Next Article in Journal
Lithium-Ion Battery State-of-Charge Estimation Using Electrochemical Model with Sensitive Parameters Adjustment
Next Article in Special Issue
C4S Nanosheet: A Potential Anode Material for Potassium-Ion Batteries
Previous Article in Journal
Textile PAN Carbon Fibers Cathode for High-Voltage Seawater Batteries
Previous Article in Special Issue
Battery-Type Lithium-Ion Hybrid Capacitors: Current Status and Future Perspectives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Dehydrogenation of Alkali Metal Aluminum Hydrides MAlH4 (M = Li, Na, K, and Cs): Insight from First-Principles Calculations

1
School of Physical Science & Technology, State Key Laboratory of Featured Metal Materials and Life-Cycle Safety for Composite Structures, Guangxi University, Nanning 530004, China
2
School of Mathematics and Physics, Key Laboratory for Ionospheric Observation and Simulation, Guangxi University for Nationalities, Nanning 530006, China
*
Author to whom correspondence should be addressed.
Batteries 2023, 9(3), 179; https://doi.org/10.3390/batteries9030179
Submission received: 14 January 2023 / Revised: 28 February 2023 / Accepted: 15 March 2023 / Published: 19 March 2023
(This article belongs to the Special Issue Advances in Carbon-Based Materials for Energy Storage)

Abstract

:
Complex aluminum hydrides with high hydrogen capacity are among the most promising solid-state hydrogen storage materials. The present study determines the thermal stability, hydrogen dissociation energy, and electronic structures of alkali metal aluminum hydrides, MAlH4 (M = Li, Na, K, and Cs), using first-principles density functional theory calculations in an attempt to gain insight into the dehydrogenation mechanism of these hydrides. The results show that the hydrogen dissociation energy (Ed-H2) of MAlH4 (M = Li, Na, K, and Cs) correlates with the Pauling electronegativity of cation M (χP); that is, the Ed-H2 (average value) decreases, i.e., 1.211 eV (LiAlH4) < 1.281 eV (NaAlH4) < 1.291 eV (KAlH4) < 1.361 eV (CsAlH4), with the increasing χP value, i.e., 0.98 (Li) > 0.93 (Na) > 0.82 (K) > 0.79 (Cs). The main reason for this finding is that alkali alanate MAlH4 at higher cation electronegativity is thermally less stable and held by weaker Al-H covalent and H-H ionic interactions. Our work contributes to the design of alkali metal aluminum hydrides with a favorable dehydrogenation, which is useful for on-board hydrogen storage.

1. Introduction

The rapidly diminishing supply of fossil fuel and increasing environmental awareness have precipitated a growing demand for clean, safe, and renewable energy. Hydrogen is believed to be an ideal clean energy carrier due to its abundance, high energy density (142 MJ/kg), and clean combustion [1]. However, hydrogen is a highly flammable, explosive, and diffusible gas at room temperature and pressure. Thus, it is very important to store hydrogen safely and effectively, yet this task remains a major challenge in hydrogen utilization [2]. Currently, hydrogen storage approaches involve the storage of: (1) compressed hydrogen in high-pressure containers, (2) liquid hydrogen in cryogenic tanks, and (3) hydrogen in solid-state materials via physisorption/chemisorption. Among these approaches, solid-state hydrogen storage provides high hydrogen capacity, moderate operating pressures and temperatures, and favorable safety, and is a promising storage solution [3,4,5].
Complex metal hydrides, such as alanates, borohydrides, and amides, are considered good solid-state hydrogen storage candidates for on-board applications due to the high gravimetric and volumetric hydrogen densities required in this regard [3,4]. These complex hydrides, unfortunately, are often thermodynamically very stable and dehydrogenate at extremely high temperatures, thereby restricting their practical applications. To overcome these drawbacks, many investigations have been devoted to tuning the thermodynamic properties of complex metal hydrides [6,7,8,9,10].
Nakamori et al. [11] systematically investigated the thermodynamical stabilities of a series of metal borohydrides, M(BH4)n (M = Li, Na, K, Cu, Mg, Zn, Sc, Zr, and Hf; n = 1–4), via first-principles calculations combined with an experimental study. They reported that the stability of M(BH4)n is related to the ionic interaction between M and the [BH4] complex, the charge transfer from the cation Mn+ to the anion [BH4] (Mn+→[BH4]), and the Pauling electronegativity of the cation M. In particular, a gradual increase in cation Pauling electronegativity χP is accompanied by a linear decrease in the dehydrogenation temperature (Td). Their works on other borohydrides, M(BH4)n (M = Ca, Sc, Ti, V, Cr, Mn, Zn, and Al; n = 2–4), also revealed a linear relation between χP and Td [12]. These works by Nakamori and co-workers contribute to the design of metal borohydrides with appropriate stability for favorable dehydrogenation through the combination of M(BH4)n with more electronegative metals or metal compounds, which is useful for hydrogen storage applications [6,13,14,15,16,17,18].
Recently, in the review by Weidenthaler, the decomposition temperature (Td) of alkali metal aluminum hydrides (alkali alanates) of the MAlH4 variety (M = Li, Na, K, and Cs) is reported to decrease linearly with the Pauling electronegativity, χP, of alkali cations [19]. As an example from this review, the decrease in the starting decomposition temperature (the first step of dehydrogenation reaction), specifically, from 443 K (LiAlH4) [20] < 503 K (NaAlH4) [21] < 573 K (KAlH4) [20] < 600 K (CsAlH4) [19], is accompanied by an increase in cation Pauling electronegativity, specifically, from 0.98 (Li) > 0.93 (Na) > 0.82 (K) > 0.79 (Cs). From these results, the electronegativity of alkali metal cations was identified as dominant with respect to the thermal stability and the decomposition temperature of MAlH4 (M = Li, Na, K, and Cs) [19]. However, the dehydrogenation mechanism of MAlH4 associated with cation electronegativity was not provided in the review. Furthermore, to date, there have only been a few reports on this topic. Therefore, in this study, we apply first-principles density functional theory calculations on MAlH4 (M = Li, Na, K, and Cs) since the calculations can reliably be used to study the micro-mechanisms of hydrogen storage materials [11,22,23]. The formation enthalpy, cohesive energy, Hirshfeld charge, hydrogen dissociation energy, density of states, charge density distribution, and Mulliken population of MAlH4 (M = Li, Na, K, and Cs) are investigated in detail. We believe that our work can provide new insights into the dehydrogenation of MAlH4 (M = Li, Na, K, and Cs) and is, therefore, helpful for exploring solid-state alkali aluminum hydrides with favorable H-desorption properties for hydrogen storage applications.

