Open Access Article
G. Krishnamurthy Grandhia,
Noolu. Srinivasa Manikanta Viswanath
b,
Marcello Righetto
c,
Sara Domenici
d,
Mokurala Krishnaiaha,
Marco Moroni
e,
Adriana Pecoraro
f,
Ana Belén Muñoz-García
f,
Michele Pavone
g,
Lorenzo Malavasi
e,
Teresa Gatti
dh and
Paola Vivo
*a
aHybrid Solar Cells, Faculty of Engineering and Natural Sciences, Tampere University, P.O. Box 541, Tampere, FI-33014, Finland. E-mail: paola.vivo@tuni.fi
bDivision of Materials Science and Engineering, Hanyang University, 222 Wangsimni-ro, Seongdong-gu, Seoul 04763, Republic of Korea Hybrid
cDepartment of Chemical Science, University of Padova, Via Marzolo 1, I-35131 Padova, Italy
dDepartment of Applied Science and Technology, Politecnico di Torino, Corso Duca degli Abruzzi 24, 10129 Torino, Italy
eDepartment of Chemistry and INSTM, University of Pavia, Via Taramelli 12, 27100, Italy
fDepartment of Physics “Ettore Pancini”, University of Naples Federico IIComp. Univ. Monte Sant'Angelo, Naples 80126, Italy
gDepartment of Chemical Sciences, University of Naples Federico IIComp. Univ. Monte Sant'Angelo, Naples 80126, Italy
hCenter for Materials Research, Justus Liebig University, Heinrich-Buff-Ring 17, 35392 Giessen, Germany
First published on 20th October 2025
Silver–bismuth perovskite-inspired materials (Ag–Bi PIMs) encompass halide double perovskites, vacancy-ordered Cs2AgBi2I9, the (Cu)–Ag–Bi–I family, and structurally related chalcogenides and mixed-anion chalcohalides. Despite their structural diversity, these materials share key electronic features with lead halide perovskites, such as octahedral MX6 motifs and similar band edge physics, and have emerged as promising non-toxic alternatives. This review explores the structural and chemical diversity of this semiconductor family, showing how cation disorder (CD), crystal vacancies, and reduced electronic dimensionality (ED)—leading to flat bands and heavy carriers—contribute to their indirect bandgaps, high exciton binding energies, and moderate charge-carrier mobilities. Recent advances in defect passivation, CD engineering, and ED control have led to promising photovoltaic efficiencies (∼10% for AgBiS2 under 1 sun illumination and ∼8% for Cs2AgBi2I9 under indoor lighting), alongside unique functional properties, such as pronounced second-harmonic generation, broadband photocatalysis, resistive switching for neuromorphic devices, and high-sensitivity X-ray detection. Emerging insights reveal that homogeneous CD can reduce bandgaps and enhance light absorption, while controlled crystal vacancies induce local structural modifications critical for nonlinear optical responses. By systematically linking the atomic-scale structure to photophysical behaviour and device-level performance, this review traces clear design guidelines—such as enhancing the ED, minimizing deep trap states, and leveraging mixed-anion chemistry—to advance Ag–Bi PIMs from promising lead-free absorbers to versatile platforms for sustainable energy, photonics, and intelligent electronics. We propose a roadmap outlining a three-stage development model focused on material innovation and device optimization for system-level integration, positioning Ag–Bi PIMs as environmentally friendly semiconductors with broad potential in next-generation optoelectronics.
A particularly effective strategy to harness the potential of Bi3+ involves pairing it with monovalent silver cations (Ag+), forming a diverse family of Ag–Bi PIMs. This heterovalent Ag+–Bi3+ combination maintains charge neutrality, as the combination of Ag+ and Bi3+ approximates that of 2 Pb2+ ions, enabling stable halide-based crystal structures such as double perovskites (elpasolite-type, e.g., Cs2AgBiBr6), layered variants (e.g., Cu2AgBiI6, also known as CABI), vacancy-ordered perovskite derivatives (e.g., Cs2AgBi2I9), and rudorffite-type phases.5–8 Ag–Bi PIMs are further categorized into three distinct chemical families based on their anion chemistry—halides, chalcogenides, and chalcohalides. Halide-based compounds such as Cs2AgBiBr6 possess wide indirect bandgaps (∼2.0–2.2 eV) alongside remarkable air stability.7 In contrast, chalcogenides (e.g., AgBiS2) feature narrower direct bandgaps (∼1.0–1.2 eV), exceptionally strong optical absorption, and intrinsic chemical stability, making them highly attractive for solar cell applications.9 Chalcohalides, combining halide and chalcogenide anions (e.g., AgBiSCl2 or sulfur-alloyed halides), bridge these extremes, enabling tuneable optoelectronic properties and optimized bandgaps (∼1.6–2.0 eV).10 By systematically studying these distinct chemistries, researchers are unveiling how compositional variation influences electronic dimensionality (ED), defect chemistry, and cation ordering—key factors that govern carrier transport and device performance.
Despite their promising attributes, unlocking the full potential of Ag–Bi PIMs requires overcoming fundamental scientific challenges that limit their optoelectronic functionality. Most notably, the combination of filled d10 (Ag+) and s2 (Bi3+) orbitals frequently leads to weak orbital connectivity, creating electronically isolated sub-lattices and low ED. This structural-electronic scenario manifests through indirect or quasi-indirect bandgaps, high exciton binding energies, and low charge-carrier mobilities. Additionally, intrinsic cation disorder (CD)—arising from the similar ionic radii of Ag+ and Bi3+—and associated antisite defects and vacancies generate deep-level traps and extensive band-edge tailing, severely limiting carrier lifetimes and device efficiency. Recent developments, however, demonstrate that strategic compositional tuning (e.g., Cu or Sb substitutions), refined film processing, and advanced defect management are effective approaches for significantly improving the electronic connectivity and mitigating these limitations.
Intriguingly, the same structural and electronic properties that challenge traditional photovoltaic performance can be advantageous in emerging technology domains. The wide bandgaps (∼1.8–2.2 eV) and defect-tolerant nature of Ag–Bi halide PIMs enable encouraging indoor photovoltaic (IPV) performance, exemplified by Cs2AgBi2I9, reaching PCEs up to ∼8% under indoor LED illumination.8 Similarly, AgBiS2 solar cells now demonstrate outdoor efficiencies exceeding 10%, driven by optimized CD engineering and extraordinary absorption coefficients.11 Beyond photovoltaics, these materials exhibit fascinating multifunctionality. Disorder-driven local symmetry breaking in Cu–(Ag)–Bi–I crystals yields unusually strong nonlinear optical (NLO) phenomena, such as significant second-harmonic generation (SHG), valuable for frequency conversion and integrated photonics.12 Ionic conductivity and resistive-switching behaviours in CABI-based compounds further underpin promising memristor and neuromorphic device applications.13 Additionally, heavy atomic constituents and charge-carrier self-trapping processes position Ag–Bi halides as attractive candidates for radiation detection and scintillator technologies, significantly broadening their technological scope.14,15
Given the rapid diversification and accelerating research momentum on Ag–Bi PIMs since 2024—including breakthroughs such as record indoor efficiencies approaching 8% (Cs2AgBi2I9), outdoor efficiencies exceeding 10% (AgBiS2), exceptional NLO responses in Cu–(Ag)–Bi–I systems, and the memristors based on CABI—an up-to-date and comprehensive review is now essential. Although recent reviews, notably by Zhu et al. (2024) on Cu–(Ag)–Bi–I semiconductors16 and Grandhi et al. (2023) on pnictogen-based halide PIMs,17 have broadened the scope beyond photovoltaics, a deeper integration of state-of-the-art insights into ED, CD, and defect chemistry across the full halide–chalcohalide–chalcogenide spectrum remains necessary. This minireview precisely bridges this gap. By systematically correlating fundamental structure–property relationships with representative device demonstrations (e.g., photovoltaics, photocatalysis, memristors, and nonlinear optics), we extract practical material design principles to accelerate the development of Ag–Bi PIMs toward impactful applications in sustainable energy conversion, multifunctional optoelectronics, and emerging electronic and photonic technologies.
m) with rock-salt ordering of Ag+ and Bi3+ (Fig. 1a), which lowers the symmetry and shapes its electronic structure.21 Its 2D analogue, (BA)2AgBiBr6 (monoclinic P21/c), retains cation ordering within perovskite slabs separated by bulky BA+ cations (Fig. 1b).22 This layering breaks 3D connectivity, introduces octahedral tilting, and enlarges the indirect bandgap due to increased VBM-CBM mismatch. The result is a high exciton binding energy and strong charge-carrier localization. Although moisture-stable, its wide bandgap of ∼2.3 to 2.5 eV and poor interlayer transport hinder its application in photovoltaics. Another important class is the vacancy-ordered double perovskite Cs3Bi2I9, which can crystallize in either a structurally zero-dimensional (0D) dimer phase (P63/mmc) or a 2D layered phase (C2/c), depending on the synthesis conditions. Here, half of the B-sites are vacant, leading to isolated or layered [Bi2I9]3− clusters separated by Cs+ ions (Fig. 1c). The lack of 3D octahedral connectivity severely restricts charge-carrier transport, resulting in a large bandgap and low mobility of 2.4–2.5 eV and ∼10−2 to 10−1 cm2 V−1 s−1, respectively.23–25 However, partial substitution of Cs+ with Ag+, forming Cs2AgBi2I9-type compositions, significantly improves the electronic properties. Ag+ incorporation introduces rock-salt-type ordering between Ag+ and Bi3+ on B-sites, reinstating 3D connectivity and enhancing structural dimensionality (Fig. 1d). Moreover, Ag+ alloying reduces the formation of deep trap states associated with isolated [Bi2I9]3− units by improving lattice continuity, making Ag–Bi hybrid phases superior optoelectronic materials compared to their vacancy-ordered parent.
![]() | ||
| Fig. 1 Crystal structures of various Ag–Bi PIMs and related compounds. (a) 3D double perovskite structure of Cs2AgBiBr6, featuring corner-sharing [AgBr6] and [BiBr6] octahedra. (b) 2D layered perovskite (BA)2AgBiBr6, where BA = butylammonium, showing alternating inorganic slabs separated by organic spacer (BA) layers. Reproduced with permission from ref. 22. Copyright 2025, American Chemical Society. (c) Layered vacancy-ordered perovskite Cs3Bi2I9. (d) Layered Cs2AgBi2I9 structure with face-sharing octahedra and a non-centrosymmetric arrangement. (e) Crystal structure of CABI with a mixed Cu/Ag/Bi framework and distorted octahedra. (f) Quasi-1D chain-like structure of AgBiI4, highlighting the linear connectivity. (g) Layered Ag2BiI5 composed of interconnected polyhedra forming 2D planes. The panels (e–g) are reproduced with permission from ref. 12. Copyright 2025. The Author(s). Published by Elsevier Inc. (h) Ordered superstructure of AgBiS2, consisting of edge-sharing [AgS6] and [BiS6] units. (i) Disordered AgBiSCl2 lattice with mixed S/Cl anions and CD, leading to structural complexity. | ||
The (Cu)–Ag–Bi–I family (e.g., CABI, AgBiI4, and Ag2BiI5) shows complex structures shaped by CD and vacancies, which significantly impact their optoelectronic properties. CABI (R
m) features layered, more precisely classified as a rudorffite-type phase characterized by a CdCl2-like octahedral motif, defect-perovskite-like structures with Cu+, Ag+, and Bi3+ sharing B-sites, leading to significant CD and vacancies (Fig. 1e).12 These induce potential fluctuations, broaden band edges, and create trap states that limit mobility, although the material retains a wide bandgap (∼2.0 eV). AgBiI4 (Fd
m) has a cubic rock-salt structure with half of the B-sites being vacant and Ag/Bi sharing sites (Fig. 1f),12 but antisite defects and vacancy disorder introduce deep traps, reducing the carrier lifetime and open-circuit voltage (VOC). Although Ag2BiI5 and NaVO2 share the R
m space group, their octahedral motifs differ: NaVO2 features edge-sharing VO6 octahedra, while Ag2BiI5 adopts a rudorffite-type structure with CdCl2-like edge-sharing [BiI6] and [AgI6] units. This highlights that identical symmetry can mask differences in local connectivity—Ag2BiI5 has higher Ag content and more disorder (Fig. 1g),12 promoting non-radiative recombination and poor diffusion lengths despite a narrower bandgap. Among these, CABI stands out for its favourable bandgap, good absorption, and potential for defect engineering, making it more promising for photovoltaic applications despite some structural disorder.
