3.1 Geometry optimization and structural stability:
Geometry optimization refers to finding the most energetically favorable configuration or arrangement of atoms in a molecule or solid. This involves the calculation of the potential energy surface of the system, which represents the relationship between the atoms' positions and the system's potential energy. Different methods for performing geometry optimization include force-field calculations, DFT, and ab initio methods. The choice of method depends on the system's complexity and the desired accuracy level. KXH3 (X = Ca, Sc, Ti, & Ni) and most metal hydrides belong to the Pm3m space group (#221). This space group's cubic structure is almost completely packed [20]. The stability of a crystal structure is an important parameter; therefore, we have calculated the energy-volume curve, which is shown in Fig. 1.
Similarly, the thermal stability is checked with the help of phonon spectra, as shown in Fig. 2. The electronic crystal structures of KXH3 (X = Ca, Sc, Ti, & Ni) hydride perovskites are shown in Fig. 3. From Fig. 3, atoms are located at the positions: K atoms at (0, 0, 0), X (X = Ca, Sc, Ti, & Ni) atoms at (0.5, 0.5, 0.5), and H atoms at (0, 0.5, 0.5) as shown Table 1 [21]. The determined values of lattice parameters and volumes are presented in Table 1, which agree with the previously reported values. Geometry optimization and structural stability are closely related since a system's stability depends on its geometry. By performing geometry optimization calculations, one can obtain the most stable geometry of a system and predict its behavior under different conditions. For the KCaH3, KScH3, KTiH3, and KNiH3 compositions, the energy optimization curves are illustrated in Fig. 2, where Birch-equation Murnaghan's of states is used to plot the energy released during the compound's creation versus the unit cell volume. Energy-Volume graphs calculate the ground state's energy corresponding to the minimal volume, which gives information on the optimum values of lattice constants.
In our study, we thoroughly examined the stability criteria of the KXH3 (X = Ca, Sc, Ti, & Ni) hydride perovskite materials. We employed two simple and practical methods to assess the structural stability of these crystal structures. The first method, known as the tolerance factor (τ), and the second, referred to as the octahedral factor (H), provide valuable insights into the formation of stable perovskite crystal structures. For a highly symmetric and stable perovskite crystal structure to exist, these two parameters, denoted as τ and H, should fall within the ranges of (0.813–1.107) and (0.442–0.895), respectively. We calculated these crucial parameters using well-established mathematical relations [22]:
$${\tau }=\frac{{r}_{K}+{r}_{H}}{\sqrt{2}{(r}_{X}+{r}_{H})}$$
1
……………….
$$H=\frac{{r}_{x}}{{r}_{H}}$$
2
……………….
In the above equation, (1 & 2) rK, rX, and rH are represents the effective ionic radiuses of the K, X (X = Ca, Sc, Ti, Ni), and H ions. We assume hydrogen's ionic radius (RH) is 0.140 nm [23]. The calculated values of the tolerance factor are 0.822, 0.983, 0.905, and 1.031 for KCaH3, KScH3, KTiH3, and KNiH3, respectively. The calculated values of the octahedral factor are 0.714, 0.535, 0614, and 0.492 for KCaH3, KScH3, KTiH3, and KNiH3, respectively. These calculated values indicate that the crystal structure of KXH3 (X = Ca, Sc, Ti, & Ni) perovskite materials is stable. We conducted an analysis of the phonon dispersion curves for perovskite hydrides KXH3 (X = Ca, Sc, Ti, & Ni) in order to ascertain their dynamic stability. The computed graphs of phonon dispersion curves along high symmetry points within the first Brillion zone are shown in Fig. 2. As we can see, no negative frequency is present in the entire Brillion zone of calculated phonon dispersion curves. This confirms that the KXH3 (X = Ca, Sc, Ti, & Ni) compounds are also dynamically stable.
Table 1
The computed and previously reported lattice parameters, volume, and band gap of KXH3 (X = Ca, Sc, Ti, & Ni).
