Next Article in Journal
Recovery of Iron, Silver and Lead from Zinc Ferrite Residue
Previous Article in Journal
Study of an SSA-BP Neural Network-Based Strength Prediction Model for Slag–Cement-Stabilized Soil
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Quantum Chemical Investigation on the Material Properties of Al-Based Hydrides XAl2H2 (X = Ca, Sr, Sc, and Y) for Hydrogen Storage Applications

1
Department of Physics, Shanxi Agricultural University, Jinzhong 030801, China
2
School of Materials Science and Engineering, Lanzhou University of Technology, Lanzhou 730050, China
3
Institute of Computational (Digital) Materials, Lanzhou University of Technology, Lanzhou 730050, China
4
Jiuquan Iron and Steel (Group) Corporation, Jiuquan 735000, China
*
Author to whom correspondence should be addressed.
Materials 2025, 18(15), 3521; https://doi.org/10.3390/ma18153521
Submission received: 27 June 2025 / Revised: 14 July 2025 / Accepted: 25 July 2025 / Published: 27 July 2025
(This article belongs to the Section Energy Materials)

Abstract

Aluminum–hydrogen compounds have drawn considerable interest for applications in solid-state hydrogen storage. The structural, hydrogen storage, electronic, mechanical, phonon, and thermodynamic properties of XAl2H2 (X = Ca, Sr, Sc, Y) hydrides are investigated using density functional theory. These hydrides exhibit negative formation energies in the hexagonal phase, indicating their thermodynamic stability. The gravimetric hydrogen storage capacities of CaAl2H2, SrAl2H2, ScAl2H2, and YAl2H2 are calculated to be 1.41 wt%, 0.94 wt%, 1.34 wt%, and 0.93 wt%, respectively. Analysis of the electronic density of states reveals metallic characteristics. Furthermore, the calculated elastic constants satisfy the Born stability criteria, confirming their mechanical stability. Additionally, through phonon spectra analysis, dynamical stability is verified for CaAl2H2 and SrAl2H2 but not for ScAl2H2 and YAl2H2. Finally, we present temperature-dependent thermodynamic properties. This research reveals that XAl2H2 (X = Ca, Sr, Sc, Y) materials represent promising candidates for solid-state hydrogen storage, providing a theoretical foundation for further studies on XAl2H2 systems.

1. Introduction

The exploitation of traditional energy resources, such as coal, natural gas, and oil, has propelled societal progress. However, this dependence has led to numerous challenges, including environmental deterioration, the depletion of energy reserves, rising carbon emissions, and the worsening of climate change [1,2]. To address these issues, increasing attention is being directed toward the exploration of abundant and clean alternative energy sources [3]. Hydrogen energy, characterized by its high calorific value, non-polluting properties, and abundant availability, has emerged as one of the most promising alternatives [4]. Hydrogen storage, as one of the four key components of the hydrogen energy economy (the others being hydrogen production, transportation, and utilization), is the primary factor constraining the widespread adoption of hydrogen energy [5]. Despite the presence of well-established hydrogen storage techniques such as compression and liquefaction, solid-state hydrogen storage has emerged as a highly promising field, drawing increasing research interest owing to its potential to enhance both safety and storage density [6,7]. In this context, solid-state hydrogen storage materials have garnered significant interest as promising sustainable energy solutions in recent years [8].
Among the wide array of these materials, the light metal hydrides, particularly those represented by aluminum-based hydrogen compounds [9,10,11,12,13,14,15,16], have consistently received considerable focus and sustained research attention owing to their high hydrogen capacity, reversible storage properties, and lower cost. In 2000, Gingl et al. [17] first reported the synthesis of a novel aluminum–hydrogen compound, SrAl2H2, which marked the discovery of the first Zintl phase hydride. As a critical class of solid-state materials, Zintl phases provide a theoretical framework for designing multifunctional compounds with diverse physical properties [18]. For instance, these materials exhibit remarkable characteristics such as thermoelectricity, superconductivity, and anomalous/spin Hall effects, rendering them highly attractive for applications in emerging fields like spintronics and optoelectronics [19]. The SrAl2H2 compound crystallizes in a hexagonal crystal structure containing a two-dimensional polyanionic [Al2H2]2− layer, where one aluminum atom is covalently bonded to each hydrogen atom. Björling et al. systematically investigated the hydrogenation reaction of the intermetallic compound AeE2 (Ae = Ca, Sr, Ba; E = Al, Ga, In), synthesizing and characterizing SrAl2H2, in addition to two isomorphic hydrides SrGa2H2 and BaGa2H2 [20]. Additionally, the first-principles method was employed to calculate the energy of these compounds, thereby theoretically confirming their stability. Lee et al. performed a comprehensive investigation into the lattice vibrational properties of SrAl2H2 and SrAlSiH by combining experimental and theoretical methods, revealing that the stability of solid aluminum hydride is inversely related to the strength of Al-H bonding [21]. Based on the results of the electronic structure and phonon spectral frequency calculated from first principles, Subedi et al. evaluated the chemical bonding properties of compounds such as SrAl2H2, SrGa2H2, and BaGa2H2, and they found that these compounds display a combination of covalent and ionic bonding characteristics [22].
To the best of our knowledge, apart from one report on the BaAl2−xSixH2−x (0.4 < x < 1.6) series hydrides [23], which are intermediate in composition and structure between BaSi2 and BaAl2H2, very few additional aluminum–hydrogen compounds with the same type of structure have been documented, either experimentally or computationally, since the synthesis of SrAl2H2 was reported in 2000. Moreover, there is a notable scarcity of research specifically focusing on their hydrogen storage capacity. On the one hand, investigation into the novel structure of aluminum hydride compounds facilitates the expansion of hydrogen storage types and paves the way for practical material innovations. On the other hand, analyzing complex hydrides, such as SrAl2H2, provides deeper insights into the atomic bonding mechanisms within these materials [24].
To the best of our knowledge, among the XAl2H2 family, only SrAl2H2 has been experimentally synthesized to date, with subsequent theoretical studies primarily focusing on Al-H bonding interactions. However, for the isomorphic XAl2H2-type compounds, no prior investigations have systematically examined their physical properties, particularly in the context of hydrogen storage applications. Against this background, we selected Ca, Sc, and Y as candidate elements for our study. These elements are located near Sr in the periodic table (either in the same period or adjacent groups) and exhibit similar electronic configurations and chemical bonding characteristics, indicating their potential to form stable compounds analogous to SrAl2H2. Guided by this rationale, we investigate the Al-based hydrides XAl2H2 (X = Ca, Sr, Sc, Y) as potential hydrogen energy carriers. Using first-principles calculations, we conduct a systematic analysis of their structural, hydrogen storage, electronic, mechanical, lattice dynamical, and thermodynamic properties.

2. Computational Details

The structural and electronic properties of XAl2H2 (X = Ca, Sr, Sc, Y) hydrides were systematically investigated using first-principles density functional theory (DFT) [25] calculations implemented in the Wien2k package (version: Wien2k_23) [26]. The all-electron full-potential linearized augmented plane wave (FP-LAPW) method with the generalized gradient approximation (GGA-PBE) [27] functional was employed for exchange-correlation treatment. Muffin-tin radii of 2.0 a.u. were used for metallic atoms (Ca/Sr/Sc/Y/Al), while 1.0 a.u. was applied for hydrogen atom. A cutoff energy of −6.0 Ry was set to separate valence-core states, and RMT·KMAX = 4.0 ensured plane-wave convergence. Structural optimizations were performed for all compounds using the 2DRoptimize module within Wien2k, which is specifically designed to handle hexagonal symmetry. This module facilitates anisotropic relaxation of lattice parameters (a and c) while maintaining hexagonal symmetry constraints, enabling efficient optimization of both the unit cell volume and the c/a ratio. Internal atomic coordinates were relaxed using the MINI module within Wien2k, which employs the “reverse-communication trust-region quasi-Newton” method [28] to minimize the total energy. Both lattice parameters and atomic positions were optimized using 8000 k-points in the full Brillouin zone, with convergence thresholds set to 10−4 Ry for total energy and 10−4 e/Å3 for charge density. These parameters ensure consistent accuracy between structural relaxations and subsequent property calculations. Elastic properties were derived from second derivatives of energy-based calculations using the IRelast code [29], providing polycrystalline elastic moduli, Poisson’s ratio, wave velocities, and Debye temperature. Anisotropy analysis was conducted using ELATool [30]. Phonon spectra and density of states were calculated via the finite displacement method [31] implemented in Phonopy [32] (1 × 1 × 1 supercell, 5 atoms), followed by thermodynamic property determination through harmonic approximation analysis of phonon density of states.

