DOI:
10.1039/C6RA13856J
(Paper)
RSC Adv., 2016,
6, 78678-78683
Pressure-induced formation of hydrogen bonds in KNH2 studied by first principles
Received
28th May 2016
, Accepted 9th August 2016
First published on 9th August 2016
Abstract
Using particle swarm optimization technique implemented in the CALYPSO code, we have performed systematic research for the structures of KNH2 at pressures up to 20 GPa. Here, phase transition from the ground-state α-KNH2 (monoclinic P21/m) to β-KNH2 (monoclinic P21) occurs at 4 GPa; then, with pressure increasing up to 6.8 GPa, β-KNH2 transforms into γ-KNH2 (monoclinic Pc). By analyzing the partial density of states and charge density, ionic bonding nature between K+ and [NH2]− is revealed and there exists a strong covalent bonding between N and H atoms in NH2 groups in the three structures. Moreover, N–H⋯N hydrogen bonding between neighboring NH2 groups is suggested in β- and γ-KNH2 by investigating the structural details and charge density, which could be favorable for accelerating dehydrogenation behavior of complex metal hydrides.
1. Introduction
Great effort has been devoted to developing safe, efficient, and reversible hydrogen storage materials such as boronhydrides M(BH4)n,1,2 alanates M(AlH4)n,3,4 and amides M(NH2)n.5,6 These complex metal hydrides have been considered the promising materials for hydrogen storage due to their innately high hydrogen content. However, many fundamental scientific and technological challenges still remain, such as hydrogen desorption under near-ambient conditions.7–10 Due to strong metal–hydrogen bonding (B–H, Al–H, and N–H) in the anion subunits (i.e., BH4−, AlH4−, and NH2−), these metal hydrides are so stable that releasing hydrogen can only be achieved at elevated temperatures. Therefore, accelerating the dehydrogenation behavior of the light complex metal hydrides has great scientific and practical significance in this field. Doping with Ti in NaAlH4 has been considered as reversible with respect to hydrogen absorption/desorption.11 Otherwise, improvement of the hydrogen reversibility of NaAlH4 has been also achieved by ball milling experiments.12 Recently, Vajeeston et al.4 have performed a theoretical study of NaAlH4, which indicates high pressure could accelerate dehydrogenation kinetics accompanied with the structural changes. Thus, exploring high pressure structures is necessary for complex metal hydrides.
Alkali amides have become the scientific focus as hydrogen storage materials. LiNH2 has been considered a potential complex hydride proposed by Chen et al.7,13–16 The ambient structure of α-LiNH2 phase with I
symmetry initially transforms into a monoclinic β-LiNH2 phase with P21 symmetry confirmed by experimental works.17 Meanwhile, the structure of γ-LiNH2 phase with hydrogen bonding between the neighboring NH2 groups has been predicted by theoretical and experimental studies under high pressures.18 Recently, sodium amide (NaNH2) has also attracted considerable attention, although its theoretical hydrogen capacity is only ∼5.1 wt%.19 Under ambient conditions, α-NaNH2 crystallizes in an orthorhombic lattice with the space group Fddd.20 With the pressure increasing, α-NaNH2 transforms to β-NaNH2 (space group P21212) phase at 2.2 GPa and then to γ-NaNH2 (space group C2/c) phase at 9.4 GPa.21,22 As the same as γ-LiNH2 phase, existence of hydrogen bond has been conformed in the high pressure phases of NaNH2.22 As for potassium amide (KNH2), the ambient structure has a monoclinic phase (hereafter, α-KNH2, space group P21/m).23 However, the crystal structures of high pressure phases have not been observed. Besides, it is still an unknown question whether the high pressure phases of KNH2 have more favorable properties than ambient structure such as the existence of hydrogen bond. Hydrogen bond is expected to weaken the N–H polar covalent bonds in amides ions and can accelerate dehydrogenation behavior, which prompts the candidate promising as hydrogen storage materials for complex metal hydrides.
