Abstract
We present a methodology for accelerating the estimation of the free energy from path integral Monte Carlo simulations by considering an intermediate artificial reference system where interactions are inexpensive to evaluate numerically. Using the spherically averaged Ewald interaction as this intermediate reference system for the uniform electron gas, the interaction contribution for the free energy was evaluated up to 18 times faster than the Ewald-only method. Furthermore, an extrapolation technique with respect to the quantum statistics was tested and applied to alleviate the Fermion sign problem. Combining these two techniques enabled the evaluation of the free energy for a system of 1000 electrons, where both finite-size and statistical errors are below chemical accuracy. The general procedure can be applied to systems relevant for planetary and inertial confinement fusion modeling with low to moderate levels of quantum degeneracy.
The description of thermal systems of interacting Fermions is a cornerstone of our understanding for a wide range of quantum systems, including ultracold atoms, , quantum dots, , and dense plasmas. In particular, dense quantum plasmas are abundant in astrophysics, where they are found in gas giants − such as Jupiter, Saturn and some classes of exoplanets, and stars most notable in later stages of stellar evolution in the form of red giants, white dwarfs , and the atmospheres of neutron stars. , However, high-density plasmas are also central in human-made applications such as inertial confinement fusion (ICF) − and the synthesis of novel materials. In recent groundbreaking experiments, ICF implosions have exceeded the Lawson criteria and achieved capsule gain, a key step toward achieving energy production through the ICF concept.
A formidable regime of dense plasmas to model theoretically is warm dense matter (WDM), which is characterized by a complex interplay between interactions, quantum degeneracy, and thermal excitations. ,, All the previously mentioned effects must be taken into account as both r s – the ratio between the Wigner-Seitz radius and the Bohr radius – and θ – the ratio of the thermal excitation energy and the electronic Fermi energy – are of order unity, which characterizes the strength of interactions and quantum degeneracy, respectively. Therefore, there remain uncertainties in the fundamental properties of WDM, such as the equation of state (EOS) and transport properties, which limit predictive modeling of, for example, the Jovian interior , and ICF implosions.
The most widely used description for WDM systems is a hybrid method (DFT-MD), where electrons are described using density functional theory (DFT), , while ions are treated by molecular dynamics (MD). The fidelity of DFT strongly depends on the unknown exchange-correlation functional and practical calculations often resort to approximations based on the properties of the uniform electron gas (UEG). − Path integral Monte Carlo (PIMC) ,, provides suitable benchmark at finite temperature, since it is exact within the statistical error. However, for Fermionic systems, PIMC is limited by the Fermion sign problem (FSP) in the number of particles and the level of quantum degeneracy it can model. The FSP arises because all Fermionic observables are ratios where the denominator is the average sign S, which decreases exponentially with particle number and the inverse temperature. , This vanishing sign causes computations of large or cold systems to be dominated by statistical errors.
A large number of methods have been introduced to address the FSP, e.g., restricted PIMC, , permutation blocking PIMC, configuration PIMC, Fermionic PIMC , and methods based on the Wigner formulation of quantum mechanics, but no method is able to fully alleviate the FSP in the whole phase space without approximation. Here we consider the ξ-extrapolation method suggested by Xiong and Xiong, where an additional ξ parameter that continuously interpolates from the bosonic (ξ = 1) to the Fermionic (ξ = −1) limit is introduced. By introducing an empirical model for the ξ-dependence, calculations can be carried out in the FSP-free parameter regime and extrapolated to the Fermionic results, circumventing the exponential computational cost with respect to the particle number. , The validity of the extrapolation method is limited to weak and moderate degenerate systems (θ ≥ 1.0) where the statistical attraction and repulsion for bosons and Fermions, respectively, are not dominant. The extrapolation will fail if the structural properties between the bosonic and Fermionic systems are too different, and more formal attempts to understand the ξ-dependence in the ground state have been made recently. However, the method has been successfully applied at moderate degeneracy for the computation of energy, ,, static structure, ,− imaginary time correlation functions, , density response, , and the average sign itself.
The (Helmholtz) free energy is central for our understanding of thermal systems, for example it is directly related to the exchange-correlation functional in DFT ,, where finite-temperature corrections are key at intermediate temperatures, ,− but the free energy is also commonly used to investigate the stability of different phases. − As the free energy is a thermodynamic potential, a free energy parametrization automatically yields a self-consistent EOS where all thermodynamic properties are obtained through differentiation. So far, first-principles tabulations of the EOS have focused on energy and pressure, but semiempirical constructions commonly model the free energy. , By accelerating first-principles computations of the free energy, we are moving closer to reliable and internally consistent equation of state tables in the WDM regime.
In this letter, we present computations for the free energy of the spin-unpolarised UEG from PIMC with unprecedentedly large system sizes and low statistical errors. Using a combination of robust extrapolation techniques and the introduction of an intermediate reference system where interactions are computationally cheap, we are able to model N = 1000 electrons. The efficiency of this scheme allows us to evaluate the free energy to well within chemical accuracy (i.e., 1 kcal/mol ≈ 1.6 mHa). In the main text, we focus on the condition r s = 3.23 and θ = 1.0 characteristic of the electronic conditions possible to achieve in hydrogen jet experiments, − but the methodology is general and applicable for either bosons and not too degenerate Fermi systems. The complete analysis for the UEG at r s = 10 is given in the Supporting Information.
The partition function or the free energy is not a thermodynamical average per se, but relates to a volume in phase space. Therefore, the free energy is not readily available from an Monte Carlo (MC) or molecular dynamics (MD) simulation, and the thermodynamic integration (TI) , method or the adiabatic connection (AC) formula has traditionally been used for its computation. Both methods require multiple computations, e.g., with an interaction that can be smoothly turned from that of a reference systemcommonly the ideal systemto the target system. Moreover, the application of the AC method to inhomogeneous systems, such as the electronic problem in the external potential of a fixed ion configuration, poses an additional obstacle. Recently, Dornheim et al. introduced the extended ensemble technique in which the free energy differences between systems 1 and 2 can be directly computed. The extended partition function in question is
| 1 |
where Z i is the partition function of system i, and c is an arbitrary coefficient that is chosen to optimize the ergodicity. In the extended ensemble, the difference in free energy per particle f i between the two systems is directly related to the thermal averages in the extended ensemble ⟨·⟩ext via
| 2 |
where δ̂ i is one in system i and zero otherwise, k B T is the temperature in energy units, and N is the number of particles.
