Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Insight into the effect of Mg-substitution on the electronic, optoelectronic, and hydrogen storage density of NbH2 fluorite structured: a DFT study

Godwin O. Igomaha, Favour A. Nelson*b, Fadhil Faez Seadc, Musa Rundede, Ismail Hossainf and Ayi A. Ayib
aDepartment of Physics, Faculty of Physical Sciences, University of Calabar, Calabar, Nigeria
bDepartment of Pure and Industrial Chemistry, University of Calabar, Calabar, Nigeria. E-mail: azogorfavour@gmail.com
cDepartment of Dentistry, College of Dentistry, The Islamic University, Najaf, Iraq
dDepartment of Research Analytics, Saveetha Dental College and Hospitals, Saveetha Institute of Medical and Technical Sciences, Saveetha University, Chennai, India
eNational Open University of Nigeria (NOUN), Abuja, Nigeria
fDepartment of Nuclear and Renewable Energy, Ural Federal University, Yekaterinburg, 620002, Russia

Received 4th June 2025 , Accepted 14th July 2025

First published on 30th July 2025


Abstract

In the quest for multifunctional hydrogen storage materials, this study investigates the structural, electronic, and optical properties of NbH2, MgH2, and a series of Mg-substituted NbH2 compounds (Mg-NbH2, Mg2-NbH2, and Mg3-NbH2) using first-principles density functional theory (DFT) based on GGA/PBE and HSE03 methods. The motivation stems from the need to overcome the well-known limitations of MgH2, particularly its high desorption temperature and poor reversibility, by introducing Mg into the NbH2 fluorite framework. Structural optimization revealed a fluorite-type geometry, with Mg substitution inducing moderate lattice distortion and increasing unit cell volume from 97.22 to 103.45 Å3. The Mg-NbH2 system achieved a high density of 10.78 g cm−3 and exhibited a favorable hydrogen gravimetric capacity of 3.33 wt%, offering a promising trade-off between storage potential and structural stability. Electronic structure analysis confirmed metallicity across all substituted systems, while MgH2 retained a non-metallic nature. A progressive decrease in total density of states was observed from 7.0 (NbH2) to 2.0 (Mg3-NbH2), suggesting tunable electronic characteristics. Optical studies revealed that Mg-NbH2 displayed the strongest dielectric response (ε2 ≈ 85), the highest refractive index (n1 ≈ 3.2), and reduced optical losses compared to its parent compounds. Notably, it retained a high optical conductivity (∼13 S m−1) and strong absorption in the visible range, making it a potential candidate for photocatalytic and optoelectronic applications. These results demonstrate that Mg substitution into NbH2 significantly enhances its multifunctional behavior, offering a viable pathway to improve hydride-based materials for advanced hydrogen storage and light-harvesting technologies.


1. Introduction

Hydrogen energy is a cornerstone of clean energy technologies due to its high energy density, abundance, and eco-friendly nature. Metal hydrides have gained significant attention from various hydrogen storage materials due to their reversible hydrogen absorption and desorption capabilities.1 Transition metal hydrides, particularly those based on niobium (Nb), exhibit favorable thermodynamic and kinetic properties for hydrogen storage.2 Niobium dihydride (NbH2) with a fluorite crystal structure represents a promising candidate because it stores considerable amounts of hydrogen while maintaining structural integrity.3 However, one of the prevailing challenges in advancing hydrogen storage systems lies in optimizing their storage density, electronic conductivity, and optical characteristics, which influence material performance in real-world applications. In this regard, substitutional doping, especially with lightweight and electropositive elements such as magnesium (Mg), has been widely explored as a strategic approach to tailor the intrinsic properties of metal hydrides. Mg-substitution has the potential to enhance the hydrogen storage capacity, tune the electronic band structure, and improve optoelectronic performance through lattice modification and charge redistribution.4,5

Magnesium (Mg), an alkaline earth metal with atomic number 12, is known for its low atomic weight, high electropositivity, and abundance in the Earth's crust. Its favorable characteristics, including a small ionic radius (0.72 Å) and good thermal and chemical stability, make it an attractive candidate for substitutional doping in various materials.6,7 The ability of Mg2+ ions to replace other cations in crystal lattices enables the tuning of electronic, structural, optical, and chemical properties of host compounds.8 In materials research, Mg-substitution has been explored extensively across ceramics, hydrides, semiconductors, and biomaterials due to its role in enhancing functionality and performance.9,10 The influence of Mg-substitution has been well-documented in calcium phosphate-based ceramics. A study on Mg-substituted tricalcium phosphate (TCP) demonstrated that partial replacement of Ca2+ ions with Mg2+ leads to measurable structural changes.11 Specifically, the substitution resulted in a contraction of the unit cell, as indicated by shifts in X-ray diffraction (XRD) peaks and a linear decrease in unit cell parameters (a0 and c0) and molar volume (V0) with increasing Mg content (up to 10 mol%). These structural modifications suggest that Mg2+ ions are readily incorporated into the TCP lattice. Additionally, the substitution significantly reduced the dissolution rate of the resulting material in aqueous solution, indicating improved chemical stability, a desirable feature for biomedical applications. Hydroxyapatite (Ca10(PO4)6(OH)2), the principal mineral component of human bone and teeth, has also been the subject of extensive Mg-substitution studies.12 High-precision hybrid DFT calculations and experimental synthesis revealed that Mg incorporation into HAP causes a reduction in lattice parameters and a noticeable distortion in the structure.13 The substitution at different calcium sites (Ca1 and Ca2) alters the electronic structure, energy levels, band gap, and bulk modulus. Particularly, substitution at the Ca2 site was more thermodynamically favorable and had a stronger effect on the infrared (IR) spectra and local structural asymmetry. These changes in local bonding environments directly influence the optical and electronic properties of HAP, making Mg-substituted HAP suitable for controlled reactivity in biological and photocatalytic systems. Interestingly, while low Mg content had a limited impact on basic catalytic reactivity, higher Mg concentrations led to structural disorder and nonstoichiometry, reducing the number of active basic sites. In such cases, new phases such as whitlockite were observed, highlighting a concentration-dependent tradeoff between structural stability and functionality. This aligns with the broader understanding that substitution levels must be carefully optimized to avoid undesired phase transitions or defect formation.

The motivation for this study stems from the growing need to enhance the hydrogen storage capacity and optoelectronic performance of existing fluorite-type metal hydrides. Among them, NbH2 stands out due to its stable fluorite structure, favorable electronic conductivity, and intrinsic hydrogen affinity.14,15 However, its hydrogen storage density and band gap characteristics remain suboptimal for advanced applications. Magnesium (Mg) was chosen as a dopant owing to its light atomic weight, high hydrogen gravimetric capacity, and known ability to enhance sorption kinetics when alloyed with transition metals. Unlike heavier dopants, Mg offers the dual benefit of reducing system weight while potentially modifying the band structure and boosting hydrogen density.16 Thus, this study aims to explore how Mg-substitution affects the structural stability, electronic properties, and hydrogen storage potential of NbH2. The novelty of the work lies in its systematic application of density functional theory (DFT) to uncover the atomistic effects of Mg incorporation in NbH2, a relatively unexplored yet promising material for multifunctional energy applications.

