Abstract
Current studies show that oxygen does not aggregate into a polymeric phase even under pressures up to 10 TPa. To address the critical knowledge gap in understanding dense oxygen, here we show the complete polymerization process of oxygen, by using structure prediction methods. We determine the crystal structures of oxygen up to 1 PPa (1000 TPa), identifying a novel two-dimensionally bonded body-centered tetragonal (bct) phase and a fully polymerized hexagonal close-packed (hcp) phase. Electronic structure analysis reveals significant bond softening in the bct phase with increasing pressure, which may affect the dynamic behavior under finite temperatures. So, we employ the machine learning potential molecular dynamics and the two-phase method to construct the melting curve of oxygen up to 200 TPa (200 TPa, 23,740 K) and identify abnormal melting behavior beyond 100 TPa. We find oxygen exhibits higher thermal conductivity and lower isochoric heat capacity than helium at identical pressures. These results indicate that oxygen-rich envelopes may accelerate the cooling process of white dwarfs.
Similar content being viewed by others
Introduction
Oxygen is one of the most abundant elements in the universe, possessing fascinating properties. Under pressures exceeding 96 GPa1,2, molecular oxygen transitions into a metallic phase and transforms into the ζ phase at around 100 GPa3,4. Sun et al. thoroughly investigated oxygen at pressures up to 10 TPa5. They predicted a transition from the molecular phase to a tetragonal phase with polymerized helical chains at 1.9 TPa, followed by a layered monoclinic phase known as the chain-molecular phase (space group Fmmm) at 9.3 TPa. Moreover, they demonstrated that oxygen maintains a superconducting temperature of 0.6 K to 2.1 K at TPa pressures. Building upon the work of Sun et al., Cogollo-Olivo et al. re-examined the oxygen phase boundaries at finite temperatures within the 1–10 TPa pressure range6. These studies comprehensively demonstrate oxygen’s properties and states below 10 TPa. However, even under such extreme pressures, the lone pair electrons in oxygen structures persist, and the polymerization process of oxygen remains incompletely elucidated.
With increasing pressure, diatomic molecules such as oxygen, hydrogen, fluorine and nitrogen gradually aggregate from the molecular phase into the polymeric phase. Hydrogen may exhibit an alternately interlayered structure7,8,9, while nitrogen, due to its unique outer electron configuration, possesses highly stable nitrogen-nitrogen triple bonds. Beyond 100 GPa, nitrogen transforms into the cubic-gauche phase10,11, and there are many metastable polymeric nitrogen structures realized. Only fluorine and oxygen are known to remain undissociated and form polymeric phases under TPa pressure5,12. Fluorine undergoes multi-step dissociation under TPa pressures. Surprisingly, it is at 30 TPa that fluorine finally breaks down its molecular structure to form a polymeric phase12. Intermediate structures during this transition exhibit superconducting behaviour, affirming fluorine as a superconducting element despite its critical temperature (Tc) of only several Kelvin. This has sparked interest in the electronic properties of TPa pressure structure states.
However, current studies on the high-pressure polymerization of diatomic molecules have predominantly focused on static-pressure structure calculations and have not addressed finite-temperature dynamics. On the one hand, pressures exceeding the TPa regime require the use of pseudopotentials with small cutoff radius to achieve convergence, with the necessity for higher energy cutoffs further increasing computational costs; on the other hand, conventional force fields fail to accurately capture the finite-temperature dynamical behavior of dense oxygen. These limitations constrain our investigations into dense matter, particularly considering that dense astronomical objects, such as white dwarfs, experience pressures on the order of TPa or even higher. In this context, exploring the structural and property changes of materials under ultrahigh-pressure conditions provides novel insights into the extreme states of dense astronomical objects.
As a “cosmic clock”13,14,15, the white dwarf has its material state and cooling process that affect galaxy evolution and age. Carbon-oxygen white dwarf is the main form of white dwarfs. As compact objects marking the final stages of stellar evolution, white dwarfs exhibit extremely high pressures. However, the methods of studying such dense matter are very limited. Kritcher A. L. et al.16 measured the material state of the carbon envelope of white dwarfs through impact experiments, with pressures reaching up to 45 TPa. However, such shock experiments are extremely complex and challenging, and the finite-temperature dynamics of oxygen polymerization under TPa pressures and its implications for white dwarf cooling remain unexplored, which motivates our machine learning-assisted investigation. Inspired by above studies, we turned our attention to the state of oxygen. Unlike carbon, no research has yet been conducted on oxygen under extreme pressures exceeding 10 TPa. Nevertheless, earlier studies have indicated the existence of a white dwarf with an extremely high oxygen content. This indicate that the cooling process of this type of white dwarf may differ significantly from that of traditional hydrogen/helium envelope white dwarfs, which may lead to some yet-to-be-expected evolutionary results.