The Hamiltonian Ĥ η = K̂ + ηV̂ where K̂ is the kinetic energy operator and V̂ is the Ewald summation, , interpolates between the ideal (η = 0) and interacting systems (η = 1). By considering η = 0 and η = 1 for the two systems in eq along with exact results for noninteracting systems, , the free energy for bosons can be computed , in what we will refer to as the η-ensemble. However, for a large number of particles, it was found practically difficult to ergodically explore the entire extended ensemble due to the presence of configurations in the ideal system that are strongly suppressed in the interacting case. Therefore, multiple intermediate η-steps are introduced, a prevailing strategy in free energy calculations with substantially different configurational spaces. Structurally, the η-ensemble becomes reminiscent of the TI with the in-between steps. However, in the η-ensemble, η-values with a finite difference are considered, whereas in TI a continuous function of the coupling constant is integrated.
The η-ensemble is performed in the bosonic sector and is therefore FSP free, but a large number of intermediate η-steps will result in a prohibitive computational cost for accurate free energy calculations for large N. The majority of the computational cost in each MC step comes from the evaluation of the Ewald summation. To avoid this problem, we evaluate the η-ensemble using a nonphysical interaction or artificial interaction, V̂ → V̂ art, which is computationally cheap, and any error is corrected for in a second step henceforth referred to as the a-ensemble. The a-ensemble concerns the Hamiltonian
| 3 |
were a = 0 and a = 1 is used for the two systems in eq . If the physical interaction V̂ and the artificial one V̂ art are sufficiently similar, no intermediate a steps are required, as no substantial energy penalty is incurred when altering a. This procedure accelerates the computation as the majority of the data collection is performed with a fast artificial interaction, but it does not constitute any approximation as it can simply be viewed as establishing a transitional reference system, as is common practice when performing TI. ,,
The artificial interaction in question is in principle a free choice, but it should be both computationally efficient and close to the physical one to avoid unnecessary computations in the a-ensemble. Working with the Coulomb interaction, a variety of cutoff based approximations have been developed, which all could be used as the artificial interaction; see review by Fukuda and Nakamura and references therein. In this work, we have used the spherically averaged Ewald potential, , in particular the formulation by Yakub and Ronchi (YR) , that has been successfully applied in MD, MC , and PIMC, , and recently has attracted new theoretical interest. By construction, the YR interaction yields energies similar to the Ewald summation, and its simple algebraic structure makes it cheap to evaluate allowing for classical MC simulation with up to 106 particles. The new a-ensemble with the YR potential as the artificial interaction has been implemented in the ISHTAR code, which employs the canonical adaptation , of the worm algorithm. , All reported computations have been performed using the primitive factorization.
Up to this point, only the bosonic sector has been considered to avoid the FSP. To calculate the free energy with Fermi statistics f 1 rather than Bose statistics f 1 for a system with interaction 1, the sign S 1 in the corresponding system should be resolved:
| 4 |
Equation completes our methodology for free energy computations which is schematically shown in Figure highlighting the steps and Hamiltonians.
1.
Schematic showing the different systems used to compute the free energies f and the Hamiltonians Ĥ of each system. Arrows indicate an ensemble or sign computation to go between systems, and labels the associated free energy change Δf. Green arrows indicate computations which are computational cheap, while orange arrows indicate moderate cost either due to a costly Ewald computation or the FSP. Red arrows are severely affected by the FSP.
For the sign evaluation in eq , the above-mentioned ξ-extrapolation was used based on the functional form:
| 5 |
where the primary scaling with N and ξ is factored out, and the remaining function a S (N, ξ) shows only small deviations from being constant. Dornheim et al. successfully showed that the extrapolation from ξ = – 0.2 based on eq with a S (N, ξ) = a S (N) is highly accurate for θ = 1. Figure greatly extends the validation of this extrapolation method by considering a 2 orders of magnitude range for ξ for N ≤ 66. Validation of the method to substantially smaller ξ is crucial for modeling larger system sizes, since keeping ξN roughly constant maintains a resolvable sign. System sizes up to N = 1000 are investigated in Figure , and N ≥ 264 is needed to converge the finite-size effect to within the statistical error bars. This highlights the need to model large systems to approach the thermodynamic limit.
2.
Average sign S Ew for the UEG at r s = 3.23 and θ = 1.0, for different system sizes N and permutation weight ξ. Note that ξ < 0 and all values of S Ew are less than unity. The error bars, correspond to 95%-confidence intervals estimated from simulations with varying initial conditions. The extrapolation of the confidence interval assuming a S is independent of ξ is shown in the highlighted areas. The point from which the extrapolation is performed is described in the Supporting Information. Good agreement with the extrapolation is demonstrated, validating the computational model for S Ew.
The minor systematic error observed in the ξ-extrapolation with N = 14 is 0.3% and corresponds to a 0.05 mHa error in free energy. These errors are expected to decrease with the size of the system, where the permutation structure is less affected by boundary effects and self-exchanges; this makes the generalization of the corresponding free energy difference via the ξ-extrapolation more straightforward. This can be seen particularly well for the more strongly coupled case of r s = 10 shown in the Supporting Information.
The nonideal contribution to the Fermionic free energy is the exchange correlation free energy:
| 6 |
which in our accelerated scheme (second line) has three distinct contributions. The contribution of the η-ensemble with the artificial interaction Δf η,art , the correction from the a-ensemble Δf a,art‑Ew and the difference between the sign contribution for the interacting and noninteracting system Δf S,Ew – Δf S,id . In the standard Ewald-only approach, the first two contributions are given by a single term Δf η,Ew = Δf η,art + Δf a,art‑Ew . The origin of each term is also shown in Figure .
