Strong quartic anharmonicity and thermoelectric performance in antiperovskite Sr3XN (X = P, As, Sb, and Bi)

Wenling Ren , Jincheng Yue , Shuyao Lin , Chen Shen , Yanhui Liu , Tian Cui , Hongbin Zhang

Front. Phys. ›› 2026, Vol. 21 ›› Issue (6) : 064203

PDF (6202KB)
Front. Phys. ›› 2026, Vol. 21 ›› Issue (6) : 064203 DOI: 10.15302/frontphys.2026.064203
RESEARCH ARTICLE

Strong quartic anharmonicity and thermoelectric performance in antiperovskite Sr3XN (X = P, As, Sb, and Bi)

Author information +
History +
PDF (6202KB)

Abstract

Antiperovskite (APV) materials have garnered significant attention due to their diverse physical properties. The self-consistent phonon (SCP) theory and the Boltzmann transport equation (BTE) were combined to calculate thermal and electronic transport properties of APV Sr3XN (X = P, As, Sb, Bi) based on density functional theory (DFT). Pronounced quartic anharmonicity was identified in Sr3PN and Sr3AsN, attributed to the vibrations of Sr atoms. This distinctive characteristic induces four-phonon (4ph) scattering, significantly reducing the lattice thermal conductivity (κL), particularly at elevated temperatures such as 900 K, with values as low as 0.58 W·m−1·K−1 for Sr3PN. Additionally, Sr3BiN exhibits remarkable electrical conductivity and a high Seebeck coefficient, leading to a promising figure of merit (ZT) of up to 1.03, underscoring its potential as a thermoelectric material. These findings demonstrate that the ultra-low lattice thermal conductivity driven by 4ph scattering, coupled with excellent thermoelectric properties, makes these materials promising candidates for high-efficiency thermoelectric applications.

Graphical abstract

Keywords

antiperovskite / four-phonon scattering / self-consistent phonon / thermoelectric

Cite this article

Download citation ▾
Wenling Ren, Jincheng Yue, Shuyao Lin, Chen Shen, Yanhui Liu, Tian Cui, Hongbin Zhang. Strong quartic anharmonicity and thermoelectric performance in antiperovskite Sr3XN (X = P, As, Sb, and Bi). Front. Phys., 2026, 21(6): 064203 DOI:10.15302/frontphys.2026.064203

登录浏览全文

4963

注册一个新账户 忘记密码

1 Introduction

Perovskite (PV) materials initially attracted widespread interest due to their excellent physical properties [14]. Classical inorganic PVs crystallise in a cubic ABX3 lattice, where A and B are cations, and X is a highly electronegativity anion (e.g., O, N, or S) [5]. Interest has since broadened to hybrid organic–inorganic perovskites — exemplified by CH3NH3PbBr3 (MAPbBr3) and MAPbI3, which offer exceptional compositional tunability and functional versatility [6, 7]. More recently, antiperovskites (APVs) have emerged as an inverse-structured analogue with the general formula X3AB, in which X is usually a transition-, alkali- or alkaline-earth-metal cation [8, 9]. APVs have garnered considerable attention because of their unique properties [8, 10], including their magnetic properties [11, 12], negative thermal expansion [1315], ionic conductivity [16, 17], superconductivity [3, 18], optoelectronic, phase stability [9], and thermoelectric properties [1923].

Interestingly, Sr-based APVs exhibit intriguing physical properties. For instance, Sr3PN has been reported to exhibit remarkable optical properties, including a high light absorption coefficient, low reflectivity, and exceptional photoconductivity, with a notable extinction coefficient [24]. Additionally, Sr3AsN is predicted to exhibit substantial mechanical robustness, energetic stability, and dynamic equilibrium within the visible spectrum [25]. Sr3MN (M = Sb or Bi) has also attracted attention as a potential Pb-free material for thin-film solar cells [2], while Sr3BiN exhibits a strong optical response in the ultraviolet regime [26]. Despite these promising attributes, there has been limited investigation of the thermal and electronic transport properties of these materials.

Anharmonicity describes deviations of a potential energy surface (PES) from an ideal harmonic shape near a system’s equilibrium configuration [27]. While the harmonic approximation considers only second-order terms in the PES expansion, anharmonicity arises from higher-order contributions, enabling phonon–phonon interactions [28]. In materials science, these anharmonic interactions significantly influence both thermal and electronic transport properties [2934]. For example, enhanced phonon–phonon scattering shortens phonon lifetimes, thereby lowering lattice thermal conductivity and amplifying its sensitivity to temperature changes [21, 35, 36]. On the other hand, anharmonicity-induced phonon hardening can increase phonon mobility, thus improving heat transport characteristics [37, 38]. Furthermore, strong anharmonicity significantly modulates electron−phonon interactions, affecting phonon scattering processes [39, 40] and enhancing electron−phonon coupling — a key mechanism underlying superconductivity [41].

A comprehensive treatment of anharmonicity encompasses several methodological tiers. Initially, quasi-harmonic and free-energy-perturbation approaches incorporate thermal expansion effects while sampling higher-order terms through stochastic integration [42]. Perturbative methods, such as linearised Boltzmann-transport solvers exemplified by Phono3py [43] and ShengBTE [22, 4448], evaluate three- and four-phonon collision integrals. At higher complexity, renormalised-phonon frameworks — including self-consistent phonon (SCPH) [49], the stochastic self-consistent harmonic approximation (SSCHA) [50] and the temperature-dependent effective-potential (TDEP) [51] — iteratively update the dynamical matrix using fnite-temperature force constants. Furthermore, approaches based on Green–Kubo and direct non-equilibrium molecular dynamics methodologies capture anharmonicity by computing thermal conductivity directly from heat-flux correlations obtained from classical or ab initio trajectories [52].

Four-phonon scattering processes, despite their computational complexity, have recently become essential for accurately describing thermal transport in strongly anharmonic materials or at elevated temperatures [53]. This additional scattering mechanism significantly reduces thermal conductivity relative to predictions based solely on three-phonon interactions. Studies on materials such as silicon [54], diamond [54], and single-layer graphene [55] have explicitly demonstrated the substantial reduction in anomalously high thermal conductivity due to four-phonon scattering. This additional scattering mechanism significantly reduces thermal conductivity beyond predictions based solely on three-phonon interactions. Studies on materials such as silicon [54], diamond [54], and single-layer graphene [55] have explicitly demonstrated the substantial reduction in anomalously high thermal conductivity due to four-phonon scattering. The pronounced impact of quartic interactions underscores the crucial relationship between lattice anharmonicity and phonon scattering processes, highlighting the necessity for continued exploration of quartic anharmonic effects to achieve more accurate thermal transport predictions.

This work provides a comprehensive investigation of four APV materials, previously predicted by high-throughput calculations [56] to exhibit promising thermoelectric performance, with a focus on their phase stability, lattice dynamics, and thermal and electronic transport properties. It emphasizes the significant role of quartic anharmonicity effects in achieving low lattice thermal conductivity (κL) and high figure-of-merit (ZT) values. A combined approach employing self-consistent phonon (SCPH) theory [49, 57], the Boltzmann transport equation (BTE), and four-phonon scattering mechanisms was used. Notably, Sr3PN and Sr3AsN exhibit pronounced anharmonicity, and the inclusion of four-phonon scattering substantially reduces their κL. Our analysis, supported by phonon evaluations at the theoretical model level, underscores the pivotal role of four-phonon processes in accurately assessing anharmonic effects. These insights have contributed significantly to achieving high ZT values, reinforcing the importance of quartic anharmonic interactions in designing of thermoelectric materials. Thus, our findings highlight the essential role of four-phonon processes in developing thermoelectric materials with exceptionally low lattice thermal conductivity, critical for optimizing material performance.

