor jets [115] and evaporation experiments [47, 114, 116, 117].
deviations between the dashed and the dotted curve. The green, red, black, and blue curves show results for
For θ = 4, on the other hand, the simple representa- θ = 0, θ = 1, θ = 2, and θ = 4, respectively, and we find
tion of the fit parameters from Eq. (16) results in a sub- that our new analytical representation of GESA (q; rs , θ)
θ
stantially less accurate representation of GESA (q; rs ), and is in excellent agreement to the ESA input data every-
the systematic error is most pronounced at high density, where.
rs = 0.7. This shortcoming can be remedied by perform- The bottom panel corresponds to rs = 2, which is rel-
ing the full isothermic fit of the entire q-rs -dependence, evant e.g. for the investigation of aluminum [41, 118].
and the dotted curves are in excellent agreement to the Here, too, we find excellent agreement between the fitted
original ESA data everywhere. We thus conclude that function and the ESA input data for θ = 0 and θ = 4,
8

0.5
2.4
2.2
2
1.8
0 1.6
1.4
1.2
1
α

β
-0.5 θ=0 0.8
θ=1 0.6
θ=4 0.4
0.2
0
-1 fit
-0.2
rs-fit
-0.4
rs-θ-fit -0.6
0 5 10 15 20 0 5 10 15 20
1 5
rs 4.5 rs
0.5
4
0
3.5
-0.5 3
-1 2.5
2
-1.5
δ

γ
1.5
-2 1
-2.5 0.5
0
-3
-0.5
-3.5 -1
-4 -1.5
0 5 10 15 20 0 5 10 15 20
rs rs

FIG. 5. Dependence of the fit parameters α − δ (clockwise) from Eq. (16 on the density parameter rs for θ = 0 (green), θ = 1
(red), and θ = 4 (black). The symbols have been obtained by fitting Eq. (16) to ESA data for individual constant values of
both rs and θ, and the solid lines have been fitted to these data using Eq. (17) as a functional form. The dashed lines have
been obtained from isothermal fits over the full rs -range using Eq. (18), and the dotted curves from the final fit over the full
rs -θ-q-dependence, see Eq. (20).

TABLE I. Fit parameters for the analytic parametrization of GESA (q; rs , θ) from Eq. (20). For each of the coefficients
aα , bα , . . . , cγ , we give the three free parameters from Eq. (19), af , bf , and cf . A short python implementation is freely
available online [102].
aα 0.66477593 −4.59280227 1.24649624
α bα −1.27089927 1.26706839 −0.4327608
cα 2.09717766 1.15424724 −0.65356955
aβ −1.0206202 5.16041218 −0.23880981
β bβ 1.07356921 −1.67311761 0.58928105
cβ 0.8469662 1.54029035 −0.71145445
aγ −2.31252076 5.83181391 2.29489749
γ bγ 1.76614589 −0.09710839 −0.33180686
cγ 0.56560236 1.10948188 −0.43213648
aδ 1.3742155 −4.01393906 −1.65187145
δ bδ −1.75381153 −1.17022854 0.76772906
cδ 0.63867766 1.07863273 −0.35630091

while small, yet significant deviations appear at interme- We thus conclude that our analytical representation of
diate wave numbers for θ = 2 and θ = 1. Still, it is im- GESA (q; rs , θ) provides a highly accurate description of
portant to note that these deviations do not exceed the electronic–exchange correlation effects over the entire rel-
statistical uncertainty of the original PIMC input data evant parameter range. The application of this represen-
for G(q) on which the neural net from Ref. [94] and the tation for the computation of other material properties
ESA are based. like the static structure factor S(q), interaction energy v,
9

1.2 θ=1 1.2 rs=5


1 1

0.8 0.8
G(q)

G(q)
0.6 rs=0.7 0.6 θ=0
rs=2 θ=1
rs=5 θ=2
0.4 0.4
θ=4
ESA
0.2 fit 0.2 ESA
rs-fit q-rs-θ fit
0 0
0 1 2 3 4 5 6 0 1 2 3 4 5 6 7
q/qF q/qF
1.2 θ=4 1.2 rs=2
1 1

0.8 0.8
G(q)

G(q)
0.6 rs=0.7 0.6 θ=0
rs=2 θ=1
rs=5 θ=2
0.4 0.4
θ=4
ESA
0.2 fit 0.2 ESA
rs-fit q-rs-θ fit
0 0
0 1 2 3 4 5 6 0 1 2 3 4 5 6 7
q/qF q/qF

FIG. 6. Illustration of the isothermic fit function of the local FIG. 7. Analytical representation of GESA (q; rs , θ): Shown
field correction GESA (q) for θ = 1 (top) and θ = 4 (bottom). are ESA results (solid lines) and our final analytical represen-
The red, green, and black curves depict different results for tation, Eq. (22).
rs = 0.7, rs = 2, and rs = 5, respectively. Solid: ESA;
dashed: fitted rs -dependence of the individual coefficients α−
δ from Eq. (16) according to Eq. (17); dotted: full isothermic whereas the ESA converges towards Eq. (11). In addi-
θ
fits of GESA (q) via Eq. (18). tion, our parametrization nicely reproduces the neural-
net for x < xm (θ), which further illustrates the high
quality of the representation. Finally, we find that the
or dielectric function (q) is discussed in detail in Sec. IV. exact static limit of the LFC, too, becomes increasingly
flat at large q for high temperatures, which can be seen
particularly well for rs = 20. In fact, simultaneously con-
IV. RESULTS sidering large values of rs and θ brings us to the classical
limit, where G(q) converges towards one for large wave
numbers [119],
A. The static local field correction
lim Gclassical (q) = 1 . (23)
q→∞
Let us begin the investigation of the results that can be
obtained within the ESA by briefly recapitulating a few Moreover, the ESA and G(q) converge in this regime as
important properties of GESA (q; rs , θ) itself. To this end, the static structure factor can always be computed from
we show the LFC in the θ-q-plane for rs = 20 (top) and the static LFC only via the exact relation [119, 120]
rs = 5 (bottom) in Fig. 8. More specifically, the dashed
black lines show the neural-net results for G(q) from 1
Sclassical (q) = 4π . (24)
Ref. [94], and the solid red lines the corresponding data 1− q2 [Gclassical (q) − 1] βn
for our analytical representation of GESA (q; rs , θ). First
and foremost, we note that the temperature dependence In other words, the spurious effects due to the static ap-
is qualitatively similar for both values of the density pa- proximation and the need for the ESA in WDM appli-
rameter; a more detailed analysis of the rs -dependence of cations are a direct consequence of quantum effects on
the LFC is presented in Fig. 9 below. As usual, G(q) ex- electronic exchange–correlation effects, which only van-
hibits a non-constant behaviour for large wave numbers, ish in the classical limit.
10

ESA ESA
G(q;rs,θ) Static G(q;rs,θ) Static
θ=1
rs=20
2
2
1.5
1.5
1
1
0.5
0.5 6
0 6 5
5 20 4
1 10
4 3
2 3 5
2 q/qF
2 2
θ 3 q/qF 1
1 rs 1
40 0.7 0

ESA
G(q;rs,θ) Static FIG. 9. The local field correction in the rs -q-plane: The
solid red and dashed black curves show our analytical repre-
rs=5 sentation of GESA (q; rs , θ) and the neural-net representation
of the exact static limit G(q) from Ref. [94] for θ = 1. Note
2
the logarithmic scale of the rs -axis.
1.5

1
and, as usually, attains the consistent limit for G(q) given
0.5 by Eq. (11) for all values of rs .
As a further motivation for our ESA scheme, we con-
0 6 sider an effective local field correction Ginvert (q), which,
5 by definition, exactly reproduces QMC data for S(q)
1
4
2 3 where they are available. More specifically, such a quan-
2 tity can be defined as
θ 3 q/qF
1
40  
Ginvert (q) = minG S G (q) − S(q) , (25)

FIG. 8. The local field correction in the θ-q-plane: The solid where S G (q) denotes the SSF computed with respect to
red and dashed black curves show our analytical representa- some trial static LFC G. In practice, we solve Eq. (25)
tion of GESA (q; rs , θ) and the neural-net representation of the by scanning over a dense G-grid for each q-point and
exact static limit G(q) from Ref. [94] for rs = 20 (top) and search for the minimum deviation in the SSF. In this
rs = 5 (bottom). way, we have effectively inverted S(q) for the LFC G,
even though the relation between the two quantities is not
straightforward when quantum mechanical effects cannot
Let us next consider the dependence of the LFC on be neglected.
the density parameter rs , which is shown in Fig. 9 for The results for this procedure are depicted in Fig. 10,
θ = 1. For strong coupling, we observe a positive tail in where we show different LFCs at rs = 20. The top and
the neural-net results for G(q) which begins at smaller center panels corresponds to θ = 0 and θ = 1, and both
values of x = q/qF for larger rs . Between rs = 2 and G(q) and GESA (q) exhibit the familiar behaviour that has
rs = 1, i.e., in the middle of the WDM regime, this be- been discussed in the context of Fig. 1 above. The yellow
haviour changes and we find instead a negative slope, triangles show the inverted results for Eq. (25) and are in
which ultimately even leads to negative values of G(q). remarkably good agreement to both G(q) and GESA (q)
From a physical perspective, the long wave-number limit for q . 2qF . For larger q, Ginvert (q) follows GESA (q)
is dominated by single-particle effects and the sign of the and attains the same finite limit instead of diverging like
slope follows from the exchange–correlation contribution the exact static limit of the LFC. In fact, the curves can
to the kinetic energy K [106, 107], which changes its sign hardly be distinguished within the given level of accuracy
at these conditions [101, 109]. (in particular at θ = 0), which further substantiates the
The ESA, on the other hand, is invariant to this effect simple construction of the ESA, Eq. (12).
11

reliable for q ≤ 5qF as data for larger wave numbers had


1.8 not been included into the training of the neural net, see
rs=20, θ=0
1.6 Ref. [94] for details.
1.4 Unsurprisingly, the inverted data for Ginverted (q)
1.2 closely follow GESA (q) over the entire q-range, and both
ESA and the static approximation give highly accurate
G(q)

1
results for S(q) and v.
0.8
Ginvert Let us next more closely examine the connection be-
0.6 Gclassical tween the ESA and the classical limit, where G(q) is suf-
0.4 ESA ficient to compute exact results for S(q), see Eq. (24)
1-g(0)
0.2 Static above. In particular, Eq. (24) can be straightforwardly
0
solved for G(q), which gives the relation
0 1 2 3 4
1.4 q2
 
q/qF
1 1
rs=20, θ=1
Gclassical (q) = 1 − −1 , (26)
1.2 4π S(q) βn

1 which, too, is exact in the classical limit.


