Next Article in Journal
Optimal Charge Planning Model of Steelmaking Based on Multi-Objective Evolutionary Algorithm
Next Article in Special Issue
First-Principles Approaches to Metals, Alloys, and Metallic Compounds
Previous Article in Journal
Effect of Silicon on the Microstructure and Performance of the New Binary Deep Eutectic Ti–Cu–Zr–Ni-Based Filler Metal
Previous Article in Special Issue
Multiscale Modelling of Hydrogen Transport and Segregation in Polycrystalline Steels
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Stability, Electronic Structure, and Dehydrogenation Properties of Pristine and Doped 2D MgH2 by the First Principles Study

College of Science, Beijing University of Chemical Technology, Beijing 100029, China
*
Author to whom correspondence should be addressed.
Metals 2018, 8(7), 482; https://doi.org/10.3390/met8070482
Submission received: 30 May 2018 / Revised: 15 June 2018 / Accepted: 18 June 2018 / Published: 25 June 2018
(This article belongs to the Special Issue First-Principles Approaches to Metals, Alloys, and Metallic Compounds)

Abstract

:
Based on first principles calculations, we theoretically predict the new two-dimensional (2D) MgH2. The thermodynamic stability, partial density of states, electron localization function, and Bader charge of pure and the transition metal (Ti, V, and Mn) doped 2D MgH2 are investigated. The results show that all the systems are dynamically stable, and the dehydrogenation properties indicate that the decomposition temperature can be reduced by introducing the transition metal, and the Mn doped system exhibits good performance for better hydrogen storage and dehydrogenation kinetics.

Graphical Abstract

1. Introduction

Hydrogen energy is considered to be the most promising alternative because it is lightweight, environmentally friendly, highly efficient, renewable, and abundant on earth. However, the storage limits the application of hydrogen. Metal hydrides are considered as the most promising materials for hydrogen storage and have been widely investigated in the past decades [1]. Among them, magnesium-based alloys and magnesium hydrides can achieve the hydrogen storage capacity of 7.6 wt % [2,3,4,5,6,7,8]. However, the high thermodynamic stability (the heat of formation is around −75.99 kJ/mol·H2), high desorption temperatures (above 573 K), and slow dehydrogenation kinetics seriously limit the practical applications [3,9,10]. Therefore, it is always a central task to design new materials or adopt efficient strategies for achieving lower desorption temperatures and good dehydrogenation performances.
Previous studies show that the bonding nature of MgH2 is a mixture of strong ionic and weak covalent bonding [11], and weakening the interactions may be an effective strategy to improve dehydrogenation performance. It has been reported that doping with transition metal elements or their oxides mixtures with MgH2 can effectively reduce its stability and improve the hydrogen desorption thermodynamics [3,12,13,14,15,16,17]. Oelerich [15] et al. have reported that MgH2 milled with Fe3O4, V2O5, Mn2O3, or Cr2O3, etc. can accelerate the hydrogen desorption kinetics. Shang [3] et al. have studied the hydrogen storage performance of (MgH2 + M) systems (M = Al, Ti, Fe, Ni, Cu, and Nb) experimentally and theoretically, and they found that MgH2 mixed with those metals can reduce the stability and improve the hydrogen desorption kinetics. Nonetheless, the MgH2 systems still have a high desorption temperature around 500 K. It is noted that the bulk MgH2 has been extensively investigated, however, the single-layer magnesium hydrides have been largely ignored. Motivated by the above mentioned details, we focus on exploring new structures with good dehydrogenation performance in this work.
In this paper, the new two-dimensional (2D) MgH2 structure is theoretically predicted and studied by first principles calculations. The stabilities of pure and Ti/V/Mn doped MgH2 are discussed by the phonon spectra and heat of formation. The calculated heat of formation for pure and Ti/V/Mn doped 2D MgH2 are −37.57, −25.67, −18.14, and −23.90 kJ/mol·H2, respectively, which are significantly lower than that of −75.99 kJ/mol·H2 of bulk MgH2. The electronic structure and hydrogen desorption kinetics results show that the predicted two-dimensional magnesium hydride are promising candidates for hydrogen storage.