2 Methodology

Density functional theory (DFT) calculations [58] were conducted using the Vienna Ab initio Simulation Package (VASP) [59], utilizing a plane-wave basis set and the Perdew−Burke−Ernzerhof revised for solids (PBEsol) exchange-correlation functional [60]. The Sr (4s4p5s), X (s2p3), and N (s2p3) shells were treated as valence states. The PBEsol functional was chosen because it modifies the gradient expansion of PBE to better reproduce the exchange-correlation energy density in dense electron systems, providing more accurate predictions of lattice constants.

The electronic transport properties were calculated utilizing the AMSET code [61], with a band gap correction applied via the Heyd−Scuseria−Ernzerhof (HSE06) hybrid functional. HSE06 combines Hartree−Fock (HF) exact exchange within the PBE generalized gradient approximation, providing a more accurate estimation of band gaps that are typically underestimated by PBE. The electrical conductivity (σ), Seebeck coefficient (S), and charge carrier contribution to the thermal conductivity (κe) were obtained using the following equations [61]:

σαβ=e2αβ(ε)(f0ε)dε,

Sαβ=1eTαβ(ε)(εεF)(f0ε)dεαβ(ε)(f0ε)dε,

κeαβ=1T{20[αβ(ε)(εεF)(f0ε)]2dεαβ(ε)(f0ε)dεαβ(ε)(εεF)(f0ε)dε},

where α and β denote Cartesian coordinates. e, ε, and f0 represent the electron charge, energy, and Fermi–Dirac distribution, respectively. Σαβ(ε) is the spectral conductivity. T is temperature, and εF is the Fermi level for a given doping concentration.

High-standing convergence criteria were applied for structural optimization, as phonon properties are highly sensitive to the precision of the underlying electronic structure and forces between atoms. A kinetic energy cutoff of 600 eV and a Monkhorst−Pack k-mesh of 10 × 10 × 10 for Brillouin zone integration were used in a system containing five atoms. The thresholds were set to 106eV/Å for the Hellmann−Feynman forces acting on each atom and 108eV for the total energy. For all single-point self-consistent calculations, a kinetic energy cutoff of 600 eV was used, along with a k-mesh of 5 × 5 × 5 and a 2 × 2 × 2 supercell. A convergence criterion of 108eV for total energy was applied.

The frozen phonon approach [62] was employed to evaluate the harmonic interatomic force constants (IFCs). To obtain higher-order IFCs, a compressive-sensing method [6365] was use to get higher order IFCs so ab initio molecular dynamics (AIMD) simulations were performed, using 2 × 2 × 2 supercells with a 2 × 2 × 2 Monkhorst−Pack k-point grid. Subsequently, 100 atomic structures from the trajectories were selected, and random displacements of 0.1 Å were introduced to reduce cross-correlations among sampled configurations. The resulting displacement-force datasets were used to determine the third order and fourth order. Cutoff radii of 5 Å and 4 Å were applied for the 3rd and 4th IFCs, respectively, to ensure convergence while maintaining computational efficiency. Combined this with the self-consistent phonon (SCPH) theory [66], the temperature-dependent phonon bands were obtained. Calculation details can be found in the supplementary.

The FourPhonon [67, 68] package was employed to iteratively solve the linearized phonon BTE, enabling the calculation of the κL. The lattice temperature-dependent κL is obtained as [44]

κlαβ=1κBT2ΩNλf0(f0+1)(ωλ)2υλαFλβ,

where λ comprises both a phonon branch index and a wave vector, α and β represent Cartesian coordinates, is the reduced Planck constant, κB represents the Boltzmann constant, T is the temperature, Ω is the unit cell volume, and N is the number of atoms. ωλ and υλ are the angular frequency and the group velocity of phonon mode λ, respectively. At thermal equilibrium, in the absence of a temperature gradient or other thermodynamical forces, phonons are distributed according to Bose−Einstein statistics, f0. For a certain mode λ, the linearized BTE is expressed as [69]

Fλ=τλ0(υλ+Δλ),

where τλ0 denotes the relaxation time of mode λ, and Δλ is the deviation of the distribution function, characterizing the phonon population. The deviation of the distribution function arises due to the competition between the driving force (temperature gradient) and scattering mechanisms. The iterative process begins with an initial guess for the Δλ, which is subsequently used to update the equations by refining Fλ at each step until the conductivity tensor converges. For the BTE calculations, the number of grid planes in the inner q space was tested, as shown in Fig. S7. The blue line represents the results for 3ph scattering, which clearly converge when the q-mesh is larger than 7. The red line corresponds to the 4ph scattering, which shows convergence when the q-mesh exceeds 6. Therefore, an 8×8×8 q-mesh was adopted for all calculations to ensure accuracy while maintaining reasonable computational cost.

For materials with strong anharmonicity, the quasiparticle (phonon‐gas) picture becomes incomplete: incoherent vibrational transport can make a sizable contribution to the lattice thermal conductivity and often raises κL above the value predicted by purely coherent (propagating) phonons. Within the unified framework, these incoherent contributions are captured by the off-diagonal terms of the heat-flux operator and lead to the following expression for κL [7072]:

κL=2kBT2VNqqs,sΩqs+Ωqs2vqs,svqs,sΩqsnqs(nqs+1)+Ωqsnqs(nqs+1)4(ΩqsΩqs)2+(Γqs+Γqs)2(Γqs+Γqs).

Here, , kB, T, V, and Nq denote the reduced Planck constant, the Boltzmann constant, the absolute temperature, the unit-cell volume, and the number of sampled phonon wave vectors, respectively. Ωqs is the (possibly renormalized) phonon frequency of branch s at wave vector q, nqs is the Bose–Einstein occupation, Γqs is the scattering rate (linewidth, i.e., the inverse lifetime), and vqs,s is the generalized group-velocity tensor. The dyadic and the centered dot indicate a tensor product and a double contraction. The Lorentzian denominator encodes spectral broadening from finite lifetimes, while the off-diagonal (ss) terms account for incoherent (diffuson-like) energy transfer that becomes prominent in strongly anharmonic or disordered crystals.

The thermoelectric efficiency depends on the thermoelectric ZT, which is defined as [73, 74]

ZT=σS2Tκe+κl.

3 Results and discussion

3.1 Phase stability and lattice dynamics

The APV crystallizes in a face-centered cubic structure with space group Pm3¯m (No. 221), as illustrated in the inset of Fig. 1(e). As shown in Table 1, the optimized lattice constants for four compounds, namely Sr3PN, Sr3AsN, Sr3SbN, and Sr3BiN are 5.02, 5.05, 5.14, and 5.18 Å, respectively. These results are consistent with previous studies [2, 7679], showing differences of less than 0.08 Å. Apparently, the lattice parameters of these compounds increase as the atomic radius of the X atom grows (P: 107 pm, As: 119 pm, Sb: 139 pm, Bi: 148 pm) and its electronegativity decreases (P: 2.19, As: 2.18, Sb: 2.05, Bi: 2.02). Table 1 also shows the computed dielectric tensors and Born effective charges, which are used to correct the dynamical matrices via the non-analytical term correction (NAC). For example, the dielectric tensor and Born effective charges for Sr3PN are: ϵ=9.05, Z(Sr)||=2.60, Z(Sr)=2.71, Z(X)=3.45, and Z(N)=4.46.