Compound | Lattice Parameters (Å) | Volume (Å)3 | Band gap (eV) | References |
KCaH3 | 4.482 | 92.407 | 3.31 | Present study |
KScH3 | 4.154 | 72.251 | 0.00 |
KTiH3 | 3.974 | 63.808 | 0.00 |
KNiH3 | 3.686 | 50.039 | 0.00 |
RbCaH3 | 4.532 | 93.082 | 3.32 | Experimental [24] |
LiScH3 | 3.864 | 57.699 | 0.00 | Theoretical [25] |
KTiH3 | 3.999 | 63.952 | 0.00 | Theoretical [26] |
CaNiH3 | 3.699 | 50.612 | 0.00 | Experimental [24] |
3.2 Hydrogen storage properties:
It is imperative to note that the formation energy of a compound is merely one of several factors determining its suitability for hydrogen storage. Other crucial factors encompass the kinetics of hydrogen uptake and release, the stability of the compound under operating conditions, and the capacity of the material to store hydrogen. The formation energy (ΔHf) of the suggested compounds was calculated employing the following relation [22]:
$${\Delta }{\text{H}}_{\text{f}}\left({\text{K}\text{X}\text{H}}_{3}\right)=\left[{\text{E}}_{\text{tot. }}\left({\text{K}\text{X}\text{H}}_{3}\right)-{\text{E}}_{\text{s}}\left(\text{K}\right)-{\text{E}}_{\text{s}}\left(\text{X}\right)-3{\text{E}}_{\text{s}}\left(\text{H}\right)\right]$$
3
……………….
In the above equation, Es (K), Es(X), and Es(H) represent the energy of single atom of K, X = (Ca, Sc, Ti & Ni) and H, respectively. Etotal shows the total energy of the compound, and N indicates the total number of atoms present in the compound. Our investigation revealed that all of the examined compounds possess a negative value of formation energy, indicating their thermodynamic stability and feasibility for experimental synthesis. The stability order is as follows: KNiH3 (-80.358 KJ/mol.H2) > KTiH3 (-77.437 KJ/mol.H2) > KScH3 (-67.792 KJ/mol.H2) > KCaH3 (-57.822 KJ/mol.H2). The gravimetric ratio refers to the weight of hydrogen that can be stored per unit weight of the storage material. Hydrogen storage properties can be described in gravimetric capacity, as it denotes the amount of hydrogen that can be stored per unit mass of the storage medium. This is typically measured in units of weight percent or wt%. The gravimetric capacity of hydrogen storage materials is influenced by various factors, including the type of material used, the temperature and pressure of storage, and the method of hydrogen storage. Several promising materials under development could meet or exceed this target, including metal hydrides, complex metal hydrides, and porous materials. Gravimetric hydrogen storage capabilities of KXH3 (X = Ca, Sc, Ti, & Ni) perovskite-type hydrides have been investigated by the gravimetric ratio, denoting the quantity of deposited hydrogen per unit mass of the substance, can be computed using the provided equation [22]:
$${C}_{wt\%}=\left(\frac{\frac{H}{M}{m}_{{H}_{2}}}{{m}_{Host}+\left(\frac{H}{M}\right){m}_{{H}_{2}}}\times 100\right)\%$$
4
……………….
The constituents of the given equation include \({m}_{{H}_{2}}\), signifying the molar mass of hydrogen, \({m}_{Host}\), representing the molar mass of the host material, and H/M, indicating the hydrogen-to-material atom ratio. The H/M is investigated by using the simulation package. Table 2 illustrates the investigated values of gravimetric ratios, which are 3.646 wt% for KCaH3, 3.452 wt% for KScH3, 3.346 wt% for KTiH3, and 3.005 wt% for KNiH3. The hydrogen desorption temperature of hydrides perovskite must be determined in addition to the gravimetric ratio. We can calculate the hydrogen desorption temperature by using the equation given below [27]:
$$T=-\frac{{\Delta }\text{H}}{{\Delta }\text{S}}$$
5
……………….
In the above equation, T, ΔH, and ΔS represent the desorption temperature, formation enthalpy, and change in entropy, respectively. The Standard conditions for the dehydrogenation reaction result in a change in entropy of ΔS (ΔHHydrogen = 130.7 J mol− 1 K− 1). Table 2 illustrates the calculated values of desorption temperature, which are 442.40 K, 518.68 K, 592.47 K, and 614.82 K for KCaH3, KScH3, KTiH3, and KNiH3, respectively. Notably, this temperature is still higher than the temperature at which decomposition begins for commercial use of proton exchange membrane fuel cells (PEMFC) or for the most cutting-edge automobile engines, which typically operate between 289 and 393 Kelvin and 363 and 377 Kelvin, respectively [28].