2. Computational Details

The present calculations on the alkali alanates of MAlH4 (M = Li, Na, K, and Cs) were executed using the density functional theory (DFT) method with on-the-fly generation (OTFG) ultrasoft pseudopotential, as implemented in the Cambridge Serial Total Energy Package (CASTEP) code in Materials Studio 2017 [24]. The exchange-correlation function was used with the generalized gradient approximation (GGA)of Perdew–Burke–Ernzerhof (PBE) [25]. The dependence of total energy on the plane–wave cutoff energy and Monkhorst–Pack k-point mesh were tested carefully. Subsequently, a plane–wave cutoff energy of 900 eV and a k-point grid of 3 × 2 × 2 were adopted, thereby ensuring a convergence accuracy with a total energy difference below 3 meV/atom. Atomic valence electrons, namely, 1s22s1 (Li), 2s22p63s1 (Na), 3s23p64s1 (K), 5s25p66s1 (Cs), 3s23p1 (Al), and 1s1 (H), were used for corn electrons. Geometry optimization using the Broyden–Fletcher–Goldfarb–Shanno (BFGS) method [26] allowed the lattice and all atoms to relax, with the convergence tolerance of 1.0 × 10−5 eV/atom, 0.03 eV/Å, 0.05 GPa, and 0.001 Å for energy, maximum force, maximum stress, and maximum displacement, respectively. The single-point energies and electronic structures of all the considered systems were calculated following geometric optimization.
The alkali metal aluminum hydrides of MAlH4 (M = Li, Na, K, and Cs) considered in our studies have space groups of P21/C (monoclinic structure, LiAlH4) [27], I41/A (tetragonal structure, NaAlH4) [28], and Pnma (orthorhombic structure, KAlH4 and CsAlH4) [29,30], as shown in Table 1. The geometric optimization of these aluminum hydrides gave the relaxed lattice parameters and volume (Table 1), whose level of agreement with the experimental data [27,28,29,30] is fairly high. Using relaxed MAlH4, a 2 × 1 × 1 supercell consisting of 8 MAlH4 units (M8Al8H32) was established for calculations, as illustrated in Figure 1. In this figure, two [AlH4] units in MAlH4 bulk are considered for hydrogen desorption, with their H atoms labeled as HA, HB, HC, HD, HE, HF, and HG, and Al-H and H-H distances within 1.625–1.643 Ǻ (Al-H) and 2.639–2.754 Ǻ (H-H) in each [AlH4] unit. From these Al-H and H-H distances, the H bonding (Al-H and H-H bonding) in the [AlH4] group can be predicted [31]. The previous experimental and theoretical studies on aluminum hydrides also provide support for the formation of Al-H and H-H bonds in the [AlH4] unit [31,32,33].
The formation enthalpy (ΔH) and cohesive energy (Ecoh) of MAlH4 aluminum hydrides (LiAlH4, NaAlH4, KAlH4, and CsAlH4) are calculated using the following formulae (Equations (1) and (2)), wherein the zero-point energy (ZPE) correction has been considered [34]:
Δ H = E MAlH 4 E M E Al 2 E H 2
E coh = E MAlH 4 ε M ε Al 4 ε H
where E(MAlH4) denotes the total energy of MAlH4 (M = Li, Na, K, and Cs); E(M/Al) represents the energy of M/Al atom in each crystal structure of bcc-Li/Na/K/Cs and fcc-Al; ε(M/Al/H) denotes the isolated M/Al/H atom’s energy. E(H2), the energy of hydrogen molecular, is estimated to be −31.407 eV by placing two H atoms 0.741 Ǻ apart [35] in a 10 × 10 × 10 Ǻ (1000 Ǻ3) cubic box. The result is in good agreement with those of −31.292 eV [36] and −31.592 eV [37] from the literature.
The hydrogen dissociation energy is defined as the energy cost of removing hydrogen molecules (hydrogen pair) from the mother bulk, since the initial decomposition of MAlH4 (M = Li, Na, K, and Cs) takes place via the reaction of 3MAlH4→M3AlH6 + 2Al + 3H2 with the release of hydrogen gas [19,38,39]. In this case, the hydrogen pairs used for H-dissociation include (HA and HB), (HA and HC), and (HA and HD) from one [AlH4] unit, and (HA and HE), (HA and HF), (HA and HG), and (HA and HH) from two [AlH4] units (Figure 1). In addition, the two hydrogen atoms, HA and HX (HB, HC, HD, HE, HF, HG or HH), in these hydrogen pairs can be removed in two possible ways: (1) HA and HX are taken away one by one (asynchronous hydrogen desorption, expressed as HA→HX). That is, the atom HA is first removed from the M8Al8H32 bulk forming M8Al8H31 with a HA vacancy, and then atom HX is removed from relaxed M8Al8H31 bulk forming M8Al8H30 with HA and HX vacancies. In this case, the hydrogen dissociation energy is calculated by Equation (3) (the first step for HA removal, Ed-HA), Equation (4) (the second step for HX removal, Ed-HX), and Equation (5) (the total energy for HA→HX removal, Ed-H2); (2) HA and HX are taken away simultaneously (synchronous hydrogen desorption, denoted as HA–HX), with the hydrogen dissociation energy (Ed-H2) determined by the following Equation (6) [40]
E d H A = [ E M 8 Al 8 H 31 + 1 2 E H 2 ] E M 8 Al 8 H 32
E d H X = [ E M 8 Al 8 H 30 + 1 2 E H 2 ] E M 8 Al 8 H 31
E d H 2 = E d H A + E d H X
E d H 2 = E M 8 Al 8 H 30 + E H 2 E M 8 Al 8 H 32
in which E(H2) is the same as the previous definition; E(M8Al8H32), E(M8Al8H31), and E(M8Al8H30) are the total energy of corresponding systems.