The thermodynamic stability determines whether a compound can be synthesized. Therefore, an accurate evaluation of the stability of the APVs is essential. The thermodynamic stability can be assessed using the convex hull constructed from the formation energies of all relevant competing phases within a given chemical system. In this work, the formation energies and convex hull data of the Sr–P–N, Sr–As–N, Sr–Sb–N, and Sr–Bi–N ternary systems were obtained from the OQMD database [80]. The formation energy of a compound with composition x is defined as Efphase(x)=Etotphase(x)iνiμielement, where Etotphase(x) is the total energy of the compound, νi is the stoichiometric coefficient of element i, and μielement is the total energy per atom of the elemental reference. The negative formation energies shown in Table 2 indicate that the compound can, in principle, form exothermically from its elements. The stability of each compound is further quantified by its energy above the convex hull, defined as Ehull=Efphase(x)Efhull(x), where Efhull(x) represents the lowest formation energy achievable at the same composition by any combination of stable competing phases [81, 82]. As shown in Fig. S1, Sr3SbN and Sr3BiN lie on the convex hull (ΔEH=0), indicating thermodynamic stability at 0 K. Sr3PN and Sr3AsN are 26 and 82 meV/atom above the hull, respectively, placing them within the commonly accepted metastability range (100meV/atom) [56]. Compounds within this range are often experimentally accessible due to entropic stabilization or kinetic trapping. For Sr3AsN, an orthorhombic Pnma polymorph is predicted, consistent with other antiperovskites [83]. This Pnma phase is 29 meV/atom lower in energy than the cubic Pm3¯m structure. However, it exhibits a relatively low ZT value [4] and is therefore not considered in subsequent analysis. The cubic Pm3¯m phase remains metastable with respect to the convex hull but may be stabilized at finite temperatures. To further assess their resistance to decomposition, the decomposition enthalpies were evaluated, defined as ΔHd=EcompiνiEi, where Ecomp and Ei denote the total energies of the compound and the lowest-enthalpy mixture of competing phases at the same composition, respectively [84]. Under this convention, ΔHd<0 indicates that a compound is stable relative to its decomposition products. For Sr3PN, Sr3AsN, Sr3SbN, and Sr3BiN, the calculated ΔHd values are 0.123, −0.251, −0.680, and −0.731 eV/f.u., respectively. The slightly positive value for Sr3PN suggests that it is metastable at 0 K but may be kinetically or entropically stabilized at finite temperatures, consistent with its small energy above the convex hull (26 meV/atom). In contrast, the negative ΔHd values for Sr3AsN, Sr3SbN, and Sr3BiN confirm their thermodynamic stability. Taken together, these results indicate that Sr3PN and Sr3AsN are near the stability boundary and represent plausible synthesis targets, whereas Sr3SbN and Sr3BiN are robust thermodynamic ground states [78].

To further assess the thermal stability of these APV materials, AIMD simulations were conducted to track changes in free energy over time. As shown in Fig. 1(c), the free energy stabilizes over time, and variations in free energy across different temperatures are minimal, confirming the thermal stability of the four APV materials under varying conditions.

Dynamic stability refers to a material’s ability to resist distortions and maintain its structure under dynamic perturbations. To evaluate this, lattice dynamics were analyzed using harmonic phonon dispersion relations, as represented by the yellow lines in Figs. 1(a)−(d). For Sr3PN and Sr3AsN, imaginary frequencies appear at the M(0.5,0.5,0) and R(0.5,0.5,0.5) points, indicating dynamic instability (ωq2<0). These unstable modes correspond to the irreducible representations M2+ at M-pont and R5 at R-point, respectively. In contrast, the harmonic phonon dispersion of Sr3SbN and Sr3BiN reveals no imaginary frequencies, confirming their dynamic stability. Furthermore, the imaginary frequencies observed in Sr3PN and Sr3AsN are primarily induced by the vibrations of Sr atoms, highlighting the role of atomic contributions to dynamic instability.

To further explore the phonon modes, frozen phonon PES were calculated for the M2+ and R5 modes at the M and R points, as shown in Figs. 1(e) and (f). Both Sr3PN and Sr3AsN exhibit soft modes at these points, characterized by double-well PES, confirming their dynamic instability. In this scenario, the minimum energy points are not located at the zero-tilt amplitude (Q1=Q2=0) but occur at non-zero coordinates, indicating structural distortions [85]. The mode diagram further shows that the vibrations at the M and R points are primarily driven by Sr atoms, consistent with the earlier findings that phonon soft modes are dominated by Sr contributions. To address these instabilities, SCP theory incorporates anharmonic effects into the lattice dynamics. This iterative approach refines the effective potential by accounting for higher-order interactions, renormalizing the phonon frequencies, and removing imaginary components, thereby dynamically stabilizing the structure. Additionally, SCP theory captures temperature-dependent stabilization, where anharmonic contributions at finite temperatures smooth the double-well PES, enabling accurate predictions of thermodynamic and vibrational properties. As a result, the imaginary frequencies in Sr3PN and Sr3AsN can be resolved using SCP theory, making it feasible to study these materials despite their thermodynamic metastability.

In addition, the stability and symmetry of APV structures can be predicted using the tolerance factor (t) [8689]. The tolerance factor is given by t=(rX+rSr)/[2(rN+rSr)] [90], where rSr, rN, and rX represent the ionic radii of the Sr cation, N anion, and X(P, As, Sb, Bi) anion, respectively. For Sr3PN and Sr3AsN, t values of 0.825 and 0.846 suggest significant lattice distortion, aligning with their observed dynamic instability [9]. In contrast, Sr3SbN and Sr3BiN exhibit t values of 0.904 and 0.928, respectively, which stabilize the lattice in a cubic structure due to the larger ionic radii of the Sb and Bi anions.

The inclusion of the quartic anharmonic term in SCP theory plays a crucial role in solving the self-consistent equation, providing the phonon frequencies and polarization vectors at different temperatures. As shown by the red lines in Figs. 1(a)−(d), the soft modes of Sr3PN and Sr3AsN shift, primarily due to the dominant contribution of the diagonal terms in the quartic term coefficient matrix. Off-diagonal terms can redistribute anharmonic effects among different modes. Although they may indirectly influence the soft modes, their impact is generally weaker compared to the direct contributions of the diagonal terms. This shift indicates that the anharmonic contributions have a substantial impact on phonon behavior. It is also observed that the low-lying modes of Sr3PN and Sr3AsN harden with increasing temperature. Moreover, with increasing temperature, the phonon frequencies over 250 (cm1) of Sr3SbN and Sr3BiN shift upward, while phonon self-energy corrections significantly affect medium and high-frequency optical phonons. The disappearance of imaginary frequencies dominated by Sr atoms further suggests notable anharmonicity in these materials. At 600 K, the harmonic and anharmonic phonon density of states reveal that collective motions involving Sr and X atoms primarily govern the acoustic and low-frequency optical modes, while N atoms control the behavior of the two high-frequency phonon branches. The inclusion of Born effective charges in self-consistent phonon (SCP) calculations leads to frequency splitting, particularly at the Γ point, highlighting the importance of accurately modeling anharmonic interactions in lattice dynamics. This frequency splitting, known as longitudinal optical–transverse optical (LO-TO) splitting, arises from long-range Coulomb interactions associated with macroscopic electric fields generated by LO phonons. To accurately capture this effect, NAC must be incorporated when calculating phonon frequencies. NAC accounts for the contributions of Born effective charges and the dielectric constant, enabling precise predictions of phonon dispersion, as illustrated in Figs. 1(a)–(d).

The mechanical stability of a material is described by its strain energy (W), which is influenced by its elastic constants [91, 92]:

W=EstrEunstr=12i,jCi,jϵiϵj,

where Cij represents the second-order elastic tensor and ϵ denotes strain. Estr and Eunstr refer to the strained and unstrained states, respectively. Further details about the elastic tensors are provided in the Supplementary Materials. The corresponding mechanical stability criteria can be expressed as [92]

C11C12>0,C11+2C12>0,C44>0,

where the elastic constants Cij, shown in Table 3, satisfy the mechanical stability criteria, confirming the mechanical stability of the materials. Additionally, various mechanical property parameters were examined, including the bulk modulus (B), shear modulus (G), Pugh’s ratio (B/G) and Young’s modulus (E), as presented in Table 3. The results indicate that all four APVs exhibit similar mechanical properties, with B of around 48 GPa, G around 40 GPa, and E around 93 GPa. These values suggest that all four APVs possess significant resistance to deformation and low compressibility. Moreover, based on the Pugh ratio (B/G) (approximately 1.2), these APVs exhibit brittle behavior, as the values fall below the critical threshold defined by the Pugh criterion [93].