In the present work, we perform the systematic research for high-pressure structures of KNH2 by using particle swarm optimization technique implemented in the CALYPSO code and density functional theory (DFT). Here, we predict two new structures at different pressures that never reported, and discuss their structures, structural stability, electronic structures and chemical bonding.
2. Computational methods
The high-pressure structure searches for KNH2 were explored by using the particle swarm optimization technique implemented in the CALYPSO code.24,25 The simulation method has been successfully used to investigate the structures of a great variety of materials26–29 at high pressures. In our calculations, ab initio structure relaxations were performed using density functional theory30,31 within the Perdew–Burke–Ernzerhof (PBE) parameterization of the generalized gradient approximation (GGA),32 which is implemented in the Vienna ab initio simulation package (VASP).33 The projector augmented wave potentials34 were used to describe the ionic potentials with valence electrons of 3s23p64s1, 2s22p3 and 1s1 for K, N and H atoms, respectively. The cutoff energy (1000 eV) for the expansion of the wave function into plane waves and a Monkhorst–Pack kmesh35 of 0.03 × 2π Å−1 were chosen to ensure that all the enthalpy calculations are well converged to better than 1 meV per atom. The convergence tests have been described elsewhere.36–39 The elastic constants was calculated using density functional theory within the (PBE) parameterization of the (GGA)32 as implemented in the Vienna ab initio simulation package CASTEP code.40 Vanderbilt ultrasoft pseudopotentials41 were generated for K, N, and H with the valence configuration of 3s23p64s, 2s22p3 and 1s1, respectively. Convergence tests concluded that suitable values would be a 1000 eV kinetic energy cutoff and a Monkhorst–Pack (MP) k-point mesh of 0.03 × 2π Å−1 for the electronic Brillouin zone (BZ) integration. Phonon dispersion of the structures predicted was obtained using direct supercell method42 with the PHONOPY code.43
3. Results and discussion
CALYPSO structure predictions are performed for simulation cells containing up to 4 unit cells at pressures range from 0 to 20 GPa. Besides, the earlier predicted structures of LiNH2 and NaNH2 are considered in our structural searching. The ground-state α-KNH2 is successfully obtained at 0 GPa. Other five energetic competing structures are detected in our structure searches with different unit cells. They are, namely, orthorhombic Amm2, orthorhombic P21212, monoclinic P21(1) (β-LiNH2 structure), monoclinic P21(2) (2 f.u. per cell), and monoclinic Pc structures as depicted in Fig. 1. Except for monoclinic P21(1) (β-LiNH2 structure), NH2 amide group in all proposed structures are orientational order. In the Amm2 structure, the NH2 groups arrange in the same orientation and a layer of K+ ions is inserted between two NH2 layers. The ordered NH2 groups in different orientations appear in P21212, monoclinic P21(2), and Pc structures. Moreover, we can see that the orientations of two adjacent NH2 molecules are opposite and K+ ions are surrounded by NH2 molecules in the three structures.
