DOI:
10.1039/C5RA11459D
(Paper)
RSC Adv., 2015,
5, 59292-59296
Compressed sodalite-like MgH6 as a potential high-temperature superconductor
Received
16th June 2015
, Accepted 1st July 2015
First published on 2nd July 2015
Abstract
Recently, an experimental work reported a very high Tc of ∼190 K in hydrogen sulphide (H2S) at 200 GPa. The search for new superconductors with high superconducting critical temperatures in hydrogen-dominated materials has attracted significant attention. Here we predict a candidate phase of MgH6 with a sodalite-like framework in conjunction with first-principles electronic structure calculations. The calculated formation enthalpy suggests that it is thermodynamically stable above 263 GPa relative to MgH2 and solid hydrogen (H2). Moreover, the absence of imaginary frequency in phonon calculations implies that this MgH6 structure is dynamically stable. Furthermore, our electron–phonon coupling calculation based on BCS theory indicates that this MgH6 phase is a conventional superconductor with a high superconducting critical temperature of ∼260 K under high pressure, which is even higher than that of the recently reported compressed H2S. The present results offer insight in understanding and designing new high-temperature superconductors.
Introduction
The search for new superconductors with high critical temperatures (Tc) has attracted great intention due to the discovery of superconductivity of mercury in 1911.1 For conventional Bardeen–Cooper–Schrieffer (BCS) superconductors, MgB2 was found to possess a Tc of 39 K.2 However, there are no other conventional superconducting compounds to be found to have a higher Tc than that of MgB2. For other unconventional superconducting compounds, the last several decades have witnessed an exciting revolution in searching high-temperature superconductivity materials, such as cuprates and iron-base superconductivity materials, but their superconducting mechanism cannot be explained by Bardeen–Cooper–Schrieffer (BCS) theory and still raises a source of debate.3–5 Ashcroft suggested the metallic phase of solid hydrogen should possess high-temperature conductivity.6 Until now, solid hydrogen has not been found to become metallic until up to 360 GPa in a laboratory.7 It is noted that a recent theoretical work proposed that H2S may become a good superconductor above 100 GPa.8 Soon after that a high-pressure electrical measurement proved that the compressed H2S can become a superconductor with a very high Tc of ∼190 K at 200 GPa.9 On the other hand, another theoretical work also suggest that (H2S)2H2 is a good superconductor with a Tc of 191–204 K at 200 GPa.10 There have also been many theoretical efforts studying the superconducting mechanism of H2S and investigating whether the compressed H2S is stable and dissociates into other sulfur hydrides.11–18 At present, the superconducting mechanism of compressed H2S is not quite clear and remains elusive. Therefore, investigates of other relevant hydrides are important as well as helpful in shedding light on the intricate superconducting mechanism and providing ideas for new potential high-temperature superconductor design.
Recently, CaH6 has been predicted to have a good superconductivity of 235 K at 150 GPa.19 This structure consists of a sodalite-like hydrogen cage filled with interstitial calcium (Ca) atoms. Due to the similarity between Mg and Ca, Mg may also adopt a sodalite-like structure with hydrogen atoms. Therefore, it is interesting to explore the stability of sodalite MgH6 structure and its potential superconducting behaviors. Moreover, another recent theoretical work also suggests that YH6 adopts this sodalite structure with a Tc of 251–264 K at 120 GPa.20 A recent study based on evolutionary structure searches suggested that MgH4 can be stable with respect to decomposition into MgH2 and H2 near 100 GPa, and it remains the most stable stoichiometry until at least 200 GPa.21
In this work, we study the stability of MgH6 relative to MgH2 and H2 on the basis of first-principle density functional theory. The calculations indicate that the MgH6 is dynamically stable and thermodynamically stable above 263 GPa. Further electron–phonon coupling calculations imply that this MgH6 phase possesses very good superconductivity with a high superconducting critical temperature of ∼260 K above 300 GPa. Our current results also show that the charge transfer between Mg and H atoms and H–H interactions are responsible for this good superconductivity. We find that the calculated Tc of this MgH6 phase increases up to above 300 GPa.