2. Computational methods

First-principles calculations based on density functional theory (DFT) were performed using the CASTEP module implemented in Materials Studio 2020.17 The calculations in this study were performed using density functional theory (DFT) within the generalized gradient approximation (GGA), employing the Perdew–Burke–Ernzerhof (PBE) functional to describe the exchange–correlation interactions.18 Relativistic effects were treated using the Koelling–Harmon scalar relativistic approximation.19 The electronic wavefunctions were expanded in a plane-wave basis set with a cutoff energy of 381 eV. Electronic convergence was achieved with an energy tolerance of 2 × 10−6 eV per atom and an eigenvalue convergence criterion of 8.889 × 10−7 eV. The charge density was mixed using the Pulay scheme,20 with a maximum mixing history length of 20, a mixing amplitude of 0.5, and a g-vector cut-off of 1.5 Å−1. Geometry optimization was carried out using the Broyden–Fletcher–Goldfarb–Shanno (BFGS) algorithm under fixed basis quality conditions.21 The maximum number of ionic steps was set to 100, with convergence criteria for the total energy, maximum ionic force, and atomic displacement set at 2 × 10−5 eV per atom, 0.05 eV Å−1, and 0.002 Å, respectively. To improve the accuracy of the calculated band structures and structural parameters (lattice constants), the Heyd–Scuseria–Ernzerhof (HSE03) hybrid functional was employed as a post-processing correction.22 The HSE03 functional provides a more reliable description of the electronic structure by incorporating a portion of exact Hartree–Fock exchange, which is particularly important for systems where bandgap and structural precision are critical.

3. Results and discussion

3.1 Bonding characteristics, population analysis, and hybridization

To understand the effect of Mg substitution on the bonding nature within the NbH2 fluorite structure, a detailed population and bond length analysis was performed using the Mulliken scheme. The goal was to quantify the bonding strength between H, Nb, and Mg atoms, as well as to assess how increasing Mg content influences electron distribution and interatomic interactions within the lattice. In pristine NbH2, the H–Nb bond length is 1.978 Å with a bond population of 0.35, indicating a moderately covalent character. The H–H interaction shows a negative bond population of −0.07, suggesting weak repulsive or nonbonding interactions between hydrogen atoms within the lattice.23,24 Upon Mg doping, significant changes in bonding patterns are observed. For instance, in the Mg-NbH2 compound, while the H–Nb bond population drops sharply to 0.06 (nearly nonbonding), new H–Mg interactions form with a bond population of 0.17, indicating weak ionic bonding. As Mg content increases in Mg2-NbH2 and Mg3-NbH2, the H–Nb bond population remains low (0.12 and 0.03 respectively), while the H–Mg bond population increases to 0.29 and then decreases to 0.15. These trends imply that Mg substitution progressively disrupts the original Nb–H bonding network and introduces ionic H–Mg interactions, which may facilitate hydrogen release due to weaker bonding. A comparative analysis with pure MgH2 supports this observation. The H–Mg bond population in MgH2 is as high as 0.57, indicating a significantly stronger ionic character compared to Mg-NbH2 variants. Moreover, the H–H population in MgH2 is highly negative (−0.68), suggesting stronger hydrogen–hydrogen repulsion, which may relate to higher desorption kinetics.

The atomic Mulliken population data (Table S2) further reinforce these conclusions. In pristine NbH2, the Nb atom exhibits a total electronic population of 15, with notable contributions from the d-orbital (4.084), reflecting its role in covalent bonding. As Mg atoms are introduced, the overall electronic populations increase with higher s- and p-orbital character (e.g., Mg-NbH2 shows a total of 56.997 electrons), suggesting a shift toward more ionic behavior. Particularly in Mg2-NbH2 and Mg3-NbH2, the electron populations become more delocalized, and d-orbital participation drops significantly (to ∼4.1), indicating weakened Nb-based bonding. In pure MgH2, the electron population is predominantly from s- and p-orbitals (5.217 and 6.783, respectively), aligning with its known ionic bonding nature.25,26

In summary, these bonding analyses reveal that Mg substitution in NbH2 reduces the covalent character of the Nb–H bonds and introduces weaker, more ionic Mg–H interactions. This change in bonding environment weakens the hydrogen binding energy, potentially improving desorption properties. Additionally, the disruption of the Nb–H network by Mg alters the electron distribution within the lattice, which may also influence the material's optoelectronic and catalytic behavior.

3.2 Structural analysis

NbH2 and MgH2 both adopt a fluorite-type structure within the cubic Fm[3 with combining macron]m space group. In NbH2, Nb2+ ions exhibit a face-centered cubic coordination with eight H1− atoms, featuring Nb–H distances of 1.98 Å. Each hydride ion connects to four Nb2+ centers, forming mixed corner- and edge-sharing HNb4 tetrahedra. Similarly, in MgH2, Mg2+ ions are surrounded by eight H1− atoms in a comparable geometry, with bond lengths of 2.06 Å. Here, each H1− ion coordinates with four Mg2+ atoms, resulting in HMg4 tetrahedral units that also share corners and edges.27 The structural analysis of NbH2, MgH2, and their magnesium-substituted derivatives reveals significant differences depending on the exchange–correlation functional used in the DFT calculations, namely, PBE and HSE03. Under the PBE functional, NbH2 displays a distorted orthorhombic symmetry with lattice constants a = c = 4.65 Å and b = 4.47 Å, resulting in a unit volume of 96.65 Å3 and a density of 6.52 g cm−3. This closely aligns with the 4.57 Å lattice constant reported by Xiaobing et al. for NbH2.27 MgH2, on the other hand, exhibits a larger unit cell volume (112.60 Å3) with lattice constants around 4.81–4.86 Å and a much lower density of 1.55 g cm−3, which slightly exceeds the 4.77 Å reported by Vajeeston et al.28 A known characteristic of the PBE functional is its tendency to slightly overestimate lattice dimensions. With increasing Mg content in the Mg1–3-NbH2 series, there is a gradual increase in unit volume (from 97.22 to 103.45 Å3) and density (up to 10.78 g cm−3), accompanied by angular distortions in Mg-NbH2 (α = 94°, β = 92°, γ = 88°), indicating a loss of perfect cubic symmetry and increasing structural distortion due to doping.

In contrast, the HSE03 functional maintains a constant unit volume (95.39 Å3) and lattice parameter (4.57 Å) across all compounds, regardless of composition. While this value aligns well with the experimental lattice constant for NbH2 reported by Xiaobing et al., the uniformity in volume and geometry across all doped systems, including MgH2, appears unrealistic. This suggests that the structures under HSE03 may not have undergone full relaxation or were constrained during optimization, making the results less physically meaningful for structural evaluation. Correspondingly, the densities calculated under HSE03 decrease with increasing Mg content, from 6.61 g cm−3 (NbH2) to 3.03 g cm−3 (Mg3-NbH2), which contradicts the expected trend of increasing mass and decreasing volume.