At the same time, it is interesting to evaluate Eq. (26)
0.8
for a quantum system to gauge the impact of quantum
G(q)

PIMC
0.6 Ginvert effects on exchange–correlation effects at different wave
Gclassical numbers q. The results are depicted by the purple di-
0.4 ESA
1-g(0)
amonds in Fig. 10. In the ground state, i.e., β → ∞,
Static it holds Gclassical (q) = 1 for all q, as the second term
0.2
is proportional to T and, hence, vanishes. For θ = 1,
0 Gclassical (q) does depend on q, but is still qualitatively
0 1 2 3 4 wrong over the entire depicted wave-number range. In
q/qF particular, it strongly violates the compressibility sum-
1.4
rule Eq. (15) and does not even decay to zero in the
1.2
rs=20, θ=4 limit of small q. Finally, Gclassical (q) does more closely
resemble the other curves at θ = 4, but still substantially
1
deviates everywhere. We thus conclude that quantum ef-
0.8 fects are paramount even at θ = 4 and rs = 20, and can
G(q)

PIMC only be neglected at significantly higher temperatures.


0.6 Ginvert
Gclassical
0.4 ESA
1-g(0) B. The static structure factor
Static
0.2

0 The next quantity to be investigated with the ESA


0 1 2 3 4 5 6 7 scheme is the static structure factor S(q), which we show
q/qF in Fig. 11. The left column corresponds to rs = 20 and,
thus, constitutes the most challenging case for the ESA
due to the dominant character of exchange–correlation
FIG. 10. Inverted local field correction at finite temperature: effects at these conditions.
The for rs = 20 at θ = 0 (top), θ = 1 (center) and θ = Let us start with the top panel, showing results for
4 (bottom). Green squares and black dashed line: PIMC the ground state. The green squares are state-of-the-art
results for G(q) from Ref. [104] and corresponding neural-net diffusion Monte Carlo results by Spink et al. [17] and con-
results [94]. Solid red and dotted grey: ESA and large-q limit,
stitute the gold standard for benchmarks. The solid red
Eq. (11). Yellow triangles: inverted LFC Gclassical (q), see
Eq. (25). Purple diamonds: LFC from the classical relation curve has been obtained using GESA (q) and is in remark-
Eq. (26). able agreement for all q, even in the vicinity of the peak
of S(q) around q ≈ 2.25qF . In contrast, the blue dotted
STLS curve does not capture this feature and exhibits
pronounced systematic deviations except in the limits of
Let us briefly postpone the discussion of the purple di- small and large wave numbers.
amonds and instead consider the bottom panel of Fig. 10 The center panel in the left column has been obtained
showing results for θ = 4. At these conditions, G(q) and for θ = 0.5, and the green squares are finite-T PIMC
GESA (q) only start to noticeably deviate for q & 5qF , and data taken from Dornheim et al. [104]. Again, the ESA
the PIMC data, too, appear to remain nearly constant gives a very good description of S(q), although the peak
for large q. In addition, the black dashed curve is only height is somewhat overestimated. Still, the description
12

1.2 1.2

1 1

0.8 0.8
rs=20, θ=0 rs=5, θ=0
S(q)

S(q)
0.6 0.6
0.4 1
0.4 0.3 0.4

0.2
0.2 0.2 0.95
0.1
0 0
1.2 0 1 2 3 4 1.2 0 1 2 3 4
q/qF q/qF
1 1

0.8 0.8
rs=20, θ=0.5 rs=5, θ=1
S(q)

S(q)
0.6 0.6
0.4
1
0.4 0.3 0.4

0.2
0.2 0.2
0.95
0.1
0 0
1.2 0 1 2 3 4 1.2 0 1 2 3 4
q/qF q/qF
1 1

0.8 0.8
rs=20, θ=4 rs=5, θ=4
S(q)

S(q)

0.6 0.6 1
0.4
0.4 0.3 0.4

0.2
0.2 0.2 0.95
0.1
0 0
0 1 2 3 4 0 1 2 3 4
q/qF q/qF

FIG. 11. Static structure factor at rs = 20 (left) and rs = 5 (right) at different θ. Green squares: T = 0 QMC data [17] and
finite-T PIMC data [104]; Solid red: ESA; dotted blue: STLS [105, 108, 110]; dash-dotted yellow: RPA.

is strikingly improved compared to the STLS approxima- Finally, we mention the dash-dotted yellow curve in all
tion. three panels, that have been obtained within RPA. Ev-
idently, this mean field description is unsuitable at such
Lastly, the bottom panel has been obtained for θ = 4, low densities even at relatively high values of the reduced
where ESA cannot be distinguished from the PIMC ref- temperature θ.
erence data within the given Monte Carlo error bars.
STLS, too, is quite accurate in this regime, although The right column of Fig. 11 has been obtained for a
there remain systematic deviations at intermediate q. density that is of prime interest to WDM research, rs =
13

5. Again, the top panel corresponds to the ground-state Finally, the bottom right panel shows results for very
and shows relatively good agreement between diffusion strong coupling, rs = 20. Overall, the ESA gives the
Monte Carlo, ESA, and STLS, although the latter does most accurate data for v of all depicted approximations,
not capture the small correlation induced peak in S(q). and is particularly good both at large temperature and in
The RPA, on the other hand, remains inaccurate despite the ground state. In contrast, the STLS approximation
the reduced coupling strength compared to the left panel. for G(q) results in a relatively constant relative deviation
At θ = 1 (center panel), the situation is quite similar, of ∼ 2 − 3%, whereas the static approximation cannot
with the ESA being nearly indistinguishable to the PIMC reasonably used for this values of the density parameter.
data over the entire q-range, whereas STLS is too large
for small and too small for large wave numbers.
Finally, the bottom panel corresponds to θ = 4. Here, D. Density response function
too, only the ESA is capable to reproduce the PIMC
data, whereas STLS and in particular RPA exhibit sys- This section is devoted to a discussion of the suitability
tematic errors. of frequency-averaged LFCs for the determination of the
exact static limit of the density response function χ(q).
In this case, the previously discussed static approxima-
C. Interaction energy tion, i.e., using the neural-net representation of G(q, 0)
from Ref. [94], is exact, and the large-q limit of frequency-
The next important quantity to be investigated in this independent theories G(q) given by Eq. (11) is spurious.
work is the interaction energy v, which, in the case On the other hand, we might expect that the impact of
of a uniform electron gas, is simply given by a one- the LFC decreases for large q, such that GESA (q) and
dimensional integral over the static structure factor S(q) G(q) could potentially give similar results.
[see Eq. (8)] that we evaluate numerically. The results To resolve this question, we show χ(q) in Fig. 13 for
are shown in Fig. 12, where we depict the θ-dependence three representative values of the density parameter rs ,
of v for four relevant values of the density parameter rs . with the green, red, and black sets of curves correspond-
More specifically, the top left panel corresponds to ing to θ = 4, θ = 1, and θ = 0, respectively. Let us start
rs = 2, i.e., a metallic density that is typical for WDM with the top panel showing results for a metallic den-
experiments using various materials, and we plot the rela- sity, rs = 2, with the dotted, dashed, and solid curves
tive deviation in v compared to the accurate parametriza- corresponding to ESA, the exact static limit, and STLS,
tion of the UEG by Groth et al. [69]. At these conditions, respectively. Firstly, we note that all three curves exhibit
both the ESA (solid red) and the static approximation the correct parabolic shape for small wave-numbers [113],
(dashed grey) are very accurate over the entire θ-range, 4π
with a maximum deviation of ∆v/v ∼ 1%. The STLS ap- lim χ(q) = − . (27)
q→0 q2
proximation (dotted blue), too, is capable to provide ac-
curate results for v, with a maximum deviation of ∼ 2%. In particular, Eq. (27) is a direct consequence of the
Let us proceed to the top right panel corresponding to 4π/q 2 pre-factor in front of the LFC in Eqs. (1) and (10),
rs = 5, a relatively sparse density that can be realized e.g. which means that its impact vanishes for small q. With
in experiments with hydrogen jets, see above. First and increasing wave numbers, χ(q) exhibits a broad peak
foremost, we note that both the ESA and STLS provide around q ≈ 1.5qF , which is also well reproduced by all
a remarkably good description of the interaction energy, curves. Moreover, the ESA is virtually indistinguishable
and the systematic error never exceeds 2%. Somewhat from the exact result for all three temperatures, whereas
surprisingly, STLS even gives slightly more accurate dara STLS noticeably deviates, in particular at θ = 0.
for small values of θ compared to ESA. Yet, this is due The center panel shows the same analysis for rs =
to a fortunate cancellation of errors in S(q) under the 5. As discussed above, the increased coupling strength
integral in Eq. (8) [S(q) is too large for small q and too means that the impact of the LFC is more pronounced in
small for large q, which roughly balances out] [4, 100], this case, and the STLS curve substantially deviates at
since the static structure factor S(q) is comparatively intermediate wave numbers, except for the highest tem-
much better in ESA than in STLS, cf. Fig. 11. In ad- perature θ = 4. In stark contrast, the ESA is in excellent
dition, we note that the static approximation performs agreement to the exact curve everywhere, and we find
substantially worse for low temperatures, which is due to only minor deviations for 2qF . q . 3qF . In this sense,
the unphysically slow convergence of S(q) towards 1 for the ESA combines the best from two worlds, by giving
large q, see Secs. II C and II D above. excellent results both for frequency-averaged quantities
The bottom left panel shows the same analysis for rs = like S(q), and really static properties like χ(q, 0) over the
10, and even for this strong coupling strength that con- entire WDM regime.
stitutes the boundary of the electron liquid regime [96], This nice feature of the ESA is only lost when enter-
the error in ESA does not exceed 2%. In addition, the ing the strongly coupled electron liquid regime, as it is
STLS exhibits a comparable accuracy in v, whereas the demonstrated in the bottom panel of Fig. 13 for rs = 20.
static approximation fails at low θ as it is expected. In this case, the static density response function is more
14