3.2 Lattice thermal conductivity

Lattice thermal conductivity (κL) is a key property for evaluating the thermal transport capabilities of materials, especially in applications such as thermoelectrics and thermal barrier coatings. In this study, κL was estimated using the Boltzmann Transport Equation (BTE) with inputs from self-consistent phonon (SCP) frequencies and eigenvectors, νq(T)=Ωq(T)/q. The SCP framework enables the renormalization of second-order force constants at each temperature, leading to effective harmonic force constants.

The computed κL for the four cubic APVs are illustrated in Fig. 2, where both three-phonon κ3phSCP and combined three- and four-phonon scattering κ3,4phSCP processes were considered. At 900 K, κ3phSCP for Sr3PN, Sr3AsN, Sr3SbN, and Sr3BiN amount to 1.48, 1.10, 1.38, and 1.74 Wm1K1, respectively. However, κ3,4phSCP shows significant reductions, with values decreasing to 0.57, 0.55, 0.84, and 1.09 Wm1K1, respectively, at the same temperature. Comparing the two cases, κ3,4phSCP is significantly lower than κ3phSCP, with reductions at 300 K of 37.2%, 38.5%, 17.6%, and 18.1% for Sr3PN, Sr3AsN, Sr3SbN, and Sr3BiN respectively. As the temperature increases to 900 K, these reductions become even more pronounced, reaching 61.5%, 50%, 39.1%, and 37.4%, respectively. This suggests that, even in weakly anharmonic materials, 4ph scattering plays a critical role in reducing thermal conductivity. Compared to other APV materials, these APVs exhibit relatively low κL. For instance, experimental data show that Ca3SnO has a κL of 17 Wm1K1 at 290 K [94], and calculations for SnMn3N report 3.593 Wm1K1 at 300 K [95]. The low thermal conductivity in these four APVs can be attributed to strong anharmonicity, which suppresses heat transport by enhancing phonon-phonon interactions. Furthermore, the κL(ω) is primarily concentrated in the low-frequency region, accounting for 77.8%, 79.2%, 84.1%, and 82.3% of the total thermal conductivity for Sr3PN, Sr3AsN, Sr3SbN, and Sr3BiN, respectively. The significant role of four-phonon scattering in suppressing κL makes these APVs promising candidates for applications where low thermal conductivity is essential. The low κL observed in APVs, combined with their strong anharmonicity, highlights their potential for enhanced thermoelectric performance.

Furthermore, the lattice thermal conductivity κL including the off-diagonal contributions was also calculated, as shown in Fig. S2. The off-diagonal terms slightly increase κL, indicating a moderate contribution from incoherent phonon transport. At 900 K, the calculated κL3,4ph,off values for Sr3PN, Sr3AsN, Sr3SbN, and Sr3BiN are 0.81, 0.73, 1.00, and 1.24 W·m−1·K−1, respectively, suggesting that the off-diagonal contributions are negligible.

Another notable observation is the unusually weak temperature dependence of κL within each cubic APV, particularly in Sr3PN and Sr3AsN. In general, lattice thermal conductivity (κL) follows a power law relationship, κLTγ, where γ is approximately 1 for most materials. However, in Sr3PN and Sr3AsN, the observed γ values deviate significantly from this universal behavior, with γ values of 0.58 and 0.71, respectively. This suggests a much weaker temperature dependence in these materials compared to the expected T1 trend. The presence of 4ph scattering further amplifies the temperature sensitivity, especially in materials with strong anharmonicity. In Sr3PN and Sr3AsN, anharmonic effects are so pronounced that imaginary phonon frequencies appear even at 0 K, leading to γ values of 1.02 and 0.94, respectively, under the influence of 4ph scattering. This increased temperature sensitivity results in a sharper reduction of thermal conductivity as temperature rises. In Sr3SbN and Sr3BiN, the γ values also increase due to 4ph scattering, from 0.8 to 1.04 in Sr3SbN, and from 0.87 to 1.09 in Sr3BiN. This heightened temperature sensitivity results in significant suppression of κL, highlighting the critical role of four-phonon scattering in limiting heat transport in these materials [96, 97].

3.3 Mode level phonon analysis

For a deeper understanding of the effects of 4ph scattering, a mode-level phonon analysis is necessary, starting with an examination of scattering rates (SRs). As shown in Fig. 4(a), the 4ph SRs are comparable in magnitude to the 3ph SRs, especially in materials with strong anharmonicity, such as Sr3PN and Sr3AsN. In fact, in the low-frequency region, the SRs for 4ph scattering even surpass those of 3ph scattering, underscoring the dominant role that 4ph processes play in these materials.

Similar to the behavior of 3ph scattering rates, the influence of 4ph SRs becomes more pronounced at higher temperatures, particularly for Sr3SbN and Sr3BiN, shown in Fig. 4(b). The temperature-dependent increase in SRs emphasizes the growing impact of 4ph processes as temperature rises, which further limits thermal conductivity. Furthermore, Sr3AsN has the highest SRs among the materials studied, consistent with its lowest κL. This relationship highlights the critical role of phonon scattering in determining thermal conductivity, particularly in materials where anharmonicity is strong and 4ph processes become significant.

Before conducting calculations, it can be predicted that 4ph scattering will reduce the lattice thermal conductivity κL of the material, as the Umklapp process is a key factor influencing thermal resistance. In 4ph scattering processes, the Umklapp (U) process plays a dominant role, as shown in Fig. 4(c). There are three types of 4ph scattering, q+q1q2+q3, qq1+q2+q3 and q+q1+q2q3, with the first type being the most prevalent, as shown in Fig. S4. The splitting and combination processes for 3ph scattering are illustrated in Fig. S3. In general, 3ph scattering leads to longer relaxation times. However, incorporating 4ph scattering into the calculations yields more accurate predictions of relaxation times, particularly at room temperature. This highlights the risk of overestimating relaxation times when only 3ph scattering is considered, underscoring the importance of accounting for 4ph scattering mechanisms [54]. Additionally, the scattering phase spaces for 3ph and 4ph processes were calculated, as shown in Fig. S5 and Fig. S6, respectively. The decomposed phase spaces and corresponding scattering rates exhibit similar trends, confirming that these four APVs exhibit large anharmonic scattering rates, which strongly affects thermal transport.

Phonon group velocity (υph) is a key parameter for analyzing the lattice thermal conductivity (κL) of materials, as it directly influences heat transport. Figure 5 is the υph at 300, 600, and 900 K. Notably, the υph values remain largely consistent across these temperatures, which reflects the similarity in the phonon spectrum. Observing low υph values across all materials supports the overall low κL. Although Sr3PN exhibits the smallest group velocity, this does not directly result in the lowest κL. This divergence can be attributed to the role of phonon lifetime, which significantly affects κL and is inversely correlated with the SRs. The relatively high SRs observed in Sr3AsN, which lead to shorter phonon lifetimes, may explain why Sr3AsN has the lowest κL despite not having the smallest υph. This emphasizes the importance of both group velocity and phonon lifetime in determining thermal conductivity.

To gain a deeper understanding of the microscopic mechanisms contributing to the low κL, the Grüneisen parameter (γ) was computed, which measures anharmonicity in the material. As illustrated in Fig. 5(b), a notable observation is that, compared to the optical phonon modes, the acoustic phonon modes exhibit higher values of γ. This suggests that acoustic phonons experience stronger cubic anharmonicity, directly impacting heat transport by enhancing phonon−phonon scattering.