 |
| Fig. 1 Crystal structures of competing KNH2 phases: (a) orthorhombic Amm2, (b) orthorhombic P21212, (c) monoclinic P21 β-LiNH2-type, (d) monoclinic P21, (e) monoclinic Pc. The pink, blue and black spheres represent K, N and H, respectively. | |
For exploring the stability of the structures, the free energies of competing phases must be investigated. In our work, all total-energy calculations are performed at zero temperature. Therefore, the Gibbs free energy becomes equal to the enthalpy, H = E0 + PV, where E0 is the internal energy of the system. Fig. 2 displays the calculated enthalpy difference (ΔH) relative to α-KNH2 as a function of pressure. Below 4 GPa, α-KNH2 is the most stable structure. With the pressure increasing, the first phase transition from α-KNH2 to P21(2) (hereafter, β-KNH2) occurs at 4 GPa. Above 6.8 GPa, the Pc phase (hereafter, γ-KNH2) becomes the most stable until 20 GPa. Our results show that two phase transformations occurs at 4 GPa and 6.8 GPa, respectively, in which the zero point energy (ZPE) has not been included. It is known that the effect of ZPE plays a non-negligible role in the total energy of the hydrogen rich materials, but the relative ordering of the structures is not influenced by the ZPE in our calculations. As shown above, we can see that a decrease in symmetry of KNH2 with pressure, which is very similar to the high-pressure behavior of the related compounds LiNH2 (I
→ Fddd → P21212) and NaNH2 (Fddd → P21212 → C2/c). The pressure–volume relation of KNH2 is derived by fitting the third-order Birth–Murnaghan equation of state44 as shown in Fig. 3. The α → β and β → γ phase transformations for KNH2 are accompanied by large volume shrinkages of 5.0% and 3.8%, respectively. Volume discontinuous under high pressures indicates the two phase transformations for KNH2 are both the first-order phase transition.
 |
| Fig. 2 Calculated enthalpies relative to α-KNH2 as a function of pressure for various structures. | |
 |
| Fig. 3 Calculated volume–pressure relationship for α-KNH2, β-KNH2 and γ-KNH2. | |
It is importance of calculating the phonon spectra of a crystal for further confirming the dynamical stability of the structures predicted. Fig. 4 lists the phonon dispersion curves for β-KNH2 and γ-KNH2 at 5 GPa and 8 GPa, respectively. No imaginary phonon frequencies are observed, which indicates the β-KNH2 and γ-KNH2 is dynamically stable. The calculations of the elastic constants are essential, which can help to provide valuable information for the mechanical stability of a structure. The well-known mechanical stability criteria45 for monoclinic system is:
C11 > 0, C22 > 0, C33 > 0, C44 > 0, C55 > 0, C66 > 0, |
[C11 + C22 + C33 + 2(C12 + C13 + C23)] > 0, |
(C33C55 − C352) > 0, (C44C66 − C462) > 0, (C22 + C33 − 2C23) > 0 |
 |
| Fig. 4 Calculated phonon dispersion curves for high-pressure phases: (a) β-KNH2 at 5 GPa, (b) γ-KNH2 at 8 GPa. | |
The independent elastic stiffness constants of the β-KNH2 and γ-KNH2 are shown in Table 1, which indicates the mechanical stability of the two phases.
Table 1 Elastic constants Cij (GPa) for β-KNH2 at 5 GPa and γ-KNH2 at 8 GPa
|
C11 |
C22 |
C33 |
C44 |
C55 |
C66 |
C12 |
C13 |
C15 |
C23 |
C25 |
C35 |
C46 |
β-KNH2 |
48.1 |
59.7 |
53.9 |
16.9 |
16.4 |
23.1 |
32.8 |
25.3 |
6.3 |
30.8 |
−0.9 |
−3.7 |
1.3 |
γ-KNH2 |
93.6 |
94.2 |
93.0 |
20.1 |
21.8 |
35.2 |
39.3 |
24.7 |
−0.1 |
23.1 |
−0.1 |
0.5 |
−3.6 |
Table 2 summarize the optimized equilibrium crystal structural details for α-(0 GPa), β-(5 GPa) and γ-KNH2 (8 GPa). The characteristic geometries of amide ions in α-(0 GPa), β-(5 GPa) and γ-KNH2 (8 GPa) are shown in Fig. 5. In α-KNH2, each N–H bond length is 1.032 Å and H–N–H bond (∠H–N–H) angle is 143.1° in NH2 molecular. Two adjacent NH2 molecules are parallel to each other and the intermolecular N⋯H distance is 3.112 Å, which is larger than the sum of the van der Waals radii of N and H (2.75 Å).46 Therefore, hydrogen bond is difficult to be observed in α-KNH2. Fig. 5(b) shows that each N–H bond length is 1.028 Å and ∠H–N–H angle is 101.8° in NH2 molecular. The neighboring NH2 groups are perpendicular to each other, and each N–H bond is connected almost linearly with the N atom in its neighboring NH2 group. The calculated N–H⋯N bond (∠H–N⋯H) angle and length are 171.6° and 2.286 Å at 5 GPa, respectively, which can easily form the hydrogen bond. The neighboring NH2 groups for γ-KNH2 are shown in Fig. 5(c). In each NH2 group, each N–H bond length and ∠H–N–H angle can reach 1.035 Å and 99.8° at 8 GPa, respectively. The four neighboring NH2 groups lie in two different planes and the NH2 groups in each plane are parallel to each other. Each N–H bond is aligned almost linearly with the N atom in its neighboring NH2 group. The obtained ∠H–N⋯H angle and N–H⋯N bond length are 169.9° and 2.162 Å at 8 GPa in each plane. Besides, the ∠H–N⋯H angle and N–H⋯N bond length between the neighboring planes can reach 174.8° and 2.266 Å. As a result, the formation of hydrogen bond can be observed in β-KNH2 and γ-KNH2, which can also be understood from our latter analysis of electronic structure.