Table 2
The computed results of formation enthalpy (ΔHf), gravimetric ratio (Cwt%), and desorption temperature (Td) of KXH3 (X = Ca, Sc, Ti, & Ni)
Compound | ΔHf (KJ/mol.H2) | Cwt% (wt%) | Td (K) |
KCaH3 | -57.822 | 3.646 | 442.40 |
KScH3 | -67.792 | 3.452 | 518.68 |
KTiH3 | -77.437 | 3.346 | 592.47 |
KNiH3 | -80.358 | 3.005 | 614.82 |
3.3 Electronic Properties:
To investigate the electronic characteristics of the materials, a thorough investigation of electronic properties was undertaken, encompassing the electronic band structure (Eg), the total density of states (TDOS), and the partial density of states (PDOS) at high symmetry points within the first Brillouin zone using energy scales ranging from − 10 eV to 10 eV. Hydrogen molecules are adsorbed onto a material's surface in the physisorption method, while in the chemisorption method, hydrogen atoms are attached to the material's atoms; both methods are utilized for hydrogen storage. The strength of the interaction between the surface of the material and hydrogen molecules can be evaluated in the physisorption method by analyzing the material's electronic properties. If the density of states exhibits a large value at the Fermi level and the electronic band is near the Fermi level, the material possesses stronger hydrogen adsorption energies. In the chemisorption-based hydrogen storage method, electronic properties elucidate the bonding behavior between the material and the hydrogen atom. The conduction band (CB) and valence band (VB) positions explain the bonding behavior between the material and hydrogen atoms. If the VB and CB overlap, it suggests strong bonding energy between the material and the hydrogen atom.
Table 3
Mulliken electronic population analysis of KXH3 (X = Ca, Sc, Ti, & Ni):
Compound | Species | s | p | d | Total | Charge | Bond | Population | Bond Length (Å) |
KCaH3 | K | 2.11 | 6.31 | 0.00 | 8.420 | 0.58 | --- | --- | --- |
Ca | 2.35 | 6.00 | 0.63 | 8.97 | 1.03 | H-Ca | 0.31 | 2.24769 |
H | 1.53 | 0.00 | 0.00 | 1.53 | -0.53 | --- | --- | --- |
KScH3 | K | 2.06 | 5.62 | 0.00 | 7.67 | 0.27 | H-K | -0.36 | 2.98799 |
Sc | 2.54 | 7.03 | 1.59 | 11.16 | -0.16 | H-Sc | 0.97 | 2.11283 |
H | 1.39 | 0.00 | 0.00 | 1.39 | -0.39 | H-H | -0.04 | 2.98799 |
KTiH3 | K | 2.11 | 5.41 | 0.00 | 7.53 | 1.47 | H-K | -0.54 | 2.80099 |
Ti | 2.60 | 7.11 | 2.67 | 12.38 | -0.38 | H-Ti | 1.09 | 1.98060 |
H | 1.36 | 0.00 | 0.00 | 1.36 | -0.36 | H-H | -0.04 | 2.80099 |
KNiH3 | K | 2.49 | 4.94 | 0.00 | 7.42 | 1.58 | H-K | -0.56 | 2.60908 |
Ni | 0.82 | 1.04 | 8.80 | 10.66 | -0.66 | H-Ni | 1.06 | 1.84490 |
H | 1.31 | 0.00 | 0.00 | 1.31 | -0.31 | H-H | -0.03 | 2.60908 |
The studied materials' electronic band structures were computed by utilizing the GGA-PBE technique, as illustrates in Fig. 5. Electronic band structures suggest that KCaH3 is semiconducting material with a large band gap value of 3.310 eV, while the conduction band and valance band of KScH3, KTiH3, and KNiH3 are overlapping, which suggests that these materials are metallic and strong hydrogen bonding energy. The computed graphs of TDOS for KXH3 (X = Ca, Sc, Ti, & Ni) are shown in Fig. 6. A Vertical dashed indicates the Fermi level (EF), which is set at zero and taken as a reference point. The maximum values of TDOS at EF are 7.45, 6.13, 3.17, and 0.83 states/eV for KCaH3, KTiH3, KScH3, and KNiH3, respectively. A large value of DOS at EF shows that these materials are the metallic behavior and best candidates for hydrogen storage applications. The PDOS of the material gives the information of the electronic state in a solid material at any point of energy level. A solid's electronic structure is defined by its energy bands comprising multiple electronic states. The curves help to investigate the involvement of certain atoms or orbitals to such energy bands and also examine the bonding behavior on these curves, giving information regarding hybridization between states.