In order to explore the above issues efficiently, in this work, we employed the machine learning accelerated crystal structure prediction method MAGUS17,18,19 combined with NEP20 potential to predict oxygen structures under pressures up to 1 PPa; these new structures were cross-validated using ab initio random structure search (AIRSS)21,22. We have summarized the multi-step complete polymerization process of oxygen. By calculating the electron localization function (ELF), we identified two-dimensional bonding in the bct phase persisting up to 248 TPa and observed that, as pressure increases, oxygen’s lone pair electrons gradually vanish while the bonding softens. To further explore the impact of this electronic structure on finite-temperature dynamical behavior, we utilized the GPUMD molecular dynamics package23,24 to compute the melting curve of oxygen up to 200 TPa via the two-phase method, and constructed the oxygen phase diagram by determining the phase boundaries between the bct and hcp phases using the stochastic self-consistent harmonic approximation (SSCHA)25,26,27. Compared with the melting point predicted by the Coulomb coupling parameter Γ, our computational approach successfully revealed abnormal melting of the bct phase beyond 100 TPa. On this basis, to elucidate the influence of such an oxygen envelope on the thermal evolution at white dwarf conditions, we calculated the thermal conductivity and isochoric heat capacity of oxygen and helium at 200 TPa using the Kubo-Greenwood method, and found that the oxygen envelope can release heat more efficiently than the helium envelope.
Results and discussion
Figure 1(a) illustrates the oxygen polymerization pathway we constructed, in which the bct and hcp phases were observed for the first time. Further calculations indicate that oxygen transforms from the Fmmm phase to the bct phase after 20 TPa, and then the bct phase aggregates into the hcp phase after 248 TPa. As shown in Fig. 1(b), the two short lattice axes of the orthorhombic Fmmm phase gradually become equal in the bct phase, with no abrupt changes in bond lengths; additionally, the enthalpy of the Fmmm phase converges to that of the bct phase after 20 TPa, and the recognition result of the space group also becomes the bct phase. Then, with increasing pressure, oxygen aggregates into the hcp phase after 248 TPa and exhibits various bonding modes—from molecular, to chain, to layered, and finally to polymeric structures. This behavior is primarily attributed to the involvement of lone pair electrons; their strong repulsion increases the pressure required for oxygen to form a three-dimensional, densely packed polymeric structure and also affects charge transfer and chemical bond reconstruction.
a The first bond length of oxygen varies from 1 TPa to 1 PPa (red line). b The relative enthalpy of Fmmm and bct (EFmmm-Ebct) varies with pressure (orange line), and the lengths of both short axes b (lb) and c(lc) of the Fmmm phase converge with pressure increase (red dash line). The blue shading area indicates that the difference in enthalpy between the structures is within ±10 meV, which is generally considered to be degenerate. The inset is a schematic diagram of the transition from Fmmm phase to bct phase. Source data are provided as a Source Data file.
After elucidating the high-pressure phase transition pathway of oxygen, we further investigated the electronic properties of these new structures. As shown in Fig. 2(a), the ELF indicates that bonding in the body-centered tetragonal (bct) phase is predominantly established within a two-dimensional plane, while the charge distribution outside this plane exhibits an interlocking pattern due to the mutual repulsion among lone pair electrons. This two-dimensional concave–convex interlocking arrangement of charge effectively fills the interlayer gaps with electrons, thereby enhancing the structural stability under high-pressure conditions. As the pressure increases (see Figs. 2(c) and 2(d)), the charge density along the bond directions in the bct phase decreases, indicating bond softening, and the population of lone pair electrons gradually decreases. Upon transforming into the hexagonal close-packed (hcp) phase, weak interstitial anionic electrons (IAEs) emerge at the lattice interstices. These IAEs not only alleviate the interlayer repulsion but also facilitate the formation of three-dimensional bonding, thus driving the phase transition from the bct to the hcp structure.