2. Computational Details

The structural optimization and electronic property calculations were performed using the projector augmented plane-wave method (PAW) based on the density functional theory (DFT) in the Vienna ab initio simulation package (VASP) [18,19]. The exchange-correlation potential was approximated by generalized gradient approximation (GGA) in the Perdew-Burke-Ernzerhof (PBE) form [20,21]. To avoid the interlayer effects of the c-axis, the vacuum region around 15 Å was set in all the systems. The energy cutoff of 600 eV and the 9 × 9 × 1 Γ-centered Monkhorst-Pack k-points [22] were employed for all calculations. The atomic positions were fully relaxed and the force tolerance between each atom was less than 0.01 eV/Å for the structural optimization. The convergence criteria of 10−6 eV per atom was applied to be self-consistent. Meanwhile, for calculation of electronic structures, we also applied the local density approximation (LDA) [23] and HSE06 [24] was functional. The kinetic stability was discussed using the phonon spectra calculations in PHONOPY code coupled with VASP using the density functional perturbation theory (DFPT) method [25,26,27].

3. Results and Discussion

Figure 1 shows the fully relaxed structure of the top and side view of pure 2D MgH2 of the hexagonal structure with space group P-3m1 ( D 3 d 3 ). The primitive cell has the lattice constant of a = b = 3.01 Å, the Mg-H bond length of l = 1.97 Å, and the buckled height of d = 1.86 Å. The next calculations were performed for the 3 × 3 × 1 supercell of 2D MgH2, named Mg9H18. The corresponding lattice parameters, Wyckoff [28] and atomic positions, are shown in Table 1. As is seen, there are nine Mg atoms located at 1b (Mg1), 6h (Mg2), and 2d (Mg3) sites, while the eighteen H atoms are located in three identical Wyckoff positions, i.e., 6i, as shown in Figure 1.
In this work, three different Mg sites are considered as possible positions for substitution doping. Meanwhile, defects are inevitable in synthesis or processing and can usually affect their properties [29,30,31,32,33]. The most common types of defect are vacancy defects, so we also considered the vacancies of Mg (Mg8H18) for comparison to the doped systems. The formation energies were calculated to determine the favorable positions of doping elements of Ti/V/Mn, which is defined as ΔE = Etot(Mg8H18Xn) − Etot(Mg9H18) − n Etot(X) + Etot(Mg), where Etot is the total energy of the system, the parameter n = 0/1 represents Mg vacancy, and X (X = Ti, V, and Mn) doped. The energies are listed in Table 2. It is noticed that the Mg8H18 and Ti/V/Mn doped systems have positive energy, indicating that the stability of all the systems are lower than that of pure Mg9H18. In addition, for the three high symmetry sites of Mg1 (1b), Mg2 (6h), and Mg3 (2d), the ΔE are nearly identical, therefore, we assume that all the doped-sites are located at the Mg1 site in the following work. The relaxed parameters and bond lengths of Mg9H18 and Mg8H18X (X = Ti, V, and Mn) are listed in Table 2, and for the detailed lattice parameters, see Table A1 (Appendix A). As is seen, the bond length of Mg-H is changed, which indicates that the doped X atoms break the symmetry of the 2D MgH2 structure.
Structural stability is discussed by the phonon spectra calculations using the DFPT method, as is shown in Figure 2. Clearly, there are no imaginary frequencies in the whole Brillouin zone, indicating that all the systems are dynamically stable. Meanwhile, the heat of formation (ΔH) [7,34,35,36] is one of the most fundamentally thermodynamic properties. The heat of formation can be obtained directly from the equation ΔH = [Etot(Mg9-nH18Xn+m) − (n + m) Etot(X) − (9-n) Etot(Mg) − 9 Etot(H2)]/9, where the parameters (n = 0, m = 0), (n = 1, m = −1), and (n = 1, m = 0), represent pure, Mg vacancy, and X (X = Ti, V and Mn) doped Mg9H18, respectively. The value of Etot(H2) of −6.762 eV in a 10 × 10 × 10 Å3 cubic cell is very close to −6.773 eV reported in Ref. [37].
