TY - JOUR AU1 - Gao,, Hao AU2 - Liu,, Cong AU3 - Hermann,, Andreas AU4 - Needs, Richard, J AU5 - Pickard, Chris, J AU6 - Wang,, Hui-Tian AU7 - Xing,, Dingyu AU8 - Sun,, Jian AB - Abstract Helium and methane are major components of giant icy planets and are abundant in the universe. However, helium is the most inert element in the periodic table and methane is one of the most hydrophobic molecules, thus whether they can react with each other is of fundamental importance. Here, our crystal structure searches and first-principles calculations predict that a He3CH4 compound is stable over a wide range of pressures from 55 to 155 GPa and a HeCH4 compound becomes stable around 105 GPa. As nice examples of pure van der Waals crystals, the insertion of helium atoms changes the original packing of pure methane molecules and also largely hinders the polymerization of methane at higher pressures. After analyzing the diffusive properties during the melting of He3CH4 at high pressure and high temperature, in addition to a plastic methane phase, we have discovered an unusual phase which exhibits coexistence of diffusive helium and plastic methane. In addition, the range of the diffusive behavior within the helium-methane phase diagram is found to be much narrower compared to that of previously predicted helium-water compounds. This may be due to the weaker van der Waals interactions between methane molecules compared to those in helium-water compounds, and that the helium-methane compound melts more easily. crystal structure prediction, ab initio molecular dynamics, ab initio calculations, high pressure and high temperature, melting and phase transition, collective motion INTRODUCTION Hydrogen and helium are the most abundant elements in the universe and are significant constituents of the planets in our solar system [1]. Superionicity of hydrogen in ice and ammonia was discovered in early works [2–4]. In the superionic state, some atoms form a fixed sublattice while others diffuse as in a liquid. These states have attracted much interest in planetary and high-pressure science for some decades [5–19], because ionic mobility affects thermal and electrical conductivity deep inside planets, and therefore their thermal evolution and ability to sustain magnetic fields. For example, it has been reported that the body centered cubic (bcc) phase of superionic ice transforms to a face centered cubic (fcc) phase, which is also superionic [11]; the latter has reportedly been seen in recent shock wave experiments [18,19]. Sun et al. [14] reported a superionic state in close-packed and P21/c phases of ice at higher pressures, and French et al. [15] constructed thermodynamic potentials for the superionic bcc and fcc ice phases and calculated the phase boundary between them. Superionicity of hydrogen in ammonia and ammonia compounds has also been studied [10,12,13,16,20,21]. Methane is also an important component of giant planets in addition to water and ammonia [17]. Superionic states have not been found in methane, due to the polymerization and release of hydrogen that occurs in methane at high pressures and temperatures [22–25]. Such polymerization may result in ‘diamond rain’ in the icy planets [26]. In addition to the superionic states, the plastic phase in molecular crystals such as ammonia, in which protons rotate around the fixed nitrogen atoms, has been reported to emerge at certain pressure and temperature ranges [10]. The plastic phase and rotational motion in water have been studied under high pressure [27–29]. Very recently, Li et al. [30] found that colossal barocaloric effects in plastic crystals may have potential applications in solid-state refrigeration technologies. While hydrogen-rich compounds have been studied extensively, this does not hold for helium-containing compounds. Traditionally helium is seen as the most inert element because of its closed-shell electronic configuration, but recently helium has been found to react under high pressure with metals [31,32] and ionic compounds [33–35]. Liu et al. [33] attributed the reactivity of helium with ionic compounds to the lowering of the Madelung energy between ions arising from the insertion of helium. At high pressures, helium has also been reported to form van der Waals (vdW) compounds with atoms such as neon [36] and covalent molecules such as ammonia [21], water [37–39], nitrogen [40–42], carbon dioxide [43] and arsenolite [44]. Superionicity involving helium has rarely been studied. Technically, helium is expected to diffuse as a neutral entity rather than an ionic entity, but a partially diffusive state involving helium has significant implications for thermal conductivities and viscosities. One such example is the helium-iron compound FeHe [32]. The superionic phase of FeHe occurs at pressures above 2 TPa and temperatures higher than 12000 K. Several superionic phases have recently been found in helium-water mixtures, which