Results
Inspired by our previous work19 predicting high superconductivity of CaH6 where hydrogen atoms adopt a sodalite-like framework that is responsible for high superconductivity, we have replaced Ca atoms by Mg atoms in this sodalite-like structure. This structure was optimized with fully relaxing the atoms and lattice parameters within first-principles electronic structure framework, which produces a body centered cubic (bcc) crystal structure and space group of Im
m with 14 atoms per unit cell, in agreement with our previous work.19 To investigate the thermodynamically stability, we have calculated the formation enthalpy relative to MgH2 and H2 (Fig. 1), where previously predicted P63/mmc Ni2In-type structure21 was used for MgH2 and C2/c and Cmca
2 structure22 were used for solid hydrogen. Interestingly, this MgH6 is stable above 322 GPa which is higher than 150 GPa in CaH6. This is not unreasonable since the light elements adopt low-pressure structures for heavy elements.23 On the other hand, MgH6 is a hydrogen-dominated material where hydrogen is the lightest element and has a large zero-point (ZP) energy that may revise the stability range at high pressures. To account for the ZP effects, we estimated the ZP energies for MgH6, MgH2 and H2, using a harmonic method.24,25 As shown in the inset of Fig. 1, the inclusion of the ZP energies does shift down the stability of this MgH6 phase from 322 GPa to 263 GPa. This indicates that the ZP effect is further helpful to stabilize this newly predicted MgH6 structure.
 |
| Fig. 1 The calculated formation enthalpy of MgH6 relative to MgH2 and H2. The inset shows the formation energy of MgH6 with considering zero-point energy effects. | |
The sodalite-like MgH6 structure is shown in Fig. 2a, each hydrogen atom has four neighbor hydrogen atoms with distance of 1.1 Å at 300 GPa. The H and Mg atom situate 2d position (0, 0.25, 0.5) and 2a position (0, 0, 0), respectively. The distance between hydrogen atoms is similar to the distance in atomic phases of solid hydrogen (e.g. 1.0 Å in Cs-IV structure of solid hydrogen at 500 GPa). It is noted that one previous theoretical work suggests that the atomic phases of solid hydrogen possess good superconductivity under high pressure.26 The similar bonding behaviors suggest that MgH6 may be a good superconductor as well. To further gain insight into the interactions between hydrogen atoms, we have calculated the electron localization function (ELF).27 The topological analysis of ELF is commonly used to determine the degree of electron localization and subsequently, the tendency to form two electron covalent bonds in molecules and solids. Represented in a convenient scale, ELF = 1 corresponds to perfect localization and at the low limit ELF = 0.5 reflects the behavior of a homogeneous electron gas. The contour plot of the ELF for this MgH6 structures is shown in Fig. 2b. The strong interaction between hydrogen atoms can be clearly seen. Moreover, we have carried out Bader calculations on this MgH6 structure using the quantum theory of atoms-in-molecules (AIM).28 In the AIM theory, an atom within a solid is defined through the “zero-flux” condition of the electron density gradient. The results suggest that each Mg atom donates 1.63 electrons and each hydrogen atom attracts 0.27 electrons. This charge transfer may help enhance the interactions between hydrogen atoms.
 |
| Fig. 2 (a) The structure of MgH6. The small and big spheres represent H and Mg atoms, respectively. (b) The 2-dimensional electron localization function in {100} plane is shown with a 2 × 2 supercell at 300 GPa. | |
Discussion
To investigate the electronic states, band structure for this MgH6 structures was calculated as shown in Fig. 3. The results clearly show that this phase is a metallic phase, which is in good agreement with the previous results.19 The projected density of states in Fig. 3 show that the H_s orbit is also important for the contribution of this metallization. There are two bands crossing the Fermi level (Fig. 3). This results in a high electronic density states at the Fermi energy (Fig. 3). Moreover, there are two parallel bands along the H–N–P direction, which is helpful to form Fermi nesting that can induce strong electron–phonon coupling. The band dispersion along P–Γ–N direction is parabolic-like and close to the Fermi level. One band crossing along the N–P direction with steep slopes indicates the existence of mobile electrons along these symmetry directions. The co-existence of electrons with large effect masses (flat band around Γ and P) and itinerant electrons with high mobility (steep bands in H–N, N–P and P–Γ) suggested the possibility of high superconductivity.
