The dynamics of an inertial confinement fusion capsule, or of a stellar or planet interior, obey a very similar set of equations: magneto-radiative-hydrodynamic equations. The solutions of these equations, however, depend entirely on the transport properties associated with the different materials at play. To properly model the dynamics of these systems, it is necessary to determine with high accuracy the transport coefficients of several materials over a large range of thermodynamic conditions. Experimental capabilities in this respect are still limited due to the nature of the microphysics at play and the extreme conditions involved. Numerical simulations are thus necessary, and in this respect, molecular dynamics simulations based on density functional theory offer exquisite possibilities to constrain the transport properties in the warm to hot dense matter regime. In this paper, we report the methodology used to extract different transport properties based on molecular dynamics performed with the software Abinit. The examples shown are based on the specific cases identified for the purpose of the second charged-particle transport code comparison workshop.
I. INTRODUCTION
The modeling of high energy density phenomena, such as inertial confinement fusion (ICF) capsule implosion,1 planetary interiors,2 and stellar interiors,3 relies on magneto-radiation-hydrodynamic simulations. However, to be representative of the microscale physics, it is crucial to capture the proper transport properties of the systems from the initial conditions and throughout their evolution. Transport phenomena play a very important role in the aforementioned processes. For instance, heat conductivity in ICF cavities not only determines the efficiency of the energy transfer and deposit within the cavity4 but also impacts the growth of hydrodynamic instabilities at the ablator–fuel interface.5 In planetary interiors, electrical conductivity is key to characterize the existence of a dynamo process,6 and interdiffusion and self-diffusion coefficients are of first importance to determine the cooling efficiency.7–9 In stars, mixing between the core and the stably stratified layers is driven by convective overshoot and the transport coefficients strongly influence the efficiency of the mixing of the energy transport.10
Density and temperature conditions can vary significantly whether we consider planets, stars, or ICF capsules; the spatial and temporal scales vary significantly as well. The microphysics at play is, however, similar. The use of theoretical models, such as Spitzer–Härm, for thermal conductivities,11 Chapman–Enskog for viscosities,12 or Drude for electrical conductivities,13 provides asymptotic limits. Serious efforts have been made in the recent past to improve our understanding and quantitative characterization of the transport coefficients in the warm dense regime. Average-atom models provided a fair evaluation of electrical conductivities for single species systems.14 To include mixing effects as well as ionic contributions for the viscosity and diffusion coefficients, N-body simulation tools are required.
The advent of molecular dynamics simulations based on density functional theory (DFT), whether using an orbital-free or a Mermin–Kohn–Sham (KS) scheme, offered an exquisite opportunity to explore the N-body effects on transport properties. For instance, orbital-free molecular dynamics simulations of asymmetric mixtures allowed to exhibit very different diffusion mechanisms depending on the concentration of light and heavy elements.15 At much lower temperatures, KS-DFT molecular dynamics shed some light on several processes relevant for planetary interiors. Diffusion coefficients have been determined for the interior of gaseous giants and allowed to constrain the efficiency of the core erosion in Jupiter.16 They also allowed to provide an explanation of the depth of the zonal winds on Uranus and Neptune by invoking ionic electrical conductivities.17 KS-DFT also allowed the computation of precise electrical conductivities for several mixtures,18,19 which can be used in geodynamo simulations.
Until recently, KS-DFT molecular dynamics simulations could only be performed at moderate temperatures because of the orbital wall20 faced by this method. The development of hybrid methods, such as the mixed deterministic and stochastic DFT approach21 or the extended method,22–24 alleviates this issue. It is now possible to simulate warm dense to hot dense matter in a computationally tractable manner using KS-DFT. This provides insight into the role of electronic structure calculations in the computation of transport properties.