The calculated graphs of PDOS for KXH3 (X = Ca, Sc, Ti, & Ni) are represented in Fig. 7. From Fig. 3, all examined materials have flat-going core states, mostly involving the f-orbital, which are ignored. From − 10 to -5 eV energy, the s-state has shown a small contribution. From − 5 to 0 eV, the energy s-state of KXH3 (X = Sc, Ti & Ni) and the d-state of the d-state of KNiH3 show maximum contribution. At the fermi level, the s-state of KCaH3, d-state of KScH3, and KTiH3 have shown maximum values, which show strong hydrogen bonding energy. Mulliken atomic population analysis is the computational technique utilized to investigate the bonding behavior, length of bonds, and electronic structures of solids and molecules. The Mulliken analysis gives information about the electron density distribution in the material and can help predict the intensity of the material's interaction with hydrogen molecules. The + ve value of the population shows the nature of the covalent bonding, and the negative value shows the nature of the ionic bonding of the material. Mulliken atomic population analysis for KXH3 (X = Ca, Sc, Ti, & Ni) is calculated and presented in Table 3. Furthermore, by population analysis, we can also determine the population ionicity (Pi), which gives information about the "percentage of the covalence behavior of the bond," which can be calculated by utilizing the given formula [29]:
$${\text{P}}_{i}=1-{e}^{-\left|\frac{{P}_{c}-P}{P}\right|}$$
6
……………….
The above equation shows all the studied materials show covalent bonding behavior. In addition, the bond lengths for H-Sc, H-Ti, and H-Ni in KScH3, KTiH3, and KNiH3 are 2.11283 Å, 1.98060 Å, and 1.84490 Å, respectively. The electronic structure of KXH3 (X = Ca, Sc, Ti, & Ni) perovskite materials enable strong interactions between the hydrogen molecules and the host lattice. A weak chemical bond between the hydrogen atoms and the nearby atoms in the lattice can result in hydrogen storage. This procedure improves hydrogen storage or absorption due to the expansion of interstitial sites. These compounds have the potential to undergo a chemical reaction with hydrogen, resulting in the formation of intermetallics that offer improved storage capabilities.
3.4 Mechanical Properties:
Hydrogen storage materials are employed to store and release hydrogen for diverse applications, including hydrogen-powered vehicles, fuel cells, and energy storage systems. Mechanical stability emerges as a crucial parameter for H2-storage materials, as it determines their durability and safety during operation. Mechanical stability refers to the ability of a material to resist deformation or fracture under mechanical stress or strain. In the context of hydrogen storage materials, this encompasses resistance to pressure changes, temperature fluctuations, and repeated hydrogen absorption and desorption cycles. To ensure mechanical stability, hydrogen storage materials are often engineered to exhibit high strength, toughness, and resistance to fatigue and corrosion. Materials such as metal hydrides, which can absorb and release substantial amounts of hydrogen, are frequently reinforced with other materials to enhance their mechanical properties. Through the evaluation of mechanical properties, we can discern the strength and bonding behavior of the crystal structure.
