Pol
Benítez
*ab,
Siyu
Chen
cd,
Ruoshi
Jiang
c,
Cibrán
López
ab,
Josep-Lluís
Tamarit
ab,
Jorge
Íñiguez-González
ef,
Edgardo
Saucedo
bg,
Bartomeu
Monserrat
cd and
Claudio
Cazorla
*ab
aGroup of Characterization of Materials, Departament de Física, Universitat Politècnica de Catalunya, Campus Diagonal Besòs, Av. Eduard Maristany 10–14, 08019 Barcelona, Spain. E-mail: pol.benitez@upc.edu; claudio.cazorla@upc.edu
bResearch Center in Multiscale Science and Engineering, Universitat Politècnica de Catalunya, Campus Diagonal-Besòs, Av. Eduard Maristany 10–14, 08019 Barcelona, Spain
cDepartment of Materials Science and Metallurgy, University of Cambridge, Cambridge, CB30FS, UK
dCavendish Laboratory, University of Cambridge, Cambridge, CB30HE, UK
eMaterials Research and Technology Department, Luxembourg Institute of Science and Technology (LIST), Avenue des Hauts-Fourneaux 5, L-4362 Esch/Alzette, Luxembourg
fDepartment of Physics and Materials Science, University of Luxembourg, 41 Rue du Brill, L-4422 Belvaux, Luxembourg
gMicro and Nanotechnologies Group, Emerging Thin Film Photovoltaics Lab, Departament dEnginyeria Electrònica, Universitat Politècnica de Catalunya, Campus Diagonal Besòs, Av. Eduard Maristany 10–14, 08019 Barcelona, Spain
First published on 14th April 2025
Silver chalcohalide antiperovskites (CAP), Ag3XY (X = S, Se; Y = Br, I), are a family of highly anharmonic inorganic compounds with great potential for energy applications. However, a substantial and unresolved discrepancy exists between the optoelectronic properties predicted by theoretical first-principles methods and those measured experimentally at room temperature, hindering the fundamental understanding and rational engineering of CAP. In this work, we employ density functional theory, tight-binding calculations, and anharmonic Fröhlich theory to investigate the optoelectronic properties of CAP at finite temperatures. Near room temperature, we observe a giant band-gap (Eg) reduction of approximately 20–60% relative to the value calculated at T = 0 K, bringing the estimated Eg into excellent agreement with experimental measurements. This relative T-induced band-gap renormalization is roughly twice the largest value previously reported in the literature for similar temperature ranges. Low-energy optical polar phonon modes, which break inversion symmetry and enhance the overlap between silver and chalcogen s electronic orbitals in the conduction band, are identified as the primary drivers of this significant Eg reduction. Furthermore, when temperature effects are considered, the optical absorption coefficient of CAP increases by nearly an order of magnitude in the visible light spectrum. These findings not only bridge a critical gap between theory and experiment but also pave the way for future technologies where temperature, electric fields, and light dynamically modulate optoelectronic properties, establishing CAP as a versatile platform for energy and photonic applications.
Likewise, the band gap (Eg) of semiconducting and dielectric materials can be significantly affected by EPC, typically decreasing with increasing temperature (the so-called Varshni effect7). This common Eg behavior can be explained by the Allen–Heine–Cardona perturbative theory, which attributes it to a larger T-induced energy increase in the valence band compared to the conduction band due to a greater sensitivity to phonon population variations (i.e., larger second-order electron–phonon coupling constants).8–10 Representative examples of this thermal Eg dependence include diamond, which exhibits a ∼5% band-gap reduction at around 1000 K;11,12 antimony sulfide (Sb2S3), which shows a Eg reduction of 200 meV in the temperature range of 10 ≤ T ≤ 300 K;13 MgO, which displays a band-gap reduction of ∼15% in the temperature interval 0 ≤ T ≤ 1500 K;14 SrTiO3, which exhibits a ∼15% band-gap reduction from 300 to 1000 K;15 and molecular crystals, which display record band-gap reductions of 15–20% at low temperatures.16 Anomalous band-gap thermal behaviour, in which Eg increases with increasing temperature, has also been observed in a variety of materials such as black phosphorus,17 halide perovskites,18 and chalcopyrite19 and hydride12 compounds.