The enthalpy and total energy parameters derived from both PBE and HSE03 functionals provide essential insights into the thermodynamic stability of NbH2 and its Mg-substituted derivatives. Under the PBE functional, the total energy of NbH2 is −6.75 eV, with a corresponding enthalpy of −6.75 eV. As Mg is gradually introduced into the NbH2 framework to form Mg-NbH2, Mg2-NbH2, and Mg3-NbH2, the total energy becomes increasingly negative (−6788.06 eV, −6822.37 eV, and −6855.15 eV, respectively), indicating the growing size and atomic complexity of the systems. Although total energy values are not directly comparable across compounds of different compositions, the enthalpy values, which range from −6.79 eV to −6.86 eV across these substituted structures, suggest a trend toward improved thermodynamic stability with increasing Mg content. MgH2 exhibits the most negative enthalpy value at −6.89 eV, suggesting it is the most stable compound among those studied. A similar trend is observed under the HSE03 functional. The total energy of NbH2 is −699.49 eV with an enthalpy of −0.70 eV. As Mg substitution increases, the total energies of Mg-NbH2, Mg2-NbH2, and Mg3-NbH2 decline significantly to −2139.97 eV, −3582.32 eV, and −5022.71 eV, respectively, while their enthalpy values follow a corresponding downward trend from −2.14 eV to −5.02 eV. MgH2 again shows the most negative enthalpy at −6.46 eV, confirming its superior thermodynamic stability. While both functionals support the conclusion that Mg incorporation enhances the stability of NbH2-based compounds, PBE tends to yield slightly more negative enthalpy values, possibly due to its tendency to overbind. In contrast, HSE03, a hybrid functional, provides more moderate estimates but still reinforces the general stability trend. The consistent decrease in enthalpy with increasing Mg content across both computational methods underscores the favorable energetic landscape of Mg-doped NbH2 hydrides, with MgH2 emerging as the most stable configuration.

From a mechanical standpoint, the PBE functional shows a progressive decrease in bulk modulus from 318.63 GPa (NbH2) to just 6.27 GPa (Mg3-NbH2), indicating significant softening and structural weakening upon doping. Notably, MgH2 exhibits an unusually high bulk modulus (869.24 GPa), possibly due to tighter packing or bonding interactions. HSE03 did not report bulk modulus values, limiting direct mechanical comparison. Vibrational frequency trends further support the reliability of PBE for structural behavior, showing consistent values for undoped compounds (1668 cm−1), while HSE03 reveals a progressive increase in frequency with doping from 1163.69 cm−1 in Mg-NbH2 to 3836.40 cm−1 in Mg3-NbH2 though the unusually high values, especially in PBE for Mg3-NbH2 (6151.39 cm−1), suggest computational anomalies or strong local modes. The vibrational frequencies listed in Tables 1 and 2 represent the maximum optical phonon mode at the Γ-point, derived from lattice dynamics calculations. These frequencies provide insight into the vibrational behavior of the H atoms in the lattice, which is critical for evaluating hydrogen bonding strength, lattice dynamics, and the potential for reversible hydrogen desorption.29 Higher frequencies are typically associated with strong metal–H interactions and dynamic stability, while lower frequencies may reflect weakened bonding or soft modes due to structural distortion or increased Mg substitution.

Table 1 Calculated structural, energetic, and vibrational properties of NbH2, MgH2, and Mg-substituted NbH2 compounds (Mg-NbH2, Mg2-NbH2, Mg3-NbH2) using the GGA–PBE functional. Reported parameters include unit cell volume, density, lattice constants (a, b, c), total energy, enthalpy of formation, bulk modulus, and the highest vibrational mode (Γ-point frequency in cm−1)
Compounds Unit volume (A3) Unit density (g cm−3) Lattice constants Total energy (eV) Enthalpy (eV) Bulk modulus (GPa) Frequency (cm−1)
a b c
NbH2 96.65 6.52 4.65 4.47 4.65 −6754.24 −6.75 318.63 1668
Mg-NbH2 97.22 8.15 4.71 4.56 4.55 −6788.06 −6.79 209.42 1668
Mg2-NbH2 98.11 9.72 4.52 4.81 4.52 −6822.37 −6.82 156.96 1321.48
Mg3-NbH2 103.45 10.78 4.67 4.76 4.65 −6855.15 −6.86 6.27 6151.39
MgH2 112.60 1.55 4.86 4.81 4.81 −6888.58 −6.89 869.24 1668


Table 2 Structural, energetic, and vibrational properties of the studied hydrides computed using the HSE03 hybrid functional. This includes unit cell volume, density, lattice constants, total energy, enthalpy of formation, and maximum vibrational frequency (in cm−1). The data are used to validate and complement the GGA–PBE results
Compounds Unit volume (A3) Unit density (g cm−3) Lattice constants Total energy (eV) Enthalpy (eV) Bulk modulus (GPa) Frequency (cm−1)
a = b = c
NbH2 95.39 6.61 4.57 −699.49 −0.70 1668
Mg-NbH2 95.39 5.24 4.57 −2139.97 −2.14 1163.69
Mg2-NbH2 95.39 4.22 4.57 −3582.32 −3.58 1255.92
Mg3-NbH2 95.39 3.03 4.57 −5022.71 −5.02 3836.40
MgH2 95.39 1.83 4.57 −6463.42 −6.46 1668


The differences between the GGA–PBE values (Table 1) and those obtained using HSE03 (Table 2) arise from the known limitations of the GGA functional, which tends to underestimate bandgaps and yields functional-dependent absolute energies. Nonetheless, GGA–PBE is suitable for assessing structural trends, relative stabilities, and vibrational characteristics across chemically similar compounds, and all values within each table are internally consistent. In summary, the PBE functional more accurately captures the structural evolution and distortion due to Mg substitution in NbH2, reflecting realistic trends in lattice parameters, unit cell volume, density, and mechanical stability. Meanwhile, HSE03, although aligning precisely with some literature values for NbH2, offers limited structural flexibility and possibly constrained optimization, making it less suitable for analyzing doping-induced changes. Therefore, for structural analysis involving substitutional effects, PBE appears to provide more reliable and physically consistent results (Fig. 1).


image file: d5ra03949e-f1.tif
Fig. 1 Optimized crystal structures of NbH2, MgH2, and Mg-substituted NbH2 (Mg-NbH2, Mg2-NbH2, Mg3-NbH2) showing their fluorite-derived configurations. Atoms are color-coded by element to emphasize structural distortion due to Mg substitution.

3.3 Electronic analysis: band structure and PDOS

The electronic properties of the studied compounds were investigated through both GGA–PBE and HSE03 functionals, providing insights into their band structure and density of states (DOS) characteristics. The GGA–PBE band structure in Fig. 2 reveals that the compounds exhibit conducting behavior, with the conduction band minimum (CBM) and valence band maximum (VBM) overlapping one another. This is except for MgH2, whose CBM and VBM are located at different k-points (indicating an indirect band gap) or at the same k-point (for direct band gap systems). However, the band gaps obtained from GGA–PBE are generally underestimated, a known limitation of this functional.18 This shortcoming is addressed in the HSE03 band structure plots in Fig. 3. The HSE03 functional shows precisely the same results as those from GGA–PBE, which confirms that the GGA–PBE method used is accurate. Conducting materials like Mg-NbH2, which exhibit metallic behavior due to the presence of electronic states at the Fermi level, are valuable in a wide range of technological applications that require efficient charge transport. The use of such materials is in electrical contacts and interconnects within electronic devices, where their low resistivity ensures minimal energy loss and high current-carrying capacity.30 In addition, conducting hydride-based compounds have potential in hydrogen storage and sensing applications, where their electrical conductivity can facilitate real-time monitoring of hydrogen uptake or release via resistive or impedance-based methods.31 Generally, the metallic nature of these compounds makes them suitable for integration into next-generation energy devices, sensors, and electronic systems where conductivity and functional reactivity are both required.
image file: d5ra03949e-f2.tif
Fig. 2 Electronic band structures of NbH2, MgH2, and Mg-substituted NbH2 compounds calculated using the GGA–PBE functional. The band dispersion is plotted along high-symmetry directions of the Fm[3 with combining macron]m Brillouin zone. The Fermi level is aligned to 0 eV.

image file: d5ra03949e-f3.tif
Fig. 3 Total and partial density of states (TDOS and PDOS) for the studied compounds obtained with the GGA–PBE method. The contribution of individual atomic orbitals (Nb d, Mg s, H s) is highlighted to reveal electronic characteristics and bonding behavior.

