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

Structural stabilities, robust half-metallicity, magnetic anisotropy, and thermoelectric performance of the pristine/Ir-doped Sr2CaOsO6: strain modulations

Samia Shahzadia, Ihab Mohamed Moussab, Sohail Mumtaz*c and S. Nazir*a
aDepartment of Physics, University of Sargodha, 40100 Sargodha, Pakistan. E-mail: safdar.nazir@uos.edu.pk; Tel: +92 334 9719060
bDepartment of Botany and Microbiology, College of Science, King Saud University, P.O. Box 2455, Riyadh, 11451, Saudi Arabia
cDepartment of Chemical and Biological Engineering, Gachon University, 1342 Seongnamdaero, Sujeong-gu, Seongnam-si 13120, Republic of Korea. E-mail: sohail.ahmed2015@gmail.com

Received 9th April 2025 , Accepted 22nd April 2025

First published on 21st May 2025


Abstract

Half-metallic (HM) ferromagnetic (FM)/ferrimagnetic (FIM) materials with a large energy-gap (Eg) and high magnetocrystalline anisotropy energy (MAE) are receiving consideration for their potential usage in solid-state electronic devices. This study explores various traits of the pristine (prs.)/Ir-doped (dop.) Sr2CaOsO6 structure using ab initio calculations, where Ir is doped at the Os-site. To determine the synthesis feasibility of the structures under ambient conditions, the formation energy, elastic constants, and phonon curves are determined. The prs. structure manifests a FM semiconducting nature with an Eg of 0.048 eV. Strikingly, the Ir-dop. structure becomes HM FIM because additional electrons provided by the dopant (Ir) cause a repulsion in the Os t22g spin-minority channel, resulting in conductivity. Conversely, an Eg of 1.15 eV in the spin-majority channel exists, which is high enough to keep the HM state stable. The computed partial spin-moment on the Os in the prs. system is 1.19 μB. In the Ir-dop. system it is 1.09/−1.39 μB on the Os/Ir ion holding an Os+6/Ir+4 state with electronic distributions of 5d2(t22g↑t02g↓e0g↑e0g)/5d5(t32g↑t22g↓e0g↑e0g) with image file: d5ra02453f-t1.tif. Further, the spin-magnetization density isosurfaces assist in determining the ms values and FM/FIM state of the prs./Ir-dop. system holding a Curie temperature (TC) of 185/171 K. Besides this, we computed the thermoelectric properties of the prs./Ir-dop. motifs; the figure of merit (0.33/0.02), Seebeck coefficient (147/30 μV K−1), and low thermal conductivity (0.21/0.71 × 1019 Ωm−1 s−1) at 300 K highlight their potential for conversion devices. Interestingly, a semiconducting-to-HM transition is predicted at a crucial compressive strain of −3% in the prs. structure. Conversely, the HM state in the dop. structure displays robustness against strain. Additionally, it is shown that an applied tensile strain can significantly improve ZT, while compressive strains illustrate a positive impact on the TC value.


1 Introduction

Depending on the B and B′ cations, double perovskite oxides (DPOs) display a wide range of ferromagnetic (FM), ferrimagnetic (FIM), and anti-ferromagnetic (AFM) spin ordering (SO) in addition to half-metal (HM)1,2 and semiconducting (SC)/insulating phases,3–5 and an insulator-to-HM transition.6 Materials research continues to focus on the insulator-to-metal transitions (IMTs) because of the conceptual gaps that lead to an inability to regulate the electronic conduction in prospective magnetic memory devices.7 Likewise, structure modifications in the crystals allow the material to transform from a band insulator to Mott insulator (MI).8 In particular, DPOs based on the 5d transition metals (TMs) provide an ideal environment for orbital, charge, and lattice degrees of freedom. In this respect, osmium-based materials have attracted a lot of interest due to the successful synthesis process of the oxides, which generates a variety of unusual phases. For example, a MI FM state in Ba2NaOsO6 (ref. 9) and a magnetic insulating state in Sr2MOsO6 (where M = Cu/Ni)10 are predicted. Interestingly, the HM FIM state has been discovered with an energy gap (Eg) of 0.40 eV in the spin-minority channel (N) of Sr2CuOsO6,11 where the spin-majority channel (N) is metallic. The combined effect of electron correlation and spin–orbit coupling (SOC) results in a FIM MI character of the Ca2FeOsO6 structure.12 Similarly, a half semi-metallic FIM phase in Sr2CrOsO6 is theoretically predicted within the GGA+U+SOC method13 owing to a FIM MI state with a massive TC (Curie temperature (temp.)) of 725 K.14 Moreover, the Sr2CrOsO6 system displays a compensatory HM behavior without the influence of SOC. However, when SOC is included, a net magnetic moment of 0.54 μB is obtained,13 which exposes that the dominating influence of SOC cannot be properly attributed to the magnetic moment correction.

Systems with high TC and HM FM/FIM stable ground states are in huge demand for the fields of spintronics,15–17 magnetoelectronics,18 magnetodielectric capacitors,19,20 and data storage devices.21 It has been discovered that enhancing the antisite disorder deficiencies and grain boundaries may improve the physical aspects of DPOs.22,23 Hence, the d-orbital occupancy can be effectively regulated by doping the B/B′ site with various TMs to significantly improve magnetic exchange interactions between them, which often results in high-TC HM systems.24,25 The TC for Sr2CrOsO6 jumps dramatically from 490 to 660 K when the larger Sr cation is substituted at the Ca-site.26–28 Furthermore, the inclusion of disordered BO6 and B′O6 octahedra accumulating irregularly leads to a crystal field splitting and enhances the unusual electronic and magnetic features.29 The structural distortions are reduced when Ba is substituted at the Sr-site in ALaNiOsO6 (A = Sr, Ba) and the Weiss temp. (θw) changes from negative to positive.30 Experimental observations reveal that doping of Ni2+ at the Fe-site in Sr2Fe1.5Mo0.5O6−δ leads to competition between Fe3+/Mo5+ and Fe2+/Mo4+ along with induced conductivity.31 By adopting a similar strategy, Bhandari et al.,32 utilizing density functional theory (DFT) calculations, predicted that the system enters into an HM state when Ni2+ is doped (dop.) at the Cr3+ site (electron doping) in the FIM MI Ca2CrOsO6 system. In addition to doping of various elements at the A or B site in A2BB'O6, the doping of osmium at the B′ site in Sr2CrReO6 significantly increases the TC and causes an HM-to-insulator transition.33 However, partial magnetizations in Sr2CrReO6 at the Re-site dramatically decrease at a 40% W-amount and a drop in TC is also observed with an increase in dopant (W) concentration.34