Highly anharmonic silver chalcohalide antiperovskites (CAP)20 with chemical formula Ag3XY (X = S, Se; Y = Br, I) are structurally similar to lead halide perovskites (e.g., CsPbI3), with the “anti” designation indicating the exchange of anions and cations compared to the typical ionic perovskite arrangement. Analogous to lead halide perovskites, CAP are highly promising materials for energy and optoelectronic applications,21–26 offering low toxicity due to their lead-free composition.27,28 The two most extensively studied CAP compounds, Ag3SBr and Ag3SI, possess experimentally determined band gaps of approximately 1.0 eV,29,30 making them favorable for photovoltaic applications. These materials have also been recognized as high temperature superionic conductors.21,22 Additionally, CAP have been investigated as potential thermoelectric materials25,26 owing to their substantial vibrational anharmonicity and unique charge transport properties.23,24
Intriguingly, for both Ag3SBr and Ag3SI, there is an enormous disagreement between the Eg predicted by first-principles methods (at T = 0 K, under static lattice conditions) and those measured experimentally at room temperature. In particular, high-level density functional theory (DFT) calculations employing hybrid functionals and including spin–orbit coupling (SOC) effects estimate the band gap of these two archetypal CAP to be 1.8 and 1.4 eV, respectively.29–31 The Eg discrepancies between theory and measurements amount to 60–80% of the experimental values (i.e., differences of 0.5–0.8 eV), which are unusually large and call for a careful inspection of the factors causing them.
In this study, we assessed the influence of EPC effects on the Eg and optical absorption spectra of CAP using first-principles DFT methods, tight-binding calculations, and anharmonic Fröhlich theory. Near room temperature, our computational investigations revealed a giant Eg reduction of 20–60% relative to the value calculated at T = 0 K, bringing the estimated band gap into excellent agreement with the experimental values. Low-energy optical polar phonons, which cause large symmetry-breaking structural distortions and promote the overlap between silver and chalcogen s electronic orbitals in the conduction band, were identified to be the primary mechanism driving this substantial T-induced band-gap reduction. Furthermore, at finite temperatures the optical absorption spectra of CAP were significantly enhanced, in some cases by nearly an order of magnitude. The polar nature of the phonons causing these effects opens up new technological possibilities, where the optoelectronic properties of materials could be effectively manipulated by external electric fields and light.
![]() | ||
Fig. 1 General physical properties of the archetypal CAP Ag3SBr. (a) The cubic Pm![]() ![]() ![]() |
As discussed in the Introduction, the discrepancies between the experimentally measured (at T = 300 K) and theoretically determined (at T = 0 K) band gaps of Ag3SBr and Ag3SI are tremendously large (i.e., 60–80% of the experimental values). Therefore, we investigated their potential causes by assessing the impact of electron–phonon coupling (EPC) and temperature on the band gap of CAP.
Band gap renormalization due to electron–phonon interactions is typically estimated using two main approaches: density functional perturbation theory (DFPT)1 and finite-differences.2 In DFPT, electron–phonon interactions are treated as a perturbation, with band energy variations (and consequently, band gap shifts) derived from the Fan-Migdal and Debye–Waller self-energies, as described by Allen and Heine.36 A key advantage of DFPT is its computational efficiency, as it does not require the use of supercells. This method has been widely employed in band gap renormalization studies37 and is implemented in popular ab initio codes such as EPW.38
Finite-differences approaches, on the other hand, are computationally more demanding, requiring supercells to reproduce phonons in real space and dynamical simulations to correctly sample the electronic response to ionic fluctuations. However, they offer distinct advantages over DFPT. One key benefit is their flexibility, as they can be applied with any underlying electronic structure method. Additionally, finite-differences methods naturally incorporate terms beyond the lowest order in the electron–phonon interaction, making them particularly useful for capturing higher-order effects.18 Readers seeking a more comprehensive discussion of DFPT and finite-differences methods are referred to the review articles,1,2 which extensively cover these techniques.
In this study, we employ the finite-differences approach to estimate temperature-renormalized band gaps, as CAP materials exhibit strong anharmonicity.20 Consequently, renormalizing their electronic band energies at the harmonic level would be inadequate. Moreover, since accurate CAP band gap predictions require the use of hybrid functionals and spin–orbit coupling, the finite-differences approach emerges as the most practical and reliable choice.