3. Results and Discussion

3.1. Structural and Hydrogen Storage Properties

The crystal structure of XAl2H2 (X = Ca, Sr, Sc, Y) hydrides, as illustrated in Figure 1a, exhibits a hexagonal structure with the P-3m1 (No. 164) space group. The unit cell comprises five atoms: the metal atom X (X = Ca, Sr, Sc, Y) occupies the 1a (0, 0, 0) Wyckoff site at the hexagonal corners; Al atoms occupy the 2d (1/3, 2/3, ZAl) Wyckoff site; and H atoms occupy another 2d (1/3, 2/3, ZH) Wyckoff sites. From the combination of Figure 1b,c, it can be observed that the XAl2H2 crystal structure consists of alternately stacked layers: a tetrahedral layer of hydrogen atoms (with aluminum atoms located within this tetrahedral framework) and a planar triangular layer of X atoms.
To determine the structural parameters of the equilibrium state for the system, the energy–volume relationship and the energy–c/a relationship have been optimized. The optimization curves are presented in Figure 2. Subsequently, the lattice constants, c/a ratio, unit cell volume V0, bulk modulus B0, and its pressure derivative B’ are calculated by fitting the data to the Birch–Murnaghan (B-M) equation of state, expressed as follows [33]:
E ( V ) = E 0 + 9 V 0 B 0 16 V 0 V 2 3 1 3 B + V 0 V 2 3 1 2 6 4 V 0 V 2 3
where E0 is the total energy, V0 is the equilibrium unit cell volume, B0 is the bulk modulus, and B’ is the pressure derivative of the bulk modulus.
The optimized internal coordinates of Al and H atoms, as well as the shortest interatomic distances between Al and H in XAl2H2 (X = Ca, Sr, Sc, Y), are summarized in Table 1, together with the experimental [17] and theoretical [22] values for SrAl2H2. First, it is evident that for SrAl2H2, the calculated coordinates of Al and H atoms align very closely with the experimental values. The distances between nearest neighbor Al and H atoms are marginally smaller than the theoretical values reported in reference [22] but remain closer to the experimental values. This not only validates the reliability of our computational methodology but also implies that the calculated values for other compounds, such as CaAl2H2, ScAl2H2, and YAl2H2, can serve as dependable references when experimental data are unavailable. Second, compared to CaAl2H2 and SrAl2H2, the positions of Al atoms in ScAl2H2 and YAl2H2 show minor variations, whereas the positions of H atoms in ScAl2H2 and YAl2H2, especially in YAl2H2, undergo significant shifts closer to the ab-plane. Consequently, this leads to an increase in the Al-H bond distances in YAl2H2, potentially weakening the Al-H chemical bond strength. Furthermore, the similar Al-H interatomic distances observed in the three materials, excluding YAl2H2, suggest that the Al-H bond strength in these three materials is approximately equivalent.
The calculated lattice parameters, unit cell volume, bulk modulus, formation enthalpy, and gravimetric hydrogen storage capacities for the series are presented in Table 2, along with reference values of the lattice parameters for SrAl2H2 [17,22]. It should be noted that our optimized lattice parameters for SrAl2H2 show better agreement with the experimental data, with errors of 0.08% and 0.38% for the parameters a and c, respectively. Meanwhile, the formation enthalpy of the series is calculated using Equation (2):
Δ H = E ( XAl 2 H 2 ) E ( X ) 2 E ( Al ) E ( H 2 )
where E(XAl2H2) denotes the total energy per formula unit of XAl2H2 (X = Ca, Sr, Sc, Y), E(X) and E(Al) represent the solid-phase elemental energies of X and Al, respectively, and E(H2) corresponds to the ground-state energy (2.32 Ry) of an isolated H2 molecules. The calculated formation enthalpies of −1.46, −1.80, −0.47, and −1.33 eV/f.u. for CaAl2H2, SrAl2H2, ScAl2H2, and YAl2H2, respectively, demonstrate thermodynamic stability through their negative values. This stability hierarchy, with SrAl2H2 exhibiting the most negative ΔH (−1.80 eV/f.u.), is indicative of enhanced structural robustness and implies that these hydrides may possess feasible synthesis potential under ambient conditions. Moreover, the quantitative evaluation of hydrogen storage efficiency generally relies on gravimetric analysis, a method that quantifies the mass ratio of stored hydrogen relative to the host material’s mass. Gravimetric storage capacity, a key metric for evaluating hydrogen storage performance, represents the mass of hydrogen stored per unit mass of the material and is defined by Equation (3) [34]:
C wt % = H M m H m Host + H M m H × 100 %
Here, H/M denotes the ratio of hydrogen atoms to material atoms, while mHost and mH represent the molar mass of the host material and hydrogen, respectively. As shown in Table 2, the gravimetric hydrogen storage capacities of CaAl2H2, SrAl2H2, ScAl2H2, and YAl2H2 are 1.41 wt%, 0.94 wt%, 1.34 wt%, and 0.93 wt%, respectively. Although the gravimetric hydrogen storage capacity obtained in this research is marginally below the 5.5 wt% target established by the U.S. Department of Energy (DOE), the aluminum-based compounds investigated here offer distinct advantages rooted in aluminum’s abundant natural reserves and cost-effectiveness, which serve as key strengths for practical applications. Furthermore, the hydrogen storage capacity is anticipated to be enhanced through emerging strategies, including doping, catalysis, or nanostructuring [8]. Notably, CaAl2H2 exhibits a relatively higher gravimetric hydrogen storage capacity due to the lower atomic mass of the Ca atom compared to the other elements. Moreover, desorption temperature can be calculated using the equation presented in [35]:
T des = Δ H Δ S
where ΔH represents the computed formation enthalpy, and ΔS stands for the entropy variation during the dehydrogenation reaction, which is roughly 130.7 J/(mol·K). As shown in Table 2, the desorption temperatures of CaAl2H2, SrAl2H2, ScAl2H2, and YAl2H2 are 1076 K, 1326 K, 349 K, and 983 K, respectively. The hydrogen desorption temperature of ScAl2H2 lies well within the acceptable range (233 K–333 K) specified by the DOE [36]. By combining the data from Table 1 and Table 2, it can be observed that for XAl2H2 materials, there is no direct correlation between the Al-H bond distance and hydrogen desorption temperatures. This observation is consistent with findings reported in previous studies [16,37]. In contrast, the desorption temperatures of the other compounds significantly exceed the upper limit of 333 K, indicating that CaAl2H2, SrAl2H2, and YAl2H2 require higher temperatures for hydrogen release. This may adversely impact the efficiency of hydrogen storage and release processes.