Chalcogenide PIMs such as AgBiS2 and AgBiSxSe1−x are another important Ag–Bi material family. The structural and electronic properties of AgBiS2 are strongly influenced by its intrinsic CD and the formation of sulfur vacancies (VS).26 AgBiS2 typically crystallizes in a rock-salt-derived structure with cubic symmetry, most commonly associated with the Fm
m space group (No. 225), akin to the ideal NaCl-type lattice. While the fully disordered schapbachite phase is stable at room temperature due to entropy-driven CD, the ordered matildite phase with checkerboard A-site occupation is also stable, stabilized by favourable electrostatic interactions between Ag+ and Bi3+. In AgBiS2, Ag+ and Bi3+ statistically share the 4a site, while S2− occupies the 4b site, forming a close-packed lattice (Fig. 1h). The similar ionic radii and moderate charge difference favor CD over long-range ordering. This disorder induces local charge imbalances, particularly in Bi-rich regions, which in turn stabilize sulfur vacancies. These vacancies distort the valence band, introducing deep trap states that facilitate non-radiative recombination, shorten carrier lifetimes, and ultimately limit the photovoltaic efficiency.27,28 In contrast to disordered AgBiS2, AgBiSCl2 crystallizes in an orthorhombic Cmcm structure with distinct, ordered Ag+ and Bi3+ sites. Its layered anion sublattice, with S2− and Cl− ions, favours Bi–S and Ag–Cl coordination (Fig. 1i), promoting both cation and anion ordering. This structural order suppresses the formation of deep-level defects, particularly sulfur vacancies, reducing non-radiative recombination. The anisotropic bonding and orthorhombic distortions enhance lattice stability, exciton binding, and defect tolerance.10 Additionally, the mixed-anion framework allows tuneable bandgaps via S/Cl electronegativity, making AgBiSCl2 a more stable and efficient alternative to AgBiS2 for optoelectronic applications.
From a structural perspective, AgBiS2 stands out as a highly promising photovoltaic candidate for sunlight harvesting, owing to its optimal direct bandgap (∼1.5 eV), strong absorption, and relatively defect-tolerant chemistry—provided that CD is well-controlled and sulfur vacancy formation is suppressed. CABI is a strong secondary option, though its layered, disordered structure introduces mid-gap states that limit performance without defect management. AgBiSCl2 benefits from structural ordering and improved stability but still needs optimization to enhance the carrier mobility. Cs2AgBi2I9 offers better connectivity than Bi-only phases but is hindered by partial vacancy disorder. Other materials like Cs2AgBiBr6, (BA)2AgBiBr6, AgBiI4, and Ag2BiI5 are structurally disadvantaged due to large indirect bandgaps, poor transport, and deep trap states, making them less suitable for light-harvesting applications.
To provide a concise overview, the main Ag–Bi compounds discussed in this review are summarized in Table 1. The table organizes structural families, representative compositions, bandgaps, electronic-dimensionality notes, defect trends, mobility ranges, and benchmark device metrics. This comparative picture complements the structural discussion above and serves as a reference point for the optoelectronic properties, defect chemistry, and device performance that will be discussed in later sections.
| Ag–Bi family | Representative composition(s) | Structure (space group) | Bandgap (eV) & character | Effective masses | Mobility range (cm2 V−1 s−1) | ED (qualitative) | Dominant native defects | Best device metric (context) |
|---|---|---|---|---|---|---|---|---|
| Halide, double perovskite | Cs2AgBiBr6 | Elpasolite (Fm m) |
≈2.2 (indirect); first strong direct transition at higher E | 0.33/0.35 | ∼3 → 1.3 (deloc → loc) | Low ED: VBM (Ag-4d/Br-p) vs. CBM (Bi-6p) separation | VBr; AgBi/BiAg antisites | 6.37% (1-sun, H-treated); ∼7% (1000 lux WLED, IPV) |
| Vacancy-ordered iodide | Cs2AgBi2I9 | Layered, face-sharing motifs | ≈1.8 (quasi-direct) | 0.6/2.1 | —(Mobility-lifetime product (µτ) of 3.4 × 10−3 cm2 V−1) | Higher ED (quasi-2D) than Cs3Bi2I9 via Ag–I–Bi connectivity | VI (iodine vacancies) | ∼3% (1-sun); ∼8% (1000 lux, IPV) |
| Cu–Ag–Bi–I | CABI (Cu2AgBiI6) | Layered CdCl2-like (R m) |
≈1.9–2.0 (direct) | 0.6/1.0 | ∼2.1 → 1.0 (deloc → loc) | Cu 3d raises the VBM → improved ED | VCu/VAg (shallow), VBi (deeper) | ∼2.2–3% (1-sun); 5.66% (1000 lux, IPV) |
| Iodide (Ag-rich rudorffite) | Ag3BiI6 | Rudorffite-related (edge-sharing octahedra) | ≈1.8–2.0 (direct/quasi-direct) | 0.4/0.4 | ∼0.41 → 0.02 (deloc → loc) | Moderate ED; disorder-broadened edges | Antisites + vacancies | ∼4.3% (1-sun); 5.6% with S-alloying |
| Chalcohalide (mixed anion) | AgBiSCl2 | Orthorhombic (Cmcm) | >2.0 (direct) | —/— | — | Intermediate ED; ordered Ag/Cl–Bi/S layers | VS suppressed vs. AgBiS2; VCl; mixed-anion (SCl/ClS) | Projected >10% (theory) under 1-sun; no PEC demonstrated yet |
| Chalcogenide | AgBiS2 | Rock-salt-derived (Fm m) (CD-tunable) |
≈1.5 (direct when homogeneously disordered) | 0.35/0.722 (segregated case) | ∼0.43 → 0.11 (segregated) to 2.7 → 2.2 (disordered) | ED increases with homogeneous CD; strong absorption | VS, antisites; CD-controlled | 10.20% (0.06 cm2), 9.53% (1.00 cm2), and ∼10% NC-based |
However, Bi3+ ions differ significantly from Pb2+ ions in having a smaller ionic radius, a higher oxidation state, and deeper-lying 6s2 orbitals, all of which lead to notable differences in the lattice structure. These structural differences consequently alter the electronic properties and defect chemistry of Ag–Bi-based materials.
Cations with higher oxidation states (e.g., Bi3+ compared to Pb2+) introduce more transition levels within the bandgap, thereby enhancing the likelihood of deep trap formation.39 Additionally, unlike LHPs where the metal 6s2 orbitals directly contribute to the VBM, the Bi3+ 6s2 lone-pair orbitals in Ag–Bi PIMs lie approximately 1 eV below the valence-band maximum (VBM) and are poorly coupled to the halide p states.4,39 This energetic mismatch leads to a less dispersive valence band and higher effective masses for charge carriers, significantly increasing the tendency to form deep trap states. Specifically, as the energetic separation between the cation and anion orbitals grows, dangling-bond orbitals become strongly localized, pushing defect states deeper into the bandgap. Furthermore, shorter spatial distances between neighbouring dangling orbitals—such as those between Bi atoms around an oxygen vacancy compared to those around an iodine vacancy—strengthen the orbital overlap and deepen the resulting trap states (illustrated by shallow iodine vacancies in BiOI versus deep oxygen vacancies in Bi–O materials).39 Consequently, careful structural and compositional engineering is essential to achieve Ag–Bi PIMs with high defect tolerance.
Ag–Bi double perovskites, such as Cs2AgBiBr6, typically exhibit an indirect bandgap. This arises from the differing orbital contributions at the VBM and conduction band minimum (CBM): the VBM is primarily composed of anti-bonding Ag 4d and halide p orbitals, while the CBM is dominated by Bi 6p orbitals (with negligible Ag 5s character).40 These states occur at different k-points in the Brillouin zone, resulting in reduced band-edge absorption. Additionally, Jahn–Teller-like distortions associated with Ag+ broaden the density of states and lower the charge-carrier mobility. Antisite defects and Br vacancies further introduce non-radiative recombination pathways, limiting photovoltaic efficiency despite the material's excellent stability. In contrast, Cs2AgBi2I9 exhibits a different electronic structure due to its vacancy-ordered lattice. Here, Ag+ (4d10) states contribute more significantly to the VBM, raising the VBM energy and narrowing the bandgap. Moreover, Ag 5s and I 6p orbitals contribute to the CBM in Cs2AgBi2I9, forming an intermediate band that further reduces the bandgap.20 This greater orbital overlap leads to increased band dispersion and lower effective masses, resulting in enhanced carrier mobility compared to Cs2AgBiBr6.8,41
Recent studies have focused on mixed-metal iodides such as CABI, which exhibit a direct bandgap of ∼2 eV and strong light absorption. In CABI, both theoretical and experimental studies show that the VBM is primarily formed by Cu 3d states, mixed I 5p orbitals—similar to AgBiI4—while the CBM is dominated by Bi 6p and I 5p states.6,42 This orbital alignment reduces the mismatch at the band edges and increases the optical absorption efficiency of the material. In contrast, in Cu-free Ag–Bi–I compositions, the VBM is dominated by I-5p states.
Bandgap tunability in Ag–Bi PIMs can be achieved through halide mixing. Ag–Bi iodide PIMs typically exhibit smaller bandgaps (∼1.8 eV), which enhance light absorption and charge transport. In contrast, bromide-rich compositions display wider bandgaps and higher exciton binding energies. Recent density functional theory (DFT) studies have shown that mixing of Br with CABI enables fine-tuning of the bandgap from around 1.9 eV (iodine-rich) to about 2.3 eV (bromide-rich), without introducing problematic mid-gap states.43 Similarly, modifying the A-site cations (such as using mixtures of Cs+, Rb+, and formamidinium) also influences the electronic properties. DFT calculations on similar Sb-halide PIMs indicate that mixed A-site cations improve structural stability and reduce energetic disorder at grain boundaries, indirectly enhancing charge transport.44 Similar strategies are expected to be effective in Ag–Bi PIMs as well.
Ag–Bi PIMs frequently show notable CD, which will be discussed in detail in Section 3.4. This disorder sometimes can positively influence the electronic structure of the PIMs. For example, CD can broaden electronic bands and create localized energy states at the band edges. Recent theoretical simulations of CABI showed partial ordering of Cu and Ag at room temperature, and this partial ordering directly affects the electronic structure near the band edges.43 This quaternary PIM also frequently contains vacancies, particularly at Cu and Ag sites. These vacancies typically form shallow acceptor levels, which may facilitate p-type conductivity. Importantly, such defects generally do not introduce deep trap states. DFT calculations have confirmed this defect-tolerant behaviour, showing a low density of recombination centres.
Ag–Bi halide PIMs can adopt various structural dimensionalities (atom connectivity in spatial coordinates), including 3D double perovskites, 2D layered structures, and even 0D molecular-like configurations. Dimensionality plays a critical role in determining their electronic properties. Low-dimensional structures typically exhibit high exciton binding energies and reduced carrier mobilities, but they can be advantageous for applications such as nonlinear optics and light emission (e.g., scintillators). For instance, CABI possesses a near-3D crystal structure and improved electronic connectivity compared to ternary Ag–Bi–I systems, enabling relatively isotropic hole transport. Theoretical calculations have predicted hole mobilities exceeding 10−2 cm2 V−1 s−1 in ideal single crystals.6 However, structural complexities arising from mixed-metal sites still introduce anisotropy in the electronic bands, making control over crystallinity during synthesis essential for optimizing performance.
Further tuning of electronic properties can be achieved through strain engineering or element alloying. For example, partially substituting Bi3+ with Sb3+ lowers the CBM, improving energy level alignment with electron transport layers and enhancing device performance. Studies on antimony–bismuth and Ag–Bi PIM alloys have shown that Sb3+ incorporation reduces structural disorder and sharpens absorption edges, resulting in improved photovoltaic efficiency.45 Additionally, applying strain, via chemical doping or external pressure, can beneficially modify the electronic structure without compromising defect tolerance.