This fundamental change in the electronic structure and bonding mode is likely to have a significant impact on the thermodynamic behavior at finite temperatures. So far, the melting curve of oxygen has been roughly estimated using the Lindemann criterion28. However, a more reliable approach is to use large-scale molecular dynamics simulations. In order to avoid the latent heat of the phase transition in molecular dynamics calculations, we employed the two-phase method to determine the melting point. The two-phase method assumes that the solid and liquid phases coexist at the melting point and are in thermodynamic equilibrium. In this case, the latent heat of the phase transition is eliminated because of the equality of the chemical potentials of these two phases, which has been well validated in previous studies29,30. Since the two-phase method for calculating the melting point is subject to significant size effects, we used systems containing more than 105 atoms for the melting point calculations.
As shown in Fig. 3(a), our results align closely with the melting point estimated by Montoya, J. A. et al.28 using the Lindemann coefficient L = 0.11. However, the Lindemann criterion yields significantly divergent results when different L-values are adopted, providing only a rough range and trend for the melting curve. In contrast, the melting curve derived from molecular dynamics calculation is more reliable, as it incorporates explicit molecular-scale processes such as pre-melting defect accumulation and the evolution of solid-liquid interfaces. Additionally, we employed the SSCHA method to calculate the Gibbs free energy and calibrate the phase boundaries between Fmmm-bct and bct-hcp, as illustrated in Fig. 3(b). The SSCHA effectively captures significant high-order anharmonic effects at elevated temperatures, circumventing the limitations of the quasi-harmonic approximation (QHA) in addressing high-temperature systems. The results indicate that the Fmmm-bct phase boundary is minimally influenced by temperature, while higher temperatures reduce the pressure required for the bct-to-hcp phase transition.
a The melting curve of oxygen in the pressure range of 1–10 TPa (red line), compared with DFT results (light bule line) by Montoya, J. A. et al.28, the black line is the phase boundary calculated by Cogollo-Olivo, B. H. et al.6. b The melting curve of oxygen in the pressure range of 10–200 TPa (red line). The melting curve determined according to the Coulomb coupling parameters (cyan line). We calculated the melting point of helium (red symbol) and Fmmm-bct phase boundary (brown line). The inset shows the phase boundary of bct-hcp (The error bars in the figure represent the upper and lower limits of the melting point, with the center being the average value). Source data are provided as a Source Data file.
One of the most important parameters for white dwarf cooling models is the Coulomb coupling parameter \(\Gamma={\left({{\rm{Ze}}}\right)}^{2}/{{\rm{a}}}{{{\rm{k}}}}_{{{\rm{B}}}}{{\rm{T}}}\), where a is the distance between ions. For large \(\Gamma\) values (\(\simeq 175\))31, the transition occurs with the release of corresponding latent heat. The melting temperatures are also derived from the Coulomb coupling parameters 169 < \(\Gamma\)< 183. As shown in Fig. 3(b), the melting points deduced from \(\Gamma=175\) are very close to those calculated by molecular dynamics within the range of 100TPa. However, such a straightforward method cannot capture the abnormal melting behaviour of oxygen and shows higher melting temperatures for high pressures of > 100 TPa. The bond softening in the bct phase reduces the energy barrier for atomic motion, thereby lowering the melting temperature and explaining the anomalous melting behavior observed beyond 100 TPa. At 200 TPa, the density of bct-phase oxygen increases upon entering the liquid state (see Fig. S2). According to the Clausius-Clapeyron relation, this density behavior implies a negative slope in the melting curve of bct-phase oxygen, which aligns with our computational predictions.
After accurately constructing the phase diagram, we calculated the transport properties and potential influence on the thermal evolution of the oxygen envelope at white dwarfs32. The recent asteroseismology analysis has demonstrated that the total oxygen content of white dwarfs is much higher than expected in typical stellar evolution calculations33. It is of great interest to compare the distinct thermal effects of helium and oxygen at white dwarfs’ conditions by calculating their thermal conductivity and specific heat capacity. In Fig. 4, we use the Kubo-Greenwood formula to calculate thermal conductivity. It is based on linear response theory and can describe the behaviour of free electrons in metals. It is a good approximation for calculating the transport properties of warm dense matter34,35,36,37,38. Compared with helium, oxygen has a higher thermal conductivity at the same pressure, which may lead to premature cooling of oxygen-rich envelopes. On the other hand, we calculated the isochoric heat capacity, where the ionic part obtained through the NVT ensemble on average (\({{{\rm{C}}}}_{{{\rm{vi}}}}=\frac{ < {{{\rm{E}}}}^{2} > -{ < {{\rm{E}}} > }^{2}}{{{{\rm{K}}}}_{{{\rm{B}}}}{ < {{\rm{T}}} > }^{2}}\)), and the electron part obtained by Sommerfeld expansion (\({{{\rm{C}}}}_{{{\rm{ve}}}}=\frac{{{{\rm{\pi }}}}^{2}{{\rm{N}}}{{{{\rm{K}}}}_{{{\rm{B}}}}}^{2}{{\rm{T}}}}{2{{{\rm{\varepsilon }}}}_{{{\rm{F}}}}}\)). The ion and electron parts of oxygen are lower than that of helium. The energy loss of cold white dwarfs mainly comes from the thermal energy stored by ions because of their higher specific heat for electrons. Therefore, oxygen with lower specific heats can release energy more effectively, thereby possibly affecting the cooling process at white dwarfs conditions14.