3.2. Electronic Properties

The electronic structure of materials is crucial in determining hydrogen adsorption energy, as the quantum-level charge distribution directly influences the binding mechanisms between hydrogen species and storage substrates.
The calculated band structures of the XAl2H2 compounds are presented in Figure 3. The horizontal axis denotes the high-symmetry points in the Brillouin zone, while the vertical axis corresponds to the energy of electron states. Overall, certain bands intersect the Fermi level, which is denoted by the red dashed line. This intersection implies that there is no energy gap for electrons in the valence band to overcome when transitioning to the conduction band. Consequently, all XAl2H2 compounds exhibit metallic characteristics, a feature that is highly conducive to improving hydrogen storage performance. Specifically, this metallic behavior is expected to significantly enhance hydrogen diffusion kinetics by promoting electron mobility, thereby reducing the activation energy barriers for interstitial hydrogen migration within the crystalline lattice. From Figure 3b, it is evident that the calculated band structure of SrAl2H2 exhibits remarkable consistency with the results reported in previous studies [22]. Additionally, the band structure distribution of CaAl2H2 depicted in Figure 3a exhibits a high degree of similarity to that of SrAl2H2, with the trend characteristics of each energy band along the high-symmetry points being essentially consistent. It can be clearly observed in Figure 3a,b that for both compounds mentioned above, the highest point of the valence band is situated at the A point, where the valence band intersects the Fermi level. Furthermore, the lowest point of the conduction band is positioned at the M point, and the conduction band also crosses the Fermi level in close proximity to this point. However, in contrast to the band structures of the first two compounds, the energy bands of ScAl2H2 and YAl2H2 exhibit extensive and complex overlaps near the Fermi level, as illustrated in Figure 3c,d. These overlaps originate from the orbital contributions of the main constituent elements. For instance, Ca in CaAl2H2, with a valence electron configuration of 4s2, predominantly contributes electrons from the 4s orbital, while Al (3s23p1) and H contribute electrons from their respective 3s/3p and 1s orbitals. This leads to a band structure primarily composed of s-like and p-like bands. In contrast, Sc in ScAl2H2, with an electron configuration of 3d14s2, not only contributes electrons from the 4s orbital but also introduces significant contributions from the 3d orbital, resulting in additional d-like bands. Consequently, the band structure of ScAl2H2 becomes more intricate, with numerous bands crossing the Fermi level. Similarly, since Y in YAl2H2 has an electron configuration of 4d15s2, its energy bands resemble those of ScAl2H2.
Figure 4 displays the total densities of states (TDOS) and partial densities of states (PDOS) projected onto the s, p, and d orbitals of X atoms (X = Ca, Sr, Sc, and Y); the s, p, and d orbitals of Al; and the s orbitals of H in the series of compounds. As observed in Figure 4a–c, the non-zero DOS at the Fermi level provides conclusive evidence of metallic characteristics for these materials, which is consistent with the band structure analysis results. Firstly, the DOS distribution for SrAl2H2 closely aligns with those reported in references [20,21,22], thereby validating the reliability of the calculated results. As shown in Figure 4, the X (X = Ca, Sr, Sc, Y) atom predominantly contributes via its d orbitals, whereas the Al atom primarily contributes through its s and p orbitals. Notably, the contribution from the d orbitals of Al is relatively minor and predominantly located above the Fermi level. Secondly, as illustrated in Figure 4a,b, the density of states (DOSs) for CaAl2H2 and SrAl2H2 exhibits a high degree of similarity due to their belonging to the same group of elements. For CaAl2H2 and SrAl2H2, the DOS near the Fermi level is predominantly attributed to the 3p orbitals of Al atoms. It is noteworthy that CaAl2H2 displays an apparent pseudo-band gap nearby the Fermi level, suggesting that, similar to SrAl2H2, its conductivity is relatively weak. As depicted in Figure 4a, the d orbitals of the Ca atom are mainly distributed above the Fermi level, while the s and p orbitals of the Al atom and the s orbital of the H atom are predominantly situated below the Fermi level. In the vicinity of the −8 to −2 eV region, multiple peaks of equivalent energy are observed in the s and p orbitals of Al and the s orbital of H, indicating significant hybridization among these orbitals and the formation of an Al-H covalent bond. Such hybridization also occurs in SrAl2H2, as shown in Figure 4b. Thirdly, the DOS of ScAl2H2 shows similarities with those of CaAl2H2 and SrAl2H2, particularly in the distribution of atomic states across different energy regions and in the hybridization between Al s/p orbitals and H s orbitals. However, a notable distinction exists in the DOS of ScAl2H2. In detail, while the Fermi level in CaAl2H2 and SrAl2H2 is predominantly influenced by Al 3p orbitals, the DOS near the Fermi level in ScAl2H2 arises from both Al 3p and significant Sc 3d orbital contributions, as illustrated in Figure 4c. Although other orbitals, such as H s and Al s/d, also contribute, their influence is relatively minor compared to the combined effect of Al 3p and Sc 3d orbitals. The presence of Sc 3d orbitals introduces additional electronic states near the Fermi level, resulting in an increased DOS occupancy and a shift of the pseudo-band gap to lower energies. This leads to enhanced conductivity in ScAl2H2 compared to CaAl2H2 and SrAl2H2. The enhanced DOS occupation at the Fermi level also accounts for the complex band crossings observed near the Fermi level in Figure 3c for ScAl2H2. In the case of YAl2H2, as shown in Figure 4d, although the pseudo-band gap becomes less defined, similar changes are observed in the DOS. Specifically, the contribution of Y 4d orbitals at the Fermi level increases, and the pseudo-band gap shifts toward lower energy levels, which also indicates a trend of improved conductivity relative to CaAl2H2 and SrAl2H2. Additionally, the hybridization between the s and p orbitals of Al and the s orbitals of H in compound YAl2H2 is significantly reduced. This phenomenon correlates with the increased interatomic distance dAl-H observed in structural analysis and the consequent weakening of interactions between Al and H atoms. However, as previously noted, there is no direct correlation between the Al-H bond and the hydrogen desorption temperature in the systems examined nor in certain other multicomponent aluminum hydrides [16,37]. Consequently, although the Al-H hybridization weakens in YAl2H2, its influence on the hydrogen desorption temperature is not a determining factor.
Furthermore, the electron density distributions of XAl2H2 on the (1 1 0) plane are calculated, as shown in Figure 5. These charge density maps clearly reveal significant charge accumulation between adjacent Al and H atoms, as well as between Al-Al pairs, thereby confirming the presence of covalent bonding interactions. Moreover, our results indicate that all X atoms in each compound exhibit ionic bonding characteristics. Thus, a combination of covalent and ionic bonds in these compounds is revealed. Notably, for YAl2H2, the charge accumulation between Al and H atoms is substantially reduced compared to CaAl2H2, SrAl2H2, and ScAl2H2, indicating weaker Al-H covalent bonds in YAl2H2. This observation is consistent with the conclusions drawn from the DOS analysis.

