Skip to main content
UKPMC Funders Author Manuscripts logoLink to UKPMC Funders Author Manuscripts
. Author manuscript; available in PMC: 2017 May 26.
Published in final edited form as: Nat Phys. 2016 May 9;12(9):835–838. doi: 10.1038/nphys3760

Crystal Structure of the Superconducting Phase of Sulfur Hydride

Mari Einaga 1,*,#, Masafumi Sakata 1, Takahiro Ishikawa 1, Katsuya Shimizu 1,#, Mikhail I Eremets 2,#, Alexander P Drozdov 2, Ivan A Troyan 2, Naohisa Hirao 3, Yasuo Ohishi 3
PMCID: PMC5446087  EMSID: EMS68272  PMID: 28553364

Abstract

A superconducting critical temperature above 200 K has recently been discovered in H2S (or D2S) under high hydrostatic pressure1, 2. These measurements were interpreted in terms of a decomposition of these materials into elemental sulfur and a hydrogen-rich hydride that is responsible for the superconductivity, although direct experimental evidence for this mechanism has so far been lacking. Here we report the crystal structure of the superconducting phase of hydrogen sulfide (and deuterium sulfide) in the normal and superconducting states obtained by means of synchrotron X-ray diffraction measurements, combined with electrical resistance measurements at both room and low temperatures. We find that the superconducting phase is mostly in good agreement with theoretically predicted body-centered cubic (bcc) structure for H3S (Ref.3). The presence of elemental sulfur is also manifest in the X-ray diffraction patterns, thus proving the decomposition mechanism of H2S to H3S + S under pressure46.


Recently, a very high Tc of 200 K has been discovered in hydrogen sulfide system1, 2. This work was initiated by prediction of a substantial superconductivity H2S7, which in turn arises from the idea that hydrogen dominant metallic alloys might be superconductors with high critical temperature similar to pure metallic hydrogen.8

The superconducting transition was proved by the sharp drop of the resistance to zero, a strong isotope effect in study of D2S, a shift of the superconducting transition with magnetic field, and finally by measuring the magnetic susceptibility and magnetization. As a likely explanation, the authors1, 2 suggested that H2S decomposes under pressure (with the assistance of temperature) to pure sulfur and some sulfur hydride with a higher content of hydrogen (such as SH4 or similar). At the same time, a theoretical work appeared which considered a different starting material (H2S)2H2 (stoichiometry H3S) and found R3m and Im-3m structures under pressure above 111 GPa and 180 GPa, respectively.3 These structures and other stoichiometric compounds were further carefully studied theoretically by different groups in numerous works4, 6, 925 and Tc ∼ 200 K was consistently obtained for the Im-3m structure. The calculated Tc as well as its pressure dependence9 are close to the experimental data1, 2. This suggests that the high Tc observed in the experiments relates not to H2S, but to the H3S in the Im-3m structure. Later calculations supported this idea: H2S is indeed unstable at high pressures and should decompose to sulfur and higher hydrides, most likely to H3S4, 6, 12. The goal of the present work is to check experimentally the structure of the superconducting hydrogen sulfide and compare with the theoretically predicted structure.

Samples were prepared in the same way as described in Ref.1, 2– H2S was loaded at temperatures of ∼ 200 K, then pressure was increased to ∼ 150 - 170 GPa and sample was annealed at room temperature. Typical X-ray diffraction (XRD) images of sulfur hydride and sulfur deuteride pressurized to 150 - 173 GPa are shown in Fig. 1. The XRD patterns of sulfur hydride and sulfur deuteride samples do not differ from each other. The diffraction patterns appear to be produced by two major phases. This clearly follows from the different pressure dependence of the peaks (Fig. 2 and Fig. SI3) and different variation of intensities while scanning the sample over its diameter (Fig. SI1): One group is fitted by elemental sulfur of β-Po structure26 and another group is described by the bcc-structure of H3S from the theoretical work3. We can conclude that H2S (D2S) solid most likely decomposes under pressure with route: 3H2S → 2H3S + S.

Figure 1. XRD of the sulfur hydride and sulfur deuteride samples.

Figure 1

a, Unrolled powder diffraction image of sulfur hydride at 150 GPa at room temperature recorded on the imaging plate. b, c, The integrated XRD patterns obtained with extraction of the background. The patterns of bcc H3S and β-Po elemental sulfur at 150 GPa and 170 GPa calculated according to Ref.5, 26 lie below the obtained patterns. The marks of star indicate the peaks which do not belong to the sample as follows from the scan over the sample (Fig. SI1): these peaks remain unchanged while the sample peaks change with radius of the sample both in position and intensity. Open circle indicates a reflection from high-pressure phase IV of elemental sulfur (incommensurately modulated body centered monoclinic structure). d, XRD patterns of sulfur deuteride at 173 GPa at 300 K and 13 K. The peaks marked by star (⋆) which do not the reflection from sample. The results of analyses are shown in table SI1 in supplementary information.