By replacing halides with chalcogenides, the electronic structure of Ag–Bi PIMs is markedly altered in terms of band dispersion, dimensionality, and defect tolerance. Halide-based Ag–Bi PIMs (e.g., Cs2AgBiBr6) generally feature indirect bandgaps with flat, dispersal-limited band edges due to the spatial separation of Ag-centred and Bi-centred orbitals. In contrast, Ag–Bi chalcogenides such as AgBiS2 and AgBiSe2 typically exhibit narrower, more direct bandgaps (on the order of 1.0–1.3 eV) and exceptionally strong optical absorption. The greater covalency of Bi–S/Se bonds can promote stronger orbital overlap at the band edges, but the ED in these chalcogenides typically remains low when Ag and Bi cations are ordered or phase-segregated, yielding quasi-0D electronic structures analogous to the halide double perovskites. This often manifests as large exciton binding energies, pronounced carrier localization (self-trapping), and indirect-like optical behaviour in disordered lattices. Notably, AgBiS2 crystallizes in a rock-salt-derived structure where Ag+ and Bi3+ randomly occupy the cation sublattice, and if this CD is homogeneous, it can increasingly lead to more dispersive bands. Indeed, recent experiments show that engineering a more homogeneously disordered AgBiS2 (i.e., avoiding Ag/Bi clustering) results in bandgap narrowing and enhanced absorption.9 Alloyed chalcogenides AgBiSxSe1−x allow bandgap tuning across 1.0–1.3 eV (absorption in the 700–1200 nm wavelength regime).46
Chalcohalide Ag–Bi PIMs, which combine halide and chalcogenide anions (e.g., AgBiSCl2), tend to bridge the above behaviours. Due to the mixed anion chemistry, these materials have intermediate bandgaps (often in the 1.1–1.7 eV range) that are direct in nature and moderately dispersive. For instance, AgBiSCl2 adopts an ordered orthorhombic structure with alternating Bi–S and Ag–Cl layers, yielding a direct bandgap at the Brillouin zone Y-point and greater in-plane band dispersion than the fully halide analogues. Orbital-projected density of states and charge density analyses reveal that the valence band of AgBiSCl2 is predominantly composed of sulfur and chlorine p orbitals, along with significant contributions from silver d orbitals. In contrast, the conduction band is chiefly governed by bismuth 6p orbital character. The incorporation of the more electronegative halide (Cl−) alongside S2− also allows tuning of band-edge positions via anion electronegativity differences.10
Importantly, AgBiSCl2 is reported to be a more structurally stable and defect-tolerant semiconductor than pure AgBiS2. In AgBiS2, sulfur vacancies (VS) are the dominant native defects (owing to the relatively weak Ag–S and Bi–S bonds) and can act as deep traps under S-poor conditions.26 By contrast, the presence of a halide in AgBiSCl2 may suppress or compensate such anion-vacancy formation, leading to fewer mid-gap states and consequently a clear band-edge photoluminescence (PL),10 whereas AgBiS2 typically exhibits only broadband trap-state dominant emission.47 Overall, replacing halides with chalcogenides tends to narrow the bandgap and increase optical absorption but at the cost of more localized carriers and potential deep defect states, while chalcohalide materials offer a compromise between improved band connectivity and stability. These comparisons underscore how anion substitution (X = S, Se, and/or halogen) profoundly influences the electronic structure of Ag–Bi PIMs, providing handles to tailor band dispersion, dimensionality, and defect chemistry for better optoelectronic performance.
In summary, the electronic structure of Ag–Bi PIMs is strongly influenced by factors such as CD, vacancies, halide composition, and ED. Their defect-tolerant nature, combined with the ability to tune electronic properties, makes these materials promising not only for solar cells but also for broader applications in optoelectronics and emerging technologies. Continued computational and experimental efforts will be essential to fully realize their potential.
![]() | ||
| Fig. 2 (a) Charge density isosurfaces of the valence band maximum (VBM) and conduction band minimum (CBM), at the Brillouin zone centre. These isosurfaces illustrate the different orbital connectivity in Cs2AgBiBr6 and CsPbBr3. While in the former, the charge density is localised on Bi atoms in the conduction band and on Ag atoms in the valence band, the charge density is more evenly spread in the latter. Ag atoms are shown in light gray, Bi in purple, Pb in dark gray, and Br in brown. Reproduced with permission from ref. 57, licensed under a Creative Commons Attribution (CC-BY) license. (b) and (c). HSE06 + SOC band structures of Cs2AgBiBr6 and CsPbBr3, respectively. Reproduced with permission from ref. 57, licensed under a Creative Commons Attribution (CC-BY) license. | ||
Within this framework, the lower ED of Ag–Bi PIMs is attributed to the combination of two metal cations with distinct electronic configurations: (Bi(III) with configuration [Xe]4f145d106s2 and Ag(I) with configuration [Kr]4d105s0).56
As described in the previous section, the conduction band edge of Ag–Bi PIMs is generally dominated by Bi 6p orbitals, whereas the valence band edge is mainly composed of halide p orbitals and Ag 4d orbitals (the Bi 6s2 lone-pair levels lie deeper in energy and do not set the VBM).58,59 This orbital arrangement contributes to a lower ED, which has been thoroughly investigated and demonstrated for Cs2AgBiBr6. In this material, the weak participation of Ag 5s orbitals results in poor orbital connectivity at the conduction band, leaving the Bi 6p orbital states relatively isolated.54 The weak Bi–Ag overlap results in a flat, poorly connected conduction band, contributing to a large electron effective mass and an indirect optical gap. This spatial isolation between conduction and valence band states—localized on [AgX6] and [BiX6] octahedra, respectively—has been observed across a range of Ag–Bi PIMs (e.g., Cs2AgBiBr6, AgBiS2, and (AgI)x(BiI3)y).9,52,54,60,61 It has been linked to a quasi-0D electronic structure close to the band edges. Consequently, the low ED in these materials has been associated with unfavourable electronic structures (Fig. 2b and c), yielding poor optoelectronic properties such as indirect bandgaps, reduced absorption coefficients (<103 cm−1 at the band edge), and low charge-carrier mobilities (<2 cm2 V−1 s−1).62,63 Recently, Biega et al. further demonstrated that this spatial separation of electron and hole wavefunctions of different cations in Cs2AgBiBr6 yields localised excited states (e.g., excitons) and has been referred to as a “chemical localisation” effect.58,64 Importantly, this localization weakens quantum confinement effects65,66—for instance, when transitioning from 3D Cs2AgBiBr6 to 2D layered PEA4AgBiBr8—thereby limiting the effectiveness of dimensionality reduction strategies in Ag–Bi PIMs.
To address ED-related limitations, compositional engineering and doping strategies are gaining increasing attention for Ag–Bi PIMs. However, alloying or substituting Ag(I) and Bi(III) with alternative isoelectronic cations, such as Au(I), Na(I) and Sb(III), has not yielded significant improvements in the ED,59,67–69 reinforcing the idea that the distinct electronic configurations of these ions are a primary cause of low ED in PIMs. In contrast, Hossain et al. recently demonstrated that incorporating Ag(I) cations into the 0D Cs3Bi2I9 (A-site compositional mixing) yields a novel 2D semiconductor, Cs2AgBi2I9, with higher ED and improved optoelectronic properties.8,41 Similarly, Cu(I) introduction has emerged as a promising route to enhanced ED. While Cu(I) is unsuitable as an A-site cation, several stable Cu-based NaVO2-type rudorffites (Cu4x(AgBi)1−xI4) have been reported by Herz, Snaith and coworkers.6,55,70,71 Crucially, the contribution of Cu(I) 3d orbitals to the upper valence band increases the electronic connectivity and reduces the electronic isolation of Ag- and Bi-octahedra.6,69 The resulting direct bandgap and low exciton binding energies further underscore the importance of high ED in achieving photovoltaic-relevant properties.6,71 Among these, CABI stands out as a particularly promising composition, offering a direct bandgap and a high absorption coefficient of 1.0 × 105 cm−1 just above the band edge.6,72,73
| Material | Space group | Bandgap character | Effective mass | e–h sum mobility (cm2 V−1 s−1) |
|---|---|---|---|---|
| CsPbBr3 | Fm m |
Direct82 | 0.22 (electrons);83 0.24 (holes)83 | 13 (ref. 84) |
| Cs2AgBiBr6 | Fm m |
Indirect58,76 | 0.33 (electrons);59 0.35 (holes)59 | 3 (delocalized);85 1.3 (localized)85 |
| Cu2AgBiI6 (CABI) | R m |
Direct6,7 | 0.6 (electrons);6 1.0 (holes)6 | 2.1 (delocalized);71 1 (localized)71 |
| Cs2AgBi2I9 | P63/mmc | Indirect (quasidirect)41 | 0.6 (electrons);20 2.1 (holes)20 | 36 (ref. 20) |
| CuAgBiI5 | R m |
Direct55 | — | 1.7 (ref. 70) |
| Cu0.4AgBiI4.4 | R m |
Direct70 | — | 0.6 (ref. 70) |
| Cu6AgBiI10 | R m |
Direct70 | — | 7.3 (ref. 70) |
| AgBi2I7 | Indirect86 | — | 0.49 (delocalized);60 0.01 (localized)60 | |
| AgBiI4 | R m |
Indirect87 | 0.4 (delocalized);60 0.02 (localized)60 | |
| Ag2BiI5 | R m |
Direct88 | — | 0.49 (delocalized);60 0.03 (localized)60 |
| Ag3BiI6 | R m |
Direct61,89 | 0.41 (delocalized);60 0.02 (localized)60 | |
| AgBiS2 (segregated) | P m1 |
Indirect90,91 | 0.43 (delocalized);92 0.11 (localized)92 | |
| AgBiS2 (disordered) | Fm m |
Direct90 | — | 2.7 (delocalized);92 2.2 (localized)92 |
The low ED of Ag–Bi implies poorer orbital connectivity, which yields less dispersive electronic bands and increased effective masses.59,76 However, although higher effective masses were reported for Ag–Bi PIMs with respect to LHPs (see Table 2), these values do not entirely account for the low reported mobilities. Increased charge-carrier scattering rates and a crucial role played by strong electron-phonon scattering have therefore been proposed.77–79 In 2021, Wu et al. proposed that self-trapping in Cs2AgBiBr6 is a major intrinsic limitation to the photovoltaic performance of these materials.80 Different from early reports indicating predominantly Fröhlich coupling in Ag–Bi halide double perovskites, Wu et al. have proposed that a strong coupling with acoustic phonons dominates the electron–phonon interaction via the acoustic deformation potential interaction (Fig. 3a).80,81 At the same time, Wright et al. have reported the presence of an ultrafast (∼ps) decay of photoconductivity in Cs2AgBiBr6 thin films. As demonstrated by fluence-dependent and temperature-resolved photoconductivity measurements, such decay is also associated with a change in the charge-carrier transport regime from bandlike to hopping transport (Fig. 3b).85 Such a swift localization of photogenerated charge-carriers in Cs2AgBiBr6 not only yields a reduction in charge-carrier mobility—from ∼3 to ∼1 cm2 V−1 s−1—but also imposes a less-efficient hopping transport regime. Crucially, while an energetic barrier to self-trapping is expected for higher-dimensional 3D and 2D semiconductors,93,94 the lower dimensionality of the Cs2AgBiBr6 electronic structure (i.e., its lower ED) has been proposed as a cause of barrierless trapping.71,85
![]() | ||
| Fig. 3 (a) Temperature-dependent photoluminescence (PL) broadening measured for Cs2AgBiBr6 single crystals. The dashed line indicates the conventional phonon coupling model (weak coupling regime), while the solid line indicates the fit with the Toyozawa strong-coupling model. Reproduced from ref. 80, licensed under a Creative Commons Attribution (CC-BY-NC) license. (b) Temperature-dependent mobility for delocalised (green triangles and line) and localised states (blue triangles and line) in Cs2AgBiBr6 thin films. Reproduced from ref. 85 under a Creative Commons Attribution (CC-BY 4.0) license. | ||
Recently, charge-carrier localization processes have been reported for a wide variety of Ag–Bi PIMs, including CABI, (Cu4x(AgBi)1−xI4), Cs2AgBi2I9, AgBiS2, (4FPEA)4AgBiI8, and (AgI)x(BiI3)y,8,60,66,70,71,92,95 all showing a similar ultrafast loss in charge-carrier mobility. Furthermore, such localization has been reported in analogous systems containing Au(I), Na(I) and Sb(III)59,67,68—showing a similarly low ED—thus further indicating the role played by low ED arising from the combination of d10 and s2 metal cations.