3. Results and Discussion

3.1. Thermal Stability

The formation enthalpy, ΔH, refers to the formation heat in a hydriding reaction and is helpful for evaluating the thermal stability of metal hydrides [41,42]. A negative formation enthalpy (ΔH < 0) suggests an exothermic reaction. Furthermore, a compound is more thermally stable if it has a more negative formation enthalpy [41,42,43]. Table 1 lists the formation enthalpy of MAlH4 (M = Li, Na, K, and Cs) with ZPE correction. It has been found that the formation enthalpies, ΔH, of MAlH4 (M = Li, Na, K, and Cs) are always negative, and the values −117.156 kJ/mol (LiAlH4), −120.468 kJ/mol (NaAlH4), −157.2 kJ/mol (KAlH4) and −171.9 kJ/mol (CsAlH4) are in reasonable agreement with the available literature findings, that is, −113.42 kJ/mol (LiAlH4) [44], −155.5 kJ/mol (NaAlH4) [45], and −183.7 kJ/mol (KAlH4) [45]. In particular, these ΔH values become more negative when the Pauling electronegativity of cation M (χP, Table 1) decreases, with a linear relation ΔH = 296.033χP − 402.19 obtained via least square fitting in Figure 2. A similar linear correlation, ΔH = 248.7χP − 390.8, is also presented in borohydrides [11]. It is clear from the results above that MAlH4 (M = Li, Na, K, and Cs) aluminum hydrides may have reduced negative formation enthalpy as their alkali cation M has greater electronegativity, such as ΔH = −117.156 kJ/mol and χP = 0.98 for LiAlH4 vs. ΔH = −171.9 kJ/mol and χP = 0.79 for CsAlH4. The cohesive energy Eoch (Table 1), interestingly, has the same characteristics as the formation enthalpy ΔH, for which a linear relationship between Eoch and χP, Ecoh = 296.489χP − 1968.247, is also achieved. These ΔH and Eoch results suggest that the thermal stability corresponds to a descending order, wherein LiAlH4 < NaAlH4 < KAlH4 < CsAlH4 [41,42,43,46], followed by the decomposition temperature [11,19].
The charge transfer of a metal cation is related to its electronegativity, and it can help assess the thermal stability of metal hydrides. As mentioned in [11,47,48], the Mn+→[BH4] charge transfer is an important characteristic of the stability of M(BH4)n borohydrides. The suppression of the charge transfer by the substitution of a metal cation with a more electronegative element is expected to tailor the stability, thereby effectively lowering the dehydrogenation temperature of borohydrides. In this regard, it is interesting to find that the M+→[AlH4]- charge (Hirshfeld charge) transfer increases from 0.21 e (LiAlH4) to 0.28 e (NaAlH4), 0.35 e (KAlH4), and 0.42 e (CsAlH4), with the cation electronegativity (χP) decreasing from 0.98 (Li) to 0.93 (Na), 0.82 (K), and 0.79 (Cs), as shown in Table 1. The results indicate that the increase in cation electronegativity helps to suppress the charge transfer of the alkali cation; as a consequence, the thermal stability of MAlH4 (M = Li, Na, K, and Cs) is reduced [11,47,48]. This, combined with the analysis of the formation enthalpies ΔH and cohesive energies Ecoh (as described above), leads us to the conclusion that alkali alanates of MAlH4 (M = Li, Na, K, and Cs) containing higher χP are expected to use less energy for hydrogen desorption in the following hydrogen dissociation energy calculation.

3.2. Hydrogen Dissociation Energy

Dehydrogenation ability can be characterized theoretically by the hydrogen dissociation energy at which one or more hydrogen atoms are removed from a mother bulk. Table 2 lists the atomic hydrogen dissociation energies Ed-H for HA (Equation (3)) and HX (HB, HC, HD, HE, HF, HG, and HH) removal (Equation (4)), as well as the molecular hydrogen dissociation energies Ed-H2 for HA→HX (Equation (5)) and HA–HX removal (Equation (6)), and their minimum and average values for MAlH4 (M = Li, Na, K, and Cs). As can be seen in Table 2, for atomic hydrogen HA or HX desorption (an asynchronous HA→HX process with HA releasing first followed by the release of HX), the energy cost for HX desorption is much lower than that for HA desorption, such as in LiAlH4 −0.080 eV (Ed-HB) << 1.815 eV (Ed-HA). This indicates that it is easier for alanates of MAlH4 (M = Li, Na, K, and Cs) with an HA vacancy to release another hydrogen HX and even induce spontaneous HX dissociation as the hydrogen dissociation energies Ed-HX are negative (Table 2) [49,50,51]. Shi et al. [52] had proposed that the hydrogen diffusion of sodium aluminum hydrides is mediated by hydrogen vacancies. Here, the favorable HX desorption benefits from the HA vacancy in the MAlH4 bulk. For molecular hydrogen HA and HX desorption, on the one hand, asynchronous HA→HX desorption delivers a hydrogen dissociation energy very close to that for the corresponding synchronous HA–HX desorption, as noted in the example of LiAlH4 with Ed-H2 = 1.735 eV (HA→HB) and 1.733 eV (HA–HB). On the other hand, both asynchronous HA→HX and synchronous HA–HX desorption achieve lower hydrogen dissociation energy as HA and HX from two [AlH4] units compared to HA and HX from one [AlH4] unit, e.g., in LiAlH4 Ed-H2 = 0.953 eV (HA→HE) vs. 1.735 eV (HA→HB). In general, the formation of a stable [AlH3] unit/cluster followed by hydrogen release plays an important role in lowering the hydrogen removal energy in sodium alanate [53,54]. In the present work on alkali alanates of MAlH4, [AlH3] units are found to form upon HA→HX and HA–HX desorption. Moreover, for a given system, the formed [AlH3] units are more stable with HA and HX from two [AlH4] units because, in this case, HA and HX desorption has lower hydrogen dissociation energy (Ed-H2) (as described above); thus, the corresponding HA and HX desorbed system has lower total energy (E) according to Equations (3)–(6). This can be observed in the examples shown in Figure 3, where HA and HX desorption from one/two [AlH4] units has relative low hydrogen dissociation energy among all corresponding types of HA→HX and HA–HX desorption, including HA–HC and HA–HE desorption for LiAlH4, HA–HC and HA→HF desorption for NaAlH4, HA–HC and HA→HG desorption for KAlH4, and HA–HD and HA–HG desorption for CsAlH4 (Table 2). Obviously, our findings support the notion that the formation of stable [AlH3] units upon H-desorption is responsible for the reduction in the hydrogen dissociation energy in the alkali alanates of MAlH4.
In Table 2, it is worth noting that the hydrogen dissociation energies Ed-H2 of the considered hydrides of MAlH4 (M = Li, Na, K, and Cs), with respect to the minimum and average values, decrease in the order of LiAlH4 < NaAlH4 < KAlH4 < CsAlH4, i.e., 0.832 eV (LiAlH4) < 0.972 eV (NaAlH4) < 0.979 eV (KAlH4) < 1.053 eV (CsAlH4), for the minimum value, and 1.211 eV (LiAlH4) < 1.281 eV (NaAlH4) < 1.291 eV (KAlH4) < 1.361 eV (CsAlH4) for the average value. These characteristics are also achieved in M4Al4H16 (primitive cell) in addition to M8Al8H32 (2 × 1 × 1 supercell, present work). In particular, the descending order of hydrogen dissociation energies (Ed-H2) agrees well with that of the calculated formation enthalpies (ΔH) and cohesive energies (Ecoh) (Table 1), as well as the experimental onset dehydriding temperature, 443 K (LiAlH4) [20] < 503 K (NaAlH4) [21] < 573 K (KAlH4) [20] < 600 K (CsAlH4) [19] (as described above). However, this order is opposite to that of the cation electronegativity, namely, 0.98 (Li) > 0.93 (Na) > 0.82 (K) > 0.79 (Cs). The results further verify the fact that the alkali metal aluminum hydrides of MAlH4 with more electronegative alkali cations are thermally less stable and, therefore, energetically favorable for hydrogen desorption.