As both a conceptual and practical validation of the acceleration method, Figure shows the exchange correlation free energy computed both via the standard Ewald-only method and our accelerated scheme for N ≤ 30. The results cannot be distinguished from each other on the scale of Figure , and any deviation lies within the statistical error margins. As the system size increases, the accelerated method can perform up to 18 times as many MC steps as the standard method in a given time; see the Supporting Information for additional information. The additional MC samples reduce the statistical error but more crucially allow us to investigate larger system sizes.
3.
Finite size and finite P corrected exchange correlation free energies for the UEG at r s = 3.23 and θ = 1.0 shown as the difference from the GDSMFB parametrization with a value of f xc = −0.15529 Ha (Ref.). The remaining N dependence is a fraction of a percent. The computations have been performed in a variety of ways, using the physical interaction (Ewald-only) or accelerated method (Accelerated), with (ξ > −1) and without (ξ = −1) ξ-extrapolation, and varying number of propagators P. Overlapping data points are shown when P is reduced or extrapolation techniques are employed, demonstrating the correctness of the procedure. The dashed line is a fit on the form f xc (N) = c 0 + c 1 N –c 2 for N ≥ 30, where c 0, c 1 and c 2 are fitting coefficients. The extrapolated free energy is reduced by 0.2% compared to the reference and c 2 ≈ 1.3. Error bars as described in Figure .
In Figure , the exchange correlation free energy computations are scaled up to 1000 electrons using the accelerated method. To limit computational expense, a reduced number of imaginary time slices P is used to factorize the density matrix for large N. The finite P error has been systematically investigated for smaller N with Ps between 8 and 200 as demonstrated in the Supporting Information. Empirically, we find that the corresponding P-correction that connects a finite P to the limit of P –1 → 0 is independent of N, reflecting the local nature of the factorization error, which is ultimately due to the quantum delocalization of individual particles. The correction has been applied in Figure . As a further validation of the finite-P correction, duplicate data points are shown when P is reduced and the results are always within the statistical error.
To further reduce the size dependence of the free energy, the results in Figure have been finite-size corrected using the method introduced by Groth et al. (see further details in the Supporting Information). The finite-size correction is highly efficient and at the investigated condition removes 93% of the finite-size effect already at the smallest system used, resulting in a remaining finite-size error of the order of 1 mHa per electron. In Figure , it is shown that the surviving size-dependent error scales roughly linearly with N –1 (dashed black), and for the largest systems investigated, this error is expected to be one hundredth of a mHa. The results are within 0.3% of the GDSMFB parametrization computed by the adiabatic connection formula, well within the expected error margins of their parametrization. In conclusion, the finite size correction method by Groth et al. is highly efficient and virtually any remaining finite size error has been eliminated by reaching system sizes with 1000 electrons, now numerically feasible with our accelerated technique for free energy calculations.
The magnitude of each of the contributions to the exchange correlation free energy is shown in Figure (a). The dominant contribution under the investigated condition is the interaction contribution from the η-ensemble (Δf η,YR ) followed by the sign contribution (Δf S,Ew – Δf S,id ). The correction for using the artificial interaction in the η-ensemble (Δf a,YR‑Ew ) is 3 orders of magnitude smaller than the overall contribution of interactions and vanishes with increasing system size, as the YR potential tend to the Coulomb form. This highlights the efficiency of the YR interaction in mimicking the full Ewald summation with respect to energy, even if some artifacts are present for spatially resolved quantities. The Δf a,YR‑Ew correction is small enough for the present system that it could have been neglected for the computations at larger N, but the a-ensemble will have a substantial contribution for more ordered systems, for other choices of artificial interaction, and for systems with different V̂. For the r s = 10 system, the picture is broadly the same, but the interaction contribution is even more dominant for this strongly interacting case.
4.
Size (a) and scaling (b) of the N dependent free energy for the UEG at r s = 3.23 and θ = 1.0. The correction term Δf a,YR-Ew and FSC vanish as N → ∞ while the contribution from the η-ensemble and the sign converges to a finite value. The subtraction of the infinite system size contribution N = ∞ in (b) was facilitated by a fit for N ≥ 66 on the form Δf = d 0 + d 1 N –d 2 , the same functional form used by Demyanov and Levashov. The interaction components are seen to converge sublinearly (d 2 < 1). Error bars based on statistical error as described in Figure .
The approach to the thermodynamic limit for the three contributions to the free energy is highlighted in Figure (b) by subtracting the (fitted) thermodynamic limit. The size of the finite N errors generally follows the magnitude of each respective term. The finite size error in the sign, which is by far the hardest contribution to compute in practice, is seen to scale linearly with particle number; this might be exploited for further extrapolation and optimization in future works. The two interaction contributions scale sublinearly, with approximate exponents of 0.7 and 0.85, respectively. However, these two exponents are not universal as they increase for the r s = 10 conditions. In the Supporting Information, the sublinear scaling is discussed in terms of the finite-size correction model. For the simulation with N = 1000, all finite-size errors are below 1 mHa and the chemical accuracy is reached even without any finite-size correction procedure.
To summarize, we have introduced and exemplified the use of an accelerated method for free energy estimation based on ab initio PIMC. The method accelerates the computation in two primary ways. First, an intermediate “artificial” reference system is introduced in which interactions are numerically evaluated more efficiently. The majority of interaction effects can be captured in this artificial system, and any remaining error can be corrected by the introduced a-ensemble which in our work only required a single computation with the numerically more costly physical interaction. In this work, the use of the artificial interaction reduced the computation effort by a factor of up to 18 for the interaction contribution. Second, a ξ-extrapolation methodology is employed to resolve the sign for larger system sizes that are otherwise prevented by the Fermionic sign problem. This extrapolation was shown to be accurate to 0.3% over 2 orders of magnitude in ξ for θ = 1. The generality of the procedure was demonstrated by considering two different density conditions.