 |
| Fig. 3 The calculated band structure and density of states of MgH6 structure at 300 GPa. | |
As mentioned above, the electronic band structure of MgH6 shows the existence of sharp and flat bands crossing the Fermi energy level, indicative of its possible superconductivity. In order to investigate the superconductivity of MgH6, phonon band structures, phonon density of states (PHDOS) and Eliashberg spectral function α2F(ω) for MgH6 were carried out. As shown in Fig. 4. There is no imaginary frequency to be found in the whole Brillouin zone (BZ), confirming that it is a dynamically stable structure.
 |
| Fig. 4 The calculated phonon band structure, phonon density of states and Eliashberg spectral function of MgH6 at 300 GPa. | |
Circles with a radius proportional to the electron–phonon coupling strength were also plotted in Fig. 4, to illustrate the contributions associated with different phonon modes. One can observe that nearly all phonon modes contribute to the whole Brillouin zone, reflecting a three-dimensional superconducting nature of the MgH6 structure. Moreover, the superconductivity in MgH6 is associated with the Kohn anomalies observed in the phonon dispersion (Fig. 4). At 300 GPa (Fig. 4b), the electron–phonon coupling parameter (λ) for the MgH6 structure is 3.29 with an average phonon frequency ln(ω) of 1450 K. Using the Allen–Dynes equation,29 which is an extension of the McMillan theory,30 with a nominal Coulomb pseudopotential parameter (μ*) of 0.12 the estimated superconducting critical temperature Tc is 263 K.
We have also studied the Tc of MgH6 as a function of pressure. As shown in Fig. 5, the Tc increases at the pressure range of 300 to 400 GPa. Furthermore, it is found that the calculated λ decreases with increasing pressure, but the average frequency increases as expected. The competing between the λ and the average frequency results in the increased Tc with pressure. The origin of the superconductivity could be traced back by a comparison of the calculated Eliashberg spectral function (α2F(ω)) and PHDOS. As shown in Fig. 4b, 15% of the electron phonon coupling is contributed to the low-frequency vibrations in the region from 0 to 600 cm−1, which are mostly from the vibrations of Mg atoms. The remained 85% derives from high-frequency vibrations from 600 to 2600 cm−1 which are predominately H–H stretching and bending modes. The superconducting critical temperature of 263 K at 300 GPa is higher than 39 K2 in MgB2 and even 190 K9 in recently measured solid H2S.
 |
| Fig. 5 Superconducting temperature (Tc) vs. pressure. The coupling parameters, the average phonon frequencies ωln and superconducting temperatures of MgH6 as a function of pressure. | |
Finally, we should point out that our calculated phase transition pressure (263 GPa) of MgH6 is higher than that (∼100 GPa) of MgH4,21 which means that MgH4 is more thermodynamically stable than MgH6 at low pressures. Furthermore, we have carried out the calculations on formation enthalpy of MgH6 relative to MgH4 and H2 as shown in Fig. 6. It shows that MgH6 actually thermodynamically stable relative to MgH4 and H2 above 325 GPa with considering ZP effects. However, it is noteworthy that MgH6 is thermodynamically stable above 263 GPa relative to MgH2 and H2. Moreover, our phonon calculation confirms that this newly predicted phase is still a dynamically stable structure. These results suggest MgH6 may be synthesized above 263 GPa with starting materials of MgH2 and H2. We also envisage that our predicted MgH6 structure could possess a very high superconducting temperature. This work will benefit the study of superconducting mechanism for conventional BCS superconductors.