3.3. Electronic Structures

As described above, Al-H and H-H bonds can be detected in the [AlH4] groups of MAlH4 (M = Li, Na, K, and Cs). Thus, upon dehydrogenation, the release of HA or/and HX from the considered [AlH4] units (Figure 1) may be accompanied by the separation/rupture of Al-H and H-H bonds. In addition, weakened Al-H bonds are believed to be beneficial to the dehydrogenation of aluminum hydrides [55,56,57]. In consideration of these facts, determined the bonding features between Al-H and H-H according to the density of states, charge density distributions, and Mulliken populations should be help clarify the dehydriding mechanism of the alkali alanates of MAlH4 (M = Li, Na, K, and Cs).
Figure 4 presents the total (TDOS) and partial density of states (PDOS) of MAlH4 (M = Li, Na, K, and Cs), with a Fermi level (EF) at 0 eV and H atoms (HA and HB) from the considered [AlH4] unit in Figure 1. As can be seen, the DOS pictures of the studied hydrides of MAlH4, especially M = Li, Na and K, are very similar to each other. There are orbital hybridizations between the Al and H atoms whether below Fermi level or above, thereby demonstrating a bonding interaction between Al-H. Similarly, the bonding precipitated by the interaction between H and H atoms is achieved through the hybridizations between the s states of the H atoms. These bonds of Al-H and H-H described in the DOS pictures (Figure 4) can also be detected by the charge density distribution of MAlH4 (M = Li, Na, K, and Cs) (Figure 5). As shown in the Figures, the overlapping electronic clouds with appropriate distances between Al and H atoms (1.625–1.641 Ǻ), and H and H atoms (2.650–2.721 Ǻ) in Figure 5, may contribute to the formation of Al-H and H-H bonds, respectively [31,40]. In the DOS pictures, it is worth noting that the peaks of the TDOS contribution from Al and H electronic states (marked as I and II in Figure 4) tend to increase at lower cation electronegativity due to the enhanced PDOS of Al s and p states and H s states (such as KAlH4 and CsAlH4), so one or both Al-H and H-H bonding interactions may become stronger with a decreasing cation electronegativity [50,51,58].
To further quantitatively elucidate the bonding characteristics between Al-H and H-H, a Mulliken population analysis was performed on the MAlH4 (M = Li, Na, K, and Cs) compounds. The results, including the average bond order (BO), the average bond length (BL), and the scaled bond order (BOs) between Al-H and H-H, are listed in Table 3. Here, BO indicates the overlapping electron population between atoms, and is useful when considering bonding characteristics with an ionic (BO < 0, BO = 0) or covalent nature (BO > 0) [40,41,58,59,60,61]. BOs, defined as BOs = BO/BL, is helpful for assessing the relative bonding strength between atoms. A bond with a higher BOs value is expected to be stronger [40,41,58]. It can be seen from Table 3 that for all the studied compounds, Al-H bonds with a positive bond order (BOAl-H > 0) show a covalent character, while H-H bonds with a negative bond order (BOH-H < 0) exhibit an ionic nature. The H-H ionic interactions were found to be weakened at a higher electronegativity of the cation M. As shown, the scaled bond order between H-H (BOsH-H) decreases linearly (−0.013 Å−1 (LiAlH4) < −0.014 Å−1 (NaAlH4) < −0.021 Å−1 (KAlH4) < −0.022 Å−1 (CsAlH4)) with the increase in cation electronegativity (0.98 (Li) > 0.93 (Na) > 0.82 (K) > 0.79 (Cs)). The trend of the Al-H covalent interactions among the studied aluminum hydrides (except CsAlH4) is similar to that of the H-H ionic interactions. Araújo et al. [55] reported that hydrogen atoms held by weak covalent and ionic bonds may lead to lower dissociation temperatures for complex alkali metal aluminum hydrides. In addition, many previous studies have shown that reducing the Al-H covalent bonding strength in metal aluminum hydrides facilitates their decomposition for H-desorption [55,56,57,62]. Our calculated results with respect to MAlH4 (M = Li, Na, K, and Cs) mainly support these findings. That is, aluminum hydrides of MAlH4 with larger cation electronegativity and weaker H-H ionic and Al-H covalent interactions exhibit lower hydrogen dissociation energies (Table 2). As an example, the LiAlH4 hydride containing the weakest H-H ionic (BOsH-H = −0.013 Å−1, Table 3) and the Al-H covalent interactions (BOsAl-H = 0.476 Å−1, Table 3) show the lowest hydrogen dissociation energies for hydrogen desorption (Ed-H2 = 1.211 eV, Table 2) relative to the other three hydrides, namely, NaAlH4, KAlH4, and CsAlH4. The results described above lead us to the conclusion that weakening the ionic and covalent H bonds (such as H-H ionic and Al-H covalent bonds) by increasing cation electronegativity helps to modify the dehydrogenation performance of the alkali metal aluminum hydrides of MAlH4 (M = Li, Na, K, and Cs). The incorporation of more electronegative elements (compared to M) into the MAlH4 bulk could be used for this purpose.

4. Conclusions

First-principles calculations were performed on the alkali alanates of MAlH4 (M = Li, Na, K, and Cs) to investigate their thermal stability, hydrogen dissociation energy, and electronic structures. The results show that cation electronegativity (χP) is a good indicator with which to assess the thermal stability and dehydrogenation ability of MAlH4 (M = Li, Na, K, and Cs). For MAlH4 with a higher χP, on the one hand, it is thermally less stable because the formation enthalpy ΔH and cohesive energy Eoch become less negative and the M+→[AlH4] charge transfer is suppressed. On the other hand, it is energetically favorable for hydrogen desorption (HA→HX and HA–HX, especially HA and HX from two [AlH4] units), which is associated with its poor stability and weaker H-H ionic and Al-H covalent interactions. Our work provides new insights into the dehydrogenation of MAlH4 (M = Li, Na, K, and Cs) and is useful for designing advanced aluminum hydrides with favorable H-desorption properties.