3.4 Electronic transport

The electronic band structure and density of states for Sr3PN, Sr3AsN, Sr3SbN, and Sr3BiN provide crucial insights into the semiconducting behavior of these materials. As shown in Fig. 6, all four materials exhibit characteristics of direct band gap semiconductors, with the conduction band minimum (CBM) and valence band maximum (VBM) both located at the Γ point. The calculated band gap values using the PBEsol functional with spin−orbit coupling (SOC) are 0.38 eV for Sr3PN, 0.35 eV for Sr3AsN, 0.15 eV for Sr3SbN, and 0.17 eV for Sr3BiN. To improve the accuracy of band gap predictions, the HSE06 hybrid functional was employed. The refined band gap values are significantly higher, at 0.88 eV for Sr3PN, 0.92 eV for Sr3AsN, 1.19 eV for Sr3SbN, and 1.25 eV for Sr3BiN, reflecting the well-known tendency of PBEsol to underestimate band gaps. In terms of orbital contributions, the CBM is primarily dominated by Sr d orbital, while the VBM is mainly contributed by the p orbital of N and p orbital of X. This distribution of electronic states further characterizes the electronic structure of these APV materials and their potential for semiconducting applications.

In contrast to traditional perovskites (PVs), antiperovskites (APVs) exhibit a unique structural arrangement where cations occupy the positions typically reserved for anions in the octahedral structure. This distinctive arrangement, coupled with the abundance of X-site cations, imparts unconventional physical and chemical properties, particularly affecting d-spin states, band structure, and ion transport. These structural differences set APVs apart from traditional PVs, enabling unique electronic behavior. The valence bands of Sr3XN (X = P, As, Sb, Bi) arise from the hybridization of Sr 4d-N 2p and Sr 4d-X p states, as shown in previous studies [4, 98]. This hybridization is a key factor in shaping the electronic structure and influencing the material’s conductive properties. Additionally, the presence of Sr s states is constrained by the symmetry of the p states originating from X and N atoms [98]. This restriction further defines the behavior of the electronic states, contributing to its distinctive electronic properties.

Figure 7 displays an overview of the electronic transport properties, encompassing both p-type and n-type materials. To model rational electronic transport behavior, the band gap obtained from the HSE06 hybrid functional was used as the input parameter for these materials. As the carrier concentration (n) increases, both the electrical conductivity (σ) and the electronic thermal conductivity (κe) increase, while the Seebeck coefficient (S) decreases. This behavior is attributed to the growing participation of charge carriers in the conduction (σneμ) due to the elevated carrier concentration n. Conversely, at higher temperatures, σ decreases while S increases. When comparing p-type and n-type materials, n-type materials demonstrate larger σ, whereas p-type materials exhibit a larger S. This trend is consistent with the greater band dispersion observed at the conduction CBM for n-type materials and the greater degeneracy at the VBM for p-type materials, as shown in Fig. 6. In terms of power factor, p-type materials achieve values of 0.62, 0.81, 0.83, and 1.03 W·m−1·K−2 for Sr3PN, Sr3AsN, Sr3SbN, and Sr3BiN, respectively. However, the n-type counterparts display even higher maximum power factors of 0.78, 1.13, 1.13, and 1.62 W·m−1·K−2, respectively. There are highest power factors (>0.8) at 900 K for these two materials.

The ultimate goal of the analysis is to achieve excellent thermoelectric properties, which are quantified by the ZT, as illustrated in Fig. 8, particularly at nh at 900 K. This remarkable performance is driven by the combination of ultralow lattice thermal conductivity (κL) and a high power factor. The substantial electronic band dispersion at the CBM further enhances ZT for n-type materials, with ZT values of 0.62, 0.81, 0.83, and 1.03 at 900 K for Sr3PN, Sr3AsN, Sr3SbN, and Sr3BiN, respectively. For instance, Sr3SbN and Sr3BiN exhibit exceptional ZT values at ne around 3×1018 cm−3. Importantly, the highest ZT for n-type is significantly larger than p-type, suggesting that n-type variants are more promising for thermoelectric applications. In practical experiments, factors such as thermal radiation, air-induced thermal convection, and the impact of grain boundary scattering on carrier mobility (μ) may lead to slight reductions in the calculated ZT values [99]. Nevertheless, n-type materials remain strong candidates for high-performance thermoelectric devices due to their superior thermoelectric properties.

4 Conclusion

To summarize, this study investigated into the thermal and electrical transport characteristics and the thermoelectric potential of APV materials (Sr3XN; X = P, As, Sb, Bi). As previously explained, the significance of quartic anharmonicity has been emphasized in all the materials under consideration. Similar to 3ph scattering, the prominence of 4ph scattering is dictated by two crucial factors: the anharmonicity of the energy landscape and the available scattering phase space. The influence of 4ph interactions extends beyond phonon self-energy adjustments, as it also diminishes the temperature sensitivity of κL, ultimately impacting the overall thermoelectric performance. It is important to highlight that the computation of both three- and four-phonon scattering rates is a challenging task involving intricate interactions among matrix elements, eigenvectors, and frequencies, which collectively encompass various phonon modes. Moreover, the unique combination of highly dispersive band edges and electronic band degeneracy in Sr3XN leads to outstanding electronic conductivity and a substantial Seebeck coefficient. The optimal power factor is realized in Sr3BiN, ultimately resulting in the maximum ZT value of 1.03.

Beyond bulk APVs, emerging two-dimensional APV monolayers have been predicted to be structurally robust and to exhibit electronic transport favorable for thermoelectricity, highlighting 2D APVs as an intriguing platform worthy of attention [100102]. Our study not only highlights the exceptional thermoelectric properties of APV Sr3BiN but also underscores the importance of quartic anharmonicity for κL and ZT, which significantly impacts the thermoelectric performance of the material. This research provides valuable insights for designing and optimizing thermoelectric properties in strongly anharmonic materials moving forward.

References

[1]

P. Mahajan , B. Padha , S. Verma , V. Gupta , R. Datt , W. C. Tsoi , S. Satapathi , and S. Arya , Review of current progress in hole-transporting materials for perovskite solar cells, J. Energy Chem. 68, 330 (2022)

[2]

Y. Kang , Antiperovskite Sr3MN and Ba3MN (M = Sb or Bi) as promising photovoltaic absorbers for thin-film solar cells: A first-principles study, J. Am. Ceram. Soc. 105(9), 5807 (2022)

[3]

J. Zheng , B. Perry , and Y. Wu , Antiperovskite superionic conductors: A critical review, ACS Mater. Au 1(2), 92 (2021)

[4]

M. Ochi and K. Kuroki , Comparative first-principles study of antiperovskite oxides and nitrides as thermoelectric material: Multiple Dirac cones, low-dimensional band dispersion, and high valley degeneracy, Phys. Rev. Appl. 12(3), 034009 (2019)

[5]

S. Pokrant , A. E. Maegli , G. L. Chiarello , and A. Weidenkaff , Perovskite-related oxynitrides in photocatalysis, Chimia (Aarau) 67(3), 162 (2013)

[6]

W. Tang , J. Zhang , S. Ratnasingham , F. Liscio , K. Chen , T. Liu , K. Wan , E. S. Galindez , E. Bilotti , M. Reece , M. Baxendale , S. Milita , M. A. McLachlan , L. Su , and O. Fenwick , Substitutional doping of hybrid organic–inorganic perovskite crystals for thermoelectrics, J. Mater. Chem. A 8(27), 13594 (2020)

[7]

M. R. Ahmadian-Yazdi , S. Lin , and Z. Cai , Unveiling heavy heterovalent doping-modulated microstructure and thermoelectric performance in bulk hybrid perovskite single crystals, Chem. Eng. J. 487, 150477 (2024)

[8]

X. Li , Y. Zhang , W. Kang , Z. Yan , Y. Shen , and J. Huo , Anti-perovskite nitrides and oxides: Properties and preparation, Comput. Mater. Sci. 225, 112188 (2023)

[9]

Y. Mochizuki , H. J. Sung , A. Takahashi , Y. Kumagai , and F. Oba , Theoretical exploration of mixed-anion antiperovskite semiconductors M3XN (M = Mg, Ca, Sr, Ba; X = P, As, Sb, Bi), Phys. Rev. Mater. 4(4), 044601 (2020)