Table 2 Predicted lattice constants and atomic coordinates for α-KNH2, β-KNH2, and γ-KNH2 at 8 GPa at the selected pressures
Pressure (GPa) |
Space group |
Lattice parameter (Å) |
Atomic coordinates (fractional) |
0 |
P21/m |
a = 4.635, b = 3.835 |
K 2e (0.2189, 0.25, 0.3121) |
c = 6.334, β = 96.3° |
N 2e (0.2804, 0.25, 0.7575) |
a = 2.870, c = 2.409 |
H 4f (0.3089, 0.0435, 0.8622) |
5 |
P21 |
a = 3.698, b = 4.622 |
K 2a (0.0609, 0.1460, 0.2523) |
c = 5.112, β = 82.4° |
N 2a (0.6208, 0.6561, 0.2683) |
H1 2a (0.5931, 0.1673, 0.8407) |
H2 2a (0.5224, 0.5117, 0.4149) |
8 |
Pc |
a = 5.976, b = 4.982 |
K 2a (0.7153, 0.2031, 0.8437) |
c = 4.915, β = 147.9° |
N 2a (0.2409, 0.7207, 0.3731) |
H1 2a (0.9213, 0.2829, 0.5781) |
H2 2a (0.2610, 0.5831, 0.2366) |
 |
| Fig. 5 Geometries of amide ions in (a) α-KNH2 at 0 GPa, (b) β-KNH2 at 5 GPa, and (c) γ-KNH2 at 8 GPa. | |
To further understand the nature of chemical bonding in KNH2, we calculate the partial DOS (PDOS) of α, β, and γ phases at different pressures as depicted in Fig. 6. The three phases have the insulator character due to the finite energy band gap (Eg) between the valence band (VB) and conduction band (CB). The calculated Eg can reach 2.1 eV (α-KNH2) at 0 GPa, 3.7 eV (β-KNH2) at 5 GPa, and 4.1 eV (γ-KNH2) at 8 GPa, respectively. The results show that N 2p and H 1s contribute mostly to the VBs. However, the contributions to the VBs from K are very little for the three phases. Therefore, we conclude that the interaction between K+ cations and [NH2]− anions is ionic. Otherwise, energetically degenerate between N 2p and H 1s in the VB region can be observed, which indicates the hybridization between N and H atoms is very remarkable and there exists strong covalent bond between N and H atoms in the [NH2]− anions. Fig. 6 shows that VBs for the three phases are split into three separate regions in the range from −8 to 0 eV. With increasing pressure, from α- to β-, to γ-KNH2 there is a little difference in peak profiles and the PDOS peaks for β-KNH2 and γ-KNH2 become broader, which indicates that electronic delocalization in KNH2 increases under the influence of pressure.