In summary, we explored the polymerization, phase transitions, and thermal transport properties of oxygen under extreme pressures using machine learning-assisted structure prediction and molecular dynamics simulations. We identified a novel high-pressure phase transition pathway of oxygen, from the Fmmm phase to a body-centered tetragonal (bct) phase at 20 TPa, which remains stable up to 248 TPa before transforming into a hexagonal close-packed (hcp) phase. Electronic structure analysis revealed significant bond softening in the bct phase, impacting its thermodynamic behavior. Using molecular dynamics and the two-phase method, we constructed the melting curve of oxygen up to 200 TPa and observed an anomalous melting behavior beyond 100 TPa that is not captured by classical white dwarf models. Furthermore, our calculations of the thermal conductivity and isochoric heat capacity indicate that oxygen-rich envelopes at white dwarf conditions can release energy more efficiently than those with helium envelopes. This enhanced thermal transport capability may lead to a faster cooling process for oxygen-dominated white dwarfs, impacting their evolution and lifespan. Our findings provide new insights into the extreme states of matter in compact astrophysical objects and suggest a new research pathway for understanding the cooling process of white dwarfs.
Methods
The geometry optimizations calculations are realized through the density functional theory package QUANTUM ESPRESSO39,40. We use the ultra-soft pseudopotential based on the Perdew-Burke-Ernzerhof (PBE) functional41,42. We use the CASTEP code43,44 to generate similar pseudopotentials as used in QUANTUM ESPRESSO for structure prediction and verification. This method has been successfully applied in studying a series of substances at TPa pressure45,46,47,48,49. The structure prediction results are cross-checked by MAGUS17,18,19 and AIRSS21,22. Thermal conductivity calculations are performed using Kubo Greenwood formula, implemented through KGEC code50,51,52. Machine learning force field NEP20,53 and GPU accelerated molecular dynamics package GPUMD23,24 are employed for the molecular dynamic calculation. Please refer to the supplementary information for more details.
DFT calculation details
The pseudopotential of the entire valence electron case is generated using the ld1.x module. For the structures under 25 TPa, we use the pseudopotential with a core cutoff radius of 0.7 Bohr, and the energy cutoff is 1600 eV. For the structures under pressures ranging from 25 TPa to 300 TPa, we use the pseudopotential with a core cutoff radius of 0.56 Bohr. An energy cutoff of 2500 eV. For the structure of 300 TPa-1 PPa, a pseudopotential with a core cutoff radius of 0.42 Bohr is used. The energy cutoff is 2800 eV. The pseudopotential with an internal cutoff smaller than half the distance between atoms is generally considered a reliable pseudopotential. In all calculations, the kinetic energy cutoff energy of the charge density and potential is eight times that of the kinetic energy of the wave function. A high-precision k-point grid of 2π × 0.03 Å has been used. We examined the influence of multiple exchange-correlation functionals on the Equation of State (EOS) in the Supplementary Materials (Fig. S5). The PBE functional exhibits close agreement with both the metaGGA functional R2SCAN and the hybrid Hartree-Fock functional B3LYP over the studied pressure range, indicating that PBE delivers reasonably high accuracy while maintaining significantly higher computational efficiency.
Structural prediction method
Using the crystal structure search method accelerated by machine learning17,18, we have predicted a series of planetary materials54,55,56,57,58,59. In this work, each search uses 1000 entirely random structures, with a total search structure approaching 20,000. The structure searches were employed at 1 TPa, 5 TPa, 10 TPa, 50 TPa, 100 TPa, 200 TPa, 400 TPa, and 1000 TPa, respectively. The energy cutoff used during the search process is 0.7 times that of high-precision DFT calculations.
Machine learning force field details
In this study, we employed a machine learning force field to overcome the size effect on describing the melting curve. Moreover, we used the NEP to perform molecular dynamics simulations.