show novel behaviors such as simultaneous superionicity of hydrogen and helium [39]. However, the possibility of compounds in the helium-methane system under high pressures and the nature of their high-temperature behavior are still open questions. This is despite these species forming significant portions of icy planetary atmospheres and mantles, respectively. Their miscibility is then relevant for the atmosphere-mantle boundary region, which is expected to feature large compositional gradients [45,46]. We have therefore investigated helium-methane mixtures at high pressures and temperatures. RESULTS In this work, searches for helium-methane compounds were performed using a machine learning accelerated crystal structure prediction method [47]. Structural optimizations, ab initio molecular dynamics (AIMD) simulations, enthalpies and electronic structures were calculated using the Vienna ab initio simulation (VASP) code [48] with projector augmented-wave potentials and the Perdew-Burke-Ernzerhof exchange-correlation functional (PBE) [49]. Further details of the calculations are provided in the Supplementary Material. Thermodynamic stabilities of the newly predicted compounds were estimated from their relative enthalpies of formation compared with those of polymorphs of carbon, hydrogen and helium, and C—H compounds predicted in previous works [23,25,50–52]. Using an optB88-vdW functional [53,54] and hard pseudopotentials, we predicted that a new structure with a helium-methane stoichiometry of 3:1 is stable from 55 to 155 GPa (as shown in Fig. 1(a)) which is much wider than the stable pressure region of pure methane molecular crystals. Because methane decomposes above 100 GPa, here we showed a C-H-He phase diagram rather than a He-CH4 phase diagram. The formation energy of the compound is about 32 meV/f.u. at 110 GPa (Fig. 1(b)). We have also computed the enthalpy-pressure relations using the PBE and PBE + D3 functionals [55], as shown in the Supplementary Material, all of the results confirmed that He3CH4 is energetically stable. He3CH4 is a molecular crystal with hexagonal space group (SG) P63mc composed of helium and methane molecules. As shown in Fig. 1(c) and (d), helium chains are inserted into the open channels formed by the methane molecules. The packing of methane molecules in He3CH4 is the same as in the experimental hexagonal closed packed (HCP) methane phase at low pressure [56], which is very different from the high-pressure phases of methane with orthorhombic space groups. Another compound, HeCH4 with SG P21/c, is stable over a narrow range of pressures around 105 GPa. In addition, we have also found several metastable compounds, including HeCH4 with SG P21, He2CH4 with SG P31m and P21/m, and He2(CH4)3 with SG Cm. These metastable phases are very close to the convex hull. The phase diagrams under different pressures are shown in the Supplementary Material. Figure 1. Open in new tabDownload slide Energetic stability and crystal structures of He-CH4 compounds. (a) C-H-He phase diagram at 155 GPa, (b) enthalpy-pressure relations for the C-H-He compounds and the crystal structure of He3CH4 viewed along [110] (c) and [001] (d). The solid and open circles represent stable and metastable phases, respectively. Figure 1. Open in new tabDownload slide Energetic stability and crystal structures of He-CH4 compounds. (a) C-H-He phase diagram at 155 GPa, (b) enthalpy-pressure relations for the C-H-He compounds and the crystal structure of He3CH4 viewed along [110] (c) and [001] (d). The solid and open circles represent stable and metastable phases, respectively. He3CH4 is an insulator with a large band gap of 8.9 eV at 50 GPa, as shown in Fig. S11 in the Supplementary Material, it is higher than that of pure methane molecular crystals. The phonon band structures of He3CH4 at 0 and 50 GPa shown in the Supplementary Material indicate the dynamical stability of this compound. HeCH4 is also found to be an insulator with a large band gap and is dynamically stable. The conditions in the upper mantle regions of icy planets reach tens to hundreds of GPa and thousands of K [45]. To obtain the dynamical properties of He3CH4 we performed AIMD simulations within the pressure range 50–200 GPa and the temperature range 500–3000 K. The averaged mean squared displacements (MSD) of hydrogen, helium and carbon atoms were calculated to study phase transitions induced by temperature and pressure. According to the diffusive behaviors of the different atoms, the states of He3CH4 can be divided into four types: the solid phase, a plastic CH4 phase, a phase of plastic CH4 plus diffusive He, and the fluid phase. We used three representative trajectories to reveal the differences between the states by comparing their MSD and motions, as shown in Fig. 2. All of the simulations start from the relaxed configuration at 150 GPa, which are independent, but with different temperatures. At 1000 K, the atoms