Table 4
The computed results of elastic constants (Cij) and Cauchy’s pressure (CP) of KXH3 (X = Ca, Sc, Ti, & Ni):
Compound | C11 | C12 | C44 | CP |
KCaH3 | 42.080 | 7.439 | 16.893 | -9.454 |
KScH3 | 36.349 | 7.362 | 20.286 | -12.924 |
KTiH3 | 70.994 | 19.896 | 35.254 | -15.358 |
KNiH3 | 59.339 | 24.124 | 52.302 | -28.178 |
Table 5
The computed mechanical parameters Bulk modulus (B), Shear modulus (G), Young modulus (E), Pugh ratio (B/G), Pugh’s modulus (G/B), Poisson ratio (v), Material’s Hardness (Hv), Machinability index (µM), and Anisotropic index (AU):
Compound | B | G | E | B/G | G/B | V | HV | µM | AU |
KCaH3 | 18.985 | 17.062 | 39.388 | 1.113 | 0.899 | 0.155 | 13.903 | 1.124 | 0.975 |
KScH3 | 17.025 | 17.729 | 39.482 | 0.961 | 1.042 | 0.114 | 17.081 | 0.840 | 1.401 |
KTiH3 | 36.929 | 30.988 | 72.645 | 1.192 | 0.840 | 0.173 | 23.355 | 1.048 | 1.380 |
KNiH3 | 35.862 | 33.836 | 77.223 | 1.060 | 0.944 | 0.142 | 28.137 | 0.686 | 2.970 |
The microstructure and crystal structure of the material can also influence its mechanical stability. For instance, materials with a highly ordered crystal structure, such as zeolites, can exhibit better mechanical stability than those with a more disordered structure. The prospective mechanical parameters and anisotropic factor of all studied hydride perovskite materials have been investigated. These mechanical properties can be identified by utilizing the stiffness constants. Stiffness constants like C11, C12, and C44 are reduced to only three for a cubic phase but vary widely across other crystal structures. C11, C12, and C44 were investigated using the CASTEP simulation code. The three elastic constants, namely C11, C12, and C44, correspond to the material's resistance to longitudinal deformation, transverse expansion, and hardness, respectively. For mechanical stability, the elastic constants must fulfill the Born stability criteria, which are given as follows:
C11 + 2C12 > 0; C11 − C12 > 0; C11 > 0; C44 > 0 ………………. (7)
Table 4 shows that the calculated values of elastic constants for KXH3 (X = Ca, Sc, Ti, & Ni) fulfill the abovementioned requirement, demonstrating the mechanical stability of the materials.
Cauchy’s pressure (CP) can be computed as:
CP = C12 - C44 ………………. (8)
By using the VRH technique, B is computed as:
$${B}_{V}=\frac{{C}_{11}+2{C}_{12}}{3}$$
9
……………….
$${B}_{R}=\frac{{C}_{11}+2{C}_{12}}{3}$$
10
……………….
$$B=\frac{{B}_{V}+{B}_{R}}{2}$$
11
……………….
The following relation can determine Young's modulus:
……………….
Shear modulus can be computed by the equation given below:
$$G=\frac{{G}_{R}+{G}_{V}}{2}$$
13
……………….
Where Gv and Gr are:
$${G}_{V}=\frac{1}{3}\left(3{C}_{44}+{C}_{11}-{C}_{12}\right)$$
14
……………….
$${G}_{R}=\frac{5\left({C}_{11}-{C}_{12}\right){C}_{44}}{3\left({C}_{11}-{C}_{12}\right)+4{C}_{44}}$$
15
……………….
$$v=\frac{3B-2G}{3(3B+G)}$$
16
……………….
The computed values of Cauchy’s pressure for KXH3 (X = Ca, Sc, Ti, & Ni) are represented in Table 4. The negative values illuminate each material's brittleness and angular bonding. The highest bulk modulus value indicates the material is most resistant to volume change. The calculated Bulk modulus values for KCaH3, KScH3, KTiH3, and KNiH3 are 18.985, 17.025, 36.929, and 35.862, respectively. This shows that KTiH3 and KNiH3 are the most resistant to volume change. The computed results of Bulk modulus (B) are represented in Table 5. Shear modulus indicates the material’s resistance toward shape change under applied stress.