Accelerating the calculation of free energies paves the way for scaling up computations to remove the final systematic error – the finite size effects – at warm dense matter conditions. The presented method can be combined with other acceleration techniques to consider even large systems, e.g., employing GPUs, hierarchical energy evaluation, and contraction schemes. High-precision free energy estimates for the UEG open for the possibility to explore a potential spin phase transition at finite temperature, which have been intensely studied in the ground state. , Future work might also explore the long-wavelength physics with the presented method via the density stiffness theorem, , which relates the static linear and nonlinear density response to free energy differences. In this regard, the simulation of large systems is crucial to study the optical limit of |k⃗| → 0, where the minimum wavenumber |k⃗| = 2π/L is determined by the box length L. Lastly, the present study focuses on the UEG, but it is straightforward to apply our methodology to light elements such as hydrogen and beryllium to inform planetary and inertial confinement fusion modeling. Moreover, our approach can easily be applied to the simulation of inhomogeneous systems such as electrons in a fixed ionic configuration, which might be of great value for the benchmarking of DFT and potentially even for the data-driven construction of novel exchange correlation functionals.
Supplementary Material
Acknowledgments
This work was partially supported by the Center for Advanced Systems Understanding (CASUS), financed by Germany’s Federal Ministry of Education and Research and the Saxon state government out of the State budget approved by the Saxon State Parliament. Further support is acknowledged for the CASUS Open Project Guiding dielectric theories with ab initio quantum Monte Carlo simulations: from the strongly coupled electron liquid to warm dense matter. This work has received funding from the European Research Council (ERC) under the European Union’s Horizon 2022 research and innovation programme (Grant agreement No. 101076233, “PREXTREME”). Views and opinions expressed are however those of the authors only and do not necessarily reflect those of the European Union or the European Research Council Executive Agency. Neither the European Union nor the granting authority can be held responsible for them. Computations were performed on a Bull Cluster at the Center for Information Services and High-Performance Computing (ZIH) at Technische Universität Dresden and at the Norddeutscher Verbund für Hoch- und Höchstleistungsrechnen (HLRN) under grant mvp00024.
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jpclett.5c02193.
The authors declare no competing financial interest.
References
- Giorgini S., Pitaevskii L. P., Stringari S.. Theory of ultracold atomic Fermi gases. Rev. Mod. Phys. 2008;80:1215. doi: 10.1103/RevModPhys.80.1215. [DOI] [Google Scholar]
- Ceperley D. M.. Path-integral calculations of normal liquid He 3. Phys. Rev. Lett. 1992;69:331. doi: 10.1103/PhysRevLett.69.331. [DOI] [PubMed] [Google Scholar]
- Reimann S. M., Manninen M.. Electronic structure of quantum dots. Rev. Mod. Phys. 2002;74:1283. doi: 10.1103/RevModPhys.74.1283. [DOI] [Google Scholar]
- Dornheim T., Yan Y.. Abnormal quantum moment of inertia and structural properties of electrons in 2D and 3D quantum dots: An ab initio path-integral Monte Carlo study. New J. Phys. 2022;24:113024. doi: 10.1088/1367-2630/ac9f29. [DOI] [Google Scholar]
- Lorenzen, W. ; Becker, A. , Redmer, R. In Frontiers and Challenges in Warm Dense Matter; Graziani, F. , Desjarlais, M. P. , Redmer, R. , Trickey, S. B. , Eds.; Springer International Publishing: 2014; pp 203–234. [Google Scholar]
- Guillot T.. Interiors of giant planets inside and outside the solar system. Science. 1999;286:72–77. doi: 10.1126/science.286.5437.72. [DOI] [PubMed] [Google Scholar]
- Fortney J. J., Nettelmann N.. The Interior Structure, Composition, and Evolution of Giant Planets. Space Science Reviews. 2010;152:423–447. doi: 10.1007/s11214-009-9582-x. [DOI] [Google Scholar]
- Helled R., Mazzola G., Redmer R.. Understanding dense hydrogen at planetary conditions. Nature Reviews Physics. 2020;2:562–574. doi: 10.1038/s42254-020-0223-3. [DOI] [Google Scholar]
- French M., Becker A., Lorenzen W., Nettelmann N., Bethkenhagen M., Wicht J., Redmer R.. Ab initio simulations for material properties along the Jupiter adiabat. Astrophysical Journal Supplement Series. 2012;202:5. doi: 10.1088/0067-0049/202/1/5. [DOI] [Google Scholar]
- Preising M., French M., Mankovich C., Soubiran F., Redmer R.. Material properties of Saturn’s interior from ab initio simulations. Astrophysical Journal Supplement Series. 2023;269:47. doi: 10.3847/1538-4365/ad0293. [DOI] [Google Scholar]
- Kippenhahn, R. ; Weigert, A. ; Weiss, A. . Stellar Structure and Evolution, 2nd ed.; Springer: 2012. [Google Scholar]
- Fortov V. E.. Extreme states of matter on Earth and in space. Physics-Uspekhi. 2009;52:615. doi: 10.3367/UFNe.0179.200906h.0653. [DOI] [Google Scholar]
- Chandrasekhar S.. XLVIII. The density of white dwarf stars. London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science. 1931;11:592–596. doi: 10.1080/14786443109461710. [DOI] [Google Scholar]
- Chabrier G., Brassard P., Fontaine G., Saumon D.. Cooling sequences and color-magnitude diagrams for cool white dwarfs with hydrogen atmospheres. Astrophysical Journal. 2000;543:216. doi: 10.1086/317092. [DOI] [Google Scholar]
- Gudmundsson E. H., Pethick C. J., Epstein R. I.. Structure of neutron star envelopes. Astrophys. J. 1983;272:286–300. doi: 10.1086/161292. [DOI] [Google Scholar]
- Haensel, P. ; Potekhin, A. Y. ; Yakovlev, D. G. . Neutron Stars 1: Equation of State and Structure; Springer: 2007. [Google Scholar]
- Nuckolls J., Wood L., Thiessen A., Zimmerman G.. Laser compression of matter to super-high densities: Thermonuclear (CTR) applications. Nature. 1972;239:139–142. doi: 10.1038/239139a0. [DOI] [Google Scholar]
- Betti R., Hurricane O. A.. Inertial-confinement fusion with lasers. Nat. Phys. 2016;12:435–448. doi: 10.1038/nphys3736. [DOI] [Google Scholar]
- Hurricane O. A., Patel P. K., Betti R., Froula D. H., Regan S. P., Slutz S. A., Gomez M. R., Sweeney M. A.. Physics principles of inertial confinement fusion and US program overview. Rev. Mod. Phys. 2023;95:025005. doi: 10.1103/RevModPhys.95.025005. [DOI] [Google Scholar]
- Miao M., Sun Y., Zurek E., Lin H.. Chemistry under high pressure. Nature Reviews Chemistry. 2020;4:508–527. doi: 10.1038/s41570-020-0213-0. [DOI] [Google Scholar]
- Abu-Shawareb H., Acree R., Adams P., Adams J., Addis B., Aden R., Adrian P., Afeyan B. B., Aggleton M., Aghaian L.. et al. Lawson criterion for ignition exceeded in an inertial fusion experiment. Phys. Rev. Lett. 2022;129:075001. doi: 10.1103/PhysRevLett.129.075001. [DOI] [PubMed] [Google Scholar]
- Bonitz M., Dornheim T., Moldabekov Z. A., Zhang S., Hamann P., Kählert H., Filinov A., Ramakrishna K., Vorberger J.. Ab initio simulation of warm dense matter. Physics of Plasmas. 2020;27:042710. doi: 10.1063/1.5143225. [DOI] [Google Scholar]
- Vorberger, J. ; Graziani, F. ; Riley, D. ; Baczewski, A. D. ; Baraffe, I. ; Bethkenhagen, M. ; Blouin, S. ; Böhme, M. P. ; Bonitz, M. ; Bussmann, M. . et al. Roadmap for warm dense matter physics. 2025; https://arxiv.org/abs/2505.02494 (accessed 2025-08-28).