are restricted to their equilibrium positions, and exhibit small vibrations. As shown in Fig. 2(a), the MSD of the atoms are extremely small in the simulations, thus He3CH4 remains in the solid phase at this temperature. Figure 2. Open in new tabDownload slide Dynamical behaviors of He3CH4 at high pressure from AIMD simulations at 1000 K, 1900 K and 2350 K. (a-c) The averaged MSD of H, He and C atoms at different temperatures. (d-f) Representations of trajectories at different temperatures in the last 10 ps. Blue, cyan and red dots represent H, He and C atoms, respectively. At 2350 K, the trajectories of H and He atoms overlap with one another, and therefore we only show He and C trajectories here. Figure 2. Open in new tabDownload slide Dynamical behaviors of He3CH4 at high pressure from AIMD simulations at 1000 K, 1900 K and 2350 K. (a-c) The averaged MSD of H, He and C atoms at different temperatures. (d-f) Representations of trajectories at different temperatures in the last 10 ps. Blue, cyan and red dots represent H, He and C atoms, respectively. At 2350 K, the trajectories of H and He atoms overlap with one another, and therefore we only show He and C trajectories here. When the temperature increases to 1900 K, the compound transforms to the plastic CH4 phase. As shown in Fig. 2(b), the averaged MSD of H atoms increases rapidly in a short timescale, and stays about 2.0 Å2 afterwards. This is deemed to be a plastic phase when the methane molecules are free to rotate around the carbon atoms with small fluctuations of C-H bond lengths and angles (Fig. 2(e)). Plastic phases have also been reported in pure methane [57], ice [27,58], ammonia [10,12] and helium-ammonia compounds [21]. Meanwhile, the averaged MSD values of He and C atoms are still very small. Using the rigid molecule approximation [57] we have been able to calculate the theoretical MSD of the H atoms. Suppose that the methane molecules are rigid with fixed radius |${r_{CH}}$|(the length of a C-H bond). For one of the H atoms, its initial position is |${{\boldsymbol{r}}_0}( {{r_{CH}},0,0} )$| in spherical coordinates. Since methane molecules rotate freely, the position of the H atom |${\boldsymbol{r}}( {{r_{CH}},\ \theta ,\ \phi } )$| is distributed uniformly over the sphere of radius |${r_{CH}}$|⁠. Therefore, the analytical MSD is $$\begin{eqnarray}\left\langle {{{\left( {{\boldsymbol{r}} - {{\boldsymbol{r}}_0}} \right)}^2}} \right\rangle& =& \frac{1}{{4\pi }}\int_{0}^{\pi }{d}\theta \ \int_{0}^{{2\pi }}{{d\phi \sin \theta \ \left( {{\boldsymbol{r}} {-} {\boldsymbol{r}}_0^2} \right)}}\nonumber\\ &=& \frac{{2r_{CH}^2}}{{4\pi }}\int_{0}^{\pi }{d}\theta \ \int_{0}^{{2\pi }}{{d\phi \sin \theta}} \nonumber\\ &&\times\, {{(1 - \ \cos \theta )}} = \ 2r_{CH}^{2}. \end{eqnarray}$$(1) For the plastic CH4 phase, the resulting MSD is|$\langle {{{( {r - {r_0}} )}^2}} \rangle$| = 2.142 Å2, which is very close to the value from our AIMD calculations (2.169 Å2). At higher temperatures (2350 K) methane molecules rotate while helium atoms are diffusive (Fig. 2(c) and (f)), which leads to the formation of a superionic-like He state in He3CH4. The diffusion coefficient of helium atoms DHeis 4.81 × 10−10m2/s from the velocity auto-correlation functions (VACFs). We obtain very similar results DHe = 4.48 × 10−10m2/s from the MSD. Diffusion coefficients of helium along different directions were also calculated: |$D_{He}^x = \ 3.79\ \times {10^{ - 10}}{m^2}/s$|⁠, |$D_{He}^y = \ 4.02\ \times {10^{ - 10}}{m^2}/s$|⁠,|$\ \ D_{He}^z = \ 6.62\ \times {10^{ - 10}}{m^2}/s$|⁠. The existence of open channels along the z axis in He3CH4 (Fig. 1(d)) can explain the anisotropy of diffusion coefficients. Heating the superionic phases of He3CH4 eventually leads to melting of the methane sublattice, which gives rise to the fluid phase. We can further understand the dynamical processes of the He3CH4 structure at different temperatures by using the radial distribution function (RDF). C-H and C-He partial RDFs are shown in Fig. 3(a) and (b) and others are shown in the Supplementary Material. The C-He, He-He and H-He partial RDFs of the partially diffusive phase (2350 K, green lines) are very similar to those of the plastic phase (1900 K, red lines), but are clearly different to those of the fluid phase (2700 K, blue lines). We analyzed some trajectories of the partially diffusive phase and found that the diffusion of helium atoms maintains the order of the helium chains. In some cases of the partially diffusive phase we can see the plateaus in the averaged MSD of He. The performance of the diffusive helium atoms is similar to collective diffusion of ions in lithium battery cathode materials [59,60]. The atoms jump along the chains or hop from one chain to another and therefore the RDFs remain solid-like. However, in the fluid phase, the sublattices of helium and methane have already melted and the RDFs are liquid-like. Finally, the first peak of the