The modulus of rigidity is a mechanical characteristic that offers insight into the hardness properties of a material. Pugh’s ratio (B/G) is a significant parameter that more comprehensively elucidates the ductility and brittleness characteristics of a solid material. B/G is the ratio of the material between resistance to fracture and deformation. In more specific terms, the component B pertains to the fracture resistance of the material, while G pertains to its deformation resistance. The literature commonly recognizes a critical threshold of 1.75 as the standard value for distinguishing between the brittleness and ductility characteristics of materials. The B/G value exceeding 1.75 signifies the material's ductile behavior, whereas its value falling below 1.75 implies excessively brittle behavior. Table 5 shows that all materials are less brittle because their B/G values fall below the critical value (1.75). The Poisson ratio is another significant parameter in assessing the plastic properties as well as the brittleness and ductility behavior of solid materials.
In the evaluation of a material's ductility, the Poisson ratio serves as a crucial parameter, with a value exceeding 0.25 indicating ductility while a value below 0.25 indicates potentially brittle behavior [22]. All the studied materials show brittle behavior because their calculated values of Poisson ratio are less than the critical values (0.25), as shown in Table 5. The Poisson ratio is a potential parameter for identifying the ionic or covalent bonding of a material. The critical values for covalent and ionic behavior are 0.1 and 0.25, respectively. The Poisson ratio can provide insights into the nature of a material's bonding, with a value in proximity to 0.1 indicative of covalent bonding while a value near 0.25 suggestive of ionic behavior [22].
From Table 5, it can be seen that the studied materials show covalent behavior. Pugh’s modulus is calculated by the G/B ratio, which is also used to determine the bonding behavior of the material. A critical value for ionic and covalent behavior through Pugh’s modulus (G/B) is 0.6 and 1.1, respectively. If the value of Pugh’s modulus is near about 0.6, the material has ionic bonding. If the value of Pugh's modulus is near about 1.1, the material has covalent bonding. From Table 5, it can be seen that all the studied materials have covalent bonding [22].
$${A}^{U}=\frac{{2C}_{44}}{{C}_{11}-{C}_{12}}$$
17
……………….
Isotropic crystals have a value of A = 1, whereas anisotropic crystals have 1 < A < 1. The computed values of A are 0.975, 1.401, 1.380, and 2.970 for KCaH3, KScH3, KTiH3, and KNiH3, respectively—the computed values of A confirming the anisotropic behavior of studied materials.
3.5 Thermodynamic Properties:
The assessment of thermodynamic properties, including longitudinal, transverse, and average velocities, Debye temperature, and melting temperature, is crucial for understanding the hydrogen storage capabilities of materials. Higher Debye temperatures indicate that the solid can better store hydrogen at high temperatures and needs more energy to disrupt. Melting temperature is also related to the Debye temperature. If the Debye temperature is very high, then that material's melting temperature (Tm) and thermal conductivity must also be high. The determined thermodynamic parameters are listed in Table 6.
Table 6
The computed results of density (ρ), longitudinal velocity (vl), transverse velocity (vt), average sound velocity (vm), Debye temperature (θD), and melting temperature (Tm) of KXH3 (X = Ca, Sc, Ti, & Ni):
Compound | ρ (g/cm3) | vt (km/s) | vm (km/s) | θD (K) | Tm (K) |
KCaH3 | 1.472 | 3.407 | 3.493 | 405.39 | 505.57 ± 300 |
KScH3 | 2.001 | 2.977 | 3.023 | 381.46 | 493.47 ± 300 |
KTiH3 | 2.342 | 3.638 | 3.746 | 492.65 | 619.86 ± 300 |
KNiH3 | 3.346 | 3.180 | 3.250 | 463.57 | 610.47 ± 300 |
Thermodynamic parameters can be evaluated by using the following equations.
Debye temperature \({{\theta }}_{\text{D}}\) can be evaluated [30].
$${{\theta }}_{\text{D}}=\frac{ħ}{{k}_{B}}{\left[\frac{3n{N}_{a}\rho }{4\pi M}\right]}^{\frac{1}{3}} \times {v}_{m}$$
18
……………….
The quantities denoted as \(ħ\), \({k}_{B}\), \(n\), \({v}_{m}, {N}_{a}, \text{a}\text{n}\text{d} V\) represent fundamental physical parameters, namely Planck’s constant, Boltzmann’s constant, number of atoms, average velocity of sound, Avogadro number, and unit cell volume, respectively.