Author Contributions

Conceptualization, R.Z. and W.J.; Methodology, R.Z., X.M., Y.H., C.H. and W.J.; Data Curation, R.Z., X.M., Y.H., C.H. and X.Z.; Formal Analysis, R.Z., X.M. and W.J.; Validation, R.Z., Y.H., X.Z., Y.M. and Q.W.; Funding Acquisition, W.J.; Writing—Original Draft, R.Z. and W.J.; Supervision, X.M. and W.J.; Writing—Review and Editing, W.J. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the National Natural Science Foundation of China (51661002), and the Natural Science Foundation of Guangxi (2018GXNSFAA138189).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

This work was supported by the National Natural Science Foundation of China (51661002), the Natural Science Foundation of Guangxi (2018GXNSFAA138189), and the high-performance computing platform of Guangxi University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Schlapbach, L.; Züttel, A. Hydrogen-storage materials for mobile applications. Nature 2001, 414, 353–358. [Google Scholar] [CrossRef] [PubMed]
  2. Zhu, M.; Lu, Y.S.; Ouyang, L.Z.; Wang, H. Thermodynamic tuning of Mg-based hydrogen storage alloys: A review. Materials 2013, 6, 4654–4674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Rusman, N.A.A.; Dahari, M. A review on the current progress of metal hydrides material for solid-state hydrogen storage applications. Int. J. Hydrogen Energy 2016, 41, 12108–12126. [Google Scholar] [CrossRef]
  4. Li, L.; Huang, Y.K.; An, C.H.; Wang, Y.J. Lightweight hydrides nanocomposites for hydrogen storage: Challenges, progress and prospects. Sci. China Mater. 2019, 62, 1597–1625. [Google Scholar] [CrossRef] [Green Version]
  5. Ali, N.A.; Ismail, M. Modification of NaAlH4 properties using catalysts for solid-state hydrogen storage: A review. Int. J. Hydrogen Energy 2021, 46, 766–782. [Google Scholar] [CrossRef]
  6. Li, H.W.; Orimo, S.; Nakamori, Y.; Miwa, K.; Ohba, N.; Towata, S.; Züttel, A. Materials designing of metal borohydrides: Viewpoints from thermodynamical stabilities. J. Alloys Compd. 2007, 446–447, 315–318. [Google Scholar] [CrossRef] [Green Version]
  7. Ren, Z.H.; Zhang, X.; Huang, Z.G.; Hu, J.J.; Li, Y.Z.; Zheng, S.Y.; Gao, M.X.; Pan, H.G.; Liu, Y.F. Controllable synthesis of 2D TiH2 nanoflakes with superior catalytic activity for low-temperature hydrogen cycling of NaAlH4. Chem. Eng. J. 2022, 427, 131546. [Google Scholar] [CrossRef]
  8. Guo, Y.H.; Yu, X.B.; Gao, L.; Xia, G.L.; Guo, Z.P.; Liu, H.K. Significantly improved dehydrogenation of LiBH4 destabilized by TiF3. Energy Environ. Sci. 2010, 3, 465–470. [Google Scholar] [CrossRef] [Green Version]
  9. Li, Y.; Zhang, Y.; Gao, M.X.; Pan, H.G.; Liu, Y.F. Preparation and catalytic effect of porous Co3O4 on the hydrogen storage properties of a Li-B-N-H system. Prog. Nat. Sci.-Mater. 2017, 27, 132–138. [Google Scholar] [CrossRef]
  10. Wei, S.; Liu, J.X.; Xia, Y.P.; Zhang, H.Z.; Cheng, R.G.; Sun, L.X.; Xu, F.; Huang, P.R.; Rosei, F.; Pimerzin, A.A.; et al. Remarkable catalysis of spinel ferrite XFe2O4 (X = Ni, Co, Mn, Cu, Zn) nanoparticles on the dehydrogenation properties of LiAlH4: An experimental and theoretical study. J. Mater. Sci. Technol. 2022, 111, 189–203. [Google Scholar] [CrossRef]
  11. Nakamori, Y.; Miwa, K.; Ninomiya, A.; Li, H.W.; Ohba, N.; Towata, S.; Züttel, A.; Orimo, S. Correlation between thermodynamical stabilities of metal borohydrides and cation electronegativites: First-principles calculations and experiments. Phys. Rev. B 2006, 74, 045126. [Google Scholar] [CrossRef] [Green Version]
  12. Nakamori, Y.; Li, H.W.; Kikuchi, K.; Aoki, M.; Miwa, K.; Towata, S.; Orimo, S. Thermodynamical stabilities of metal-borohydrides. J. Alloys Compd. 2007, 446–447, 296–300. [Google Scholar] [CrossRef]
  13. Zhang, B.J.; Liu, B.H. Hydrogen desorption from LiBH4 destabilized by chlorides of transition metal Fe, Co, and Ni. Int. J. Hydrogen Energy 2010, 35, 7288–7294. [Google Scholar] [CrossRef]
  14. Mo, X.H.; Jiang, W.Q. Dehydrogenation properties of LiBH4 modified by Mg from first-principles calculations. J. Alloys Compd. 2018, 735, 668–676. [Google Scholar] [CrossRef]
  15. Cai, W.T.; Yang, Y.Z.; Tao, P.J.; Ouyang, L.Z.; Wang, H. Correlation between structural stability of LiBH4 and cation electronegativity in metal borides: An experimental insight for catalyst design. Dalton Trans. 2018, 47, 4987–4993. [Google Scholar] [CrossRef] [PubMed]
  16. Chen, X.; Li, Z.; Zhang, Y.; Liu, D.M.; Wang, C.Y.; Li, Y.T.; Si, T.Z.; Zhang, Q.G. Enhanced low-temperature hydrogen storage in nanoporous Ni-based alloy supported LiBH4. Front. Chem. 2020, 8, 283. [Google Scholar] [CrossRef] [PubMed]
  17. Nakamori, Y.; Orimo, S.I. Destabilization of Li-based complex hydrides. J. Alloys Compd. 2004, 370, 271–275. [Google Scholar] [CrossRef]
  18. Bai, Y.; Pei, Z.W.; Wu, F.; Wu, C. Role of metal electronegativity in the dehydrogenation thermodynamics and kinetics of composite metal borohydride-LiNH2 hydrogen storage materials. ACS Appl. Mater. Interfaces 2018, 10, 9514–9521. [Google Scholar] [CrossRef] [PubMed]
  19. Weidenthaler, C. Crystal structure evolution of complex metal aluminum hydrides upon hydrogen release. J. Energy Chem. 2020, 42, 133–143. [Google Scholar] [CrossRef] [Green Version]
  20. Mamatha, M.; Weidenthaler, C.; Pommerin, A.; Felderhoff, M.; Schüth, F. Comparative studies of the decomposition of alanates followed by in situ XRD and DSC methods. J. Alloys Compd. 2006, 416, 303–314. [Google Scholar] [CrossRef]