C-H RDFs remains essentially unchanged deep into the fluid phase. This indicates that the integrity of the methane molecules is maintained up to the highest temperature. Figure 3. Open in new tabDownload slide Dynamical structure and interaction analysis. Radial distribution functions (RDFs) for C-H (a) and C-He (b) pairs in He3CH4 at around 150 GPa and heating to about 1000 K (solid phase), 1900 K (plastic phase), 2350 K (coexistence of plastic and partially diffusive phase) and 2700 K (fluid phase). (c) Plots of the reduce density gradient (RDG) versus the electron density multiplied by the sign of the second largest eigenvalue of the electron-density Hessian matrix. (d) ELF plotted in the (110) plane. Figure 3. Open in new tabDownload slide Dynamical structure and interaction analysis. Radial distribution functions (RDFs) for C-H (a) and C-He (b) pairs in He3CH4 at around 150 GPa and heating to about 1000 K (solid phase), 1900 K (plastic phase), 2350 K (coexistence of plastic and partially diffusive phase) and 2700 K (fluid phase). (c) Plots of the reduce density gradient (RDG) versus the electron density multiplied by the sign of the second largest eigenvalue of the electron-density Hessian matrix. (d) ELF plotted in the (110) plane. The superionic state in helium-methane differs from the few other examples of superionic helium [32,39] because plastic CH4 and diffusive He states coexist. Plastic and superionic phases also appear in the phase diagram of ammonia and ice [10,12,58], but they cannot coexist since both the plastic and diffusive atoms in ammonia and water are protons. From the results of extensive AIMD simulations, we proposed a phase diagram of He3CH4 between 50 and 200 GPa and below 3000 K (Fig. 4). The colored dots represent independent NVT (Canonical Ensemble) simulations which are classified by their averaged MSDs and RDFs (tests with the NPT (Isothermal-Isobaric Ensemble) confirm these classifications, see the Supplementary Material). The phase boundaries divide the diagram into four regions: solid, plastic CH4, plastic CH4 + diffusive He and fluid phases. The partially diffusive phase appears at pressures above 70 GPa within a narrow range. Compared with helium-water compounds, He3CH4 is dominated by dispersion interactions between methane molecules, which are much weaker than the hydrogen bonds between water molecules. Therefore the ice frameworks in helium-water compounds are more stable than the sublattice of methane molecules. There are many open channels in helium-water compounds, which results in a much wider partially diffusive region in helium-water than that in the helium-methane system. To validate the superionicity of helium we have performed AIMD calculations at 150 GPa using hard pseudopotentials. The results are shown in the Supplementary Material and the appearance of the partially diffusive phase at high temperatures is unaffected. Figure 4. Open in new tabDownload slide Phase diagram of He3CH4 as determined in this work. Each symbol represents an AIMD simulation. The dashed lines are phase boundaries. Figure 4. Open in new tabDownload slide Phase diagram of He3CH4 as determined in this work. Each symbol represents an AIMD simulation. The dashed lines are phase boundaries. To investigate the interactions in He3CH4 we have applied a real-space analysis to the helium-methane compound. According to the electron localization function (ELF) shown in Fig. 3(d), strong intramolecular covalent bonds persist between carbon and hydrogen atoms while the interactions between methane and helium molecules are of closed-shell character. The types of intermolecular interactions in He3CH4 can be determined using the atoms-in-molecules (AIM) theory [61] and reduced density gradient (RDG) analysis [62]. Based on AIM, a topological analysis of the electron density ρ(r) was carried out to compute atomic Bader charges and search for bond critical points (BCPs). We found that the charge transfer in He3CH4 is less than 0.03e between carbon and hydrogen. This is consistent with the fact that the electronegativities of carbon and hydrogen are similar and methane cannot participate in hydrogen bonding. A BCP connecting a pair of atoms provides important information about the bonding between atoms. We show all of the non-equivalent BCPs and their properties in the Supplementary Material. The first and second BCPs represent the bonds inside methane molecules. The electron densities at these BCPs are much larger than others and the negative Laplacian values indicate the concentration of electrons between the carbon and hydrogen atoms. These are typical features of covalent bonding. Other BCPs with small densities and positive Laplacian values are characteristic of dispersion interactions. Furthermore, reduced density gradient analysis has been applied to the electron density. As shown in Fig. 3(c), the spikes at low density, the low-gradient region (inside the red