The PDOS and TDOS plots in Fig. 3 offer a more detailed understanding of the electronic structure by showcasing the contribution of specific atomic orbitals to the electronic states near the Fermi level.32 In the PDOS, the valence band is predominantly composed of s-orbital contributions from the smaller atoms (such as Mg and H, depending on the composition), while the conduction band shows significant contributions from the d-orbitals of the metal center of Nb. All the compounds show three peaks, except MgH2, which is fully occupied by the s-orbitals in bonding. NbH2 shows a TDOS of 7, a d-orbital contribution of 3.5, and an s-orbital contribution of 4. MgH2 shows a TDOS contribution of 3.5 and an s-orbital contribution of 2.8. Mg-NbH2 shows a TDOS of 5.8, d-orbital contribution of 3, and s-orbital contribution of 3.2. Mg2-NbH2 shows a TDOS contribution of 5, a d-orbital contribution of 2.9, and an s-orbital contribution of 3. Mg3-NbH2 shows a TDOS contribution of 2, a d-orbital contribution of 3.8 and 3.3. From this, it is observed that the substitution of Mg in NbH2 reduces the TDOS contribution as well as the d-orbital contribution and but increases the s-orbital contribution as compared to the parent compound NbH2. This indicates strong hybridization between the metal d-states and the non-metal s-states, suggesting covalent bonding interactions that could influence charge mobility.33 The TDOS further supports these observations by displaying the overall electronic density distribution, with pronounced peaks near the valence and conduction band edges, implying localized states that are important for optical transitions. In summary, the combination of band structure and DOS analyses illustrates that the compounds are conductors with substantial orbital hybridization. The discrepancy between the PBE and HSE03 results underlines the importance of using hybrid functionals for accurate electronic property predictions (Fig. 4).


image file: d5ra03949e-f4.tif
Fig. 4 Electronic band structures of NbH2 and Mg-substituted NbH2 calculated using the HSE03 functional, along high-symmetry directions of the Fm[3 with combining macron]m Brillouin zone. The Fermi level is set to 0 eV. These results are used to cross-validate GGA–PBE predictions.

3.4 Optoelectronic analysis

3.4.1 Optical absorption, reflectivity, and loss function. The optoelectronic properties of the studied compounds NbH2, Mg-NbH2, Mg2-NbH2, Mg3-NbH2, and MgH2 were evaluated based on their absorption coefficient, reflectivity, and energy loss function across the photon energy range of 0 to 80 eV, as shown in the provided plots. In the absorption plot, NbH2 exhibits the highest absorption coefficient across the spectrum, peaking around 5.5 × 105 in the energy range of approximately 30–40 eV, indicating its superior capability for absorbing high-energy photons. This suggests potential for optoelectronic applications where strong photon–material interaction is required.34 The incorporation of magnesium into NbH2 generally reduces the absorption strength, with Mg2-NbH2 and Mg3-NbH2 showing the lowest absorption intensities, suggesting that increased Mg content diminishes the optical absorption performance. MgH2, although a separate phase, shows moderate absorption primarily in the lower energy range (5–25 eV), reflecting different electronic transitions compared to the Nb-containing compounds.

Reflectivity analysis discloses that MgH2 demonstrates an exceptionally high reflectivity (approaching 1.0) in the low-energy region (below 10 eV), significantly higher than all other compounds. This implies strong surface reflection of incident light, which may reduce its suitability for applications requiring light penetration.35 NbH2 shows moderate reflectivity throughout the spectrum, while the Mg-doped variants, especially Mg2-NbH2 and Mg3-NbH2, show reduced reflectivity, indicating improved photon absorption capabilities in those regions.

The loss function plot, which provides understanding on the energy loss of fast electrons traversing the material,36 is particularly dominated by a sharp peak for MgH2 around 10 eV, reaching a maximum value of 60. This signifies a strong plasmon resonance in MgH2, indicative of collective oscillations of the electron gas. In contrast, the other compounds exhibit lower and broader peaks, suggesting less intense plasmonic activity. NbH2 and its Mg-doped variants show moderate loss function values in the 10–30 eV range, with multiple smaller peaks, pointing to more complex electronic transitions and weaker plasmonic behavior. Inclusively, NbH2 stands out for its high absorption and moderate reflectivity, making it a promising candidate for energy harvesting or UV shielding applications. The Mg incorporation tunes the optical response, potentially allowing for the engineering of materials with tailored optoelectronic characteristics. MgH2, with its intense reflectivity and sharp energy loss peak, is more reflective and plasmonically active, which may be beneficial in applications involving light reflection or plasmonic enhancement (Fig. 5).


image file: d5ra03949e-f5.tif
Fig. 5 Optical absorption spectra, reflectivity, and energy loss function of the studied compounds, calculated using the GGA–PBE method. The plots reveal light-harvesting potential and dielectric loss behavior relevant to photocatalysis and optoelectronics.
3.4.2 Optical conductivity. Optical conductivity refers to a material's response to an electromagnetic field (light) in terms of its ability to conduct electricity at various frequencies, especially in the visible and near-visible spectrum. It is a frequency-dependent counterpart of electrical conductivity denoted as σ(ω), where σ is the conductivity and ω is the angular frequency of the light.37 This physical property is described by the real part of the optical conductivity (Re[σ(ω)]) which involves energy dissipation (i.e., absorption of light) and the imaginary part of the optical conductivity ([Im σ(ω)]) which is related to energy storage in the medium.38 In Fig. 6, the plots for the real and imaginary optical conductivities are shown for the studied compounds. Starting with the real conductivity plot, MgH2 (blue) shows the highest peak at 13 s m−1 followed by NbH2 (black) at 9 s m−1, Mg-NbH2 (red) shows a peak at 8 s m−1, Mg3-NbH2 (green) at 6.3 s m−1 and Mg2-NbH2 (purple) at 6 s m−1. This means that the subsequent addition of Mg to NbH2 gradually reduces the dissipation energy thus affecting the absorption capacity of the compounds. On the other hand, complete substitution of Nb with Mg in MgH2 shows that MgH2 strongly interacts with light matter to enhance optical conductivity. The second plot which contains the imaginary optical conductivity shows MgH2 to possess the highest peak at 7 s m−1 followed by NbH2 at 5 s m−1, then Mg-NbH2 at 4.8 s m−1, Mg3-NbH2 at 4 s m−1 and Mg2-NbH2 at 3.8 s m−1. Thus, the imaginary plot follows the same route as the real plot which the concurrent reduction in conductivity as the substitution of Mg in NbH2 increases.
image file: d5ra03949e-f6.tif
Fig. 6 Real and imaginary parts of optical conductivity for the studied compounds under GGA–PBE. These parameters indicate photon-induced charge carrier response and material conductivity in different energy regions.
3.4.3 Optical dielectric function. The dielectric function is a complex quantity that describes how a material responds to an external electric field, especially in the presence of an oscillating electromagnetic wave such as light. It is fundamental in understanding a material's optical, electronic, and energy absorption properties. Mathematically, the dielectric function is expressed as:39
 
ε(ω) = ε1(ω) + iε2(ω) (1)
where ε1(ω) is the real part, representing the material's ability to store electric energy (related to polarization and refractive index) while ε2(ω) is the imaginary part, which accounts for the energy loss (i.e., absorption of the electromagnetic wave by the material).40–42

The dielectric function plots comprising the real and imaginary parts reveal vital information on the optical response and electronic polarizability of the investigated hydride compounds.43 Among the materials, Mg-NbH2 demonstrates the most pronounced peaks in both the real and imaginary components of the dielectric function, indicating a strong interaction with incident electromagnetic radiation and high polarizability. Specifically, in the real part, Mg-NbH2 peaks sharply around 0 eV with a maximum value exceeding 30, suggesting substantial static dielectric response, which aligns with its metallic behavior and delocalized charge carriers. In the imaginary part, which reflects absorption characteristics, Mg-NbH2 again exhibits the highest peak (around 85), emphasising its strong interband transitions and enhanced optical activity.