In the same way, the strain approach is one of the most effective means of adjusting or controlling a material’s traits by altering the lattice parameters under various growth circumstances.35,36 For example, when a Sr2FeMoO6 film is formed on a SrTiO3 substrate, lattice mismatch causes compressive (comp.) strain of −1.2%, leading to an HM FM state.36,37 Furthermore, at −8% comp. and +1% to +5% tensile (tens.) strains, a FM SC to metal and HM transition is verified in La2FeMnO6, respectively.38 Likewise, in the Lu2NiIrO6 DPO, a FIM SC to HM transition is noted under hydrostatic stress of −6% and biaxial strain of −8%.39 Besides this, experimental observations reveal that lattice mismatch occurs when Re2NiMnO6 thin films (where Re = La, Pr, Nd, Sm, and Y) grow on a LaAlO3 substrate; the B(B′)O6 octahedra are further deformed by induced strain, which improves the films’ magnetic qualities and serves as an effective approach to get the maximum TC in DPOs.40 Interestingly, a magnetic transition from the FM SC to the HM FM state in La2FeMnO6 is demonstrated by DFT investigations with the application of a biaxial ([110]) strain ranging from 0% to +10%.38 Moreover, under volume compression, a HM state with an extremely high TC of 624 K is predicted in the Ca2MoOsO6 structure.41

The above-mentioned considerations make it abundantly evident that B′-site doping in DPOs is also a suitable method for altering physical aspects. Thus, we theoretically study the effect of Ir-doping at the Os-site on the distinct traits of FM SC Sr2CaOsO6 (SCOO). Because of the partial filling of the 5d orbitals of the dopant (Ir) and the host (Os), a SC-to-metal transition (MT) is observed in the Ir-doped (dop.) structure. Each system shows an appropriate amplitude of magnetic anisotropy energy (MAE), which improves the system’s functionality for the data storage devices.21 Along with this, we do a thorough examination of the material's thermoelectric (TE) qualities for its use in renewable energy devices. Additionally, detailed investigations of the biaxial ([110]) strain effect on the physical behavior of the systems were conducted, which give interesting outputs.

2 Computational and structural details

A full-potential linearized augmented plane-wave approach based on spin-polarized (SP) DFT, as implemented in the WIEN2K code,42 was employed for the current calculations. The exchange–correlation functional utilized in this study, which is derived from the generalized gradient approximation (GGA) combined with on-site Coulomb interaction (GGA+U), is used, keeping U as 2.8/2.6 eV on the 5d states of the Ir/Os ion.43 Additionally, because of the heavy Os/Ir element, SOC effects are also employed in the scalar relativistic form. In the wave-function expansion within the atomic spheres, łmax = 12, Rmt × Kmax = 7, and Gmax = 24 within the irreducible wedge of the Brillouin zone is taken. A 6 × 6 × 4 k-mesh with 76 points is found to be highly converged. Also, the atomic positions of the ions are fully relaxed by lowering the overall forces below 5 mRy a.u.−1. Self-consistency of the system is assumed for a total energy (Et)/charge convergence to 10−5 Ry/10−5 C. Moreover, Boltzmann's theory is used to compute the TE parameters using the relaxation-time approximation, as defined in the BoltzTrap code.44

The monoclinic SCOO keeps CaO6 and OsO6 octahedra alternately organized in a rock-salt pattern, providing a full 1[thin space (1/6-em)]:[thin space (1/6-em)]1 ordering with space group no. 14 (P21/n). The experimental lattice constants are a = 5.7643, b = 5.8191, and c = 8.1796 Å with β = 90.22°.45 In the primitive unit cell of SCOO, 4/2/2/12 Sr/Ca/Os/O ions exist. The atomic coordinates for the Sr, Ca, Os, O1, O2, and O3 are (0.0104, 0.0334, 0.2490), (0.5, 0, 0), (0.5, 0, 0.5), (0.265, 0.312, 0.028), (−0.186, 0.234, −0.038), and (−0.088, 0.479, 0.227), correspondingly. The Ir-dop. SCOO structure is designed by replacing one of the Os ions with one Ir in its primitive unit, as reported in the previous experimental46,47 and theoretical5,48 works, where it changes the P21/n to P1 structural symmetry. The crystal structure of the prs./Ir-dop. motif is presented in Fig. 1(a) and (b).


image file: d5ra02453f-f1.tif
Fig. 1 Crystal diagram of the (a) prs. and (b) Ir-dop. Sr2CaOsO6 structures.

3 Results and discussion

3.1 Unstrained systems

First, the thermodynamic stability of the structures is examined by determining the formation enthalpy (ΔHf) as follows:
 
image file: d5ra02453f-t2.tif(1)
 
image file: d5ra02453f-t3.tif(2)
image file: d5ra02453f-t4.tif represent the Et of the prs./Ir-dop. SCOO, Sr (Fm[3 with combining macron]m-225), Ca (Fm[3 with combining macron]m-225), Os (P63/mmc-194) and Ir (Fm[3 with combining macron]m-225) atoms and oxygen (C2/m-12) molecule in their respective ground states, correspondingly. To prevent the overestimation of the ΔHf value, the suggested adjustments by Wang et al.,49 are also considered. The estimated ΔHf of the prs./Ir-dop. structure is −29.25/−25.85 eV, where the “−” sign ensures the thermodynamic stability of the system. Next, to determine the mechanical stability of both structures, we calculated the elastic tensors (Cij) by generating six finite constants using conventional strain and stress relations.50,51 The computed 13 independent elastic stiffness tensors for the monoclinic systems are listed in Table 1, which fulfilled the basic requirements and Born stability criteria52 for mechanical stability. Additionally, Fig. 1S of the ESI displays the bulk modulus (B), shear modulus (G), and Young modulus (Y) of both systems. B measures a compound's ability to deform in response to the pressure surrounding its surface, commonly known as hardness. Along with this, a higher value of G suggests that the material is stiffer and less likely to deform under shear forces, whereas lower values indicate that it is more flexible or ductile. Y is also called the modulus of elasticity and is a measure of a material's stiffness. A high value means that the materials are stiff and resistant to deformation under stress, while lower denotes a more elastic and easily deformed material. Hence, the B, G, and Y values of the prs./Ir-dop. system are 129/128, 145/142, and 55/54, respectively. It shows that the prs. system is stiffer than the doped one. Furthermore, we computed Pugh's ratio (B/G), Cauchy's pressure (CP), and Poisson's ratio (ν), which illustrate whether a compound is ductile or brittle,53 as displayed in Fig. 2. If the B/G value is less than 1.75, ν is less than 0.25, and the CP value is less than 0, this reveals a brittle nature. In contrast, if the B/G value is greater than 1.75, ν is greater than 0.25, and the CP value is greater than 0, this results in a ductile nature.54 Hence, our computed values of the B/G (see Fig. 2(a)), CP (see Fig. 2(b)) and ν (see Fig. 2(c)) for both structures demonstrate that they are ductile. Finally, phonon calculations were performed for both structures to confirm their dynamical stability. The determined phonon spectra for the prs./Ir-dop. systems are plotted in Fig. 3(a) and (b). In principle, each atom typically contributes three phonon branches with the total number of phonon branches in a unit cell being precisely proportional to the three times number of atoms. There are 3n total branches, which include 3n − 3 optical modes and 3 acoustic modes.55 Both structures contain 20 atoms in their basic unit cell and have 60 vibrational modes including 3 acoustic modes along with 57 optical modes. The phonon dispersion curves do not contain any negative frequencies for both structures (see Fig. 3), indicating that they are dynamically stable as well.
Table 1 Computed 13 independent elastic constants (Cij) of the prs. and Ir-dop. Sr2CaOsO6 structures under −5%/0% (unstrained)/+5% biaxial ([110]) strain
Systems Strain C11 C12 C13 C15 C22 C23 C25 C33 C35 C44 C46 C55 C66
prs. −5% 203.70 144.29 164.91 6.71 204.85 162.51 16.06 258.55 −4.48 69.14 −4.86 62.03 101.73
  0% 193.76 113.58 90.42 3.94 188.95 78.33 −3.98 222.13 4.44 59.37 −6.03 49.71 67.45
  +5% 144.22 68.45 52.47 −0.95 128.78 56.99 −2.54 183.14 0.98 46.39 −2.03 32.00 45.07
Ir-dop. −5% 201.35 144.26 165.27 5.13 200.23 162 14.39 252.84 −6.38 66.81 −3.98 60.87 102.21
  0% 190.53 113.98 90.20 4.09 183.40 75.84 −3.81 222.68 3.36 57.46 −5.58 47.99 68.69
  +5% 189.78 94.99 76.84 7.37 149.68 96.63 14.68 187.91 −17.41 51.62 −4.55 25.27 67.67