In the present paper, we show the current capabilities of KS-DFT molecular dynamics in computing transport coefficients for warm dense matter using the Abinit code.25–27 We explain the strategies developed to design the simulations, to extract, and to post-process the data so as to provide the relevant transport coefficients. We illustrate the discussion using different cases as defined for the second charged-particle transport code comparison workshop.28
II. NUMERICAL SIMULATION METHODOLOGY
To compute the different transport coefficients, we relied on molecular dynamics (MD) simulations in the isokinetic ensemble. The ions are advanced in time with a Verlet algorithm, and we ensured conservation of temperature by utilizing a velocity rescaling thermostat:29 both of which are implemented in Abinit. The time step was adapted to ensure that the trajectory of each atom was properly sampled with respect to the chemical bond oscillation frequency and to the mean-free path. The simulation cell contained 64–256 atoms. Previous studies showed that such a simulation cell would have limited finite-size effects on bulk properties at least [such as equation of states (EOS)].30 Ionic transport properties are more prone to finite-size effects,31,32 but they should be acceptable within the bounds of the other uncertainties. However, small simulation cells allow for longer simulations which contributes to reducing the uncertainty on the ionic transport coefficients as discussed in the following sections. Our simulations were at least 3000 time steps long (up to 15 000 for the longest).
While the nuclei are all treated as classical particles, the electronic density is self-consistently computed at each time step using the Kohn–Sham density functional theory (KS-DFT)33 with the Mermin finite-temperature functional form.34 Following previous work, we opted for either the generalized gradient approximation (GGA) Perdew, Burke and Ernzerhof exchange-correlation functional35 or for a local density approximation (LDA) (see Table I for further details). These functionals are originally designed for 0 K calculations which constitute one source of uncertainty. Temperature-dependent exchange-correlation functionals are under active development,36–39 and codes like Abinit must be adapted to make use of such exchange-correlation functionals. We used specifically designed projector augmented-wave (PAW) pseudo-potentials.27,40 They were designed using the Atompaw code41 and have been utilized for similar plasma conditions as presented here (see, e.g., Refs. 42–44) We used short cutoff radii to prevent the PAW spheres from overlapping including most of the electrons in the valence (see Table I for details). With such hard pseudo-potentials, the energy cutoff of the plane wave expansion had to be set above 50 Ha. To sample the Brillouin zone, we used the Baldereschi mean-value point.45 We chose the number of Kohn–Sham orbitals in order to have a last level occupancy of . For the highest temperatures, we used the extended DFT version of the Abinit code23 to increase the efficiency of the calculations by reducing the number of Kohn–Sham orbitals. The extended DFT model replaces high energy orbitals with pure plane waves, assuming their electrons as free fermions, while keeping a Kohn–Sham description of inner shell electrons with the full plane wave basis set. Much fewer numbers of KS orbitals need to be considered, reducing the computational cost of the calculation. It allowed for simulations containing more particles that were carried out for longer time lengths at multi-eV temperatures.
KS-DFT molecular dynamics input parameters for ionic transport properties analysis. Mean-value and gamma points are, respectively, referred to as MP and Γ. GGA stands for the Perdew–Burke–Ernzerhof (PBE) exchange-correlation functional.35 stands for the number of KS orbitals (excluding extended DFT pure plane waves). stands for the energy cutoff of the plane wave expansion. is the radius of the PAW spheres, is the number of electrons kept in the valence (while other electrons are frozen in the core), and stands for the mean voluminal overlap ratio of the PAW spheres.