- Wahl S. M., Hubbard W. B., Militzer B., Guillot T., Miguel Y., Movshovitz N., Kaspi Y., Helled R., Reese D., Galanti E.. et al. Comparing Jupiter interior structure models to Juno gravity measurements and the role of a dilute core. Geophys. Res. Lett. 2017;44:4649–4659. doi: 10.1002/2017GL073160. [DOI] [Google Scholar]
- Hu S. X., Goncharov V. N., Boehly T. R., McCrory R. L., Skupsky S., Collins L. A., Kress J. D., Militzer B.. Impact of first-principles properties of deuterium–tritium on inertial confinement fusion target designs. Physics of Plasmas. 2015;22:056304. doi: 10.1063/1.4917477. [DOI] [Google Scholar]
- Bonitz M., Vorberger J., Bethkenhagen M., Böhme M., Ceperley D., Filinov A., Gawne T., Graziani F., Gregori G., Hamann P.. et al. Toward first principles-based simulations of dense hydrogen. Physics of Plasmas. 2024;31:110501. doi: 10.1063/5.0219405. [DOI] [Google Scholar]
- Hohenberg P., Kohn W.. Inhomogeneous electron gas. Phys. Rev. 1964;136:B864. doi: 10.1103/PhysRev.136.B864. [DOI] [Google Scholar]
- Mermin N. D.. Thermal properties of the inhomogeneous electron gas. Phys. Rev. 1965;137:A1441. doi: 10.1103/PhysRev.137.A1441. [DOI] [Google Scholar]
- Rapaport, D. C. The Art of Molecular Dynamics Simulation; Cambridge University Press: 2004. [Google Scholar]
- Perdew J. P., Zunger A.. Self-interaction correction to density-functional approximations for many-electron systems. Phys. Rev. B. 1981;23:5048. doi: 10.1103/PhysRevB.23.5048. [DOI] [Google Scholar]
- Karasiev V. V., Sjostrom T., Dufty J., Trickey S.. Accurate homogeneous electron gas exchange-correlation free energy for local spin-density calculations. Phys. Rev. Lett. 2014;112:076403. doi: 10.1103/PhysRevLett.112.076403. [DOI] [PubMed] [Google Scholar]
- Dornheim T., Groth S., Bonitz M.. The uniform electron gas at warm dense matter conditions. Phys. Rep. 2018;744:1–86. doi: 10.1016/j.physrep.2018.04.001. [DOI] [PubMed] [Google Scholar]
- Ceperley D. M.. Path integrals in the theory of condensed helium. Rev. Mod. Phys. 1995;67:279. doi: 10.1103/RevModPhys.67.279. [DOI] [Google Scholar]
- Böhme M., Moldabekov Z. A., Vorberger J., Dornheim T.. Static electronic density response of warm dense hydrogen: Ab initio path integral Monte Carlo Simulations. Phys. Rev. Lett. 2022;129:066402. doi: 10.1103/PhysRevLett.129.066402. [DOI] [PubMed] [Google Scholar]
- Dornheim T.. Fermion sign problem in path integral Monte Carlo simulations: Quantum dots, ultracold atoms, and warm dense matter. Phys. Rev. E. 2019;100:023307. doi: 10.1103/PhysRevE.100.023307. [DOI] [PubMed] [Google Scholar]
- Troyer M., Wiese U. J.. Computational Complexity and Fundamental Limitations to Fermionic Quantum Monte Carlo Simulations. Phys. Rev. Lett. 2005;94:170201. doi: 10.1103/PhysRevLett.94.170201. [DOI] [PubMed] [Google Scholar]
- Hatano N.. Data analysis for quantum Monte Carlo simulations with the negative-sign problem. J. Phys. Soc. Jpn. 1994;63:1691–1697. doi: 10.1143/JPSJ.63.1691. [DOI] [Google Scholar]
- Brown E. W., Clark B. K., DuBois J. L., Ceperley D. M.. Path-integral Monte Carlo simulation of the warm dense homogeneous electron gas. Phys. Rev. Lett. 2013;110:146405. doi: 10.1103/PhysRevLett.110.146405. [DOI] [PubMed] [Google Scholar]
- Dornheim T., Groth S., Filinov A., Bonitz M.. Permutation blocking path integral Monte Carlo: A highly efficient approach to the simulation of strongly degenerate non-ideal fermions. New J. Phys. 2015;17:073017. doi: 10.1088/1367-2630/17/7/073017. [DOI] [Google Scholar]
- Schoof T., Groth S., Vorberger J., Bonitz M.. Ab initio thermodynamic results for the degenerate electron gas at finite temperature. Phys. Rev. Lett. 2015;115:130402. doi: 10.1103/PhysRevLett.115.130402. [DOI] [PubMed] [Google Scholar]
- Filinov V., Levashov P., Larkin A.. Monte Carlo simulations of the electron short-range quantum ordering in Coulomb systems and the ‘fermionic sign problem’. Journal of Physics A: Mathematical and Theoretical. 2022;55:035001. doi: 10.1088/1751-8121/ac40e3. [DOI] [Google Scholar]
- Filinov V. S., Syrovatka R. S., Levashov P. R.. Solution of the ‘sign problem’in the path integral Monte Carlo simulations of strongly correlated Fermi systems: Thermodynamic properties of helium-3. Mol. Phys. 2022;120:e2102549. doi: 10.1080/00268976.2022.2102549. [DOI] [Google Scholar]
- Larkin A. S., Filinov V. S., Fortov V. E.. Peculiarities of the momentum distribution functions of strongly correlated charged fermions. Journal of Physics A: Mathematical and Theoretical. 2018;51:035002. doi: 10.1088/1751-8121/aa98d0. [DOI] [Google Scholar]