image file: d5ra02453f-f2.tif
Fig. 2 Computed (a) Pugh's ratio (B/G), (b) Cauchy's pressure (CP), and (c) Poisson’s ratio (ν) in the prs./Ir-dop. Sr2CaOsO6 structure for −5% (compressive)/0% (unstrained)/+5% (tensile) strain.

image file: d5ra02453f-f3.tif
Fig. 3 Computed phonon dispersion curves for the (a) prs. and (b) Ir-dop. Sr2CaOsO6 structures.

To find the magnetic ground state (MGS) of the prs. system, we plotted the computed Et for the non-magnetic (NM), FM, and AFM MGS spin ordering (SO) in Fig. 2S of the ESI within the GGA+U/GGA+U+SOC scheme. In the case of the NM phase, both Os ions remain non-SP. In contrast, they are aligned (↑↑)/anti-aligned (↑↓) for the FM/FIM SO. It is demonstrated that the Et of the FM SO is lower than that of the NM/AFM state in both methods (see Fig. 2S of the ESI). However, it is experimentally observed that the system persists in a paramagnetic state for 600 K to 2 K temp. and there is no magnetic phase transition occurring.45 Additionally, the computed Et values of the FM and FIM SO are compared to determine the MGS for the Ir-dop. SCOO DPO. In the case of the FM/FIM SO, Os and Ir ions, the spins remain parallel (image file: d5ra02453f-u1.tif)/anti-parallel (image file: d5ra02453f-u2.tif) to each other, where the small and large arrow lengths represent their respective spin magnitudes. It is established that the FIM state is more stable than the FM one with an energy difference of ΔE = EFIMEFM = −96 meV (see Table 2). This indicates that Ir and Os ion spins favor anti-alignment in both the in-plane and out-of-plane orientations. Hence, for further investigations only FM/FIM SO is considered for the prs./Ir-dop. motif.

Table 2 Computed total energy difference (ΔE), energy gap in the spin majority/minority channel (EN↑g/EN↓g), total/partial spin/orbital moment (mt/ms/morb.) in μB, magnetocrystalline anisotropy energy (MAE) in meV, Curie temperature (TC) and MAE constant (K) per unit volume (×107 erg cm−3) within the GGA+U/GGA+U+SOC method for the prs. and Ir-dop. Sr2CaOsO6 structures. The symbol M symbol represents the metal
Property GGA+U GGA+U+SOC
prs. Ir-dop. prs. Ir-dop.
ΔE −51 −131.8 −96.4
EN↑g 0.048 1.15 0.004 M
EN↓g 1.86 M
MAE 2.95 0.79
K 1.72 0.46
TC 185 171
mt 4.0 1.0 3.84 0.91
Os1 Os2 Os Ir Os1 Os2 Os Ir
ms 1.19 1.19 1.01 −1.39 1.09 1.09 0.89 1.25
morb. −0.5 −0.5 0.44 0.083


Next, to examine the electronic structure of the prs./Ir-dop. SCOO system, we computed the total density of states (TDOS) in the stable FM/FIM SO within the GGA+U method. Fig. 4(a) shows that the prs. motif is SC with an Eg of 0.048 eV (as listed in Table 2). The remarkable finding of the present work is that a transition from the SC to HM state occurs in the Ir-dop. structure (see Fig. 4(b)), where a few states cross the Fermi level (EF) in the N, while it contains a definite large Eg of 1.15 eV in the N.


image file: d5ra02453f-f4.tif
Fig. 4 GGA+U computed non-degenerate total density of states (TDOS) in the (a) prs. and (b) Ir-dop. Sr2CaOsO6 structures.

Additionally, we presented the orbital resolved partial density of states (PDOS) on the 5d-states of the Os and Ir/Os ions in the prs. and Ir-dop. system in Fig. 3S of the ESI, to better examine the states close to EF. The Os 5d states are dominant at the valence and the conduction band edges (VBE and CBE) as demonstrated in Fig. 3S(a) of the ESI. As Os is in a +6(5d2) oxidation state, this results in the filling of two t2g states in the N and the eg states are empty. Hence, the filled states are in the VB, while the two empty states shift towards the CB. For the Ir-dop. structure, metallicity in the N mainly arises from the Os-5d states along with substantial contributions from Ir-5d states (see Fig. 3S(b) of the ESI). It is found that the Os-5d states become partially occupied and significantly shift towards lower energies in the CB, which leads the system into the HM phase. This happens because the Ir ion is in a +4(5d5) oxidation state and it contributes 3 additional electrons to the system; one of the additional electrons occupies the t2g and two occupy the t2g state. Hence, they produce a repulsive force in the Os t2g states, causing the Os states to move towards the CB from the VB in the N. Along with this, we plotted the computed non-degenerate band structures for both systems within the GGA+U scheme in Fig. 5 to provide further confirmation of their electronic states. The corresponding MI/HM behavior in the prs./Ir-dop. structure is well illustrated by Fig. 5(a, a′) and (b, b′), which also support the estimated TDOS in Fig. 4(a) and (b). Moreover, we plotted the GGA+U+SOC computed TDOS in Fig. 4S of the ESI for the prs./Ir-dop. SCOO system in the most stable FM/FIM SO. A small Eg of 0.0045 eV exists in the prs. system (see Fig. 4S(a) of the ESI), while the Ir-dop. motif turns metallic (see Fig. 4S(b) of the ESI). It is predicted that the Eg computed within the GGA+U+SOC scheme for the prs. motif decreases (see Fig. 4S(a) of the ESI) as compared to the GGA+U one (see Fig. 4(a)). This is because relativistic effects and SOC interactions often lower the Eg by mixing the orbitals in a way that reduces the energy splitting of the electronic states close to the EF.


image file: d5ra02453f-f5.tif
Fig. 5 GGA+U computed non-degenerate spin-majority (left column)/spin-minority (right column) band structures for the (a and a′) prs. and (b and b′) Ir-dop. Sr2CaOsO6 structures.