3.3. Mechanical Properties

The fundamental mechanical properties of materials play a critical role in both material design and applications. Through computational characterization of the XAl2H2 hydrides (X = Ca, Sr, Sc, and Y), the six independent single-crystalline elastic constants (C11, C12, C13, C33, C44, and C66) are determined as presented in Table 3. Although no direct experimental or theoretical reference data are currently available, the reliability of the present computational results is validated by incorporating theoretically calculated values for SrAlSiH and CaAlSiH [38]. These compounds possess identical crystallographic structures and are chemically analogous in composition to the studied system, thereby providing a robust basis for comparison. Moreover, it is crucial to note that the stability of the hexagonal phase requires satisfying Born’s mechanical stability conditions [39,40]:
C 11 + 2 C 12 C 33 2 C 13 2 > 0
Upon comparing these values with the stability criteria, it is confirmed that all compounds meet the requirements for mechanical stability assessment. Furthermore, as shown in Table 3, for this series of materials, the slight difference between the values of C11 and C33 indicates that the compressive resistance along the a-axis and c-axis in these compounds is nearly equivalent. However, a notable difference is observed: for YAl2H2, C11 is considerably larger than C33, indicating that the material is more compressible along the c-axis direction compared to the a-axis. Additionally, the relatively low values of C12 and C13 in these compounds imply that when pressure is applied along the a-axis, the materials tend to exhibit shear deformation along the b-axis and c-axis of the crystal. Moreover, for the compounds CaAl2H2, SrAl2H2, and ScAl2H2, the elastic constant C44 is smaller than C66, indicating that shear deformation is more easily achieved in the (0 0 1) plane compared to the (1 0 0) plane. In contrast, the compound YAl2H2 shows an opposite trend, with C44 exceeding C66, suggesting that shear deformation is more favorably facilitated in the (1 0 0) plane than in the (0 0 1) plane.
In addition, the elastic moduli of the polycrystalline XAl2H2 (X = Ca, Sr, Sc, and Y) compounds, including bulk modulus (B) and shear modulus (G), are calculated using the Voigt–Reuss–Hill (VRH) approximations [41]. Young’s modulus (E) and Poisson’s ratio (ν) are mathematically expressed as follows [39]:
E = 9 B G 3 B + G
ν = 3 B 2 G 2 ( 3 B + G )
From Table 4, it can be seen clearly that the computed bulk moduli for CaAl2H2 (55.35 GPa), SrAl2H2 (52.44 GPa), ScAl2H2 (75.47 GPa), and YAl2H2 (79.15 GPa) exhibit excellent agreement with the values obtained from the B-M equation of state fitting, which are 55.16 GPa for CaAl2H2, 51.64 GPa for SrAl2H2, 76.01 GPa for ScAl2H2, and 74.82 GPa for YAl2H2, as presented in Table 2. This consistency confirms the reliability of the calculated elastic constants. Our computational results reveal that in the studied hexagonal crystals XAl2H2 (X = Ca, Sr, Sc, and Y), the relatively smaller lattice parameter ratio c/a serves as a crucial factor leading to an enhanced bulk modulus. The optimized c/a ratios follow the order SrAl2H2 (1.05) > CaAl2H2 (1.03) > ScAl2H2 (0.99) > YAl2H2 (0.95), while the corresponding calculated bulk moduli display an inverse progression: SrAl2H2 (52.44 GPa) < CaAl2H2 (55.35 GPa) < ScAl2H2 (75.47 GPa) < YAl2H2 (79.15 GPa). The same phenomenon has also been observed in similar systems, such as the compounds XAlSiH (X = Ca, Sr, Ba) reported in Reference [38] and XGaSiH (X = Ca, Sr, Ba) described in Reference [18]. Furthermore, a material possessing both a high shear modulus and Young’s modulus exhibits excellent resistance to deformation and superior stiffness. Among the studied series, it is observed that the shear modulus (G = 38.53 GPa) and Young’s modulus (E = 98.78 GPa) of ScAl2H2 are slightly lower than those of the other three compounds, indicating relatively weaker resistance to deformation and reduced stiffness. However, as shown in Table 4, the shear modulus and Young’s modulus values for other aluminum-based hydrides are also presented. A comparative analysis reveals that, although the mechanical moduli of ScAl2H2 are slightly lower than those of the other three compounds examined in this study, they still exhibit overall superior performance compared to other aluminum-based hydrides such as Rb2AlTlH6 [42] and NaAlH3 [15].
Pugh’s mechanical stability criterion [43] employs the bulk-to-shear modulus ratio (B/G > 1.75 for ductility) to classify material behavior. Our calculations reveal distinct patterns: CaAl2H2 (B/G = 0.99) and SrAl2H2 (B/G = 1.25) show intrinsic brittleness, whereas ScAl2H2 (B/G = 2.04) and YAl2H2 (B/G = 1.79) exhibit superior ductility, as shown in Table 4. Complementary analysis through Poisson’s ratio confirms this contrast in mechanical properties: values below the 0.26 threshold [44] for CaAl2H2 (ν = 0.176) and SrAl2H2 (ν = 0.152) indicate limited plastic deformation capacity, while ScAl2H2 (ν = 0.281) and YAl2H2 (ν = 0.265) surpass this critical value, demonstrating enhanced ductile behavior that aligns consistently with B/G ratio predictions. Moreover, Poisson’s ratio can serve as a criterion for distinguishing between the characteristics of ionic and covalent bonds in compounds. Specifically, a Poisson’s ratio approaching 1 is indicative of predominantly ionic bonding, whereas a value of approximately 0.25 suggests covalent bonding [45]. The calculated results reveal that covalent bonding dominates in CaAl2H2 and SrAl2H2 compounds, while ionic bonding is predominant in ScAl2H2 and YAl2H2 compounds. Furthermore, the elastic anisotropic properties of materials are quantitatively assessed using the anisotropy index A, which is determined through the following equation [46]:
A = 2 C 44 C 11 C 12
A value of A = 1 indicates perfect isotropy, whereas deviations from unity signify increasing anisotropy. The computed A values for CaAl2H2, SrAl2H2, ScAl2H2, and YAl2H2 are 0.775, 0.152, 0.281, and 0.265, respectively, suggesting pronounced anisotropic characteristics in these materials. To more directly and thoroughly demonstrate the characteristics of elastic anisotropy, Figure 6 presents two-dimensional graphs that show the directional dependence of the bulk modulus, shear modulus, Young’s modulus, and Poisson’s ratio. As shown in Figure 6, apart from the (0 0 1) plane exhibiting relatively isotropic behavior as indicated by its nearly circular plot, the (0 1 0) plane and (1 0 0) plane both display varying degrees of anisotropy, with their plots deviating from circular symmetry to different extents.
In addition, mass density ρ, average sound velocity vm, longitudinal wave sound velocity vl, transverse wave sound velocity vt, and Debye temperature θD are shown in Table 3 and can be estimated using the following relationships [39]:
θ D = h k B 3 n 4 π N A ρ M 1 3 v m
v m = 1 3 2 v t 3 + 1 v l 3 1 3
v l = B + 4 3 G / ρ
v t = G / ρ
Here, M, NA, ρ, h, k, and n represent the molecular weight, Avogadro’s number, the density of the material, Planck’s constant, Boltzmann’s constant, and the number of atoms in the unit cell, respectively.
The Debye temperature can be regarded as an indicator of the stiffness or rigidity of a solid lattice. The Debye temperature of CaAl2H2 is calculated to be 620.6 K, significantly higher than that of the other three compounds. This suggests that CaAl2H2 exhibits superior crystalline lattice rigidity.