[10]

H. K. Singh , Z. Zhang , I. Opahle , D. Ohmer , Y. Yao , and H. Zhang , High-throughput screening of magnetic antiperovskites, Chem. Mater. 30(20), 6983 (2018)

[11]

R. Masrour , A. Jabar , L. Bahmad , E. Hlil , M. Hamedoun , A. Benyoussef , A. Hourmatallah , N. Benzakour , A. Rezzouk , and K. Bouslykhane , Magnetic properties of Mn3ZnN antiperovskite nanoparticles: A Monte Carlo simulations, J. Cluster Sci. 32(1), 163 (2021)

[12]

H. K. Singh , I. Samathrakis , N. M. Fortunato , J. Zemen , C. Shen , O. Gutfleisch , and H. Zhang , Multifunctional antiperovskites driven by strong magnetostructural coupling, npj Comput. Mater. 7(1), 98 (2021)

[13]

P. Tong , D. Louca , G. King , A. Llobet , J. Lin , and Y. Sun , Magnetic transition broadening and local lattice distortion in the negative thermal expansion antiperovskite Cu1−xSnxNMn3, Appl. Phys. Lett. 102(4), 041908 (2013)

[14]

K. Takenaka , T. Hamada , D. Kasugai , and N. Sugimoto , Tailoring thermal expansion in metal matrix composites blended by antiperovskite manganese nitrides exhibiting giant negative thermal expansion, J. Appl. Phys. 112(8), 083517 (2012)

[15]

Y. Sun , C. Wang , Y. Wen , L. Chu , H. Pan , M. Nie , and M. Tang , Negative thermal expansion and magnetic transition in anti-perovskite structured Mn3Zn1−xSnxN compounds, J. Am. Ceram. Soc. 93(8), 2178 (2010)

[16]

W. Xia , Y. Zhao , F. Zhao , K. Adair , R. Zhao , S. Li , R. Zou , Y. Zhao , and X. Sun , Antiperovskite electrolytes for solid-state batteries, Chem. Rev. 122(3), 3763 (2022)

[17]

K. Kim and D. J. Siegel , Multivalent ion transport in anti-perovskite solid electrolytes, Chem. Mater. 33(6), 2187 (2021)

[18]

T. He , Q. Huang , A. Ramirez , Y. Wang , K. Regan , N. Rogado , M. Hayward , M. Haas , J. Slusky , K. Inumara , H. W. Zandbergen , N. P. Ong , and R. J. Cava , Superconductivity in the non-oxide perovskite MgCNi3, Nature 411(6833), 54 (2001)

[19]

H. M. Tahir Farid , A. Mera , T. I. Al-Muhimeed , A. A. AlObaid , H. Albalawi , H. H. Hegazy , S. R. Ejaz , R. Y. Khosa , S. Mehmood , and Z. Ali , Optoelectronic and thermoelectric properties of A3AsN (A = Mg, Ca, Sr and Ba) in cubic and orthorhombic phase, J. Mater. Res. Technol. 13, 1485 (2021)

[20]

H. Lin , G. Tan , J. N. Shen , S. Hao , L. M. Wu , N. Calta , C. Malliakas , S. Wang , C. Uher , C. Wolverton , and M. G. Kanatzidis , Concerted rattling in CsAg5Te3 leading to ultralow thermal conductivity and high thermoelectric performance, Angew. Chem. Int. Ed. 55(38), 11431 (2016)

[21]

Y. Zhao , S. Zeng , G. Li , C. Lian , Z. Dai , S. Meng , and J. Ni , Lattice thermal conductivity including phonon frequency shifts and scattering rates induced by quartic anharmonicity in cubic oxide and fluoride perovskites, Phys. Rev. B 104(22), 224304 (2021)

[22]

C. Shen , M. Dai , X. Xiao , N. Hadaeghi , W. Xie , A. Weidenkaff , T. Tadano , and H. Zhang , Impact of quartic anharmonicity on lattice thermal transport in EuTiO3: A comparative theoretical and experimental investigation, Mater. Today Phys. 34, 101059 (2023)

[23]

M. Bilal , S. Jalali-Asadabadi , R. Ahmad , and I. Ahmad , Electronic properties of antiperovskite materials from state-of-the-art density functional theory, J. Chem. 2015, 1 (2015)

[24]

P. Sreedevi , R. Vidya , and P. Ravindran , Antiperovskite materials as promising candidates for efficient tandem photovoltaics: First-principles investigation, Mater. Sci. Semicond. Process. 147, 106727 (2022)

[25]

E. Haque and M. A. Hossain , DFT based study on structural stability and transport properties of Sr3AsN: A potential thermoelectric material, J. Mater. Res. 34(15), 2635 (2019)

[26]

J. Wakini,C. Songa,S. Chege,F. O. Saouma,E. Wabululu,P. Nyawere,V. Odari,J. Sifuna,G. S. Manyali, Low born effective charges, high covalency and strong optical activity in X2+Bi3−N3− (X = Ca, Sr, Ba) inverse-perovskites, arXiv: 2022)

[27]

C. Chang and L. D. Zhao , Anharmoncity and low thermal conductivity in thermoelectrics, Mater. Today Phys. 4, 50 (2018)

[28]

P. Aynajian, Phonons and Their Interactions, Springer Berlin Heidelberg, Berlin, Heidelberg, 2011, pp 7–13

[29]

P. G. Klemens, Thermal Conductivity and Lattice Vibrational Modes, in: Solid state physics, Vol. 7, Elsevier, 1958, pp 1–98

[30]

A. A. Maradudin,E. W. Montroll,G. H. Weiss,I. Ipatova, Theory of Lattice Dynamics in the Harmonic Approximation, Vol. 3, Academic press New York, 1963

[31]

J. Yue , Y. Liu , and J. Zheng , Interlayer thermal transport and glasslike behavior in crystalline CsCu4Se3, Phys. Rev. B 111(2), 024313 (2025)

[32]

N. Wang , J. Yue , S. Guo , H. Zhang , S. Li , M. Cui , Y. Liu , and T. Cui , Pressure induced enhancement of anharmonicity and optimization of thermoelectric properties in n-type SnS, Front. Phys. (Beijing) 20(3), 034206 (2025)

[33]

H. Yu , H. Zhang , J. Zhao , J. Liu , X. Xia , and X. Wu , Thermal conductivity of micro/nanoporous polymers: Prediction models and applications, Front. Phys. (Beijing) 17(2), 23202 (2022)

[34]

Y. Chen , D. Wang , Y. Zhou , Q. Pang , J. Shao , G. Wang , J. Wang , and L. D. Zhao , Enhancing the thermoelectric performance of Bi2S3: A promising earth-abundant thermoelectric material, Front. Phys. (Beijing) 14(1), 13601 (2019)

[35]

X. Yang , J. Tiwari , and T. Feng , Reduced anharmonic phonon scattering cross-section slows the decrease of thermal conductivity with temperature, Mater. Today Phys. 24, 100689 (2022)

[36]

R. Guo , X. Wang , Y. Kuang , and B. Huang , First-principles study of anisotropic thermoelectric transport properties of IV−VI semiconductor compounds SnSe and SnS, Phys. Rev. B 92(11), 115202 (2015)

[37]

J. H. Fetherolf , P. Shih , and T. C. Berkelbach , Conductivity of an electron coupled to anharmonic phonons: Quantum-classical simulations and comparison of approximations, Phys. Rev. B 107(6), 064304 (2023)

[38]

X. Ding , X. Jin , Z. Chang , D. Li , X. Zhou , X. Yang , and R. Wang , Anharmonicity-induced phonon hardening and anomalous thermal transport in ScZn, Phys. Rev. B 110(5), 054304 (2024)

[39]