Following Mg-NbH2, MgH2 shows the next highest peaks in both real and imaginary components, indicative of relatively strong optical absorption and dielectric response, consistent with its semiconducting nature. NbH2 and Mg3-NbH2 exhibit moderate peaks, suggesting intermediate dielectric behavior and limited but noticeable interaction with electromagnetic fields. Lastly, Mg2-NbH2 shows the least pronounced peaks across both dielectric plots, implying the lowest optical absorption and static dielectric constant among the series. These trends point to Mg-NbH2 being the most optically active and polarizable, making it a strong candidate for optoelectronic or plasmonic applications, while Mg2-NbH2 would be more optically inert (Fig. 7).


image file: d5ra03949e-f7.tif
Fig. 7 Real and imaginary components of the dielectric function (ε1 and ε2) of the studied compounds obtained from GGA–PBE optical calculations, showing their polarizability and optical transparency across energy ranges.
3.4.4 Optical refractive index. The refractive index (denoted as n) is a fundamental optical property of a material that describes how light propagates through it. It is defined as the ratio of the speed of light in a vacuum (c) to the speed of light in the material (v):44
 
image file: d5ra03949e-t1.tif(2)

There are two components, which include the real part (n1) representing how much the light is bent or refracted when entering the material.45 A higher n1 means greater bending and stronger optical density. The imaginary part (n2) is associated with the absorption of light within the material. A higher n indicates that the material absorbs more light and transmits less. Together, the complex refractive index is written as:46

 
ñn1 + in2 (3)

The refractive index plots presented illustrate both the real (n1) and imaginary (n2) parts of the refractive indices of the compounds NbH2, MgH2, Mg-NbH2, Mg3-NbH2, and Mg2-NbH2 over a wide energy range.47 These indices offer vital insights into the optical transparency, reflectivity, and absorption of the materials. In the real part of the refractive index (n1) plot, Mg-NbH2 exhibits the highest peak value of about 3.2, indicating strong light bending capability and high optical density in the low-energy region. This is followed closely by MgH2, which also shows a significant peak above 3.0, confirming its notable dielectric response. NbH2 presents a slightly lower peak around 2.5, while Mg3-NbH2 and Mg2-NbH2 have more moderate peaks, suggesting comparatively lower light-retardation effects. These values gradually taper off with increasing photon energy, reflecting a decrease in refractive behavior at high energies. In the imaginary part of the refractive index (n2) plot, which relates to absorption, Mg-NbH2 again shows the most prominent peak, exceeding 9.0, indicating it has the highest optical absorption in the visible-to-UV region. This implies greater interband transitions and energy dissipation in this material. Following this are NbH2, MgH2, and Mg3-NbH2, all exhibiting moderate absorption behaviors. Mg2-NbH2 has the lowest imaginary peak, suggesting it is more optically transparent and has reduced electronic losses compared to the others. These variations in the refractive indices reflect differences in the electronic structure and bonding nature of the compounds, particularly the extent of hybridization and electron delocalization. The high optical activity of Mg-NbH2 positions it as a strong candidate for optoelectronic or photonic applications, while Mg2-NbH2 may be more suitable for transparent or low-loss dielectric uses (Fig. 8).


image file: d5ra03949e-f8.tif
Fig. 8 Refractive index (n) and energy loss function (L) spectra (both real and imaginary parts) for the studied compounds. These optical properties provide insight into the interaction of light with the material and the onset of plasmonic behavior.

3.5 Elastic and mechanical properties

The elastic and mechanical properties of the studied compounds reveal significant insights into their structural rigidity, ductility, and stability. For all systems, the elastic constants C11, C12, and C44 play critical roles in defining their resistance to deformation. Notably, NbH2 exhibits high stiffness with C11 = 257.15 GPa and C44 = 111.15 GPa, indicating a strong and mechanically stable structure. The Born mechanical stability criteria—C11 > 0, C44 > 0, C11–C12 > 0, and C11 +2C12 > 0—are satisfied by NbH2, Mg-NbH2, and Mg2-NbH2, confirming their mechanical stability.48 However, Mg3-NbH2 and MgH2 fail to meet the C44 > 0 criterion, with negative values (−8.03 and −9.27 GPa, respectively), indicating shear instability and potential mechanical failure under certain deformations.

Among the compounds, NbH2 also possesses the highest bulk modulus (156.96 GPa), shear modulus (106.45 GPa), and Young's modulus (260.47 GPa), highlighting its superior resistance to volume change, shear deformation, and overall elastic deformation, respectively. As magnesium is incrementally added, these mechanical parameters decline, with MgH2 having the lowest values (B = 23.76 GPa, G = 20.19 GPa, and E = 47.21 GPa), indicating a much softer and more ductile material. The Pugh ratio (B/G), which measures ductility, further differentiates the compounds. A value above 1.75 typically indicates ductility, while values below suggest brittleness.49 Mg-NbH2 has the highest B/G ratio (3.27), suggesting it is the most ductile among the set, whereas Mg3-NbH2 (0.97) and Mg2-NbH2 (1.35) lean towards brittleness. NbH2, with a B/G of 1.47, straddles the brittle–ductile borderline. The anisotropic factor (A), which describes directional dependency of the elastic response,50 ranges from 0.21 for NbH2 (indicating near isotropy) to a highly negative −9.67 for Mg3-NbH2, suggesting severe elastic anisotropy and instability. Similarly, Poisson's ratio (υ), which reflects the degree of lateral expansion when compressed,51 ranges anomalously, with Mg3-NbH2 and MgH2 showing unrealistic values (1.59 and −10.14, respectively), further confirming their mechanical instability or computational artefacts due to soft shear modes.

In summary, the data indicate that NbH2 is the most mechanically stable and rigid among the studied compounds. The addition of Mg modifies the mechanical behavior, with Mg-NbH2 showing enhanced ductility but reduced stiffness, while higher Mg concentrations lead to mechanical instability, as reflected by the failure to satisfy some of the Born criteria and the emergence of negative or extreme values in key elastic parameters (Table 3).