3.4. Lattice Dynamical and Thermodynamic Properties

The calculated phonon spectra and projected phonon density of states (PPDOSs) are illustrated in Figure 7a–d, providing valuable insights into the lattice dynamical properties and enabling a deeper exploration of the thermodynamic properties. The absence of imaginary frequencies in the phonon dispersion curves of CaAl2H2 and SrAl2H2 confirms their dynamical stability, whereas the presence of imaginary frequencies in those of ScAl2H2 and YAl2H2 suggests their potential dynamical instability in the structure. For SrAl2H2, the phonon dispersion curves are found to be in good agreement with those reported in previous theoretical studies [21,22]. As shown in Figure 7b, the phonon spectrum of SrAl2H2 is distinctly distributed across three frequency regions: the low-frequency region (0–348.4 cm−1), the intermediate-frequency region (540.8–734.5 cm−1), and the high-frequency region (1324.4–1412.7 cm−1). Most of the phonon DOS in the low-frequency region originates from Al atoms, which predominantly contribute to the range of approximately 100–350 cm−1, while the remaining contributions are from Sr atoms that are primarily responsible for the range of about 0–100 cm−1. In both the intermediate-frequency and high-frequency regions, the contributions are predominantly from H atoms, with minor contributions from Al atoms. Crystal structures that are similar tend to exhibit phonon dispersion curves with comparable characteristics. As shown in Figure 7a, the phonon dispersion curves for CaAl2H2 can also be divided into three frequency zones: the low-frequency region (0–376.3 cm−1), the intermediate-frequency region (414.2–670.1 cm−1), and the high-frequency region (1388.2–1473.2 cm−1), respectively. It is noteworthy that for both CaAl2H2 and SrAl2H2, in the high-frequency region, the primary contributions predominantly originate from the vibrations between Al and H atoms. This can be primarily attributed to the stronger covalent character of the Al-H bonds. It is observed that in the phonon DOS for both compounds, the contribution of Ca or Sr atoms is only distributed within a narrow frequency range of 0–100 cm−1. This indicates that the vibrations of Ca or Sr atoms are more localized compared to those of other atoms. Consequently, this further suggests that the bonding strength of X atoms (e.g., Ca or Sr) is relatively weak, providing only a limited restoring force to sustain interatomic resonance. For ScAl2H2, as clearly shown in Figure 7c, the phonon dispersion curves exhibit significant imaginary modes throughout the entire Brillouin zone, with the contributions to the phonon DOS associated with these imaginary modes predominantly originating from Sc atoms. For YAl2H2, as illustrated in Figure 7d, the phonon dispersion curves display substantial imaginary modes across most of the Brillouin zone, except along the high-symmetry directions M-K and L-H. The contributions to the phonon DOS linked to these imaginary modes are primarily attributed to Al atoms, and minor contributions come from Y and H atoms.
Additionally, the hexagonal structure of the P-3m1 space group belongs to O3d point group. Therefore, as detailed in Table 5, the characters of the irreducible representation for the XAl2H2 (X = Ca, Sr, Sc, and Y) compounds at the Γ point are determined according to factor group theory [47]. Therein, it is observed that the modes belong to single states characterized by the A representation and double-degenerate states characterized by the E representation. Among the optic modes, these degenerate states are both Raman (R)-active and infrared (IR)-active. The calculated frequencies of modes at the Γ point for CaAl2H2 and SrAl2H2 are listed in Table 5. The calculation results for SrAl2H2 are in good agreement with the available experimental values [48]. However, the data for ScAl2H2 and YAl2H2 are excluded because of the presence of significant imaginary modes within the Brillouin zone.
Furthermore, temperature plays a pivotal role in material performance, necessitating an exploration of thermodynamic property variations with temperature. The temperature-dependent behavior of Helmholtz free energy (F), internal energy (E), entropy (S), and constant-volume specific heat (Cv) is investigated from 0 K to 1000 K using the harmonic approximation, as presented in Figure 8. The computational framework relies on the following equations [39]:
F = 3 n N k B T 0 ω max ln 2 sinh ω 2 k B T g ω d ω
E = 3 n N 2 0 ω max ω coth ω 2 k B T g ω d ω
S = 3 n N k B 0 ω max ω 2 k B T coth ω 2 k B T ln 2 sinh ω 2 k B T g ω d ω
C V = 3 n N k B 0 ω max ω 2 k B T 2 csc h 2 ω 2 k B T g ω d ω
Here, kB is the Boltzmann constant, ħ is the reduced Planck constant, n is the number of atoms per unit cell, N is the number of unit cell, T is the temperature, ω is the phonon frequency, and ωmax is the maximum phonon frequency, respectively. In addition, the normalized phonon DOS g(ω) satisfies the normalization condition [39]:
0 ω max g ω d ω = 1
The thermodynamic data of ScAl2H2 and YAl2H2 are omitted in Figure 8, as the formulas used to calculate these data rely on the phonon DOS, and the phonon spectra of these two compounds exhibit imaginary frequencies. It is worth noting that the thermodynamic data for CaAl2H2 and SrAl2H2 are very close, with their curves nearly overlapping in Figure 8, attributable to their similar chemical compositions. As shown in Figure 8, the computed free energy F is found to decrease steadily as temperature rises. Meanwhile, the internal energy E increases linearly at temperatures above room temperature. At absolute zero, the internal energy E equals the free energy F, a value referred to as the zero-point energy E0, which can be calculated via the equation below [39]:
F 0 = E 0 = 3 n N 0 ω max ω 2 g ω d ω
The zero-point energies E0 for CaAl2H2 and SrAl2H2 are calculated as 40.49 kJ/mol and 41.51 kJ/mol, respectively. Notably, the entropy S values exhibit a monotonic increase with temperature. At 300 K, the entropies of CaAl2H2 and SrAl2H2 are 104.4 J/(mol∙K) and 105.1 J/(mol∙K), respectively. For the constant-volume heat capacity Cv, both the compounds show a temperature-dependent increase, following Debye’s T3 law in the low-temperature regime. At 300 K, the Cv values for CaAl2H2 and SrAl2H2 are determined to be 89.0 J/(mol∙K) and 86.4 J/(mol∙K), respectively. As temperature continues to rise, Cv gradually approaches the classical limit predicted by the Dulong–Petit law: Cv = 3nR = 3 × 5 × 8.31 = 124.65 J/(mol∙K).

4. Conclusions

In summary, a comprehensive investigation was carried out on the structural, hydrogen storage, electronic, mechanical, lattice dynamics, and thermodynamic properties of the Zintl phase hydrides XAl2H2 (X = Ca, Sr, Sc, and Y) using the full-potential linearized augmented plane wave (FP-LAPW) method within the framework of density functional theory (DFT). The optimized structural parameters (a, c, and c/a) for SrAl2H2 are in good agreement with the experimental data and other theoretical results, thereby validating the reliability of the structural calculations for CaAl2H2, ScAl2H2, and YAl2H2 reliability. The calculated formation enthalpies of −1.46, −1.80, −0.47, and −1.33 eV/f.u. for CaAl2H2, SrAl2H2, ScAl2H2, and YAl2H2, respectively, indicate their thermodynamic stability due to the negative values. The gravimetric hydrogen storage capacity of CaAl2H2 is 1.41%, which is the highest among the studied systems. In comparison, the capacities of SrAl2H2, ScAl2H2, and YAl2H2 are 0.94%, 1.34%, and 0.93%, respectively. The hydrogen desorption temperatures of CaAl2H2, SrAl2H2, ScAl2H2, and YAl2H2 are 1076 K, 1326 K, 349 K, and 983 K, respectively. Electronic structure analysis reveals that all these compounds exhibit metallic characteristics and feature a combination of covalent and ionic bonding. CaAl2H2 and SrAl2H2 exhibit an apparent pseudo-band gap near the Fermi level, suggesting relatively weak conductivity compared to ScAl2H2 and YAl2H2. The mechanical stability of the materials is confirmed by satisfying Born’s criteria. ScAl2H2 and YAl2H2 exhibit good ductility, whereas CaAl2H2 and SrAl2H2 display intrinsic brittleness. However, all these materials show anisotropic characteristics to varying degrees. The absence of imaginary frequencies in the phonon dispersion curves of CaAl2H2 and SrAl2H2 confirms their dynamical stability, whereas the presence of imaginary frequencies in the phonon dispersion curves of ScAl2H2 and YAl2H2 indicates potential dynamical instability in these structures. Furthermore, for CaAl2H2 and SrAl2H2, we calculated the phonon frequencies at the center of the first Brillouin zone, identifying Raman-active and infrared-active vibrational modes. In addition, we provide temperature-dependent thermodynamic properties, including Helmholtz free energy, internal energy, entropy, and heat capacity. Overall, our investigation into hexagonal XAl2H2 compounds reveals that CaAl2H2 possesses promising properties as a hydrogen storage material. This study investigated the fundamental physical properties of the XAl2H2 system for solid-state hydrogen storage, providing a theoretical foundation for understanding the material’s characteristics and facilitating further research in this field.