The estimated heats of formation are listed in Table 3. As is seen, the heat of formation of Mg9H18, Mg8H18, Mg8H18Ti, Mg8H18V, and Mg8H18Mn are −37.57, 31.71, −25.67, −18.14, and −23.90 kJ/mol·H2, respectively. The results show that the stability decreased for the doped 2D MgH2, followed by Mg8H18Ti, Mg8H18Mn, and Mg8H18V, and Mg8H18 is the most unstable. In comparison, we also obtained the heat of formation of the bulk MgH2 of ΔH = −54.56 kJ/mol·H2, which is close to the theoretical values −54.4 in Ref. [36] and −53.85 kJ/mol·H2 in Ref. [38]. At the same time, we estimated the decomposition temperature according to the following relationship:   ln P P 0 =   Δ H R T Δ S R , where P, P0, R, T, and ΔS represent the pressure, the standard pressure, the gas constant, the decomposition temperature, and the entropy change, respectively. At the standard pressure, the ΔH is defined as ΔH = TΔS [39,40]. For most of the dehydrogenation reactions of simple metal hydrides, the ΔS is in the range of 95 J/mol·K < ΔS(H2) < 140 J/mol·K [41]. Consequently, the decomposition temperatures are 268 K < T(Mg9H18) < 396 K, 183 K < T(Mg8H18Ti) < 270 K, 130 K < T(Mg8H18V) < 191 K, 171 K < T(Mg8H18Mn) < 252 K, which are significantly lower than that of 573~673 K of bulk MgH2. The discussions mentioned above show that 2D MgH2 has better dehydrogenation thermodynamic properties than that of bulk MgH2, and doping with Ti, V, and Mn elements can reduce the stability and improve the dehydrogenation thermodynamics properties of 2D MgH2.
To understand the effect of Ti/V/Mn-doped 2D MgH2 well, we analyzed the electronic structures. The band structures were obtained using PBE, LDA, and HSE06 functionals and are shown in Figure A1. It can be seen that the pure Mg9H18 with the energy gap is 4.87 eV, which is smaller than the experimental values 5.16 eV [42] or 5.6 eV [43] of bulk MgH2. For comparison, we found that the bandgaps using HSE06 functional are larger than those using PBE and LDA functionals, and the bandgaps using LDA functional are close to those of the PBE functional for pure and vacancies of Mg9H18. For the doped systems, there are few energy bands across the Fermi level due to the d orbitals of the dopants. Figure 3 shows the total and partial density of states (PDOS) of Mg9H18 and Mg8H18, which are calculated using the PBE functional. We can see the stronger hybridization between H and Mg atoms near the Fermi level, which indicates the strong interaction between H and Mg atoms. For the Mg8H18X, the electronic structure is different from that of the pure Mg9H18, as is shown in Figure 4. We can see that the d orbitals of Ti/V/Mn are mainly located near the Fermi level in doped-2D MgH2, and there are few H-s orbitals and states of Mg atoms at the Fermi level, which indicates that the interactions between Ti/V/Mn and H/Mg atoms are relatively weaker. Meanwhile, since the H-s orbitals are reduced at the Fermi level, the hybridization between H and Mg atoms is weaker compared to pure Mg9H18.
In order to analyze the chemical bond of all the systems, the electron localization function (ELF) was calculated and is shown in Figure 5 with the isosurfaces of 0.6 e/Å3. As shown in Figure 5b–f, all the systems have similar features in that the ELF values are lower between Mg and H atoms. These suggest that the ionic bonds exist between Mg and H atoms, and the Mg atoms act as the charge donor, according well with the discussions of PDOS and ELF of bulk MgH2. The Bader charges were also calculated (see Table 3), and it can be seen that Mg atoms contributed two electrons, and H atoms acquired electrons to form anions. Meanwhile, the H atoms obtained fewer electrons due to the Mg vacancy and doping with Ti/V/Mn elements, thereby weakening the interactions between H and the metal atoms.
As mentioned above, doping with Ti/V/Mn elements reduces the stability of 2D MgH2 and weakens the interactions between H and metal atoms, which facilitates the release of hydrogen. To further understand the dehydrogenation behavior, the dehydrogenation energy was estimated by the formula: Ed = Etot(Mg9-nH17Xn+m) − Etot(Mg9-nH18Xn+m) + 1/2 Etot(H2), where (n = 0, m = 0), (n = 1, m = −1), and (n = 1, m = 0) represent the pure, Mg vacancy, and X (X = Ti, V, and Mn) doped Mg9H18, respectively. The dehydrogenation energies are listed in Table 3. The results show that the dehydrogenation energy of the Mg8H18 was significantly reduced compared to the pure and doped Mg9H18, while there are high ΔE of 2.968 eV and positive ΔH of 31.71 kJ/mol·H2, indicating that it is almost impossible to steadily occur. For doped systems, their dehydrogenation energies are significantly smaller than 1.589 eV of pure Mg9H18, especially Mg8H18Mn with the dehydrogenation energy of 0.853 eV. Therefore, doping with Ti/V/Mn elements can improve the dehydrogenation thermodynamic properties of 2D MgH2.

