ArticlePDF Available

A viscous quantum hydrodynamics model based on dynamic density functional theory

Authors:

Abstract and Figures

Dynamic density functional theory (DDFT) is emerging as a useful theoretical technique for modeling the dynamics of correlated systems. We extend DDFT to quantum systems for application to dense plasmas through a quantum hydrodynamics (QHD) approach. The DDFT-based QHD approach includes correlations in the the equation of state self-consistently, satis es sum rules and includes irreversibility arising from collisions. While QHD can be used generally to model non-equilibrium, heterogeneous plasmas, we employ the DDFT-QHD framework to generate a model for the electronic dynamic structure factor, which o ers an avenue for measuring hydrodynamic properties, such as transport coe cients via x-ray Thomson scattering.
This content is subject to copyright. Terms and conditions apply.
1
SCiEntifiC REPORTS | 7: 15352 | DOI:10.1038/s41598-017-14414-9
www.nature.com/scientificreports
A viscous quantum hydrodynamics
model based on dynamic density
functional theory
Abdourahmane Diaw & Michael S. Murillo
Dynamic density functional theory (DDFT) is emerging as a useful theoretical technique for modeling
the dynamics of correlated systems. We extend DDFT to quantum systems for application to dense
plasmas through a quantum hydrodynamics (QHD) approach. The DDFT-based QHD approach includes
correlations in the the equation of state self-consistently, satises sum rules and includes irreversibility
arising from collisions. While QHD can be used generally to model non-equilibrium, heterogeneous
plasmas, we employ the DDFT-QHD framework to generate a model for the electronic dynamic
structure factor, which oers an avenue for measuring hydrodynamic properties, such as transport
coecients via x-ray Thomson scattering.
Access to high-power laser sources, such as the Linac Coherent Light Source (LCLS)1, National Ignition Facility
(NIF)2 and Omega Laser3, has opened the path to investigating essential properties of non-ideal plasmas such
as ionization potential depression4, transport coecients5 and ionization state6. Understanding the dynamical
properties of non-ideal plasmas is critical for modeling and designing high energy-density science experiments,
including inertial-connement fusion7, cluster explosions8, laser-produced ion beams9, hypervelocity impacts10,
in nanotechnology11,12 and astrophysics13.
Among all the approaches to modeling heterogeneous, non-equilibrium quantum systems, quantum hydro-
dynamics (QHD) is a computationally attractive approach with rich history in statistical mechanics. Shortly aer
the development of quantum mechanics, Bloch14 proposed the first QHD model by simply choosing the
omas-Fermi pressure for the electrons in an otherwise classical hydrodynamics model. In 1964, Hohenberg
and Kohn15 developed ground-state density functional theory for the inhomogeneous electron gas, which was
immediately generalized to nite temperature by Mermin16. Combining the ideas of Bloch with DFT, Ying17 pro-
posed a new quantum hydrodynamic model via an adiabatic generalization of the density functionals. In Ying’s
model, the pressure is represented by P[n(r,t)] with n(r,t) a time-dependent density described by the continuity
equation. Ying’s QHD model includes explicitly all correlation and exchange eects included in the chosen energy
functional. Using an alternate approach, Gasser and Jüngel18 derived QHD equations using the Schrodinger equa-
tion with Wentzel-Kramers-Brillouin (WKB) wave functions. is approach yields the classical momentum equa-
tion with the Bohm potential but it does not account for correlations. Correlations effects and quantum
degeneracy can be included in an ad hoc manner in this model by replacing the Bohmian potential with quantum
potentials12 or self-consistently through orbital-free density functional theory (OF-DFT)19,20. In yet another
approach, using the moment expansion of the Wigner-Boltzmann equation, Gardner21 proposed a QHD model
for semiconductor devices that extends the classical hydrodynamic model to include
O()
2
quantum corrections.
Similar results were obtained with the Wigner-Poisson system by Manfredi and Haas22 for a quantum electron
gas. Following Levermore23, Degond and Ringhofer24 used a non-commutative version of the entropy externali-
zation principle to build a QHD model starting from the quantum Liouville equation. e moment equations are
closed by a quantum Wigner distribution function that minimizes the entropy.
Despite these important advances, describing collisional processes in moderately coupled quantum plasmas
remains a challenge10,22,25. Here, we explore an alternate approach based on a new formulation of quantum hydro-
dynamics (QHD). QHD approaches have the advantage of including equation-of-state and transport quantities
more naturally than response-function approaches. Apart from these potential modeling advantages, QHD mod-
els of DSF therefore also provide access to experimental measurements of these quantities, thereby extending the
utility of DSF. We develop a QHD framework based on the extension of the classical dynamical density functional
Department of Computational Mathematics, Science and Engineering, Michigan State University East Lansing,
Michigan, 48823, USA. Correspondence and requests for materials should be addressed to A.D. (email: rahmane@
melix.org)
Received: 9 February 2017
Accepted: 11 October 2017
Published: xx xx xxxx
OPEN
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
2
SCiEntifiC REPORTS | 7: 15352 | DOI:10.1038/s41598-017-14414-9
theory (DDFT)2628, a variant of time-dependent density functional theory (TDDFT)29. DDFT provides a set of
hydrodynamics equations by taking the velocity moments of Liouville equation and closes the system using den-
sity functional theory2628. A fundamental assumption of this theory is that the equilibrium energy functional of
the system can be used to guess the correlation energy functional when the system is out of equilibrium. While
DDFT has found wide use in many-body classical systems3032, we extend its use in quantum systems17 to viscous
quantum systems, in general, and to DSF, specically.
We apply the DDFT-QHD model to stationary, homogeneous and isothermal plasmas for which the dynamic
structure factor (DSF) is well dened. While the DSF is of interest in its own right, it is also connected to x-ray
omson scattering (XRTS) experiments; XRTS yields much essential information about plasmas, including den-
sity, temperature and atomic physics information (e.g., ionization state6, ionization potential depression4, etc.).
Results
Dense strongly coupled plasmas are characterized by large collisional contributions and degenerate electrons.
ese features make the DDFT-QHD approach a reliable tool for accurately describing the dynamical properties
of these systems. For simplicity, here, we consider a quantum plasma comprising only electrons with density
distribution n and mass m interacting through a pairwise Coulomb potential
|−′|vrr()
. We use atomic units (i.e.,
πε====e m 41
0
) for the remainder of this work. e hydrodynamic equations for the electrons can be
written generally as
+∇⋅=
n
t
nu() 0,
(1)
+∇⋅=−∇
n
t
n
u
uu
()
() ,
(2)
which are continuity and momentum equations written in terms of a generalized force tensor
. Note that the
continuity equation (1) and the le-hand side of (2) are generic, with the physical properties of the quantum
electron gas entering through terms on the right-hand side of (2). In the DDFT approach17,28,33, it is assumed that
the system is close enough to equilibrium that an adiabatic closure can be chosen for
; that is,
=nu[, ] 
. e
primary assumption of this model is that the system is near equilibrium, a condition well satised in highly colli-
sional plasmas. Further, the equilibrium density is forced to be consistent with the thermodynamic ground state
of the system by choosing the diagonal portion of the tensor to be of the form
, where  is the free
energy of the system. When
n[]
is expressed using orbital-free density-functional theory (OFDFT), that portion
of
is closed. e o-diagonal portion of
can be written in its long-wavelength form to yield a generalized
Navier-Stokes equation of the form
δ
δ
ηξ
η
−∇ ⋅=−∇
+∇⋅∇ +
+
∇∇n
n
n
uu,
[]
3
(),
(3)
where η is the shear viscosity, and ξ is the bulk viscosity; all other symbols have their usual meanings. Provided η
and ξ can be expressed in terms of
nu(, )
, the hydrodynamic equations are closed.
In DDFT, one writes the total free-energy functional as
Ω= +Ω +ΩnTnt nt ntrrr[] [(,)][]( ,) [(,)], (4)
Hxc
where
is the free energy of the noninteracting system,
ntr[](, )
H
is the Hartree free-energy functional,
and
ntr[](, )
xc
is the exchange-correlation (xc) functional. e Hartree term is exactly known and is an explicit
function of space and time.
A key advantage to the DDFT approach to QHD is that all thermodynamic properties are included
self-consistently through the total free energy , for which a wide range of approximations are available3437.
In fact, this approach is very similar to the well-known generalized hydrodynamics, developed by Frenkel38,
that extends the classical Navier-Stokes equation to describe the properties of both solid and liquid bodies.