 |
| Fig. 6 Calculated total and partial density of states for α-(at 0 GPa), β-(at 5 GPa), and γ-KNH2 (at 8 GPa). Fermi levels are set to zero energy and marked by dotted vertical lines; s, p and d states are red, green, and blue lines, respectively. | |
In complex metal hydrides, hydrogen bond is advantageous for accelerating dehydrogenation behavior. For further probing the existence of hydrogen bond, we calculate the charge density for the three phases. The calculated charge densities for some selected planes across the NH2 group for α-(0 GPa), β-(5 GPa), and γ-KNH2 (8 GPa) are displayed in Fig. 7(a–c). We can see that there is a strong covalent bonding between N and H atoms in NH2 groups. Additionally, hydrogen bond between the neighboring NH2 groups can be confirmed by investigating the charge density (ρ) at the bond critical point (bcp).47 Scheiner et al. have proposed that existence of hydrogen bond can be achieved if the electron density (ρbcp) at the bond critical point is in the range from 0.01 to 0.03 e Å−3. In α-KNH2, the ρbcp in the N–H⋯N bond is 0.001 e Å−3 as depicted in Fig. 7(a), which indicates there is no hydrogen bond in this phase. Fig. 7(b) shows that the ρbcp is 0.017 e Å−3 in β-KNH2, meanwhile we can see that the ρbcp can reach 0.017 e Å−3 in the same plane and 0.018 e Å−3 in the N–H⋯N bond between the neighboring NH2 groups for γ-KNH2 as shown in Fig. 7(b and c), respectively, which indicates the hydrogen bond can be formed in high pressure phases for KNH2. Our results are similar to those of NaNH2 (ref. 22) under high pressures.
 |
| Fig. 7 Charge density distribution of KNH2 within the scale of 0 to 0.01 e Å−3: (a) the slice across the NH2 group for α-KNH2 at 0 GPa, (b) the slice across the neighboring NH2 group for β-KNH2 at 5 GPa, and (c) the slice across the neighboring NH2 groups for γ-KNH2 at 8 GPa. The hydrogen bond paths are denoted by the dotted lines. | |
4. Conclusions
We have performed the predictions of new high-pressure structures of KNH2 by using particle swarm optimization technique implemented in the CALYPSO code and density functional theory (DFT). Five energetically competing structures have been selected as the candidates in our structural predictions. The calculations of enthalpy difference relative to α-KNH2 have suggested two pressure-induced structural phase transformations occur with increasing pressure. The ground-state α-KNH2 transforms into a monoclinic β-KNH2 with space group P21 at 4 GPa and then into a monoclinic γ-KNH2 with space group Pc at 6.8 GPa. Moreover, these phase transformations are companied with volume reductions of 5.0% and 3.8%, respectively, which reveals the nature of the first-order structural phase transition. The predicted structural sequence (P21/m → P21 → Pc) of KNH2 is very similar to the high-pressure behavior of the related compounds LiNH2 (I
→ Fddd → P21212) and NaNH2 (Fddd → P21212 → C2/c), which further verify that the trend of lowering symmetry induced by high pressure may be common in alkali metal amides. The results of phonon dispersion curves and elastic constants indicate the new predicted high-pressure structures are stable dynamically and mechanically. By analyzing partial density of states and charge density, we can see that there exists a strong covalent bonding between N and H atoms in NH2 groups in the three structures. Meanwhile, N–H⋯N hydrogen bonding between neighboring NH2 groups have been proposed in β- and γ-KNH2 by investigating the structural details and charge density. The existence of hydrogen bonding can weak the covalent bonding between N and H atoms in NH2 groups, which could be favorable for accelerating dehydrogenation behavior of complex metal hydrides.
Acknowledgements
This work was supported by Open Project of State Key Laboratory of Superhard Materials (Jilin University), the National Natural Science Foundation of China (No. 11104019, 2011CB808200), Program for Changjiang Scholars and Innovative Research Team in University (No. IRT1132), and National Found for Fostering Talents of basic Science (No. J1103202). Parts of calculations were performed in the Scientific Computation and Numerical Simulation Center of Changchun University of Science and Technology.