4. Conclusions

In summary, we theoretically predicted two-dimensional MgH2 and studied the electronic and dehydrogenation properties of pure and Ti/VMn doped 2D MgH2. The phonon spectra calculations indicate that all the systems are dynamically stable. The results of heat of formation suggests that Ti/V/Mn doping can reduce the thermodynamic stability, followed by Mg8H18Ti, Mg8H18Mn, and Mg8H18V, and Mg8H18 is the most unstable. Importantly, the dehydrogenation temperatures for all the systems are significantly lower than that of bulk MgH2 at 573~673 K. Especially, Mg8H18Ti (183~270 K), Mg8H18V (130~191 K), and Mg8H18Mn (171~252 K) have much lower decomposition temperature than that of pure 2D MgH2 (268~396 K), which is important for practical applications. The partial densities of states, electron localization function, and Bader charge calculation results show that Ti, V, and Mn elements can weaken the interaction between H and the metal atoms, which is favorable to dehydrogenation and better than that of the bulk MgH2.

Author Contributions

All authors contributed equally to this work.

Acknowledgments

This work was supported by the National Natural Science Foundation of China (Grant No. 11604008) and by BUCT Fund for Disciplines Construction (Project No. XK1702).

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A

Table A1. The relaxed structure parameters of pure, vacancies, and doped Mg9H18 using Perdew-Burke-Ernzerhof (PBE) and local density approximation (LDA) functionals.
Table A1. The relaxed structure parameters of pure, vacancies, and doped Mg9H18 using Perdew-Burke-Ernzerhof (PBE) and local density approximation (LDA) functionals.
Hydride PBELDA
a (Å)α (°)β (°)γ (°)a (Å)α (°)β (°)γ (°)
Mg9H189.03390.090.0120.08.89490.090.0120.0
Mg8H189.06290.090.0120.08.88890.090.0120.0
Mg8H18Ti9.02790.090.0120.08.88390.090.0120.0
Mg8H18V8.95190.090.0120.08.80390.090.0120.0
Mg8H18Mn8.81590.090.0120.08.66290.090.0120.0
Figure A1. Band structures of Mg9H18 (a); Mg8H18 (b); Mg8H18Ti (c); Mg8H18V (d); and Mg8H18Mn (e) calculated using PBE (black line), LDA (red dot line), and HSE06 (blue dot line) functionals, respectively.
Figure A1. Band structures of Mg9H18 (a); Mg8H18 (b); Mg8H18Ti (c); Mg8H18V (d); and Mg8H18Mn (e) calculated using PBE (black line), LDA (red dot line), and HSE06 (blue dot line) functionals, respectively.
Metals 08 00482 g0a1