Author Contributions

Conceptualization, software, writing—original draft preparation, resources, and funding acquisition, Y.G.; writing—review and editing, validation, and methodology, R.G.; data curation and visualization, L.W.; investigation, Y.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Science and Technology Innovation Project of Higher Education in Shanxi Province (grant number 2022L097) and the Natural Science Foundation of Gansu Province (grant number 24JRRA961).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

Youyu Zhang was employed by Jiuquan Iron and Steel (Group) Corporation. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Ouyang, L.; Chen, K.; Jiang, J.; Yang, X.; Zhu, M. Hydrogen storage in light-metal based systems: A review. J. Alloys Compd. 2020, 829, 154597. [Google Scholar] [CrossRef]
  2. Murphy, R. What is undermining climate change mitigation? how fossil-fuelled practices challenge low-carbon transitions. Energy Res. Soc. Sci. 2024, 108, 103390. [Google Scholar] [CrossRef]
  3. Li, R.; Hu, F.; Xia, T.; Li, Y.; Zhao, X.; Zhu, J. Progress in the application of first principles to hydrogen storage materials. Int. J. Hydrogen Energy 2024, 56, 1079–1091. [Google Scholar] [CrossRef]
  4. Schlapbach, L.; Züttel, A. Hydrogen-storage materials for mobile applications. Nature 2001, 414, 353–358. [Google Scholar] [CrossRef]
  5. Paul, A.R.; Mehla, S.; Bhargava, S. Intermetallic compounds for hydrogen storage: Current status and future perspectives. Small 2024, 20, 2408889. [Google Scholar] [CrossRef]
  6. Xu, Y.; Zhou, Y.; Li, C.; Dong, S.; Liu, H.; Yang, W.; Li, Y.; Jiang, H.; Ding, Z.; Li, H.; et al. Unraveling the potential of solid-state hydrogen storage materials: Insights from first principle calculations. Fuel 2024, 373, 132340. [Google Scholar] [CrossRef]
  7. Xu, X.; Dong, Y.; Hu, Q.; Si, N.; Zhang, C. Electrochemical Hydrogen Storage Materials: State-of-the-Art and Future Perspectives. Energy Fuels 2024, 38, 7579–7613. [Google Scholar] [CrossRef]
  8. Salman, M.; Rambhujun, N.; Pratthana, C.; Lai, Q.; Sapkota, P.; Aguey-Zinsou, K.F. Chapter 12-Solid-state hydrogen storage as a future renewable energy technology. In Micro and Nano Technologies; Elsevier: Amsterdam, The Netherlands, 2021; pp. 263–287. [Google Scholar] [CrossRef]
  9. Ke, X.; Kuwabara, A.; Tanaka, I. Cubic and orthorhombic structures of aluminum hydride AlH3 predicted by a first-principles study. Phys. Rev. B 2005, 71, 184107. [Google Scholar] [CrossRef]
  10. Mei, Z.; Zhao, F.; Xu, S.; Ju, X. Theoretical investigations on the phase transition of pure and li-doped AlH3. RSC Adv. 2017, 7, 42024–42029. [Google Scholar] [CrossRef]
  11. Resan, M.; Hampton, M.D.; Lomness, J.K.; Slattery, D.K. Effect of tixaly catalysts on hydrogen storage properties of LiAlH4 and LiAlH4. Int. J. Hydrogen Energy 2005, 30, 1417–1421. [Google Scholar] [CrossRef]
  12. Ianni, E.; Sofianos, M.V.; Rowles, M.R.; Sheppard, D.A.; Humphries, T.D.; Buckley, C.E. Synthesis of NaAlH4/Al composites and their applications in hydrogen storage. Int. J. Hydrogen Energy 2018, 43, 17309–17317. [Google Scholar] [CrossRef]
  13. Weidenthaler, C. Crystal structure evolution of complex metal aluminum hydrides upon hydrogen release. J. Energy Chem. 2020, 42, 133–143. [Google Scholar] [CrossRef]
  14. Dragojlović, M.; Radaković, J.; Batalović, K. DFT study of crystal structure and electronic properties of metal-doped AlH3 polymorphs. Int. J. Hydrogen Energy 2022, 47, 6142–6153. [Google Scholar] [CrossRef]
  15. Xu, N.; Song, R.; Zhang, J.; Chen, Y.; Chen, S.; Li, S.; Jiang, Z.; Zhang, W. First-principles study on hydrogen storage properties of the new hydride perovskite XAlH3 (X=Na, K). Int. J. Hydrogen Energy 2024, 60, 434–440. [Google Scholar] [CrossRef]
  16. Umer, M.; Murtaza, G.; Ahmad, N.; Ayyaz, A.; Raza, H.H.; Usman, A.; Liaqat, A.; Manoharadas, S. First principles investigation of structural, mechanical, thermodynamic, and electronic properties of Al-based perovskites XAlH3 (X=K, Rb, Cs) for hydrogen storage. Int. J. Hydrogen Energy 2024, 61, 820–830. [Google Scholar] [CrossRef]
  17. Gingl, F.; Vogt, T.; Akiba, E. Trigonal SrAl2H2: The first Zintl phase hydride. J. Alloys Compd. 2000, 306, 127–132. [Google Scholar] [CrossRef]
  18. Ammi, H.; Charifi, Z.; Baaziz, H.; Ghellab, T.; Bouhdjer, L.; Adalla, S.; Ocak, H.Y.; Uğur, Ş.; Uğur, G. Investigation on the hydrogen storage properties, electronic, elastic, and thermodynamic of Zintl Phase Hydrides XGaSiH (X = sr, ca, ba). Int. J. Hydrogen Energy 2024, 87, 966–984. [Google Scholar] [CrossRef]
  19. Tang, F.; Chen, Y.; Yin, X.; Zhao, W.; Zhang, L.; Han, Z.; Zheng, R.; Zhang, X.; Fang, Y. Colossal magnetoresistance and Fermi surface topology in the layered Zintl-phase compound YbAl2Si2. Phys. Rev. B 2024, 110, 174408. [Google Scholar] [CrossRef]
  20. Björling, T.; Noréus, D.; Häussermann, U. Polyanionic hydrides from polar intermetallics AeE2 (Ae = Ca, Sr, Ba; E = Al, Ga, In). J. Am. Chem. Soc. 2006, 128, 817–824. [Google Scholar] [CrossRef] [PubMed]
  21. Lee, M.H.; Sankey, O.F.; Björling, T.; Moser, D.; Noréus, D.; Parker, S.T.; Häussermann, U. Vibrational properties of polyanionic hydrides SrAl2H2 and SrAlSiH: New insights into Al-H bonding interactions. Inorg. Chem. 2007, 46, 6987–6991. [Google Scholar] [CrossRef]
  22. Subedi, A.; Singh, D.J. Bonding in Zintl phase hydrides: Density functional calculations for SrAlSiH, SrAl2H2, SrGa2H2 and BaGa2H2. Phys. Rev. B 2008, 78, 045106. [Google Scholar] [CrossRef]
  23. Moser, D.; Häussermann, U.; Utsumi, T.; Björling, T.; Noréus, D. A series of BaAl2−xSixH2−x (0.4<x<1.6) hydrides with compositions and structures in between BaSi2 and BaAl2H2. J. Alloys Compd. 2010, 505, 1–5. [Google Scholar] [CrossRef]
  24. Iwaszczuk, J.; Baj, A.; Walejko, P. Sialic acids-structure and properties. Prospect. Pharm. Sci. 2024, 22, 31–38. [Google Scholar] [CrossRef]
  25. Hohenberg, P.; Kohn, W. Inhomogeneous electron gas. Phys. Rev. 1964, 136, B864. [Google Scholar] [CrossRef]
  26. Blaha, P.; Schwarz, K.; Tran, F.; Laskowski, R.; Madsen, G.; Marks, L. WIEN2k: An APW+lo program for calculating the properties of solids. J. Chem. Phys. 2020, 152, 074101. [Google Scholar] [CrossRef]
  27. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef]
  28. Gay, D.M. Algorithm 611: Subroutines for unconstrained minimization using a model/trust-region approach. ACM Trans. Math. Softw. 1983, 9, 503–524. [Google Scholar] [CrossRef]
  29. Jamal, M.; Bilal, M.; Ahmad, I.; Jalali-Asadabadi, S. IRelast package. J. Alloys Compd. 2018, 753, 569–579. [Google Scholar] [CrossRef]
  30. Yalameha, S.; Nourbakhsh, Z.; Vashaee, D. ElATools: A tool for analyzing anisotropic elastic properties of the 2D and 3D materials. Comput. Phys. Commun. 2022, 271, 108195. [Google Scholar] [CrossRef]
  31. Parlinski, K.; Li, Z.Q.; Kawazoe, Y. First-principles determination of the soft mode in cubic ZrO2. Phys. Rev. Lett. 1997, 78, 4063–4066. [Google Scholar] [CrossRef]
  32. Togo, A.; Tanaka, I. First principles phonon calculations in materials science. Scr. Mater. 2015, 108, 1–5. [Google Scholar] [CrossRef]
  33. Song, R.; Chen, Y.; Chen, S.; Xu, N.; Zhang, W. First-principles to explore the hydrogen storage properties of XPtH3 (X=Li, Na, K, Rb) perovskite type hydrides. Int. J. Hydrogen Energy 2024, 57, 949–957. [Google Scholar] [CrossRef]
  34. Surucu, G.; Gencer, A.; Candan, A.; Gullu, H.H.; Isik, M. CaXH3 (X= Mn, Fe, Co) perovskite-type hydrides for hydrogen storage applications. Int. J. Energy Res. 2019, 44, 2345. [Google Scholar] [CrossRef]
  35. Masood, M.K.; Elaggoune, W.; Chaoui, K.; Bibi, S.; Khan, M.I.; Usman, M.; Alothman, A.A.; Rehman, J. The investigation of the physical properties of nickel-based hydrides and the prediction of their suitability for hydrogen storage application. Mat. Sci. Semicon. Proc. 2024, 186, 109094. [Google Scholar] [CrossRef]
  36. Murtaza, H.; Ain, Q.; Alshgari, R.A.; Akhter, T.; Munir, J. Exploring hydrogen storage attributes of alkali metal XNH6 (X=Li, Na, K) perovskite hydrides using DFT calculations. J. Power Sources 2025, 641, 236788. [Google Scholar] [CrossRef]
  37. 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. [Google Scholar] [CrossRef]
  38. Ammi, H.; Charifi, Z.; Baaziz, H.; Ghellab, T.; Bouhdjer, L.; Adalla, S. Electronic, elastic, and thermodynamic properties of complex hydrides XAlSiH (X = Sr, Ca, and Ba) intended for hydrogen storage: An ab-initio study. Phys. Scr. 2024, 99, 0659a2. [Google Scholar] [CrossRef]
  39. Guo, Y.; Gao, T.; Li, S. The structural, electronic, mechanical, lattice dynamics and thermodynamic properties of Rh5B4: First-principles calculations. Mod. Phys. Lett. B 2017, 31, 1750245. [Google Scholar] [CrossRef]
  40. Mouhat, F.; Coudert, F.X. Necessary and sufficient elastic stability conditions in various crystal systems. Phys. Rev. B 2014, 90, 224104. [Google Scholar] [CrossRef]
  41. Hill, R. The Elastic behavior of a crystalline aggregate. Proc. Phys. Soc. 1952, 65, 349. [Google Scholar] [CrossRef]
  42. Murtaza, H.; Ain, Q.; Jbara, A.S.; Munir, J.; Aldwayyan, A.S.; Ghaithan, H.M.; Ahmed, A.A.A.; Qaid, S.M. The prediction of hydrogen storage capacity and solar water splitting applications of Rb2AlXH6 (X= In, Tl) perovskite halides: A DFT study. J. Phys. Chem. Solids 2025, 198, 112427. [Google Scholar] [CrossRef]
  43. Pugh, S.F. Relations between the elastic moduli and the plastic properties of polycrystalline pure metals. Philos. Mag. 1954, 45, 823–843. [Google Scholar] [CrossRef]
  44. Connétable, D.; Thomsa, O. First-principles study of the structural, electronic, vibrational, and elastic properties of orthorhombic NiSi. Phys. Rev. B 2009, 79, 094101. [Google Scholar] [CrossRef]
  45. Zheng, W.; Liu, F.; Lu, Y.; Liu, Z.; Liu, W.; Liu, Q. First-principles calculations of the structural, mechanical, electronic, and optical properties of BaX2 (X=O, S, Se and Te) compounds. Mat. Sci. Semicon. Proc. 2022, 147, 106755. [Google Scholar] [CrossRef]
  46. Ranganathan, S.I.; Ostoja-Starzewski, M. Universal Elastic Anisotropy Index. Phys. Rev. Lett. 2008, 101, 055504. [Google Scholar] [CrossRef]
  47. Hahn, T. Space Group Symmetry; Springer: Berlin, Germany, 2005; Volume A. [Google Scholar]
  48. Vajeeston, P.; Fjellvg, H. Revised electronic structure, Raman and IR studies of AB2H2 and ABCH (A = Sr, Ba; B = Al, Ga; C = Si, Ge) phases. RSC Adv. 2014, 4, 22–31. [Google Scholar] [CrossRef]
Figure 1. (Color online.) The crystal structure of XAl2H2 (X = Ca, Sr, Sc, Y): (a) the side view of 1 × 1 × 1 unit cell; (b) the front view of 2 × 2 × 2 supercell; (c) the polyhedral view of 2 × 2 × 2 supercell.
Figure 1. (Color online.) The crystal structure of XAl2H2 (X = Ca, Sr, Sc, Y): (a) the side view of 1 × 1 × 1 unit cell; (b) the front view of 2 × 2 × 2 supercell; (c) the polyhedral view of 2 × 2 × 2 supercell.
Materials 18 03521 g001
Figure 2. (Color online.) The total energy versus volume curve of (a) CaAl2H2, (b) SrAl2H2, (c) ScAl2H2 and (d) YAl2H2.
Figure 2. (Color online.) The total energy versus volume curve of (a) CaAl2H2, (b) SrAl2H2, (c) ScAl2H2 and (d) YAl2H2.
Materials 18 03521 g002
Figure 3. (Color online.) Band structures of (a) CaAl2H2, (b) SrAl2H2, (c) ScAl2H2, and (d) YAl2H2.
Figure 3. (Color online.) Band structures of (a) CaAl2H2, (b) SrAl2H2, (c) ScAl2H2, and (d) YAl2H2.
Materials 18 03521 g003
Figure 4. (Color online.) The calculated total and partial DOS of (a) CaAl2H2, (b) SrAl2H2, (c) ScAl2H2, and (d) YAl2H2. The Fermi levels have been set to 0 eV and marked by dot lines.
Figure 4. (Color online.) The calculated total and partial DOS of (a) CaAl2H2, (b) SrAl2H2, (c) ScAl2H2, and (d) YAl2H2. The Fermi levels have been set to 0 eV and marked by dot lines.
Materials 18 03521 g004
Figure 5. (Color online.) The distribution of charge density of (a) CaAl2H2, (b) SrAl2H2, (c) ScAl2H2, and (d) YAl2H2 in the (1 1 0) plane.
Figure 5. (Color online.) The distribution of charge density of (a) CaAl2H2, (b) SrAl2H2, (c) ScAl2H2, and (d) YAl2H2 in the (1 1 0) plane.
Materials 18 03521 g005
Figure 6. (Color online.) Two-dimensional plots of calculated anisotropy mechanical parameters (bulk, shear, Young’s modulus and Poisson’s ratio) at (0 0 1), (0 1 0) and (1 0 0) plane for XAl2H2 (X = Ca, Sr, Sc, Y).
Figure 6. (Color online.) Two-dimensional plots of calculated anisotropy mechanical parameters (bulk, shear, Young’s modulus and Poisson’s ratio) at (0 0 1), (0 1 0) and (1 0 0) plane for XAl2H2 (X = Ca, Sr, Sc, Y).
Materials 18 03521 g006
Figure 7. (Color online.) Calculated phonon dispersion and partial phonon density of states of (a) CaAl2H2, (b) SrAl2H2, (c) ScAl2H2, and (d) YAl2H2.
Figure 7. (Color online.) Calculated phonon dispersion and partial phonon density of states of (a) CaAl2H2, (b) SrAl2H2, (c) ScAl2H2, and (d) YAl2H2.
Materials 18 03521 g007
Figure 8. (Color online.) Calculated free energy, internal energy, entropy S, and the specific heat at constant volume Cv as a function of temperature for CaAl2H2 and SrAl2H2.
Figure 8. (Color online.) Calculated free energy, internal energy, entropy S, and the specific heat at constant volume Cv as a function of temperature for CaAl2H2 and SrAl2H2.
Materials 18 03521 g008
Table 1. The internal coordinates of Al and H atoms and the shortest interatomic distances of Al-H for XAl2H2 (X = Ca, Sr, Sc, Y).
Table 1. The internal coordinates of Al and H atoms and the shortest interatomic distances of Al-H for XAl2H2 (X = Ca, Sr, Sc, Y).
CompoundsAlHdAl-H (Å)Refs.
CaAl2H2(1/3, 2/3, 0.456027)(1/3, 2/3, 0.085312)1.697Present
SrAl2H2(1/3, 2/3, 0.460439)(1/3, 2/3, 0.097497)1.716Present
SrAl2H2(1/3, 2/3, 0.4589)(1/3, 2/3, 0.0976)1.71Exp. [17]
SrAl2H2(1/3, 2/3, 0.4608)(1/3, 2/3, 0.0964)1.721Theo. [22]
ScAl2H2(1/3, 2/3, 0.441807)(1/3, 2/3, 0.033534)1.720Present
YAl2H2(1/3, 2/3, 0.500094)(1/3, 2/3, 0.000166)2.067Present
Table 2. The lattice constants (a and c), c/a ratio, unit cell volume (V0), bulk modulus (B0) and its pressure derivative B’, formation enthalpy (ΔH), gravimetric hydrogen storage capacities (Cwt%), and desorption temperature Tdes of XAl2H2 (X = Ca, Sr, Sc, Y).
Table 2. The lattice constants (a and c), c/a ratio, unit cell volume (V0), bulk modulus (B0) and its pressure derivative B’, formation enthalpy (ΔH), gravimetric hydrogen storage capacities (Cwt%), and desorption temperature Tdes of XAl2H2 (X = Ca, Sr, Sc, Y).
Compoundsacc/aV0B0B’ΔHCwt%TdesRefs.
(Å)(Å) 3)(GPa) (eV/f.u.) (K)
CaAl2H24.43284.57731.0377.9055.163.80−1.461.411076Present
SrAl2H24.53234.73951.0584.1851.643.93−1.800.941326Present
SrAl2H24.52834.72151.04 Exp. [17]
SrAl2H24.5284.7221.04 Theo. [22]
ScAl2H24.23434.21230.9965.4176.013.94−0.471.34349Present
YAl2H24.34404.13640.9567.6074.824.01−1.330.93983Present
Table 3. Calculated elastic constants (GPa) of XAl2H2 (X = Ca, Sr, Sc, Y).
Table 3. Calculated elastic constants (GPa) of XAl2H2 (X = Ca, Sr, Sc, Y).
CompoundsC11C12C13C33C44C66Refs.
CaAl2H2124.2126.3016.31131.9237.9448.96Present
SrAl2H2118.2025.1515.71122.4645.0946.52Present
ScAl2H2137.3752.4341.60133.9531.1642.47Present
YAl2H2146.9571.8446.83105.5654.4037.54Present
CaAlSiH175.623739.121223.780794.06143.117768.25125Theo. [38]
SrAlSiH163.262432.589527.253197.460146.255065.3364Theo. [38]
Table 4. Calculated bulk modulus (B), shear modulus (G), Young’s modulus (E), Pugh ratio (B/G), Frantsevich ratio (G/B), Poisson’s ratio (ν), anisotropy factor (A), density ρ (g/cm3), Debye temperature (θD), transverse elastic wave velocity (vt), longitudinal elastic wave velocity (vl), and average wave velocity (vm) of XAl2H2 (X = Ca, Sr, Sc, Y).
Table 4. Calculated bulk modulus (B), shear modulus (G), Young’s modulus (E), Pugh ratio (B/G), Frantsevich ratio (G/B), Poisson’s ratio (ν), anisotropy factor (A), density ρ (g/cm3), Debye temperature (θD), transverse elastic wave velocity (vt), longitudinal elastic wave velocity (vl), and average wave velocity (vm) of XAl2H2 (X = Ca, Sr, Sc, Y).
Present WorkOther Al-Base Hydrides
CompoundsCaAl2H2SrAl2H2ScAl2H2YAl2H2CaAlSiH [38]SrAlSiH [38]Rb2AlTlH6 [42]NaAlH3 [15]
B (GPa)55.3552.4475.4779.1565.35064.46827.443.669
G (GPa)45.7747.3838.5343.9753.09152.91918.28.702
E (GPa)107.64109.2398.78111.30125.332124.65044.624.479
B/G1.211.111.961.801.23091.21821.505.018
G/B0.830.900.510.56
ν0.1760.1520.2810.2650.1800.1770.230.407
A0.7750.9690.7341.450 0.72
ρ2.052.832.563.56
θD (K)620.6521.9546.0488.6606.615502.165208.18
vt (m/s)4727.664089.593877.553514.884756.124048.461870.3
vl (m/s)7538.606388.367035.456221.77616.096466.873160.11
vm (m/s)5206.064492.984321.333909.415239.814458.982071.09
Table 5. Vibrational assignments and phonon frequencies ω (cm−1) at zone-center (Γ point) for CaAl2H2 and SrAl2H2. The notations IR and R refer to infrared-active and Raman-active modes, respectively.
Table 5. Vibrational assignments and phonon frequencies ω (cm−1) at zone-center (Γ point) for CaAl2H2 and SrAl2H2. The notations IR and R refer to infrared-active and Raman-active modes, respectively.
The Point Group: O3d (−3m)
Γ acoustic = A 2 u E u
Γ optic = 2 A 1 g 2 A 2 u 2 E u 2 E g
ModeCaAl2H2SrAl2H2SrAl2H2 [48]
A2u (IR)132, 1397134, 1332132, 1333
Eu (IR)121, 414135, 541143, 593
A1g (R)247, 1466254, 1406267, 1412
Eg (R)376, 670348, 735377, 765
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