References
- R. L. Davis and C. H. Kennard, J. Solid State Chem., 1985, 59, 393–396 CrossRef CAS.
- E. Hazrati, G. Brocks and G. A. de Wijs, J. Phys. Chem. C, 2012, 116, 18038–18047 CAS.
- B. Hauback, H. Brinks and H. Fjellvåg, J. Alloys Compd., 2002, 346, 184–189 CrossRef CAS.
- P. Vajeeston, P. Ravindran, R. Vidya, H. Fjellvag and A. Kjekshus, Appl. Phys. Lett., 2003, 82, 2257–2259 CrossRef CAS.
- S. Orimo, Y. Nakamori, G. Kitahara, K. Miwa, N. Ohba, T. Noritake and S. Towata, Appl. Phys. A: Mater. Sci. Process., 2004, 79, 1765–1767 CrossRef CAS.
- H. Zhu, F. Zhang, C. Ji, D. Hou, J. Wu, T. Hannon and Y. Ma, J. Appl. Phys., 2013, 113, 033511 CrossRef.
- P. Chen, Z. Xiong, J. Luo, J. Lin and K. L. Tan, Nature, 2002, 420, 302–304 CrossRef CAS PubMed.
- C. Jensen and K. Gross, Appl. Phys. A: Mater. Sci. Process., 2001, 72, 213–219 CrossRef CAS.
- A. Züttel, S. Rentsch, P. Fischer, P. Wenger, P. Sudan, P. Mauron and C. Emmenegger, J. Alloys Compd., 2003, 356, 515–520 CrossRef.
- A. Züttel, P. Wenger, S. Rentsch, P. Sudan, P. Mauron and C. Emmenegger, J. Power Sources, 2003, 118, 1–7 CrossRef.
- T. Kiyobayashi, S. S. Srinivasan, D. Sun and C. M. Jensen, J. Phys. Chem. A, 2003, 107, 7671–7674 CrossRef CAS.
- A. Leon, O. Kircher, J. Rothe and M. Fichtner, J. Phys. Chem. B, 2004, 108, 16372–16376 CrossRef CAS.
- Y. Song and Z. Guo, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 74, 195120 CrossRef.
- J. Yang, X. Zhou, Q. Cai, W. J. James and W. B. Yelon, Appl. Phys. Lett., 2006, 88, 041914 CrossRef.
- T. Tsumuraya, T. Shishidou and T. Oguchi, J. Phys.: Condens. Matter, 2009, 21, 185501 CrossRef PubMed.
- C. Zhang and A. Alavi, J. Phys. Chem. B, 2006, 110, 7139–7143 CrossRef CAS PubMed.
- R. S. Chellappa, D. Chandra, M. Somayazulu, S. A. Gramsch and R. J. Hemley, J. Phys. Chem. B, 2007, 111, 10785–10789 CrossRef CAS PubMed.
- X. Huang, D. Li, F. Li, X. Jin, S. Jiang, W. Li, X. Yang, Q. Zhou, B. Zou and Q. Cui, J. Phys. Chem. C, 2012, 116, 9744–9749 CAS.
- T. Ichikawa and S. Isobe, Z. Kristallogr., 2008, 223, 660–665 CrossRef CAS.
- M. Nagib, H. Kistrup and H. Jacobs, Atomkernenergie, 1975, 26, 87–90 CAS.
- A. Liu and Y. Song, J. Phys. Chem. B, 2010, 115, 7–13 CrossRef PubMed.
- Y. Zhong, H. Y. Zhou, C. H. Hu, D. H. Wang and A. R. Oganov, J. Phys. Chem. C, 2012, 116, 8387–8393 CAS.
- R. Juza, H. Jacobs and W. Klose, Z. Anorg. Allg. Chem., 1965, 338, 171–178 CrossRef CAS.