References

  1. Sakintuna, B.; Lamari-Darkrim, F.; Hirscher, M. Metal hydride materials for solid hydrogen storage: A review. Int. J. Hydrogen Energy 2007, 32, 1121–1140. [Google Scholar] [CrossRef]
  2. Mohammed, Z.; Ahmed, R.; Mohammed Benali, K.; Bakhtiar ul, H.; Ahmad Radzi Mat, I.; Souraya, G.-S. First principle investigations of the physical properties of hydrogen-rich MgH2. Phys. Scr. 2013, 88, 065704. [Google Scholar] [CrossRef]
  3. Shang, C.X.; Bououdina, M.; Song, Y.; Guo, Z.X. Mechanical alloying and electronic simulations of (MgH2 + M) systems (M = Al, Ti, Fe, Ni, Cu and Nb) for hydrogen storage. Int. J. Hydrogen Energy 2004, 29, 73–80. [Google Scholar] [CrossRef]
  4. Ul Haq, B.; Kanoun, M.B.; Ahmed, R.; Bououdina, M.; Goumri-Said, S. Hybrid functional calculations of potential hydrogen storage material: Complex dimagnesium iron hydride. Int. J. Hydrogen Energy 2014, 39, 9709–9717. [Google Scholar] [CrossRef]
  5. Kumar, D.; Singh, A.; Prasad Tiwari, G.; Kojima, Y.; Kain, V. Thermodynamics and kinetics of nano-engineered Mg-MgH2 system for reversible hydrogen storage application. Thermochim. Acta 2017, 652, 103–108. [Google Scholar] [CrossRef]
  6. Trivedi, D.R.; Bandyopadhyay, D. Study of adsorption and dissociation process of H2 molecule on MgnRh clusters: A density functional investigation. Int. J. Hydrogen Energy 2016, 41, 20113–20121. [Google Scholar] [CrossRef]
  7. Song, Y.; Guo, Z.X.; Yang, R. Influence of selected alloying elements on the stability of magnesium dihydride for hydrogen storage applications: A first-principles investigation. Phys. Rev. B 2004, 69, 094205. [Google Scholar] [CrossRef]
  8. Vajeeston, P.; Ravindran, P.; Kjekshus, A.; Fjellvåg, H. Pressure-Induced Structural Transitions in MgH2. Phys. Rev. Lett. 2002, 89, 175506. [Google Scholar] [CrossRef] [PubMed]
  9. Kurko, S.; Matović, L.; Novaković, N.; Matović, B.; Jovanović, Z.; Mamula, B.P.; Grbović Novaković, J. Changes of hydrogen storage properties of MgH2 induced by boron ion irradiation. Int. J. Hydrogen Energy 2011, 36, 1184–1189. [Google Scholar] [CrossRef]
  10. Song, M.Y.; Kwon, S.N.; Park, H.R.; Hong, S.-H. Improvement in the hydrogen storage properties of Mg by mechanical grinding with Ni, Fe and V under H2 atmosphere. Int. J. Hydrogen Energy 2011, 36, 13587–13594. [Google Scholar] [CrossRef]
  11. Noritake, T.; Aoki, M.; Towata, S.; Seno, Y.; Hirose, Y.; Nishibori, E.; Takata, M.; Sakata, M. Chemical bonding of hydrogen in MgH2. Appl. Phys. Lett. 2002, 81, 2008–2010. [Google Scholar] [CrossRef]
  12. Liang, G.; Huot, J.; Boily, S.; Van Neste, A.; Schulz, R. Catalytic effect of transition metals on hydrogen sorption in nanocrystalline ball milled MgH2 − Tm (Tm = Ti, V, Mn, Fe and Ni) systems. J. Alloys Compd. 1999, 292, 247–252. [Google Scholar] [CrossRef]
  13. Shang, C.X.; Bououdina, M.; Guo, Z.X. Structural stability of mechanically alloyed (Mg + 10Nb) and (MgH2 + 10Nb) powder mixtures. J. Alloys Compd. 2003, 349, 217–223. [Google Scholar] [CrossRef]
  14. Rivoirard, S.; de Rango, P.; Fruchart, D.; Charbonnier, J.; Vempaire, D. Catalytic effect of additives on the hydrogen absorption properties of nano-crystalline MgH2(X) composites. J. Alloys Compd. 2003, 356, 622–625. [Google Scholar] [CrossRef]