Now, the system’s magnetism is described by calculating the total/partial spin magnetic moment (mt/ms) in each case along with the three-dimensional (3D) spin magnetization density isosurfaces. The estimated mt for the prs./Ir-dop. system is 4.0/1.0μB f.u.−1 (see Table 2). The computed ms for the Os1/Os2 ion in the prs. system is 1.19μB, whereas the ms on the Os/Ir is 1.01/−1.39μB for the Ir-dop. structure, as displayed in Table 2. The “−” sign shows that Ir and Os ions’ ms are aligned antiparallel (image file: d5ra02453f-u3.tif) to each other, which indicates that an AFM coupling is dominant, and turns out to be a FIM SO in the Ir-dop. structure. Further, the computed ms on the Os/Ir assures that they are in a +6(t22g↑t02g↓e0g↑e0g↓)/+4(t32g↑t22g↓e0g↑e0g↓) state. Moreover, the computed morb. on the Os is −0.5μB in the prs. and 0.44/0.08μB on the Os/Ir in the Ir-dop. structure. Also, the 3D spin magnetization density isosurfaces, keeping an iso-value of ± 0.05% e Å−1, are plotted in Fig. 6 for the direct observation of ms as well as to further affirm the SO in both structures. An appropriate density arises around the Os ions in the prs. system (see Fig. 6(a)), where the same density colors confirm that their spins’ parallel (↑↑) alignment and t22g orbital characterization is visualized due to its +6 state. Meanwhile in the Ir-dop. motif, the magnitude of density on the Ir ion is greater than that on the Os ion, which further verifies the magnitude of the calculated ms on each ion. Besides this, the color contrast of the densities also assures that the spins of both ions align antiparallel (image file: d5ra02453f-u4.tif) to each other.


image file: d5ra02453f-f6.tif
Fig. 6 Computed spin magnetization density iso-surfaces for the (a) prs. and (b) Ir-dop. Sr2CaOsO6 structures with an iso-value of ±0.05% e Å−1.

Now, the FM SO in the prs. structure is explained by virtual hopping between less-than-half-filled d-orbitals via oxygen (Os+6–O−2–Os+6), as depicted in Fig. 7(a). Since the B-site atom is a non-magnetic cation, the magnetic aspects are determined by exchange coupling within the Os+6 ion, and a FM superexchange interaction occurs at 180°. But in the Ir-dop. system, the electron hopping takes place between the half-filled d-orbitals of Os+6 and half-filled d-orbitals of the Ir+4 ion, which occurs 180° between Os and Ir ions as Os+6–O−2–Ir+4 (see Fig. 7(b)). This supports the strong AFM superexchange mechanism in this structure, resulting in the FIM SO. Now, it is a well-established fact that MAE plays an important part in defining coercivity, which is required for permanent magnets with long-term magnetization.56,57 Therefore, we investigated the MAE and MAE constant image file: d5ra02453f-t7.tif as well as its relation to the structural distortions. To do this, we set the magnetization direction along the [001]/[010]/[100]-axis and compare Et. Moreover, a dense k-mesh of 12 × 12 × 9 is utilized to verify that the Et and MAE values converged correctly. Our results reveal that the prs. system has a MAE/K of 2.95 meV/1.72 × 107 erg per cm3 with an easy axis of [100] and average ∠Os–O–Os of 153.2° In comparison, the Ir-dop. system has a lower MAE/K of 0.79 meV/0.42 × 107 erg per cm−3, as listed in Table 2, with an easy axis of [010] and an average ∠Os–O–Ir of 153.6°. This clearly illustrates that the MAE value is highly dependent on the structural distortions as reported in previous works.5,21


image file: d5ra02453f-f7.tif
Fig. 7 The superexchange interaction process between (a) Os+6 5d and Os+6 5d via oxygen results in a ferromagnetic (FM) coupling in the prs. and that between (b) Ir+4 5d and Os+6 5d via oxygen results in a strong antiferromagnetic (AFM) coupling in the Ir-doped Sr2CaOsO6 structure, which leads to ferromagnetic ordering.

Next, to evaluate the TC of the prs./Ir-dop. SCOO motif, the exchange constants (J) are determined using the Heisenberg model, as follows:58

 
image file: d5ra02453f-t8.tif(3)
where image file: d5ra02453f-t9.tif specifies the spin vectors and can be calculated as S(S + 1)1/2. The S indicates the ms on each magnetic ion, where J respresnts the interactions between inos. The Et of FM, FIM, and AFM SO can be written as:
 
EFMt = E0 + (S1(S1 + 1))1/2(S2(S2 + 1))1/2(2J1 + 4J2) (4)
 
EFIMt = E0 + (S1(S1 + 1))1/2(S2(S2 + 1))1/2(−2J1 + 4J2) (5)
 
EAFMt = E0 + (S1(S1 + 1))1/2(S2(S2 + 1))1/2(−2J1 − 4J2) (6)
E0 refers to the energy in the spin-degenerate system and S1(S2) = 1(1) for Os1+6/Os2+6 in the prs. system and image file: d5ra02453f-t10.tif for Os+6(Ir+4) in the Ir-dop. motif. Hence, TC can be determined as follows:59
 
image file: d5ra02453f-t11.tif(7)
 
image file: d5ra02453f-t12.tif(8)
Our results revealed that the determined TC is 185/171 K for the prs./Ir-dop. motif, as listed in Table 2. As structural distortions are reduced in the Ir-dop. system as compared to the prs. one, this lowers the TC a bit, which is also consistent with the previous reports that TC decreases with a reduction in structural distortion.5