  21. Finholt, A.E.; Barbaras, G.D.; Barbaras, G.K.; Urry, G.; Wartik, T.; Schlesinger, H.I. The preparation of sodium and calcium aluminum hydrides. J. Inorg. Nucl. Chem. 1955, 1, 317–325. [Google Scholar] [CrossRef]
  22. Nyahuma, F.M.; Zhang, L.T.; Song, M.C.; Lu, X.; Xiao, B.B.; Zheng, J.G.; Wu, F.Y. Significantly improved hydrogen storage behaviors in MgH2 with Nb nanocatalyst. Int. J. Min. Met. Mater. 2022, 29, 1788–1797. [Google Scholar] [CrossRef]
  23. Liu, Y.; Huang, Z.N.; Gao, X.; Wang, Y.Q.; Wang, F.; Zheng, S.S.; Guan, S.N.; Yan, H.L.; Yang, X.; Jia, W.H. Effect of novel La-based alloy modification on hydrogen storage performance of magnesium hydride: First-principles calculation and experimental investigation. J. Power Sources 2022, 551, 232187. [Google Scholar] [CrossRef]
  24. Segall, M.D.; Lindan, P.J.D.; Probert, M.J.; Pickard, C.J.; Hasnip, P.J.; Clark, S.J.; Payne, M.C. First-principles simulation: Ideas, illustrations and the CASTEP code. J. Phys. Condes. Matter 2002, 14, 2717–2744. [Google Scholar] [CrossRef]
  25. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Pfrommer, B.G.; Côté, M.; Louie, S.G.; Cohen, M.L. Relaxation of crystals with the quasi-newton method. J. Comput. Phys. 1997, 131, 233–240. [Google Scholar] [CrossRef] [Green Version]
  27. Hauback, B.C.; Brinks, H.W.; Fjellvåg, H. Accurate structure of LiAlD4 studied by combined powder neutron and X-ray diffraction. J. Alloys Compd. 2002, 346, 184–189. [Google Scholar] [CrossRef]
  28. Ozolins, V.; Majzoub, E.H.; Udovic, T.J. Electronic structure and Rietveld refinement parameters of Ti-doped sodium alanates. J. Alloys Compd. 2004, 375, 1–10. [Google Scholar] [CrossRef]
  29. Hauback, B.C.; Brinks, H.W.; Heyn, R.H.; Blom, R.; Fjellvåg, H. The crystal structure of KAlD4. J. Alloys Compd. 2005, 394, 35–38. [Google Scholar] [CrossRef]
  30. Bernert, T.; Krech, D.; Kockelmann, W.; Felderhoff, M.; Frankcombe, T.J.; Weidenthaler, C. Crystal structure relation between tetragonal and orthorhombic CsAlD4: DFT and time-of-flight neutron powder diffraction studies. Eur. J. Inorg. Chem. 2015, 2015, 5545–5550. [Google Scholar] [CrossRef]
  31. Sirsch, P.; Che, F.N.; Titah, J.T.; McGrady, G.S. Hydride-hydride bonding interactions in the hydrogen stoage materials AlH3, MgH2, and NaAlH4. Chem. Eur. J. 2012, 18, 9476–9480. [Google Scholar] [CrossRef]
  32. Sorte, E.G.; Emery, S.B.; Majzoub, E.H.; Ellis-Caleo, T.; Ma, Z.L.; Hammann, B.A.; Hayes, S.E.; Bowman, R.C.; Conradi, M.S. NMR Study of Anion Dynamics in Solid KAlH4. J. Phys. Chem. C 2014, 118, 5725–5732. [Google Scholar] [CrossRef]
  33. Ravindran, P.; Vajeeston, P.; Vidya, R.; Fjellvåg, H.; Kjekshus, A. Modeling of hydrogen storage materials by density-functional calculations. J. Power Sources 2006, 159, 88–99. [Google Scholar] [CrossRef]
  34. Setten, M.J.V. Electronic structure and formation enthalpy of hydroaluminates and hydroborates. Encycl. Mater. Sci. Technol. 2008, 043152, 1–6. [Google Scholar]
  35. Fukai, Y. The Metal-Hydrogen System; Volume 21 of Springer Series in Material Science; Springer: Berlin/Heidelberg, Germany, 1993. [Google Scholar]
  36. Shi, B.; Song, Y. Influence of transition metals Fe, Ni, and Nb on dehydrogenation characteristics of Mg(BH4)2: Electronic structure mechanisms. Int. J. Hydrogen Energy 2013, 38, 6417–6424. [Google Scholar] [CrossRef]
  37. Zhang, J.; Sun, L.Q.; Zhou, Y.C.; Peng, P. Dehydrogenation thermodynamics of magnesium hydride doped with transition metals: Experimental and theoretical studies. Comput. Mater. Sci. 2015, 98, 211–219. [Google Scholar] [CrossRef]
  38. Kumar, A.; Muthukumar, P.; Sharma, P.; Kumar, E.A. Absorption based solid state hydrogen storage system: A review. Sustain. Energy Technol. Assess. 2022, 52, 102204. [Google Scholar] [CrossRef]
  39. Zhao, L.; Xu, F.; Zhang, C.C.; Wang, Z.Y.; Ju, H.Y.; Gao, X.; Zhang, X.X.; Sun, L.X.; Liu, Z.W. Enhanced hydrogen storage of alanates: Recent progress and future perspectives. Prog. Nat. Sci. 2021, 31, 165–179. [Google Scholar] [CrossRef]
  40. Mo, X.H.; Liang, J.Q.; Wang, W.H.; Jiang, W.Q. First-principles study on the dehydrogenation of Li4BN3H10 modified by Co. Int. J. Hydrogen Energy 2021, 46, 11815–11823. [Google Scholar] [CrossRef]
  41. Jiang, W.Q.; Chen, Y.J.; Mo, X.H.; Li, X.L. First-principles investigation of LaMg2Ni and its hydrides. Sci. Rep. 2020, 10, 12167. [Google Scholar] [CrossRef]
  42. Jiang, W.Q.; Qin, C.S.; Zhu, R.R.; Guo, J. Annealing effect on hydrogen storage property of Co-free La1.8Ti0.2MgNi8.7Al0.3 alloy. J. Alloys Compd. 2013, 565, 37–43. [Google Scholar] [CrossRef]
  43. Wang, H.Y.; Zhang, N.; Wang, C.; Jiang, Q.C. First-principles study of the generalized stacking fault energy in Mg-3Al-3Sn alloy. Scr. Mater. 2011, 65, 723–726. [Google Scholar] [CrossRef]
  44. Løvvik, O.M.; Opalka, S.M.; Brinks, H.W.; Hauback, B.C. Crystal structure and thermodynamic stability of the lithium alanates LiAlH4 and Li3AlH6. Phys. Rev. B 2004, 69, 134117. [Google Scholar] [CrossRef]
  45. Lide, D.R. CRC Handbook of Chemistry and Physics, 80th ed.; CRC Press: Boca Raton, FL, USA, 2000. [Google Scholar]
  46. Wang, X.; Yu, X.H.; Zhong, Y.; Rong, J.; Li, X.Y.; Feng, J.; Zhan, Z.L. Stability and mechanical peroperties of high-La content La-Ni phases by first-principles study. Int. J. Mod. Phys. B 2018, 32, 1850242. [Google Scholar] [CrossRef]