![]() | ||
| Fig. 4 (a) Illustration of different types of cation ordering in AgBiS2. The partial occupancy of the Ag(I) and Bi(III) cation sites is represented as fractional filling of the circles at each site. These schematics of AgBiS2 structures depict ideal segregated (top), ordered (middle), and homogeneously disordered cases (bottom). Reproduced from ref. 92, licensed under a Creative Commons Attribution (CC-BY 4.0) license. (b) Illustration of cation vacancies in AgBiI4. Adopting the CdCl2-type structure, AgBiI4 shows a layered structure with alternating fully occupied and entirely vacant layers. Reproduced from ref. 87, licensed under a Creative Commons Attribution (CC-BY) license. (c) Charge-carrier mobility in AgBiS2 nanocrystals thin films, as a function of annealing temperature. With increasing temperatures, the increasingly homogeneous CD yields improved charge-carrier mobilities for both delocalised (squares) and localised (triangles) states. Reproduced with permission from ref. 92, licensed under a Creative Commons Attribution (CC-BY 4.0) license. (d) Ratio between the mobility of localised and delocalised charge carriers in (AgI)x(BiI3)y thin films as a function of the fractional vacancy concentration in the rudorffite structure. Reproduced from ref. 60 under the terms of Creative Commons Attribution (CC-BY 4.0) license. | ||
Beyond the disordered distribution of Ag(I) and Bi(III) cations, neutral structural vacancies have also been proposed to contribute to the disordered cation distribution in Ag–Bi PIMs.6,61,70,87,102 Turkevych et al. have described the structure of the Ag–Bi rudorffite series (AgBi2I7, AgBiI4, Ag2BiI5, and Ag3BiI6) as an edge-sharing octahedral network, with [MX6] octahedra exhibiting fractional occupancy modelled as AgaBibΔc (Δ represents a vacancy), where a, b, and c are the fractional occupation ratios.7 Similarly, Sansom et al. and Buizza et al. reported structural vacancies for AgBiI4,87 CABI,6 and within the Cu4x(AgBi)1−xI4 series (see Fig. 4b).70 Importantly, Ag and Bi vacancies can break local symmetries, create local strains, and modulate lattice softness in Ag–Bi PIMs.12,60
For instance, Annurakshita et al. recently demonstrated that vacancy-induced lattice symmetry breaking can be engineered to tune the NLO properties of Ag–Bi PIMs.12
CD and cation vacancies also significantly influence charge-carrier transport in Ag–Bi PIMs. Antisite defects (AgBi) have been shown to introduce deep trap states that act as non-radiative charge-carrier recombination centres.101 Maiti et al. further suggested that increased electrostatic repulsion between neighbouring octahedra and lattice distortion occur in disordered Cs2AgBiBr6.100 Cation-segregated configurations (see Fig. 4a) have been linked to the formation of localized states in both AgBiS2 and NaBiS2.9,68 Importantly, Righetto et al. demonstrated that charge-carrier localization is mitigated in more homogeneously disordered AgBiS2 (Fig. 4c),92 highlighting how disorder-induced ED-tuning can help overcome localization effects in Ag–Bi PIMs. Furthermore, Lal et al. recently showed a strong correlation between the extent of charge-carrier localization and the abundance of cation vacancies in the lattice (Fig. 4d),60 suggesting that vacancy-induced lattice softness may hinder charge-carrier transport. While the study of CD and cation vacancies in Ag–Bi PIMs is still emerging, engineering cation distribution—e.g., through compositional fine-tuning,12 thermal annealing,9,98 and coordination chemistry100—has become a highly promising strategy for optimizing performance in these materials.
Here, we analyse the defect chemistry of Ag–Bi PIMs and correlate it with device performance metrics where available. We formulate defect-formation reactions using the Kröger–Vink notation for representative Ag–Bi materials, including Cs2AgBiBr6, Cs2AgBi2Br9, CABI, AgBiI4, Ag2BiI5, AgBiS2, and AgBiSCl2. This approach helps elucidate the impact of native point defects on carrier dynamics, recombination losses, and optoelectronic performance, ultimately guiding material optimization strategies.
In Cs2AgBiBr6, bromine vacancies (VBr), common due to their low formation energies, act as shallow trap states via
and contribute to radiative recombination. Cesium vacancies (VCs) act as acceptors and induce slight p-type behaviour through CsCs → VCs− + Cs (g or sol). Silver vacancies, VAg, similarly generate holes: → AgAg → VAg + Ag (g or sol). Antisite defects like AgBi (silver on a bismuth site) and BiAg (bismuth on a silver site) introduce deep trap states that enhance nonradiative recombination and lattice distortion formed by AgAg + BiBi → AgBi + BiAg. Bromine interstitials (Bri), though less frequent due to high formation energy, form under Br-rich conditions and follow
, increasing the ionic conductivity. Device performance is primarily limited by deep traps from AgBi and BiAg, as shown by DFT calculations (Fig. 5a).110 Controlling growth conditions is key; Xiu et al. found that Ag-rich and Br-poor conditions optimize n-type conductivity and suppress secondary phases.111
![]() | ||
| Fig. 5 Energy level diagrams and defect energetics in Ag–Bi halides and chalcohalides. (a) Calculated charge transition levels of native point and antisite defects in Cs2AgBiBr6, including VCs, VAg, VBi, BiCs, and AgBi, with their respective charge states across the bandgap. Reproduced with permission from ref. 103. Copyright 2016, WILEY-VCH. (b) Thermodynamic transition levels (Eq/q′) of intrinsic point defects in Cs3Bi2I9, referenced to the valence band maximum (VBM), showing the depth and recombination potential of each defect. Reproduced with permission from ref. 104. Copyright 2017, American Chemical Society. (c) Distribution of key defect levels in Cs3Bi2I9, with mid-gap states such as VI and BiAg expected to act as nonradiative recombination centers. Reproduced with permission from ref. 105 under the terms of the CC-BY license. Copyright 2024. The Author(s), published by Wiley-VCH GmbH. (d) Defect transition levels in Cu2AgBiI6 (CABI), AgBiI4, and Ag2BiI5, where shallow acceptors like VCu and VAg support p-type conductivity, while deeper defects such as VBi, CuBi, and AgBi introduce trap states that degrade carrier lifetimes. (e) Comparative defect chemistry in AgBiS2 and AgBiSCl2, where VS acts as a dominant deep donor in AgBiS2, while VAg, VBi, and AgBi serve as acceptor-type defects. In AgBiSCl2, halide incorporation introduces additional defects such as VCl and mixed-anion substitutions (SCl and ClS), modifying the defect landscape and potentially mitigating deep-level recombination pathways. Panels (d and e) were fabricated primarily based on the reported defect energy level data presented in ref. 70, 106, 107, 92 and 107–109, respectively. | ||
In Cs3Bi2I9, iodine vacancies—one of the most common and energetically favourable defects—form via the reaction
, creating positively charged traps that promote nonradiative recombination. Cesium vacancies act as acceptors and lead to p-type conductivity, following the reaction CsCs → VCs− + Cs(g). Bismuth vacancies, though less favorable energetically, can also act as acceptors via BiBi → VBi3− + Bi(g). Antisite defects such as BiCs and CsBi result in local lattice distortions and deep-level trap states, represented by the reaction BiBi + CsCs → BiCs2+ + CsBi2−. Alloying Ag+ at the Cs+ site offers a promising defect mitigation strategy (Fig. 5b).112 First, Ag+ forms stronger Ag–I bonds than Cs–I, thereby stabilizing iodide ions and suppressing iodine vacancy formation; the reaction
becomes less favorable in the presence of Ag. Second, Ag+ substitution reduces Cs vacancy formation due to site occupancy, as described by AgCs + CsCs → AgCs0 + VCs− + Cs(g), making VCs generation energetically less favourable (Fig. 5c).105 Lastly, Ag+ ions tend to avoid substituting for Bi3+ sites, which limits the formation of antisite defects. This selective site preference helps preserve the crystal structure and minimizes the creation of mid-gap electronic states, thereby enhancing the optoelectronic performance of the material.
The defect chemistry of CABI, AgBiI4, and Ag2BiI5 is governed by native point defects, with cationic vacancies at the Cu+ and Ag+ sites playing a dominant role in determining their optoelectronic properties. In CABI, copper vacancies (VCu) are the most energetically favourable acceptor-type defects, contributing to p-type conductivity via the reaction CuCu → VCu− + Cu(g), and these shallow acceptors aid hole transport with minimal nonradiative recombination. Likewise, silver vacancies (VAg), which are common across CABI, AgBiI4, and Ag2BiI5, form through the reaction AgAg → VAg− + Ag(g) and also contribute to p-type conductivity, though at high concentrations they may introduce instability or recombination losses (Fig. 5d).16,70,113 Iodine vacancies (VI), which act as deep-level donor defects, are formed via
, and are detrimental as they promote nonradiative recombination and degrade photovoltaic performance. In Ag-rich phases such as AgBiI4 and Ag2BiI5, antisite defects (e.g., BiAg and AgBi) can arise due to charge and size mismatches, creating deep traps and inducing local lattice distortions. While CABI largely tolerates CD, precise stoichiometry and defect passivation—via mild alloying or halide-rich growth—are key to reducing deep defects and improving carrier lifetimes, stability, and performance.
The defect chemistry of AgBiS2 and AgBiSCl2 is primarily governed by intrinsic point defects, with sulfur vacancies (VS) being the most prevalent due to the relatively weak Ag–S and Bi–S bonds, particularly under sulfur-poor or high-temperature conditions. These sulfur vacancies act as donor-type defects, forming via the reaction
, where
denotes a doubly positively charged vacancy and the released electrons contribute to n-type conductivity. However, these defects introduce deep-level trap states that enhance nonradiative recombination and degrade carrier transport. In AgBiS2, silver vacancies (VAg) commonly form under Ag-deficient conditions through the reaction AgAg → VAg + Ag(g), contributing to p-type behaviour, while bismuth vacancies (VBi3−) may form via the reaction BiBi → VBi3− + Bi(g), though with higher formation energy due to Bi's stronger bonding. Additionally, antisite defects such as AgBi2+ and BiAg2− can emerge through AgAg + BiBi → AgBi2+ + BiAg2−, inducing local distortions and deep traps (Fig. 5e).92,114–116 In AgBiSCl2, the presence of highly electronegative Cl− modifies the defect landscape by partially stabilizing the lattice and suppressing VS formation, although sulfur vacancies can still form under S-deficient conditions by the same reaction. Cl vacancies (VCl) may also appear via the reaction
, acting as shallow donors. Furthermore, mixed-anion defects such as SCl− and ClS+, arising from site exchange reactions (SS + ClCl → SCl− + ClS+), can significantly perturb the electronic structure and affect charge transport. AgBiS2 displays strong n-type behaviour with sulfur vacancies, while AgBiSCl2 has more complex defect chemistry with VS, VCl, and anion substitutions. Optimizing the growth conditions (sulfur-rich for AgBiS2 and Cl-rich for AgBiSCl2) and controlling the Ag
:
Bi ratio are crucial for reducing recombination losses and improving performance.
In summary, AgBiS2, AgBiSCl2, and CABI possess lattices with greater defect tolerance than many other Ag–Bi PIMs, resulting in fewer intrinsic deep traps and, consequently, enhanced optoelectronic device performance. In contrast, AgBiI4 and Cs2AgBiBr6 suffer from abundant deep trap states and Ag/Bi antisite defects that significantly impair the device efficiency. Whether these intrinsic differences ultimately limit—or can be engineered to enhance—optoelectronic performance depends on how charge carriers interact with both the defect landscape and the soft, highly polarizable lattice. The following section moves from static defect energetics to their dynamic consequences, briefly examining how carrier trapping, small-polaron formation and exciton binding energies shape the observable photophysics of the Ag–Bi family.
:
Bi ratio control suppresses antisite defects like AgBi and BiAg, leading to enhancement in both stability and optoelectronic quality.117,118 Alloying approaches also offer a promising means to neutralize deep traps and stabilize lattice environments; incorporation of Ag+ into Cs3Bi2I9 has been demonstrated to reinforce Ag–I interactions and inhibit iodine vacancy generation, while mild alloying with trace amounts of heterovalent or isovalent cations induces local strain fields that passivate shallow states and prolong carrier lifetimes.8,20 In addition to intrinsic defect control, extrinsic defect passivation strategies, including surface functionalization or post-synthesis annealing, provide further means of control, with sulfur-rich annealing of AgBiS2 as a prototypical system where defect healing directly enhances carrier transport.119,120 At the processing stage, sophisticated synthesis techniques—like growth under controlled inert atmospheres—can mitigate oxidation-induced defects in materials such as AgBiSCl2, and thorough growth temperature optimization can reduce deep-level state formation in sulfide materials. Real-time feedback is increasingly being integrated into synthesis through advanced characterization techniques, as in situ monitoring allows for dynamic modification of growth parameters, and defect mapping through PL imaging reveals local defect concentrations to inform subsequent optimization pathways. Concurrent work in compositional engineering reveals that the addition of stabilizing anions, such as chloride in AgBiSCl2, modifies the defect landscape to favour higher electronic conductivity, and mixed-anion approaches (e.g., S2−/Cl− co-substitution) lead to beneficial changes in the electronic structure that may suppress non-radiative processes. Significantly, the defect chemistry engineering towards shallow traps is an effective approach to enhance radiative recombination, and the higher halide content in Cs2AgBiBr6 leads to shallow trap states, which significantly improve photoluminescence and charge transport. Collectively, these advances provide an integrated toolbox for defect control in Ag–Bi PIMs, spanning growth optimization, alloying, passivation, processing control, and compositional engineering that can be viewed as a unified approach to performance enhancement. However, although progress has been made, challenges remain in relating defect formation energetics to full-device stability under operational stress, underscoring the need for systematic, long-term investigations that integrate computational defect modelling with in situ experimental verification. With this knowledge, the community can progress towards design rules for the rational design of defect-tolerant Ag–Bi PIMs, which will enable their use as a robust, lead-free option for application in next-generation solar and optoelectronic technologies.