The characteristics of NEP machine learning force field include GPU computation capability, enabling long-time simulations of large systems. Its descriptors consist of radial and angular components, with radial part adopting radial distribution functions, while angular part utilizes ACE-like method, the loss function of NEP is the root mean square error of forces, energies, and virials. For more information, please refer to https://github.com/brucefan1983/GPUMD.
The initial dynamic trajectory set was obtained through AIMD. The AIMD uses NPT ensemble and supercells with over 50 atoms, including sufficient solid and liquid configurations. We train the force field at different pressures to maintain accuracy, including 1 TPa, 3 TPa, 5 TPa, 7.5 TPa, 10 TPa, 50 TPa, 100 TPa, and 200 TPa (as shown in Fig. S10).
Firstly, we generated a rough force field through equal interval sampling. Subsequently, we obtained the most representative configuration in the total data through farthest point sampling and retrained it to obtain a more accurate force field, each force field consists of at least 2000 structures, at least 40% of which are liquid structures. At very high temperatures, it becomes necessary to account for the influence of electronic entropy on the dynamics processes. Therefore, we describe high-temperature molecular dynamics more accurately by introducing electron temperature into the NEP force field60. The melting point is calculated by the two-phase method, and each phase uses a supercell with more than 60,000 atoms, the total number of atoms is over 120,000. To obtain a solid-liquid interface, we immobilized one set of atoms while melting the other, thus avoiding the introduction of a vacuum layer and its impact on the melting point. We conduct all dynamic calculations for at least 50 ps, by which time a good equilibrium has been reached.
KGEC details
We performed molecular dynamics simulations on 36-atom supercells containing oxygen and helium. After achieving structural stabilization, we selected 10 configurations at 50 ps intervals to calculate the electronic thermal conductivity and isochoric heat capacity (constant-volume heat capacity). This sampling strategy aimed to mitigate the influence of thermal fluctuations in the atomic configurations. The final results were statistically analyzed, including the average value along with the maximum and minimum values observed at each temperature point Fig. 4.
a The thermal conductivity of oxygen (bct phase) and helium (hcp phase) at 200 TPa (red line for Oxygen, blue line for Helium, the error bars in the figure represent the upper and lower limits of the thermal conductivity, with the center being the average value). b The isochoric specific heat capacity of oxygen (red) and helium (blue), the solid and dashed lines represent ionic and electronic contributions, respectively. c, d Schematic diagram of the potential distinct cooling strength of oxygen-rich or helium-rich envelops of white dwarf, due to the more efficient transfer of heat by oxygen, white dwarfs in the oxygen envelope cool faster. Source data are provided as a Source Data file.
SSCHA method
We used the stochastic self-consistent harmonic approximation (SSCHA) method to calculate the anharmonic effects. The convergence of SSCHA minimization has been verified through 200 supercells. The SSCHA calculations were performed using a 3 × 3 × 3 supercell with 54 atoms. The trial harmonic dynamical matrices were obtained from the QUANTUM ESPRESSO (QE) code combined with the Density Functional Perturbation Theory (DFPT) method, and the SSCHA iteration was used to minimize the free energy corresponding to the crystal structure (lattice and atomic coordinates).
Lindemann criterion
The Lindemann criterion is a purely phenomenological vibrational approach, which posits that as temperature increases, the vibrational amplitude of atoms within a crystal lattice also increases. When the amplitude reaches a certain critical value, the atoms deviate from their equilibrium positions, leading to the disruption of the lattice structure and the onset of melting in the solid. Furthermore, this critical amplitude is considered to be related to the interatomic distance.
Data availability
Source data for Figs. 1, 3–4, Extended Data Fig. 1–10 are available with this paper. All data are available on Zenodo at https://doi.org/10.5281/zenodo.1569441461 Source data are provided with this paper.
Code availability
MAGUS source code can be accessed from gitlab (https://gitlab.com/bigd4/magus) after registration (https://www.wjx.top/vm/m5eWS0X.aspx), or on Zenodo at https://doi.org/10.5281/zenodo.1473087462. GPUMD and NEP source code can be accessed from github (https://github.com/brucefan1983/GPUMD), or on Zenodo at https://doi.org/10.5281/zenodo.1112233963.
References
Weck, G., Desgreniers, S., Loubeyre, P. & Mezouar, M. Single-crystal structural characterization of the metallic phase of oxygen. Phys. Rev. Lett. 102, 255503 (2009).