Figure 2. Pressure dependence of XRD in sulfur hydride and sulfur deuteride samples.

Figure 2

a, b, XRD patterns taken at room temperature and different pressures. Upper (red) and lower (green) ticks indicates the peak position of the predicted bcc structure of H3S and β-Po elemental sulfur, respectively. The peaks marked with star do not belong to the sample as follows from Fig. SI1. On decreasing pressure in sulfur hydride, it is considerable that the phase transition of elemental sulfur is observed - the peak from β-Po sulfur gradually disappears and that from phase IV (open circle) is enhanced. c, Pressure dependence of the atomic volume of sulfur hydride and deuteride. The experimental data were obtained with increasing of pressure and are fitted with first-order Birch EoS (the black solid line). The volumes of hexagonal (R3m) and bcc (Im-3m) phases obtained from the theoretical work3 are shown with solid squares (■) and triangle (▲), respectively, and the broken line. The estimated standard deviations are smaller than the size of the symbols.

The pressure dependence of the atomic volume, Vatm, of sulfur hydride and sulfur deuteride are shown in Fig. 2c. It is fitted by a first-order Birch equation of state27 with the bulk modulus B0 = 506 (30) GPa, and its pressure derivative B0’ = 6 (fixed). The value of experimentally observed Vatm is slightly larger but the compressibility is in good agreement with the Duan’s calculation3. The pressure dependence of the normalized atomic volume V/V0 of elemental sulfur in the β-Po structure is shown in Fig. SI3. It is in a good agreement with the experimental data of Ref.26 at high pressures P > 170 GPa, and with our DFT calculations (see Methods).

Our powder X-ray diffraction measurements do not allow us to distinguish between the predicted bcc-structures: Im-3m and R3m. In these structures the positions of sulfur atoms are the same and the only difference is position of hydrogen atoms: hydrogen atoms are situated symmetrically between the neighbor sulfur atoms in the Im-3m structure and slightly asymmetrical in the R3m structure (Fig. SI2). However, the position of the hydrogen atoms cannot be determined from the powder measurements as hydrogen atoms are extremely weak scatterers.

The low temperature data help with further analysis. We measured simultaneously the XRD and electrical resistance at the same setup28 (Fig. 3). The transition to the superconducting state was determined from the sharp drop of the resistance (Fig. 3a and 3b). We found that the normal and the superconducting state have the same structure as the XRD patterns are the same at room and low temperatures (Fig. 1d). Moreover, the structure of the sample visibly does not change over the pressure range 92 - 173 GPa. This is in a contrast with the dependence of the critical temperature on pressure which has a pronounced kink at 150 GPa for H3S and 160 GPa for D3S (Fig. 3c). This kink finds a natural explanation in the theoretical predictions9, 23: the pressure dependence of the critical superconducting temperature is different in the R3m phase at lower pressures and in the Im-3m phase at higher pressures. Our X-ray diffraction measurements support this interpretation as R3m and Im-3m phases differ only in ordering of hydrogen atoms and the same XRD patterns should be the same in the both pressure domains. Thus one can conclude that the highest critical temperature of 203 K2 corresponds to the Im-3m phase.

Figure 3. Pressure dependence of SC transition in sulfur hydride and sulfur deuteride.

Figure 3

a, b, Temperature dependence of resistance in sulfur hydride (on decease of pressure) and sulfur deuteride (on increase of pressure). c, Pressure dependence of critical temperature of superconductivity Tc of sulfur hydride (black points) and sulfur deuteride (red points). Open circles and squares are taken from Ref.2. The points marked with solid symbols are from the present work: the circles represent data on decreasing of pressure, the squares and triangles – at increasing of pressure. Broken lines (black for sulfur hydride and red for sulfur deuteride) indicate the phase boundary between R3m and Im-3m structural phase. The error bars indicate the difference between Tc onset and the zero-resistance temperature at each pressures (Fig. 3a and 3b).