The low-dimensional crystal or electronic structure yield pronounced excitonic effects, evident as distinct exciton resonances in absorption spectra. As a result, very high exciton binding energies have been consistently reported for Cs2AgBiBr6 (>200 meV) as well as AgBiI4 and Ag2BiI5 (>150 meV)—significantly exceeding those of LHPs (∼25 meV for MAPbI3)—severely limiting efficient charge separation at room temperature.121,122 However, for CABI, a remarkably low exciton binding energy of approximately 25 meV was initially reported (estimated by the Elliott fitting of absorption spectra),6 suggesting efficient free-carrier generation. However, a more recent study reported a considerably higher exciton binding energy (>100 meV), which was reduced to ∼100 meV through partial Sb3+ incorporation.45 Another recently discovered material, Cs2AgBi2I9, has been found to exhibit a relatively low exciton binding energy (∼40 meV),8 consistent with the enhanced ED discussed previously. These findings highlight how targeted compositional tuning is a crucial strategy to mitigate exciton-related limitations in Ag–Bi halide PIMs and thereby enhance their optoelectronic performance. AgBiS2 is expected to exhibit a high exciton binding energy, especially in its quantum dot form, due to enhanced quantum confinement effects.
In fully 3D-ordered double perovskites like Cs2AgBiBr6, the fundamental bandgap is indirect (∼2.2 eV), with the first allowed optical transition being at higher energy (∼3.0 eV). Consequently, this shows a weak absorption onset, with strong absorption emerging only at higher-energy direct transitions.123 In contrast, solution-processed direct bandgap CABI films achieve ∼1 × 105 cm−1 near the band edge, due to allowed direct transitions.6
Nonetheless, low ED usually broadens PL spectra and suppresses PL quantum yields, since photo-excitations tend to localize. In Cs2AgBiBr6, a broad red PL band peaking around 600–700 nm (1.9–2.0 eV) is observed at room temperature (Fig. 6a). This emission is strongly Stokes-shifted from the absorption edge and was initially attributed to an indirect exciton transition or sub-gap defect emission, but now known to originate from self-trapped excitons (STEs) and small polarons via advanced calculations and spectroscopic studies.85,123
![]() | ||
| Fig. 6 Representative absorption and photoluminescence (PL) characteristics of Ag–Bi PIMs. (a) Deconvoluted PL spectrum showing multiple recombination channels in a representative Ag–Bi PIM, Cs2AgBiBr6. Reproduced with permission from ref. 127. Copyright 2016, WILEY-VCH. (b) Absorption coefficient and normalized PL intensity of Cs3Bi2I9, highlighting low absorption near the band edge and broad PL emission. (c) Optical properties of Cs2AgBi2I9, including absorption onset and broad PL emission, indicating sub-bandgap recombination. Panels (b and c) are reproduced with permission from ref. 105 under the terms of the CC-BY license. Copyright 2024. The Author(s), published by Wiley-VCH GmbH. (d) Time-resolved PL decay curves of single crystal and powder samples of Cs2AgBiBr6. Reprinted (adapted) with permission from ref. 125. Copyright 2016, American Chemical Society. (e) PL comparison between the AgBiS2 core and AgBiS2/ZnS core–shell nanocrystals, revealing strong PL enhancement upon surface passivation with ZnS, indicating the suppression of surface trap states. Reproduced from ref. 47 under the terms of CC-BY 4.0. Copyright 2024, American Chemical Society. (f) Steady-state PL spectrum of AgBiSCl2 NCs (inset: schematic representing the origin of the trap-assisted PL component of the PL spectrum of the NCs). Reproduced from ref. 10 under the terms of CC-BY 4.0. Copyright 2023. The Author(s), published by American Chemical Society. | ||
Cs3Bi2I9 exhibits strong carrier localization and weak PL (Fig. 6b) due to deep defects and an indirect gap (∼2.0–2.2 eV); incorporating Ag+ to form Cs2AgBi2I9 slightly improves the dimensionality and reduces the bandgap (∼1.8 eV) but still displays emission characteristics due to polaron-induced effects (Fig. 6c).8 The appearance of a shoulder peak in the absorption spectrum after Ag incorporation causes the reduced bandgap in Cs2AgBi2I9 compared to Cs3Bi2I9.
(Cu)–Ag–Bi–I systems, such as CABI and (AgI)x(BiI3)y, exhibit ultrafast self-localization of charge carriers, characterized by picosecond-scale conductivity losses and broad, weak, redshifted PL,60,71,95 alongside additional non-radiative recombination pathways linked to Cu/Ag vacancies and antisite defects. Such self-trapping arises from strong coupling between charge carriers and lattice distortions (high deformation potentials) associated with electronically soft, low-dimensional lattices. Photoexcited carriers in many Ag–Bi PIMs typically localize on ultrafast (sub-picosecond) timescales. Time-resolved studies have shown carrier self-localization rates around 1–2 ps, along with similarly barrierless, sub-picosecond trapping in AgBiS2 nanocrystals.60,71,85,124 However, homogeneous CD in AgBiS2 notably reduces the extent of charge-carrier localization.124 Conversely, Sb incorporation into CABI leads to broader STE emission, indicative of enhanced carrier self-trapping due to structural modifications. This reduces trap-mediated recombination, as evidenced by transient absorption studies, ultimately improving the photovoltaic performance.45
Since the ultrafast self-trapping driven PL tends to be broad and of low intensity in which case radiative recombination occurs from a distorted, lower-energy state, the STE PL in Ag–Bi PIM lifetimes often becomes multi-exponential, with a dominant short component from non-radiative traps and a long-lived tail (up to hundreds of nanoseconds or even microseconds) from radiative STE recombination. Single crystals of Cs2AgBiBr6, for instance, show a long-lived PL component of approximately 690 ns (Fig. 6d), explained by slow radiative decay of STEs despite a high defect density.125 However, a combination of ultrafast self-trapping and trapping at the surface and bulk of the carriers in polycrystalline limited the diffusion lengths to just 30 nm and mobilities to <1 cm2 V−1 s−1.126
High defect concentrations and deep trap states (common in these materials) contribute to the lack of band-edge PL and weak STE PL intensities, funnelling carriers into non-radiative paths. Intrinsic defect chemistry in Ag–Bi PIMs is complex, as these compounds readily form vacancies and antisite defects that introduce deep recombination centres. Without deliberate passivation, non-radiative recombination via such defects is a major loss pathway, severely reducing PL quantum yields and carrier transport properties. For AgBiS2 nanocrystals, coating with a ZnS shell induces near-infrared PL emission at 764 nm (Fig. 6e) by passivating surface defects (Ag/S vacancies and antisites), enhancing carrier confinement through a type-I band alignment, and suppressing small-polaron forssmation and ultrafast trapping.
Interestingly, AgBiSCl2 NCs exhibit PL originating from both band edge excitons and midgap states—marking the first report of band edge emission in chalcohalide nanomaterials—where the partial suppression of self-trapping and a more favourable electronic structure enable radiative recombination, although non-radiative decay via midgap states introduced by Cl vacancies, disorder, and antisites still dominates (Fig. 6f).
In summary, the photophysical features of Ag–Bi PIMs are strongly influenced by their low ED, intrinsic defects, and pronounced electron–phonon coupling. These characteristics result in substantial exciton binding energies, rapid carrier self-localization into small polarons, limited mobility, and weak, broad PL. Strategies such as compositional tuning (e.g., Sb incorporation into CABI), CD engineering (as in the case of AgBiS2), and defect passivation (e.g., ZnS coating on AgBiS2 nanocrystals) effectively mitigate these limitations, highlighting pathways toward improved optoelectronic performance in this intriguing family of materials.
For Cs2AgBiBr6,129–134 solution films benefit from antisolvent-assisted nucleation (e.g., IPA and methyl acetate) and hot-casting/gas-quenching, which promote dense, ultra-smooth coverage; the first inverted devices (2.23%, VOC ≈ 1.01 V) were realized using IPA drip + high-T anneal, and subsequent work shows that post-annealing ≥250–285 °C is often required for phase-pure films from solution. Optimizing preheat and anneal setpoints (and the timing of the antisolvent drop) consistently improves the grain size and crystallinity. Dry/semidry methods (pulsed layer deposition and chemical vapor deposition) add scalability by tuning the deposition pressure/temperature for highly crystalline layers. Even so, the current record 6.37% stems from hydrogenation of spin-cast films, which narrows Eg from ≈2.18 eV to ≈1.64 eV and passivates halide-vacancy traps—evidence that defect energetics, not morphology alone, set the efficiency ceiling.
For Cu–Ag–Bi–I—especially CABI—the thin-film picture diverges by route. Co-evaporation is highly sensitive to the anneal temperature (≈110–150 °C) and iodine fugacity: modest increases drive CuI-rich impurity domains, visible as a sharpened ≈415 nm absorption shoulder and a strong ≈730 nm PL band.33 These domains shorten electron-transport lengths and can produce deceptively bright PL while degrading the device current (front/back EQE measurements reveal electron-extraction bottlenecks in planar cells prepared under such conditions).33 In contrast, solution processing has yielded repeatable improvements via additive and solvent engineering: hydroiodic acid (HI) additions increase the surface coverage and reduce interfacial recombination (early devices ≈1.3% AM 1.5G),135 while hypophosphorous acid (H3PO2) generates in situ Ag nanoparticles and strengthens Bi–(H3PO2/H3PO3) coordination, giving smooth films (RMS ∼24 → 14 nm), lowered ideality factor (1.64 → 1.15) and improved Vbi (0.81 → 0.86 V). These CABI devices delivered improved photovoltaic performance.136 Cation engineering (partial Sb3+ alloying on the Bi site) also reduced trap densities and delivered a 1-sun PCE of 1.82%,45 while halide engineering tuned lattice/defect chemistry in mixed-iodide/bromide analogues.43 In short, co-evaporation demands a narrow thermal window, whereas solution routes profit from controlled chemistry for uniformity and reduced non-radiative recombination loss.
For AgBiS2 (chalcogenide),9,11,137 the decisive control is densification. Vapor-assisted solution and chemical vapor deposition/co-evaporation now produce sub-micron-grain, pinhole-free films with 10.20% (0.06 cm2) and 9.53% (1.00 cm2) PCE and strong durability (e.g., ≥94% retained after ∼3000 h ambient, ≈87% after 1000 h at 85 °C). Parallel progress in nanocrystal inks shows that post-deposition in situ passivation (e.g., ligand exchange/halogenated agents) can make ultrathin, trap-lean films with FF ≈ 72% and >10% PCE. Solution-crystallized “thick” films deliver JSC > 31 mA cm−2 under AM 1.5G. Thermal co-evaporation has also yielded phase-pure AgBiS2, underscoring that both vacuum and solution tracks can meet the film-quality bar when voiding/cracking are suppressed by proper rheology and drying-kinetics control.
While AgBiS2 has been synthesized in nanocrystal form (limited synthesis efforts so far),10 other Ag–Bi–I materials (e.g., AgBiI4, Ag2BiI5, and AgSb2I7) are gaining process-specific insights. Rapid thermal/microwave iodization of stacked metal-halide precursors produces AgSb2I7 thin films,138 while data-driven optimization (self-driving labs) has screened >1700 synthesis conditions for AgBiI4 in minutes,139 yielding pin-hole-free films with larger grains than historical baselines. These show that precursor sequence, iodization kinetics, and anneal ramps are equally central to reproducible morphology in emerging Ag–Bi–I systems.