L. D. Zhao , S. H. Lo , Y. Zhang , H. Sun , G. Tan , C. Uher , C. Wolverton , V. P. Dravid , and M. G. Kanatzidis , Ultralow thermal conductivity and high thermoelectric figure of merit in SnSe crystals, Nature 508(7496), 373 (2014)

[40]

M. Zacharias , G. Volonakis , F. Giustino , and J. Even , Anharmonic electron−phonon coupling in ultrasoft and locally disordered perovskites, npj Comput. Mater. 9(1), 153 (2023)

[41]

T. Yildirim , O. Gülseren , J. Lynn , C. Brown , T. Udovic , Q. Huang , N. Rogado , K. Regan , M. Hayward , J. Slusky , T. He , M. K. Haas , P. Khalifah , K. Inumaru , and R. J. Cava , Giant anharmonicity and nonlinear electron−phonon coupling in MgB2: A combined first-principles calculation and neutron scattering study, Phys. Rev. Lett. 87(3), 037001 (2001)

[42]

P. B. Allen , Anharmonic phonon quasiparticle theory of zero-point and thermal shifts in insulators: Heat capacity, bulk modulus, and thermal expansion, Phys. Rev. B 92(6), 064106 (2015)

[43]

A. Togo , L. Chaput , and I. Tanaka , Distributions of phonon lifetimes in Brillouin zones, Phys. Rev. B 91(9), 094306 (2015)

[44]

W. Li , J. Carrete , N. A. Katcho , and N. Mingo , ShengBTE: A solver of the Boltzmann transport equation for phonons, Comput. Phys. Commun. 185(6), 1747 (2014)

[45]

T. Tadano and S. Tsuneyuki , First-principles lattice dynamics method for strongly anharmonic crystals, J. Phys. Soc. Jpn. 87(4), 041015 (2018)

[46]

Y. Oba , T. Tadano , R. Akashi , and S. Tsuneyuki , First-principles study of phonon anharmonicity and negative thermal expansion in ScF3, Phys. Rev. Mater. 3(3), 033601 (2019)

[47]

T. Tadano and S. Tsuneyuki , Quartic anharmonicity of rattlers and its effect on lattice thermal conductivity of clathrates from first principles, Phys. Rev. Lett. 120(10), 105901 (2018)

[48]

X. Xiao , W. Xie , K. Philippi , Y. Liu , K. Skokov , I. A. Radulov , M. Widenmeyer , A. Kovalevsky , C. Shen , H. Zhang , S. Checchia , M. Scavini , J. He , and A. Weidenkaff , Anomalous thermal conductivity of alkaline-earth-metal-substituted EuTiO3 induced by resonant scattering, Mater. Today Phys. 35, 101132 (2023)

[49]

T. Tadano and S. Tsuneyuki , Self-consistent phonon calculations of lattice dynamical properties in cubic SrTiO3 with first-principles anharmonic force constants, Phys. Rev. B 92(5), 054301 (2015)

[50]

I. Errea , M. Calandra , and F. Mauri , Anharmonic free energies and phonon dispersions from the stochastic self-consistent harmonic approximation: Application to platinum and palladium hydrides, Phys. Rev. B 89(6), 064302 (2014)

[51]

F. Knoop , N. Shulumba , A. Castellano , J. P. Alvarinhas Batista , R. Farris , M. J. Verstraete , M. Heine , D. Broido , D. S. Kim , J. Klarbring , . TDEP: Temperature dependent effective potentials, J. Open Source Softw. 9(94), 6150 (2024)

[52]

D. B. Zhang , J. G. Li , Y. H. Ren , and T. Sun , Green-kubo formalism for thermal conductivity with Slater-Koster tight binding, Phys. Rev. B 108(10), 104307 (2023)

[53]

J. Yue , S. Guo , J. Li , J. Zhao , C. Shen , H. Zhang , Y. Liu , and T. Cui , Pressure-induced remarkable four-phonon interaction and enhanced thermoelectric conversion efficiency in CuInTe2, Mater. Today Phys. 39, 101283 (2023)

[54]

T. Feng , L. Lindsay , and X. Ruan , Four-phonon scattering significantly reduces intrinsic thermal conductivity of solids, Phys. Rev. B 96(16), 161201 (2017)

[55]

T. Feng and X. Ruan , Four-phonon scattering reduces intrinsic thermal conductivity of graphene and the contributions from flexural phonons, Phys. Rev. B 97(4), 045202 (2018)

[56]

H. K. Singh , A. Sehrawat , C. Shen , I. Samathrakis , I. Opahle , H. Zhang , and R. Xie , High-throughput screening of half-antiperovskites with a stacked Kagome lattice, Acta Mater. 242, 118474 (2023)

[57]

N. R. Werthamer , Self-consistent phonon formulation of anharmonic lattice dynamics, Phys. Rev. B 1(2), 572 (1970)

[58]

P. Hohenberg and W. Kohn , Inhomogeneous electron gas, Phys. Rev. 136(3B), B864 (1964)

[59]

G. Kresse and J. Furthmüller , Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set, Phys. Rev. B 54(16), 11169 (1996)

[60]

G. I. Csonka , J. P. Perdew , A. Ruzsinszky , P. H. Philipsen , S. Lebègue , J. Paier , O. A. Vydrov , and J. G. Ángyán , Assessing the performance of recent density functionals for bulk solids, Phys. Rev. B 79(15), 155107 (2009)

[61]

A. M. Ganose , J. Park , A. Faghaninia , R. Woods-Robinson , K. A. Persson , and A. Jain , Efficient calculation of carrier scattering rates from first principles, Nat. Commun. 12(1), 2222 (2021)

[62]

K. Esfarjani and H. T. Stokes , Method to extract anharmonic force constants from first principles calculations, Phys. Rev. B 77(14), 144112 (2008)

[63]

F. Zhou , B. Sadigh , D. Åberg , Y. Xia , and V. Ozolins , Compressive sensing lattice dynamics. ii. Efficient phonon calculations and long-range interactions, Phys. Rev. B 100(18), 184309 (2019)

[64]

L. J. Nelson , G. L. W. Hart , F. Zhou , and V. Ozolins , Compressive sensing as a paradigm for building physics models, Phys. Rev. B 87(3), 035125 (2013)

[65]

F. Zhou , W. Nielson , Y. Xia , and V. Ozolins , Lattice anharmonicity and thermal conductivity from compressive sensing of first-principles calculations, Phys. Rev. Lett. 113(18), 185501 (2014)

[66]

T. Tadano , Y. Gohda , and S. Tsuneyuki , Anharmonic force constants extracted from first-principles molecular dynamics: Applications to heat transfer simulations, J. Phys.: Condens. Matter 26(22), 225402 (2014)

[67]

Z. Han , X. Yang , W. Li , T. Feng , and X. Ruan , Fourphonon: An extension module to shengbte for computing four-phonon scattering rates and thermal conductivity, Comput. Phys. Commun. 270, 108179 (2022)

[68]

T. Feng and X. Ruan , Quantum mechanical prediction of four-phonon scattering rates and reduced thermal conductivity of solids, Phys. Rev. B 93(4), 045202 (2016)

[69]

M. Omini and A. Sparavigna , An iterative approach to the phonon Boltzmann equation in the theory of thermal conductivity, Physica B 212(2), 101 (1995)

[70]

Y. Xia,V. Ozolins,C. Wolverton, Microscopic mechanisms of glasslike lattice thermal transport in cubic Cu12Sb4S13 tetrahedrites, Phys. Rev. Lett. 125(8), 085901 (2020)

[71]

Y. Xia , D. II Gaines , J. He , K. Pal , Z. Li , M. G. Kanatzidis , V. Ozolins , and C. Wolverton , A unified understanding of minimum lattice thermal conductivity, Proc. Natl. Acad. Sci. USA 120(26), e2302541120 (2023)

[72]

J. Yue , Y. Liu , and J. Zheng , Interlayer thermal transport and glasslike behavior in crystalline CsCu4Se3, Phys. Rev. B 111(2), 024313 (2025)