Methods

The sample and electrical probes were prepared similar to Ref.2. Angle-dispersive powder x-ray diffraction measurements were carried out at SPring-8 (beamline BL10XU) with monochromatic beam of the energy of ∼ 30.0 keV (λ ∼ 0.412 0.414 Å). XRD and electrical resistance were measured simultaneously with the aid of a cryostat28. The diffraction patterns were recorded by using an imaging plate with an exposure time between 120 and 300 seconds. Four probe electrical measurements were performed with a AC-resistance bridge (Linear Research Inc., LR-700). We determined the pressure-volume dependence of β-Po sulfur by using first-principles calculations based on the density functional theory. The Quantum ESPRESSO code29 was used for the calculations, in which the Perdew-Burke-Ernzerhof generalized gradient approximation30 and the Vanderbilt ultrasoft pseudopotential31 were employed. The k-space integration over the Brillouin zone was performed on a 24 × 24 × 24 grid, and the energy cut-off of the plane wave basis was set at 80 Ry.

Supplementary Material

Supplementary Information

Acknowledgements

This work was performed under proposal No. 2015A0112 of the SPring-8. This research was supported by Japan Society for the Promotion of Science Grant-in-Aid for Specially Promoted Research, No. 26000006, JSPS KAKENHI, Grant-in-Aid for Young Scientists (B), No.15K17707 and the European Research Council 2010-Advanced Grant 267777.

Footnotes

Data Availability Statement

Raw data were generated at the SPring-8 synchrotron radiation facility (beamline BL10XU). Derived data supporting the findings of this study are available from the corresponding author upon request.

Author contributions

M. E. performed the whole XRD measurement and the data interpretation and writing the manuscript. M. S. performed the cryogenic operations and XRD date collections. K. S. performed the in situ electrical resistance measurements in XRD measurements and writing the manuscript. T. I. performed the support calculations for the data interpretation. M. I. E. designed the study and participated in XRD experiments and writing of the manuscript. A. P. D. prepared the sample in DAC for whole experiments. I. A. T. participated in building Raman setup. N. H. and Y. O. performed the optimization of synchrotron XRD and cryogenic operations. M. E., K. S. and M. I. E. contributed equally to this paper.

Competing financial interests

The authors declare that they have no competing financial interests.