  47. Miwa, K.; Ohba, N.; Towata, S.; Nakamori, Y.; Orimo, S. First-principles study on lithium borohydride LiBH4. Phys. Rev. B 2004, 69, 245120. [Google Scholar] [CrossRef]
  48. Miwa, K.; Ohba, N.; Towata, S.; Nakamori, Y.; Orimo, S. First-principles study on copper-substituted lithium borohydride, (Li1−xCux)BH4. J. Alloys Compd. 2005, 404–406, 140–143. [Google Scholar] [CrossRef]
  49. Dai, J.H.; Song, Y.; Yang, R. Intrinsic mechanisms on enhancement of hydrogen desorption from MgH2 by (001) surface doping. Int. J. Hydrogen Energy 2011, 36, 12939–12949. [Google Scholar] [CrossRef]
  50. Mo, X.H.; Long, L.X.; Tan, W.B.; Huang, Y.P.; Lu, J.W.; Zhao, Y.W.; Jiang, W.Q. First-principles investigation of dehydrogenation of Cu-doped LiBH4. Solid State Commun. 2021, 326, 114184. [Google Scholar] [CrossRef]
  51. Mo, X.H.; Jiang, W.Q.; Cao, S.L. First-principles study on the dehydrogenation characteristics of LiBH4 modified by Ti. Results Phys. 2017, 7, 3236–3242. [Google Scholar] [CrossRef]
  52. Shi, Q.; Voss, J.; Jacobsen, H.S.; Lefmann, K.; Zamponi, M.; Vegge, T. Point defect dynamics in sodium aluminum hydrides–a combined quasielastic neutron scattering and density functional theory study. J. Alloys Compd. 2007, 446–447, 469–473. [Google Scholar] [CrossRef]
  53. Araújo, C.M.; Li, S.; Ahuja, R.; Jena, P. Vacancy-mediated hydrogen desorption in NaAlH4. Phys. Rev. B 2005, 72, 165101. [Google Scholar] [CrossRef] [Green Version]
  54. Wang, H.; Tezuka, A.; Ogawa, H.; Ikeshoji, T. First-principles study of hydrogen vacancies in sodium alanate with Ti substitution. J. Phys. Condes. Matter 2010, 22, 205503. [Google Scholar] [CrossRef]
  55. Araújo, C.M.; Ahuja, R.; Guillén, J.M.O.; Jena, P. Role of titanium in hydrogen desorption in crystalline sodium alanate. Appl. Phys. Lett. 2005, 86, 251913. [Google Scholar] [CrossRef]
  56. Zhang, X.; Ren, Z.H.; Zhang, X.L.; Gao, M.X.; Pan, H.G.; Liu, Y.F. Triggering highly stable catalytic activity of metallic titanium for hydrogen storage in NaAlH4 by preparing ultrafine nanoparticles. J. Mater. Chem. A 2019, 7, 4651–4659. [Google Scholar] [CrossRef]
  57. Liu, Z.Y.; Liu, J.X.; Wei, S.; Xia, Y.P.; Cheng, R.G.; Sun, L.X.; Xu, F.; Huang, P.R.; Bu, Y.T.; Cheng, J.; et al. Improved hydrogen storage properties and mechanisms of LiAlH4 doped with Ni/C nanoparticles anchored on large-size Ti3C2Tx. J. Alloys Compd. 2023, 931, 167353. [Google Scholar] [CrossRef]
  58. Jiang, W.Q.; Cao, S.L. Effect of Al on the dehydrogenation of LiBH4 from first-principles calculations. Int. J. Hydrogen Energy 2017, 42, 6181–6188. [Google Scholar]
  59. Segall, M.D.; Shah, R.; Pickard, C.J.; Payne, M.C. Population analysis of plane-wave electronic structure calculations of bulk materials. Phys. Rev. B 1996, 54, 16317–16320. [Google Scholar] [CrossRef] [Green Version]
  60. Wang, H.P.; Wang, X.M.; Ge, F.F.; Zhou, M.J.; Wu, W.D.; Lu, T.C. Density function study of H2 adsorption on LiB (010) surface. Phys. B 2010, 405, 1792–1795. [Google Scholar]
  61. Zhu, C.Y.; Liu, Y.H.; Duan, D.F.; Cui, T. Structural transitions of NaAlH4 under high pressure by first-principles calculations. Phys. B 2011, 406, 1612–1614. [Google Scholar] [CrossRef]
  62. Berseth, P.A.; Harter, A.G.; Zidan, R.; Blomqvist, A.; Araújo, C.M.; Scheicher, R.H.; Ahuja, R.; Jena, P. Carbon nanomaterials as catalysts for hydrogen uptake and release in NaAlH4. Nano Lett. 2009, 9, 1501–1505. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The crystal models of MAlH4 (M = Li, Na, K, and Cs) with 2 × 1 × 1 supercell: (a) LiAlH4, (b) NaAlH4, (c) KAlH4, and (d) CsAlH4. Li, Na, K, Cs, Al, and H atoms are denoted by red, green, blue, orange, pink, and white spheres, respectively. The H atoms labeled as HA, HB, HC, HD, HE, HF, HG, and HH in two [AlH4] units are considered for hydrogen desorption.
Figure 1. The crystal models of MAlH4 (M = Li, Na, K, and Cs) with 2 × 1 × 1 supercell: (a) LiAlH4, (b) NaAlH4, (c) KAlH4, and (d) CsAlH4. Li, Na, K, Cs, Al, and H atoms are denoted by red, green, blue, orange, pink, and white spheres, respectively. The H atoms labeled as HA, HB, HC, HD, HE, HF, HG, and HH in two [AlH4] units are considered for hydrogen desorption.
Batteries 09 00179 g001
Figure 2. The formation enthalpy (ΔH) as a function of cation electronegativity (χP) for MAlH4 (M = Li, Na, K, and Cs) alanates. The straight line, ΔH = 296.033χP − 402.19, indicates least square fitting.
Figure 2. The formation enthalpy (ΔH) as a function of cation electronegativity (χP) for MAlH4 (M = Li, Na, K, and Cs) alanates. The straight line, ΔH = 296.033χP − 402.19, indicates least square fitting.
Batteries 09 00179 g002
Figure 3. The formed AlH3 group, hydrogen dissociation energy (Ed-H2), and total energy (E) for MAlH4 (M = Li, Na, K, and Cs) with HA and HX desorption: (a) HA and HX from one [AlH4] unit; (b) HA and HX from two [AlH4] units. The bond lengths between Al and H atoms in AlH3 group are described (in Ǻ).
Figure 3. The formed AlH3 group, hydrogen dissociation energy (Ed-H2), and total energy (E) for MAlH4 (M = Li, Na, K, and Cs) with HA and HX desorption: (a) HA and HX from one [AlH4] unit; (b) HA and HX from two [AlH4] units. The bond lengths between Al and H atoms in AlH3 group are described (in Ǻ).
Batteries 09 00179 g003