In particular, we performed first-principles calculations and ab initio molecular dynamics (AIMD) simulations based on DFT (Methods). Additionally, to capture long-range EPC effects, we employed anharmonic Fröhlich theory39–42 considering long-range dipole–dipole interactions and T-renormalized phonons (Methods). Furthermore, the optical absorption spectra of all CAP were assessed at T ≠ 0 K conditions and the main EPC mechanisms underlying the Eg discrepancies were identified with the help of a tight-binding model.
Eg(T) = Eg(0) + ΔEg(T), | (1) |
ΔEg(T) = ΔESg(T) + ΔELg(T). | (2) |
The short-wavelength phonon correction was estimated through AIMD simulations using a supercell (Methods), where the band-gap value was averaged over multiple generated configurations, as described in ref. 39:
![]() | (3) |
In polar materials, there is an additional contribution to the band-gap renormalization stemming from long-range Fröhlich coupling that is not fully captured by the finite size of the supercells employed in the AIMD simulations.39–43 This long-wavelength phonon band-gap correction can be expressed as follows:
![]() | (4) |
For a 3D polar material, the T-induced energy level shifts appearing in eqn (4) can be computed as follows:39
![]() | (5) |
![]() | (6) |
Fig. 2a presents the anharmonic phonon spectrum calculated for Ag3SBr under finite-temperature conditions, accounting for long-range dipole–dipole interactions (i.e., including non-analytical corrections), which result in LO–TO splitting near the reciprocal space point Γ. Fig. 2b shows the corresponding short- and long-wavelength phonon band-gap corrections expressed as a function of temperature, which are always negative. Since in this study the ΔELg correction term has been calculated using the material's anharmonic phonon spectrum, we refer to this method as anharmonic Fröhlich theory (Methods).
![]() | ||
Fig. 2 Anharmonic phonon spectrum and thermal band-gap corrections estimated for the archetypal CAP Ag3SBr. (a) Anharmonic phonon spectrum obtained at T = 200 K neglecting (black solid lines) and considering (red dashed lines) non-analytical corrections (NAC). (b) Short- and long-wavelength phonon band-gap corrections, ΔESg and ΔELg, respectively, expressed as a function of temperature (excluding quantum nuclear effects). The short-range correction term was evaluated at several temperature points (blue circles and error bars); as a guide to the eye, the ΔESg data points were fitted to an arbitrary polynomial function (blue dashed line). Calculations were performed at the HSEsol + SOC level.35 |
In Fig. 2b, it is observed that near room temperature the ΔESg correction is dominant and significantly larger than ΔELg, approximately six times greater in the absolute value. Notably, at T = 400 K, the total band-gap correction for Ag3SBr amounts to 0.7 eV, which is of giant proportions, representing roughly 40% of the Eg value calculated at zero temperature (excluding quantum nuclear effects).
Fig. 3 shows the relative band-gap variation, referenced to the value calculated at zero temperature and expressed as a function of temperature, for the four CAP compounds Ag3SBr, Ag3SI, Ag3SeBr and Ag3SeI. In all cases, the band gap significantly decreases as the temperature increases (Table 1). The relative T-induced Eg reduction is largest for Ag3SeBr and smallest for Ag3SI. In particular, near room temperature, the band gap of Ag3SBr and Ag3SI is reduced by 39% and 29% while those of Ag3SeBr and Ag3SeI decrease by 56% and 38%, respectively (Fig. 3). As shown in Table 1, the agreement between the experimental and theoretical Eg values for Ag3SBr and Ag3SI improves as the temperature increases. In Ag3SeBr and Ag3SeI, the liquid phase is stabilized over the crystal phase at moderate temperatures (Fig. 3c and d); thus no band gaps were estimated for these two compounds under T > 400 K conditions.