Akahama, Y., Kawamura, H., Häusermann, D., Hanfland, M. & Shimomura, O. New high-pressure structural transition of oxygen at 96 GPa associated with metallization in a molecular solid. Phys. Rev. Lett. 74, 4690 (1995).
Shimizu, K., Suhara, K., Ikumo, M., Eremets, I. & Amaya, K. Superconductivity in oxygen. Nature 393, 767–769 (1998).
Ma, Y., Oganov, A. R. & Glass, C. W. Structure of the metallic ζ -phase of oxygen and isosymmetric nature of the ε − ζ phase transition: Ab initio simulations. Phys. Rev. B 76, 064101 (2007).
Sun, J., Martinez-Canales, M., Klug, D. D., Pickard, C. J. & Needs, R. J. Persistence and eventual demise of oxygen molecules at terapascal pressures. Phys. Rev. Lett. 108, 045503 (2012).
Cogollo-Olivo, B. H., Biswas, S., Scandolo, S. & Montoya, J. A. Phase diagram of oxygen at extreme pressure and temperature conditions: an ab initio study. Phys. Rev. B 98, 094103 (2018).
Pickard, C. J. & Needs, R. J. Structure of phase III of solid hydrogen. Nat. Phys. 3, 473 (2007).
Pickard, C. J., Martinez-Canales, M. & Needs, R. J. Density functional theory study of phase IV of solid hydrogen. Phys. Rev. B 85, 214114 (2012).
Howie, R. T., Guillaume, C. L., Scheler, T., Goncharov, A. F. & Gregoryanz, E. Mixed molecular and atomic phase of dense hydrogen. Phys. Rev. Lett. 108, 125501 (2012).
Eremets, M. I., Hemley, R. J., Mao, H. & Gregoryanz, E. Semiconducting non-molecular nitrogen up to 240 GPa and its low-pressure stability. Nature 411, 170 (2001).
Goncharov, A. F., Gregoryanz, E., Mao, H., Liu, Z. & Hemley, R. J. Optical evidence for a nonmolecular phase of nitrogen above 150 GPa. Phys. Rev. Lett. 85, 1262 (2000).
Duan, D. et al. Multistep dissociation of fluorine molecules under extreme compression. Phys. Rev. Lett. 126, 225704 (2021).
Córsico, A. H., Althaus, L. G., Bertolami, M. M. M. iller & Kepler, S. O. Pulsating white dwarfs: new insights. Astron. Astrophys. Rev. 27, 7 (2019).
Althaus, L. G., Córsico, A. H., Isern, J. & García-Berro, E. Evolutionary and pulsational properties of white dwarf stars. Astron. Astrophys. Rev. 18, 471 (2010).
Winget, D. E. & Kepler, S. O. Pulsating white dwarf stars and precision asteroseismology. Annu. Rev. Astron. Astrophys. 46, 157 (2008).
Kritcher, A. L. et al. A measurement of the equation of state of carbon envelopes of white dwarfs. Nature 584, 51 (2020).
Xia, K. et al. A novel superhard tungsten nitride predicted by machine-learning accelerated crystal structure search. Sci. Bull. 63, 817 (2018).
Wang, J. et al. MAGUS: machine learning and graph theory assisted universal structure searcher. Natl Sci. Rev. 10, nwad128 (2023).
Han, Y. et al. Efficient crystal structure prediction based on the symmetry principle. Nat. Comput. Sci. 5, 255 (2025).
Fan, Z. Improving the accuracy of the neuroevolution machine learning potential for multi-component systems. J. Phys.: Condens. Matter 34, 125902 (2022).
Pickard, C. J. & Needs, R. J. High-pressure phases of silane. Phys. Rev. Lett. 97, 045504 (2006).
Pickard, C. J. & Needs, R. J. Ab initio random structure searching. J. Phys.: Condens. Matter 23, 053201 (2011).
Fan, Z. et al. GPUMD: A package for constructing accurate machine-learned potentials and performing highly efficient atomistic simulations. J. Chem. Phys. 157, 114801 (2022).
Liu, J., Byggmästar, J., Fan, Z., Qian, P. & Su, Y. Large-scale machine-learning molecular dynamics simulation of primary radiation damage in tungsten. Phys. Rev. B 108, 054312 (2023).
Monacelli, L. et al. The stochastic self-consistent harmonic approximation: calculating vibrational properties of materials with full quantum and anharmonic effects. J. Phys.: Condens. Matter 33, 363001 (2021).
Bianco, R., Errea, I., Paulatto, L., Calandra, M. & Mauri, F. Second-order structural phase transitions, free energy curvature, and temperature-dependent anharmonic phonons in the self-consistent harmonic approximation: theory and stochastic implementation. Phys. Rev. B 96, 014111 (2017).
Monacelli, L., Errea, I., Calandra, M. & Mauri, F. Pressure and stress tensor of complex anharmonic crystals within the stochastic self-consistent harmonic approximation. Phys. Rev. B 98, 024106 (2018).
Montoya, J. A. & Cogollo-Olivo, B. H. Estimation of suitable upper-limits for temperature, in stability comparisons between solid phases at high pressures. Study cases: carbon, oxygen, and fluorine. Rev. Acad. Colomb. Cienc. Exactas Fís. Nat. 47, 182 (2023).
Hong, Q.-J. & Van De Walle, A. Solid-liquid coexistence in small systems: A statistical method to calculate melting temperatures. J. Chem. Phys. 139, 094114 (2013).
Asadi, E., Asle Zaeem, M., Nouranian, S. & Baskes, M. I. Quantitative modeling of the equilibration of two-phase solid-liquid Fe by atomistic simulations on diffusive time scales. Phys. Rev. B 91, 024105 (2015).
Potekhin, A. Y. & Chabrier, G. Equation of state of fully ionized electron-ion plasmas. II. Extension to relativistic densities and to the solid phase. Phys. Rev. E 62, 8554 (2000).
Chen, J. et al. Slowly cooling white dwarfs in M13 from stable hydrogen burning. Nat. Astron.5, 1170 (2021).
Giammichele, N. et al. A large oxygen-dominated core from the seismic cartography of a pulsating white dwarf. Nature 554, 73 (2018).
Chen, Z. et al. Evolution of ac conductivity in nonequilibrium warm dense gold. Phys. Rev. Lett. 110, 135001 (2013).
Xu, J. et al. Thermal conductivity and electrical resistivity of solid iron at Earth’s core conditions from first principles. Phys. Rev. Lett. 121, 096601 (2018).
Pozzo, M., Davies, C., Gubbins, D. & Alfè, D. Thermal and electrical conductivity of iron at Earth’s core conditions. Nature 485, 355 (2012).
Pozzo, M., Davies, C., Gubbins, D. & Alfè, D. Thermal and electrical conductivity of solid iron and iron–silicon mixtures at Earth’s core conditions. Earth Planet. Sci. Lett. 393, 159 (2014).
Holst, B., French, M. & Redmer, R. Electronic transport coefficients from ab initio simulations and application to dense liquid hydrogen. Phys. Rev. B 83, 235120 (2011).
Giannozzi, P. et al. Advanced capabilities for materials modelling with QUANTUM ESPRESSO. J. Phys. Condens. Matter 29, 465901 (2017).
Giannozzi, P. et al. Quantum ESPRESSO toward the exascale. J. Chem. Phys. 152, 154105 (2020).
Rappe, A. M., Rabe, K. M., Kaxiras, E. & Joannopoulos, J. D. Optimized pseudopotentials. Phys. Rev. B 41, 1227 (1990).
Perdew, J. P., Burke, K. & Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865 (1996).
Clark, S. J. et al. First principles methods using CASTEP. Z. Kristallogr. Crystalline Mater. 220, 567 (2005).
Segall, M. D. et al. First-principles simulation: ideas, illustrations and the CASTEP code. J. Phys.: Condens. Matter 14, 2717 (2002).
Martinez-Canales, M., Pickard, C. J. & Needs, R. J. Thermodynamically stable phases of carbon at multiterapascal pressures. Phys. Rev. Lett. 108, 045704 (2012).
Pickard, C. J. & Needs, R. J. Aluminium at terapascal pressures. Nat. Mater. 9, 624 (2010).
Monserrat, B., Martinez-Canales, M., Needs, R. J. & Pickard, C. J. Helium-iron compounds at terapascal pressures. Phys. Rev. Lett. 121, 015301 (2018).
Monserrat, B., Drummond, N. D., Pickard, C. J. & Needs, R. J. Electron-phonon coupling and the metallization of solid helium at terapascal pressures. Phys. Rev. Lett. 112, 055504 (2014).
Pickard, C. J., Martinez-Canales, M. & Needs, R. J. Decomposition and terapascal phases of water ice. Phys. Rev. Lett. 110, 245701 (2013).
Calderín, L., Karasiev, V. V. & Trickey, S. B. Kubo–Greenwood electrical conductivity formulation and implementation for projector augmented wave datasets. Comput. Phys. Commun. 221, 118 (2017).