Table 3 Calculated elastic constants (C11, C12, C44) and derived mechanical parameters including bulk modulus (B), shear modulus (G), Young's modulus (E), and Poisson's ratio (ν) for NbH2, MgH2, and Mg-substituted NbH2 compounds (Mg-NbH2, Mg2-NbH2, Mg3-NbH2) using the GGA–PBE method. These parameters provide insight into the mechanical stability, ductility, and stiffness of the hydride materials
Elastic moduli (GPa) NbH2 Mg-NbH2 Mg2-NbH2 Mg3-NbH2 MgH2
C11 > 0 257.15 179.28 176.68 360.03 95.60
C12 102.94 102.59 39.92 165.62 −26.12
C44 > 0 111.15 44.32 38.33 −8.03 −9.27
C11 + 2C12 > 0 463.03 384.46 256.52 691.27 43.36
C11–C12 > 0 154.21 76.69 136.76 194.41 121.72
Bulk modulus (B) 156.96 126.64 78.83 68.56 23.76
Shear modulus (G) 106.45 38.74 58.38 70.53 20.19
Young modulus (E) 260.47 105.46 140.47 157.56 47.21
Pugh ratio (B/G) 1.47 3.27 1.35 0.97 1.18
Anisotropic factor (A) 0.21 0.37 0.55 −9.67 1.07
Poisson ratio (υ) 0.23 0.41 0.22 1.59 −10.14


3.6 Hydrogen storage capacity

The hydrogen gravimetric storage capacity of a material refers to the mass percentage of hydrogen that can be stored within the compound, which is a crucial parameter for evaluating materials for hydrogen storage applications. A higher gravimetric capacity indicates that a compound can store more hydrogen per unit mass, making it more efficient for lightweight and compact hydrogen storage systems, especially in the context of fuel cells and clean energy technologies. The formula for calculating the hydrogen gravimetric capacity is given as:52
 
image file: d5ra03949e-t2.tif(4)

From the calculations, MgH2 exhibits the highest hydrogen storage capacity at 7.66 wt%, consistent with its lightweight elemental nature and strong hydrogen affinity.53,54 This high gravimetric capacity underscores MgH2's long-standing position as a benchmark hydrogen storage material, despite its high desorption temperature and sluggish kinetics. In contrast, pristine NbH2 has a significantly lower hydrogen capacity of 2.124 wt%, which can be attributed to the heavier atomic weight of niobium and a lower hydrogen-to-metal ratio. However, when Mg is incrementally introduced into the NbH2 lattice to form ternary hydrides—Mg-NbH2 (3.326%), Mg2-NbH2 (2.797%), and Mg3-NbH2 (2.378%), a nonlinear but significant improvement in hydrogen capacity is observed relative to pure NbH2. This increase is primarily due to the incorporation of lighter Mg atoms into the matrix, which reduces the overall molecular weight of the hydride without proportionally sacrificing the hydrogen content. Interestingly, the capacity peaks at Mg-NbH2 and slightly declines as more Mg is added (Mg2 and Mg3), likely due to saturation of substitutional sites or a shift in phase stability toward less hydrogen-rich or more stable configurations. This suggests that while Mg improves the H/M (hydrogen-to-metal) ratio initially, excessive substitution may hinder additional hydrogen uptake.

Compared to the state-of-the-art systems, these Mg-NbH2 variants offer moderate hydrogen capacities, bridging the gap between high-capacity but kinetically limited MgH2, and low-capacity but stable intermetallic hydrides like LaNi5H6 (∼1.4%) or TiFe (∼1.9%).55,56 Moreover, the bonding and electronic analysis in earlier sections indicate that Mg incorporation weakens the H–H-metal bonding strength, potentially enhancing desorption kinetics, an important advantage over pure MgH2. In conclusion, although the gravimetric hydrogen capacities of the Mg-NbH2 systems are lower than that of pure MgH2, they present a promising trade-off between capacity, bonding strength, and structural stability. Their performance may be further enhanced through doping, nanostructuring, or hybridization, making them viable candidates for next-generation solid-state hydrogen storage materials.

Upon desorption of hydrogen from the compounds, as modelled using the PBE functional, several significant structural and energetic changes are observed. For all systems, there is a notable decrease in total energy and enthalpy compared to the hydrogenated structures, though the trend of increasing thermodynamic stability with increasing Mg content remains consistent. The total energies for NbH2, Mg-NbH2, Mg2-NbH2, Mg3-NbH2, and MgH2 after hydrogen desorption are −6625.08 eV, −6658.70 eV, −6692.73 eV, −6726.11 eV, and −6760.63 eV, respectively, with corresponding enthalpy values ranging from −6.63 eV to −6.76 eV. These values suggest that desorption slightly lowers the enthalpic stability of each compound compared to its hydrogenated counterparts but does not significantly disrupt the stability hierarchy. Structurally, desorption leads to a reduction in unit volume for all compounds except Mg3-NbH2, which expands slightly. NbH2 shows the largest contraction in unit volume from 96.65 Å3 to 76.12 Å3, indicative of lattice shrinkage upon hydrogen removal. The increase in unit density from 6.52 g cm−3 to 8.28 g cm−3 for NbH2 and up to 13.67 g cm−3 for Mg3-NbH2 also supports this compaction behavior. Lattice constants adjust accordingly, with varying degrees of anisotropic distortion depending on Mg content. Notably, NbH2 exhibits angular distortions after desorption, with α = 80.95°, β = 89.98°, and γ = 90.04°, indicating a departure from ideal cubic symmetry and a transition toward a slightly distorted or monoclinic-like structure. All other compounds exhibit 90°.

The bulk modulus values post-desorption exhibit notable variation. For example, Mg-NbH2 shows a dramatic drop in bulk modulus to 133.54 GPa, indicating a softer and more compressible structure, whereas NbH2 maintains a relatively high modulus of 1172.09 GPa. These differences suggest that Mg incorporation affects the mechanical rigidity of the structure more significantly after hydrogen release. Vibrational frequency data reveal sharp contrasts as well. The frequency remains at 1668 cm−1 for NbH2 and MgH2, implying similar vibrational modes post-desorption. However, Mg-NbH2 shows an extremely high frequency of 7709.70 cm−1, suggesting the emergence of highly localized or stiff vibrational modes, potentially due to structural rearrangements or residual stress within the desorbed lattice. Overall, hydrogen desorption leads to structural compaction, modest decreases in enthalpic stability, and changes in mechanical and vibrational properties. The general trend of increasing Mg content enhancing the stability of the system is preserved, while significant modifications in lattice structure and dynamics point to complex desorption-driven behavior in these metal hydride materials (Table 4).

Table 4 Structural parameters of NbH2, MgH2, and Mg-substituted NbH2 compounds after hydrogen desorption, calculated using the GGA–PBE method. Reported values include unit cell volume, density, lattice constants, total energy, enthalpy, and bulk modulus. The data highlight the structural evolution and thermodynamic implications of hydrogen release
Compounds Unit volume (A3) Unit density (g cm−3) Lattice constants Total energy (eV) Enthalpy (eV) Bulk modulus (GPa) Frequency (cm−1)
a b c
NbH2 76.12 8.28 4.91 3.60 4.36 −6625.08 −6.63 1172.09 1668
Mg-NbH2 74.99 10.56 3.78 4.65 4.27 −6658.70 −6.66 133.54 7709.70
Mg2-NbH2 77.45 12.31 4.01 4.81 4.01 −6692.73 −6.69 94.60 1321.48
Mg3-NbH2 81.58 13.67 4.30 4.42 4.30 −6726.11 −6.73 355.18 6151.39
MgH2 93.09 1.88 4.54 4.53 4.53 −6760.63 −6.76 385.95 1668


4. Conclusions

This study presents a comprehensive density functional theory (DFT) investigation of NbH2, MgH2, and Mg-substituted NbH2 systems (Mg-NbH2, Mg2-NbH2, and Mg3-NbH2), aimed at exploring their structural, electronic, optical, and hydrogen storage properties for multifunctional energy applications. The substitution of Mg into the fluorite-type NbH2 matrix leads to notable structural distortions, increasing unit cell volume and density, and introducing flexibility in the lattice geometry. Among the computational approaches, GGA/PBE proved more effective than HSE03 in capturing these trends realistically, especially under varying Mg content. Electronic analysis revealed a transition from metallicity in NbH2 and its Mg-substituted counterparts to semiconducting behavior in MgH2, underscoring the tunability of electronic properties via controlled doping. The total density of states (TDOS) decreased with increasing Mg content, suggesting reduced carrier density and modified electronic transport characteristics. Optically, Mg-NbH2 exhibited superior dielectric strength and refractive index, positioning it as the most promising candidate for optoelectronic and plasmonic applications, while MgH2 showed dominant reflectivity and optical conductivity in the UV-visible range.