[73]

A. F. Ioffe , L. Stil’Bans , E. Iordanishvili , T. Stavitskaya , A. Gelbtuch , and G. Vine-yard , Semiconductor thermoelements and thermoelectric cooling, Phys. Today 12(5), 42 (1959)

[74]

H. J. Goldsmid, ., Introduction to Thermoelectricity, Vol. 121, Springer, 2010

[75]

K. Momma and F. Izumi , Vesta for three-dimensional visualization of crystal, volumetric and morphology data, J. Appl. Crystall. 44(6), 1272 (2011)

[76]

I. Ullah , G. Murtaza , R. Khenata , A. Mahmood , M. Muzzamil , N. Amin , and M. Saleh , Structural and optoelectronic properties of X3ZN (X = Ca, Sr, Ba; Z = As, Sb, Bi) anti-perovskite compounds, J. Electron. Mater. 45(6), 3059 (2016)

[77]

M. Hichour , R. Khenata , D. Rached , M. Hachemaoui , A. Bouhemadou , A. Reshak , and F. Semari , FP-APW+lo study of the elastic, electronic and optical properties for the cubic antiperovskite ANSr3 (A = As, Sb and Bi) under pressure effect, Physica B 405(7), 1894 (2010)

[78]

F. Gäbler , M. Kirchner , W. Schnelle , U. Schwarz , M. Schmitt , H. Rosner , and R. Niewa , (Sr3N)E and (Ba3N)E (E = Sb, Bi): Synthesis, crystal structures, and physical properties, Z. Anorg. Allg. Chem. 630(13−14), 2292 (2004)

[79]

M. Y. Chern , F. DiSalvo , J. Parise , and J. A. Goldstone , The structural distortion of the anti-perovskite nitride ca3asn, J. Solid State Chem. 96(2), 426 (1992)

[80]

S. Kirklin , J. E. Saal , B. Meredig , A. Thompson , J. W. Doak , M. Aykol , S. Ruhl , and C. Wolverton , The open quantum materials database (OQMD): assessing the accuracy of DFT formation energies, npj Comput. Mater. 1(1), 1 (2015)

[81]

C. Shen , Q. Gao , N. M. Fortunato , H. K. Singh , I. Opahle , O. Gutfleisch , and H. Zhang , Designing of magnetic mab phases for energy applications, J. Mater. Chem. A 9(13), 8805 (2021)

[82]

J. Yue , J. Zheng , X. Shen , K. Maji , C. C. Yang , S. Lin , P. Lemoine , E. Guilmeau , Y. Liu , and T. Cui , Diffuson-dominated thermal transport crossover from ordered to liquid-like Cu3BiS3: The negligible role of ion hopping, Small 21(42), e06386 (2025)

[83]

C. Feng , C. Wu , X. Luo , T. Hu , F. Chen , S. Li , S. Duan , W. Hou , D. Li , G. Tang , and G. Zhang , Pressure-dependent electronic, optical, and mechanical properties of antiperovskite X3NP (X = Ca, Mg): A first-principles study, J. Semicond. 44(10), 102101 (2023)

[84]

C. J. Bartel , A. W. Weimer , S. Lany , C. B. Musgrave , and A. M. Holder , The role of decomposition reactions in assessing first-principles predictions of solid stability, npj Comput. Mater. 5(1), 4 (2019)

[85]

J. Zheng , D. Shi , Y. Yang , C. Lin , H. Huang , R. Guo , and B. Huang , Anharmonicity-induced phonon hardening and phonon transport enhancement in crystalline perovskite BaZrO3, Phys. Rev. B 105(22), 224303 (2022)

[86]

V. M. Goldschmidt , Die gesetze der krystallochemie, Naturwissenschaften 14(21), 477 (1926)

[87]

H. Yokokawa , N. Sakai , T. Kawada , and M. Dokiya , Thermodynamic stabilities of perovskite oxides for electrodes and other electrochemical materials, Solid State Ion. 52(1−3), 43 (1992)

[88]

K. Kreuer , On the development of proton conducting materials for technological applications, Solid State Ion. 97(1−4), 1 (1997)

[89]

G. Nagabhushana , R. Shivaramaiah , and A. Navrotsky , Direct calorimetric verification of thermodynamic instability of lead halide hybrid perovskites, Proc. Natl. Acad. Sci. USA 113(28), 7717 (2016)

[90]

N. H. Alotaibi , Frist principle study of double perovskites Cs2AgSbX6 (X = Cl, Br, I) for solar cell and renewable energy applications, J. Phys. Chem. Solids 171, 110984 (2022)

[91]

M. Born , K. Huang , and M. Lax , Dynamical theory of crystal lattices, Am. J. Phys. 23(7), 474 (1955)

[92]

F. Mouhat and F. X. Coudert , Necessary and sufficient elastic stability conditions in various crystal systems, Phys. Rev. B 90(22), 224104 (2014)

[93]

S. Pugh , Xcii. relations between the elastic moduli and the plastic properties of poly-crystalline pure metals, Lond. Edinb. Dublin Philos. Mag. J. Sci. 45(367), 823 (1954)

[94]

Y. Okamoto , A. Sakamaki , and K. Takenaka , Thermoelectric properties of antiperovskite calcium oxides Ca3PbO and Ca3SnO, J. Appl. Phys. 119(20), 205106 (2016)

[95]

Y. C. Alaoui , A. Jabar , N. Tahiri , O. El Bounagui , and H. Ez-Zahraouy , A large magnetocaloric effect and thermoelectric properties of anti-perovskite snmn3n material: A DFT study and Monte Carlo simulation, Phys. Scr. 98(9), 095942 (2023)

[96]

J. Yue , Y. Liu , W. Ren , S. Lin , C. Shen , H. K. Singh , T. Cui , T. Tadano , and H. Zhang , Role of atypical temperature-responsive lattice thermal transport on the thermoelectric properties of antiperovskites Mg3XN (X= P, As, Sb, Bi), Mater. Today Phys. 41, 101340 (2024)

[97]

S. Lin , J. Yue , W. Ren , C. Shen , and H. Zhang , Strong anharmonicity and medium-temperature thermoelectric efficiency in antiperovskite Ca3XN (X = P, As, Sb, Bi) compounds, J. Mater. Chem. A 12(30), 19567 (2024)

[98]

Y. Wang , H. Zhang , J. Zhu , X. , S. Li , R. Zou , and Y. Zhao , Antiperovskites with exceptional functionalities, Adv. Mater. 32(7), 1905007 (2020)

[99]

T. Yue , Y. Zhao , J. Ni , S. Meng , and Z. Dai , Strong quartic anharmonicity, ultralow thermal conductivity, high band degeneracy and good thermoelectric performance in Na2TlSb, npj Comput. Mater. 9(1), 17 (2023)

[100]

S. M. Alay-e-Abbas , G. Abbas , W. Zulfiqar , M. Sajjad , N. Singh , and J. A. Larsson , Structure inversion asymmetry enhanced electronic structure and electrical transport in 2D A3SnO (A = Ca, Sr, and Ba) anti-perovskite monolayers, Nano Res. 16(1), 1779 (2023)

[101]

Y. Zhang , B. Feng , H. Hayashi , C. P. Chang , Y. M. Sheu , I. Tanaka , Y. Ikuhara , and H. Ohta , Double thermoelectric power factor of a 2D electron system, Nat. Commun. 9(1), 2224 (2018)

[102]

D. Li , Y. Gong , Y. Chen , J. Lin , Q. Khan , Y. Zhang , Y. Li , H. Zhang , and H. Xie , Recent progress of two-dimensional thermoelectric materials, Nano-Micro Lett. 12(1), 36 (2020)

RIGHTS & PERMISSIONS

Higher Education Press

AI Summary AI Mindmap
PDF (6202KB)

Supplementary files

supporting information

371

Accesses

0

Citation

Detail

Sections
Recommended

AI思维导图

/