Furthermore, our DDFT-QHD approach can be connected with other approaches based on Bohmian dynam-
ics. If we set the viscous terms equal to zero in (3) and choose the gradient-corrected Thomas-Fermi (TF)
functional for T[n], one recovers the well-known Bohmian QHD20 form; again, however, the DDFT approach
enforces self-consistency of its form with the other terms in the free energy. e connection between DFT and the
Bohmian potential will be briey shown below.
Density uctuations are not readily available in density-functional theories, and our DDFT-QHD approach
suers from this limitation. However, in equilibrium, the uctuation-dissipation theorem allows us to connect the
linear response of the system to density uctuations. We write the DSF of the electrons as
ω
π
χω
=−
βω
S
m
e
k
k
(, )
1I(,)
1
,
(5)
ee
ee
where β is the inverse electron temperature and
χ ωk(, )
ee
is the susceptibility of the free electrons. A large body of
literature39 focuses on the calculation of the system DSF
ωS k(, )
, with most work based on the Chihara40
decomposition
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
3
SCiEntifiC REPORTS | 7: 15352 | DOI:10.1038/s41598-017-14414-9
ωωωω ωω ω=| +| +′′−′+ .Sk fk qk Sk ZdSk Sk ZS k(, )()()(,) (, )(,) (, )
(6)
ii bs ce fe
2
e
e quantities
fk()
and
qk()
are the Fourier components of the density of bound and free electrons. e rst
term of (6) corresponds to low-frequency electron-density uctuations arising from ion dynamics and is propor-
tional to the ion-ion DSF
ωS k(, )
ii
. e factor
ωS k(, )
ce
in the second term describes the contribution from core
electrons41 and is modulated by the ion self-motion
ωS k(, )
s
. e third term is the free-electron DSF
ωS k(, )
ee
in
the presence of a uniform ionic background. e quantity
ωS k(, )
ee
can be obtained from the standard Lindhard
dielectric function within the random-phase approximation (RPA), or extended to include collisions as proposed
by Mermin42. iele et al. generalized the Mermin form to include a dynamic collision frequency within the Born
approximation43, and Arkhipov et al. generalized the Mermin form to two-component plasmas, including sum
rules44.
e RPA results were also improved by including exchange and correlations through the local eld corrections4552.
e ionic correlations contributions in a warm dense matter have been considered by Gregori and Gericke53. In
this scheme, the strongly coupled eects of the ions are included through the dierent components of the mem-
ory function constrained by the sum rules54. is phenomenological approach has been applied successfully in
Coulomb liquid5457 community for systems where the memory functions have a Gaussian or exponential form.
However, for more complex systems, the form of the memory becomes mathematically intractable. Schmidt and
coworkers58 have proposed a hydrodynamic model that begins with moments of the Wigner-Poisson system with
a collision term added. In such an approach you cannot describe correlations properly since the resulting pressure
term is of an ideal gas. e DDFT-QHD approach we introduce here accounts for self-consistently many-body
physics eects and also non-local hydrodynamic eects through the choice of the free-energy functional.
The linear susceptibility associated with a weak external potential
δωvk(, )
ext
that induces a disturbance
δωnk(, )
in the electronic density
ωnk(, )
is dened as
χω
δω
δω
=.
n
v
k
k
k
(, )
(, )
(, )
(7)
e
ext
e
us, the susceptibility can be determined by linearizing the quantum hydrodynamics equation and using (7). To
do, we rst expand the density and velocities about a uniform mean as
δ=+nt nntrr(, )(,),(8)
0
δ=ttur ur(, )(,),(9)
which yields the linearized QHD equations in Fourier space:
ωδ δ−+ ⋅=nnku0, (10)
0
ωδ δ
δ
δξηδ δ−=−|+
+
+
nn
V
n
niknvuk uk
4
3
,
(11)
xt00020e
where
δδ=ΩVnnr() []/
, and the tilde sign denotes the Fourier transform. By combining (10) and (11) and using
(7), we obtain an expression for the electron susceptibility:
χω ωδ
δ
ηω=−+|
kknknV
ni
k
n
k
(, )() ,
(12)
el
e2022
00
2
0
where
ηηξ=+(4 /3 )
l
is the longitudinal viscosity. To proceed, we need to choose specic forms for the dierent
contributions of the free-energy functional [n]. e free-energy functional is typically chosen to ensure that an
accurate equilibrium density is recovered, although exact analytical forms are generally not known. However, the
contributions of the excess free-energy functional to the free energy of the system,
=Ω +Ωnnn[] [] []
ex Hxc
, can
be expressed formally in terms of the direct correlation function
|− |crr()
ee
as follows59:
∫∫
μ
β
′′Ω=Ω+ ∆− ∆∆ |− |+nn dn dd nncOnrr rr rrrr[] [] ()
1
2
() () ()(),
(13)
xxxeee0ee2
where
∆= nnnrr() () 0
, and μex is the excess chemical potential. Once the pair potential
v q()
has been speci-
fied, the self-consistent contributions of the excess free-energy functional can be calculated using the
direct-correlation function via path integral quantum Monte Carlo (PIMC) simulations60,61, integral equations62
or analytical ts63. Let us now evaluate the free-energy functional of the non-interacting electron gas] T[n].
Many approximations for T[n]6466, have been described in the literature- omas-Fermi (TF), Kirzhnits gra-
dient correction (TFK)65, von Weizsäcker funtional (vW)64, Perrot functional66 to name a few. Most of these
models are based on an extension of the TF functional
πβ αα α=
Tn dI Ir[] 2() 2
3(),
(14)
FT25/2 1/23/2
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
4
SCiEntifiC REPORTS | 7: 15352 | DOI:10.1038/s41598-017-14414-9
where
=
+
α
I
x
e
dx
1(15)
p
p
x
0
is the Fermi-Dirac integral of order p,
αr()
is the chemical potential normalized with the kBT and the electron
density is given by
πβ
α=.nIrr()
2
[()]
(16)
23/2 1/2
We consider here the functional form for an electron gas based on the TF functional with the nite-temperature
Kirzhnits gradient correction (TFK)65:
γ
π
β
α
α
=+
|∇ |+ .
Tn Tn d
I
Inr[] []
32
8
()
()
(17)
FK FTT
2
3/21/2
1/2
22
Here, we introduced a coecient γ that allows to capture a variety of results. First, the systematic gradient expan-
sion of Kirzhnits yields the prefactor γ = 1/9. Second, the von Weizsäcker result follows by a partial integration
and γ = 1. e assumption in this gradient-correction expansion is that the error made by neglecting the third-
and higher-order terms is very small. For high-density plasmas, interface-mixing problems or shock structures
in which temperature and density gradients can be large, this expansion ceases to be valid. In such circum-
stances, it may be important to include higher-order terms for the thermal terms through higher gradient cor-
rections q in the TFK functional. e non-interacting free-energy functional T[n] can also be expressed in terms
of the Lindhard function, which is exactly known, instead of using OFDFT66,67. Let us now show the connection
between DFT and the Bohm description. e functional derivative of the kinetic energy functional (17) is given
by20
δ
δ
α
βγπβξαξα=+ ′|∇| +∇
T
n
r
nn
() 32
8
[()2() ],
(18)
FKT
2
3/22 2
where
ξααα=′
−−
II() ()/()
1/21/2
2
, its derivative with respect to the density is denoted
ξ α
()
. e rst term of (18) is
the Fermi pressure while the second term corresponds to the generalized Bohm potential in the nite-temperature
regime, revealing the underlying connection between DFT and the Bohm description. e connection between
DFT and the Bohmian picture has recently been discussed in a paper by Stanton and Murillo20 where a Kirzhnits65
correction was used to get the “Bohm” term quite generally.
Aer linearizing and taking the Fourier transform of (13) and (17), the contributions of the non-interacting
and interacting electron gases to the susceptibility become
δ
δπλ πν λβ|= +−
V
n
kCk
k
k
()
()
4(),
(19)
FFe0T
22
T
41
e
where the TF length and the parameter
ν
are given by
λπβ
ανβ
πα==
II
2
4()and
8
3(),
(20)
FT
2
1/20
3/20
respectively. Substituting (19) into (12), we nd
χω πωλ ληω
=−+ +− .
ν
Γ
−−
qa
q
qq nC qiq
(, )
1
4()
(21)
e
FF
q
e
l
e2
2
22
T2
4
4T4
3
0e 2
2
Here the frequency ω is in units of the electron plasma frequency,
=| |qka
is the wave number,
π=an(3/4 )1/3
is
the Wigner-Seitz radius,
λ λ=a/
FFTT
is the omas-Fermi length in the units of the Wigner-Seitz radius, the
viscosity is in units of
ωn a
p2
, and the coupling parameter, dened as the ratio of the potential energy to the average
kinetic energy, is given by
βΓ= a/
.
Substituting (21) into (5) yields the free-electron DSF
()
Sq
nr r
q
nC qq qq
(, )
1
3
1
1exp(3/) () ()
(22)
pe
ss
l
q
eFFl
e
4
230e 2T244T4222
2
ωω
πω
ηω
ωλληω
=−−Γ+−−+
.
Γ
ν
−−
Equation (22) is the main result of this work. e second factor is the usual Bose function. e denominator
of the third factor includes quantum degenerate plasma eects through the direct correlation function
C q()
ee
,
thermal eects with high-order gradient terms, and viscous damping through
ηl
. It is worth noting that when the
degeneracy parameter
θ∼Γr/
s
is very large, electrons can be considered to be in a non-degenerate, classical state.
If we then replace the exponential in the Bose function by its Taylor expansion, we recover the dynamic structure
of non-degenerate electrons given by the Navier-Stokes model68.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
5
SCiEntifiC REPORTS | 7: 15352 | DOI:10.1038/s41598-017-14414-9
e form of the DSF obtained from this theory, without the dissipative eects, is connected to the approaches
based on the local field corrections49. The direct correlation function which is the main ingredient of this
approach is related to the local eld correction (LFC)
G q()
as:
β
=− −.Gk
vk
Ck() 1
1
()
()
(23)
ee
In the random phase approximation (RPA), the direct correlation function is given by
β=−Gkvk() ()
ee
con-
sequently the LFC vanishes,
=G k() 0
. us, the direct correlation function describes the strongly Coulomb cor-
relation and exchange eects beyond the RPA. Several approximations for the LFC45,46,49,69, has been proposed
starting from the formulation of local eld corrections due to Coulomb correlations and exchange eects by
Hubbard45. Utsumi and Ichimaru49 formula has been widely used to investigated the static properties of systems
at metallic densities. Holas, Aravind and Singwi70 have suggested an expression for the dynamical LFC in strongly
coupled electron gas. Although, we can use existing analytical ts for the LFC to obtain the direct correlation
function, we choose here to computed this quantity directly using Ornstein-Zernike equations with the
hypernetted-chain approximation closure71. Furthermore, matter under extreme conditions of temperature and
pressure undergoes large spatial gradients (i.e., shocks structure, interface problems, etc.). e heterogeneity can
greatly altered the ompson spectra with respect to the uniform case as recently discussed by Kozlowski and
coworkers72. In our approach, the constitutive equations will relax to the correct DFT thermodynamic ground
state, which other methods cannot guarantee. is means we have the full non-local correlations absent from
most other approaches opening up the possibility of studying the cases for which the usual homogeneous and
isotropic forms like
ε ωk(, )
are not applicable. is can be done through the introduction of the inhomogeneous
direct correlation function
crr(, )
ee
. In the next section we will focus on the characteristic features of our main
result (22).
Discussion
Computing the DSF (22) requires knowledge of the viscosity and the direct correlation function. In the literature,
the DSF is oen expressed in terms of the local eld correlation, and the latter quantity is evaluated using analyt-
ical ts52,73,74. e direct correlation function
C q()
ee
can also be obtained directly through numerical simulations;
that is the avenue pursued in this work, using hypernetted-chain calculations62,75, with a quantum statistical
potential (QSP)7678. We use QSP approach here to merely have easy access to results for which we can illustrate
the DDFT method, which is the main point of the paper; other methods can be also used to get the structure
information for the DDFT model. In fact, we see a strength of QSPs in this regard: the key quantity is the
electron-electron
c r()
ee
, which is not accessible from DFT approaches. Electron-electron correlation functions are
available from PIMC, however, and that provides validation for our input quantities. Jones and Murillo79 have
shown the theoretical underpinnings of QSPs and reviewed their extension to fully degenerate quantum systems.
Dutta and Duy60 have compared compared QSP-based RDFs from the modied Kelbg QSP and PIMC and show
that over an extremely wide range of physical conditions the QSP predictions are nearly perfect; it is only at very
low densities that we can see a modest deviation. Here, we choose the QSP from the pioneering work of Hansen
and McDonald76 since they yield results similar to the more complicated Kelbg potentials. Comparisons between
Coulomb, HM and Kelbg potentials are shown in the online Supplementary Material. Next, we turn to the
Figure 1. Eects of the coupling parameter on the spectra. We show the variation in the DSF for dierent
values of the coupling parameter Γ and the normalized wavelengths
=| |qak
. e coupling parameter Γ ranges
from 0.2 to 0.8, and rs ranges from 1.0 to 4.0. e dynamic structure factor
ωS q(, )
ee
is normalized by its
maximum value. e two plasmon peaks are symmetric and the ratio of their amplitudes gives a measure of the
electrons temperature.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
6
SCiEntifiC REPORTS | 7: 15352 | DOI:10.1038/s41598-017-14414-9
electronic viscosity is needed, which is determined by both electron-electron and electron-ion collisions. e
electron viscosity is obtained by interpolating the zero-temperature viscosity proposed by Conti and Vignale80
and the nite-temperature viscosity for classical plasmas by Stanton and Murillo33. When building this t, we
considered contributions only from electron-electron interactions. However, the electron viscosity should also
take into account electron-ion8183, contributions, which can be more signicant than the electron-electron vis-
cosity in the regimes of interest. Please see the online Supplementary Material for a detailed description of the
calculation of these two quantities.
Figure1 shows the spectra of the DSF for dierent values of the wave number q, the coupling strength Γ and
the density parameter rs. e free electron DSF is normalized by its maximum value. In Fig.1a, the positions of
the plasmon peaks (Stokes and anti-Stokes) and its amplitude remain almost unchanged when the quantum
parameter rs increases from 1.0 to 4.0. e reason is, in this regime
θ(1)
, the correlations and quantum degen-
eracy eects are negligible and consequently have no impact on the propagation of the plasmon. Furthermore, the
width of the plasmon peak reduces when q and rs increase owing to the fact that the viscosity which acts to
broaden the width of the peak is very sensitive to the density parameter rs. Figure1a–c show that the position and
width of the plasma peak vary strongly with Γ. ese gures also display some of the standard features of the
plasmons peaks; they are symmetric with respect to the zero frequency, and the dierence between their ampli-
tudes gives a measure of the electron temperature through the detailed balance relation. It is worth noting that for
large values of the coupling parameter and density, the plasmon peak is at a frequency smaller than the plasma
frequency
ωp
.
For a given value of the wave number q, the peak of
ωS q(, )
ee
corresponds to the dispersion relation of the
plasmon
ω ω=q()
q
. According to (22), the dispersion relation is approximately given by
ωω Γ
νη
−+λ+λ−
.
‐‐
qnC qq qq
3() 42
(24)
qp eF Fl
22
2
0e T
22T
44
24
In the RPA limit, with a pure Coulomb potential, the direct correlation function is given by
β=−Cqvq() ()
ee
.
By substituting this result into (24) and setting the viscosity equal to zero, we recover the Bohm-Gross dispersion
relation84 for an isothermal plasma. e dispersion relation of the plasmon is shown in Fig.2 for
=.r186
s
and for
Γ = 1.0 and Γ = 0.7. e red triangles lines show the DDFT-QHD result (24), and the data points indicated with
blue line corresponds to (24), with the viscosity set to zero, ηl = 0. From our basic result (24) we can explore sev-
eral limits that yield other models. For example, because the direct correlation function is related to the local-eld
correction through the relation
β=− cqvqGq() ()[1 ()]
ee ee
, we can neglect the three higher-order terms in (24)
to obtain the local-eld correction (LFC) result. We show this limit in Fig.2 as a green line. Next, we can retain
the second and third terms to example quantum corrections to the LFC result, and that model is shown as blue
triangles. e full result, including viscosity is shown as the red line; note that the viscous correction is large,
suggesting that the power series in q of (24) is probably not converged at the largest values of q in the plot.
e dispersion relation (24) implies that the width of the peak of the DSF is broadened by the viscosity.
erefore, by measuring the width of the peak of a spectrum, information about the electron viscosity can be
inferred. We can obtain a good estimate of the width in the following way. We know the location of the peak from
the dispersion relation (24). Near that peak, we know what ω is, and this information can be put into
ηωqlpeak
2
to
Figure 2. Plasmon dispersion relation. We show the frequency as a function of the wave number q for
=.r186
s
and (a) Γ = 1.0 and (b) Γ = 0.7. DDFT-QHD refers to (24), which accounts for strong correlations and viscosity.
e label “DDFT-QHD:
η=0
l
” corresponds to (24) with the viscosity set to zero. e green curve shows the
local eld correction dispersion relation.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
7
SCiEntifiC REPORTS | 7: 15352 | DOI:10.1038/s41598-017-14414-9
obtain the width of the DSF. is scaling gives the width in terms of both ηl and
C q()
ee
, which means that ηl cannot
be determined without knowledge of the direct correlation function. However, by tting the entire spectrum of
the DSF, all quantities can be obtained: the density, the temperature, and the viscosity. e DSF spectra suggest
that measurements at a few q values is best.