. | H1 . | C1 . | CH1 . | Al1 . | Cu1 . | HCu1 . | CH2 . | Au1 . | C2 . |
---|---|---|---|---|---|---|---|---|---|
ρ (g/cm3) | 1.0 | 10.0 | 1.0 | 2.7 | 8.97 | 1.8 | 0.9 | 19.32 | 100.0 |
(eV) | 2.0 | 2.0 | 2.0 | 1.0 | 1.0 | 1.0 | 7.8 | 10.0 | 0.2 |
256 H | 64 C | 48 H–48 C | 108 Al | 108 Cu | 50 H/50 Cu | 48 H/48 C | 108 Au | 108 C | |
-point(s) | MP | 23 | 23 | MP | 23 | MP | 23 | Γ | 33 |
512 | 512 | 512 | 1024 | 2048 | 1024 | 1024 | 3072 | 512 | |
(Ha) | 62.0 | 320.0 | 60.0 | 50.0 | 30.0 | 15.0 | 60.0 | 50.0 | 400.0 |
(atu) | 5.0 | 8.0 | 5.0 | 100.0 | 50.0 | 50.0 | 3.0 | 10.0 | 5.0 |
XC Func. | GGA | GGA | GGA | GGA | LDA | GGA | GGA | LDA | GGA |
(in Bohrs) | 0.7 | 0.3 | 0.8–1.1 | 1.5 | 2.0 | 1.0–2.0 | 0.8–1.1 | 2.25 | 0.3 |
/atom | 1 | 6 | 1–4 | 11 | 19 | 1–19 | 1–4 | 33 | 6 |
18 787 | 7475 | 22 494 | 4148 | 3359 | 1429 | 8334 | 5948 | 5370 | |
Total time (ps) | 2.272 | 1.446 | 2.721 | 10.034 | 4.063 | 1.728 | 0.605 | 1.439 | 0.649 |
0.00% | 8.76% | 0.00% | 6.04% | 20.17% | 15.58% | 13.12% | 0.00% | ||
Notes | Ext. DFT | Ext. DFT | FCC crystal |
. | H1 . | C1 . | CH1 . | Al1 . | Cu1 . | HCu1 . | CH2 . | Au1 . | C2 . |
---|---|---|---|---|---|---|---|---|---|
ρ (g/cm3) | 1.0 | 10.0 | 1.0 | 2.7 | 8.97 | 1.8 | 0.9 | 19.32 | 100.0 |
(eV) | 2.0 | 2.0 | 2.0 | 1.0 | 1.0 | 1.0 | 7.8 | 10.0 | 0.2 |
256 H | 64 C | 48 H–48 C | 108 Al | 108 Cu | 50 H/50 Cu | 48 H/48 C | 108 Au | 108 C | |
-point(s) | MP | 23 | 23 | MP | 23 | MP | 23 | Γ | 33 |
512 | 512 | 512 | 1024 | 2048 | 1024 | 1024 | 3072 | 512 | |
(Ha) | 62.0 | 320.0 | 60.0 | 50.0 | 30.0 | 15.0 | 60.0 | 50.0 | 400.0 |
(atu) | 5.0 | 8.0 | 5.0 | 100.0 | 50.0 | 50.0 | 3.0 | 10.0 | 5.0 |
XC Func. | GGA | GGA | GGA | GGA | LDA | GGA | GGA | LDA | GGA |
(in Bohrs) | 0.7 | 0.3 | 0.8–1.1 | 1.5 | 2.0 | 1.0–2.0 | 0.8–1.1 | 2.25 | 0.3 |
/atom | 1 | 6 | 1–4 | 11 | 19 | 1–19 | 1–4 | 33 | 6 |
18 787 | 7475 | 22 494 | 4148 | 3359 | 1429 | 8334 | 5948 | 5370 | |
Total time (ps) | 2.272 | 1.446 | 2.721 | 10.034 | 4.063 | 1.728 | 0.605 | 1.439 | 0.649 |
0.00% | 8.76% | 0.00% | 6.04% | 20.17% | 15.58% | 13.12% | 0.00% | ||
Notes | Ext. DFT | Ext. DFT | FCC crystal |
In order to compute the electronic transport properties, we had to select several ionic configurations from the molecular dynamics trajectories and perform a more converged DFT calculation. For instance, we increased the Brillouin zone sampling to a 23 Monkhorst–Pack grid.46 We also needed to increase the number of Kohn–Sham orbitals to include a large number of weakly occupied states. For these calculations we opted out of the extended DFT and performed a full Kohn–Sham calculation. We then applied a Kubo–Greenwood formalism to the eigenstates and eigenvalues to extract the real part of the electrical conductivity.
III. EQUILIBRIUM PROPERTIES
A. Equation of state
Pressure evolution of case CH1. (Top) Time evolution of the instantaneous pressure in the molecular dynamics. The dashed line indicates the average value over the time range considered for the statistics. (Bottom) Time evolution of the pressure autocorrelation function. An exponential decay fit is also plotted. The first vertical dashed line shows the characteristic decay time. The second dashed line shows the approximated decorrelation time considered.