From a mechanical standpoint, NbH2 demonstrated high rigidity and elastic stability, whereas Mg doping progressively enhanced ductility but at the cost of mechanical stability, especially in Mg2-NbH2 and Mg3-NbH2, where instability indicators such as negative shear constants and unrealistic Poisson ratios were observed. In terms of hydrogen storage, MgH2 retained the highest gravimetric capacity (7.66 wt%), but its known drawbacks such as slow desorption kinetics highlight the need for alternatives. The Mg-NbH2 compounds showed moderate hydrogen capacity (2.4–3.3 wt%) and offer a promising trade-off between structural stability, storage potential, and functional versatility. In summary, this work identifies Mg-NbH2 as a tunable material system with balanced performance across hydrogen storage, optoelectronics, and mechanical robustness. The insights gained here provide a rational design strategy for developing next-generation metal hydrides with multifunctional energy applications.

Data availability

Data will be made available upon request.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

We acknowledged the centre for High-Performance Computing Centre, South Africa for the computational resources.

References

  1. F. Qureshi, M. Yusuf, H. Kamyab, D. V. N. Vo, S. Chelliapan, S. W. Joo and Y. Vasseghian, Latest eco-friendly avenues on hydrogen production towards a circular bioeconomy: Currents challenges, innovative insights, and future perspectives, Renewable Sustainable Energy Rev., 2022, 168, 112916 CrossRef CAS.
  2. A. Y. Esayed and D. O. Northwood, Metal hydrides: a review of group V transition metals—niobium, vanadium and tantalum, Int. J. Hydrogen Energy, 1992, 17, 41–52 CrossRef CAS.
  3. M. R. Ghaani, Study of New Materials and Their Functionality for Hydrogen Storage and Other Energy Applications, PhD dissertation, University of Milano-Bicocca, 2014.
  4. M. Babar, A. W. Anwar, M. Moin, M. A. Ahmad, U. Thumu, A. Ali and A. Qadoos, Comparative analysis of inorganic lead halide perovskites with promising (Mg+2)-doped for optoelectronic applications: a computational insights, Opt. Quantum Electron., 2024, 56, 732 CrossRef CAS.
  5. H. T. Rouf, T. Zeeshan, M. Khalil, F. Ullah, S. M. Ramay and M. Saleem, DFT and experimental study of Mg substituted strontium oxide for optoelectronic applications, J. Phys. Chem. Solids, 2025, 201, 112656 CrossRef CAS.
  6. R. C. Ropp, Encyclopedia of the Alkaline Earth Compounds. Encyclopedia of the Alkaline Earth Compounds. Newnes, 2012 Search PubMed.
  7. S. Tekumalla and M. Gupta, in Magnesium—The Wonder Element for Engineering/Biomedical Applications, IntechOpen, 2020 Search PubMed.
  8. M. Halka and B. Nordstrom, Alkali and Alkaline Earth Metals, Infobase Publishing, 2010 Search PubMed.
  9. M. I. Khan, Z. Saman, A. Mujtaba, M. S. Hasan, M. Zhang and S. Ezzine, Tailoring the structural, morphological, optical, photocatalysis, and electrical characteristics of Zn0.5Mg0.5−xCoxLa0.1Fe1.9O4 ferrites across cobalt concentration from x = 0.0–0.2, Mater. Sci. Eng., B, 2025, 319, 118352 CrossRef CAS.
  10. P. Schouwink, M. B. Ley, A. Tissot, H. Hagemann, T. R. Jensen, Ľ. Smrčok and R. Černý, Structure and properties of complex hydride perovskite materials, Nat. Commun., 2014, 5, 5706 CrossRef CAS PubMed.
  11. C. Tardei, F. Grigore, I. Pasuk and S. Stoleriu, The study of Mg2/Ca2 substitution of-tricalcium phosphate, J. Optoelectron. Adv. Mater., 2006, 8, 568–571 CAS.
  12. M. D. Y. S. Beqain, Novel Substituted Hydroxyapatites, PhD Thesis, The University of Waikato, 2023.
  13. G. Costentin, C. Drouet, F. Salles and S. Sarda, Structure and Surface Study of Hydroxyapatite-Based Materials: Experimental and Computational Approaches, Design and Applications of Hydroxyapatite-Based Catalysts, Elsevier, 2022, pp. 73–140 Search PubMed.
  14. Z. Chuanzhao, S. Guoliang, W. Jingjing, L. Cheng, J. Yuanyuan, K. Xiaoyu and H. Andreas, Prediction of Novel High-Pressure Structures of Magnesium Niobium Dihydride, ACS Appl. Mater. Interfaces, 2017, 9(31), 26169–26176 CrossRef PubMed.
  15. H. Wenzl, Ordered and Disordered Hydrogen Interstitials in Niobium, J. Phys. Colloq., 1978, 277, 221 Search PubMed.
  16. F. D'Errico, M. Tauber and M. Just, Magnesium alloys for sustainable weight-saving approach: A brief market overview, new trends, and perspectives, Curr. Trends Magnesium (Mg) Res., 2022, 13–45 Search PubMed.
  17. S. J. Clark, M. D. Segall, C. J. Pickard, P. J. Hasnip, M. J. Probert, K. Refson, M. C. Payne and Z. Kristallogr, First principles methods using CASTEP, Z. Kristallogr. Cryst. Mater., 2005, 220, 567–570 CrossRef CAS.
  18. J. P. Perdew, K. Burke and M. Ernzerhof, Perdew, burke, and ernzerhof reply, Phys. Rev. Lett., 1998, 80, 891 CrossRef CAS.
  19. I. Grinberg, N. J. Ramer and A. M. Rappe, Transferable relativistic dirac-slater pseudopotentials, Phys. Rev. B:Condens. Matter Mater. Phys., 2000, 62, 2311 CrossRef CAS.
  20. P. Pulay and W. Meyer, An efficient reformulation of the closed-shell self-consistent electron pair theory, J. Chem. Phys., 1984, 81, 1901–1905 CrossRef CAS.
  21. J. D. Head and M. C. Zerner, A Broyden—Fletcher—Goldfarb—Shanno optimization procedure for molecular geometries, Chem. Phys. Lett., 1985, 122, 264–270 CrossRef CAS.
  22. K. Hummer, J. Harl and G. Kresse, Heyd-Scuseria-Ernzerhof hybrid functional for calculating the lattice dynamics of semiconductors, Phys. Rev. B: Condens. Matter Mater. Phys., 2009, 80, 115205 CrossRef.
  23. S. J. Grabowski, What is the covalency of hydrogen bonding?, Chem. Rev., 2011, 111, 2597–2625 CrossRef CAS PubMed.
  24. J. Contreras-García, W. Yang and E. R. Johnson, Analysis of hydrogen-bond interaction potentials from the electron density: integration of noncovalent interaction regions, J. Phys. Chem. A, 2011, 115, 12983–12990 CrossRef PubMed.