- Y. Wang, J. Lv, L. Zhu and Y. Ma, Phys. Rev. B: Condens. Matter Mater. Phys., 2010, 82, 094116–094123 CrossRef.
- Y. Wang, J. Lv, L. Zhu and Y. Ma, Comput. Phys. Commun., 2012, 183, 2063–2070 CrossRef CAS.
- Y. Li, Y. Wang, C. J. Pickard, R. J. Needs, Y. Wang and Y. Ma, Phys. Rev. Lett., 2015, 114, 125501 CrossRef PubMed.
- Y. Wang, H. Liu, J. Lv, L. Zhu, H. Wang and Y. Ma, Nat. Commun., 2011, 2, 563 CrossRef PubMed.
- M. Zhang, H. Liu, Q. Li, B. Gao, Y. Wang, H. Li, C. Chen and Y. Ma, Phys. Rev. Lett., 2015, 114, 015502 CrossRef PubMed.
- L. Zhu, Z. Wang, Y. Wang, G. Zou, H. K. Mao and Y. Ma, Proc. Natl. Acad. Sci. U. S. A., 2012, 109, 751–753 CrossRef CAS PubMed.
- S. Baroni, P. Giannozzi and A. Testa, Phys. Rev. Lett., 1987, 58, 1861–1864 CrossRef CAS PubMed.
- P. Giannozzi, S. De Gironcoli, P. Pavone and S. Baroni, Phys. Rev. B: Condens. Matter Mater. Phys., 1991, 43, 7231 CrossRef CAS.
- J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef CAS PubMed.
- G. Kresse and J. Furthmüller, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 11169 CrossRef CAS.
- P. E. Blöchl, Phys. Rev. B: Condens. Matter Mater. Phys., 1994, 50, 17953 CrossRef.
- H. J. Monkhorst and J. D. Pack, Phys. Rev. B: Solid State, 1976, 13, 5188–5192 CrossRef.
- D. Li, K. Bao, F. Tian, Z. Zeng, Z. He, B. Liu and T. Cui, Phys. Chem. Chem. Phys., 2012, 14, 4347–4350 RSC.
- D. Li, F. Tian, D. Duan, K. Bao, B. Chu, X. Sha, B. Liu and T. Cui, RSC Adv., 2014, 4, 10133–10139 RSC.
- C. Chen, Y. Xu, X. Sun, S. Wang and F. Tian, RSC Adv., 2014, 4, 55023–55027 RSC.
- F. Tian, D. Duan, D. Li, C. Chen, X. Sha, Z. Zhao, B. Liu and T. Cui, Sci. Rep., 2014, 4, 5759 Search PubMed.
- M. Segall, P. J. Lindan, M. Probert, C. Pickard, P. Hasnip, S. Clark and M. Payne, J. Phys.: Condens. Matter, 2002, 14, 2717 CrossRef CAS.
- D. Vanderbilt, Phys. Rev. B: Condens. Matter Mater. Phys., 1990, 41, 7892 CrossRef.
- K. Parlinski, Z. Li and Y. Kawazoe, Phys. Rev. Lett., 1997, 78, 4063–4066 CrossRef CAS.
- A. Togo, F. Oba and I. Tanaka, Phys. Rev. B: Condens. Matter Mater. Phys., 2008, 78, 134106 CrossRef.
- W. Kohn and L. J. Sham, Phys. Rev., 1965, 140, A1133 CrossRef.
- Z. J. Wu, E. J. Zhao, H. P. Xiang, X. F. Hao, X. J. Liu and J. Meng, Phys. Rev. B: Condens. Matter Mater. Phys., 2007, 76, 054115 CrossRef.
- H. Lutz, J. Mol. Struct., 2003, 646, 227–236 CrossRef CAS.
- S. Scheiner, Hydrogen bonding: a theoretical perspective, Oxford University Press, Oxford, U.K., 1997 Search PubMed.
|
This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.