dashed box) reflects the existence of non-covalent interactions [62] in He3CH4. The distinct spikes are very near zero, indicating that the type is vdW interactions, which are weaker than hydrogen bonds in helium-water [39] and helium-ammonia compounds [21]. DISCUSSION The isotopic effects in hydrogen are important in some cases [63] and we have accounted for them in the helium-methane compound by comparing the vibrational densities of states (VDOS) of He3CH4 and He3CD4 (Supplementary Material). Within the harmonic approximation, isotopic effects lead to changes in the atomic mass and in the phonon frequencies. The frequency of the highest peak in the partial VDOS drops from about 3700 cm−1 for hydrogen to about 2700 cm−1 for deuterium, with a ratio of about |$1{/}\sqrt{2}$|⁠. Since strong covalent bonds exist between carbon and hydrogen/deuterium, the highest phonon peaks of carbon have the same tendencies as the isotopic effects. In contrast, the frequency region of the helium VDOS does not change substantially, because the vdW interactions between helium and methane are weak. The isotopic effects can also influence dynamical properties, and therefore we conducted AIMD calculations for He3CD4 at 2350 K and compared the VDOS of He3CH4 and He3CD4 at high temperatures. For carbon and hydrogen/deuterium the reduction of the frequencies also appears with the ratio of |$1{/}\sqrt{2}$|⁠. For helium, the partial VDOS is very similar. This demonstrates that isotopic effects do not substantially affect the superionicity of the helium-methane compound. In previous studies nuclear quantum effects were considered for ice [15] and mixtures of methane, ammonia and water [64]. The nuclear quantum corrections for He3CH4 at different densities and temperatures are shown in the Supplementary Material. The quantum corrections for the helium-methane compounds are slightly larger than those for ice [15] because of the higher proportion of hydrogen and helium in the compound and the lower temperature region in our simulations. CONCLUSION In summary, we have predicted a helium-methane compound (He3CH4) that is stable at pressures relevant to upper mantle conditions of icy planets. The inclusion of He atoms highly changes the packing of methane molecules and the stable pressure region of the helium-methane compound is much wider than that of pure methane. He insertion also changes the electronic properties of methane. For example, the band gap of He3CH4 is larger than that of pure methane. Moreover, the phase diagram of He3CH4 has been investigated and a novel phase of coexistence of plastic methane and diffusive helium has been found. The temperature range of the partially diffusive regime in helium-methane is narrower than that in the helium-water system, due to the weaker interactions between the methane molecules, which results in a relatively fragile framework and an easier transition to the fluid state than in the helium-water system. In addition, we observed anisotropy in the He diffusion, which is related to the structure of the compound. We have also analyzed the interactions in He3CH4 and their effects on the phase diagram in comparison with previously predicted helium-water compounds. This work should be helpful in constructing models of icy giant planets, and it would be very instructive to investigate how much the finite-temperature miscibility of helium and methane reflects our ground state results [65]. METHODS We used a Bayesian-Optimization-based crystal structure prediction method [47] combined with VASP to predict new compounds for the He-CH4 system. The prediction results have been checked by ab initio random structure searching (AIRSS) [66]. The optB88-vdW functional [53] was employed to account for vdW interactions in the calculations of enthalpies and electronic structures. A basis set energy cutoff of 720 eV was used except for the AIMD calculations, for which a lower cutoff energy of 625 eV was used for reducing the computational cost of extensive AIMD simulations. The Brillouin zone was sampled with a Monkhorst-Pack k-point mesh with a spacing of 2π × 0.03 Å−1. The phonon dispersion curves were calculated using |$2\ \times \ 2\ \times \ 2$| supercells with PHONOPY code [67] to validate the dynamical stabilities of the predicted structures. The AIM [61] and RDG [62] analysis of the electron density ρ(r) was performed using the critic2 code [68]. We used orthorhombic supercells containing 256 atoms to perform AIMD simulations for He3CH4 in the NVT ensemble with Γ-centered k-points sampling. The time step of AIMD was set to 1 fs and all simulations were carried out with at least 3000 steps. Some trajectories were extended to more than 10 ps to confirm the stabilities. In addition, to validate our results we employed hard pseudopotentials for hydrogen and carbon and repeated the calculations. The cutoff energy of 1000 eV was used to calculate pressure-composition phase