Furthermore, \({v}_{m}\) can be evaluated by using the following relation [30]:
$${v}_{m}=\frac{1}{3}{\left[\frac{2}{{v}_{t}^{3}}+\frac{1}{{v}_{l}^{3}}\right]}^{\raisebox{1ex}{$-1$}\!\left/ \!\raisebox{-1ex}{$3$}\right.}$$
19
……………….
Here, \({v}_{t}\) and \({v}_{l}\) can be evaluated by using the given relations [30]:
$${v}_{t}={\left[\frac{G}{\rho }\right]}^{\raisebox{1ex}{$1$}\!\left/ \!\raisebox{-1ex}{$2$}\right.}\text{a}\text{n}\text{d} {v}_{l}={\left[\frac{3B+4G}{3\rho }\right]}^{\raisebox{1ex}{$1$}\!\left/ \!\raisebox{-1ex}{$2$}\right.}$$
20
……………….
The determined Debye temperatures for KCaH3, KScH3, KTiH3, and KNiH3 were found to be 405.39 K, 381.46 K, 492.65 K, and 463.57 K, respectively. The melting temperature, which is also referred to as the melting point, denotes the temperature at which a solid material undergoes a phase transition from its solid state to a liquid state. This transition is marked by a state of equilibrium where both the solid and liquid phases of the substance coexist. We can evaluate the melting point of solid material by using the following relation [30]:
Tm(K) = [553(K) + 5.911 (C12)] GPa ± 300K ………………. (21)
The melting temperature values for KCaH3, KScH3, KTiH3, and KNiH3 are 505.57 K, 493.47 K, 619.86 K, and 610.47 K, respectively. The thermodynamic parameters were determined by utilizing the elastic constants (C11, C12, C44), and the corresponding calculated values are presented in Table 6.
3.6 Optical Properties:
The optical properties of a material are studied to learn more about how light energy interacts with it. Optical properties of the material such as absorption, reflection, conductivity, energy loss function, refractive index, and complex dielectric functions are important parameters of the material used in a couple of applications like hydrogen storage applications, photocatalytic applications, coatings, solar cell devices, and optoelectronic applications. All the optical parameters are based on the complex dielectric function, which can be computed by using the following relation [31]:
ε(ω) = ε1(ω) + iε2(ω) ………………. (22)
The real component ε1(ω) of the dielectric function ε(ω) may be obtained from the imaginary component ε2(ω) using the Kramer-Kronig relation, and the imaginary component can be obtained by adding an extensive number of unoccupied states. The optical parameters could be calculated precisely by examining the dielectric function and considering the electronic transitions. A detailed description of real part ε1(ω) and imaginary part ε2(ω) of complex dielectric function against photon energy range 0 to 20 eV is shown in Fig. 8 (a & b). The real part ε1(ω) of the dielectric function measures the ability of the material to store the electric charge and also explains the dispersion effects that occur inside a material. Hydrogen storage could be improved by a greater value of ε1(ω) because it can provide stronger attractive forces between the hydrogen and the relevant material. The static values of zero photon energy of εo(ω) are 45.23, 34.74, 03.80, and 03.19 for KScH3, KTiH3, KNiH3, and KCaH3, respectively. Then ε1(ω) decreases sharply to zero at 0.78 eV for KScH3, 5.65 eV for KNiH3, 7.11 eV for KCaH3, and 11.14 eV for KNiH3. The negative values of ε1(ω) indicate the metallic nature of the material. The highest value of KScH3 and KTiH3 shows that stored energy in these materials can be utilized for useful purposes. It also suggests that stored energy can be used for optoelectronic applications. ε2(ω) explains the adsorptive behavior of the material to incident photons. The maximum calculated values of ε2(ω) are 22.39 at 0.31 eV, 11.62 at 0.18 eV, 5.22 at 6.89 eV, and 3.33 at 9.08 eV for KScH3, KTiH3, KCaH3, and KNiH3, respectively. The refractive index describes the material's ability to absorb light at a certain wavelength and the material's transparency to the incident photon. The maximum values of the refractive index n(ω) are 2.78 at 4.36 eV, 2.41 at 5.00 eV, 2.19 at 3.31 eV, and 2.17 at 3.70 eV for KTiH3, KScH3, KCaH3, and KNiH3, respectively. The maximum values of the extinction coefficient are 2.79, 1.71, 1.58, and 1.08 for KScH3, KTiH3, KCaH3, and KNiH3, respectively. The calculated graph of the refractive index and extinction coefficient as a function of photon energy is shown in Fig. 8(c & d).