![]() | ||
Fig. 3 Temperature-induced relative band-gap variation in CAP. Percentages are referenced to the band gap calculated at T = 0 K conditions (excluding quantum nuclear effects), namely, ΔEg(T) = Eg(T) − Eg(0), for (a) Ag3SBr, (b) Ag3SI, (c) Ag3SeBr, and (d) Ag3SeI. Error bars indicate numerical uncertainties and dashed lines are a guide to the eye. Shaded areas indicate regions of thermodynamic stability of the liquid phase (theory). Calculations were performed at the HSEsol + SOC level.35 |
CAP | E 0Kg [eV] | E 200Kg [eV] | ΔESg [meV] | ΔELg [meV] | E 400Kg [eV] | ΔESg [meV] | ΔELg [meV] | E 600Kg [eV] | ΔESg [meV] | ΔELg [meV] | E expg [eV] |
---|---|---|---|---|---|---|---|---|---|---|---|
Ag3SBr | 1.8 | 1.3 ± 0.1 | −440 | −42 | 1.1 ± 0.1 | −570 | −108 | 0.9 ± 0.2 | −680 | −175 | 1.0 |
Ag3SI | 1.4 | 1.1 ± 0.1 | −260 | −29 | 1.0 ± 0.1 | −290 | −75 | 0.8 ± 0.1 | −490 | −122 | 0.9 |
Ag3SeBr | 1.6 | 0.9 ± 0.1 | −630 | −42 | 0.7 ± 0.1 | −770 | −105 | Liquid | — | — | — |
Ag3SeI | 1.3 | 0.9 ± 0.1 | −370 | −37 | 0.8 ± 0.2 | −400 | −90 | Liquid | — | — | — |
Notably, our theoretical Eg results obtained at T = 400 K are fully consistent with the available experimental data obtained at room temperature. This excellent agreement near ambient conditions strongly suggests that the neglect of EPC effects is the main reason for the huge theoretical–experimental Eg discrepancies discussed in the Introduction. The T-induced relative band-gap renormalization found in CAP are of giant proportions, ranging from 20 to 60% near room temperature, setting a new record previously held by molecular crystals, which exhibited a 15 to 20% band-gap renormalization for similar temperature ranges.16
Table 1 also presents the value of the ΔESg and ΔELg correction terms estimated for each CAP at three different temperatures. In all cases, both the short- and long-wavelength phonon corrections are negative, with the former term considerably surpassing the latter in absolute value. For example, at T = 400 K, the short-range band-gap corrections are seven and four times larger than the long-range ones calculated for Ag3SeBr and Ag3SI, respectively. As the temperature is raised, the size of the two band-gap correction terms increases in absolute value, with |ΔELg| exhibiting the largest relative enhancement (e.g., approximately a 320% relative increase for Ag3SBr from 200 to 600 K).
To provide further insights into the impact of thermal effects on the electronic band structure of CAP, we also examined how band morphology evolves with temperature. Since band-gap renormalization in this study is assessed using the finite-differences approach based on AIMD simulations, it is convenient to unfold the energy bands calculated for the supercell into the reciprocal space of the primitive unit cell.44 This was done using the Easyunfold software,45 focusing on the representative CAP compound Ag3SBr (Fig. S1, ESI†).
Specifically, we analyzed five uncorrelated supercell snapshots extracted from a long AIMD trajectory (∼100 ps) at T = 200 K. For this particular case, the PBEsol exchange–correlation functional was employed due to the very high computational cost of performing hybrid functional calculations on supercells. Reassuringly, we verified that the band structures obtained using semilocal and hybrid functionals are practically equivalent in morphology (Fig. S2, ESI†).
Our results indicate that thermal effects lead to a downward shift of the conduction band minimum, which remains located at the Γ point, which is consistent with the band-gap trends observed in the corresponding electronic density of states. However, the ionic disorder induced by lattice vibrations causes a noticeable flattening of the valence band near its top. As a result, the valence band maximum becomes poorly defined, unlike in the static case (Fig. S1, ESI†). The consistency of results across the five ionically disordered configurations suggests that this small sampling is sufficient to capture the key temperature-induced changes in band morphology.
Fig. 4 shows the optical absorption spectra estimated for CAP as a function of incident light wavelength and temperature. It is found that α is significantly enhanced under increasing temperature, in some cases by as much as an order of magnitude. Similarly to the band gap, the T-induced optical absorption variations are the largest for Ag3SeBr (for which α ∼ 103–105 cm−1 at zero temperature and ∼104–106 cm−1 at 200 K) and smallest for Ag3SI (for which α ∼ 103–105 cm−1 at any temperature). It is also noted that the most significant optical absorption changes generally occur at low temperatures, that is, within the 0 ≤ T ≤ 200 K interval. These T-induced α trends align well with the remarkably large influence of the EPC on the band gap, underscoring the critical role of thermal renormalization effects on the optoelectronic properties of CAP.