- Xiong Y., Xiong H.. On the thermodynamic properties of fictitious identical particles and the application to fermion sign problem. J. Chem. Phys. 2022;157:094112. doi: 10.1063/5.0106067. [DOI] [PubMed] [Google Scholar]
- Dornheim T., Tolias P., Groth S., Moldabekov Z. A., Vorberger J., Hirshberg B.. Fermionic physics from ab initio path integral Monte Carlo simulations of fictitious identical particles. J. Chem. Phys. 2023;159:164113. doi: 10.1063/5.0171930. [DOI] [PubMed] [Google Scholar]
- Dornheim T., Schwalbe S., Moldabekov Z. A., Vorberger J., Tolias P.. Ab initio path integral Monte Carlo simulations of the uniform electron gas on large length scales. The. J. Phys. Chem. Lett. 2024;15:1305–1313. doi: 10.1021/acs.jpclett.3c03193. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huang, K. Statistical Mechanics, 2nd ed.; John Wiley and Sons: 1987. [Google Scholar]
- He, R.-C. ; Zeng, J.-X. ; Yang, S. ; Wang, C. ; Ye, Q.-J. ; Li, X.-Z. . Revisiting the Fermion Sign Problem from the Structure of Lee-Yang Zeros. I. The Form of Partition Function for Indistinguishable Particles and Its Zeros at 0 K. 2025; https://arxiv.org/abs/2507.22779, (accessed 2025-09-02).
- Xiong Y., Xiong H.. Thermodynamics of fermions at any temperature based on parametrized partition function. Phys. Rev. E. 2023;107:055308. doi: 10.1103/PhysRevE.107.055308. [DOI] [PubMed] [Google Scholar]
- Morresi T., Garberoglio G.. Normal liquid 3He studied by path-integral Monte Carlo with a parametrized partition function. Phys. Rev. B. 2025;111:014521. doi: 10.1103/PhysRevB.111.014521. [DOI] [Google Scholar]
- Dornheim T., Schwalbe S., Moldabekov Z. A., Vorberger J., Tolias P.. Ab initio path integral Monte Carlo simulations of the uniform electron gas on large length scales. The. J. Phys. Chem. Lett. 2024;15:1305–1313. doi: 10.1021/acs.jpclett.3c03193. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dornheim T., Schwalbe S., Böhme M. P., Moldabekov Z. A., Vorberger J., Tolias P.. Ab initio path integral Monte Carlo simulations of warm dense two-component systems without fixed nodes: Structural properties. J. Chem. Phys. 2024;160:164111. doi: 10.1063/5.0206787. [DOI] [PubMed] [Google Scholar]
- Dornheim T., Döppner T., Tolias P., Böhme M. P., Fletcher L. B., Gawne T., Graziani F. R., Kraus D., MacDonald M. J., Moldabekov Z. A.. et al. Unraveling electronic correlations in warm dense quantum plasmas. Nat. Commun. 2025;16:5103. doi: 10.1038/s41467-025-60278-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dornheim T., Schwalbe S., Tolias P., Böhme M. P., Moldabekov Z. A., Vorberger J.. Ab initio density response and local field factor of warm dense hydrogen. Matter and Radiation at Extremes. 2024;9:057401. doi: 10.1063/5.0211407. [DOI] [Google Scholar]
- Dornheim T., Moldabekov Z., Schwalbe S., Tolias P., Vorberger J.. Fermionic free energies from ab initio path integral Monte Carlo simulations of fictitious identical particles. J. Chem. Theory Comput. 2025;21:7290–7303. doi: 10.1021/acs.jctc.5c00301. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gupta U., Rajagopal A. K.. Density functional formalism at finite temperatures with some applications. Phys. Rep. 1982;87:259–311. doi: 10.1016/0370-1573(82)90077-1. [DOI] [Google Scholar]
- Smith, J. C. ; Sagredo, F. ; Burke, K. In Frontiers of Quantum Chemistry; Wójcik, M. J. , Nakatsuji, H. , Kirtman, B. , Ozaki, Y. , Eds.; Springer Singapore: 2018; pp 249–271. [Google Scholar]
- Sjostrom T., Daligault J.. Gradient corrections to the exchange-correlation free energy. Phys. Rev. B. 2014;90:155109. doi: 10.1103/PhysRevB.90.155109. [DOI] [Google Scholar]
- Karasiev V. V., Calderín L., Trickey S. B.. Importance of finite-temperature exchange correlation for warm dense matter calculations. Phys. Rev. E. 2016;93:063207. doi: 10.1103/PhysRevE.93.063207. [DOI] [PubMed] [Google Scholar]
- Ramakrishna K., Dornheim T., Vorberger J.. Influence of finite temperature exchange-correlation effects in hydrogen. Phys. Rev. B. 2020;101:195129. doi: 10.1103/PhysRevB.101.195129. [DOI] [Google Scholar]
- Alfe D., Gillan M. J., Price G. D.. The melting curve of iron at the pressures of the Earth’s core from ab initio calculations. Nature. 1999;401:462–464. doi: 10.1038/46758. [DOI] [Google Scholar]
- Wilson H. F., Militzer B.. Solubility of water ice in metallic hydrogen: Consequences for core erosion in gas giant planets. Astrophysical Journal. 2012;745:54. doi: 10.1088/0004-637X/745/1/54. [DOI] [Google Scholar]