Pressure evolution of case CH1. (Top) Time evolution of the instantaneous pressure in the molecular dynamics. The dashed line indicates the average value over the time range considered for the statistics. (Bottom) Time evolution of the pressure autocorrelation function. An exponential decay fit is also plotted. The first vertical dashed line shows the characteristic decay time. The second dashed line shows the approximated decorrelation time considered.
To estimate the statistical uncertainty on the pressure, we use the standard deviation and divide it by the number of uncorrelated steps. The latter is estimated based on the auto-correlation function (ACF) of the pressure fluctuations. We fit the ACF with an exponential decay providing a characteristic decay time. We considered that the pressure was uncorrelated after twice the decay time. An example can be seen in Fig. 1.
B. Pair distribution functions
In Fig. 2, we plotted an example of a pair distribution function (PDF) for the equimolar CH mixture, with the C–C, C–H, and H–H respective curves. The PDF does not provide direct information on the transport properties. It does, however, provide details on the structure of the system. Figure 2 shows, for instance, that hydrogen atoms are mostly dissociated, which means their diffusion is mostly that of an atom. The first peak of the C–C PDF would, however, suggest that some carbon–carbon bonding, possibly transient, does exist, which will reduce the magnitude of the carbon interdiffusion.
IV. IONIC TRANSPORT PROPERTIES
A. Self-diffusion and interdiffusion coefficients
Mean squared displacement of aluminum nuclei in the Al1 case. In purple, the MSD was computed with a single origin, the blue case is with multiple origins. The shaded areas correspond to the one-σ deviation and the dashed lines depict an affine fit of the evolutions.
Mean squared displacement of aluminum nuclei in the Al1 case. In purple, the MSD was computed with a single origin, the blue case is with multiple origins. The shaded areas correspond to the one-σ deviation and the dashed lines depict an affine fit of the evolutions.
The second step is to extract the diffusion coefficient from the MSD time evolution. For this purpose, we used a linear regression to an affine curve. We let the algorithm find the most optimal start to the fit to minimize the uncertainty on the diffusion coefficient. We used this criterion to restrict the fit to the diffusion part omitting the ballistic section of the curve. In Fig. 3, we depicted the result of the fit by the dashed lines.
Integrated VACF and MSD give very consistent results as previously shown in Ref. 48. For instance, the cases plotted in Figs. 3 and 4 give a diffusion coefficient of m2/s for the MSD and m2/s for the integrated VACF. It is also interesting to note that we computed these quantities within an isokinetic ensemble which is not considered as the most well-suited ensemble for such calculations because it rescales the velocity components at each time step. It is known that the choice of thermostat is essential for the quality of the ionic transport properties estimates.49 Nevertheless, based on the comparison displayed in Ref. 28, it is clear that the isokinetic ensemble with our methodology gives very satisfactory results. It is possible that the cell size, the short chosen time steps, and the multi-origin technique help compensate for the crude thermalization associated with the isokinetic ensemble.
Normalized (top) and integrated (bottom) velocity autocorrelation function of aluminum nuclei in the Al1 case. In purple, the VACF was computed with a single origin, the blue case is with multiple origins. The shaded areas correspond to the one-σ deviation.
Normalized (top) and integrated (bottom) velocity autocorrelation function of aluminum nuclei in the Al1 case. In purple, the VACF was computed with a single origin, the blue case is with multiple origins. The shaded areas correspond to the one-σ deviation.
B. Viscosity
Normalized (top) and integrated (bottom) stress tensor autocorrelation function of gold nuclei in the Au1 case. The full line and shaded area correspond to the multiple origin calculation of the average SACF. The dashed line is the result of the fit presented in Eq. (10).
Normalized (top) and integrated (bottom) stress tensor autocorrelation function of gold nuclei in the Au1 case. The full line and shaded area correspond to the multiple origin calculation of the average SACF. The dashed line is the result of the fit presented in Eq. (10).