Next, TE parameters are computed to provide insight into these materials for their potential realization in devices that convert heat into electricity. The BoltzTrap code44 is used to calculate the TE factors, such as the electrical conductivity (σ) per relaxation time image file: d5ra02453f-t14.tif, Seebeck coefficient (S), electronic thermal conductivity (κe) per relaxation time image file: d5ra02453f-t15.tif, susceptibility (χ), power factor image file: d5ra02453f-t16.tif, and figure of merit image file: d5ra02453f-t17.tif at temp. ranging from 200 to 600 K, as presented in Fig. 8. As efficient TE materials require a high value of image file: d5ra02453f-t18.tif, the variation in temp.-dependent image file: d5ra02453f-t19.tif is illustrated in Fig. 8(a); it depends on the concentration of the free carriers (electrons or holes) and increases with temp. due to an increase in kinetic energy. The values of image file: d5ra02453f-t20.tif are higher for the Ir-dop. motif, because there is a greater concentration of free charge carriers at EF. Its value is 0.21/0.71 × 1019 Ωm−1 s−1 at 300 K and reaches 0.45/0.9 × 1019 at 600 K for the prs./Ir-dop. structures. The S value, which is another important TE factor for determining the electronic transport properties, measures the TE capacity of a material to generate a potential difference across its edges as a result of a gradient. The temp. difference supports the passage of carriers, creating a gradient that can be determined in μV K−1. Fig. 8(b) demonstrates that S varies significantly between 30 to 200 μV K−1 for the considered temp. range. The computed S values for the prs./Ir-dop. structure are 147/30 μV K−1 and 85/37 μV K−1 at 300 K and 600 K, respectively. Hence, one can conclude from Fig. 8(a) and (b) that the behavior of S against temp. is in contrast to that of image file: d5ra02453f-t21.tif: whereas S decreases with temp., image file: d5ra02453f-t22.tif increases. In addition to image file: d5ra02453f-t23.tif, Fig. 8(c) displays image file: d5ra02453f-t24.tif, which shows an increase with temp., being 0.41/0.84 × 1014 at 300 K and 1.6/2.1 × 1014 W m−1 K−1 s−1 at 600 K for the prs./Ir-dop. system. Next, the term χ describes the response of the TE materials to variations in external conditions such as electric or magnetic fields, which may have an impact on their TE performance. A higher temp. can enhance the TE performance of a material because the thermal energy pushes more charge carriers into the CB, increasing sensitivity to external forces. Hence, Fig. 8(d) exhibits that χ increases with the rise in temp. for the prs./Ir-dop. structure, from 0.9/1.6 × 10−9 m3 mol−1 to 1.5/1.8 × 10−9 m3 mol−1. Similarly, the power factor (PF) is also an important factor to consider when measuring a material's TE efficiency. The trend of the PF is like that of S, which decreases with an increase in temp. The calculated value of PF at 300 K is 0.45/0.064 × 1011 W m−1 K−2 s−1 and at 600 K is 0.32/0.13 W m−1 K−2 s−1 for the prs./Ir-dop. system (see Fig. 8(e)). Likewise, ZT is a critical component in determining TE efficiency for practical applications. The computed value of ZT at 300 K is 0.33/0.02 and at 600 K is 0.13/0.038 for the prs./Ir-dop. system (see Fig. 8(f)).


image file: d5ra02453f-f8.tif
Fig. 8 Computed (a) electrical conductivity per relaxation time image file: d5ra02453f-t5.tif in × 1019 Ωm−1 s−1, (b) Seebeck coefficient (S) in μV K−1, (c) thermal conductivity per relaxation time image file: d5ra02453f-t6.tif in × 1014 W m−1 K−1 s−1, (d) susceptibility (χ) in × 10−9 m3 mol−1, (e) power factor (PF) in × 1011 W m−1 K−2 s−1, and (f) figure of merit (ZT) for the prs./Ir-dop. Sr2CaOsO6 structure.

The net thermal conductivity (κ) is calculated as follows:

 
κ = κe + κl (9)
Slack's equation60 was used to calculate κl as follows:
 
image file: d5ra02453f-t25.tif(10)
Mav, δ, n, T, γ, and θ represent the average atomic mass in the crystal, cubic root of the average atomic volume, total number of atoms in the unit cell, absolute temp., Grüneisen parameter, and Debye temp., respectively. The Grüneisen parameter is determined via Poisson's ratio:
 
image file: d5ra02453f-t26.tif(11)
where A is calculated as follows:61
 
image file: d5ra02453f-t27.tif(12)
while θ can be determined as follows:62
 
image file: d5ra02453f-t28.tif(13)
where h, kB, n, NA, ρ, M, and vm serve as Planck’s constant, the Boltzmann constant, the total number of atoms per unit cell, Avogadro's number, density, molecular weight, and sound velocity, respectively. The average sound velocity (vm) is figured out using
 
image file: d5ra02453f-t29.tif(14)
Here, vl/vt is the longitudinal/transverse velocity, which can be measured as follows:
 
image file: d5ra02453f-t30.tif(15)
 
image file: d5ra02453f-t31.tif(16)

Fig. 9 shows that the estimated image file: d5ra02453f-t32.tif decreases with increasing temp., which is a desired property for TE applications. The decrease in image file: d5ra02453f-t33.tif as temp. rises is due to increased phonon scattering caused by higher lattice vibrations, which decreases the mean free route of phonons. The established values of image file: d5ra02453f-t34.tif for the prs. and Ir-dop. systems are 0.68 and 0.69 × 1014 W m−1 K−1, respectively.


image file: d5ra02453f-f9.tif
Fig. 9 Computed values of lattice thermal conductivity per relaxation time image file: d5ra02453f-t13.tif in × 1014 W m−1 K−1 with temperature, for the prs./Ir-dop. Sr2CaOsO6 structure.

Here, we would like to mention that the Slack method predominantly considers acoustic phonon contributions, disregarding optical phonon modes, and complicated anharmonic interactions, which may be crucial in materials characterized by poor symmetry or complicated bonding environments. Consequently, the model may generate considerable error in the calculation of κl, thereby impacting the precision of resultant TE parameters including ZT. Recent investigations have underscored the ambiguities linked to the Slack model. For example, Rabin et al.63 shown that the Slack formula often overestimates κ in materials characterized by low group velocities and substantial optical phonon contributions, which are particularly influential in materials exhibiting poor symmetry or intricate bonding environments. Consequently, the model may generate significant error in the calculation of κl, possibly compromising the precision of derived thermoelectric parameters, including the ZT. Likewise, J. Carrete et al.64 also noted that the empirical coefficient in the Slack equation is not universal as it changes according to the Grüneisen parameter and the bonding properties of the material. Hence, inaccuracies in κl can influence the assessment of the thermoelectric ZT, either under-representing or overstating a material’s potential, thereby impacting the precision of derived thermoelectric parameters such as the ZT. The Slack model provides a valuable first estimate. However, enhanced accuracy in predictions can be attained by resolving the Boltzmann transport equation for phonons using first-principles interatomic force constants. In future endeavors, we intend to integrate sophisticated methodologies to authenticate and enhance the outcomes derived from the empirical methodology.

Usually, the empirical Slack model offers a useful initial estimate of κ, but it fails to adequately represent phonon transport dynamics in intricate materials. Recent developments in computational materials science have resulted in the emergence of first-principles approaches that offer a more precise and physically substantiated framework for forecasting thermal transport parameters. One technique involves solving the Peierls–Boltzmann transport equation (PBTE) alongside interatomic force constants derived from density functional perturbation theory (DFPT). These approaches consider both acoustic and optical phonons and comprehensively capture phonon–phonon scattering processes, allowing more accurate predictions across a diverse array of materials. Ma et al.65 and Lindsay66 examine the theoretical basis and practical application of these first-principles methodologies. Ma et al. emphasise the accomplishments and persistent difficulties in modeling heat transport using PBTE + DFPT, particularly in low-dimensional and disordered systems. Lindsay offers a comprehensive examination of the concept and illustrates its efficacy in precisely modeling κ across several crystalline materials.