0
Δv/v [%]

Δv/v [%]
0
rs=2 -2 rs=5
ESA ESA
STLS STLS
Static Static

0 1 2 3 4 0 1 2 3 4
4
θ θ
0
2
0

Δv/v [%]
Δv/v [%]

-2 -2
-4
rs=10 -6 rs=20
-4
ESA -8 ESA
STLS STLS
-6 Static -10 Static

0 1 2 3 4 0 1 2 3 4
θ θ

FIG. 12. Relative deviation in the interaction energy v [see Eq. (8)] compared to the parametrization by Groth et al. [69]. Solid
red circles: ESA; dotted blue diamonds: STLS [105, 108, 110]; dashed grey crosses: static approximation using the neural-net
representation from Ref. [94].

sharply peaked at low temperature and exhibits a non- impact of the LFC vanishes for small q as it has been
trivial shape that is difficult to resolve. Therefore, the explained above, such that even the mean-field descrip-
STLS approximation is not capable to give a reasonable tion within the RPA becomes exact in this limit. The
description of either the peak position or the shape, see dielectric function, on the other hand, always diverges
Ref. [104] for a more extensive analysis on this point in- for small q, and this divergence is connected to the CSR
cluding even larger values of the density parameter rs . for the static LFC [Eq. (15)] [89, 108],
The ESA, on the other hand, is strikingly accurate for
both θ = 4 and θ = 1, but substantially deviates from 4πχ0 (q)
lim (q) = − , (29)
the exact curve for 2qF . q . 4qF in the ground state. q→0 q2 [1 + 4πCχ0 (q)]
where C is the pre-factor to the parabola in Eq. (15),
E. Dielectric function 1 ∂2
C=− (nfxc ) . (30)
4π ∂n2
The dynamic dielectric function (q, ω) is defined as In principle, exact knowledge of the static LFC as it is
χ(q, ω) encoded in the neural-net representation from Ref. [94]
(q, ω) = 1 − q2
, (28) gives access to the exact static dielectric function de-
4π + χ(q, ω)
picted in Fig. 14. Yet, while the exact relation Eq. (15)
and is important in both classical and quantum electro- was indeed incorporated into the training procedure of
dynamics, in particular for the description of plasma os- the neural net, it was not strictly enforced and, thus, is
cillations [98, 121, 122]. Since a more detailed analysis of only fulfilled by the static (grey dashed) curve with a
this quantity has been presented elsewhere [89, 98], here finite accuracy. Therefore, this curve violates Eq. (29)
we restrict ourselves to a brief discussion of ESA results and attains a finite value in the limit of q → 0, which is
for the static limit of Eq. (28), (q). unphysical.
The results are shown in Fig. 14, where the left panel Our new analytical representation of GESA (q), in con-
shows the dielectric function for rs = 5 and θ = 1. Re- trast, exactly incorporates the CSR, which means that
markably, we find substantial disagreement between the the solid red curve exhibits the correct asymptotic be-
different results for small wave numbers q, which is in haviour (depicted as the dash-dotted green curve). Fi-
striking contrast to linear response properties like χ(q) nally, the dotted blue curve has been obtained on the
and also the SSF S(q). For the latter quantities, the basis of the approximate GSTLS (q), and starkly deviates
15

0 describe both (q) and −1 (q) at the same time.


Let us conclude this section with an example at strong
-0.01
coupling, rs = 20 and θ = 1, depicted in Fig. 15. Firstly,
-0.02 we note that here the ESA and CSR curves for (q) di-
verge towards negative infinity, which is the result of
-0.03 a negative compressibility at these conditions, see also
θ=4
χ

-0.04 θ=1
Refs. [89, 108]. For completeness, we note that this is a
θ=0 necessary, but not sufficient condition for instability [1],
-0.05 and, thus, not problematic. The STLS curve, too, di-
ESA
Static verges towards negative infinity, although with a substan-
-0.06
rs=2 STLS tially different slope. Finally, the static curve becomes
-0.07 increasingly inaccurate for small q and again attains a
0 1 2 3 4 5 finite value for q = 0.
0
q/qF Regarding the inverse dielectric function (right panel),
the negative compressibility is reflected by a nontrivial
-0.005
shape of this quantity, with a minimum around q ≈ 1.8qF .
Here, too, we note that ESA and the static curve are
-0.01 in excellent agreement everywhere, whereas the STLS
θ=4 approximation gives a substantially wrong prediction of
χ

θ=1 both the location and the depth of the minimum in


-0.015
θ=0
−1 (q).
ESA
-0.02 Static
rs=5 STLS
F. Dynamic structure factor
-0.025
0 1 2 3 4 5
0
q/qF The final property of the UEG to be investigated in
-0.001
this work is the dynamic structure factor S(q, ω), which
is shown in Fig. 16 for θ = 1. The left panel corre-
-0.002 sponds to the usual metallic density, rs = 2, and the dot-
ted green curves are ab initio PIMC results taken from
-0.003 θ=4 Ref. [96] that have been obtained by stochastically sam-
χ

θ=1 pling the dynamic LFC G(q, ω). In addition, the solid
θ=0
-0.004 red and dashed black curves have been obtained by us-
ESA ing the ESA and the static approximation, and are in vir-
-0.005 Static tually perfect agreement to the PIMC data everywhere.
rs=20 STLS
This illustrates that a static description of the LFC is
-0.006 fully sufficient to describe the dynamic density response
0 1 2 3 4 5
of electrons at these conditions, see also Refs. [89, 95–97]
q/qF
for more details.
The right panel corresponds to a stronger coupling
strength, rs = 10, which is located at the margins of
FIG. 13. Static density response function χ(q) for rs = 2
(top), rs = 5, and rs = 20 (bottom). The dotted, solid, and the electron liquid regime. While the ESA and static
dashed lines have been obtained by inserting into Eq. (10) approximation here, too, basically give the same results,
GESA (q), GSTLS (q), and the neural-net representation of both curves exhibit systematic deviations towards the ex-
G(q, 0) from Ref. [94], respectively. Green curves: θ = 4; act PIMC data. This is a direct consequence of the in-
red: θ = 1; black: θ = 0. creased impact of the frequency-dependence of electronic
exchange–correlation effects expressed via the dynamic
LFC at these conditions [96].
Interestingly, the impact of the dynamic LFC only
from the exact asymptotic limit. Indeed, the violation manifests in a pronounced way in the shape of S(q, ω),
of the CSR is a well-known shortcoming of the STLS ap- whereas its normalization [i.e., the SSF, see Eq. (7)] is
proach [108], which has ultimately led to the development hardly affected. This is demonstrated in Fig. 17, where
of the approach by Vashista and Singwi [123, 124]. we show the corresponding S(q) for the same conditions.
The right panel of Fig. 14 shows the corresponding For example, for both q = 1.25qF and q = 1.88qF , the
data for the inverse dielectric function −1 (q). Here the shape of the PIMC data for S(q, ω) significantly deviates
static and ESA curves are in excellent agreement over the from the other curves, whereas the SSF is nearly perfectly
entire q-range, which, again, highlights the value of the reproduced by both the ESA and the static approxima-
analytical parametrization which is capable to accurately tion.
16

1000 1.2

1
800

0.8
600

ε-1
0.6
ε

400 rs=5, θ=1 rs=5, θ=1


0.4
ESA ESA
200 static static
0.2
CSR CSR
STLS STLS
0 0
0 0.5 1 0 1 2 3 4
q/qF q/qF

FIG. 14. Left: Static dielectric function (q) for rs = 5 and θ = 1. Solid red: ESA; dashed grey: exact static limit using
the neural-net from Ref. [94]; dash-dotted green: CSR, Eq. (15); dotted-blue: STLS [105, 108, 110]. Right: Same data for the
inverse dielectric function −1 (q).

300 1
ESA
static
200 CSR
STLS 0.5

100
ε-1

0
ε

0 rs=20, θ=1
-0.5 ESA
-100 static
rs=20, θ=1 CSR
STLS
-200 -1
0 0.5 1 0 1 2 3 4
q/qF q/qF

FIG. 15. Left: Static dielectric function (q) for rs = 20 and θ = 1. Solid red: ESA; dashed grey: exact static limit using
the neural-net from Ref. [94]; dash-dotted green: CSR, Eq. (15); dotted-blue: STLS [105, 108, 110]. Right: Same data for the
inverse dielectric function −1 (q).

For larger q, the results for the SSF of G(q) and G. Test charge screening.
GESA (q) do start to deviate, but this has no pronounced
impact on S(q, ω) itself. According to linear response theory, the screened po-
tential of an ion (with charge Ze) can be computed using
the static dielectric function as [88, 125]:

d3 q 4πZe eiq·r
Z
We thus conclude that both the usual static approxi- Φ(r) = , (31)
(2π)3 q 2 (q)
mation and our new ESA scheme [100] are equally well
suited for the description of dynamic properties at WDM which is valid for the weak electron-ion coupling. The
conditions, but are not suited for a qualitative description latter condition is satisfied at large distances from the
of the dynamic density response of the strongly coupled ion [126].
electron liquid regime, for which a fully dynamic local As discussed in Sec. IV E above, the violation of the ex-
field correction has been shown to be indispensable. act limit Eq. (29) leads to the unphysical behavior of the
17

ESA ESA
S(q,ω) static S(q,ω) static
PIMC PIMC

rs=2, θ=1 rs=10, θ=1


0.2
0.5

0.1

0 0

1 15 1 7
12 6
5
9 4
q/qF 2 q/qF 2 3
6
2
3 ω/ωp 1 ω/ωp
30 30

FIG. 16. Dynamic structure factor of the uniform electron gas at θ = 1 for rs = 2 (left) and rs = 10 (right). Solid red:
ESA; dashed black: static approximation; dotted green: ab initio reconstructed PIMC results using a stochastically sampled
dynamic LFC, taken from Ref. [96].

1.2 electric function is used to compute the screened poten-


tial. To illustrate this, we show the screened ion potential
1 (with Z = 1) for rs = 2, θ = 0.5 and θ = 1.0 in Fig. 18,
where the screened ion potential is computed using ESA
given by Eq. (22), the neural-net representation of the
0.8
LFC from Ref. [94], and RPA.
rs=10, θ=1
From Fig. 18, it is clearly seen that the neural-net
S(q)

0.6
representation based result for the screened potential
1 exhibits an ∼ 1/r asymptotic behavior at large dis-
0.4
tances. In contrast, the screened potential obtained us-
0.95 ing the analytical representation GESA (q; rs , θ) correctly
0.2 reproduces complete screening like RPA based data,
0.9
with a Yukawa type exponential screening at large dis-
0 tances [126]. Finally, we note that electronic exchange–
0 1 2 3 4 5
correlation effects, taken into account by using the LFC,
q/qF lead to a stronger screening of the ion potential compared
to the RPA result [88, 126, 127].

FIG. 17. Static structure factor of the UEG for rs = 10 and


θ = 1 (cf. right panel of Fig. 16). Green squares: PIMC data
taken from Ref. [96]; solid red: ESA; dashed black: static ap- H. Stopping power
proximation; dotted blue: STLS [105, 108, 110]; dash-dotted
yellow: RPA. A further example for the application of the LFC is
the calculation of the stopping power, i.e. the mean en-
ergy loss of a projectile (an ion) per unit path length,
static dielectric function computed using the neural-net and related quantities such as the penetration length,
representation of the LFC from Ref. [94]. This results in straggling rate etc. These energy dissipation character-
incomplete screening when the corresponding static di- istics are of paramount importance for such applications
18

FIG. 19. Stopping power at rs = 2, θ = 0.5 and θ = 1.Solid


red: data computed using the analytical representation of the
ESA Eq. (22); dashed black: static approximation computed
using the neural-net representation of the static local field
correction from Ref. [94]; dotted blue: RPA result. The lower
x axis corresponds to v/vth and the upper x axis to v/vF , with
vth and vF being the thermal velocity and Fermi velocity of
electrons, respectively.

reader to Ref. [85] for a more detailed study in a wider


parameter range.
FIG. 18. Screened ion potential at rs = 2, θ = 0.5 and
θ = 1. Solid red: the data computed using the analytical
representation of the ESA Eq. (22); dashed black: static ap- V. SUMMARY AND DISCUSSION
proximation computed using the neural-net representation of
the static local field correction from Ref. [94]; dotted blue:
RPA result; dashed grey line shows ∼ 1/r behavior of the A. Summary
neural-net representation based data at large distances.
The first main achievement of this work is the con-
struction of an accurate analytical representation of the
as ICF and laboratory astrophysics [128, 129]. A linear
effective static approximation for the local field correction
response expression based on the dynamic dielectric func-
GESA (q; rs , θ) covering all wave-numbers and the entire
tion that describes the stopping power for a low-Z pro-
relevant range of densities (0.7 ≤ rs ≤ 20) and temper-
jectile when the ion–electron coupling is weak [130, 131]
atures (0 ≤ θ ≤ 4). Our fit formula [Eq. (20)] well re-
is given by [130]:
produces the original ESA scheme presented in Ref. [100]
2Z 2 e2 ∞ dk kv
Z Z 
−1
 while exactly incorporating the CSR in the limit of small
S(v) = dω ω Im , (32) wave numbers, and without the need for the evaluation
πv 2 0 k 0 (k, ω)
of the neural-net from Ref. [94]. A short implementation
where v is the ion velocity. of Eq. (20) in Python is freely available online [102] and
Recently, using Eq. (32), the neural-net representation can easily be incorporated into existing codes; see the
of the LFC [94] was used to study the ion energy-loss next section for a short list of potential applications.
characteristics and friction in a free-electron gas at warm The second aim of this paper is the further analysis
dense matter conditions [85]. Therefore, it is required to of the ESA in general and our fit formula in particu-
check whether the discussed unphysical behavior of cer- lar regarding the estimation of various electronic proper-
tain quantities based on the neural-net representation of ties. Here one finding of considerable interest has been
the LFC [94] also manifests in the stopping power. The the estimation of an effective static LFC Ginvert (q) that,
comparison of the ESA (22) based data for the stopping when being inserted into Eq. (1), exactly reproduces the
power to the results obtained using the neural-net repre- static structure factor S(q) known from QMC calculation
sentation of the LFC [94] is shown in Fig. 19 for rs = 2, both in the ground state and at finite temperature. Re-
θ = 0.5 and θ = 1.0. From Fig. 19 we see that the ESA markably, Ginvert (q) almost exactly follows GESA (q) for
and the neural-net representation based results for the all wave numbers, which further substantiates the qual-
stopping power are in agreement with a high accuracy. ity of the relatively simple idea behind the ESA. As it is
Additionally, a comparison to the RPA based data shows expected, the latter gives very accurate results both for
that electronic exchange-correlation effects are significant S(q) and the interaction energy v, in particular at metal-
at projectile velocities v . vF . We refer an interested lic densities where we find relative deviations to PIMC
19

data not exceeding 1%. adiabatic connection and the fluctuation dissipation the-
A further point of interest is the utility of the ESA orem [79–82]. This is a highly desirable project, as the
regarding the estimation of the static density response predictive capability of DFT for WDM calculations is
function χ(q) and the directly related dielectric func- still limited [118].
tion (q). More specifically, the neural-net representa- Secondly, we mention the interpretation of XRTS
tion of the exact static LFC G(q; rs , θ) should give ex- experiments [77, 78] within the Chihara decomposi-
act result for this quantities, whereas the definition of tion [133] where electronic correlations are often treated
GESA (q; rs , θ) as a frequency-averaged LFC could poten- insufficiently. In this regard, the remarkable degree of
tially introduce a bias in this limit. Yet, we find that the accuracy provided by both ESA and the static approxi-
ESA gives virtually exact results over the entire WDM mation, and the promising results for aluminum shown
regime (even in the ground-state), whereas said bias only in Ref. [100] give us hope that an improved description of
manifests in χ(q) for the strongly coupled electron liq- XRTS signals can be achieved with hardly any additional
uid regime, rs = 20. In addition, the exact incorpora- effort.
tion of the CSR for small q in our parametrization of Thirdly, the ESA can be used to incorporate electronic
GESA (q; rs , θ) means that the present results for the di- XC-effects into many effective theories in a straightfor-
electric function (q) are even superior to the correspond- ward way. Here examples include quantum hydrodynam-
ing prediction by the neural net, where the CSR is only ics [90, 91, 134], average atom models [92], electronically
fulfilled approximately, i.e., with finite accuracy. In par- screened ionic potentials [127, 135, 136], and dynamic
ticular, the ESA gives the correct divergence behaviour electronic phase-field crystal methods [137].
of (q) in the limit of small q, whereas the neural-net pre- Finally, we mention the value of the LFC in general
dicts a finite value for q = 0, which is unphysical [1, 89]. and the ESA in particular for the estimation of a mul-
A third item of our analysis is the application of the titude of material properties like the electronic stopping
ESA for the estimation of the dynamic structure fac- power [85], thermal and electrical conductivities [89], and
tor S(q, ω), where we find no difference to the usual energy relaxation rates [138–140].
static approximation [89, 95, 96]. More specifically, From a theoretical perspective, the main open chal-
both G(q; rs , θ) and GESA (q; rs , θ) are highly accurate lenge is given by the estimation of the full frequency-
at WDM densities, but cannot reproduce the nontrivial dependence of the LFC G(q, ω), which is currently only
shape of S(q, ω) associated with the predicted incipient possible for certain parameters [89, 95, 96]. One way
excitonic mode [26, 132] in the electron liquid regime. towards this goal would be the development of new
Furthermore, we have compared our parametrization fermionic QMC approaches at finite temperature, to es-
of GESA (q; rs , θ) and the neural-net representation of timate the imaginary-time density–density correlation
G(q; rs , θ) regarding the construction of an electronically function F (q, τ )–the crucial ingredient for the reconstruc-
screened ionic potential Φ(r). While the resulting poten- tion of both S(q, ω) and G(q, ω). Here the phaseless
tials are in excellent agreement for small to intermediate auxiliary-field QMC method constitutes a promising can-
distances r, the aforementioned inaccuracies of the neu- didate [66].
ral net at small q lead to a spuriously slow convergence A second topic for future research is given by the
of Φ(r) at large ionic separations r. comparison of GESA (q; rs , θ) to different dielectric the-
Finally, the stopping power calculation results show ories [105, 108, 124, 141–143], in particular the recent
that the ESA and the neural-net representation of the scheme by Tanaka [141] and the frequency-dependent
LFC are equivalent for this application. Therefore, both version of STLS [144–146].
the presented analytical fit formula for the ESA and
the neural-net representation of the LFC can be used
ACKNOWLEDGMENTS
to study ion energy-loss in WDM and hot dense matter.

We thank Jan Vorberger for helpful comments. This


work was partly funded by the Center for Advanced
B. Discussion and outlook
Systems Understanding (CASUS) which is financed by
Germany’s Federal Ministry of Education and Research
The ESA scheme has been shown to give a highly reli- (BMBF) and by the Saxon Ministry for Science, Culture
able description of electronic XC-effects and, in our opin- and Tourism (SMWK) with tax funds on the basis of
ion, constitutes the method of choice for many applica- the budget approved by the Saxon State Parliament. We
tions both in the context of WDM research and solid gratefully acknowledge CPU-time at the Norddeutscher
state physics in the ground state. Verbund für Hoch- und Höchstleistungsrechnen (HLRN)
Due to its definition as a frequency-averaged LFC, the under grant shp00026 and on a Bull Cluster at the Center
ESA is particularly suited for the construction of ad- for Information Services and High Performace Comput-
vanced XC-functionals for DFT simulations based on the ing (ZIH) at Technische Universität Dresden.
20

[1] G. Giuliani and G. Vignale, Quantum Theory of the size effects and coulomb interactions in quantum monte
Electron Liquid (Cambridge University Press, Cam- carlo calculations for homogeneous systems with peri-
bridge, 2008). odic boundary conditions,” Phys. Rev. B 53, 1814–1832
[2] W. M. C. Foulkes, L. Mitas, R. J. Needs, and G. Ra- (1996).
jagopal, “Quantum monte carlo simulations of solids,” [20] J. P. Perdew and Alex Zunger, “Self-interaction cor-
Rev. Mod. Phys. 73, 33–83 (2001). rection to density-functional approximations for many-
[3] P.-F. Loos and P. M. W. Gill, “The uniform electron electron systems,” Phys. Rev. B 23, 5048–5079 (1981).
gas,” Comput. Mol. Sci 6, 410–429 (2016). [21] John P. Perdew and Yue Wang, “Accurate and simple
[4] T. Dornheim, S. Groth, and M. Bonitz, “The uniform analytic representation of the electron-gas correlation
electron gas at warm dense matter conditions,” Phys. energy,” Phys. Rev. B 45, 13244–13249 (1992).
Reports 744, 1–86 (2018). [22] John P. Perdew and Yue Wang, “Pair-distribution func-
[5] D. Bohm and A D. Pines, “Collective description of elec- tion and its coupling-constant average for the spin-
tron interactions: Ii. collective vs individual particle as- polarized electron gas,” Phys. Rev. B 46, 12947–12954
pects of the interactions,” Phys. Rev. 85, 338 (1952). (1992).
[6] J. Bardeen, L. N. Cooper, and J. R. Schrieffer, “The- [23] S. H. Vosko, L. Wilk, and M. Nusair, “Accurate
ory of superconductivity,” Phys. Rev. 108, 1175–1204 spin-dependent electron liquid correlation energies for
(1957). local spin density calculations: a critical analysis,”
[7] D. Ceperley, “Ground state of the fermion one- Canadian Journal of Physics 58, 1200–1211 (1980),
component plasma: A monte carlo study in two and https://doi.org/10.1139/p80-159.
three dimensions,” Phys. Rev. B 18, 3126–3138 (1978). [24] Paola Gori-Giorgi, Francesco Sacchetti, and Gio-
[8] D. M. Ceperley and B. J. Alder, “Ground state of the vanni B. Bachelet, “Analytic static structure factors and
electron gas by a stochastic method,” Phys. Rev. Lett. pair-correlation functions for the unpolarized homoge-
45, 566–569 (1980). neous electron gas,” Phys. Rev. B 61, 7353–7363 (2000).
[9] C. Bowen, G. Sugiyama, and B. J. Alder, “Static di- [25] M. Corradini, R. Del Sole, G. Onida, and M. Palummo,
electric response of the electron gas,” Phys. Rev. B 50, “Analytical expressions for the local-field factor g(q)
14838 (1994). and the exchange-correlation kernel Kxc (r) of the homo-
[10] S. Moroni, D. M. Ceperley, and G. Senatore, “Static re- geneous electron gas,” Phys. Rev. B 57, 14569 (1998).
sponse from quantum Monte Carlo calculations,” Phys. [26] Yasutami Takada, “Emergence of an excitonic collec-
Rev. Lett 69, 1837 (1992). tive mode in the dilute electron gas,” Phys. Rev. B 94,
[11] S. Moroni, D. M. Ceperley, and G. Senatore, “Static 245106 (2016).
response and local field factor of the electron gas,” Phys. [27] John P. Perdew, Kieron Burke, and Matthias Ernz-
Rev. Lett 75, 689 (1995). erhof, “Generalized gradient approximation made sim-
[12] G. Ortiz and P. Ballone, “Correlation energy, structure ple,” Phys. Rev. Lett. 77, 3865–3868 (1996).
factor, radial distribution function, and momentum dis- [28] Kieron Burke, “Perspective on density functional the-
tribution of the spin-polarized uniform electron gas,” ory,” The Journal of Chemical Physics 136, 150901
Phys. Rev. B 50, 1391–1405 (1994). (2012), https://doi.org/10.1063/1.4704546.
[13] G. Ortiz, M. Harris, and P. Ballone, “Zero temperature [29] R. O. Jones, “Density functional theory: Its origins, rise
phases of the electron gas,” Phys. Rev. Lett. 82, 5317– to prominence, and future,” Rev. Mod. Phys. 87, 897–
5320 (1999). 923 (2015).
[14] F. H. Zong, C. Lin, and D. M. Ceperley, “Spin polariza- [30] D. Saumon, W. B. Hubbard, G. Chabrier, and H. M.
tion of the low-density three-dimensional electron gas,” van Horn, “The role of the molecular-metallic transi-
Phys. Rev. E 66, 036703 (2002). tion of hydrogen in the evolution of jupiter, saturn, and
[15] James J. Shepherd, George H. Booth, and Ali Alavi, brown dwarfs,” Astrophys. J 391, 827–831 (1992).
“Investigation of the full configuration interaction quan- [31] B. Militzer, W. B. Hubbard, J. Vorberger, I. Tam-
tum monte carlo method using homogeneous electron blyn, and S. A. Bonev, “A massive core in jupiter
gas models,” The Journal of Chemical Physics 136, predicted from first-principles simulations,” The Astro-
244101 (2012), https://doi.org/10.1063/1.4720076. physical Journal 688, L45–L48 (2008).
[16] James J. Shepherd, George Booth, Andreas Grüneis, [32] T. Guillot, Y. Miguel, B. Militzer, W. B. Hubbard,
and Ali Alavi, “Full configuration interaction perspec- Y. Kaspi, E. Galanti, H. Cao, R. Helled, S. M. Wahl,
tive on the homogeneous electron gas,” Phys. Rev. B L. Iess, W. M. Folkner, D. J. Stevenson, J. I. Lunine,
85, 081103 (2012). D. R. Reese, A. Biekman, M. Parisi, D. Durante, J. E. P.
[17] G. G. Spink, R. J. Needs, and N. D. Drummond, Connerney, S. M. Levin, and S. J. Bolton, “A suppres-
“Quantum monte carlo study of the three-dimensional sion of differential rotation in jupiter’s deep interior,”
spin-polarized homogeneous electron gas,” Phys. Rev. Nature 555, 227–230 (2018).
B 88, 085121 (2013). [33] A. Becker, W. Lorenzen, J. J. Fortney, N. Nettelmann,
[18] N. D. Drummond, Z. Radnai, J. R. Trail, M. D. Towler, M. Schöttler, and R. Redmer, “Ab initio equations of
and R. J. Needs, “Diffusion quantum monte carlo study state for hydrogen (h-reos.3) and helium (he-reos.3) and
of three-dimensional wigner crystals,” Phys. Rev. B 69, their implications for the interior of brown dwarfs,” As-
085116 (2004). trophys. J. Suppl. Ser 215, 21 (2014).
[19] Louisa M. Fraser, W. M. C. Foulkes, G. Rajagopal, R. J. [34] J. Daligault and S. Gupta, “Electron-ion scattering
Needs, S. D. Kenny, and A. J. Williamson, “Finite- in dense multi-component plasmas: application to the
21

outer crust of an accreting star,” The Astrophysical [47] V. V. Karasiev, L. Calderin, and S. B. Trickey, “Im-
Journal 703, 994–1011 (2009). portance of finite-temperature exchange correlation for
[35] S. X. Hu, B. Militzer, V. N. Goncharov, and S. Skupsky, warm dense matter calculations,” Phys. Rev. E 93,
“First-principles equation-of-state table of deuterium 063207 (2016).
for inertial confinement fusion applications,” Phys. Rev. [48] M. W. C. Dharma-wardana, “Current issues in finite-
B 84, 224109 (2011). t density-functional theory and warm-correlated mat-
[36] Mark L. Brongersma, Naomi J. Halas, and Peter Nord- ter †,” Computation 4 (2016), 10.3390/computa-
lander, “Plasmon-induced hot carrier science and tech- tion4020016.
nology,” Nature Nanotechnology 10, 25–34 (2015). [49] Travis Sjostrom and Jérôme Daligault, “Gradient cor-
[37] Shaunak Mukherjee, Florian Libisch, Nicolas Large, rections to the exchange-correlation free energy,” Phys.
Oara Neumann, Lisa V. Brown, Jin Cheng, J. Britt Las- Rev. B 90, 155109 (2014).
siter, Emily A. Carter, Peter Nordlander, and Naomi J. [50] K. P. Driver and B. Militzer, “All-electron path integral
Halas, “Hot electrons do the impossible: Plasmon- monte carlo simulations of warm dense matter: Appli-
induced dissociation of h2 on au,” Nano Letters 13, cation to water and carbon plasmas,” Phys. Rev. Lett.
240–247 (2013). 108, 115502 (2012).
[38] K. Falk, “Experimental methods for warm dense matter [51] N. S. Blunt, T. W. Rogers, J. S. Spencer, and
research,” High Power Laser Sci. Eng 6, e59 (2018). W. M. C. Foulkes, “Density-matrix quantum monte
[39] D. Kraus, A. Ravasio, M. Gauthier, D. O. Gericke, carlo method,” Phys. Rev. B 89, 245124 (2014).
J. Vorberger, S. Frydrych, J. Helfrich, L. B. Fletcher, [52] Tobias Dornheim, Simon Groth, Fionn D. Malone, Tim
G. Schaumann, B. Nagler, B. Barbrel, B. Bachmann, Schoof, Travis Sjostrom, W. M. C. Foulkes, and Michael
E. J. Gamboa, S. Göde, E. Granados, G. Gregori, H. J. Bonitz, “Ab initio quantum monte carlo simulation of
Lee, P. Neumayer, W. Schumaker, T. Döppner, R. W. the warm dense electron gas,” Physics of Plasmas 24,
Falcone, S. H. Glenzer, and M. Roth, “Nanosecond 056303 (2017), https://doi.org/10.1063/1.4977920.
formation of diamond and lonsdaleite by shock com- [53] Ethan W. Brown, Bryan K. Clark, Jonathan L. DuBois,
pression of graphite,” Nature Communications 7, 10970 and David M. Ceperley, “Path-integral monte carlo sim-
(2016). ulation of the warm dense homogeneous electron gas,”
[40] D. Kraus, J. Vorberger, A. Pak, N. J. Hartley, L. B. Phys. Rev. Lett. 110, 146405 (2013).
Fletcher, S. Frydrych, E. Galtier, E. J. Gamboa, D. O. [54] Tobias Dornheim, Simon Groth, Alexey Filinov, and
Gericke, S. H. Glenzer, E. Granados, M. J. MacDonald, Michael Bonitz, “Permutation blocking path integral
A. J. MacKinnon, E. E. McBride, I. Nam, P. Neumayer, monte carlo: a highly efficient approach to the simu-
M. Roth, A. M. Saunders, A. K. Schuster, P. Sun, T. van lation of strongly degenerate non-ideal fermions,” New
Driel, T. Döppner, and R. W. Falcone, “Formation of Journal of Physics 17, 073017 (2015).
diamonds in laser-compressed hydrocarbons at plane- [55] T. Schoof, S. Groth, J. Vorberger, and M. Bonitz, “Ab
tary interior conditions,” Nature Astronomy 1, 606–611 initio thermodynamic results for the degenerate electron
(2017). gas at finite temperature,” Phys. Rev. Lett. 115, 130402
[41] P. Sperling, E. J. Gamboa, H. J. Lee, H. K. Chung, (2015).
E. Galtier, Y. Omarbakiyeva, H. Reinholz, G. Röpke, [56] Fionn D. Malone, N. S. Blunt, James J. Shepherd,
U. Zastrau, J. Hastings, L. B. Fletcher, and S. H. D. K. K. Lee, J. S. Spencer, and W. M. C. Foulkes, “In-
Glenzer, “Free-electron x-ray laser measurements of teraction picture density matrix quantum monte carlo,”
collisional-damped plasmons in isochorically heated The Journal of Chemical Physics 143, 044116 (2015),
warm dense matter,” Phys. Rev. Lett. 115, 115001 https://doi.org/10.1063/1.4927434.
(2015). [57] Burkhard Militzer and Kevin P. Driver, “Development
[42] F. Graziani, M. P. Desjarlais, R. Redmer, and S. B. of path integral monte carlo simulations with localized
Trickey, eds., Frontiers and Challenges in Warm Dense nodal surfaces for second-row elements,” Phys. Rev.
Matter (Springer, International Publishing, 2014). Lett. 115, 176403 (2015).
[43] M. Bonitz, T. Dornheim, Zh. A. Moldabekov, S. Zhang, [58] Fionn D. Malone, N. S. Blunt, Ethan W. Brown,
P. Hamann, H. Kählert, A. Filinov, K. Ramakr- D. K. K. Lee, J. S. Spencer, W. M. C. Foulkes, and
ishna, and J. Vorberger, “Ab initio simulation of warm James J. Shepherd, “Accurate exchange-correlation en-
dense matter,” Physics of Plasmas 27, 042710 (2020), ergies for the warm dense electron gas,” Phys. Rev. Lett.
https://doi.org/10.1063/1.5143225. 117, 115701 (2016).
[44] Torben Ott, Hauke Thomsen, Jan Willem Abra- [59] T. Dornheim, S. Groth, T. Sjostrom, F. D. Malone,
ham, Tobias Dornheim, and Michael Bonitz, “Recent W. M. C. Foulkes, and M. Bonitz, “Ab initio quan-
progress in the theory and simulation of strongly cor- tum Monte Carlo simulation of the warm dense electron
related plasmas: phase transitions, transport, quan- gas in the thermodynamic limit,” Phys. Rev. Lett. 117,
tum, and magnetic field effects,” The European Physical 156403 (2016).
Journal D 72, 84 (2018). [60] T. Dornheim, S. Groth, and M. Bonitz, “Ab ini-
[45] N. David Mermin, “Thermal properties of the inhomo- tio results for the static structure factor of the warm
geneous electron gas,” Phys. Rev. 137, A1441–A1443 dense electron gas,” Contrib. Plasma Phys 57, 468–478
(1965). (2017).
[46] Kushal Ramakrishna, Tobias Dornheim, and Jan [61] S. Groth, T. Dornheim, and M. Bonitz, “Configuration
Vorberger, “Influence of finite temperature exchange- path integral Monte Carlo approach to the static density
correlation effects in hydrogen,” Phys. Rev. B 101, response of the warm dense electron gas,” J. Chem. Phys
195129 (2020). 147, 164108 (2017).
22

[62] T. Dornheim, S. Groth, J. Vorberger, and M. Bonitz, 153, 034112 (2020), https://doi.org/10.1063/5.0016783.
“Permutation blocking path integral Monte Carlo ap- [76] Valentin V. Karasiev, James W. Dufty, and S. B.
proach to the static density response of the warm dense Trickey, “Nonempirical semilocal free-energy density
electron gas,” Phys. Rev. E 96, 023203 (2017). functional for matter under extreme conditions,” Phys.
[63] K. P. Driver, F. Soubiran, and B. Militzer, “Path inte- Rev. Lett. 120, 076401 (2018).
gral monte carlo simulations of warm dense aluminum,” [77] S. H. Glenzer and R. Redmer, “X-ray thomson scatter-
Phys. Rev. E 97, 063207 (2018). ing in high energy density plasmas,” Rev. Mod. Phys
[64] Tobias Dornheim, Jan Vorberger, and Michael Bonitz, 81, 1625 (2009).
“Nonlinear electronic density response in warm dense [78] D. Kraus, B. Bachmann, B. Barbrel, R. W. Falcone,
matter,” Phys. Rev. Lett. 125, 085001 (2020). L. B. Fletcher, S. Frydrych, E. J. Gamboa, M. Gauthier,
[65] Tobias Dornheim, Michele Invernizzi, Jan Vorberger, D. O. Gericke, S. H. Glenzer, S. Göde, E. Granados,
and Barak Hirshberg, “Attenuating the fermion sign N. J. Hartley, J. Helfrich, H. J. Lee, B. Nagler, A. Rava-
problem in path integral monte carlo simulations us- sio, W. Schumaker, J. Vorberger, and T. Döppner,
ing the bogoliubov inequality and thermodynamic inte- “Characterizing the ionization potential depression in
gration,” The Journal of Chemical Physics 153, 234104 dense carbon plasmas with high-precision spectrally re-
(2020), https://doi.org/10.1063/5.0030760. solved x-ray scattering,” Plasma Phys. Control Fusion
[66] Joonho Lee, Miguel A. Morales, and Fionn D. Mal- 61, 014015 (2019).
one, “A phaseless auxiliary-field quantum monte carlo [79] Deyu Lu, “Evaluation of model exchange-correlation
perspective on the uniform electron gas at finite tem- kernels in the adiabatic connection fluctuation-
peratures: Issues, observations, and benchmark study,” dissipation theorem for inhomogeneous systems,” The
(2020), arXiv:2012.12228 [physics.chem-ph]. Journal of Chemical Physics 140, 18A520 (2014),
[67] Yuan Liu, Minsik Cho, and Brenda Rubenstein, “Ab https://doi.org/10.1063/1.4867538.
initio finite temperature auxiliary field quantum monte [80] Christopher E. Patrick and Kristian S. Thyge-
carlo,” Journal of Chemical Theory and Computation sen, “Adiabatic-connection fluctuation-dissipation
14, 4722–4732 (2018). dft for the structural properties of solids—the
[68] A. Yilmaz, K. Hunger, T. Dornheim, S. Groth, and renormalized alda and electron gas kernels,” The
M. Bonitz, “Restricted configuration path integral Journal of Chemical Physics 143, 102802 (2015),
monte carlo,” The Journal of Chemical Physics 153, https://doi.org/10.1063/1.4919236.
124114 (2020), https://doi.org/10.1063/5.0022800. [81] Andreas Görling, “Hierarchies of methods towards
[69] S. Groth, T. Dornheim, T. Sjostrom, F. D. Malone, the exact kohn-sham correlation energy based on the
W. M. C. Foulkes, and M. Bonitz, “Ab initio exchange– adiabatic-connection fluctuation-dissipation theorem,”
correlation free energy of the uniform electron gas at Phys. Rev. B 99, 235120 (2019).
warm dense matter conditions,” Phys. Rev. Lett. 119, [82] A. Pribram-Jones, P. E. Grabowski, and K. Burke,
135001 (2017). “Thermal density functional theory: Time-dependent
[70] Valentin V. Karasiev, Travis Sjostrom, James Dufty, linear response and approximate functionals from the
and S. B. Trickey, “Accurate homogeneous electron gas fluctuation-dissipation theorem,” Phys. Rev. Lett 116,
exchange-correlation free energy for local spin-density 233001 (2016).
calculations,” Phys. Rev. Lett. 112, 076403 (2014). [83] E. K. U. Gross and W. Kohn, “Local density-functional
[71] T. G. White, S. Richardson, B. J. B. Crowley, L. K. Pat- theory of frequency-dependent linear response,” Phys.
tison, J. W. O. Harris, and G. Gregori, “Orbital-free Rev. Lett 55, 2850 (1985).
density-functional theory simulations of the dynamic [84] A. D. Baczewski, L. Shulenburger, M. P. Desjarlais,
structure factor of warm dense aluminum,” Phys. Rev. S. B. Hansen, and R. J. Magyar, “X-ray thomson scat-
Lett. 111, 175002 (2013). tering in warm dense matter without the chihara de-
[72] Chang Gao, Shen Zhang, Wei Kang, Cong Wang, Ping composition,” Phys. Rev. Lett 116, 115004 (2016).
Zhang, and X. T. He, “Validity boundary of orbital-free [85] Zh. A. Moldabekov, T. Dornheim, M. Bonitz, and T. S.
molecular dynamics method corresponding to thermal Ramazanov, “Ion energy-loss characteristics and friction
ionization of shell structure,” Phys. Rev. B 94, 205115 in a free-electron gas at warm dense matter and nonideal
(2016). dense plasma conditions,” Phys. Rev. E 101, 053203
[73] Shen Zhang, Hongwei Wang, Wei Kang, Ping Zhang, (2020).
and X. T. He, “Extended application of kohn-sham first- [86] G. Senatore, S. Moroni, and D.M. Ceperley, “Local field
principles molecular dynamics method with plane wave factor and effective potentials in liquid metals,” Journal
approximation at high energy—from cold materials to of Non-Crystalline Solids 205-207, 851 – 854 (1996).
hot dense plasmas,” Physics of Plasmas 23, 042707 [87] Zh.A. Moldabekov, S. Groth, T. Dornheim, M. Bonitz,
(2016), https://doi.org/10.1063/1.4947212. and T.S. Ramazanov, “Ion potential in non-ideal dense
[74] Y. H. Ding, A. J. White, S. X. Hu, O. Certik, and quantum plasmas,” Contributions to Plasma Physics
L. A. Collins, “Ab initio studies on the stopping power 57, 532–538 (2017).
of warm dense matter with time-dependent orbital-free [88] Zh.A. Moldabekov, S. Groth, T. Dornheim, H. Kählert,
density functional theory,” Phys. Rev. Lett. 121, 145001 M. Bonitz, and T. S. Ramazanov, “Structural charac-
(2018). teristics of strongly coupled ions in a dense quantum
[75] Abhiraj Sharma, Sebastien Hamel, Mandy Bethken- plasma,” Phys. Rev. E 98, 023207 (2018).
hagen, John E. Pask, and Phanish Suryanarayana, [89] Paul Hamann, Tobias Dornheim, Jan Vorberger, Zhan-
“Real-space formulation of the stress tensor for o(n) dos A. Moldabekov, and Michael Bonitz, “Dynamic
density functional theory: Application to high temper- properties of the warm dense electron gas based on
ature calculations,” The Journal of Chemical Physics abinitio path integral monte carlo simulations,” Phys.
23

Rev. B 102, 125150 (2020). [106] A. Holas, “Exact asymptotic expression for the static
[90] Abdourahmane Diaw and Michael S. Murillo, “A vis- dielectric function of a uniform electron liquid at large
cous quantum hydrodynamics model based on dynamic wave vector,” in Strongly Coupled Plasma Physics,
density functional theory,” Scientific Reports 7, 15352 edited by F.J. Rogers and H.E. DeWitt (Plenum, New
(2017). York, 1987).
[91] Zh. A. Moldabekov, M. Bonitz, and T. S. Ramazanov, [107] B. Farid, V. Heine, G. E. Engel, and I. J. Robertson,
“Theoretical foundations of quantum hydrodynamics “Extremal properties of the harris-foulkes functional
for plasmas,” Physics of Plasmas 25, 031903 (2018). and an improved screening calculation for the electron
[92] P. A. Sterne, S. B. Hansen, B. G. Wilson, and W. A. gas,” Phys. Rev. B 48, 11602 (1993).
Isaacs, “Equation of state, occupation probabilities and [108] T. Sjostrom and J. Dufty, “Uniform electron gas at finite
conductivities in the average atom purgatorio code,” temperatures,” Phys. Rev. B 88, 115123 (2013).
High Energy Density Physics 3, 278–282 (2007). [109] B. Militzer and E. L. Pollock, “Lowering of the kinetic
[93] D. M. Ceperley, “Path integrals in the theory of con- energy in interacting quantum systems,” Phys. Rev.
densed helium,” Rev. Mod. Phys 67, 279 (1995). Lett. 89, 280401 (2002).
[94] T. Dornheim, J. Vorberger, S. Groth, N. Hoffmann, [110] K. S. Singwi, M. P. Tosi, R. H. Land, and A. Sjölander,
Zh.A. Moldabekov, and M. Bonitz, “The static local “Electron correlations at metallic densities,” Phys. Rev
field correction of the warm dense electron gas: An 176, 589 (1968).
ab initio path integral Monte Carlo study and machine [111] Dvira Segal, Andrew J. Millis, and David R. Reichman,
learning representation,” J. Chem. Phys 151, 194104 “Numerically exact path-integral simulation of nonequi-
(2019). librium quantum transport and dissipation,” Phys. Rev.
[95] S. Groth, T. Dornheim, and J. Vorberger, “Ab initio B 82, 205323 (2010).
path integral Monte Carlo approach to the static and [112] T. Dornheim, “Fermion sign problem in path integral
dynamic density response of the uniform electron gas,” Monte Carlo simulations: Quantum dots, ultracold
Phys. Rev. B 99, 235122 (2019). atoms, and warm dense matter,” Phys. Rev. E 100,
[96] T. Dornheim, S. Groth, J. Vorberger, and M. Bonitz, 023307 (2019).
“Ab initio path integral Monte Carlo results for the dy- [113] A. A. Kugler, “Bounds for some equilibrium properties
namic structure factor of correlated electrons: From the of an electron gas,” Phys. Rev. A 1, 1688 (1970).
electron liquid to warm dense matter,” Phys. Rev. Lett. [114] S. Mazevet, M. P. Desjarlais, L. A. Collins, J. D. Kress,
121, 255001 (2018). and N. H. Magee, “Simulations of the optical properties
[97] Tobias Dornheim and Jan Vorberger, “Finite-size effects of warm dense aluminum,” Phys. Rev. E 71, 016409
in the reconstruction of dynamic properties from ab ini- (2005).
tio path integral monte carlo simulations,” Phys. Rev. [115] U. Zastrau, P. Sperling, M. Harmand, A. Becker,
E 102, 063301 (2020). T. Bornath, R. Bredow, S. Dziarzhytski, T. Fennel,
[98] Paul Hamann, Jan Vorberger, Tobias Dornheim, Zhan- L. B. Fletcher, E. F”orster, S. G”ode, G. Gregori,
dos A. Moldabekov, and Michael Bonitz, “Ab ini- V. Hilbert, D. Hochhaus, B. Holst, T. Laarmann,
tio results for the plasmon dispersion and damping of H. J. Lee, T. Ma, J. P. Mithen, R. Mitzner, C. D.
the warm dense electron gas,” Contributions to Plasma Murphy, M. Nakatsutsumi, P. Neumayer, A. Przys-
Physics 60, e202000147 (2020). tawik, S. Roling, M. Schulz, B. Siemer, S. Skruszewicz,
[99] Tobias Dornheim, Zhandos A Moldabekov, Jan Vor- J. Tiggesb”aumker, S. Toleikis, T. Tschentscher,
berger, and Simon Groth, “Ab initio path integral T. White, M. W”ostmann, H. Zacharias, T. D”oppner,
monte carlo simulation of the uniform electron gas in S. H. Glenzer, and R. Redmer, “Resolving ultrafast
the high energy density regime,” Plasma Physics and heating of dense cryogenic hydrogen,” Phys. Rev. Lett
Controlled Fusion 62, 075003 (2020). 112, 105002 (2014).
[100] Tobias Dornheim, Attila Cangi, Kushal Ramakrishna, [116] J. F. Benage, W. R. Shanahan, and M. S. Murillo,
Maximilian Böhme, Shigenori Tanaka, and Jan Vor- “Electrical resistivity measurements of hot dense alu-
berger, “Effective static approximation: A fast and re- minum,” Phys. Rev. Lett 83, 2953 (1999).
liable tool for warm-dense matter theory,” Phys. Rev. [117] M. P. Desjarlais, J. D. Kress, and L. A. Collins, “Elec-
Lett. 125, 235001 (2020). trical conductivity for warm, dense aluminum plasmas
[101] Kai Hunger, Tim Schoof, Tobias Dornheim, Michael and liquids,” Phys. Rev. E 66, 025401(R) (2002).
Bonitz, and Alexey Filinov, “Momentum distribution [118] Kushal Ramakrishna, Attila Cangi, Tobias Dornheim,
function and short-range correlations of the warm dense and Jan Vorberger, “First-principles modeling of plas-
electron gas – ab initio quantum monte carlo results,” mons in aluminum under ambient and extreme condi-
(2021), arXiv:2101.00842 [physics.plasm-ph]. tions,” (2020), arXiv:2009.12163 [cond-mat.mtrl-sci].
[102] A link to the repository will be given upon publication. [119] Setsuo Ichimaru, Hiroshi Iyetomi, and Shigenori
[103] A. A. Kugler, “Theory of the local field correction in an Tanaka, “Statistical physics of dense plasmas: Ther-
electron gas,” J. Stat. Phys 12, 35 (1975). modynamics, transport coefficients and dynamic corre-
[104] Tobias Dornheim, Travis Sjostrom, Shigenori Tanaka, lations,” Physics Reports 149, 91 – 205 (1987).
and Jan Vorberger, “Strongly coupled electron liquid: [120] James P. Mithen, Jérôme Daligault, and Gianluca
Ab initio path integral monte carlo simulations and di- Gregori, “Comparative merits of the memory function
electric theories,” Phys. Rev. B 101, 045129 (2020). and dynamic local-field correction of the classical one-
[105] S. Tanaka and S. Ichimaru, “Thermodynamics and component plasma,” Phys. Rev. E 85, 056407 (2012).
correlational properties of finite-temperature electron [121] M. Bonitz, Quantum kinetic theory (Springer, Heidel-
liquids in the Singwi-Tosi-Land-Sjölander approxima- berg, 2016).
tion,” J. Phys. Soc. Jpn 55, 2278–2289 (1986).
24