![]() | ||
Fig. 4 Optical absorption coefficient (α) of CAP calculated at different temperatures as a function of photon energy. (a) Ag3SBr, (b) Ag3SI, (c) Ag3SeBr, and (d) Ag3SeI. Solid lines represent the estimated average values and statistical errors are indicated with shaded thick curves. The rainbow-colored region denotes photons with energy in the visible spectrum. Calculations were performed at the HSEsol + SOC level.35 |
Unfortunately, we cannot directly compare our theoretical α(ω) results with experimental data, as such data are not available in the literature. Notably, Caño et al.30 measured the optical absorption coefficient of CAP films scaled by their layer thickness, d, specifically, ≡ α·d. However, since the thickness of the synthesized CAP films was not determined in work,30 we cannot access the physical quantity of interest. In this regard, performing new optoelectronic experiments on CAP films across a broad range of temperatures, including the low-T regime, would be highly desirable.
As shown in Fig. 5a, the influence of each of the fifteen Γ phonon modes on the band gap of Ag3SBr was analysed by monitoring the change in Eg driven by frozen-phonon eigenmode distortions of increasing amplitude, u. The Γ phonons were classified into acoustic (A), optical polar (P) and optical nonpolar (NP), where the P phonons break the inversion symmetry of the centrosymmetric cubic Pmm phase. It was found that Eg is unresponsive to acoustic phonon distortions, as expected, while optical P phonons produce the largest band-gap variations. As the amplitude of the optical phonon distortions increases, Eg systematically decreases in both the P and NP cases.
![]() | ||
Fig. 5 Phonon-induced band-gap variation estimated for the archetypal CAP Ag3SBr. (a) Band gap as a function of the lattice distortion amplitude u for acoustic, polar optical (P) and non-polar optical (NP) Γ phonons. (b) Derivative of the band gap with respect to the phonon distortion amplitude calculated at u0 = 0.4 Å and expressed as a function of the phonon energy. The eigenmode of the optical polar Γ phonon rendering the largest band-gap derivative in absolute value is sketched: Ag, S and Br atoms are represented with grey, yellow and brown spheres, respectively. Calculations were performed at the HSEsol + SOC level.35 (c) Γ phonon-induced relative bond length distortions in the cubic Pm![]() |
Fig. 5b shows the value of the derivative of the band gap with respect to the phonon distortion amplitude, u, expressed as a function of the phonon eigenmode energy (as obtained from T-renormalised phonon calculations, Methods). We found that low-energy polar phonon modes (∼10 meV) cause the most significant band-gap reductions, followed by high-energy lattice vibrations of the same type (∼200 meV). At room temperature, phonon excitations with the lowest energy host the highest populations and, consequently, represent the most characteristic lattice vibrations in the crystal. Therefore, based on the results shown in Fig. 3 and 5, we conclude that low-energy polar phonon modes are primarily responsible for the substantial temperature-induced Eg reduction reported in this study for CAP compounds.
The eigenmode of the optical P phonon with the lowest energy is represented in Fig. 5b. As observed therein, this frozen-phonon lattice distortion reduces the distance between the central sulfur atom and one adjacent silver atom (Ag2), while increasing the other two S–Ag1 and S–Ag3 bond lengths, compared to the undistorted cubic unit cell. Fig. 5c summarizes the relative bond length variation, in absolute value, for all pairs of atoms resulting from each of the fifteen Γ phonon modes calculated for the cubic Pmm phase. As shown therein, the optical P phonons produce the largest S–Ag distance changes (up to 20%), while the optical NP phonons cause the largest Br–Ag bond length variations (up to 12%). The Br–Ag bond lengths are also appreciably impacted by the optical P phonons (5–10%). This general behaviour is reminiscent of that observed for optical polar phonons in model perovskite oxides like BaTiO3 (with atomic substitutions Ag ↔ O, S ↔ Ti and Br ↔ Ba).47,48
After identifying the phonon modes that underpin the giant T-induced band-gap reduction reported in this study for CAP, specifically low-energy optical P modes, we further analyse the induced changes in the electronic band structure. Fig. 6a shows the electronic density of states calculated for the archetypal compound Ag3SBr (equilibrium geometry). It is observed that the top of the valence band (VB) is dominated by highly hybridized silver d and chalcohalide p electronic orbitals, while the bottom of the conduction band (CB) is dominated by isotropic and more delocalized S and Ag s orbitals. The electronic band structure in Fig. 6b shows that the VB corresponds to the high-symmetry reciprocal space point M (1/2,1/2,0), while the CB to the center of the Brillouin zone, Γ (0,0,0); thus the band gap of Ag3SBr is indirect (we have checked that the same conclusion applies to the rest of CAP compounds analyzed in this study).