- Wu J., González-Cataldo F., Militzer B.. High-pressure phase diagram of beryllium from ab initio free-energy calculations. Phys. Rev. B. 2021;104:014103. doi: 10.1103/PhysRevB.104.014103. [DOI] [Google Scholar]
- Militzer B., González-Cataldo F., Zhang S., Driver K. P., Soubiran F.. First-principles equation of state database for warm dense matter computation. Phys. Rev. E. 2021;103:013203. doi: 10.1103/PhysRevE.103.013203. [DOI] [PubMed] [Google Scholar]
- More R. M., Warren K., Young D., Zimmerman G.. A new quotidian equation of state (QEOS) for hot dense matter. Phys. Fluids. 1988;31:3059–3078. doi: 10.1063/1.866963. [DOI] [Google Scholar]
- Lyon, S. P. ; Johnson, J. D. . Technical Report No. LA-UR-92-3407; Los Alamos National Laboratory: 1995. [Google Scholar]
- Pople J. A.. Nobel lecture: Quantum chemical models. Rev. Mod. Phys. 1999;71:1267. doi: 10.1103/RevModPhys.71.1267. [DOI] [Google Scholar]
- Zastrau U., Appel K., Baehtz C., Baehr O., Batchelor L., Berghäuser A., Banjafar M., Brambrink E., Cerantola V., Cowan T. E.. et al. The High Energy Density Scientific Instrument at the European XFEL. Journal of Synchrotron Radiation. 2021;28:1393–1416. doi: 10.1107/S1600577521007335. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fletcher L. B., Vorberger J., Schumaker W., Ruyer C., Goede S., Galtier E., Zastrau U., Alves E. P., Baalrud S. D., Baggott R. A.. et al. Electron-ion temperature relaxation in warm dense hydrogen observed with picosecond resolved x-ray scattering. Frontiers in Physics. 2022;10:838524. doi: 10.3389/fphy.2022.838524. [DOI] [Google Scholar]
- Hamann P., Kordts L., Filinov A., Bonitz M., Dornheim T., Vorberger J.. Prediction of a roton-type feature in warm dense hydrogen. Phys. Rev. Res. 2023;5:033039. doi: 10.1103/PhysRevResearch.5.033039. [DOI] [Google Scholar]
- Frenkel, D. ; Smit, B. . Understanding Molecular Simulation: From Algorithms to Applications, 2nd ed.; Academic Press: 2002; Chapter Free Energy Calculations, pp 167–200. [Google Scholar]
- Zwanzig R. W.. High-temperature equation of state by a perturbation method. I. Nonpolar gases. J. Chem. Phys. 1954;22:1420–1426. doi: 10.1063/1.1740409. [DOI] [Google Scholar]
- Pribram-Jones A., Grabowski P. E., Burke K.. Thermal density functional theory: Time-dependent linear response and approximate functionals from the fluctuation-dissipation theorem. Phys. Rev. Lett. 2016;116:233001. doi: 10.1103/PhysRevLett.116.233001. [DOI] [PubMed] [Google Scholar]
- Dornheim T., Moldabekov Z. A., Schwalbe S., Vorberger J.. Direct free energy calculation from ab initio path integral Monte Carlo simulations of warm dense matter. Phys. Rev. B. 2025;111:L041114. doi: 10.1103/PhysRevB.111.L041114. [DOI] [Google Scholar]
- Ewald P. P.. Die Berechnung optischer und elektrostatischer Gitterpotentiale. Annalen der Physik. 1921;369:253–287. doi: 10.1002/andp.19213690304. [DOI] [Google Scholar]
- Zhou C.-C., Dai W.-S.. Canonical partition functions: ideal quantum gases, interacting classical gases, and interacting quantum gases. Journal of Statistical Mechanics: Theory and Experiment. 2018;2018:023105. doi: 10.1088/1742-5468/aaa37e. [DOI] [Google Scholar]
- Barghathi H., Yu J., Del Maestro A.. Theory of noninteracting fermions and bosons in the canonical ensemble. Physical Review Research. 2020;2:043206. doi: 10.1103/PhysRevResearch.2.043206. [DOI] [Google Scholar]
- Dornheim T., Tolias P., Moldabekov Z. A., Vorberger J.. η-ensemble path integral Monte Carlo approach to the free energy of the warm dense electron gas and the uniform electron liquid. Physical Review Research. 2025;7:023250. doi: 10.1103/4n7x-78fs. [DOI] [Google Scholar]
- Caillol J. M., Gilles D.. Monte Carlo simulations of the Yukawa one-component plasma. J. Stat. Phys. 2000;100:933–947. doi: 10.1023/A:1018727428374. [DOI] [Google Scholar]
- Plummer D., Svensson P., Gericke D. O., Hollebon P., Vinko S. M., Gregori G.. Ionization calculations using classical molecular dynamics. Phys. Rev. E. 2025;111:015204. doi: 10.1103/PhysRevE.111.015204. [DOI] [PubMed] [Google Scholar]
- Fukuda I., Nakamura H.. Non-Ewald methods: theory and applications to molecular systems. Biophysical Reviews. 2012;4:161–170. doi: 10.1007/s12551-012-0089-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Angelie, A. Monte-Carlo calculations of microfield distributions in a plasma. Colloquium on collisions and radiation, Orleans, France, 1983; https://inis.iaea.org/api/records/4bsv9-02263/files/15026316.pdf/content (accessed 2025-08-28).