  25. A. E. Reed, R. B. Weinstock and F. Weinhold, Natural population analysis, J. Chem. Phys., 1985, 83, 735–746 CrossRef CAS.
  26. H. Lu, D. Dai, P. Yang and L. Li, Atomic orbitals in molecules: general electronegativity and improvement of Mulliken population analysis, Phys. Chem. Chem. Phys., 2006, 8, 340–346 RSC.
  27. X. Xiaobing, Z. Weibing, Y. Weiyang, W. Na and E. Tang Biyu, Energetics and electronic properties of Mg2 TMH2 (TM = Sc, Ti, V, Y, Zr, Nb): an ab initio study, Phys. B, 2009, 404 Search PubMed.
  28. P. Vajeeston, P. Ravindran, A. Kjekshus and H. Fjellvåg, High hydrogen content complex hydrides: a density-functional study, Appl. Phys. Lett., 2006, 8, 071906 CrossRef.
  29. E. T. Nibbering and T. Elsaesser, Ultrafast vibrational dynamics of hydrogen bonds in the condensed phase, Chem. Rev., 2004, 104, 1887–1914 CrossRef CAS PubMed.
  30. M. M. H. Bhuiya, A. Kumar and K. J. Kim, Metal hydrides in engineering systems, processes, and devices: a review of non-storage applications, Int. J. Hydrogen Energy, 2015, 40, 2231–2247 CrossRef CAS.
  31. L. W. Wang, Calculating the density of states and optical-absorption spectra of large quantum systems by the plane-wave moments method, Phys. Rev. B:Condens. Matter Mater. Phys., 1994, 49, 10154 CrossRef CAS PubMed.
  32. M. S. Khan, B. Gul, G. Benabdellah, B. Ahmad, N. H. Alotaibi, S. Mohammad and H. Ahmad, Insights into optoelectronic, thermodynamic, and thermoelectric properties of novel GePtCh (Ch = S, Se, Te) semiconductors: first-principles perspective, Phys. Scr., 2024, 99, 065939 CrossRef.
  33. L. Zhang, X. Xiao, C. Xu, J. Zheng, X. Fan, J. Shao and L. Chen, Remarkably improved hydrogen storage performance of MgH2 catalyzed by multivalence NbHx nanoparticles, J. Phys. Chem. C, 2015, 119, 8554–8562 CrossRef CAS.
  34. U. Mehmood, F. A. Al-Sulaiman, B. S. Yilbas, B. Salhi, S. H. A. Ahmed and M. K. Hossain, Superhydrophobic surfaces with antireflection properties for solar applications: a critical review, Sol. Energy Mater. Sol. Cells, 2016, 157, 604–623 CrossRef CAS.
  35. K. Sturm, Adv. Electron energy loss in simple metals and semiconductors, Adv. Phys., 1982, 31, 1–64 CrossRef CAS.
  36. N. R. Abdullah, H. O. Rashid, B. J. Abdullah, C. S. Tang and V. Gudmundsson, Planar buckling controlled optical conductivity of SiC monolayer from deep-UV to visible light region: a first-principles study, Mater. Chem. Phys., 2023, 297, 127395 CrossRef CAS.
  37. I. Khan, A. Ullah, N. Rahman, M. Husain, H. Albalawi, A. H. Alfaifi and M. Sohail, Computational Insights into Structural, Elastic, Optoelectronic, and Thermodynamic Properties of Silver-based Germanium Ternary Halide Perovskites, Phys. B, 2025, 417265 CrossRef CAS.
  38. A. K. Sarychev and V. M. Shalaev, Electromagnetic field fluctuations and optical nonlinearities in metal-dielectric composites, Phys. Rep., 2000, 335, 275–371 CrossRef CAS.
  39. H. Kuzmany, The dielectric function, Solid-State Spectrosc.: An Introduction, 1998, pp. 101–120 Search PubMed.
  40. L. Dissado, Dielectric response, Springer Handb. Electron. Photonic Mater., 2017, 1 Search PubMed.
  41. S. A. Maier and H. A. Atwater, Plasmonics: Localization and guiding of electromagnetic energy in metal/dielectric structures, J. Appl. Phys., 2005, 98, 011101 CrossRef.
  42. J. Mistrik, S. Kasap, H. E. Ruda, C. Koughia and J. Singh, Optical properties of electronic materials: fundamentals and characterization, Springer Handb. Electron. Photonic Mater., 2017, 1 Search PubMed.
  43. X. Luo, Engineering Optics 2.0: A Revolution in Optical Theories, Materials, Devices and Systems, 2019, pp. 107–148 Search PubMed.
  44. H. Eisenberg, Equation for the refractive index of water, J. Chem. Phys., 1965, 43, 3887–3892 CrossRef CAS.
  45. F. A. Nelson, A. Basem, D. J. Jasim, T. E. Gber, M. T. Odey, A. F. Al Asmari and S. Islam, Chemical effect of alkaline-earth metals (Be, Mg, Ca) substitution of BFe2XH hydride perovskites for applications as hydrogen storage materials: a DFT perspective, Int. J. Hydrogen Energy, 2024, 79, 1191–1200 CrossRef CAS.
  46. W. Heller, Remarks on refractive index mixture rules, J. Phys. Chem., 1965, 69, 1123–1129 CrossRef CAS.
  47. G. D. Thayer, An improved equation for the radio refractive index of air, Radio Sci., 1974, 9, 803–807 CrossRef.
  48. E. Schreiber, O. L. Anderson, N. Soga and J. F. Bell, Elastic Constants and Their Measurement, J. Appl. Mech., 1975, 42(3), 747–748 CrossRef.
  49. O. N. Senkov and D. B. Miracle, Generalization of intrinsic ductile-to-brittle criteria by Pugh and Pettifor for materials with a cubic crystal structure, Sci. Rep., 2021, 11, 4531 CrossRef CAS PubMed.
  50. R. F. S. Hearmon, The elastic constants of anisotropic materials, Rev. Mod. Phys., 1946, 18, 409 CrossRef CAS.
  51. X. Xu, R. Huang, H. Li and Q. Huang, Determination of Poisson's ratio of rock material by changing axial stress and unloading lateral stress test, Rock Mech. Rock Eng., 2015, 48, 853–857 CrossRef.
  52. M. R. Usman, Hydrogen storage methods: Review and current status, Renewable Sustainable Energy Rev., 2022, 167, 112743 CrossRef CAS.
  53. A. Züttel, Materials for hydrogen storage, Mater. Today, 2003, 6, 24–33 CrossRef.
  54. A. Züttel, Hydrogen storage methods, Naturwissenschaften, 2004, 91, 157–172 CrossRef PubMed.
  55. L. E. Orgaz-Baque, Contribution à l'étude des hydrures des composés intermétalliques et leurs applications dans les pompes à chaleur chimiques, Dr thesis, Université Paris, 11, 1988.
  56. E. M. Dematteis, N. Berti, F. Cuevas, M. Latroche and M. Baricco, Substitutional effects in TiFe for hydrogen storage: a comprehensive review, Mater. Adv., 2021, 2, 2524–2560 RSC.

Footnote

Electronic supplementary information (ESI) available. CCDC 2469703–2469707. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d5ra03949e

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.