![]() | ||
Fig. 6 Electronic band structure properties of the archetypal CAP Ag3SBr. (a) Electronic density of states calculated for the equilibrium cubic Pm![]() ![]() |
The effects on the electronic band structure resulting from a frozen-phonon lattice distortion corresponding to the lowest-energy optical P eigenmode (u0 = 0.4 Å) are twofold (Fig. 6b). First, due to the breaking of phonon-induced inversion symmetry, the energy band degeneracy at the reciprocal space point M is lifted. However, the band gap of the system is unaffected by this energy degeneracy lifting effect since the VB remains practically invariant. Second, the CB edge experiences a significant decrease in energy and, as a consequence, the band gap of the system is reduced by approximately 30%. Therefore, we may conclude that the giant T-induced Eg reduction reported in this study for CAP is primarily caused by low-energy polar phonon modes that induce a pronounced CB energy decrease.
To better understand the electronic origins of the optical P phonon-induced CB energy lowering, we constructed a tight-binding (TB) model based on Wannier functions that accurately reproduces our DFT band structure results (Methods and Fig. S3, ESI†). Specifically, the TB model consists of s, p, and d orbitals for the five atoms in the unit cell, resulting in a total of 45 distinct Wannier orbitals. Consistently, the TB model reproduces the dominant Ag and S s character of the CB and its energy lowering under the polar lattice distortion of interest (Fig. 6c and d).
According to this TB model, the impact of the frozen-phonon distortion on the Ag and S s conduction orbitals is twofold. First, the difference in their kinetic energies, corresponding to the diagonal Hamiltonian matrix elements difference |〈Ag s|H|Ag s〉 − 〈S s|H|S s〉|, decreases (Fig. S3, ESI†). And second, the hopping s term involving the Ag2 and S atoms, represented by the off-diagonal Hamiltonian matrix element 〈Ag2 s|H|S s〉, increases (Fig. S3, ESI†). The general physical interpretation that follows from these TB results is that the polar frozen-phonon distortion enhances the hybridization of Ag2 and S s conduction orbitals, which lowers and increases the energy of the corresponding bonding (σ) and antibonding (σ*) states, respectively. Consequently, the CB, which is dominated by the Ag2-S σ interaction, is lowered. The revealed EPC mechanism, which overall produces a band-gap reduction, is schematically represented in Fig. 6e.
One may wonder whether, in addition to silver chalcohalide antiperovskites, there exist other families of materials exhibiting similarly large T-renormalization effects on the band gap and optical absorption coefficient. As discussed in previous sections, the polar nature of low-energy optical phonons appears to be essential in this regard. Consequently, a tentative set of necessary conditions for identifying potential materials that display similar T-induced effects on the optoelectronic properties may include dielectric materials exhibiting (1) centrosymmetric crystalline phases, (2) low-energy or even imaginary optical polar phonons, and (3) highly hybridized and delocalized electronic orbitals near the Fermi energy level. The availability of large DFT calculations and phonon databases may enable high-throughput material screening of such a kind.49,50
Ferroelectric oxide perovskites, exemplified by the archetypal compounds SrTiO3 (STO) and BaTiO3 (BTO), appear to satisfy the set of necessary conditions outlined above. Notably, a significant band-gap modulation has been reported for STO under biaxial strain conditions, although this phenomenon arises from different physical mechanisms than those identified in this study for CAP compounds (i.e., energy degeneracy lifting due to symmetry breaking).51 Moreover, the experimental room-temperature band gap of BTO (≈3.2 eV52) shows substantial disagreement with zero-temperature theoretical estimates obtained with hybrid functionals (≈4.0 eV53), highlighting an experiment–theory inconsistency similar to that described for Ag3SBr and Ag3SI in the Introduction. Additionally, the band gap of the multiferroic oxide perovskite BiFeO3 exhibits a remarkable temperature-dependent shrinkage, decreasing by approximately 50% within the temperature range 300 ≤ T ≤ 1200 K,54 likely influenced by the magnetic degrees of freedom.55 These findings suggest that the temperature effects and EPC mechanisms identified in this study for CAP compounds may have broader relevance, potentially extending to other well-known families of functional materials. Theoretical investigations exploring this possibility are currently underway.