Di Paola, C., Macheda, F., Laricchia, S., Weber, C. & Bonini, N. First-principles study of electronic transport and structural properties of Cu12Sb4S13 in its high-temperature phase. Phys. Rev. Res. 2, 033055 (2020).
Mazevet, S., Torrent, M., Recoules, V. & Jollet, F. Calculations of the transport properties within the PAW formalism. High. Energy Density Phys. 6, 84 (2010).
Fan, Z., Chen, W., Vierimaa, V. & Harju, A. Efficient molecular dynamics simulations with many-body potentials on graphics processing units. Comput. Phys. Commun. 218, 10 (2017).
Huang, T. et al. Metallic aluminum suboxides with ultrahigh electrical conductivity at high pressure. Research 2022, 9798758 (2022).
Liu, C. et al. Partially diffusive helium-silica compound under high pressure. Chin. Phys. Lett. 39, 076101 (2022).
Gao, H. et al. Superionic silica-water and silica-hydrogen compounds in the deep interiors of Uranus and Neptune. Phys. Rev. Lett. 128, 035702 (2022).
Liu, C. et al. Mixed coordination silica at megabar pressure. Phys. Rev. Lett. 126, 035701 (2021).
Gao, H. et al. Coexistence of plastic and partially diffusive phases in a helium-methane compound. Natl Sci. Rev. 7, 1540 (2020).
Liu, C. et al. Multiple superionic states in helium–water compounds. Nat. Phys. 15, 1065 (2019).
Zhang, Y. et al. Warm dense matter simulation via electron temperature dependent deep potential molecular dynamics. Phys. Plasmas 27, 122704 (2020).
Wang, Y. Source Data for “Machine Learning Simulations Reveals Oxygen’s Phase Diagram and Thermal Properties at Conditions Relevant to White Dwarfs”. Zenodo. https://doi.org/10.5281/zenodo.15694414 (2025).
Han, Y. et al. Source code and demo of MAGUS (Machine Learning and Graph Theory Assisted Universal Structure Searcher) v2.0.0. Zenodo. https://doi.org/10.5281/zenodo.14730874 (2025).
Fan, Z. brucefan1983/GPUMD: GPUMD-v3.9.3. https://doi.org/10.5281/zenodo.11122339 (2024).
Acknowledgements
We gratefully acknowledges the National Natural Science Foundation of China (Grant 12125404, T2495231, 123B2049), the Basic Research Program of Jiangsu (Grant BK20233001, BK20241253), the Jiangsu Funding Program for Excellent Postdoctoral Talent (Grants 2024ZB002 and 2024ZB075), the Postdoctoral Fellowship Program of CPSF (Grant GZC20240695), the AI & AI for Science program of Nanjing University, Artificial Intelligence and Quantum physics (AIQ) program of Nanjing University, and the Fundamental Research Funds for the Central Universities. C.J.P. is supported by the EPSRC through grant EP/S021981/1. The calculations were carried out using supercomputers at the High-Performance Computing Center of Collaborative Innovation Center of Advanced Microstructures, the high-performance supercomputing center of Nanjing University.
Author information
Authors and Affiliations
Contributions
J.S. conceived and designed the work. Y.W. performed primary computations and drafted the manuscript. J.Y.S. helped train machine learning potentials, molecular dynamics simulations, and free energy calculations. Z.L. and T.H. helped with the analysis of the molecular dynamics results. J.W. and C.D. contributed to the MAGUS code and helped on crystal structure prediction. C.J.P. performed crystal structure prediction with the AIRSS code. D.N. provided critical suggestions on the white dwarf-related content. Y.W., H.T.W., D.X., D.N., and J.S. wrote the manuscript. All authors participated in scientific discussions and provided critical feedback on the manuscript.
Corresponding authors
Ethics declarations
Competing interests
The authors declare no competing interests.
Peer review
Peer review information
Nature Communications thanks Sandro Scandolo, and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. A peer review file is available.
Additional information
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Supplementary information
Source data
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, 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 Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted 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 licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
About this article
Cite this article
Wang, Y., Shi, J., Liang, Z. et al. Machine Learning simulations reveal oxygen’s phase diagram and thermal properties at conditions relevant to white dwarfs. Nat Commun 16, 5504 (2025). https://doi.org/10.1038/s41467-025-61390-0
Received:
Accepted:
Published:
DOI: https://doi.org/10.1038/s41467-025-61390-0