4 Strained systems

Here, we examine the biaxial ([110]) strain consequences for the structural, mechanical, and dynamical stability of the SCOO system by optimizing the two-axis lattice parameters a and b in the range from −5% to +5. First, the thermodynamical stability of the systems is examined by calculating ΔHf under applied strain, and the results are shown in Fig. 5S of the ESI. It is clear that ΔHf is negative within all strain ranges; however, it is more negative under tens. strains, which means that these structures are more stable than the comp. strained ones. Next, we determine the mechanical stability of the structures under strain; 13 independent elastic constants are listed in Table 1, which obey the Born criteria and confirm mechanical stability against strain. Further, Fig. 1S of the ESI displays the B/Y/G values, which show that the prs. system under −5% comp. strain is stiffer and harder than the tens. strain ones. Additionally, we plotted the B/G, Cp, and ν values of both systems under −5% and +5% strain in Fig. 2, which depicted that structures are ductile and show more strength for −5% comp. strain. Moreover, we computed the phonon dispersion curves under two maximum strains of ±5%, as shown in Fig. 6S of the ESI, to ensure the dynamic stability of the motifs. This shows that no negative frequencies exist, indicating that the systems are also dynamically stable for the considered higher strain values. Subsequently, the magnetic ground state stability of the Ir-dop. SCOO system is examined by plotting ΔE = EFIMEFM against strain, as shown in Fig. 7S of the ESI. It is found that ΔE remains negative within all considered strain ranges, which indicates that FIM is more stable than the FM SO in the whole strain range.

To qualitatively describe how the electronic structure of the systems varies under strain, we plotted the computed Eg in the N and N against strain, as shown in Fig. 10. When tens. strain is varied from 0% to +5%, a slight linear increase in Eg from 0.048 to 0.050 eV in the N is observed (see Fig. 10(a)) for the prs. system. Conversely, when comp. strain is varied from 0% to −2%, the computed Eg decreases from 0.048 to 0.027 eV and ultimately becomes zero at a crucial value of −3% in the N (see Fig. 10(a)), while N maintains a definite Eg for the whole strain range. Thus, the system exhibits a SC-to-HM transition at a crucial −3% strain and for higher comp. strains as well. Likewise, when the Ir-dop. system is subjected to strain from 0% to +5%, Eg decreases from 1.15 to 1.12 eV in the N and decreases more significantly under comp. strain from 1.15 eV to 0.75 eV (see Fig. 10(b)). In contrast, N preserves its metallic character for the whole strain range, which ensures the robustness of the HM phase in a doped system. In addition, we plotted the corresponding SP TDOS from −1% to −5% for the prs. system in Fig. 11, which also assures that no states lie at EF and a small/large Eg exists in the N/N for −1 and −2% strains (see Fig. 11(a) and (b)), correspondingly. However, the states overlap at the EF in the N and Eg becomes zero for a crucial −3% strain (see Fig. 11(c)), while a large Eg of 1.86 eV exists in the N. Hence, a SC-to-HM transition occurs at a crucial comp. strain of −3% for the prs. system. Moreover, N becomes more metallic for the −4%/−5% strain (see Fig. 11(d) and (e)) as more electrons reside at EF for these strain levels. Along with this, we plotted the SP TDOS for the tens. strain of +1% to +5% in the prs. system in Fig. 8S of the ESI, which shows that no transition occurs at any strain value and a definite Eg exists in both channels. Hence, the system holds its insulating behavior for all tens. strain levels. Likewise, the calculated TDOS for the Ir-dop. under ±5% biaxial ([110]) strains is plotted in Fig. 9S of the ESI, which is consistent with Fig. 10 and no significant changes occur with strain.


image file: d5ra02453f-f10.tif
Fig. 10 GGA+U calculated energy gap (Eg) against ±5% biaxial ([110]) strain in the spin majority (N)/spin-minority (N) channel of the (a) prs. and (b) Ir-dop. Sr2CaOsO6 structures.

image file: d5ra02453f-f11.tif
Fig. 11 GGA+U computed non-degenerate total density of states (TDOS) of the prs. Sr2CaOsO6 structure for (a) −1%, (b) −2%, (c) −3%, (d) −4%, and (e) −5% biaxial ([110]) compressive strain.

Next, we investigated the strain-induces changes in the systems’ magnetism. To do this, the ms on the Os/Ir ion in the prs. and Ir-dop. SCOO motifs is plotted in Fig. 10S of the ESI. As the strain is varied from −5% to +5%, a slight change in the ms value occurs in both systems. The ms on the Os ion varies from 1.12μB to 1.18μB as strain changes from −5% to +5%. Within a similar strain limit, the ms on Os/Ir changes from 0.9 to 0.95μB/−1.35 to −1.32μB in the Ir-dop. case. Along with this, we computed TC under ±5% biaxial ([110]) strain, as presented in Fig. 12 for both systems. It shows that under −5% strain, prs./Ir-dop. has the highest TC of 245/233 K, while +5% strain gives the lowest TC of 138/129 K. Under comp./tens. strain, structural distortions increase/decrease, which has a positive/negative effect on the TC values, as discussed in previous theoretical works.5,39


image file: d5ra02453f-f12.tif
Fig. 12 Computed Curie temperature (TC) under ±5% biaxial ([110]) strain for the prs. and Ir-dop. Sr2CaOsO6 structures.

Panneerselvam et al.67 found that altering the polar optical phonon scattering mechanism via strain in ScN significantly influences the variance in lattice thermal conductivity while minimally affecting the TE power factor values. Likewise, Zhang et al.68 examined the impact of strain on two-dimensional materials, emphasizing that strain engineering can improve TE capabilities by altering the electronic structure. Additionally, Yu et al.69 showed that strain engineering may markedly influence the κl and heat flow in Bi2Te3 nano-films, therefore highlighting the significance of strain in TE materials. The κl of the Bi2Te3 nano-film may be efficiently adjusted by the application of strain. A tensile strain of 6% can decrease heat conductivity by 50%; however, a compressive strain of 4% can enhance κl by 60%. Hence, we plotted the ZT values of N/N at room temp. (300 K) as a function of ±5% strain in Fig. 13 for both systems. The total ZT is similar in the N↑/N↓ for the prs./Ir-dop. system because it is closely related to the real Eg value, in contrast to Fig. 10 (see Fig. 4). Therefore, we compare the ZT values with the calculated Eg (see Fig. 10) for the strained system, which reveals that ZT is substantially correlated with Eg, where ZT increases as Eg increases and vice versa. The computed ZT under +2/+3% tens. strain is 0.68/0.71 for the prs. and 0.04/0.15 for the Ir-dop. system at 300 K. Therefore, we computed the TE factors of the prs. system under +2% and +3% biaxial tens. strains, as shown in Fig. 11S of the ESI. As Eg increases under +2/+3% tens. strain, the values of image file: d5ra02453f-t35.tif, and χ are less than those of the unstrained system because there is a smaller number of charge carriers across the EF, but the trend remains the same with temp. as compared to the unstrained system. Also, the values of the PF at +2/+3% strain are less than that of the unstrained system, whereas the values of S also increase at +2/+3% tens. strain and it follows the same trend with temp. as followed in the unstrained system. Besides this, we calculated the image file: d5ra02453f-t36.tif of the prs. system under +2/+3% tens. strain, as presented in Fig. 12S of the ESI. This shows that its value decreases with an increase in temperature, which is a required criterion for TE applications.


image file: d5ra02453f-f13.tif
Fig. 13 Computed figure of merit (ZT) in the spin-majority and spin-minority channels for the (a) prs. and (b) Ir-dop. Sr2CaOsO6 structures as a function of ±5% biaxial ([110]) strain.