  15. Oelerich, W.; Klassen, T.; Bormann, R. Metal oxides as catalysts for improved hydrogen sorption in nanocrystalline Mg-based materials. J. Alloys Compd. 2001, 315, 237–242. [Google Scholar] [CrossRef]
  16. Aguey-Zinsou, K.F.; Ares Fernandez, J.R.; Klassen, T.; Bormann, R. Effect of Nb2O5 on MgH2 properties during mechanical milling. Int. J. Hydrogen Energy 2007, 32, 2400–2407. [Google Scholar] [CrossRef]
  17. Song, M.; Bobet, J.-L.; Darriet, B. Improvement in hydrogen sorption properties of Mg by reactive mechanical grinding with Cr2O3, Al2O3 and CeO2. J. Alloys Compd. 2002, 340, 256–262. [Google Scholar] [CrossRef]
  18. Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for Ab Initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys. Rev. B Condens. Matter 1996, 54, 11169–11186. [Google Scholar] [CrossRef] [PubMed]
  19. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  20. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed]
  21. Hammer, B.; Hansen, L.B.; Nørskov, J.K. Improved adsorption energetics within density-functional theory using revised Perdew-Burke-Ernzerhof functionals. Phys. Rev. B 1999, 59, 7413–7421. [Google Scholar] [CrossRef] [Green Version]
  22. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  23. Perdew, J.P.; Zunger, A. Self-interaction correction to density-functional approximations for many-electron systems. Phys. Rev. B 1981, 23, 5048–5079. [Google Scholar] [CrossRef]
  24. Heyd, J.; Scuseria, G.E.; Ernzerhof, M. Erratum: Hybrid functionals based on a screened Coulomb potential. J. Chem. Phys. 2006, 118, 8207. [Google Scholar] [CrossRef]
  25. Togo, A.; Oba, F.; Tanaka, I. First-Principles Calculations of the Ferroelastic Transition Between Rutile-Type and CaCl2-Type SiO2 at High Pressures. Phys. Rev. B Condens. Matter 2008, 78. [Google Scholar] [CrossRef]
  26. Gonze, X.; Lee, C. Dynamical matrices, Born effective charges, dielectric permittivity tensors, and interatomic force constants from density-functional perturbation theory. Phys. Rev. B 1997, 55, 10355–10368. [Google Scholar] [CrossRef]
  27. Baroni, S.; de Gironcoli, S.; Dal Corso, A.; Giannozzi, P. Phonons and related crystal properties from density-functional perturbation theory. Rev. Mod. Phys. 2001, 73, 515–562. [Google Scholar] [CrossRef] [Green Version]
  28. Gu, T.; Wang, Z.; Tada, T.; Watanabe, S. First-principles simulations on bulk Ta2O5 and Cu/Ta2O5/Pt heterojunction: Electronic structures and transport properties. J. Appl. Phys. 2009, 106, 262907. [Google Scholar] [CrossRef]
  29. Sun, R.; Wang, Z.; Saito, M.; Shibata, N.; Ikuhara, Y. Atomistic mechanisms of nonstoichiometry-induced twin boundary structural transformation in titanium dioxide. Nat. Commun. 2011, 6, 7120. [Google Scholar] [CrossRef] [PubMed]
  30. Wang, Z.; Saito, M.; Mckenna, K.P.; Gu, L.; Tsukimoto, S.; Shluger, A.L.; Ikuhara, Y. Atom-resolved imaging of ordered defect superstructures at individual grain boundaries. Nature 2011, 479, 380–383. [Google Scholar] [CrossRef] [PubMed]
  31. McKenna, K.P.; Hofer, F.; Gilks, D.; Lazarov, V.K.; Chen, C.; Wang, Z.; Ikuhara, Y. Atomic-scale structure and properties of highly stable antiphase boundary defects in Fe3O4. Nat. Commun. 2014, 5, 5740. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, Z.; Saito, M.; Mckenna, K.P.; Fukami, S.; Sato, H.; Ikeda, S.; Ohno, H.; Ikuhara, Y. Atomic-Scale Structure and Local Chemistry of CoFeB-MgO Magnetic Tunnel Junctions. Nano Lett. 2016, 16, 1530–1536. [Google Scholar] [CrossRef] [PubMed]