The rate at which hydrogen may be absorbed into the material and dispersed through it is referred to as the absorption coefficient, also known as the hydrogen diffusion coefficient or hydrogen permeability. The calculated graphs of absorption coefficient α(ω) for KXH3 (X = Ca, Sc, Ti, & Ni) hydride perovskite drawn against 0 to 20 eV photon energy are shown in Fig. 9(a). It is analyzed that all of the materials show zero absorption when no photons hit the surface of the compound. The absorption rate in the materials increases by increasing the photon energy. A high absorption coefficient is preferred for effective hydrogen storage since it allows for fast absorption of hydrogen and a large amount of storage. The composition and crystal structure of hydride perovskite materials are the variables that affect the absorption coefficient. The determined peaks values of absorption coefficient are 20.01×104 cm− 1, 18.84×104 cm− 1, 16.81×104 cm− 1, and 15.28×104 cm− 1 for KNiH3, KCaH3, KTiH3, and KScH3, respectively. All studied materials' α(ω) becomes zero at a high photon energy range. Optical conductivity is used to evaluate the mechanism of conduction according to the photoelectric effect, which occurs when high-energy photons (E = ħω) hit a material's surface and cause photoelectron emission. The material's conductivity in hydrogen storage determines how readily it can pass through. Also, studying the breaking of bonds that occur due to incoming radiations with the material's surface is helpful. In hydride perovskite materials, hydrogen diffusion plays a crucial role in determining the rates of hydrogen absorption and release. A substance with a high conductivity may absorb and release hydrogen more quickly because hydrogen can diffuse through it more quickly and effectively.
As a result, materials with high conductivity are chosen for applications that require effective hydrogen storage. For this purpose, the conductivity of KXH3 (X = Ca, Sc, Ti, & Ni) hydrides perovskite has been determined against the photon energy from the range 0 to 20 eV, as shown in Fig. 9(b). the calculated peak values of conductivity are 4.37 at 6.97 eV for KCaH3, 4.28 at 5.03 eV for KTiH3, 3.84 at 11.01 eV for KNiH3, and 3.71 at 5.03 eV for KScH3. So, KScH3 and KTiH3, both compounds, predict high conductivity and good material for hydrogen storage applications. The reflectivity is determined to investigate the material's behavior with the interaction of incident radiations. Some light that strikes a substance is absorbed, while the remaining is reflected. The proportion of light reflected and the amount of light that incident determines a material's reflectivity.
Reflectivity will affect the temperature of hydride perovskites when exposed to light. The material may heat up due to light absorption, promoting hydrogen desorption. Conversely, the material may stay colder if it reflects light, which can sometimes decrease hydrogen desorption. The maximum reflectivity peaks for KXH3 (Ca, Sc, & Ti) are observed in the energy range of 5 eV to 8 eV, while the maximum peak of KNiH3 is observed at 11.61 eV. The calculated graphs of reflectivity against photon energy for KXH3 (X = Ca, Sc, Ti, & Ni) are presented in Fig. 9 (c). The energy loss function is a mathematical expression that describes the amount of energy lost during the transition of electrons due to scattering or dispersion. The loss function may affect the interaction between hydride ions and metal cations in hydride perovskites, affecting the material's electronic structure. The energy loss function is proportional to the scattering probabilities during inner shell transitions. Hydrogen transport and binding within a material can be modified by the loss function, which will influence the electronic density of states. The calculated graphs of the energy loss function of KXH3 (X = Ca, Sc, Ti, & Ni) hydride perovskites against photon energy in the range of 0 to 20 eV are shown in Fig. 9(d). The maximum values of energy loss functions are 1.91 at 12.16 eV, 1.37 at 15.09 eV, 1.19 at 13.56 eV, and 0.84 at 15.40 eV for KCaH3, KTiH3, KScH3, and KNiH3, respectively.