5 Conclusion

In summary, ab initio calculations were performed to examine the thermodynamic, mechanical, dynamical stability, thermoelectric, electronic structure, and magnetic properties of the unstrained and strained systems of the pristine (prs.)/Ir-doped (dop.) Sr2CaOsO6. It is predicted that both systems are thermodynamically and mechanically stable as they have negative formation energies and follow the Born criteria. Due to non-negative phonon frequency curves, they are dynamically stable as well. The unstrained prs./Ir-dop. system has a ferromagnetic (FM)/ferrimagnetic (FIM) ground state due to FM/antiferromagnetic (AFM) interactions between Os and Os/Os and Ir ions. The prs. motif is a semiconductor, while Ir-doping results in a half-metallic state (HM). The additional electrons produced by the dopant (Ir) cause a repulsion in the half-filled Os t22g in the spin-minority channel, which moves the Os bands towards the Fermi level, resulting in conductivity. Also, a large energy-gap of 1.15 eV in the non-metallic channel ensures the stability of the HM state of the system. Moreover, the computed ZT is 0.33/0.02 at 300 K for the prs./Ir-dop. motif. Further, our results indicate that the prs. system becomes half-metallic at a crucial value of −3% compressive strain. In contrast, the Ir-dop. motif retains its half-metallic nature against ±5% biaxial ([110]) strain. The prs./Ir-dop. structure has a high magneto-crystalline anisotropy energy (MAE) and MAE constants (K) of 2.95/0.79 meV and 1.72/0.46 × 107 erg per cm3 at the Curie temperature of 185/171 K, respectively. Interestingly, ZT values increase in the prs. system under +2/+3% tens. strain because the energy gap increases against these strain values. Hence, near-unity ZT values suggest that these materials are very suitable for magnetic memory devices and thermoelectric applications.

Data availability

The datasets used and/or analyzed during the current study are available from the corresponding author on reasonable request.

Author contributions

Samia Shahzadi: writing – original draft, investigations, formal analysis, data curation. Ihab Mohamed Moussa: validation, visualization, resources, formal analysis, funding. Sohail Mumtaz: visualization, validation, formal analysis. S. Nazir: writing – review and editing, validation, supervision, project administration, conceptualization.

Conflicts of interest

The authors declare no competing interests.

Acknowledgements

The authors would like to thank the Researchers Supporting Project number (RSPD2025R741), King Saud University, Riyadh, Saudi Arabia. The computational work was supported by the University of Sargodha, Sargodha, Pakistan.

Notes and references

  1. J. Philipp, P. Majewski, L. Alff, A. Erb, R. Gross, T. Graf, M. Brandt, J. Simon, T. Walther and W. Mader, et al., Phys. Rev. B:Condens. Matter Mater. Phys., 2003, 68, 144431 CrossRef.
  2. L. Ding, D. D. Khalyavin, P. Manuel, J. Blake, F. Orlandi, W. Yi and A. A. Belik, Acta Mater., 2019, 173, 20–26 CrossRef CAS.
  3. M. T. Anderson, K. B. Greenwood, G. A. Taylor and K. R. Poeppelmeier, Prog. Solid State Chem., 1993, 22, 197–233 CrossRef CAS.
  4. M. P. Ghimire and X. Hu, Mater. Res. Express, 2016, 3, 106107 CrossRef.
  5. S. Nazir, Phys. Chem. Chem. Phys., 2024, 26, 14384–14392 RSC.
  6. A. Aligia, P. Petrone, J. Sofo and B. Alascio, Phys. Rev. B:Condens. Matter Mater. Phys., 2001, 64, 092414 CrossRef.
  7. Z. Yang, C. Ko and S. Ramanathan, Annu. Rev. Mater. Res., 2011, 41, 337–367 CrossRef CAS.
  8. N. F. Mott, Proc. Phys. Soc., 1949, 62, 416 CrossRef.
  9. H. Xiang and M.-H. Whangbo, Phys. Rev. B:Condens. Matter Mater. Phys., 2007, 75, 052407 CrossRef.
  10. C. Tian, A. C. Wibowo, H.-C. zur Loye and M.-H. Whangbo, Inorg. Chem., 2011, 50, 4142–4148 CrossRef CAS PubMed.
  11. W. Song, J. Wang and Z. Wu, Chem. Phys. Lett., 2009, 482, 246–248 CrossRef CAS.
  12. H. L. Feng, M. Arai, Y. Matsushita, Y. Tsujimoto, Y. Guo, C. I. Sathish, X. Wang, Y.-H. Yuan, M. Tanaka and K. Yamaura, J. Am. Chem. Soc., 2014, 136, 3326–3329 CrossRef CAS PubMed.
  13. K.-W. Lee and W. Pickett, Phys. Rev. B:Condens. Matter Mater. Phys., 2008, 77, 115101 CrossRef.
  14. Y. Krockenberger, K. Mogare, M. Reehuis, M. Tovar, M. Jansen, G. Vaitheeswaran, V. Kanchana, F. Bultmark, A. Delin and F. Wilhelm, et al., Phys. Rev. B:Condens. Matter Mater. Phys., 2007, 75, 020404 CrossRef.
  15. M. Azuma, K. Takata, T. Saito, S. Ishiwata, Y. Shimakawa and M. Takano, J. Am. Chem. Soc., 2005, 127, 8889–8892 CrossRef CAS PubMed.
  16. Z.-C. Wang, L. Chen, S.-S. Li, J.-S. Ying, F. Tang, G.-Y. Gao, Y. Fang, W. Zhao, D. Cortie and X. Wang, et al., npj Quantum Mater., 2021, 6, 53 CrossRef CAS.
  17. A. Maignan, B. Raveau, C. Martin and M. Hervieu, J. Solid State Chem., 1999, 144, 224–227 CrossRef CAS.
  18. G. A. Prinz, Science, 1998, 282, 1660–1663 CrossRef CAS PubMed.
  19. S.-W. Cheong and M. Mostovoy, Nat. Mater., 2007, 6, 13–20 CrossRef CAS PubMed.
  20. R. Ramesh and N. A. Spaldin, Nat. Mater., 2007, 6, 21–29 CrossRef CAS PubMed.
  21. P. C. Rout and U. Schwingenschlogl, Nano Lett., 2021, 21, 6807–6812 CrossRef CAS PubMed.
  22. J. Navarro, L. Balcells, F. Sandiumenge, M. Bibes, A. Roig, B. Martinez and J. Fontcuberta, J. Condens. Matter Phys., 2001, 13, 8481 CrossRef CAS.
  23. M. García-Hernández, J. Martínez, M. Martínez-Lope, M. Casais and J. Alonso, Phys. Rev. Lett., 2001, 86, 2443 CrossRef PubMed.
  24. P. Sanyal, A. Halder, L. Si, M. Wallerberger, K. Held and T. Saha-Dasgupta, Phys. Rev. B, 2016, 94, 035132 CrossRef.
  25. M. P. Ghimire, L.-H. Wu and X. Hu, Phys. Rev. B, 2016, 93, 134421 CrossRef.
  26. R. Morrow, J. R. Soliz, A. J. Hauser, J. C. Gallagher, M. A. Susner, M. D. Sumption, A. A. Aczel, J. Yan, F. Yang and P. M. Woodward, J. Solid State Chem., 2016, 238, 46–52 CrossRef CAS.
  27. K. Samanta, P. Sanyal and T. Saha-Dasgupta, Sci. Rep., 2015, 5, 15010 CrossRef CAS PubMed.
  28. M. M. H.-E. Saad, J. Phys. Chem. Solids, 2016, 94, 1–11 Search PubMed.
  29. G. Jackeli and G. Khaliullin, Phys. Rev. Lett., 2009, 102, 017205 CrossRef CAS PubMed.
  30. H. L. Feng, W. Schnelle, L. H. Tjeng and M. Jansen, Solid State Commun., 2016, 243, 49–54 CrossRef CAS.
  31. N. Dai, J. Feng, Z. Wang, T. Jiang, W. Sun, J. Qiao and K. Sun, J. Mater. Chem. A, 2013, 1, 14147–14153 RSC.
  32. R. S. Bhandari, K. D. Yadav, P. B. Belbase, M. Zeeshan, B. Sadhukhan, P. D. Rai, K. R. Thapa, C. G. Kaphle and P. M. Ghimire, RSC Adv., 2020, 10, 16179 RSC.
  33. O. El Rhazouani, Z. Zarhri, A. Benyoussef and A. El Kenz, Comput. Condens. Matter., 2016, 8, 1–6 CrossRef.
  34. O. Rhazouani and A. Benyoussef, J. Magn. Magn. Mater., 2018, 446, 166–169 CrossRef.
  35. H. Jalili, N. Heinig and K. Leung, Phys. Rev. B:Condens. Matter Mater. Phys., 2009, 79, 174427 CrossRef.
  36. Q. Gan, R. Rao, C. Eom, J. Garrett and M. Lee, Appl. Phys. Lett., 1998, 72, 978–980 CrossRef CAS.
  37. C. Jung, H. Yamada, M. Kawasaki and Y. Tokura, Appl. Phys. Lett., 2004, 84, 2590–2592 CrossRef CAS.
  38. Y. Qian, H. Wu, E. Kan, J. Lu, R. Lu, Y. Liu, W. Tan, C. Xiao and K. Deng, J. Appl. Phys., 2013, 114, 063713 CrossRef.
  39. S. Nazir, Phys. Chem. Chem. Phys., 2020, 22, 17969–17977 RSC.
  40. C. Xie, L. Shi, S. Zhou, J. Zhao, H. Liu, Y. Li and D. Yao, Surf. Coat. Technol., 2015, 277, 222–226 CrossRef CAS.
  41. N. Zu, R. Li and R. Ai, J. Magn. Magn. Mater., 2018, 467, 145–149 CrossRef CAS.
  42. P. Blaha, K. Schwarz, P. Sorantin and S. Trickey, Comput. Phys. Commun., 1990, 59, 399–415 CrossRef CAS.
  43. C. E. Calderon, J. J. Plata, C. Toher, C. Oses, O. Levy, M. Fornari, A. Natan, M. J. Mehl, G. Hart and M. B. Nardelli, et al., Comput. Mater. Sci., 2015, 108, 233–238 CrossRef CAS.
  44. G. K. H. Madsen and D. J. Singh, Comput. Phys. Commun., 2006, 175, 67–71 CrossRef CAS.
  45. G. S. Thakur, W. Schnelle and M. Jansen, Z. Anorg. Allg. Chem., 2023, 649, e202300061 CrossRef CAS.
  46. R. P. Madhogaria, N. S. Bingham, R. Das, M.-H. Phan and H. Srikanth, J. Alloys Compd., 2021, 888, 161624 CrossRef CAS.
  47. K. Naveen, T. Rom, S. S. Islam, M. Reehuis, P. Adler, C. Felser, A. Hoser, R. C. Nath, A. K. Yadav and S. N. Jha, et al., Phys. Chem. Chem. Phys., 2021, 23, 21769–21783 RSC.
  48. S. Nazir, Sci. Rep., 2021, 11, 1240 CrossRef CAS PubMed.
  49. L. Wang, T. Maxisch and G. Ceder, Phys. Rev. B:Condens. Matter Mater. Phys., 2006, 73, 195107 CrossRef.
  50. X. Wu, D. Vanderbilt and D. Hamann, Phys. Rev. B:Condens. Matter Mater. Phys., 2005, 72, 035105 CrossRef.
  51. S. Shang, Y. Wang and Z.-K. Liu, Appl. Phys. Lett., 2007, 90, 101909 CrossRef.
  52. F. Mouhat and F.-X. Coudert, Phys. Rev. B:Condens. Matter Mater. Phys., 2014, 90, 224104 CrossRef.
  53. S. Pugh, Lond. Edinb. Dubl. Phil. Mag., 1954, 45, 823–843 CrossRef CAS.
  54. A. Bakar, A. O. Alrashdi, M. M. Fadhali, A. Afaq, H. Yakout and M. Asif, J. Mater. Res. Technol., 2022, 19, 4233–4241 CrossRef CAS.
  55. Y. Lv, X. Zhang and W. Jiang, Ceram. Int., 2018, 44, 128–135 CrossRef CAS.
  56. J. Coey, IEEE Trans. Magn., 2011, 47, 4671–4681 CAS.
  57. J. Coey, Scr. Mater., 2012, 67, 524–529 CrossRef CAS.
  58. M. Pajda, J. Kudrnovsky, I. Turek, V. Drchal and P. Bruno, Phys. Rev. B:Condens. Matter Mater. Phys., 2001, 64, 174402 CrossRef.
  59. W. Chen, J. George, J. B. Varley, G.-M. Rignanese and G. Hautier, Npj Comput. Mater., 2019, 5, 72 CrossRef.
  60. J. Wei, Y. Guo and G. Wang, RSC Adv., 2023, 13, 11513–11524 RSC.
  61. S. A. Sofi and D. C. Gupta, Int. J. Energy Res., 2021, 45, 4652–4668 CrossRef CAS.
  62. O. L. Anderson, J. Phys. Chem. Solids, 1963, 24, 909–917 CrossRef CAS.
  63. D. Rabin, D. Fuks and Y. Gelbstein, Phys. Chem. Chem. Phys., 2023, 25, 520–528 RSC.
  64. J. Carrete, N. Mingo and S. Curtarolo, Appl. Phys. Lett., 2014, 105, 101907 CrossRef.
  65. T. Ma, P. Chakraborty, X. Guo, L. Cao and Y. Wang, Int. J. Thermophys., 2020, 41, 9 CrossRef CAS.
  66. L. Lindsay, Nanoscale Microscale Thermophys. Eng., 2016, 20, 67–84 CrossRef CAS.
  67. I. R. Panneerselvam, M. H. Kim, C. Baldo, Y. Wang and M. Sahasranaman, Phys. Chem. Chem. Phys., 2021, 23, 23288–23302 RSC.
  68. G. Zhang and Y.-W. Zhang, Mech. Mater., 2015, 91, 382–398 CrossRef.
  69. C. Yu, G. Zhang, Y.-W. Zhang and L.-M. Peng, Nano Energy, 2015, 17, 104–110 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ra02453f

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