Figure 4. The total (TDOS) and partial density of states (PDOS) for MAlH4 (M = Li, Na, K, and Cs), with Fermi level (EF, marked with vertical dotted line) at 0 eV and H atoms (HA and HB) from considered [AlH4] unit in Figure 1: (a) LiAlH4, (b) NaAlH4, (c) KAlH4, and (d) CsAlH4. The TDOS labeled by I and II are mainly contributed by Al and H electronic states.
Figure 4. The total (TDOS) and partial density of states (PDOS) for MAlH4 (M = Li, Na, K, and Cs), with Fermi level (EF, marked with vertical dotted line) at 0 eV and H atoms (HA and HB) from considered [AlH4] unit in Figure 1: (a) LiAlH4, (b) NaAlH4, (c) KAlH4, and (d) CsAlH4. The TDOS labeled by I and II are mainly contributed by Al and H electronic states.
Batteries 09 00179 g004aBatteries 09 00179 g004b
Figure 5. The electronic density contours for MAlH4 (M = Li, Na, K, and Cs) with the contour line from 0.03 to 0.18 electrons/Ǻ3: (a) LiAlH4, (b) NaAlH4, (c) KAlH4, and (d) CsAlH4. Li, Na, K, Cs, Al, and H atoms are denoted by red, green, blue, orange, pink, and white spheres, respectively. The shortest distances between Al and H atoms and H and H atoms in this figure are described (in Ǻ).
Figure 5. The electronic density contours for MAlH4 (M = Li, Na, K, and Cs) with the contour line from 0.03 to 0.18 electrons/Ǻ3: (a) LiAlH4, (b) NaAlH4, (c) KAlH4, and (d) CsAlH4. Li, Na, K, Cs, Al, and H atoms are denoted by red, green, blue, orange, pink, and white spheres, respectively. The shortest distances between Al and H atoms and H and H atoms in this figure are described (in Ǻ).
Batteries 09 00179 g005
Table 1. The space group (SP), relaxed lattice parameters (R) and cell volume (V), formation enthalpy (ΔH), cohesive energy (Ecoh), charge transfer from cation M to anion [AlH4] (C), and alkali cation electronegativity (χP) of MAlH4 (M = Li, Na, K, and Cs).
Table 1. The space group (SP), relaxed lattice parameters (R) and cell volume (V), formation enthalpy (ΔH), cohesive energy (Ecoh), charge transfer from cation M to anion [AlH4] (C), and alkali cation electronegativity (χP) of MAlH4 (M = Li, Na, K, and Cs).
CompoundsSPR (Ǻ)V (Ǻ3)ΔH (kJ/mol)Ecoh (kJ/mol)C (e)χP
abc
LiAlH4P21/C4.9088.0317.953289.762−117.156 −1684.320.210.98
NaAlH4I41/A4.9864.98611.178277.905−120.468 −1684.7040.280.93
KAlH4Pnma8.8965.8107.399382.448−157.2−1719.5520.350.82
CsAlH4Pnma10.0186.1638.077498.704−171.9−1740.7680.420.79
Table 2. The hydrogen dissociation energy of MAlH4 (M = Li, Na, K, and Cs), including atomic hydrogen dissociation energy (Ed-H) for HA and HX (HB, HC, HD, HE, HF, HG, HH) desorption and molecular hydrogen dissociation energy (Ed-H2) for HA→HX (outside the bracket) and HA–HX desorption (inside the bracket). The minimum and average values of all Ed-H2 are also listed.
Table 2. The hydrogen dissociation energy of MAlH4 (M = Li, Na, K, and Cs), including atomic hydrogen dissociation energy (Ed-H) for HA and HX (HB, HC, HD, HE, HF, HG, HH) desorption and molecular hydrogen dissociation energy (Ed-H2) for HA→HX (outside the bracket) and HA–HX desorption (inside the bracket). The minimum and average values of all Ed-H2 are also listed.
Hydrogen Dissociation Energy (eV)LiAlH4NaAlH4KAlH4CsAlH4
Atomic hydrogen desorption
Ed-H
HA1.8151.7191.7021.720
HB−0.080−0.1290.0800.142
HC−0.526−0.077−0.038−0.006
HD−0.019−0.077−0.0370.000
HE−0.862−0.736−0.721−0.658
HF−0.971−0.747−0.722−0.663
HG−0.788−0.637−0.723−0.664
HH−0.786−0.634−0.723−0.665
Molecular hydrogen desorption
Ed-H2
HA→HB (HA–HB)1.735 (1.733)1.590 (1.592)1.782 (1.773)1.862 (1.866)
HA→HC (HA–HC)1.289 (1.287)1.642 (1.545)1.664 (1.661)1.714 (1.726)
HA→HD (HA–HD)1.796 (1.704)1.642 (1.660)1.665 (1.664)1.720 (1.708)
HA→HE (HA–HE)0.953 (0.832)0.983 (0.997)0.981 (0.984)1.062 (1.057)
HA→HF (HA–HF)0.844 (0.844)0.972 (0.975)0.980 (0.982)1.057 (1.055)
HA→HG (HA–HG)1.027 (0.856)1.082 (1.085)0.979 (0.986)1.056 (1.053)
HA→HH (HA–HH)1.029 (1.029)1.085 (1.081)0.979 (0.993)1.055 (1.065)
Minimum value0.8320.9720.9791.053
Average value1.2111.2811.2911.361
Table 3. The Mulliken population for MAlH4 (M = Li, Na, K, and Cs), including the average bond order (BO), average bond length (BL), and scaled bond order (BOs) between Al-H and H-H.
Table 3. The Mulliken population for MAlH4 (M = Li, Na, K, and Cs), including the average bond order (BO), average bond length (BL), and scaled bond order (BOs) between Al-H and H-H.
CompoundsAl-HH-H
BOBL (Å)BOS−1)BOBL (Å)BOS−1)
LiAlH40.7781.6340.476−0.0362.751−0.013
NaAlH40.8551.6420.521−0.0392.755−0.014
KAlH40.9061.6340.555−0.0552.668−0.021
CsAlH40.8931.6370.545−0.0582.673−0.022
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhou, R.; Mo, X.; Huang, Y.; Hu, C.; Zuo, X.; Ma, Y.; Wei, Q.; Jiang, W. Dehydrogenation of Alkali Metal Aluminum Hydrides MAlH4 (M = Li, Na, K, and Cs): Insight from First-Principles Calculations. Batteries 2023, 9, 179. https://doi.org/10.3390/batteries9030179

AMA Style

Zhou R, Mo X, Huang Y, Hu C, Zuo X, Ma Y, Wei Q, Jiang W. Dehydrogenation of Alkali Metal Aluminum Hydrides MAlH4 (M = Li, Na, K, and Cs): Insight from First-Principles Calculations. Batteries. 2023; 9(3):179. https://doi.org/10.3390/batteries9030179

Chicago/Turabian Style

Zhou, Rui, Xiaohua Mo, Yong Huang, Chunyan Hu, Xiaoli Zuo, Yu Ma, Qi Wei, and Weiqing Jiang. 2023. "Dehydrogenation of Alkali Metal Aluminum Hydrides MAlH4 (M = Li, Na, K, and Cs): Insight from First-Principles Calculations" Batteries 9, no. 3: 179. https://doi.org/10.3390/batteries9030179

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