 |
| Fig. 6 The calculated formation enthalpy of MgH6 relative to MgH2 and H2. The inset shows the formation energy of MgH6 with considering zero-point energy effects. | |
Conclusions
In summary, we have proposed a sodalite-like MgH6 structure within first-principles electronic structure framework. This newly predicted MgH6 structure was found to be thermodynamically stable above 263 GPa on the basis of the formation enthalpy calculations relative to MgH2 and H2. The phonon calculations suggest that this MgH6 structure is dynamically stable due to the absence of imaginary frequency in the whole BZ. Furthermore, our electron–phonon coupling calculations reveal that MgH6 is a good BCS superconductor with a high superconducting critical temperature of ∼260 K above 300 GPa. The electron–phonon calculations also suggest the three-dimensional superconducting nature in this newly predicted MgH6 structure. The calculated Tc of MgH6 was found to increase at a pressure range from 300 to 400 GPa with the decrease of the calculated λ and the increase of average frequency under compression. Our finding will stimulate future experimental study on synthesis of magnesium hydrides and explore its high-Tc superconductivity under high pressure.
Computational details
Structural optimizations and enthalpy calculations were carried out with the Vienna ab initio simulation (VASP)31–33 program and projector-augmented planewave (PAW)34 potentials employing the Perdew–Burke–Ernzerhof (PBE)35 functional. The valence configurations for the H and Mg potentials are 1s1 and 2s22p63s2, respectively. A plane wave basis set with an energy cutoff of 800 eV was used. Dense k-point meshes were employed to sample the first Brillouin zone (BZ) to ensure the energies are converged within 5 meV per atom. The elements of the interatomic force constant (IFC) and electron–phonon coupling matrices were computed using the linear response method with a 4 × 4 × 4 q-point mesh and 24 × 24 × 24 k-point mesh for the first Brillouin zone integrations.36 A plane wave basis set with an energy cutoff of 80 Ry was used.
Acknowledgements
H. W. appreciatively acknowledges the financial support by Natural Science Foundation of China under Grants no. 11474128 and the 2013 Program for New Century Excellent Talents in University. Part of the calculations has been performed by the use of computing resources provided by WestGrid and Compute Canada. Hanyu Liu acknowledges the support from the University of Saskatchewan research computing group and the use of the HPC resources (Plato machine). Work at Carnegie was partially supported by EFree, an Energy Frontier Research Center funded by the DOE, Office of Science, Basic Energy Sciences under Award No. DE-SC-0001057 (salary support for H.L). The infrastructure and facilities used at Carnegie were supported by NNSA Grant No. DE-NA-00006, CDAC.
References
- D. van Delft and P. Kes, Phys. Today, 2010, 63, 38–43 CrossRef PubMed.
- J. Nagamatsu, N. Nakagawa, T. Muranaka, Y. Zenitani and J. Akimitsu, Nature, 2001, 410, 63–64 CrossRef CAS PubMed.
- G. Sun, K. Wong, B. Xu, Y. Xin and D. Lu, Phys. Lett. A, 1994, 192, 122–124 CrossRef CAS.
- Y. Kamihara, H. Hiramatsu, M. Hirano, R. Kawamura, H. Yanagi, T. Kamiya and H. Hosono, J. Am. Chem. Soc., 2006, 128, 10012–10013 CrossRef CAS PubMed.
- Y. Kamihara, T. Watanabe, M. Hirano and H. Hosono, J. Am. Chem. Soc., 2008, 130, 3296–3297 CrossRef CAS PubMed.
- N. W. Ashcroft, Phys. Rev. Lett., 1968, 21, 1748 CrossRef CAS.
- C. S. Zha, Z. Liu and R. J. Hemley, Phys. Rev. Lett., 2012, 108, 146402 CrossRef.
- Y. Li, J. Hao, H. Liu, Y. Li and Y. Ma, J. Chem. Phys., 2014, 140, 174712 CrossRef PubMed.
- A. Drozdov, M. Eremets and I. Troyan, arXiv preprint arXiv:1412.0460, 2014.