The most extensively studied Ag–Bi halide PIM for solar cells is Cs2AgBiBr6 double perovskite. Its wide bandgap (∼2.2 eV) is sub-optimal for single-junction photovoltaics, initially limiting PCEs to ≈1–2%. Optimised antisolvent casting, ammonium-halide surface passivation and molecular-dye sensitisation have pushed outdoor efficiencies to ≈4.5%.5 Under 1000 lux white-LED illumination, the same architecture delivers ≈7.2% indoor PCE.142 Deep bromine-vacancy traps drive non-radiative recombination. Low-temperature hydrogen plasma treatment mitigates these traps, narrows the bandgap to 1.64 eV and raises the record outdoor PCE to 6.3% with VOC ≈ 0.92 V—currently the highest for any Ag–Bi halide PIM.134 With a bandgap of 1.64 eV, the detailed-balance (Shockley–Queisser) limit predicts a VOC,SQ ≈ 1.35 V at 300 K;143 the champion 6.3% Cs2AgBiBr6 solar cell, which delivers VOC ≈ 0.92 V (Fig. 7a); the resulting voltage deficit (ΔV ≈ 0.43 V) still signals significant non-radiative loss pathways. Most other Ag–Bi PIM solar cells show even larger ΔV values, underlining the need for more aggressive passivation strategies.11,17
![]() | ||
| Fig. 7 (a) Current density–voltage (J–V) curves of Cs2AgBiBr6 photovoltaic cells under 1-sun illumination, comparing samples with varying hydrogenation times (0, 600, and 1200 s). Reproduced from ref. 134 under the terms of the CC-BY license. Copyright 2022. The Author(s), published by Springer Nature. (b) Simulated J–V curves of a Cu2AgBiI6/c-Si tandem solar cell, calculated using a transfer matrix optical model coupled with detailed balance. The schematic structure of the modelled solar cell is shown alongside. Reproduced from ref. 6 under the terms of the CC-BY 4.0 license. Copyright 2021. The Author(s), published by American Chemical Society. (c) Calculated radiative-limited indoor maximum efficiencies (i-SLME) for selected PIMs, alongside MAPbI3 and a-Si, under standard white LED illumination, demonstrating their potential for indoor photovoltaic applications. Data for Cs2AgBi2I9 were estimated under similar assumptions and included as an additional reference point. Reproduced from ref. 140, Copyright 2024, with permission from Elsevier. (d) Schematic illustration of the vapor-assisted solution deposition process used to form AgBiS2 absorber layers, enabling solar cells with efficiencies exceeding 10%. Reproduced from ref. 11 with permission from Springer Nature, 2025. (e) Top panel: Power conversion efficiency (PCE) of Au|PTAA|Ag3BiI6−2xSx|m-TiO2|c-TiO2|FTO solar cells as a function of sulfide modification level (x). Bottom panel: illustrates the evolution of electronic energy levels as a function of composition (x). For reference, the conduction band edge energy of TiO2 and the highest occupied molecular orbital (HOMO) level of PTAA are also indicated. Valence band edge positions were determined experimentally using photoelectron spectroscopy in air (PESA). Reproduced from ref. 141. Copyright 2021, published by Wiley-VCH. | ||
In Ag–Bi iodide absorbers, solar cell performance correlates strongly with composition. Bismuth-rich phases (with many inherent cation vacancies) show lower PCEs, whereas silver-rich phases (fewer vacancies) generally perform better. The most Ag-rich known compound—Ag3BiI6—achieved ∼4.3% PCE in 2018.7 Nevertheless, its absorption tail (indicating an Urbach energy ≈90 meV) reveals band-edge disorder linked to CD, strong electron–phonon interaction, and high density of traps.89
The quaternary iodide CABI (∼2.0 eV bandgap) initially yielded only <1% PCE in thin-film solar cells.135 Through improved crystallization, compositional tuning, and interface engineering, recent devices have reached ∼2–3% efficiency.136,144,145 Transient absorption and photothermal deflection spectroscopy reveal broad sub-gap absorption to 1.25 eV, confirming deep traps.6 These defects severely limit the VOC and carrier diffusion lengths. While theory suggests a CABI top cell could reach VOC ≈ 1.92 V and ∼30.2% efficiency (Fig. 7b) in tandem with silicon6 and >30% PCE under indoor illumination (1000 lux),140 realizing those values will require suppressing cation-site disorder, as indicated by the high Urbach energy (≈50 meV).145
Cation-alloying effects are composition-dependent. In Cs2AgBiBr6, substituting Sb3+ for Bi3+ narrows the bandgap slightly (to ∼2.08 eV at 90% Sb) but induces substantial disorder and deep defect states. The substitution caused quenching PL and added a 0.7 eV sub-gap feature, contributing to 0.7 V VOC loss and 120 meV Urbach energy, indicating severe band-tail disorder.146 By contrast, partially alloying Sb into CABI slightly reduces its Urbach energy (from 49.6 to 47.6 meV) and boosts the champion PCE from 1.31% to 1.82% (∼40% relative improvement),145 showing that judicious alloying can improve defect tolerance. Bulk KSCN passivation has likewise lowered Urbach energy from 120 to 100 meV in AgBiI4, signalling improved lattice order.
An intriguing Ag–Bi–I compound, Cs2AgBi2I9, formed by partially substituting Ag into Cs3Bi2I9, displays extended Ag–I–Bi networks that enhance ED. State-of-the-art Cs2AgBi2I9 devices deliver ≈3% outdoor PCE and ≈8% indoor PCE at 1000 lux—record value for Ag–Bi iodides.8,17 Although its wide bandgap of ≈1.8 eV would allow >40% indoor PCE, the present low VOC values of the devices (∼0.7 V) attest to the residual CD and defect-assisted recombination.8 Similar to Cs2AgBi2I9, many wide-bandgap Ag–Bi halide PIMs possess potential indoor PCEs in the 30–50% range (Fig. 7c). Following the emergence of Cs2AgBi2I9, computational studies suggest partial A-site substitution as a potential strategy to address CD in Ag–Bi iodides.147 For example, incorporating Cs+ into the CABI lattice or partially replacing Ag with Cu in the Cs3Bi2I9 structure could help stabilize the lattice structure and further narrow the bandgap. Such targeted alloying approaches may provide avenues for optimizing the Cu–Ag–Bi–I family's future optoelectronic performance.
When moving to chalcogenides, absorption is pushed towards the near-infrared. Owing to cation-disorder engineering, AgBiS2 exhibits an extraordinarily high absorption coefficient of α > 2 × 105 cm−1 at 500 nm.9 Colloidal quantum dot devices of only 30 nm thickness collect ≈27 mA cm−2 and reach 9.2% PCE (8.85% certified).9 Vapour-assisted crystallisation (Fig. 7d) recently produced ≥10% PCE from polycrystalline films—the highest Ag–Bi PIM efficiency to date.11 The Se analogue AgBiSe2 has achieved ≈2.6% PCE in quantum dot form,46 and continuous-alloy AgBiSxSe1−x enables bandgap tuning across 1.0–1.3 eV.46 Collectively, these emerging chalcogenides broaden the horizons for lead-free photovoltaics, pairing high optical performance with robust stability inherent to Ag–Bi materials.
Introducing chalcogenides (S2− and Se2−) into Ag–Bi halides further tunes the band structure, enhancing absorption spectra and narrowing the bandgap. A clear example is the sulfur-alloyed double perovskite Cs2AgBiBr6−2xSx, where adding just ∼3% sulfur significantly improved visible-light harvesting, yielding a PCE around 2% (1.3% for the pristine halide PIM).148 Computational and experimental work indicates that heavy chalcogen substitution can induce local lattice distortions and disorder, increasing Urbach tails.149 Thus, precise synthetic control is essential to obtain uniform chalcohalide phases. Even partial incorporation of sulfur can greatly enhance the performance: a recent mixed Ag–Bi–I–S thin film (Eg ∼1.8 eV) achieved 5.6% PCE (Fig. 7e)—the highest among Ag–Bi chalcohalides.141 Furthermore, recent computational studies predict that AgBiSCl2 (direct Eg > 2 eV) could surpass 10% PCE under sunlight. In fact, its wide bandgap is more suitable for indoor light harvesting.10 Overall, chalcogen-halide alloying emerges as a versatile route to strengthen absorption, suppress deep traps, and bridge the performance gap between pure halide and pure chalcogenide Ag–Bi absorbers.
In summary, Ag–Bi PIMs have progressed from proof-of-concept photovoltaic absorbers to double-digit-efficiency devices within a decade, yet their full potential remains untapped. Continued advances in defect passivation, compositional tuning, and interface engineering150 are poised to minimize the remaining voltage and current losses, positioning these lead-free systems as viable contenders for the next-generation of air-stable and environmentally benign photovoltaics.
Recent research on Ag–Bi PIMs has explored various compositions, including halides (e.g., AgBiI4 and Ag3BiI6) and chalcogenides (e.g., AgBiS2 and AgBiSSe). These materials exhibit tuneable bandgaps, making them suitable for visible-light-driven PCs and PECs. Halide-based compounds demonstrate strong light absorption and efficient charge transport,153–155 while sulfides offer improved stability and catalytic activity in aqueous environments.156 Hybrid systems, such as Ag–Bi chalcogenides combined with co-catalysts or heterostructures, are also gaining attention for enhanced performance in hydrogen evolution and CO2 reduction reactions. Zhang and co-workers utilized AgBiS2 as a catalyst for light-driven CO2 reduction, using zeolitic imidazolate frameworks as the co-catalyst.157 In this work, particular attention has been paid to engineering the growth mechanism of the nanocrystals, by inducing control of the rate of crystal growth and yielding a highly ordered structure by stabilizing the (111) facets. This was achieved by inducing atomic defects through partial Br− ion doping during the nucleation phase and hence slowing down the crystal growth, which occurred through anion exchange with S2− (Fig. 8a). This process, coupled with the capping effect of poly(vinyl pyrrolidone), yielded hollow octahedral particles with good light harvesting abilities and electron transfer capability, and hence is a very promising photoactive material with high potential as a support for zeolitic imidazolate frameworks (ZIF) co-catalysts. Moreover, AgBiS2 has demonstrated remarkable resistance to prolonged water exposure,160 making it a strong candidate for solar-to-hydrogen conversion via water splitting. Choi and co-workers were the first to report a PEC water-splitting system with AgBiS2 photoanodes, which featured broad light absorption and photocurrent generation primarily in the visible and near-infrared range.161 Their study also highlighted the significant impact of ligand selection on the PEC performance, emphasizing the importance of surface chemistry in optimizing the device efficiency.
![]() | ||
| Fig. 8 (a) Schematic illustration showing the growth mechanism of AgBiS2 nanocrystals through anion exchange. Reproduced from ref. 157 with permission from the Royal Society of Chemistry, Copyright 2024. (b) Methylene blue photocatalytic degradation by Fe-doped Cs2AgBiBr6 at different doping ratios and process mechanisms. Reproduced from ref. 158 with permission from Wiley-VCH, copyright 2023. (c) Mechanism of photocatalytic CO2 reduction by Rb-doped Cs2AgBiBr6 and effect of surface atom substitution. Reproduced from ref. 159 with permission from Wiley-VCH, copyright 2024. | ||
For AgBiS2, the effect of CD has been extensively studied in terms of bandgap engineering,90 showing that the bandgap can be significantly reduced transitioning from the hexagonal semiconducting phase to the cubic metallic phase, enabling potential modulations of the photo(electro) chemical ability of the material. However, despite being frequently mentioned, the specific influence of CD and intrinsic defects in Ag–Bi chalcogenides on PC and PEC systems remains largely unexplored, highlighting a clear need for further investigation.
Doping has emerged as a powerful strategy for enhancing the photocatalytic and photoelectrocatalytic properties of Ag–Bi PIMs by tuning their electronic structure, improving charge carrier dynamics, and increasing stability. Introducing dopants can modulate the bandgap, optimizing light absorption and extending activity into the visible spectrum.162 Additionally, doping can enhance charge separation by creating defect states or internal electric fields that facilitate carrier migration, thereby reducing recombination losses. Structural modifications induced by dopants can also improve crystallinity and introduce strain effects that benefit catalytic performance. Moreover, stability—a key challenge in PIMs—can be enhanced by doping strategies that mitigate ion migration, suppress phase degradation, and reinforce the perovskite lattice. Recent studies have explored the potential of doped Cs2AgBiBr6 double perovskites for photocatalytic applications, focusing on band structure modulation and charge carrier dynamics. Hutter and co-workers investigated Fe-doped Cs2AgBiBr6, demonstrating that Fe incorporation effectively tunes the conduction band edge, enhancing charge separation and extending light absorption into the visible range.158 Through controlled alloying, Fe doping introduces mid-gap states that facilitate electron transfer, ultimately improving the photocatalytic efficiency (Fig. 8b). Fe-doped Cs2AgBiBr6 also exhibits greater structural stability under operational conditions, as the incorporation of Fe strengthens the perovskite lattice and reduces halide volatilization. Their work provides further insights into the defect chemistry of Fe-doped Cs2AgBiBr6, showing that Fe inclusion alters the charge compensation mechanisms and affects the carrier mobility, which is crucial for optimizing photocatalytic performance. Similarly, Chen et al. examined Rb-doped Cs2AgBiBr6 hierarchical microspheres for CO2 reduction, showing that Rb-doping enhances crystallinity, carrier mobility, and structural stability.159 The improved morphology increases active sites for CO2 adsorption, leading to higher photocatalytic efficiency (Fig. 8c). Both studies address key challenges in Cs2AgBiBr6-based photocatalysis, including limited absorption and charge recombination, by leveraging dopant-induced electronic and structural modifications. Additionally, the dopants contribute to improved stability, mitigating common degradation pathways such as ion migration. These findings underscore the potential of doped Cs2AgBiBr6 as a lead-free alternative for sustainable photocatalysis. Future research should focus on optimizing dopant concentrations, exploring synergies with co-catalysts and assessing long-term durability under operational conditions.