Concluding Remarks
A general framework for electron dynamics is provided with DDFT-QHD. Our model is the specic approxima-
tion of TDDFT in which the generalized-force functional is replaced by the equilibrium functional. We estab-
lished the connection between DDFT-QHD and DSF through the uctuation-dissipation theorem, allowing for
improved QHD models to be compared with experimental data. e predicted DSF spectrum exhibits strong cor-
relations and collisions that are built self-consistently into the model; this result diers from those obtained with
more common Lindhard approaches4244, in which collisions enter through a dynamic collision frequency43,44, or
though local eld corrections85. Our result suggests that the electronic viscosity can be determined experimen-
tally by measuring the electron DSF.
Our approach is a full hydrodynamics model that can be used to simulate non-equilibrium, heterogeneous
dense plasmas72. For example, we could investigate shock physics, uid instabilities, and large-scale experiments.
Most other methods are based on stationary, homogeneous/isotropic approximations; this is explicit in functions
such as
χ ωk(, )
. us, while our DDFT formulation of QHD is signicantly beyond these simple linear response
functions, we show here that we are able to make contact with the XRTS community and connect the scattering
spectrum to transport coecients in a direct way with a hydrodynamic approach.
Finally, our model still lacks an equation for the energy uctuations. Past experience suggests5457,86, that
energy uctuations might cause a zero frequency mode. is latter was experimentally measured in a liquid
lithium by Sinn and coworkers87 conrming molecular dynamics simulations results performed by Canales,
González, and Padró86. Our DDFT-QHD approach would miss any mode originating from thermal uctua-
tions54,58,88, because it is based on an isothermal assumption. We think this approach can incorporate an energy
equation, but this is work in progress. In future work, it would be useful to include other transport quantities,
such as the viscoelastic relaxation time and the thermal conductivity. Extension of this model to the full XRTS
form factor with an electron-ion generalization of the DDFT-QHD equations89,90, is le for the future.
References
1. Bostedt, C. et al. Linac coherent light source: e rst ve years. ev. Mod. Phys. 88, 015007 (2016).
2. Lindl, J., Landen, O., Edwards, J. & Moses, E. eview of the National Ignition Campaign 2009–2012. Phys. Plasmas 21, 020501
(2014).
3. Saunders, A. M. et al. X-ray omson scattering measurements from hohlraum-driven spheres on the OMEGA laser. ev. Sci.
Instrum. 87, 11E724 (2016).
4. Crowley, B. J. B. Continuum lowering - A new perspective. High Energy Density Physics 13, 84 (2014).
5. Mcelvey, A. et al. Thermal conductivity measurements of proton-heated warm dense matter. In APS Shoc Compression of
Condensed Matter Meeting Abstracts (2015).
6. Gregori, G. et al. Measurement of carbon ionization balance in high-temperature plasma mixtures by temporally resolved X-ray
scattering. J. Quant. Spectrosc. adiat. Transf. 99, 225 (2006).
7. Meezan, N. B. et al. Indirect drive ignition at the National Ignition Facility. Plasma Physics and Controlled Fusion 59, 014021 (2017).
8. omas et al. H. Explosions of Xenon Clusters in Ultraintense Femtosecond X-ay Pulses from the LCLS Free Electron Laser. Phys.
ev. Lett. 108, 133401 (2012).
9. Hegelich, B. M. et al. Laser acceleration of quasi-monoenergetic MeV ion beams. Nature 439, 441 (2006).
10. Fletcher, A., Close, S. & Mathias, D. Simulating plasma production from hypervelocity impacts. Physics of Plasmas 22, 093504
(2015).
11. Bigot, J.-Y., Halté, V., Merle, J.-C. & Daunois, A. Electron dynamics in metallic nanoparticles. Chemical Physics 251, 181–203 (2000).
12. Wang, Y. & Eliasson, B. One-dimensional rarefactive solitons in electron-hole semiconductor plasmas. Phys. ev. B 89, 205316
(2014).
13. D avis, P. et al. X-ray scattering measurements of dissociation-induced metallization of dynamically compressed deuterium. Nat.
Commun. 7, 11189 (2016).
14. Bloch, F. B. von Atomen mit mehreren Eletronen. Zeitschri fur Physi 81, 363–376 (1933).
15. Hohenberg, P. & ohn, W. Inhomogeneous electron gas. Phys. ev. 136, B864–B871 (1964).
16. Mermin, N. D. ermal Properties of the Inhomogeneous Electron Gas. Physical eview 137, 1441–1443 (1965).
17. Ying, S. C. Hydrodynamic response of inhomogeneous metallic systems. Nuovo Cimento B Serie 23, 270 (1974).
18. Gasser, I. & Jüngel, A. e quantum hydrodynamic model for semiconductors in thermal equilibrium. Zeitschri Angewandte
Mathemati und Physi 48, 45–59 (1997).
19. Michta, D., Graziani, F. & Bonitz, M. Quantum Hydrodynamics for Plasmas - a omas-Fermi eory Perspective. Contrib. Plasma
Phys. 55, 437 (2015).
20. Stanton, L. G. & Murillo, M. S. Unied description of linear screening in dense plasmas. Phys. ev. E 91, 033104 (2015).
21. Gardner, C. L. Quantum hydrodynamic model for semiconductor devices. SIAM Journal of Applied Mathematics 54, 409–427
(1994).
22. Manfredi, G. & Haas, F. Self-consistent uid model for a quantum electron gas. Phys. ev. B 64, 075316 (2001).
23. Levermore, C. D. Moment closure hierarchies for inetic theories. Journal of Statistical Physics 83, 1021–1065 (1996).
24. Degond, P. & inghofer, C. Quantum moment hydrodynamics and the entropy principle. Journal of Statistical Physics 112, 587–628
(2003).
25. Gardner, C. L. Quantum hydrodynamic model for semiconductor devices. SIAM J. Appl. Math. 54, 409 (1994).
26. Marini Bettolo Marconi, U. & Tarazona, P. Dynamic density functional theory of uids. J. Chem. Phys 110, 8032 (1999).
27. Lutso, J. F. Density functional theory of inhomogeneous liquids. III. Liquid-vapor nucleation. J. Chem. Phys. 129, 244501–244501
(2008).
28. Diaw, A. & Murillo, M. S. Generalized hydrodynamics model for strongly coupled plasmas. Phys. ev. E 92, 013107 (2015).
29. unge, E. & Gross, E. . U. Density-functional theory for time-dependent systems. Phys. ev. Lett. 52, 997 (1984).
30. Goddard, B. D., Nold, A., Savva, N., Pavliotis, G. A. & alliadasis, S. General dynamical density functional theory for classical uids.
Phys. ev. Lett. 109, 120603 (2012).
31. Marconi, U. M. B. & Tarazona, P. Dynamic density functional theory of uids. J. Chem. Phys. 110, 8032 (1999).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
8
SCiEntifiC REPORTS | 7: 15352 | DOI:10.1038/s41598-017-14414-9
32. ex, M. & Löwen, H. Inuence of hydrodynamic interactions on lane formation in oppositely charged driven colloids. Eur. Phys. J.
E 26, 143 (2008).
33. Stanton, L. G. & Murillo, M. S. Ionic transport in high-energy-density matter. Phys. ev. E 93, 043203 (2016).
34. Perdew, J. P., Bure, . & Ernzerhof, M. Generalized gradient approximation made simple. Phys. ev. Lett. 77, 3865 (1996).
35. Malone, F. D. et al. Accurate exchange-correlation energies for the warm dense electron gas. Phys. ev. Lett. 117, 115701 (2016).
36. arasiev, V. V., Sjostrom, T., Duy, J. & Tricey, S. B. Accurate homogeneous elect ron gas exchange-correlation free energy for local
spin-density calculations. Phys. ev. Lett. 112, 076403 (2014).
37. Huang, C. & Carter, E. A. Nonlocal orbita l-free inetic energy density functional for semiconductors. Phys. ev. B 81, 045206 (2010).
38. Frenel, J. inetic eory of Liquids (Clarendon, Oxford, 1946).
39. Glenzer, S. H. & edmer, . X-ray thomson scattering in high energy density plasmas. ev. Mod. Phys. 81, 1625 (2009).
40. Chihara, J. Interaction of photons with plasmas and liquid metals - photoabsorption and scattering. J. Phys. Condens. Matter 12, 231
(2000).
41. Sahoo, S., Gribain, G. F., Shabbir Naz, G., ohano, J. & iley, D. Compton scatter proles for warm dense matter. Phys. ev. E 77,
046402 (2008).
42. Mermin, N. D. Lindhard dielectric function in the relaxation-time approximation. Phys. ev. B 1, 2362 (1970).
43. iele, . et al. omson scattering on inhomogeneous targets. Phys. ev. E 82, 056404 (2010).
44. Arhipov, Y. V. & Davletov, A. E. Screened pseudopotential and static structure factors of semiclassical two-component plasmas.
Physics Letters A 247, 339–342 (1998).
45. Hubbard, J. e Description of Collective Motions in Terms of Many-Body Perturbation eory. Proceedings of the oyal Society of
London Series A 240, 539–560 (1957).
46. Singwi, . S., Tosi, M. P., Land, . H. & Sjölander, A. Electron correlations at metallic densities. Phys. ev. 176, 589–599 (1968).
47. Vashishta, P. & Singwi, . S. Electron Correlations at Metallic Densities. V. Phys. ev. B 6, 875–887 (1972).
48. Vaishya, J. S. & Gupta, A. . Dielectric esponse of the Electron Liquid in Generalized andom-Phase Approximation: A Critical
Analysis. Phys. ev. B 7, 4300–4303 (1973).
49. Utsumi, . & Ichimaru, S. Dielectric formulation of strongly coupled electron liquids at metallic densities. II. Exchange eects and
static properties. Phys. ev. B 22, 5203–5212 (1980).
50. Geldart, D. J. W. & Voso, S. H. e screening function of an interacting electron gas. Canadian Journal of Physics 44, 2137 (1966).
51. Dharma-wardana, M. W. C. & Perrot, F. Simple classical mapping of the spin-polarized quantum electron gas: Distribution funct ions
and local-eld corrections. Phys. ev. Lett. 84, 959–962 (2000).
52. Gregori, G., avasio, A., Höll, A., Glenzer, S. H. & ose, S. J. Derivation of the static structure factor in strongly coupled non-
equilibrium plasmas for X-ray scattering studies. High Energy Density Physics 3, 99 (2007).
53. Gregori, G. & Gerice, D. O. Low frequency structural dynamics of warm dense mattera). Physics of Plasmas 16, 056306 (2009).
54. Boon, J. P. & Yip, S. Molecular hydrodynamics (Dover Publications, New Yor, 1991).
55. Pines, D. & Nozières, P. e eory of Quantum Liquids (W. A. Benjamin, New Yor, 1989).
56. ugler, A. A. Collective modes, damping, and the scattering function in classical liquids. Journal of Statistical Physics 8, 107–153
(1973).
57. Hansen, J. P., McDonald, I. . & Polloc, E. L. Statistical mechanics of dense ionized matter. iii. dynamical properties of the classical
one-component plasma. Phys. ev. A 11, 1025–1039 (1975).
58. Schmidt, ., Crowley, B. J. B., Mithen, J. & Gregori, G. Quantum hydrodynamics of strongly coupled electron uids. Phys. ev. E 85,
046408 (2012).
59. Hansen, J. & McDonald, I. inetic eory of Liquids (Academic, London, 1986).
60. Dutta, S. & Duy, J. Uniform electron gas at warm, dense matter conditions. EPL (Europhysics Letters) 102, 67005 (2013).
61. Brown, E. W., Clar, B. ., DuBois, J. L. & Ceperley, D. M. Path-Integral Monte Carlo Simulation of the Warm Dense Homogeneous
Electron Gas. Phys. ev. Lett. 110, 146405 (2013).
62. Xu, H. & Hansen, J.-P. Density-functional theory of pair correlations in metallic hydrogen. Phys. ev. E 57, 211 (1998).
63. Groth, S., Dornheim, T. & Bonitz, M. Free Energy of the Uniform Electron Gas: Testing Analytical Models against First Principle
esults. ArXiv e-prints (2016).
64. Weizsäcer, C. F. V. Zur eorie der ernmassen. Zeitschri fur Physi 96, 431 (1935).
65. irzhnits, D. Quantum Corrections to the omas-Fermi Equation. ZSoviet Phys. JETP 5, 64 (1957).
66. Perrot, F. Hydrogen-hydrogen interaction in an electron gas. J. Phys.: Cond. Mat. 6, 431 (1994).
67. Wang, L.-W. & Teter, M. P. inetic-energy functional of the electron density. Phys. ev. B 45, 13196 (1992).
68. Murillo, M. S. X-ray thomson scattering in warm dense matter at low frequencies. Phys. ev. E 81, 036403 (2010).
69. Farid, B., Heine, V., Engel, G. E. & obertson, I. J. Extremal properties of the harris-foules functional and an improved screening
calculation for the electron gas. Phys. ev. B 48, 11602–11621 (1993).
70. Holas, A., Aravind, P. . & Singwi, . S. Dynamic correlations in an electron gas. I. First-order perturbation theor y. Phys. ev. B 20,
4912–4934 (1979).
71. Wünsch, ., Hilse, P., Schlanges, M. & Gerice, D. O. Structure of strongly coupled multicomponent plasmas. Phys. ev. E 77,
056404 (2008).
72. ozlowsi, P. M., Crowley, B. J. B., Gerice, D. O., egan, S. P. & Gregori, G. eory of omson scattering in inhomogeneous media.
Scientic eports 6, 24283 (2016).
73. Ichimaru, S. Nuclear fusion in dense plasmas. ev. Mod. Phys. 65, 255 (1993).
74. Nagao, ., Bonev, S. A. & Ashcro, N. W. Cusp-condition constraints and the thermodynamic properties of dense hot hydrogen.
Phys. ev. B 64, 224111 (2001).
75. Chihara, J. Unied description of metallic and neutral liquids and plasmas. J. Phys. Condens. Matter 3, 8715 (1991).
76. Hansen, J. P. & McDonald, I. . Microscopic simulation of a strongly coupled hydrogen plasma. Phys. ev. A 23, 2041 (1981).
77. Schwarz, V. et al. Static ion structure factor for dense plasmas: Semi-classical and ab initio calculations. High Energ. Dens. Phys. 6,
305 (2010).
78. Lado, F. Eective Potential Description of the Quantum Ideal Gases. J. Chem. Phys. 47, 5369–5375 (1967).
79. Jones, C. S. & Murillo, M. S. Analysis of semi-classical potentials for molecular dynamics and Monte Carlo simulations of warm
dense matter. High Energy Density Physics 3, 379–394 (2007).
80. Conti, S. & Vignale, G. Elasticity of an electron liquid. Phys. ev. B 60, 7966 (1999).
81. Murillo, M. S. Viscosity estimates of liquid metals and warm dense matter using the Yuawa reference system. High Energ. Dens.
Phys. 4, 49 (2008).
82. Clérouin, J. e viscosity of dense hydrogen: from liquid to plasma behaviour. J. Phys. Condens. Matter 14, 9089 (2002).
83. Faussurier, G., Libby, S. B. & Silvestrelli, P. L. e viscosity to entropy ratio: From string theory motivated bounds to warm dense
matter transport. High Energ. Dens. Phys. 12, 21 (2014).
84. Gouedard, C. & Deutsch, C. Dense electron-gas response at any degeneracy. Journal of Mathematical Physics 19, 32–38 (1978).
85. Ichimaru, S. & Tanaa, S. Generalized viscoelastic theory of the glass transition for strongly coupled, classical, one-component
plasmas. Phys. ev. Lett. 56, 2815 (1986).
86. Canales, M., González, L. E. & Padró, J. À. Computer simulation study of liquid lithium at 470 and 843 . Phys. ev. E 50, 3656–3669
(1994).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
9
SCiEntifiC REPORTS | 7: 15352 | DOI:10.1038/s41598-017-14414-9
87. Sinn, H. et al. Coherent dynamic structure factor of liquid lithium by inelastic x-ray scattering. Phys. ev. Lett. 78, 1715–1718 (1997).
88. Mountain, . D. Spectral distribution of scattered light in a simple uid. ev. Mod. Phys. 38, 205–214 (1966).
89. Fu, Z.-G. et al. Dynamic properties of the energy loss of multi-mev charged particles traveling in two-component warm dense
plasmas. Phys. ev. E 94, 063203 (2016).
90. Barriga-Carrasco, M. D. Target electron collision eects on energy loss straggling of protons in an electron gas at any degeneracy.
Physics of Plasmas 15, 033103 (2008).
Acknowledgements
is work was supported by the Air Force Oce of Scientic Research (Grant No. FA9550-12-1-0344).
Author Contributions
M.M. conceived the project. A.D. and M.M. wrote the main manuscript text. A.A. prepared gures. Numerical
data were analysed by both authors.
Additional Information
Supplementary information accompanies this paper at https://doi.org/10.1038/s41598-017-14414-9.
Competing Interests: e authors declare that they have no competing interests.
Publisher's note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional aliations.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International
License, which permits use, sharing, adaptation, distribution and reproduction in any medium or
format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Cre-
ative Commons license, and indicate if changes were made. e images or other third party material in this
article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the
material. If material is not included in the article’s Creative Commons license and your intended use is not per-
mitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the
copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/.
© e Author(s) 2017
Content courtesy of Springer Nature, terms of use apply. Rights reserved
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com

Supplementary resource (1)

... Pioneering efforts on models [119][120][121] have matured into production-level codes that accelerate DFT calculations for realistic materials at finite temperature and pressure 122,123 and have recently enabled simulations at unprecedented system size. 124 In combination with complementary methods such as quantum hydrodynamics, [125][126][127][128][129][130] quantum scattering theory, 131,132 and kinetic models, 133,134 these developments have given new insights into the physics of WDM, including the EOS, 84,135,136 effective potentials, [137][138][139][140] and a number of transport properties 141 such as stopping power 109,[142][143][144] and electrical conductivity. 133,145 Many of the properties of WDM are encoded in its response to an external perturbation. ...
... Indeed, setting Gðq; xÞ 0 in Eq. (4) corresponds to the well-known random phase approximation (RPA), which describes the density response on a mean-field level. 154 Consequently, the LFC constitutes key input for a host of applications, such as the construction of advanced XC functionals for DFT via the adiabatic-connection formula, 99,155,156 the incorporation of electron-electron correlations into quantum hydrodynamics, 125,126 or the interpretation of XRTS experiments. [157][158][159][160] In addition, the LFC is formally equivalent to the XC kernel from linear response time-dependent DFT (LR-TDDFT); [161][162][163] they are related by K xc ðq; xÞ ¼ À4p=q 2 Gðq; xÞ. ...
... II A 4. From a practical perspective, we note that a gamut of numerical methods for the description of WDM systems has been presented in the literature, including molecular dynamics, classical Monte Carlo simulations, average-atom models [204][205][206][207][208][209] as well as integral equation theory approaches within the hypernetted-chain approximation for effective quantum potentials, [210][211][212][213] dielectric formalism schemes, 163,[196][197][198][199][200][214][215][216][217][218][219] and quantum hydrodynamics. 125,126,129,130,220 In Sec. II A, we focus on four particularly important methods, namely, path-integral Monte Carlo (Sec. ...
Article
Full-text available
Matter at extreme temperatures and pressures—commonly known as warm dense matter (WDM)—is ubiquitous throughout our Universe and occurs in astrophysical objects such as giant planet interiors and brown dwarfs. Moreover, WDM is very important for technological applications such as inertial confinement fusion and is realized in the laboratory using different techniques. A particularly important property for the understanding of WDM is given by its electronic density response to an external perturbation. Such response properties are probed in x-ray Thomson scattering (XRTS) experiments and are central for the theoretical description of WDM. In this work, we give an overview of a number of recent developments in this field. To this end, we summarize the relevant theoretical background, covering the regime of linear response theory and nonlinear effects, the fully dynamic response and its static, time-independent limit, and the connection between density response properties and imaginary-time correlation functions (ITCF). In addition, we introduce the most important numerical simulation techniques, including path-integral Monte Carlo simulations and different thermal density functional theory (DFT) approaches. From a practical perspective, we present a variety of simulation results for different density response properties, covering the archetypal model of the uniform electron gas and realistic WDM systems such as hydrogen. Moreover, we show how the concept of ITCFs can be used to infer the temperature from XRTS measurements of arbitrary complex systems without the need for any models or approximations. Finally, we outline a strategy for future developments based on the close interplay between simulations and experiments.
... The above list of references is by no means exhaustive and the reader is invited to see the references cited therein for more information. QHD models with viscosity have been studied under the perspective of viscous (numerical) stabilization [33], the physical description of dense plasmas [9,11,25], the existence and stability of viscous-dispersive shocks [14,15,[40][41][42][43]58], the existence of global solutions [19], and vanishing viscosity behaviors [56]. ...
Preprint
In this paper, a compressible viscous-dispersive Euler system in one space dimension in the context of quantum hydrodynamics is considered. The purpose of this study is twofold. First, it is shown that the system is locally well-posed. For that purpose, the existence of classical solutions which are perturbation of constant states is established. Second, it is proved that in the particular case of subsonic equilibrium states, sufficiently small perturbations decay globally in time. In order to prove this stability property, the linearized system around the subsonic state is examined. Using an appropriately constructed compensating matrix symbol in the Fourier space, it is proved that solutions to the linear system decay globally in time, underlying a dissipative mechanism of regularity gain type. These linear decay estimates, together with the local existence result, imply the global existence and the decay of perturbations to constant subsonic equilibrium states as solutions to the full nonlinear system.
... The influence of Landau (Kreibig) damping, which is inversely proportional to the nanoparticle size [36][37][38], is also being discussed. Simultaneously, efforts are being made to find solutions for a more accurate nonlinearized form of HDM [39,40], and the question of viscosive damping of the electron gas is gaining prominence [41][42][43][44]. The hydrodynamic model itself is based on the concept of a jellium model which can be interpreted as the electron fluid moving with respect to a positively charged background of metal ions. ...
Article
Full-text available
The response of plasmonic metal particles to an electromagnetic wave produces significant features at the nanoscale level. Different properties of the internal composition of a metal, such as its ionic background and the free electron gas, begin to manifest more prominently. As the dimensions of the nanostructures decrease, the classical local theory gradually becomes inadequate. Therefore, Maxwell’s equations need to be supplemented with a relationship determining the dynamics of current density which is the essence of nonlocal plasmonic models. In this field of physics, the standard (linearized) hydrodynamic model (HDM) has been widely adopted with great success, serving as the basis for a variety of simulation methods. However, ongoing efforts are also being made to expand and refine it. Recently, the GNOR (general nonlocal optical response) modification of the HDM has been used, with the intention of incorporating the influence of electron gas diffusion. Clearly, from the classical description of fluid dynamics, a close relationship between viscosive damping and diffusion arises. This offers a relevant motivation for introducing the GNOR modification in an alternative manner. The standard HDM and its existing GNOR modification also do not include the influence of interband electron transitions in the conduction band and other phenomena that are part of many refining modifications of the Drude–Lorentz and other models of metal permittivity. In this article, we present a modified version of GNOR-HDM that incorporates the viscosive damping of the electron gas and a generalized Drude–Lorentz term. In the selected simulations, we also introduce Landau damping, which corrects the magnitude of the standard damping constant of the electron gas based on the size of the nanoparticle. We have chosen a spherical particle as a suitable object for testing and comparing HD models and their modifications because it allows the calculation of precise analytical solutions for the interactions and, simultaneously, it is a relatively easily fabricated nanostructure in practice. Our contribution also includes our own analytical method for solving the HDM interaction of a plane wave with a spherical particle. This method forms the core of calculations of the characteristic quantities, such as the extinction cross-sections and the corresponding components of electric fields and current densities.
... The influence of Landau (Kreibig) damping, which is inversely proportional to the nanoparticle size [36][37][38], is also being discussed. Simultaneously, efforts are being made to find solutions for a more accurate nonlinearized form of HDM [39,40], and the question of viscosive damping of the electron gas is gaining prominence [41][42][43]. The hydrodynamic model itself is based on the concept of a jellium model which can be interpreted as the electron fluid moving with respect to a positively charged background of metal ions. ...
Preprint
Full-text available
The response of plasmonic metal particles to an electromagnetic wave undergoes significant features at the nanoscale level. Different properties of the internal composition of a metal, such as its ionic background and the free electron gas, begin to manifest more prominently. As the dimensions of the nanostructures decrease, the classical local theory gradually becomes inadequate. Therefore, Maxwellʹs equations need to be supplemented with a relationship determining the dynamics of current density which is the essence of nonlocal plasmonic models. In this field of physics, the standard (linearized) hydrodynamic model (HDM) has been widely adopted with great success, serving as the basis for a variety of simulation methods. However, ongoing efforts are also being made to expand and refine it. Recently, the GNOR (general nonlocal optical response) modification of the HDM has been used, with the intention of incorporating the influence of electron gas diffusion. Clearly, from the classical description of fluid dynamics, a close relationship between viscosive damping and diffusion arises. This offers a relevant motivation for introducing the GNOR modification in an alternative manner. The standard HDM and its existing GNOR modification also do not include the influence of interband electron transitions in the conduction band and other phenomena that are part of many refining modifications of the Drude‐Lorentz and other models of metal permittivity. In this article, we present a modified version of GNOR‐HDM that incorporates the viscosive damping of the electron gas and a generalized Drude‐Lorentz term. In the selected simulations, we also introduce Landau damping which corrects the magnitude of the standard damping constant of the electron gas based on the size of the nanoparticle. We have chosen a spherical particle as a suitable object for testing and comparing HD models and their modifications because it allows finding precise analytical solutions for the interaction and, simultaneously, it is a relatively easily fabricable nanostructure in practice. Our contribution also includes our own analytical method for solving the HDM interaction of a plane wave with a spherical particle. This method forms the core of calculations of the characteristic quantities, such as the extinction cross‐sections and the corresponding components of electric fields and current densities.
... Finally, we note that a consistent static XC kernel is needed for various other applications such as the construction of effective potentials [47,[109][110][111] for quantum hydrodynamics [112][113][114][115][116] and plasmonics [117] and the for computation of the energy loss characteristics of high-energy density plasmas [118][119][120]. ...
Article
Full-text available
We present a methodology for the linear-response time-dependent density functional theory (LR-TDDFT) calculation of the dynamic density response function of warm dense matter in the adiabatic approximation that can be used with any available exchange-correlation (XC) functional across Jacob's ladder and across temperature regimes. The uniqueness of the presented approach is that it can go beyond the adiabatic local density approximation and adiabatic generalized gradient approximation while preserving the self-consistency between the Kohn-Sham (KS) response function and adiabatic XC kernel for extended systems. The key ingredient to the presented method is the combination of the adiabatic XC kernel from the direct perturbation approach with the macroscopic dynamic KS response from the standard LR-TDDFT method using KS orbitals. We demonstrate the application of the method for the example of warm dense hydrogen, for which we perform a detailed analysis of the KS density response function, the random phase approximation result, the total density response function, and of the adiabatic XC kernel. The analysis is performed using local density approximation, generalized gradient approximation, and meta-generalized gradient approximation level approximations for the XC effects. The presented method is directly applicable to disordered systems such as liquid metals, warm dense matter, and dense plasmas.
... 126-136. Therefore, it constitutes the key input also for a variety of applications, including the construction of effective potentials, 69,71 quantum hydrodynamics, [137][138][139] and the interpretation of XRTS experiments. 82,83,140,141 For completeness, we note that the static linear density response function χ(q) can also be estimated from PIMC simulations of the unperturbed system via the imaginary-time version of the fluctuation-dissipation theorem, ...
Article
We study the linear energy response of the uniform electron gas to an external harmonic perturbation with a focus on resolving different contributions to the total energy. This has been achieved by carrying out highly accurate ab initio path integral Monte Carlo (PIMC) calculations for a variety of densities and temperatures. We report a number of physical insights into effects such as screening and the relative importance of kinetic and potential energies for different wave numbers. A particularly interesting finding is obtained from the observed non-monotonic behavior of the induced change in the interaction energy, which becomes negative for intermediate wave numbers. This effect is strongly dependent on the coupling strength and constitutes further direct evidence for the spatial alignment of electrons introduced in earlier works [T. Dornheim et al., Commun. Phys. 5, 304 (2022)]. The observed quadratic dependence on the perturbation amplitude in the limit of weak perturbations and the quartic dependence of perturbation amplitude corrections are consistent with linear and nonlinear versions of the density stiffness theorem. All PIMC simulation results are freely available online and can be used to benchmark new methods or as input for other calculations.
... [120][121][122][123][124][125][126][127][128][129][130]. Therefore, it constitutes key input also for a variety of applications including the construction of effective potentials [64,66], quantum hydrodynamics [131][132][133] and the interpretation of XRTS experiments [77,78,134,135]. For completeness, we note that the static linear density response function χ(q) can also be estimated from PIMC simulations of the unperturbed system via the imaginarytime version of the fluctuation-dissipation theorem, ...
Preprint
We present extensive new \emph{ab initio} path integral Monte Carlo (PIMC) simulations of the harmonically perturbed uniform electron gas (UEG) for different densities and temperatures. This allows us to study the linear response of the UEG with respect to different contributions to the total energy for different wave numbers. We find that the induced change in the interaction energy exhibits a non-monotonic behaviour, and becomes negative for intermediate wave numbers. This effect is strongly dependent on the coupling strength and can be traced back to the spatial alignment of electrons introduced in earlier works [T.~Dornheim \emph{et al.}, Communications Physics \textbf{5}, 304 (2022)]. The observed quadratic dependence on the perturbation amplitude in the limit of weak perturbations and the quartic dependence of the perturbation amplitude corrections are consistent with linear and non-linear versions of the density stifness theorem. All PIMC simulation results are freely available online and can be used to benchmark new methods, or as input for other calculations.
... Finally, we note that a consistent static XC kernel is needed for various other applications such as the construction of effective potentials [47,[109][110][111], for quantum hydrodynamics [112][113][114][115][116] and plasmonics [117], and the for computation of the energy loss characteristics of high-energy density plasmas [118][119][120]. ...
Preprint
We present a new methodology for the linear-response time-dependent density functional theory (LR-TDDFT) calculation of the dynamic density response function of warm dense matter in an adiabatic approximation that can be used with any available exchange-correlation (XC) functional across Jacob's Ladder and across temperature regimes. The main novelty of the presented approach is that it can go beyond the adiabatic local density approximation (ALDA) and generalized LDA (AGGA) while preserving the self-consistence between the Kohn-Sham (KS) response function and adiabatic XC kernel for extended systems. The key ingredient for the presented method is the combination of the adiabatic XC kernel from the direct perturbation approach with the macroscopic dynamic KS response from the standard LR-TDDFT method using KS orbitals. We demonstrate the application of the method for the example of warm dense hydrogen, for which we perform a detailed analysis of the KS density response function, the RPA result, the total density response function and of the adiabatic XC kernel. The analysis is performed using LDA, GGA, and meta-GGA level approximations for the XC effects. The presented method is directly applicable to disordered systems such as liquid metals, warm dense matter, and dense plasmas.
Preprint
Full-text available
Orbital-free density functional theory (OF-DFT) constitutes a computationally highly effective tool for modeling electronic structures of systems ranging from room-temperature materials to warm dense matter. Its accuracy critically depends on the employed kinetic energy (KE) density functional, which has to be supplied as an external input. In this work, we use an external harmonic perturbation in OF-DFT to compute the density response function. This allows us to test whether exact conditions in the limit of uniform densities (i.e. for the uniform electron gas, or UEG) are satisfied. We demonstrate the utility of this direct perturbation approach by considering different non-local and Laplacian-level KE functionals. The results illustrate that several functionals violate exact conditions in the UEG limit. Additionally, we also test KE functional approximations beyond the linear density response regime by gradually increasing the density perturbation amplitude and comparing against Kohn-Sham DFT results which employ the exact non-interacting KE functional. The results show a strong correlation between the accuracy of the KE functionals in the UEG limit and in the strongly inhomogeneous case. This empirically demonstrates the importance of the UEG limit based constraint for the construction of accurate KE functionals. This conclusion is substantiated by additional calculations for bulk Aluminum (Al) with a face-centered cubic lattice with and without an external harmonic perturbation. The analysis of the Al data follows closely the conclusions drawn for the UEG, allowing us to extend our conclusions to realistic systems that experience density inhomogeneities induced by ions. Analyzing other classes of KE functionals (e.g., based on the quadratic UEG density response function and the jellium-with-gap model) and other types of systems, such as semiconductors, is left for future studies.