Guo, Y.; Guo, R.; Wan, L.; Zhang, Y. Quantum Chemical Investigation on the Material Properties of Al-Based Hydrides XAl2H2 (X = Ca, Sr, Sc, and Y) for Hydrogen Storage Applications. Materials 2025, 18, 3521. https://doi.org/10.3390/ma18153521

AMA Style

Guo Y, Guo R, Wan L, Zhang Y. Quantum Chemical Investigation on the Material Properties of Al-Based Hydrides XAl2H2 (X = Ca, Sr, Sc, and Y) for Hydrogen Storage Applications. Materials. 2025; 18(15):3521. https://doi.org/10.3390/ma18153521

Chicago/Turabian Style

Guo, Yong, Rui Guo, Lei Wan, and Youyu Zhang. 2025. "Quantum Chemical Investigation on the Material Properties of Al-Based Hydrides XAl2H2 (X = Ca, Sr, Sc, and Y) for Hydrogen Storage Applications" Materials 18, no. 15: 3521. https://doi.org/10.3390/ma18153521

APA Style

Guo, Y., Guo, R., Wan, L., & Zhang, Y. (2025). Quantum Chemical Investigation on the Material Properties of Al-Based Hydrides XAl2H2 (X = Ca, Sr, Sc, and Y) for Hydrogen Storage Applications. Materials, 18(15), 3521. https://doi.org/10.3390/ma18153521

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