V. ELECTRONIC TRANSPORT PROPERTIES
Density functional theory within the Kohn–Sham–Mermin framework gives access to the electronic structure of the complete system. It, however, requires a higher level of convergence regarding the electronic density to offer reliable results. This is why we performed new DFT calculations on a series of snapshots extracted from the initial molecular dynamics. We used a higher number of bands to improve the convergence on the density of states and on the transition matrix elements of use in the Kubo–Greenwood formula (see below). For the MD calculations performed within the extended DFT framework,23 the snapshots were first converged with the extended approach then with the standard Kohn–Sham orbitals. The detailed parameters of the simulations are available in Table II. For all the cases reported in Ref. 28, we restricted ourselves to non spin-polarized calculations.
KS-DFT snapshots of input parameters and electronic transport property analysis. stands for the number of snapshots selected in the molecular dynamics trajectory, at least separated by the decorrelation time (Fig. 1). is the time between each snapshot, where is necessarily greater than the decorrelation time.
. | H1 . | C1 . | CH1 . | Al1 . | Cu1 . | CH2 . | Au1 . | C2 . |
---|---|---|---|---|---|---|---|---|
15 | 6 | 6 | 6 | 6 | 6 | 6 | 6 | |
(ps) | 0.121 | 0.193 | 0.242 | 1.209 | 0.605 | 0.0435 | 0.145 | 0.0726 |
256 H | 64 C | 48 H/48 C | 108 Al | 108 Cu | 48 H/48 C | 108 Au | 108 C | |
-point(s) | 23 | 33 | 33 | MP | 43 | 33 | 23 | 83 |
512 | 4096 | 2048 | 8000 | 4096 | 4096 | 12 288 | 1024 | |
(Ha) | 62.0 | 320.0 | 60.0 | 50.0 | 40.0 | 60.0 | 50.0 | 400.0 |
. | H1 . | C1 . | CH1 . | Al1 . | Cu1 . | CH2 . | Au1 . | C2 . |
---|---|---|---|---|---|---|---|---|
15 | 6 | 6 | 6 | 6 | 6 | 6 | 6 | |
(ps) | 0.121 | 0.193 | 0.242 | 1.209 | 0.605 | 0.0435 | 0.145 | 0.0726 |
256 H | 64 C | 48 H/48 C | 108 Al | 108 Cu | 48 H/48 C | 108 Au | 108 C | |
-point(s) | 23 | 33 | 33 | MP | 43 | 33 | 23 | 83 |
512 | 4096 | 2048 | 8000 | 4096 | 4096 | 12 288 | 1024 | |
(Ha) | 62.0 | 320.0 | 60.0 | 50.0 | 40.0 | 60.0 | 50.0 | 400.0 |
A. Density of states
Electronic density of states of aluminum for case Al1 ( g/cm3, eV). The zero of energy corresponds to the chemical potential. Low energy spikes correspond to 2s and 2p electron core shells, while higher energy shells are merged in the continuum, starting at eV.
Electronic density of states of aluminum for case Al1 ( g/cm3, eV). The zero of energy corresponds to the chemical potential. Low energy spikes correspond to 2s and 2p electron core shells, while higher energy shells are merged in the continuum, starting at eV.
B. Electronic and thermal conductivities
Electrical conductivity as a function of the excitation energy for aluminum (top, case Al1), copper (middle, case Cu1), and CH (bottom, case CH2). The full line and the shaded areas are the conductivity and one-sigma deviation as obtained from DFT. We fitted the results with a Drude fit [Eq. (16)], an exponent of the term [Eq. (17) labeled “Alpha-fit”], and a Drude–Smith expression [Eq. (18)].
Electrical conductivity as a function of the excitation energy for aluminum (top, case Al1), copper (middle, case Cu1), and CH (bottom, case CH2). The full line and the shaded areas are the conductivity and one-sigma deviation as obtained from DFT. We fitted the results with a Drude fit [Eq. (16)], an exponent of the term [Eq. (17) labeled “Alpha-fit”], and a Drude–Smith expression [Eq. (18)].