Ramachandran et al. reported a novel synthesis of lead-free Ag2BiI5 thin films via iodization of Ag/BiI3 layers, demonstrating excellent visible-NIR light absorption (bandgap, 1.2–1.6 eV).163 The films exhibited remarkable photocatalytic activity, degrading 96% of methylene blue (MB) dye under visible light in 120 minutes, attributed to their high crystallinity and low micro-strain. Silver(I) can also be partially replaced by copper(I), to yield mixed species. Liu et al. reported the first colloidal synthesis of water-resistant Cu1.4Ag0.6BiI5 nanocrystals (NCs), achieving a record 2.94% applied bias photon-to-current efficiency (ABPE) for lead-free photoelectrodes in water splitting.164 The layered structure and hydrophobic ligands of the NCs confer exceptional stability in aqueous media, while their bandgap (2.19 eV) aligns well with redox potentials for oxygen evolution. A photocurrent density of 4.62 mA cm−2 at 1.23 V vs. RHE was achieved without protective layers, surpassing previous Pb-free benchmarks.
Integrating Ag–Bi PIMs with co-catalysts or functional species is a convenient approach not only to enhance the photocatalytic and photoelectrochemical efficiency but also to improve the stability.162 Furthermore, heterostructure engineering with other semiconductors improves charge separation and passivates defect sites, extending material durability under operational conditions. Ravelli and co-workers recently demonstrated the synergistic combination of Cs2AgBiCl6 with graphitic carbon nitride (g-C3N4), highlighting its impact on enhancing photocatalytic efficiency in a wide series of carbohalogenation reactions.165 The heterojunction structure ensures improved charge separation and stability, addressing the common limitations of single-component photocatalysts.
In summary, Ag–Bi PIMs offer great promise for PC and PEC applications, yet several challenges remain to be addressed to fully realize their potential. Key limitations include relatively low charge carrier mobility, incomplete understanding of the role of CD and intrinsic defects, and concerns regarding long-term stability under operational conditions, especially in aqueous media. Achieving scalable and reproducible synthesis, particularly for doped systems and nanostructured heterojunctions, requires further refinement. Moreover, optimizing dopant concentrations and surface ligands to precisely tailor electronic structures and enhance charge separation remains a critical task. While combining Ag–Bi PIMs with co-catalysts or forming heterostructures has shown encouraging results, the complex interfacial charge dynamics still need to be thoroughly understood. Future work should integrate advanced in situ characterization and computational modelling to elucidate defect behaviour and provide guidance for rational material design. By addressing these aspects, Ag–Bi PIMs could emerge as key players in the development of efficient, stable, and lead-free systems for solar-driven fuel production and environmental remediation.
Compositional engineering, particularly the selection of A-site cations and X-site halides, plays a pivotal role in defining key device parameters,169,174,175 such as the RS window (on/off ratio and set/reset voltage), endurance, and retention.174,176,177 While 3D structures are advantageous for photovoltaic and photodetector applications, lower-dimensional structures (0D, 1D, and 2D) offer enhanced stability and performance in memristors.177,178 Among these, the A-site cation is particularly influential: larger cations can significantly enhance the on/off ratio (from 103 to 108) and improve both thermal and environmental stability.175,177–179 For instance, incorporating butylammonium (BA) into Cs2AgBiBr6 yields the PIM BA2CsAgBiBr7. This distorted 2D structure facilitates ion migration and enhances RS behaviour. Devices based on BA2CsAgBiBr7 exhibit high on/off ratios (>107–108), likely due to a higher Schottky barrier at the electrode interface and constrained charge transport from deep trap states.175,177–179 As discussed earlier, halide composition significantly influences the structural dimensionality of PIMs, thereby affecting their thermal stability and memristor performance. For example, Cs2AgBiBr4Cl with its high chlorine content, demonstrates markedly improved characteristics, including enhanced data retention, greater endurance, and stable operation at temperatures up to 100 °C. These advantages are primarily attributed to its high activation energy, strong cohesive forces, reduced halide vacancy migration, and lower free carrier concentration, all contributing to superior thermal stability.180 Similar trends in A-site and X-site compositional engineering and their correlation with structural dimensionality and RS performance have been observed in various PIMs, including CuxAgBiI4+x,169 BA4AgBiBr8,179 and Cs2AgBiBr6−xClx.180 The RS mechanisms in these systems are primarily governed by either metallic filaments (in devices with electrochemically active electrodes) or anionic conductive filaments based on halide vacancies.17 However, the influence of CD and cation vacancies on memristor performance remains largely unexplored and warrants further investigation. While some studies have examined the role of the crystal phase in LHPs (e.g., FAPbI3),173 further exploration of polymorphism in other PIMs,181 including Ag–Bi systems, is imperative. Additionally, polymorphism must be considered alongside device configuration, particularly electrode selection, as a high interfacial energy barrier, resulting from work function mismatch, is often desirable for optimizing performance. Collectively, these insights pave the way for a new paradigm for highly stable and robust all-inorganic perovskite-type structures, deepening our understanding of how structural dimensionality and composition influence memristor properties.
Conventional memristive materials for neuromorphic devices include transition metal oxides (e.g., TiO2 and HfO2) and halide perovskites (e.g., MAPbI3).182,183 Ag–Bi PIMs exhibit comparable resistive switching while offering the advantages of being lead-free and environmentally benign.13,184,185 Their defect-tolerant, mixed ionic-covalent structure makes them promising for low-power synaptic devices.17,184 While transition metal oxides provide excellent endurance, they require high-temperature processing,183 whereas Ag–Bi PIMs are solution-processable at low temperatures, enabling compatibility with flexible electronics. However, their long-term stability and scalability need further optimization.13,184 Recent studies also demonstrate that Ag–Bi PIMs can integrate optoelectronic and memory functionalities, drawing parallels with oxide- and halide-based systems and opening pathways for multifunctional neuromorphic hardware.184,186,187
In conclusion, Ag–Bi PIMs represent a compelling class of materials for next-generation memristors, bridging the gap between the reliability of oxides and the tunability of halide perovskites, while eliminating lead toxicity. Continued progress in compositional engineering, defect control, and interface optimization will be critical to achieving the endurance, retention, and scalability required for practical deployment in neuromorphic and in-memory computing architectures.
In another study, single crystals of the 2D Dion-Jacobson double perovskite (R-/S-4APEA)2AgBiI8·0.5H2O (R-/S-4APEA = R-/S-(4-aminophenyl)ethylamine) were employed in a circularly polarized light photodetector. The material exhibited a low defect density, which is essential for achieving efficient charge transfer while simultaneously suppressing the dark current.189 As in previous cases, the non-centrosymmetric crystal structure enabled the generation of an internal electric field, allowing the device to operate under zero-bias conditions. Furthermore, the intrinsic properties of the material facilitated the activation of the pyro-phototronic effect, which produces an additional photocurrent in response to light-induced temperature changes. This synergistic mechanism led to a 40-fold enhancement in both responsivity and detection compared to devices relying solely on the bulk photovoltaic effect.189 Interestingly, although photon energies above 520 nm are insufficient to drive charge separation and transport due to the intrinsic band structure of the semiconductor, the pyro-phototronic effect remains active up to 940 nm, enabling the extension of the operational spectral range of the circularly polarized light photodetector into the NIR region.189
The previously discussed (R-/S-MPA)4AgBiI8 was also employed to build up a self-powered X-ray photodetector, demonstrating a notably flat and smooth morphology. These high-quality films—characterized by low defect density—are advantageous for efficient charge transport during direct X-ray detection. The material exhibited a short PL lifetime, an ideal feature for photodetectors, as it reflects strong exciton binding energy together with substantial bulk resistivity, both of which contribute to minimizing dark current and noise. Benefiting from the bulk photovoltaic effect, which enables operation at zero bias, the fabricated photodetectors achieved higher sensitivity than those based on achiral perovskites even when operated under high external bias. It must also be mentioned that the detection limit degrades upon increasing the bias due to amplification of dark current and noise induced by ion migration, highlighting the benefits of developing inherent non-centrosymmetric materials for self-powered devices.190
A dicationic chiral molecule, namely R-/S-/Rac-3-aminopyrrolidine (R-/S-/Rac-3AP), was used for the synthesis of (R-/S-/Rac-3AP)4AgBiBr12 in the form of both polycrystalline powders and highly crystalline microwire arrays oriented along the 001 crystallographic axis.191 Second harmonic generation (SHG) studies were performed, observing a higher SHG conversion efficiency for the microwire arrays upon equal excitation intensity, ascribed to grain boundary suppression, inferior propagation loss, and pure (001) orientation.191
Remarkably, in 2025, non-centrosymmetric structural centers were successfully introduced into the globally centrosymmetric structures of CABI and AgBiI4 without the use of chiral cations.12 In these materials, local centrosymmetry breaking was achieved through the presence of cation vacancies. Namely, while ions or atoms are symmetrically arranged with respect to an inversion center in centrosymmetric crystals without defects, cationic or atomic vacancies can locally lead to an unbalanced displacement of atoms disrupting the centrosymmetric nature of the crystal lattice. A stronger SHG signal was observed for CABI compared to AgBiI4, consistent with a higher concentration of cationic defects in the former. This finding highlights defect engineering as a powerful strategy for inducing NLO responses in metal halide systems without the need for chiral cations.12 This approach is particularly appealing, given the intrinsic steric limitations of chiral cations, which make the attainment of 3D structures difficult and restrict efficient charge transport along all crystallographic directions.192 Moreover, the thermal instability of organic molecules significantly restricts the application range of hybrid materials.
Although research interest in Ag–Bi PIMs has grown significantly in recent years, the number of comprehensive studies on NLO properties remains limited. There is an urgent need for a more systematic approach to assess how key parameters—including chemical composition, octahedral distortion, cation vacancies, network dimensionality, and morphology—affect their NLO properties. The incorporation of chiral organic molecules ensures global non-centrosymmetry across the material, thereby enabling second-order NLO responses such as SHG. In contrast, fully inorganic compounds like CABI and AgBiI4 exhibit SHG signals that predominantly originate from localized surface regions, where cation vacancy concentrations are higher. These materials, however, benefit from a fully three-dimensional connectivity of the inorganic framework, without the insulating organic spacer layers typically present in hybrid structures. In this context, the realization of a chiral 3D Ag–Bi PIM—analogous to recent developments in Pb-based systems—represents a highly promising direction.193
Comprehensive studies are essential to elucidate how the electronic and steric characteristics of chiral cations, along with the halide composition, influence the resulting optoelectronic and NLO properties. Parallel efforts toward the development of fully inorganic Ag–Bi PIMs should focus on expanding the family of compounds studied, integrating both experimental and theoretical approaches. This includes quantifying the role of octahedral distortion indices and cation vacancy distributions, as well as engineering defect structures and optimizing film-processing conditions, with the ultimate goal of maximizing the NLO functional response.