diagrams. A cutoff energy of 910 eV was used for the AIMD simulations. The nuclear quantum corrections of free energies are calculated from the AIMD trajectories using the method proposed in reference [15]. Acknowledgements The calculations were carried out using supercomputers at the High Performance Computing Center of Collaborative Innovation Center of Advanced Microstructures, high performance supercomputing center of Nanjing University; ‘Tianhe-2’ at NSCC-Guangzhou; and the CSD3 Peta4 CPU/KNL machine in the University of Cambridge. FUNDING J.S. gratefully acknowledges financial support from the National Key R&D Program of China (2016YFA0300404), the National Natural Science Foundation of China (11974162 and 11834006), the Fundamental Research Funds for the Central Universities. C.J.P. and R.J.N. acknowledge financial support from the Engineering and Physical Sciences Research Council (EPSRC) of the U.K. under grants [EP/G007489/2] (C.J.P.) and [EP/P034616/1] (R.J.N.). C.J.P. also acknowledges financial support from EPSRC and the Royal Society through a Royal Society Wolfson Research Merit award. AUTHOR CONTRIBUTIONS J.S. conceived and led the project. H.G. and C.J.P performed the calculations. J.S, H.G. and C.L. made the analysis. J.S., H.G., A.H., and R.J.N wrote the manuscript. All authors discussed the results and commented on the manuscript. Conflict of interest statement. None declared. REFERENCES 1. Stevenson DJ . Metallic helium in massive planets . Proc Natl Acad Sci USA 2008 ; 105 : 11035 – 6 . Google Scholar Crossref Search ADS WorldCat 2. Ryzhkin IA . Superionic transition in ice . Solid State Commun 1985 ; 56 : 57 – 60 . Google Scholar Crossref Search ADS WorldCat 3. Demontis P , LeSar R, Klein ML. New high-pressure phases of ice . Phys Rev Lett 1988 ; 60 : 2284 – 7 . Google Scholar Crossref Search ADS PubMed WorldCat 4. Cavazzoni C , Chiarotti GL, Scandolo S et al. Superionic and metallic states of water and ammonia at giant planet conditions . Science 1999 ; 283 : 44 – 6 . Google Scholar Crossref Search ADS PubMed WorldCat 5. Yakushev VV , Postnov VI, Fortov VE et al. Electrical conductivity of water during quasi-isentropic compression to 130 GPa . J Exp Theor Phys 2000 ; 90 : 617 – 22 . Google Scholar Crossref Search ADS WorldCat 6. Chau R , Mitchell AC, Minich RW et al. Electrical conductivity of water compressed dynamically to pressures of 70–180 GPa (0.7–1.8 Mbar) . J Chem Phys 2001 ; 114 : 1361 – 5 . Google Scholar Crossref Search ADS WorldCat 7. Goncharov AF , Goldman N, Fried LE et al. Dynamic ionization of water under extreme conditions . Phys Rev Lett 2005 ; 94 : 125508 . Google Scholar Crossref Search ADS PubMed WorldCat 8. French M , Mattsson TR, Nettelmann N et al. Equation of state and phase diagram of water at ultrahigh pressures as in planetary interiors . Phys Rev B 2009 ; 79 : 054107 . Google Scholar Crossref Search ADS WorldCat 9. Redmer R , Mattsson TR, Nettelmann N et al. The phase diagram of water and the magnetic fields of Uranus and Neptune . Icarus 2011 ; 211 : 798 – 803 . Google Scholar Crossref Search ADS WorldCat 10. Ninet S , Datchi F, Saitta AM. Proton disorder and superionicity in hot dense ammonia ice . Phys Rev Lett 2012 ; 108 : 165702 . Google Scholar Crossref Search ADS PubMed WorldCat 11. Wilson HF , Wong ML, Militzer B. Superionic to superionic phase change in water: consequences for the interiors of Uranus and Neptune . Phys Rev Lett 2013 ; 110 : 151102 . Google Scholar Crossref Search ADS PubMed WorldCat 12. Bethkenhagen M , French M, Redmer R. Equation of state and phase diagram of ammonia at high pressures from ab initio simulations . J Chem Phys 2013 ; 138 : 234504 . Google Scholar Crossref Search ADS PubMed WorldCat 13. Bethkenhagen M , Cebulla D, Redmer R et al. Superionic phases of the 1:1 water–ammonia mixture . J Phys Chem A 2015 ; 119 : 10582 – 8 . Google Scholar Crossref Search ADS PubMed WorldCat 14. Sun J , Clark BK, Torquato S et al. The phase diagram of high-pressure superionic ice . Nat Commun 2015 ; 6 : 8156 . Google Scholar Crossref Search ADS PubMed WorldCat 15. French M , Desjarlais MP, Redmer R. Ab initio calculation of thermodynamic potentials and entropies for superionic water . Phys Rev E 2016 ; 93 : 022140 . Google Scholar Crossref Search ADS PubMed WorldCat 16. Jiang X , Wu X, Zheng Z et al. Ionic and superionic phases in ammonia dihydrate NH3·2H2O under high pressure . Phys Rev B 2017 ; 95 : 144104 . Google Scholar Crossref Search ADS WorldCat 17. Bethkenhagen M , Meyer ER, Hamel S et al. Planetary ices and the linear mixing approximation . Astrophys J 2017 ; 848 : 67 . Google Scholar Crossref Search ADS WorldCat 18. Millot M , Hamel S, Rygg JR et al. Experimental evidence for superionic water ice using shock compression . Nat Phys 2018 ; 14 : 297 . Google Scholar Crossref Search ADS WorldCat 19. Millot M , Coppari F, Rygg JR et al. Nanosecond X-ray diffraction of shock-compressed superionic water