Thermal conductivity of CH (case CH2) at 0.9 g/cm3 and 7.8 eV as a function of the excitation energy. The DC value obtained in this case is within the error bar of a recent experimental value of 465 ± 233 W/m/K.55
Thermal conductivity of CH (case CH2) at 0.9 g/cm3 and 7.8 eV as a function of the excitation energy. The DC value obtained in this case is within the error bar of a recent experimental value of 465 ± 233 W/m/K.55
For the purpose of magneto-hydrodynamic simulations, one is interested in the low frequency or even the direct current value of the electrical and thermal conductivities. Due to the finite-size of the simulation cell, the energy spectrum of the Kohn–Sham orbitals is discretized and has a finite resolution, which prevents a direct calculation of the DC conductivities. Despite recent efforts made into computing more accurately the intra-band contributions,53 it is necessary to extrapolate the conductivity values toward the lowest frequencies.
For the thermal conductivity, it is possible to use a fit similar to Eq. (17). However, because it includes higher order Onsager coefficients, the thermal conductivity tends to be more continuous near the origin. Figure 8 shows an example for the equimolar CH mixture at a high temperature, around 7.8 eV, and a density of 0.9 g/cm3. By simply averaging the DC value obtained for each snapshot, we obtain a thermal conductivity of 650 ± 10 W/m/K. This value compares very well with the recent experimental determination of 465 ± 233 W/m/K,55 which shows the capabilities of the KS-DFT-MD method in computing accurate transport coefficients.
C. Optical properties
From the Onsager coefficients, it is also possible to extract optical properties. In particular, from the frequency-dependent electrical conductivity, one can compute the absorption coefficient α. A first step is to compute the imaginary part of the conductivity using a Kramers–Kronig relationship and to then determine the optical index and its related quantities (reflectivity and absorption coefficients).52 These quantities were not explicitly requested for the workshop but they are linked to the opacities, which are related to radiative transport. It is also important to note that during shock experiments it is very often an optical quantity that is measured and we can use ab initio results to relate the optical quantities to the DC conductivity.56 An example of the absorption coefficient is plotted in Fig. 9. It corresponds to the Al1 case. As a comparison, we also added the result obtained at eV and g/cm3. We can see that a decrease in the density by a factor of 10 results in a decrease in absorption by a factor of nearly 10 as well. One could also note the pronounced peak in the absorption near 5 eV, which is mainly due to transitions from the (semi-)bound states near the continuum to the continuum, the 3s states being partly recombined at low density.
Frequency-dependent absorption coefficient as a function of the incident photon energy for Al at different densities and eV. The result was averaged over six snapshots in the two cases.
Frequency-dependent absorption coefficient as a function of the incident photon energy for Al at different densities and eV. The result was averaged over six snapshots in the two cases.
VI. CONCLUSION
To summarize, KS-DFT molecular dynamics offers a unified approach to compute ionic and electronic transport properties as well as standard thermodynamic quantities. However, the final determination of these properties requires some additional assumptions and techniques to circumvent difficulties encountered when using this theoretical scheme. We bypassed the orbital wall encountered at high temperatures using the extended DFT method, allowing for multi-eV molecular dynamics simulations with comparable accuracy as conventional KS-DFT. Additionally, to bypass the lack of statistics when determining diffusion coefficients, and viscosity, we used the multiple origin technique. While thermodynamic quantities are not so sensible to finite size effects, this is not the case for electronic transport properties such as direct current and thermal conductivities. We used physically informed extrapolation methods to determine if the finite size effects are negligible or not. These different techniques coupled with the high accuracy of the KS-DFT molecular dynamics method showed accurate results, which we were able to compare to recent experiments and other codes via the second charged-particle transport code comparison workshop.
ACKNOWLEDGMENTS
We are very thankful to Robin Piron for fruitful discussions and to the three anonymous referees for their insightful comments.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
Augustin Blanchet: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Validation (equal); Visualization (equal); Writing – original draft (supporting); Writing – review & editing (equal). Vanina Recoules: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Methodology (supporting); Validation (equal); Visualization (equal); Writing – original draft (supporting); Writing – review & editing (equal). François Soubiran: Conceptualization (equal); Data curation (equal); Formal analysis (lead); Investigation (equal); Methodology (lead); Validation (lead); Visualization (lead); Writing – original draft (lead); Writing – review & editing (lead). Mikael Tacu: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Validation (equal); Visualization (equal); Writing – original draft (supporting); Writing – review & editing (equal).
DATA AVAILABILITY
The data that support the findings of this study are available from the corresponding author upon reasonable request.