References

  • 1.Drozdov AP, Eremets MI, Troyan IA. Conventional superconductivity at 190 K at high pressures. 2014 doi: 10.1038/nature14964. Preprint at http://arxiv.org/abs/1412.0460. [DOI] [PubMed]
  • 2.Drozdov AP, Eremets MI, Troyan IA, Ksenofontov V, Shylin SI. Conventional superconductivity at 203 kelvin at high pressures in the sulfur hydride system. Nature. 2015;525:73–77. doi: 10.1038/nature14964. [DOI] [PubMed] [Google Scholar]
  • 3.Duan D, et al. Pressure-induced metallization of dense (H2S)2H2 with high-Tc superconductivity. Scientific Reports. 2014;4:6968. doi: 10.1038/srep06968. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Bernstein N, Hellberg CS, Johannes MD, Mazin II, Mehl MJ. What superconducts in sulfur hydrides under pressure, and why. Phys Rev B. 2015;91:060511(R). [Google Scholar]
  • 5.Duan D, et al. Pressure-induced decomposition of solid hydrogen sulfide. Phys Rev B. 2015;91:180502(R). [Google Scholar]
  • 6.Errea I, et al. High-Pressure Hydrogen Sulfide from First Principles: A Strongly Anharmonic Phonon-Mediated Superconductor. Phys Rev Lett. 2015;114:157004. doi: 10.1103/PhysRevLett.114.157004. [DOI] [PubMed] [Google Scholar]
  • 7.Li Y, Hao J, Liu H, Li Y, Ma Y. The metallization and superconductivity of dense hydrogen sulfide. J Chem Phys. 2014;140:174712. doi: 10.1063/1.4874158. [DOI] [PubMed] [Google Scholar]
  • 8.Ashcroft NW. Hydrogen Dominant Metallic Alloys: High Temperature Superconductors? Phys Rev Lett. 2004;92:187002. doi: 10.1103/PhysRevLett.92.187002. [DOI] [PubMed] [Google Scholar]
  • 9.Akashi R, Kawamura M, Tsuneyuki S, Nomura Y, Arita R. First-principles study of the pressure and crystal-structure dependences of the superconducting transition temperature in compressed sulfur hydrides. Phys Rev B. 2015;91:224513. [Google Scholar]
  • 10.Flores-Livas JA, Sanna A, Gross EKU. High temperature superconductivity in sulfur and selenium hydrides at high pressure. Eur Phys J B. 2016;89:63. [Google Scholar]
  • 11.Ge Y, Zhang F, Yao Y. Possible Superconductivity Approaching Ice Point. 2015 Preprint at http://arxiv.org/abs/1507.08525.
  • 12.Papaconstantopoulos DA, Klein BM, Mehl MJ, Pickett WE. Cubic H3S around 200 GPa: An atomic hydrogen superconductor stabilized by sulfur. Phys Rev B. 2015;91:184511. [Google Scholar]
  • 13.Li Y, et al. Dissociation products and structures of solid H2S at strong compression. Phys Rev B. 2016;93:020103(R). [Google Scholar]
  • 14.Quan Y, Pickett WE. Van Hove singularities and spectral smearing in high-temperature superconducting H3S. Phys Rev Lett. 2016;93:104526. [Google Scholar]
  • 15.Bianconi A, Jarlborg T. Lifshitz transitions and zero point lattice fluctuations in sulfur hydride showing near room temperature superconductivity. Nov Supercond Mater. 2015;1:37–49. [Google Scholar]
  • 16.Nicol EJ, Carbotte JP. Comparison of pressurized sulfur hydride with conventional superconductors. Phys Rev B. 2015;91:220507(R). [Google Scholar]
  • 17.Degtyarenko N, Mazur E. Causes of high-temperature superconductivity in the hydrogen sulfide electron-phonon system. 2015 Preprint at http://arxiv.org/abs/1507.05749.
  • 18.Durajski AP, Szczéşniak R, Li Y. Non-BCS thermodynamic properties of H2S superconductor. Physica C. 2015;515:1–6. [Google Scholar]
  • 19.Durajski AP, Szczéşniak R, Pietronero L. High-temperature study of super-conducting hydrogen and deuterium sulfide. Ann Phys (Berlin) 2015:1–7. [Google Scholar]
  • 20.Gor’kov L, Kresin V. Pressure and high Tc superconductivity: applications to sulfur hydrides. 2015 doi: 10.1038/srep25608. Preprint at http://arxiv.org/abs/1511.06926. [DOI] [PMC free article] [PubMed]
  • 21.Hirsch JE, Marsiglio F. Hole superconductivity in H2S and other sulfides under high pressure. Physica C: Superconductivity and its Applications. 2015;511:45–49. [Google Scholar]
  • 22.Baňackỳ P. On the mechanism of high-temperature superconductivity in hydrogen sulfide at 200 GPa: Transition into superconducting anti-adiabatic state in coupling to H-vibrations. Results in Physics. 2016;6:1–2. [Google Scholar]
  • 23.Errea I, et al. Quantum hydrogen-bond symmetrization in the superconducting hydrogen sulfide system. Nat Lett. 2016 doi: 10.1038/nature17175. published online. [DOI] [PubMed] [Google Scholar]
  • 24.Akashi R, Sano W, Arita R, Tsuneyuki S. Possible Magnéli phases and self-alloying in the superconducting sulfur hydride. 2015 doi: 10.1103/PhysRevLett.117.075503. Preprint at http://arxiv.org/abs/1512.06680. [DOI] [PubMed]
  • 25.Ishikawa T, Nakanishi A, Shimizu K, Katayama-Yoshida H, Oda T, Suzuki N. Superconducting H5S2 phase in sulfur-hydrogen system under high-pressure. Sci Rep. 2016;6:23160. doi: 10.1038/srep23160. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Luo H, Greene RG, Ruoff AL. β-Po phase of sulfur at 162 GPa: X-ray diffraction study to 212 GPa. Phys Rev Lett. 1993;71:2943. doi: 10.1103/PhysRevLett.71.2943. [DOI] [PubMed] [Google Scholar]
  • 27.Birch F. Finite strain isotherm and velocities for single-crystal and polycrystalline NaCl at high pressures and 300° K. J Geo Res: Solid Earth. 1978;83:1257–1268. [Google Scholar]
  • 28.Matsuoka T, et al. Pressure-induced reentrant metallic phase in lithium. Phys Rev B. 2014;89:144103. [Google Scholar]
  • 29.Giannozzi P, et al. QUANTUM ESPRESSO: a modular and open-source software project for quantum simulations of materials. J Phys: Cond Matt. 2009;21:395502. doi: 10.1088/0953-8984/21/39/395502. [DOI] [PubMed] [Google Scholar]
  • 30.Perdew JP, Burke K, Ernzerof M. Generalized gradient approximation made simple. Phys Rev Lett. 1996;77:3865–3868. doi: 10.1103/PhysRevLett.77.3865. [DOI] [PubMed] [Google Scholar]
  • 31.Vanderbilt D. Soft self-consistent pseudopotentials in a generalized eigenvalue formalism. Phys Rev B. 1990;41:7892–7895. doi: 10.1103/physrevb.41.7892. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary Information

RESOURCES