ice . Nature 2019 ; 569 : 251 – 5 . Google Scholar Crossref Search ADS PubMed WorldCat 20. Song X , Yin K, Wang Y et al. Exotic hydrogen bonding in compressed ammonia hydrides . J Phys Chem Lett 2019 ; 10 : 2761 – 6 . Google Scholar Crossref Search ADS PubMed WorldCat 21. Liu C , Gao H, Hermann A et al. Plastic and superionic helium ammonia compounds under high pressure and high temperature . Phys Rev X 2020 ; 10 : 021007 . Google Scholar OpenURL Placeholder Text WorldCat 22. Hirai H , Konagai K, Kawamura T et al. Polymerization and diamond formation from melting methane and their implications in ice layer of giant planets . Phys Earth Planet Inter 2009 ; 174 : 242 – 6 . Google Scholar Crossref Search ADS WorldCat 23. Gao G , Oganov AR, Ma Y et al. Dissociation of methane under high pressure . J Chem Phys 2010 ; 133 : 144508 . Google Scholar Crossref Search ADS PubMed WorldCat 24. Sherman BL , Wilson HF, Weeraratne D et al. Ab initio simulations of hot dense methane during shock experiments . Phys Rev B 2012 ; 86 : 224113 . Google Scholar Crossref Search ADS WorldCat 25. Conway LJ , Hermann A. High pressure hydrocarbons revisited: from van der Waals compounds to diamond . Geosciences 2019 ; 9 : 227 . Google Scholar Crossref Search ADS WorldCat 26. Kraus D , Vorberger J, Pak A et al. Formation of diamonds in laser-compressed hydrocarbons at planetary interior conditions . Nat Astronomy 2017 ; 1 : 606 – 11 . Google Scholar Crossref Search ADS WorldCat 27. Takii Y , Koga K, Tanaka H. A plastic phase of water from computer simulation . J Chem Phys 2008 ; 128 : 204501 . Google Scholar Crossref Search ADS PubMed WorldCat 28. Aragones JL , Vega C. Plastic crystal phases of simple water models . J Chem Phys 2009 ; 130 : 244504 . Google Scholar Crossref Search ADS PubMed WorldCat 29. Bove LE , Klotz S, Strässle TH et al. Translational and rotational diffusion in water in the gigapascal range . Phys Rev Lett 2013 ; 111 : 185901 . Google Scholar Crossref Search ADS PubMed WorldCat 30. Li B , Kawakita Y, Ohira-Kawamura S et al. Colossal barocaloric effects in plastic crystals . Nature 2019 ; 567 : 506 . Google Scholar Crossref Search ADS PubMed WorldCat 31. Dong X , Oganov AR, Goncharov AF et al. A stable compound of helium and sodium at high pressure . Nat Chem 2017 ; 9 : 440 – 5 . Google Scholar Crossref Search ADS PubMed WorldCat 32. Monserrat B , Martinez-Canales M, Needs RJ et al. Helium-iron compounds at terapascal pressures . Phys Rev Lett 2018 ; 121 : 015301 . Google Scholar Crossref Search ADS PubMed WorldCat 33. Liu Z , Botana J, Hermann A et al. Reactivity of He with ionic compounds under high pressure . Nat Commun 2018 ; 9 : 951 . Google Scholar Crossref Search ADS PubMed WorldCat 34. Zhang J , Lv J, Li H et al. Rare helium-bearing compound FeO2He stabilized at deep-earth conditions . Phys Rev Lett 2018 ; 121 : 255703 . Google Scholar Crossref Search ADS PubMed WorldCat 35. Gao H , Sun J, Pickard CJ et al. Prediction of pressure-induced stabilization of noble-gas-atom compounds with alkali oxides and alkali sulfides . Phys Rev Mater 2019 ; 3 : 015002 . Google Scholar Crossref Search ADS WorldCat 36. Loubeyre P , Jean-Louis M, LeToullec R et al. High pressure measurements of the He-Ne binary phase diagram at 296 K: evidence for the stability of a stoichiometric Ne(He)2 solid . Phys Rev Lett 1993 ; 70 : 178 – 81 . Google Scholar Crossref Search ADS PubMed WorldCat 37. Liu H , Yao Y, Klug DD. Stable structures of He and H2O at high pressure . Phys Rev B 2015 ; 91 : 014102 . Google Scholar Crossref Search ADS WorldCat 38. Teeratchanan P , Hermann A. Computational phase diagrams of noble gas hydrates under pressure . J Chem Phys 2015 ; 143 : 154507 . Google Scholar Crossref Search ADS PubMed WorldCat 39. Liu C , Gao H, Wang Y et al. Multiple superionic states in helium–water compounds . Nat Phys 2019 ; 15 : 1065 – 70 . Google Scholar Crossref Search ADS WorldCat 40. Vos WL , Finger LW, Hemley RJ et al. A high-pressure van der Waals compound in solid nitrogen-helium mixtures . Nature 1992 ; 358 : 46 – 8 . Google Scholar Crossref Search ADS WorldCat 41. Ninet S , Weck G, Loubeyre P et al. Structural and vibrational properties of the van der Waals compound (N2)11He up to 135 GPa . Phys Rev B 2011 ; 83 : 134107 . Google Scholar Crossref Search ADS WorldCat 42. Li Y , Feng X, Liu H et al. Route to high-energy density polymeric nitrogen t -N via He−N compounds . Nat Commun 2018 ; 9 : 722 . Google Scholar Crossref Search ADS PubMed WorldCat 43. Li D , Liu Y, Tian F et al. High-pressure structures of helium and carbon dioxide from first-principles calculations . Solid State Commun 2018 ; 283 : 9 – 13 . Google Scholar Crossref Search ADS WorldCat 44. Sans JA , Manjón FJ, Popescu C et al. Ordered helium trapping and bonding in compressed arsenolite: synthesis of As4O6·2He . Phys Rev B 2016 ; 93 : 054102 . Google Scholar Crossref Search ADS WorldCat 45. Nettelmann N , Wang K, Fortney JJ et al. Uranus evolution models with simple thermal boundary layers . Icarus 2016 ; 275 : 107 – 16 . Google Scholar Crossref Search ADS WorldCat 46. French M , Nettelmann N. Viscosity and prandtl number of warm dense water as in ice giant planets . Astrophys J 2019 ; 881 : 81 . Google Scholar Crossref Search ADS WorldCat 47. Xia K , Gao H, Liu C et al. A novel superhard tungsten nitride predicted by machine-learning accelerated crystal structure search . Sci Bull 2018 ; 63 : 817 – 24 . Google Scholar Crossref Search ADS WorldCat 48. Kresse G , Furthmüller J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set . Phys Rev B 1996 ; 54 : 11169 – 86 . Google Scholar Crossref Search ADS WorldCat 49. Perdew JP , Burke K, Ernzerhof M. Generalized gradient approximation made simple . Phys Rev Lett 1996 ; 77 : 3865 – 8 . Google Scholar Crossref Search ADS PubMed WorldCat 50. Wen X-D , Hand L, Labet V et al. Graphane sheets and crystals under pressure . Proc Natl Acad Sci USA 2011 ; 108 : 6833 – 7 . Google Scholar Crossref Search ADS WorldCat 51. Liu Y , Duan D, Tian F et al. Crystal structures and properties of the CH4H2 compound under high pressure . RSC Adv 2014 ; 4 : 37569 – 74 . Google Scholar Crossref Search ADS WorldCat 52. Saleh G , Oganov AR. Novel stable compounds in the C-H-O Ternary system at high pressure . Sci Rep 2016 ; 6 : 32486 . Google Scholar Crossref Search ADS PubMed WorldCat 53. Klimeš J , Bowler DR, Michaelides A. Chemical accuracy for the van der Waals density functional . J Phys Condens Matter 2010 ; 22 : 022201 . Google Scholar Crossref Search ADS PubMed WorldCat 54. Dion M , Rydberg H, Schröder E et al. Van der Waals density functional for general geometries . Phys Rev Lett 2004 ; 92 : 246401 . Google Scholar Crossref Search ADS PubMed WorldCat 55. Grimme S , Antony J, Ehrlich S et al. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu . J Chem Phys 2010 ; 132 : 154104 . Google Scholar Crossref Search ADS PubMed WorldCat 56. Bini R , Pratesi G. High-pressure infrared study of solid methane: phase diagram up to 30 GPa . Phys Rev B 1997 ; 55 : 14800 – 9 . Google Scholar Crossref Search ADS WorldCat 57. Spanu L , Donadio D, Hohl D et al. Theoretical investigation of methane under pressure . J Chem Phys 2009 ; 130 : 164520 . Google Scholar Crossref Search ADS PubMed WorldCat 58. Hernandez J-A , Caracas R. Proton dynamics and the phase diagram of dense water ice . J Chem Phys 2018 ; 148 : 214501 . Google Scholar Crossref Search ADS PubMed WorldCat 59. He X , Zhu Y, Mo Y. Origin of fast ion diffusion in super-ionic conductors . Nat Commun 2017 ; 8 : 15893 . Google Scholar Crossref Search ADS PubMed WorldCat 60. Zimmermann NER , Hannah DC, Rong Z et al. Electrostatic estimation of intercalant jump-diffusion barriers using finite-size ion models . J Phys Chem Lett 2018 ; 9 : 628 – 34 . Google Scholar Crossref Search ADS PubMed WorldCat 61. Bader RFW . A quantum theory of molecular structure and its applications . Chem Rev 1991 ; 91 : 893 – 928 . Google Scholar Crossref Search ADS WorldCat 62. Johnson ER , Keinan S, Mori-Sánchez P et al. Revealing noncovalent interactions . J Am Chem Soc 2010 ; 132 : 6498 – 506 . Google Scholar Crossref Search ADS PubMed WorldCat 63. Hermann A , Ashcroft NW, Hoffmann R. Isotopic differentiation and sublattice melting in dense dynamic ice . Phys Rev B 2013 ; 88 : 214113 . Google Scholar Crossref Search ADS WorldCat 64. Meyer ER , Ticknor C, Bethkenhagen M et al. Bonding and structure in dense multi-component molecular mixtures . J Chem Phys 2015 ; 143 : 164513 . Google Scholar Crossref Search ADS PubMed WorldCat 65. Schöttler M , Redmer R. Ab initio calculation of the miscibility diagram for hydrogen-helium mixtures . Phys Rev Lett 2018 ; 120 : 115703 . Google Scholar Crossref Search ADS PubMed WorldCat 66. Pickard CJ , Needs RJ. Ab initio random structure searching . J Phys Condensed Matter 2011 ; 23 : 053201 . Google Scholar Crossref Search ADS WorldCat 67. Togo A , Tanaka I. First principles phonon calculations in materials science . Scr Mater 2015 ; 108 : 1 – 5 . Google Scholar Crossref Search ADS WorldCat 68. Otero-de-la-Roza A , Johnson ER, Luaña V. Critic2: a program for real-space analysis of quantum chemical interactions in solids . Comp Phys Commun 2014 ; 185 : 1007 – 18 . Google Scholar Crossref Search ADS WorldCat © The Author(s) 2020. Published by Oxford University Press on behalf of China Science Publishing & Media Ltd. This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted reuse, distribution, and reproduction in any medium, provided the original work is properly cited. TI - Coexistence of plastic and partially diffusive phases in a helium-methane compound JF - National Science Review DO - 10.1093/nsr/nwaa064 DA - 2020-10-14 UR - https://www.deepdyve.com/lp/oxford-university-press/coexistence-of-plastic-and-partially-diffusive-phases-in-a-helium-7Gu8ynpfvC SP - 1540 EP - 1547 VL - 7 IS - 10 DP - DeepDyve ER -