- Caillol J. M.. Numerical simulations of Coulomb systems: A comparison between hyperspherical and periodic boundary conditions. J. Chem. Phys. 1999;111:6528–6537. doi: 10.1063/1.479947. [DOI] [Google Scholar]
- Yakub E., Ronchi C.. An efficient method for computation of long-ranged Coulomb forces in computer simulation of ionic fluids. J. Chem. Phys. 2003;119:11556–11560. doi: 10.1063/1.1624364. [DOI] [Google Scholar]
- Yakub E., Ronchi C.. A new method for computation of long ranged Coulomb forces in computer simulation of disordered systems. Journal of Low Temperature Physics. 2005;139:633–643. doi: 10.1007/s10909-005-5451-5. [DOI] [Google Scholar]
- Yakub E., Ronchi C., Staicu D.. Molecular dynamics simulation of premelting and melting phase transitions in stoichiometric uranium dioxide. J. Chem. Phys. 2007;127:094508. doi: 10.1063/1.2764484. [DOI] [PubMed] [Google Scholar]
- Demyanov G. S., Levashov P. R.. One-component plasma of a million particles via angular-averaged Ewald potential: A Monte Carlo study. Phys. Rev. E. 2022;106:015204. doi: 10.1103/PhysRevE.106.015204. [DOI] [PubMed] [Google Scholar]
- Demyanov G., Onegin A., Levashov P.. N-convergence in one–component plasma: Comparison of Coulomb, Ewald, and angular-averaged Ewald potentials. Contributions to Plasma Physics. 2024;64:e202300164. doi: 10.1002/ctpp.202300164. [DOI] [Google Scholar]
- Filinov V. S., Larkin A. S., Levashov P. R.. Uniform electron gas at finite temperature by fermionic-path-integral Monte Carlo simulations. Phys. Rev. E. 2020;102:033203. doi: 10.1103/PhysRevE.102.033203. [DOI] [PubMed] [Google Scholar]
- Dornheim T., Chuna T. M., Bellenbaum H. M., Moldabekov Z., Tolias P., Vorberger J.. Application of a spherically averaged pair potential in ab initio path integral Monte Carlo simulations of the warm dense electron gas. Phys. Rev. E. 2025;112:035203. doi: 10.1103/lj9c-bh48. [DOI] [Google Scholar]
- Demyanov G. S., Levashov P. R.. Systematic derivation of angular-averaged Ewald potential. Journal of Physics A: Mathematical and Theoretical. 2022;55:385202. doi: 10.1088/1751-8121/ac870b. [DOI] [Google Scholar]
- Dornheim, T. ; Böhme, M. ; Schwalbe, S. . ISHTAR - Imaginary-time Stochastic High-performance Tool for Ab initio Research. 2024; 10.5281/zenodo.10497098 (accessed 2025-08-28). [DOI]
- Mezzacapo F., Boninsegni M.. Structure, superfluidity, and quantum melting of hydrogen clusters. Phys. Rev. A. 2007;75:033201. doi: 10.1103/PhysRevA.75.033201. [DOI] [PubMed] [Google Scholar]
- Dornheim T., Böhme M., Militzer B., Vorberger J.. Ab initio path integral Monte Carlo approach to the momentum distribution of the uniform electron gas at finite temperature without fixed nodes. Phys. Rev. B. 2021;103:205142. doi: 10.1103/PhysRevB.103.205142. [DOI] [Google Scholar]
- Boninsegni M., Prokof’ev N., Svistunov B.. Worm algorithm for continuous-space path integral Monte Carlo simulations. Phys. Rev. Lett. 2006;96:070601. doi: 10.1103/PhysRevLett.96.070601. [DOI] [PubMed] [Google Scholar]
- Boninsegni M., Prokof’ev N. V., Svistunov B. V.. Worm algorithm and diagrammatic Monte Carlo: A new approach to continuous-space path integral Monte Carlo simulations. Phys. Rev. E. 2006;74:036701. doi: 10.1103/PhysRevE.74.036701. [DOI] [PubMed] [Google Scholar]
- Dornheim T., Groth S., Filinov A. V., Bonitz M.. Path integral Monte Carlo simulation of degenerate electrons: Permutation-cycle properties. J. Chem. Phys. 2019;151:014108. doi: 10.1063/1.5093171. [DOI] [PubMed] [Google Scholar]
- Groth S., Dornheim T., Sjostrom T., Malone F. D., Foulkes W. M. C., Bonitz M.. Ab initio exchange-correlation free energy of the uniform electron gas at warm dense matter conditions. Phys. Rev. Lett. 2017;119:135001. doi: 10.1103/PhysRevLett.119.135001. [DOI] [PubMed] [Google Scholar]
- Thorpe, B. ; Smith, M. J. ; Hasnip, P. J. ; Drummond, N. D. . Acceleration of the CASINO quantum Monte Carlo software using graphics processing units and OpenACC. 2025; https://arxiv.org/abs/2507.02888.
- Müller F., Christiansen H., Schnabel S., Janke W.. Fast, hierarchical, and adaptive algorithm for Metropolis Monte Carlo simulations of long-range interacting systems. Physical Review X. 2023;13:031006. doi: 10.1103/PhysRevX.13.031006. [DOI] [Google Scholar]
- John C., Spura T., Habershon S., Kühne T. D.. Quantum ring-polymer contraction method: Including nuclear quantum effects at no additional computational cost in comparison to ab initio molecular dynamics. Phys. Rev. E. 2016;93:043305. doi: 10.1103/PhysRevE.93.043305. [DOI] [PubMed] [Google Scholar]
- Ceperley D. M., Alder B. J.. Ground state of the electron gas by a stochastic method. Phys. Rev. Lett. 1980;45:566. doi: 10.1103/PhysRevLett.45.566. [DOI] [Google Scholar]
- Loos P.-F., Gill P. M. W.. The uniform electron gas. Wiley Interdisciplinary Reviews: Computational Molecular Science. 2016;6:410–429. doi: 10.1002/wcms.1257. [DOI] [Google Scholar]
- Giuliani, G. ; Vignale, G. . Quantum Theory of the Electron Liquid; Cambridge University Press: 2008. [Google Scholar]
- Dornheim T., Tolias P., Moldabekov Z. A., Vorberger J.. Energy response and spatial alignment of the perturbed electron gas. J. Chem. Phys. 2023;158:164108. doi: 10.1063/5.0146503. [DOI] [PubMed] [Google Scholar]
- Kasim M. F., Vinko S. M.. Learning the Exchange-Correlation Functional from Nature with Fully Differentiable Density Functional Theory. Phys. Rev. Lett. 2021;127:126403. doi: 10.1103/PhysRevLett.127.126403. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.