The polar nature of the optical phonon modes, which cause the significant T-induced reduction in the Eg of CAP, opens up exciting technological possibilities. Similar to how an electric field can stabilize a polar phase with ferroelectric polarization over a paraelectric state at constant temperature through a phase transformation,56 it is likely that polar optical phonons in CAP can also be stimulated using external electric fields. This possibility implies that the optoelectronic properties of CAP could be effectively tuned by applying an electric field rather than altering the temperature, providing a more practical approach for the development of advanced optical devices and other technological applications. Experimental validation of this hypothesis would be highly valuable.
Finally, advances in light sources and time-resolved spectroscopy have made it possible to excite specific atomic vibrations in solids and to observe the resulting changes in their electronic and electron–phonon coupling properties.57–59 These developments also suggest the possibility of tuning the optoelectronic properties of CAP, as well as of similar materials like oxide perovskites,60–62 through specific phonon excitations using optical means such as lasers. This approach may simplify the design and manufacture of practical setups by eliminating the need for electrode deposition. Therefore, the results presented in this work are significant not only from a fundamental perspective but also for envisioning potential technological applications in which the optical and electronic properties of materials could be effectively tuned by external fields and photoexcitation.
The significant Eg reduction is attributed to strong electron–phonon coupling driven by low-energy polar phonon modes, which distorts the lattice symmetry, increases the overlap between silver and chalcogen s orbitals in the conduction band, and lowers the reference energy of the resulting bonding state. With increasing temperature, the optical absorption coefficient of CAP materials also rises, enhancing their response to visible light by nearly an order of magnitude and highlighting their potential for optoelectronic applications.
This research demonstrates that CAP compounds exhibit giant band-gap renormalization primarily due to temperature and electron–phonon coupling effects, suggesting that they could be tailored for specific applications through thermal, electric field, and/or optical control. These fundamental findings open possibilities for the design of innovative optoelectronic devices and establish a foundation for exploring similar effects in other dielectric materials with strong electron–phonon coupling.
All-electron DFT calculations were also performed with the WIEN2k package69 using the local-density approximation (LDA)70 to the exchange correlation energy along using the linearized augmented plane wave method (FP-LAPW).71,72 The technical parameters for these calculations were a 10 × 10 × 10 k-point grid and a muffin-tin radius equal to RMT = 7.0/Kmax, where Kmax represents the plane-wave cutoff. Localized energy-resolved Wannier states73 were then obtained for the tight-binding calculations74–76 considering the relevant Hilbert space in the interval −10 ≤ E ≤ 20 eV around the Fermi energy.
A normal-mode decomposition technique was employed in which the atomic velocities vjl(t) (j and l represent particle and Cartesian direction indexes) generated during fixed-temperature AIMD simulation runs were expressed as follows:
![]() | (7) |
The Fourier transform of the autocorrelation function of vqs was then calculated, yielding the following power spectrum:
![]() | (8) |
Finally, this power spectrum was approximated by a Lorentzian function of the following form:
![]() | (9) |
![]() | (10) |
These calculations were performed using the hybrid HSEsol exchange–correlation functional35 and considering spin–orbit coupling effects (HSEsol + SOC). Due to the high computational expense involved, the AIMD calculations were performed using a supercell containing 40 atoms and a k-point mesh of a single point. The total number of configurations used for the average was N = 10 for each material and temperature. These values were found to be appropriate for obtaining band-gap results accurate to within 0.1 eV (Supplementary methods, ESI†).
The high-frequency and static dielectric constants calculated for the cubic Pmm phase were unusually large in static calculations, since this phase is not vibrationally stable at T = 0 K. Thus, the dielectric constants were recomputed as functions of temperature (using the same approach as we applied to the band gap and dielectric tensor) to capture thermal effects. In view of their weak temperature dependence, and for simplicity, the dielectric constants obtained at T = 200 K were employed throughout this work. We also considered the anisotropy of the dielectric tensor using the following formula:
![]() | (11) |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5tc00863h |
This journal is © The Royal Society of Chemistry 2025 |