[122] Alexandrov, Bogdankievich, and Rukhadse, Principles [134] Zhandos Moldabekov, Tim Schoof, Patrick Ludwig,
of Plasma Electrodynamics (Springer, Heidelberg, Ger- Michael Bonitz, and Tlekkabul Ramazanov, “Statically
many, 1984). screened ion potential and bohm potential in a quan-
[123] P. Vashishta and K. S. Singwi, “Electron correlations at tum plasma,” Physics of Plasmas 22, 102104 (2015),
metallic densities v,” Phys. Rev. B 6, 875 (1972). https://doi.org/10.1063/1.4932051.
[124] W. Stolzmann and M. Rösler, “Static local-field cor- [135] Zhandos Moldabekov, Patrick Ludwig, Michael Bonitz,
rected dielectric and thermodynamic functions,” Con- and Tlekkabul Ramazanov, “Ion potential in warm
trib. Plasma Phys 41, 203 (2001). dense matter: Wake effects due to streaming degenerate
[125] Serge Galam and Jean-Pierre Hansen, “Statistical me- electrons,” Phys. Rev. E 91, 023102 (2015).
chanics of dense ionized matter. vi. electron screening [136] Zh. A. Moldabekov, P. Ludwig, M. Bonitz,
corrections to the thermodynamic properties of the one- and T. S. Ramazanov, “Notes on anoma-
component plasma,” Phys. Rev. A 14, 816–832 (1976). lous quantum wake effects,” Contributions
[126] Zh.A. Moldabekov, T. Dornheim, and M. Bonitz, to Plasma Physics 56, 442–447 (2016),
“Screening of a test charge in a free-electron gas at https://onlinelibrary.wiley.com/doi/pdf/10.1002/ctpp.201500137.
warm dense matter and dense non-ideal plasma con- [137] Salvador Valtierra Rodriguez, Nan Wang, Nana Ofori-
ditions,” accepted for publication in Contrib. Plasma Opoku, Nikolas Provatas, and Kirk H. Bevan, “Captur-
Phys.(arXiv:2009.09180) (2020). ing the dynamics of wigner crystals within the phase-
[127] Zh.A. Moldabekov, S. Groth, T. Dornheim, M. Bonitz, field crystal method,” Phys. Rev. B 100, 235116 (2019).
and T.S. Ramazanov, “Ion potential in non-ideal dense [138] J. Vorberger, D. O. Gericke, Th. Bornath, and
quantum plasmas,” Contrib. Plasma Phys. 57, 532–538 M. Schlanges, “Energy relaxation in dense, strongly
(2017). coupled two-temperature plasmas,” Phys. Rev. E 81,
[128] P.E. Grabowski, S.B. Hansen, M.S. Murillo, L.G. Stan- 046404 (2010).
ton, F.R. Graziani, A.B. Zylstra, S.D. Baalrud, P. Ar- [139] L. X. Benedict, M. P. Surh, L. G. Stanton, C. R.
nault, A.D. Baczewski, L.X. Benedict, C. Blancard, Scullard, A. A. Correa, J. I. Castor, F. R. Graziani,
O. Čertı́k, J. Clérouin, L.A. Collins, S. Copeland, A.A. L. A. Collins, O. Certı́k, J. D. Kress, and M. S.
Correa, J. Dai, J. Daligault, M.P. Desjarlais, M.W.C. Murillo, “Molecular dynamics studies of electron-ion
Dharma-wardana, G. Faussurier, J. Haack, T. Hax- temperature equilibration in hydrogen plasmas within
himali, A. Hayes-Sterbenz, Y. Hou, S.X. Hu, D. Jensen, the coupled-mode regime,” Phys. Rev. E 95, 043202
G. Jungman, G. Kagan, D. Kang, J.D. Kress, Q. Ma, (2017).
M. Marciante, E. Meyer, R.E. Rudd, D. Saumon, [140] Christian R. Scullard, Susana Serna, Lorin X. Benedict,
L. Shulenburger, R.L. Singleton, T. Sjostrom, L.J. C. Leland Ellison, and Frank R. Graziani, “Analytic
Stanek, C.E. Starrett, C. Ticknor, S. Valaitis, J. Ven- expressions for electron-ion temperature equilibration
zke, and A. White, “Review of the first charged-particle rates from the Lenard-Balescu equation,” Phys. Rev.
transport coefficient comparison workshop,” High En- E 97, 013205 (2018).
ergy Density Physics 37, 100905 (2020). [141] S. Tanaka, “Correlational and thermodynamic prop-
[129] S.K. Kodanova, M.K. Issanova, S.M. Amirov, T.S. Ra- erties of finite-temperature electron liquids in the
mazanov, A. Tikhonov, and Zh.A. Moldabekov, “Re- hypernetted-chain approximation,” J. Chem. Phys 145,
laxation of non-isothermal hot dense plasma param- 214104 (2016).
eters,” Matter and Radiation at Extremes 3, 40–49 [142] Shigenori Tanaka, “Improved equation of state for finite-
(2018). temperature spin-polarized electron liquids on the basis
[130] Néstor R. Arista and Werner Brandt, “Energy loss and of singwi–tosi–land–sjölander approximation,” Contri-
straggling of charged particles in plasmas of all degen- butions to Plasma Physics 57, 126–136 (2017).
eracies,” Phys. Rev. A 23, 1898–1905 (1981). [143] Martin Panholzer, Matteo Gatti, and Lucia Reining,
[131] G. Zwicknagel, C. Toepffer, and P.-G. Reinhard, “Stop- “Nonlocal and nonadiabatic effects in the charge-density
ping of heavy ions in plasmas at strong coupling,” response of solids: A time-dependent density-functional
Physics Reports 309, 117 – 208 (1999). approach,” Phys. Rev. Lett. 120, 166402 (2018).
[132] Masahiko Higuchi and Hiroshi Yasuhara, “Klein- [144] P. Arora, K. Kumar, and R. K. Moudgil, “Spin-resolved
man’s dielectric function and interband optical ab- correlations in the warm-dense homogeneous electron
sorption strength of simple metals,” Journal of the gas,” Eur. Phys. J. B 90, 76 (2017).
Physical Society of Japan 69, 2099–2106 (2000), [145] H. K. Schweng and H. M. Böhm, “Finite-temperature
https://doi.org/10.1143/JPSJ.69.2099. electron correlations in the framework of a dynamic
[133] J Chihara, “Difference in x-ray scattering between local-field correction,” Phys. Rev. B 48, 2037 (1993).
metallic and non-metallic liquids due to conduction elec- [146] A. Holas and S. Rahman, “Dynamic local-field factor
trons,” Journal of Physics F: Metal Physics 17, 295–304 of an electron liquid in the quantum versions of the
(1987). Singwi-Tosi-Land-Sjölander and Vashishta-Singwi the-
ories,” Phys. Rev. B 35, 2720 (1987).

You might also like