AgBiS2, for instance, has demonstrated notable potential as an anode material in alkali-ion batteries. Its nanocrystals exhibit reversible capacities (∼420 mAh g−1) and remarkable cycling stability in potassium-ion battery systems, driven by reversible Ag/Bi redox reactions and efficient ion diffusion pathways.194 In addition, AgBiS2 has also been explored for charge-storage applications, achieving a specific capacitance of approximately 14 F g−1 at 2 mV s−1, an energy density of 26 Wh kg−1, and a power density of 3.6 kW kg−1.195 The inherent CD and mixed ionic-electronic conductivity significantly enhance its energy-storage capabilities, underscoring its potential for high-performance supercapacitors. Additionally, AgBiS2 demonstrates efficient near-infrared-driven photocurrents (∼2.3 mA cm−2 at ∼800 nm), sufficient to enable neuronal stimulation, thus suggesting its potential use in bioelectronic interfaces, such as neural implants and biosensors.196
Furthermore, Ag–Bi halide materials, particularly Cs2AgBiBr6, have emerged as promising candidates for solid-state X-ray detection. Their heavy-atom lattice structure combined with defect-tolerant wide bandgaps results in exceptional hard X-ray sensitivity (∼1390 µC Gy−1 cm−2 at 100 keV), ultralow dark current drift (1.03 × 10−8 nA cm−1 s−1 V−1), and remarkably low detection limits (<100 nGyair s−1), outperforming commercial CZT detectors in certain key metrics.197 These properties position Cs2AgBiBr6 and related Ag–Bi PIM-based detectors as competitive alternatives to conventional inorganic X-ray detectors, offering distinct advantages such as improved stability, lower toxicity, and enhanced environmental friendliness. Interestingly, due to their typically short charge-carrier diffusion lengths arising from strong coupling with lattice vibrations, Ag–Bi halide PIMs are particularly well-suited for drift-driven applications such as radiation detection, rather than diffusion-driven photovoltaic devices.14 Under applied electric fields, these materials can achieve notably high drift mobility-lifetime (µτ) products, making them promising candidates for high-performance radiation detectors. Notably, recent studies have highlighted Cs2AgBi2I9 for achieving the highest reported µτ product (3.4 × 10−3 cm2 V−1) among Bi-halide PIMs.41
Collectively, these emerging energy storage, sensing, and radiation detection applications highlight the multifunctional potential of Ag–Bi PIMs, suggesting significant opportunities beyond traditional photovoltaic or optoelectronic domains.
CABI and Cs2AgBi2I9 remain predominantly solution-processed (co-evaporated CABI films exist but are transport-limited),33 yet reproducible planar/mesoscopic device stacks—including indoor-PV architectures—are now routine. By contrast, chalcohalides (e.g., AgBiSCl2) are successfully synthesized at the nanocrystal level (direct bandgap, robust films)10 yet still lack module-relevant, large-area thin-film processes—an obvious near-term target for process development.
Long-term photovoltaic stability in Ag–Bi semiconductors is increasingly supported by targeted chemistry and interfaces. In Cs2AgBiBr6, hydrogenation narrows the bandgap (≈2.18 → ≈1.64 eV) and delivers steady solar cell performance under illumination at 85 °C with a record PCE of 6.37%, indicative of drift-suppressed behaviour under coupled heat/light stress. Mechanistically, interstitial H tunes band edges and reduces halide-vacancy-related traps.134 Beyond hydrogenation, interface/dopant engineering has delivered ≈90% PCE retention after ambient and heat aging (unencapsulated) in Cs2AgBiBr6—consistent with a picture where contact chemistry and defect passivation dominate operational loss channels.203 Light-stress studies further confirm that while the bulk lattice is thermally rugged (to ∼410 °C), photo-driven pathways can emerge at higher intensities or with poor interfaces, underscoring the importance of ion-migration barriers and surface passivation in stack design.204
In CABI, non-encapsulated solar cells retained most of their performance for a few weeks, and halide/cation/interface engineering further suppresses non-radiative losses and enhances device stability.43,45,135 AgBiS2 thin-film devices—when densified by solution crystallization or vapor-assisted routes—exhibit continuous 1-sun stability and months-long dark storage, mitigating the durability limits of ligand-rich NC-only layers. Vapor-assisted, solution-processed AgBiS2 solar cells have also exhibited excellent stability—from hundreds up to ∼3000 h under storage in ambient humidity and 85 °C.11,137
For AgBiS2, quantitative ISOS-type datasets show ≥94% PCE retention after ∼3000 h ambient and ∼87% after 1000 h at 85 °C, while shorter operational MPP traces highlight the importance of publishing both shelf and operational stability.11 State-of-the-art NC-based AgBiS2 solar cells demonstrated T80 of just ≈15 h (≈900 min) under maximum power point tracking.207 This is significantly lower operational stability compared to LHPs, which routinely demonstrate thousands of hours of operational stability. These examples collectively illustrate that adopting ISOS/IEC practices is essential for cross-lab comparability and benchmarking Ag–Bi PIMs against LHPs and other emerging/established PV technologies. For indoor PV, the operational stability data are very limited; Cs2AgBi2I9 maintains T80 ≈ 35 h under continuous 1000 lux illumination with MPP tracking.208
For memristors, stability manifests as low-voltage switching with quantifiable endurance and retention. Cs2AgBiBr6 devices operating near +0.4/−0.3 V with on/off ≈4.8 × 102, ≈103 cycles, and ≈104 s retention exemplify the current window; failure modes map to ionic/filamentary drift and electrode reactivity, which are mitigated by interlayers and film densification.184 AgBiS2 quantum-dot memristors demonstrate multilevel switching at very low set power (∼0.065 mW) and ≈500-cycle pulsed endurance, consistent with Ag-ion migration and S-vacancy-assisted filaments. This is again pointing to barrier/electrode chemistry as the principal means for drift suppression.210 In photocatalysis/PEC, Ag–Bi materials show encouraging operando resilience when interfaces are engineered. AgBiS2 solids are water-resistant at the film level (pre-/post-XPS/XRD show no compositional/structural change after water immersion), and neutral-electrolyte PEC operation is documented.160,161 Epitaxial ZnS shells suppress surface traps/corrosion, which can extend the steady-state operation.47 Long-run device-level datasets are currently unavailable for chalcohalides (e.g., AgBiSCl2).
To improve reproducibility, we align explicitly with ISOS (outdoor; stabilized MPP, T80, and defined light/heat/humidity) and IEC TS 62607-7-2:2023 (indoor; 50/200/1000 lux, spectral reporting, and stabilized output). Because interfaces and defect chemistry repeatedly control stability across applications, a pragmatic round-robin on one halide (Cs2AgBiBr6) and one chalcogenide (AgBiS2)—covering ISOS/IEC PV, memristor endurance/retention, and PEC steady-state with FE/ICP-MS—would separate materials from tool/process effects and accelerate the consensus on lifetimes and efficiency levers.
Under 1-sun (AM 1.5G), state-of-the-art LHPs reach ≈26–26.7% single-junction PCE, with hundreds of hours of operational stability already demonstrated in well-engineered stacks. By contrast, Ag–Bi PIM solar cells are still efficiency-limited. The best Ag–Bi halide PIM to date is hydrogenated Cs2AgBiBr6 at ∼6.4% (J–V) after narrowing the bandgap to ∼1.64 eV.134 Performance of Ag–Bi iodides remains lower outdoors (e.g., Ag–Bi–I ≈ 4.5%, Cs2AgBi2I9 ≈ 3% and CABI ≈ 2.2% in recent optimized stacks),7,8,136 consistent with wide-gap/defect-limited VOC and transport losses that remain to be solved. Ag–Bi chalcogenides now deliver ≥10% PCE (AM 1.5G), establishing the highest outdoor efficiency within Ag–Bi PIMs but still less than half of the LHP's record value.11,207 Under indoor lighting (WLED), the comparison is more nuanced because the optimal bandgap shifts to ≈1.9 eV. Ag–Bi iodides sit near this optimum bandgap and avoid the mixed-halide phase-segregation commonly encountered when LHPs are tuned to 1.9 eV for indoor use. Theoretical, radiatively limited indoor maximum efficiency (i-SLME) ceilings for many Ag–Bi iodides approach 50%.140 Experimentally, Ag–Bi PIMs have progressed from early 4–5% IPV (Ag–Bi–I) to ≈8% at 1000 lux for Cs2AgBi2I9.8 These values remain below best-in-class LHP IPV reports (>40%) but demonstrate that the Ag–Bi family is competitive in bandgap suitability and increasingly credible in device metrics under standardized indoor testing.
Although Ag–Bi PIMs significantly trail state-of-the-art LHPs for photovoltaics, additive/interface engineering continues to improve film quality and device metrics—material levers that are equally consequential for radiation detectors and memristors.
Benchmarks for Cs2AgBiBr6 include sensitivities up to ≈1974 µC Gy−1 cm−2 with appropriate interface passivation and limits-of-detection around 59.7 nGy s−1.214 Polycrystalline thick-film Cs2AgBiBr6 has also shown a sensitivity of ≈487 µC Gy−1 cm−2, and related Ag–Bi compounds (e.g., AgBi2I7) report limits-of-detection of ≈72 nGy s−1.215 For context, legacy a-Se panels offer ≈20 µC Gy−1 cm−2 sensitivity, while LHP single-crystal detectors span ≈9300–24,552 µC Gy−1 cm−2 sensitivity with limits-of-detection ≈0.22–54 nGy s−1 depending on the architecture and crystal quality.216,217 Cs2AgBiBr6 has also shown 59.5 keV γ-ray response with ≈13.9% energy resolution and measurable responses at 511 and 662 keV (ref. 218)—evidence that Ag–Bi halide double perovskites can extend beyond diagnostic X-rays into higher-energy regimes. Thus, Ag–Bi detector benchmarks already clear the thresholds used to judge a-Se flat-panel photoconductors and close the gap to the best LHP single-crystal photodiodes, while offering an intrinsically Pb-free route and device-relevant stability when contact chemistry and ion migration are controlled.
In nonlinear optics/chiroptical detection, chiral Ag–Bi double-perovskite microwire arrays ((R/S-β-MPA)4AgBiI8) yield circular-polarization photodetection with g-factor ≈0.19, responsivity >52 mA W−1, and detectivity >3.9 × 1011 Jones—values competitive for device-level circularly polarized light sensing.219 Representative Pb-based chiral perovskite comparators can reach higher g (≈0.2) and responsivity near ≈0.8 A W−1,220,221 underscoring the headroom for Ag–Bi optimization while already enabling useful circularly polarized light discrimination in lead-free systems. Recent work by Fan et al. demonstrates that lead-free Ag–Bi-based double perovskites can deliver superior SHG performance compared to LHPs, primarily due to the highly distorted [AgBr6]5− octahedra.222 These distortions break centrosymmetry more effectively than the relatively ideal [PbX6]4− configurations in lead-based systems, enabling stronger nonlinear optical responses. The compound DFPD4AgBiBr8·H2O exhibits a record SHG intensity of 13 × KH2PO4 (KDP), significantly surpassing the best-performing lead-based analogs while also offering enhanced environmental stability and a high laser-induced damage threshold.222 This study, alongside the recent observation of SHG from non-centrosymmetric Cu–(Ag)–Bi–I thin films,12 emphasizes the unique promise of lead-free Ag–Bi PIMs for nonlinear optical applications.
In memristive switching, Ag–Bi PIMs demonstrate low-voltage operation and robust figures-of-merit: AgBiI4 devices switch at ≈0.16 V with on/off ≈104, >700-cycle endurance, ≳104 s retention, and stability under 1000 bending cycles;185 2D Ag–Bi systems such as (BA)2CsAgBiBr7 report on/off >107–108 with improved retention.175 LHP memristors set the current performance ceiling—for example, quasi-2D devices with on/off ≈109 (±0.8 V), ≈230 cycles and ∼103 s retention, with broader reviews placing typical LHP on/off in the 103–106 range and endurance up to 103–104 cycles.223,224 These side-by-side values show Ag–Bi devices already competitive for low-voltage switching, with clear headroom in endurance scaling.
These diverse capabilities underscore that Ag–Bi PIMs are far more than niche photovoltaic absorbers—they are versatile functional semiconductors poised to meet the emerging needs in energy, computing, and sensing—with the important benefit of eliminating lead.
Reaching these milestones will require a deeply interdisciplinary approach. Chemists, materials scientists, theoreticians, and device engineers must work to synchronize their approaches across domains—for example, by combining computational predictions with in situ spectroscopy and advanced imaging techniques to identify and eliminate performance-limiting defects. Collaboration with application-domain experts is equally important: as the demands for IPVs, photonic computing, and radiation sensing evolve, feedback from those communities will help steer Ag–Bi PIM development toward the most impactful and relevant solutions. Notably, the lead-free, low-toxicity nature of these materials is expected to reduce barriers to deployment in sensitive environments (from living spaces to clinical settings)—aligning with the global push for sustainable and safe electronics.
In summary, Ag–Bi PIMs are at a promising turning point. Having overcome key challenges related to stability and synthesis, the field is now poised to move beyond just catching up with LHPs and begin exploring new directions that those toxic materials could not safely reach. By systematically improving material quality in the short term, innovating device architectures in the medium term, and embracing system-level design in the long term, Ag–Bi PIMs have the potential to enable a new generation of clean energy harvesters, intelligent sensors, and hybrid optoelectronic–photonic systems.
The outlook is ambitious, but with creative research and sustained cross-disciplinary collaboration, Ag–Bi PIMs can evolve from a promising alternative into a transformative technology for next-generation energy, sensing, and computing platforms.
| This journal is © The Royal Society of Chemistry 2025 |