  33. Yan, H.; Ziyu, H.; Xu, G.; Xiaohong, S. Structural, electronic and photocatalytic properties of atomic defective BiI3 monolayers. Chem. Phys. Lett. 2018, 691, 341–346. [Google Scholar] [CrossRef]
  34. García, G.N.; Abriata, J.P.; Sofo, J.O. Calculation of the electronic and structural properties of cubic Mg2NiH4. Phys. Rev. B 1999, 59, 11746–11754. [Google Scholar] [CrossRef]
  35. Chen, Y.; Dai, J.; Xie, R.; Song, Y.; Bououdina, M. First principles study of dehydrogenation properties of alkali/alkali-earth metal doped Mg7TiH16. J. Alloys Compd. 2017, 728, 1016–1022. [Google Scholar] [CrossRef]
  36. Shelyapina, M.G.; Fruchart, D.; Wolfers, P. Electronic structure and stability of new FCC magnesium hydrides Mg7MH16 and Mg6MH16 (M = Ti, V, Nb): An ab initio study. Int. J. Hydrogen Energy 2010, 35, 2025–2032. [Google Scholar] [CrossRef]
  37. Dai, J.H.; Song, Y.; Yang, R. First Principles Study on Hydrogen Desorption from a Metal (=Al, Ti, Mn, Ni) Doped MgH2 (110) Surface. J. Phys. Chem. C 2010, 114, 11328–11334. [Google Scholar] [CrossRef]
  38. Kumar, M.; Kamal, R.; Thapa, R. Screening based approach and dehydrogenation kinetics for MgH2: Guide to find suitable dopant using first-principles approach OPEN. Sci. Rep. 2017, 7, 15550. [Google Scholar] [CrossRef] [PubMed]
  39. Van Mal, H.H.; Buschow, K.H.J.; Miedema, A.R. Hydrogen absorption in LaNi5 and related compounds: Experimental observations and their explanation. J. Less Common Met. 1974, 35, 65–76. [Google Scholar] [CrossRef]
  40. Lakhal, M.; Bhihi, M.; Benyoussef, A.; El Kenz, A.; Loulidi, M.; Naji, S. The hydrogen ab/desorption kinetic properties of doped magnesium hydride MgH2 systems by first principles calculations and kinetic Monte Carlo simulations. Int. J. Hydrogen Energy 2015, 40, 6137–6144. [Google Scholar] [CrossRef]
  41. Alapati, S.V.; Johnson, J.K.; Sholl, D.S. Identification of Destabilized Metal Hydrides for Hydrogen Storage Using First Principles Calculations. J. Phys. Chem. B 2006, 110, 8769–8776. [Google Scholar] [CrossRef] [PubMed]
  42. Yu, R.; Lam, P.K. Electronic and structural properties of MgH2. Phys. Rev. B 1988, 37, 8730–8737. [Google Scholar] [CrossRef]
  43. Westerwaal, R.J.; Broedersz, C.P.; Gremaud, R.; Slaman, M.; Borgschulte, A.; Lohstroh, W.; Tschersich, K.G.; Fleischhauer, H.P.; Dam, B.; Griessen, R. Study of the hydride forming process of in-situ grown MgH2 thin films by activated reactive evaporation. Thin Solid Films 2008, 516, 4351–4359. [Google Scholar] [CrossRef]
Figure 1. The relaxed unit cell of Mg9H18. The primate cell is marked with a red dashed box.
Figure 1. The relaxed unit cell of Mg9H18. The primate cell is marked with a red dashed box.
Metals 08 00482 g001
Figure 2. The phonon spectra of Mg9H18 (a); Mg8H18Ti (b); Mg8H18V (c); and Mg8H18Mn (d).
Figure 2. The phonon spectra of Mg9H18 (a); Mg8H18Ti (b); Mg8H18V (c); and Mg8H18Mn (d).
Metals 08 00482 g002
Figure 3. The total and partial densities of states of pure (a) and defective (b) Mg9H18.
Figure 3. The total and partial densities of states of pure (a) and defective (b) Mg9H18.