Article
Full-text available
Thermal conductivity is one of the most crucial physical properties of matter when it comes to understanding heat transport, hydrodynamic evolution, and energy balance in systems ranging from astrophysical objects to fusion plasmas. In the warm dense matter regime, experimental data are very scarce so that many theoretical models remain untested. Here we present the first thermal conductivity measurements of aluminum at 0.5–2.7 g/cc and 2–10 eV, using a recently developed platform of differential heating. A temperature gradient is induced in a Au/Al dual-layer target by proton heating, and subsequent heat flow from the hotter Au to the Al rear surface is detected by two simultaneous time-resolved diagnostics. A systematic data set allows for constraining both thermal conductivity and equation-of-state models. Simulations using Purgatorio model or Sesame S27314 for Al thermal conductivity and LEOS for Au/Al release equation-of-state show good agreement with data after 15 ps. Discrepancy still exists at early time 0–15 ps, likely due to non-equilibrium conditions.
Article
Full-text available
The energy loss of multi-MeV charged particles moving in two-component warm dense plasmas (WDPs) is studied theoretically beyond the random-phase approximation. The short-range correlations between particles are taken into account via dynamic local field corrections (DLFC) in a Mermin dielectric function for two-component plasmas. The mean ionization states are obtained by employing the detailed configuration accounting model. The Yukawa-type effective potential is used to derive the DLFC. Numerically, the DLFC are obtained via self-consistent iterative operations. We find that the DLFC are significant around the maximum of the stopping power. Furthermore, by using the two-component extended Mermin dielectric function model including the DLFC, the energy loss of a proton with an initial energy of ∼15 MeV passing through a WDP of beryllium with an electronic density around the solid value ne≈3×1023cm-3 and with temperature around ∼40 eV is estimated numerically. The numerical result is reasonably consistent with the experimental observations [A. B. Zylsta et al., Phys. Rev. Lett. 111, 215002 (2013)PRLTAO0031-900710.1103/PhysRevLett.111.215002]. Our results show that the partial ionization and the dynamic properties should be of importance for the stopping of charged particles moving in the WDP.
Article
Full-text available
Density matrix quantum Monte Carlo is used to sample exact-on-average $N$-body density matrices for uniform electron gas systems of up to 10$^{124}$ matrix elements via a stochastic solution of the Bloch equation. The results of these calculations resolve a current debate over the exchange-correlation energies necessary for the parametrization of finite-temperature density functionals. Exchange-correlation energies calculated using the real-space restricted path-integral formalism and the $k$-space configuration path-integral formalism disagree by up to $\sim$$10 \%$ at certain reduced temperatures $T/T_F < 0.5$ and densities $r_s \le 1$. We are able to verify the exact-in-principle configuration path-integral Monte Carlo numbers at high density and bridge the gap between high and low densities.
Article
Full-text available
Hydrogen, the simplest element in the universe, has a surprisingly complex phase diagram. Because of applications to planetary science, inertial confinement fusion and fundamental physics, its high-pressure properties have been the subject of intense study over the past two decades. While sophisticated static experiments have probed hydrogen's structure at ever higher pressures, studies examining the higher-temperature regime using dynamic compression have mostly been limited to optical measurement techniques. Here we present spectrally resolved x-ray scattering measurements from plasmons in dynamically compressed deuterium. Combined with Compton scattering, and velocity interferometry to determine shock pressure and mass density, this allows us to extract ionization state as a function of compression. The onset of ionization occurs close in pressure to where density functional theory-molecular dynamics (DFT-MD) simulations show molecular dissociation, suggesting hydrogen transitions from a molecular and insulating fluid to a conducting state without passing through an intermediate atomic phase.
Article
Full-text available
Thomson scattering of laser light is one of the most fundamental diagnostics of plasma density, temperature and magnetic fields. It relies on the assumption that the properties in the probed volume are homogeneous and constant during the probing time. On the other hand, laboratory plasmas are seldom uniform and homogeneous on the temporal and spatial dimensions over which data is collected. This is partic- ularly true for laser-produced high-energy-density matter, which often exhibits steep gradients in temperature, density and pressure, on a scale determined by the laser focus. Here, we discuss the modification of the cross section for Thomson scattering in fully-ionized media exhibiting steep spatial inhomogeneities and/or fast temporal fluctuations. We show that the predicted Thomson scattering spectra are greatly altered compared to the uniform case, and may even lead to violations of detailed balance. Therefore, careful interpretation of the spectra is necessary for spatially or temporally inhomogeneous systems.
Article
Full-text available
Ionic transport coefficients for dense plasmas have been numerically computed using an effective Boltzmann approach. We have developed a simplified effective potential approach that yields accurate fits for all of the relevant cross sections and collision integrals. Our results have been validated with molecular-dynamics simulations for self-diffusion, interdiffusion, viscosity, and thermal conductivity. Molecular dynamics has also been used to examine the underlying assumptions of the Boltzmann approach through a categorization of behaviors of the velocity autocorrelation function in the Yukawa phase diagram. Using a velocity-dependent screening model, we examine the role of dynamical screening in transport. Implications of these results for Coulomb logarithm approaches are discussed.
Article
The uniform electron gas is a key model system in the description of matter, including dense plasmas and solid state systems. However, the simultaneous occurence of quantum, correlation, and thermal effects makes the theoretical description challenging. For these reasons, over the last half century many analytical approaches have been developed the accuracy of which has remained unclear. We have recently obtained the first \textit{ab initio} data for the exchange correlation free energy of the uniform electron gas [T. Dornheim \textit{et al.}, Phys.~Rev.~Lett.~\textbf{117}, 156403 (2016)] which now provides the opportunity to assess the quality of the mentioned approaches and parametrizations. Particular emphasis is put on the warm dense matter regime, where we find significant discrepancies between the different approaches.
Article
This paper reviews scientific results from the pursuit of indirect drive ignition on the National Ignition Facility (NIF) and describes the program's forward looking research directions. In indirect drive on the NIF, laser beams heat an x-ray enclosure called a hohlraum that surrounds a spherical pellet. X-ray radiation ablates the surface of the pellet, imploding a thin shell of deuterium/tritium (DT) that must accelerate to high velocity (v > 350 km s⁻¹) and compress by a factor of several thousand. Since 2009, substantial progress has been made in understanding the major challenges to ignition: Rayleigh Taylor (RT) instability seeded by target imperfections; and low-mode asymmetries in the hohlraum x-ray drive, exacerbated by laser-plasma instabilities (LPI). Requirements on velocity, symmetry, and compression have been demonstrated separately on the NIF but have not been achieved simultaneously. We now know that the RT instability, seeded mainly by the capsule support tent, severely degraded DT implosions from 2009-2012. Experiments using a 'high-foot' drive with demonstrated lower RT growth improved the thermonuclear yield by a factor of 10, resulting in yield amplification due to alpha particle heating by more than a factor of 2. However, large time dependent drive asymmetry in the LPI-dominated hohlraums remains unchanged, preventing further improvements. High fidelity 3D hydrodynamic calculations explain these results. Future research efforts focus on improved capsule mounting techniques and on hohlraums with little LPI and controllable symmetry. In parallel, we are pursuing improvements to the basic physics models used in the design codes through focused physics experiments.
Article
X-ray Thomson scattering (XRTS) is a powerful diagnostic for probing warm and hot dense matter. We present the design and results of the first XRTS experiments with hohlraum-driven CH2 targets on the OMEGA laser facility at the Laboratory for Laser Energetics in Rochester, NY. X-rays seen directly from the XRTS x-ray source overshadow the elastic scattering signal from the target capsule but can be controlled in future experiments. From the inelastic scattering signal, an average plasma temperature is inferred that is in reasonable agreement with the temperatures predicted by simulations. Knowledge gained in this experiment shows a promising future for further XRTS measurements on indirectly driven OMEGA targets.