- D. Duan, Y. Liu, F. Tian, D. Li, X. Huang, Z. Zhao, H. Yu, B. Liu, W. Tian and T. Cui, Sci. Rep., 2014, 4, 6968 CrossRef CAS PubMed.
- J. Hirsch and F. Marsiglio, Phys. C, 2015, 511, 45–49 CrossRef CAS PubMed.
- D. Papaconstantopoulos, B. Klein, M. Mehl and W. Pickett, Phys. Rev. B, 2015, 91, 184511 CrossRef.
- D. Duan, X. Huang, F. Tian, D. Li, H. Yu, Y. Liu, Y. Ma, B. Liu and T. Cui, Phys. Rev. B, 2015, 91, 180502 CrossRef.
- D. Papaconstantopoulos, B. Klein, M. Mehl and W. Pickett, Phys. Rev. B, 2015, 91, 184511 CrossRef.
- V. V. Struzhkin, Phys. C, 2015, 514, 77–85 CrossRef CAS PubMed.
- N. Bernstein, C. S. Hellberg, M. Johannes, I. Mazin and M. Mehl, Phys. Rev. B, 2015, 91, 060511 CrossRef.
- I. Errea, M. Calandra, C. J. Pickard, J. Nelson, R. J. Needs, Y. Li, H. Liu, Y. Zhang, Y. Ma and F. Mauri, Phys. Rev. Lett., 2015, 114, 157004 CrossRef.
- A. P. Durajski, R. Szcze and Y. Li, Phys. C, 2015, 515, 1–6 CrossRef CAS PubMed.
- H. Wang, S. T. John, K. Tanaka, T. Iitaka and Y. Ma, Proc. Natl. Acad. Sci., 2012, 109, 6463–6466 CrossRef CAS PubMed.
- Y. Li, J. Hao, H. Liu, S. T. John, Y. Wang and Y. Ma, Sci. Rep., 2015, 5, 9948 CrossRef PubMed.
- D. C. Lonie, J. Hooper, B. Altintas and E. Zurek, Phys. Rev. B, 2013, 87, 054107 CrossRef.
- C. J. Pickard and R. J. Needs, Nat. Phys., 2007, 3, 473–476 CrossRef CAS PubMed.
- M. I. McMahon and R. J. Nelmes, Chem. Soc. Rev., 2006, 35, 943–963 RSC.
- H. Liu, H. Wang and Y. Ma, J. Phys. Chem. C, 2012, 116, 9221–9226 CAS.
- H. Liu, L. Zhu, W. Cui and Y. Ma, J. Chem. Phys., 2012, 137, 074501 CrossRef PubMed.
- J. M. McMahon and D. M. Ceperley, Phys. Rev. B, 2011, 84, 144515 CrossRef.
- A. D. Becke and K. E. Edgecombe, J. Chem. Phys., 1990, 92, 5397–5403 CrossRef CAS PubMed.
- R. F. Bader, Acc. Chem. Res., 1985, 18, 9–15 CrossRef CAS.
- R. D. P. B. Allen, Phy. Rev. B, 1975, 12, 905 CrossRef.
- R. Dynes, Solid State Commun., 1972, 10, 615 CrossRef CAS.
- G. Kresse and J. Furthmüller, Phys. Rev. B, 1996, 54, 11169 CrossRef CAS.
- G. Kresse and J. Hafner, Phys. Rev. B, 1994, 49, 14251 CrossRef CAS.
- G. Kresse and J. Hafner, Phys. Rev. B, 1993, 47, 558 CrossRef CAS.
- P. E. Blochl, Phys. Rev. B, 1994, 50, 17953 CrossRef.
- J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef CAS.
- P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car, C. Cavazzoni and D. Ceresoli, J. Phys.: Condens. Matter, 2009, 21, 395502 CrossRef PubMed.
|
This journal is © The Royal Society of Chemistry 2015 |
Click here to see how this site uses Cookies. View our privacy policy here.