Metals 08 00482 g003
Figure 4. The total and partial densities of states of Mg8H18Ti (a); Mg8H18V (b); and Mg8H18Mn (c).
Figure 4. The total and partial densities of states of Mg8H18Ti (a); Mg8H18V (b); and Mg8H18Mn (c).
Metals 08 00482 g004
Figure 5. (a) Structural representation of considered systems. The big (small) ball represents Mg (H) and the red ball site is the doped site; (bf) represent the electron localization function (ELF) of Mg9H18, Mg8H18, Mg8H18Ti, Mg8H18V, and Mg8H18Mn, respectively. The color bar represents the values of ELF.
Figure 5. (a) Structural representation of considered systems. The big (small) ball represents Mg (H) and the red ball site is the doped site; (bf) represent the electron localization function (ELF) of Mg9H18, Mg8H18, Mg8H18Ti, Mg8H18V, and Mg8H18Mn, respectively. The color bar represents the values of ELF.
Metals 08 00482 g005
Table 1. The relaxed structural parameters and atomic positions of Mg9H18.
Table 1. The relaxed structural parameters and atomic positions of Mg9H18.
Lattice ParametersAtomWyckoffAtomic Positions (Fractional)
Positionsxyz
164(P-3m1)Mg11b000.5
a = b = 9.033 ÅMg26h00.333330.5
c = 15 ÅMg32d0.333330.666670.5
d = 1.86 ÅH16i0.111110.222220.43783
α = β = 90°H26i0.222220.444440.56217
γ = 120°H36i0.111110.555560.43783
Table 2. The energy (ΔE), the lattice parameter (a), and bond length of Mg9H18, Mg8H18 and Mg8H18X (X = Ti, V, and Mn).
Table 2. The energy (ΔE), the lattice parameter (a), and bond length of Mg9H18, Mg8H18 and Mg8H18X (X = Ti, V, and Mn).
HydrideΔE (eV)ParameterBond Length (Å)
Mg1Mg2Mg3a (Å)Sub-H1Mg2-H1Mg2-H2Mg2-H3Mg3-H2Mg3-H3
Mg9H180009.0331.9721.9721.9721.9721.9721.972
Mg8H182.9682.9682.9689.062-1.8942.0431.9761.9471.992
Mg8H18Ti1.1131.1141.1149.0271.9151.9971.9641.9821.9461.990
Mg8H18V1.8181.8181.8198.9511.8221.9991.9451.9921.9391.983
Mg8H18Mn1.2791.2791.2798.8151.6912.0281.9112.0081.9371.965
Table 3. The heat of formation (ΔH), the decomposition temperature (T), Bader charge of Mg and H atoms, and the dehydrogenation energies (Ed) of Mg9H18, Mg8H18, and Mg8H18X (X = Ti, V, and Mn).
Table 3. The heat of formation (ΔH), the decomposition temperature (T), Bader charge of Mg and H atoms, and the dehydrogenation energies (Ed) of Mg9H18, Mg8H18, and Mg8H18X (X = Ti, V, and Mn).
HydrideΔHTBader Charge (e)Ed
(kJ/mol·H2)(K)MgXH(eV)
Mg9H18−37.57268~396+2.000-−0.9971.589
Mg8H1831.71-+2.000-−0.886−1.931
Mg8H18Ti−25.67183~270+2.000+1.825−0.9881.305
Mg8H18V−18.14130~191+2.000+1.523−0.9711.044
Mg8H18Mn−23.90171~252+2.000+0.975−0.9400.853

Share and Cite

MDPI and ACS Style

Gong, X.; Shao, X. Stability, Electronic Structure, and Dehydrogenation Properties of Pristine and Doped 2D MgH2 by the First Principles Study. Metals 2018, 8, 482. https://doi.org/10.3390/met8070482

AMA Style

Gong X, Shao X. Stability, Electronic Structure, and Dehydrogenation Properties of Pristine and Doped 2D MgH2 by the First Principles Study. Metals. 2018; 8(7):482. https://doi.org/10.3390/met8070482

Chicago/Turabian Style

Gong, Xu, and Xiaohong Shao. 2018. "Stability, Electronic Structure, and Dehydrogenation Properties of Pristine and Doped 2D MgH2 by the First Principles Study" Metals 8, no. 7: 482. https://doi.org/10.3390/met8070482

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop