Ziming Kuang,
Baoyun Liang,
Tengcheng Huang,
Tingting Shi* and
Weiguang Xie
*
Siyuan Laboratory, Guangzhou Key Laboratory of Vacuum Coating Technologies and New Energy Materials, Department of Physics, College of Physics & Optoelectronic Engineering, Jinan University, Guangzhou 510632, China. E-mail: ttshi@jnu.edu.cn; wgxie@jnu.edu.cn
First published on 4th August 2025
Against the backdrop of intensifying global energy challenges, perovskite materials with remarkable photovoltaic potential in solar cell development continue to face dual constraints: lead toxicity and poor stability in organic–inorganic hybrid structures. As a promising alternative, all-inorganic germanium-based perovskites have garnered significant attention in recent years. Given the absence of systematic theoretical investigations on germanium-based systems, this review comprehensively examines recent theoretical advances in this emerging field. The discussion encompasses three key aspects: (i) structural characteristics, including Ge-induced lattice configurations, tolerance factors, and phase stability; (ii) optoelectronic properties, particularly Shockley–Queisser limit evaluations, theoretical bandgap predictions, germanium-induced optoelectronic modulation, and related exploratory investigations; and (iii) supplementary properties such as mechanical response, pyroelectric behavior, and ferroelectric functionality. Theoretical investigations reveal distinctive advantages, notably enhanced structural stability through non-centrosymmetric distortions and defect-dependent conductivity under various configurations. However, challenges persist in fabrication techniques and optimization of power conversion efficiency. This review concludes by proposing four strategic research directions: discovery of novel perovskite derivatives, optimization of compositional engineering, atomistic-level defect analysis, and exploration of multifunctional applications, providing a roadmap for future investigations in this burgeoning field.
However, perovskite solar cells (PSCs) still face multiple challenges in their path toward widespread commercialization, with film stability and toxicity issues posing significant barriers. Although organic–inorganic hybrid perovskites demonstrate impressive power conversion efficiencies (PCEs) in solar cells,5–8 their organic cations exhibit proneness to decomposition under high-temperature and high-humidity conditions. This inherent instability severely compromises film integrity, thereby constraining the practical deployment of hybrid perovskites in photovoltaics.9,10 Furthermore, the toxicity of lead—a critical component in high-performance perovskites—remains a major impediment to their commercial viability. To address this challenge, researchers have pursued extensive investigations focusing on replacing lead with non-toxic divalent cations that exhibit strong optical responses, thereby mitigating toxicity concerns. Promising candidate cations currently under investigation include Sn2+, Ge2+, Si2+, Sb2+, and Bi2+.11–17 Tin has gained prominence as a lead substitute candidate due to its closely analogous electronic and structural characteristics to lead,18,19 combined with the recent demonstration of tin-based perovskite solar cells achieving power conversion efficiencies above 13%.20 However, tin-based perovskites face two critical challenges such as atmospheric oxidation of Sn2+ to Sn4+ that degrades stability,21,22 alongside inherent high surface vacancy densities impairing luminescence properties in solar cell applications.23
In the evolution of perovskite research, germanium-substituted all-inorganic perovskites have emerged as a scientific focal point, addressing the dual challenges of instability and lead toxicity inherent in conventional counterparts. Fig. 1 demonstrates remarkable breakthroughs in photoelectric conversion efficiency achieved by this material system over the past decade. Krishnamoorthy et al. computationally identified germanium as a lead alternative through density functional theory (DFT) screening. Their experimental work on CsGeI3 crystals revealed a stable rhombohedral crystal structure maintaining phase integrity within operational temperature ranges. The derived solar cells exhibited photocurrent densities approximating 6 mA cm−2; yet, performance optimization was hindered by inadequate film-forming capability and atmospheric ambient instability.24 This finding demonstrates the photovoltaic potential of germanium-based perovskites while revealing critical challenges requiring resolution. Montiel's team developed CsGeI3 thin films via physical vapor deposition of CsI and GeI2 precursors followed by annealing, achieving enhanced thermal stability albeit with a modest PCE of 0.58%.25 While the efficiency remained suboptimal, this methodology pioneered new synthetic pathways for subsequent investigations. Chen et al. pioneered a simple solvothermal process for synthesizing lead-free cesium germanium halide perovskite quantum rods. The quantum rods (QRs), with larger particle volumes compared to quantum dots (QDs), provided enhanced per-particle absorbance cross sections, enabling CsGeI3 QR-based devices to attain a PCE of 4.94%.26 Liu et al. developed all-inorganic Ge-alloyed perovskite nanocrystals CsSn0.6Ge0.4I3. The derived photovoltaic devices exhibited a PCE of 4.9%. These nanocrystals demonstrated significantly enhanced stability under environmental, thermal, and optical stresses compared to CsSnI3 counterparts.27 Yang et al. synthesized novel inorganic CsPb1−xGexI2Br perovskites under humid conditions. Planar solar cells based on the CsPb0.8Ge0.2I2Br perovskite achieved a record PCE of 10.8%. Remarkably, the devices maintained nearly intact photovoltaic parameters after 7-hour exposure to 50–60% relative humidity.28 This work represents dual advancements in both efficiency enhancement and stability improvement. Chen's team employed lead-free all-inorganic CsSn0.5Ge0.5I3 solid-solution perovskites as light-harvesting layers in solar cells. The thin films were fabricated via one-step vapor-phase synthesis and integrated into planar heterojunction architectures. This configuration attained a PCE of 7.11%.29 Persistent explorations in material systems, synthetic methodology innovation, and processing condition optimization have collectively driven progressive performance enhancements. These achievements systematically demonstrate incremental breakthroughs in photoconversion efficiency for germanium-based perovskites. They unequivocally confirm the exceptional research viability and substantial development potential of this material family.
![]() | ||
Fig. 1 (a) Evolution of PCE in all-inorganic germanium-based perovskites from 2014 to 2024.24–29 |
However, it is crucial to note that, analogous to tin ions, germanium ions exist in a divalent state within germanium-based perovskites and exhibit a strong proneness to oxidation, as evidenced by multiple experimental studies.30,31 To address this challenge, researchers have undertaken in-depth theoretical investigations into the stability of germanium-based perovskites. For instance, Liu et al. conducted formation energy calculations using the Vienna Ab Initio Simulation Package (VASP).32 Their computational results revealed that the formation energies of CsGeBr3, CsGeI3, and Cs2GeCl6 are all negative, indicating the thermodynamic stability of these materials. Furthermore, Sun et al. also performed formation energy calculations on MAMX3 (M = Pb, Sn, and Ge and X = I, Br, and Cl).33 They found that germanium-based perovskites are less susceptible to oxidation compared to their tin-based counterparts, and that MAPbI3 and MAGeI3 exhibit comparable stability. Moreover, Tian et al. employed DFT to design cerium (Ce) doping ratios.34 Subsequent experimental synthesis demonstrated that the incorporation of Ce effectively mitigated moisture and air degradation of CsGeBr3, thereby enhancing its relative stability.
While organic–inorganic hybrid germanium-based perovskites have been extensively studied and demonstrate excellent potential for photovoltaic applications,35,36 their stability issues, particularly the inherent instability of organic–inorganic hybrid systems, remain a critical bottleneck hindering the development of perovskite materials. Regarding toxicity concerns, the detrimental effects of lead have been well-documented. Upon degradation and dissolution in water, lead-based perovskites release highly toxic Pb2+ ions that pose significant bioaccumulation risks.37 The toxicity profile of tin-based perovskites remains contentious: while Conings et al. initially suggested that tin halide perovskites might exhibit greater toxicity than lead-containing analogues,38 subsequent experimental validation by Xiao et al. confirmed their safety under practical operating conditions.39 In contrast, germanium's biocompatible nature renders Ge-based perovskites the most environmentally benign alternative in this materials landscape. In this context, germanium-based all-inorganic perovskites demonstrate promising potential in addressing stability and environmental concerns of conventional perovskites, making their systematic investigation crucial for advancing perovskite technologies. Theoretical approaches have emerged as pivotal tools in this domain due to their rapid implementation and methodological accessibility. Against this backdrop, this review concentrates on theoretical advancements in germanium-based all-inorganic perovskites. Structural analyses encompass a systematic examination of crystal structure stability, atomic configuration patterns, and lattice parameters, elucidating their deterministic roles in macroscopic material behavior. Optoelectronic investigations prioritize band structure engineering and gap modulation mechanisms. The scope further extends to multifunctional performance metrics including mechanical robustness, thermoelectric response, and ferroelectric functionality. Concluding with a comprehensive and in-depth analysis, we present forward-looking perspectives on the future development of germanium-based all-inorganic perovskites.
![]() | ||
Fig. 2 Crystal structures of CsGeI3: (a) R3m and (b) Pmmm.44 Reproduced with permission. Copyright 2021, the Royal Society of Chemistry. Crystal structures of CsGeCl3: (c) R3m and (d) Pm3m.45 Reproduced with permission. Copyright 2025, the Royal Society of Chemistry. (e) Relative energies of CsSn0.5Ge0.5I3 in different space groups. (f) P1-phase structure of CsSn0.5Ge0.5I3.46 Reproduced with permission. Copyright 2023, American Chemical Society. |
In 2019, Chen et al. reported a Ge–Sn hybrid perovskite solar cell achieving 7.11% PCE with exceptional operational stability.29 This breakthrough demonstrated the viability of germanium substitution for partial tin replacement in enhancing perovskite optoelectronic performance. Luo et al. focused on the energy differences between centrosymmetric and non-centrosymmetric states in CsSnI3 and CsGeI3 systems. Through first-principles calculations, they systematically compared the structural energies of CsSnxGe1−xI3 in R3m and P4/mbm phases across varying doping ratios (x). Their results revealed near-degenerate energy states between the two space groups at x = 0.5, leading to a structural phase transition. Furthermore, they proposed a stabilized distorted structural model for CsSn0.5Ge0.5I3 based on computational evidence. In this configuration, both Sn2+ and Ge2+ deviate from octahedral centers, forming distorted octahedral units. Researchers concluded that octahedral distortion combined with non-centrosymmetric displacement plays a critical role in stabilizing the low-energy state of the structure.44 In a corroborative study, Wu et al. calculated the structural energies of CsSn0.5Ge0.5I3 across multiple space groups. As shown in Fig. 2(e), they discovered a novel P1-phase Sn–Ge distorted perovskite structure exhibiting superior stability compared to other space groups. The atomic arrangement of this P1-phase CsSn0.5Ge0.5I3 is detailed in Fig. 2(f).46
![]() | (1) |
We systematically reviewed computational studies on tG values for all-inorganic germanium-based perovskites, providing insights into their structural stability. It should be emphatically noted that for germanium-based perovskites exhibiting significant octahedral distortions (e.g., asymmetric Ge–X bond stretching), the tolerance factor (tG) remains a preliminary screening indicator for structural stability; yet, it must not be considered an exclusive criterion. Comprehensive stability assessment necessitates multidimensional evaluation incorporating formation energy, cohesive energy, and lattice dynamics parameters. As summarized in Fig. 3, 11 out of 15 germanium-based perovskites exhibit tG values within the ideal cubic phase prediction range (0.9–1, shaded in pale yellow), highlighting their structural compatibility. Jong et al. conducted comprehensive theoretical investigations on AGeX3 (A = Cs and Rb and X = I, Br, and Cl) systems using VASP. Their computational results revealed that all six compounds exhibit tG within the range of 0.90 < tG < 0.97, suggesting their propensity to crystallize into stable perovskite phases. Furthermore, employing the PBEsol functional, they systematically investigated the lattice constants and bond angles of CsGeX3 (X = I, Br, and Cl). The calculations demonstrated a progressive reduction in lattice constants and a convergence of bond angles toward 90° as the ionic radius of the X-site halide anions decreases (from I to Cl). These observations indicate lattice contraction accompanied by a rhombohedral-to-cubic phase transition. This structural evolution aligns with the observed monotonic increase in tG values from X = I to X = Cl, where tG = 1 corresponds to an ideal cubic perovskite configuration. Researchers attributed this trend to the synergistic effects of the off-centering displacement of Ge atoms and the strengthening of Ge–X chemical bonds.50 These findings corroborate the structural characteristics of germanium-based perovskites discussed in preceding sections. Similarly, Pak et al. performed DFT calculations on ABF3 systems (A = Na and K and B = Si and Ge). Their results indicated that all examined structures except KSiF3 possess tolerance factors consistent with stable BF6 octahedra.51 Hasan et al.'s theoretical investigation on indium-based halide germanium perovskites revealed tolerance factors of 0.79 and 0.78 for InGeCl3 and InGeBr3, respectively. While these values fall outside the ideal tolerance factor range (0.9–1.0), they remain in close proximity to this stability window. The authors conclude that such near-optimal values suggest that these compounds retain some propensity for perovskite formation.52
![]() | ||
Fig. 3 Tolerance factors (tG) of various perovskite compounds.51,52,56,60,61 The pale yellow shaded region indicates materials predicted to form ideal cubic perovskite structures. |
Beyond the conventional Goldschmidt tolerance factor, researchers have pursued alternative approaches to assess structural stability in all-inorganic germanium-based perovskites. Ruddlesden–Popper halide perovskites (RP-HPs), characterized by distinctive two-dimensional layered architectures, have garnered significant scientific interest. Pioneering work by Li et al. demonstrated that all-inorganic two dimensional Ruddlesden–Popper halide perovskites (2DRP-HPs) exhibit reduced binding energies, enhanced optical responses, and superior stability compared to hybrid counterparts.53 To address lead toxicity in 2DRP-HPs, Haq et al. employed the full-potential linearized augmented plane wave (FPLAPW) method within the Wien2K code to investigate Cs2BX2Y2 systems (B = Ge and Si and X/Y = Cl/Cl, Br/Cl, I/Cl, Br/Br, I/Br, and I/I), systematically analyzing their structural, electronic, and optical properties. These systems crystallize in tetragonal symmetry with the space group I4/mmm (No. 139). For enhanced stability prediction, the authors introduced a modified tolerance factor τ defined as follows:54
![]() | (2) |
Moreover, in experimental research, the stability of materials is closely linked to the synthesis method employed. Chen et al. first synthesized CsGeX3 QRs via a facile solvothermal method. Transmission electron microscopy (TEM) images and energy-dispersive X-ray spectroscopy (EDS) analysis confirmed the uniform morphology and good crystallinity of these QRs. They further compared the stability of CsGeI3 and CH3NH3PbI3 perovskite thin films; testing revealed that the CsGeI3 film decomposed only about 45% after 48 hours in air, significantly superior to the CH3NH3PbI3 film, which exhibited a decomposition rate as high as 98% within 24 hours.59 Montiel et al. fabricated CsGeI3 thin films using physical vapor deposition (PVD) technology, demonstrating enhanced thermal stability compared to MAPbI3.25 However, research efforts focused on the synthesis of all-inorganic germanium-based perovskites remain relatively scarce.
![]() | ||
Fig. 4 Bandgap values of various all-inorganic germanium-based perovskites using different computational methodologies.32,50–52,61,64,68,69 The dark gray zone indicates the ideal bandgap range, with green asterisks marking materials whose calculated values fall within this regime. The light gray zone demarcates bandgap ranges achieving about 25% Shockley–Queisser efficiency, where the corresponding materials are highlighted by yellow asterisks. |
In theoretical investigations of material bandgaps, the substantial computational discrepancies arising from functional selection merit particular attention. The systematic bandgap underestimation inherent to standard DFT methodologies necessitates the careful selection of computational approaches. Jong et al. employed four computational schemes combining PBEsol/HSE06 functionals with/without spin–orbit coupling (SOC) effects to evaluate CsGeX3 (X = I, Br, and Cl) bandgaps. Comparative analysis with experimental data revealed stark contrasts: while PBE/PBEsol accurately reproduce bandgaps in organic systems like MAPbI3,65,66 these functionals systematically underestimate values in all-inorganic germanium-based perovskites. SOC inclusion exacerbates this underestimation, whereas HSE06 hybrid functionals achieve remarkable agreement with experimental measurements. Band structure analysis attributes this accuracy to the HSE06's proper treatment of conduction band uplifting and valence band depression.50 Liu et al.'s computational results revealed pronounced bandgap underestimation in CsGeX3 systems using the PBE functional, particularly exacerbated in CsGeCl3 and CsGeBr3 configurations. HSE06 functional calculations yielded bandgap values closely aligned with experimental reports for CsGeBr3 and CsGeI3,40 while persistent underestimation remained in CsGeCl3.32 In 2023, given the systematic PBE-based bandgap underestimation in all-inorganic halide perovskites, Pak's team adopted the HSE06 hybrid functional (without SOC) for ABF3 (A = Na and K; B = Si and Ge) systems to achieve enhanced computational accuracy. Parallel PBE functional calculations were systematically conducted for comparative methodology evaluation. Their results demonstrated systematically larger bandgap values obtained from HSE06 compared to PBE in fluoride perovskite systems. This discrepancy originates from HSE06-induced conduction band uplifting and valence band depression, collectively enlarging the bandgap.51 Haq et al. further demonstrated the superior accuracy of GGA-mBJ potential over conventional GGA in bandgap calculations for all-inorganic 2DRP-HPs. Standard GGA approaches exhibited substantial bandgap underestimation. Remarkably, GGA-mBJ-derived bandgaps showed consistency with HSE hybrid functional results regardless of SOC inclusion.56 Beyond DFT calculations, the GW approximation has been demonstrated to accurately describe the electronic structures of materials. It provides high-precision predictions for semiconductor bandgaps and other electronic properties. Numerous research groups have applied this method to predict the electronic structures of perovskites.37,38,67 However, its computational cost remains prohibitively high.39 Consequently, it is primarily recommended for theoretical investigations where precision is paramount.
Hasan et al. computationally compared the wavelength-dependent absorption spectra of FrBX3 (B = Ge and Sn and X = Cl, Br, and I) series, with key findings visualized in Fig. 5(a) and (b). Within the solar spectrum (380–750 nm wavelength range, accounting for 43% of total energy), comparative analysis of absorption spectra revealed that substituting germanium with tin at the B-site induces significant attenuation of absorption peaks. This conclusively demonstrates the superior visible-light energy conversion capability of germanium-based FrGeX3 perovskites over their tin-based FrSnX3 counterparts.64 Wu et al. computationally mapped the bandgap evolution of CsGe1−xSnxI3 across varying doping ratios (x) under different space groups, as shown in Fig. 5(c). Remarkably, the R3m phase exhibits the widest bandgap tunability (1.02–0.49 eV). Density of states (DOS) analysis reveals that the higher energy level of Sn-5s orbitals compared to that of Ge-4s orbitals drives progressive bandgap reduction with increasing x (i.e., decreasing Ge content).46 Employing PBE functional calculations, Pak et al. evaluated carrier effective masses in ABF3 (A = Na and K and B = Si and Ge). Notably, germanium-based AGeF3 compounds demonstrate higher electron and hole effective masses than silicon-based ASiF3 analogues. While this suggests superior light absorption in silicon-based systems, the sufficiently low effective masses in AGeF3 remain viable for charge transport applications. Calculations of the optical absorption coefficients further indicate that silicon-based ASiF3 (A = Na and K) perovskites have slightly lower photon energy absorption onsets than germanium-based counterparts. However, both materials exhibit peak absorption coefficients of 11–18 × 104 cm−1 (Fig. 5(d)), demonstrating adequate light-harvesting capabilities. These characteristics are advantageous for improving solar cell efficiency.51 Experimentally, Yang et al. developed Ge–Pb perovskite luminescence films by partially substituting lead (Pb) with germanium (Ge). Their research revealed that at Ge molar fractions of 10–20 mol%, the photoluminescence quantum efficiency (PLQE) of the films significantly increased from 53% for Ge-free reference samples to approximately 71%. Further transient photoluminescence (PL) decay kinetics analysis indicated a distinct concentration threshold effect associated with Ge incorporation: within the optimal Ge content range (10–20 mol%), radiative recombination predominates, effectively suppressing non-radiative losses and thereby achieving high PLQE; when the Ge molar fraction exceeds 30 mol%, trap-assisted recombination becomes dominant, resulting in a sharp decline of PLQE below 40%.70
![]() | ||
Fig. 5 Wavelength-dependent absorption spectra: (a) FrGeX3 and (b) FrSnX3.64 Reproduced with permission. Copyright 2022, the Royal Society of Chemistry. (c) Calculated bandgaps of CsGe1−xSnxI3 in different phases.46 Reproduced with permission. Copyright 2023, American Chemical Society. (d) Optical absorption coefficients of fluoro-perovskites ABF3 (A = Na and K and B = Si and Ge) calculated using DFPT (solid lines) and BSE/GW (dashed lines) methods.51 Reproduced with permission. Copyright 2023, the Royal Society of Chemistry. |
Liu et al. employed the HSE06 hybrid functional to calculate the bandgap of CsGeI3 under varying strains. The results revealed that the bandgap could be tuned from 0.73 eV to 2.30 eV within a strain range of −4% to 4%.60 This aligns with the bandgap variation trend observed by Xu et al. under −2% to 2% strain. Notably, their comparative analysis revealed that CsPbI3 exhibits bandgap widening (from 1.53 eV to 1.64 eV) under compressive strain (−2% to 0%), whereas it shows bandgap narrowing (from 1.64 eV to 1.61 eV) under tensile strain (0% to 2%). In striking contrast, the CsGeI3 system demonstrates the monotonic bandgap increase across the full −4% to 4% strain range, signifying superior bandgap tunability.71 Furthermore, Liu et al. discovered that, under 0.52 GPa pressure, CsGeI3 exhibits a bandgap of 1.36 eV, falling within the 1.3–1.4 eV range. This meets the Shockley–Queisser Limit criteria, indicating optimal photovoltaic performance. Using the PBE functional with a scissor operator, they also calculated the optical absorption spectra of CsGeI3 under various strains, as shown in Fig. 6(a). The figure clearly shows that compressive strain induces a significant red shift of the absorption edge, while tensile strain leads to a blue shift. Moreover, strained structures exhibit enhanced absorption intensity in the visible light region.60 Furthermore, strain engineering has also yielded significant successes in the experimental modulation of germanium-based perovskites. Yang et al. developed low-toxicity perovskite light-emitting diodes (PeLEDs) by partially substituting Pb with Ge. However, they observed that at Ge contents exceeding 30 mol%, intense spontaneous strain (particularly shear strain) exacerbated non-radiative recombination, leading to a drastic deterioration in device performance.70 To overcome this limitation, Zhou et al. proposed an A-site strain engineering strategy, which involved substituting small Cs+ ions with larger FA+ (formamidinium) and MA+ (methylammonium) ions. This approach effectively suppressed octahedral tilting and reduced lattice distortion. Experimental results confirmed that the triple-cation structure [FA0.25(Cs0.95MA0.05)0.75]GexPb1−xBr3, with 30 mol% Ge content, achieved a PLQE of 48% and a maximum external quantum efficiency (EQE) of 8.5%. The underlying mechanism was attributed to the reduction of shear strain components e4 to <1% and etx to −0.1%.72 These findings demonstrate that strain engineering offers a viable strategy for optimizing the optoelectronic properties of all-inorganic germanium-based perovskites. Previous studies have demonstrated that SnI, as the primary defect in CsSnI3, generates a deep in-gap level that severely limits the PCE.73 Zhou et al. computationally revealed that GeI defects in CsGe1−xSnxI3 alloys exhibit comparable defect formation energies to SnI. Their investigation of both defects in CsGe1−xSnxI3 showed that neither generates deep levels at x = 0.5 (CsGe0.5Sn0.5I3), whereas at x = 0.25 or 0.75, at least one defect introduces the deep level. As shown in the defect energy level diagram, GeI and SnI induce two distinct defect states: one singly occupied (occ1) near the CBM, and another doubly occupied (occ2) adjacent to the VBM. Band structure analysis of CsGexSn1−xI3 with varying x (Fig. 6(b)) reveals that Sn alloying passivates the occ2 defect state by elevating the VBM, thereby pushing it out of the bandgap. Additionally, researchers specifically investigated the evolution of occ1 with x values. As shown in Fig. 6(e), the occ1 defect state induced by GeI in CsGe0.5Sn0.5I3 exhibits delocalized and shallow characteristics. Orbital hybridization analysis (Fig. 6(d)) reveals that the average bond length between GeD and neighboring GeNN/SnNN atoms increases at x = 0.5 compared to x = 0.25/0.75, indicating weakened orbital hybridization. Consequently, the occ1 defect state induced by GeI becomes energetically elevated and shallow in CsGe0.5Sn0.5I3. These findings collectively suggest that CsGe0.5Sn0.5I3 with benign defects (SnI and GeI) represents a promising candidate for high-performance photovoltaic applications.74 Wu et al. systematically investigated defect characteristics in the CsSn0.5Ge0.5I3 alloy system. Their calculations on all intrinsic defects in P1-phase CsSn0.5Ge0.5I3 revealed that only eight point defects (ICs, ISn, IGe, CsI, SnI, GeI, Sni, and Gei) induce deep transition levels, while others exhibit shallow-level behavior. Using the chemical potential diagrams in Fig. 6(c) and (f), they selected two representative growth conditions: point A (I-poor/Cs-rich) and point G (I-rich/Cs-poor), calculating formation energies for all point defects under these conditions. The results demonstrate that, under I-poor conditions, iodine vacancies (VI) exhibit the lowest formation energy across the Fermi level range, acting as dominant defects. The Fermi level becomes pinned by VI and VGe, enabling tunable conductivity from p-type to n-type. Under I-rich conditions, the formation energies of the dominant acceptor VCs and donor VI intersect at p-type Fermi levels. Compensation effects between V2+Cs and V−I charge defects pin the Fermi level, resulting in compensated p-type conductivity. Based on these computational and thermodynamic analyses, the research team proposed that CsSn0.5Ge0.5I3-P1 thin films should be synthesized under iodine-rich conditions to achieve optimal p-type conductivity and maximize photovoltaic performance, providing critical guidance for subsequent experimental fabrication.46 Tian et al. theoretically determined that Ce and Eu doping can substitute Ge to form stable CsGe1−xLnxBr3 structures using the plane-wave ultrasoft pseudopotential (PWUS) method with GGA and PW91 exchange–correlation functional. Furthermore, the bond lengths of Ge/Ln–Br and Cs–Br exhibit variations depending on the type and concentration of Ln, providing theoretical guidance for subsequent experimental synthesis of CsGeBr3:Ln3+ with enhanced luminescence properties. In addition, the research team conducted high throughput computing on the material systems. As revealed in Fig. 6(g)–(j), with increasing Eu3+ concentration, the luminous efficacy of radiation (LER) decreases, whereas both SLER and correlated color temperature (CCT) exhibit significant enhancement. Most importantly, through the combination of CsGeBr3:Ce and Eu3+, it is possible to obtain high-quality white light with a color rendering index (Ra) as high as 94 and a color temperature of 5608 K, and this result clarifies the performance targets for experimental construction of white light-emitting diode (WLED) devices.34 In current photovoltaics, a significant gap remains between the practical device performance of perovskite solar cells and their theoretical limits. Numerical simulation software serves as a powerful theoretical tool for efficient material screening, device structure optimization, and performance limit prediction, providing critical guidance for experimental research and development. Saikia et al. simulated CsGeI3 solar cells with a conventional n–i–p structure (FTO/TiO2/CsGeI3/CuI/Ag) using SCAPS-1D software. Through meticulous optimization of charge transport layer materials and thicknesses (TiO2: 10 nm and CuI: 30 nm), perovskite layer thickness (800 nm), doping concentrations (CsGeI3 acceptor doping: 1016 cm−3; ETL/HTL donor doping: 1019 cm−3), and defect densities (bulk/interface defects: 1015 cm−3), they achieved a simulated efficiency of 10.8%. This validates the feasibility of CsGeI3 as a light-absorbing layer and defines critical parameter ranges.75 Subsequently, to further break through efficiency bottlenecks and reduce costs, Zhang et al. innovatively proposed and simulated an HTL-free all-inorganic CsGeI3 cell structure (FTO/ETL/CsGeI3/metal back electrode). They discovered that when ZnOS is employed as the ETL, its unique conduction band depression structure effectively enhances electron extraction and suppresses interface recombination; simultaneously, selecting high-work-function tungsten (W) as the back electrode creates a back-contact barrier of about 0.3 eV at the CsGeI3/W interface, which effectively blocks electron backflow and promotes hole collection. This approach eliminates the need for expensive and unstable organic HTLs, establishing a new paradigm for developing low-cost, high-efficiency lead-free germanium-based perovskite solar cells.76 This demonstrates that, guided by software simulations, advanced device structure design and precise interface/bulk-phase modulation offer substantial room for enhancing the performance of all-inorganic germanium-based perovskite solar cells. Most recently, Das et al. employed SCAPS-1D simulations with CsGeI3 as the absorber layer, predicting a promising PCE of 21.6%.77 In a parallel advancement, Dash's team innovatively utilized RbGeI3 as the light-harvesting material in all-inorganic perovskite solar cells, achieving an optimized PCE of 25.76% with a fill factor (FF) of 79.81% through systematic device parameter refinement.78 Besides DFT calculations and software simulations, machine learning (ML), as a highly efficient data mining tool, has demonstrated considerable potential in the field of materials science in recent years. The traditional “trial-and-error” approach is not only highly resource-intensive but also struggles to extract key structure–property relationships from complex high-dimensional parameter spaces. In contrast, ML, by analyzing historical experimental data to build predictive models, can transcend the limitations of empirical knowledge and significantly accelerate the processes of novel material discovery and device optimization.79 Currently, ML is gaining widespread application in perovskite research, utilizing real experimental data to train ML models to guide the design of perovskite materials and the development of high-performance solar cells.80–88 In 2024, Gao et al. pioneered the construction of a machine learning model based on an artificial neural network (ANN) algorithm, predicting a wide bandgap (∼1.6 eV) tin-based perovskite material; combined with the optimized interfacial energy level alignment (near-zero energy level mismatch), this could enable a PCE exceeding 20%.89 Similarly, Kim et al. proposed a computational framework integrating DFT with ML; they employed a crystal graph convolutional neural network (CGCNN) to conduct high-throughput screening and computational analysis of 41,400 candidate compounds, identifying CsGe0.3125Sn0.6875I3 as the optimal absorber layer material for single-junction solar cells.90 Although ML research specifically targeting all-inorganic germanium-based perovskites remains in its infancy, it has already exhibited substantial application potential in predicting the properties and designing perovskite materials.
![]() | ||
Fig. 6 (a) Computed optical absorption coefficients of CsGeI3 under varying strains.60 Reproduced with permission. Copyright 2019, the Royal Society of Chemistry. (b) Band alignment diagrams of CsGexSn1−xI3 with different alloying fractions (x). (d) Local structure around GeD with partial charge density (displayed in the top circular region). Lower panel displays the summation of bond lengths between GeD and two nearest-neighbor GeNN/SnNN atoms. (e) Orbital-projected DOS and projected crystal orbital Hamiltonian population (COHP) for GeI in CsGe0.5Sn0.5I3.74 Reproduced with permission. Copyright 2021, the Royal Society of Chemistry. (c) Point A (I-poor) and (f) Point G (I-rich).46 Reproduced with permission. Copyright 2023, American Chemical Society. (g)–(j) High throughput computing of WLED performance.34 Reproduced with permission. Copyright 2025, the authors. |
C11 > 0, C44 > 0, C11 − C12 > 0, C11 + 2C12 > 0 | (3) |
A material is deemed mechanically stable when its elastic constants satisfy these criteria. From these three fundamental elastic constants, critical mechanical parameters can be derived, including the bulk modulus (B), shear modulus (G), and Young's modulus (E), all measured in GPa. The mathematical formulations are expressed as follows:
![]() | (4) |
The scientific community has established robust theoretical frameworks for distinguishing between brittle and ductile material behaviors. In 1954, Pugh proposed the ratio of the bulk modulus to the shear modulus (B/G) as a critical indicator for materials characterization. Materials demonstrating a Pugh ratio (B/G) exceeding 1.75 exhibit enhanced ductility, whereas ratios below this threshold signify brittle characteristics.92 Moreover, Frantsevich established Poisson's ratio (v) as an additional discriminative parameter. His research identified 0.26 as the critical Poisson's ratio: materials with v < 0.26 manifest brittleness, while those with v > 0.26 display ductile behavior.93 v is mathematically defined as follows:
![]() | (5) |
Furthermore, the Cauchy pressure (C12 − C44) has emerged as a third critical parameter in this analytical framework. Empirical evidence confirms that positive Cauchy pressure values correspond to ductile behavior, whereas negative values reliably predict brittle characteristics.64
Based on the aforementioned evaluation criteria, we extensively collected and systematically analyzed numerous academic papers on the elastic constant calculations of all-inorganic germanium-based perovskites, with the statistical results summarized in Table S2. Research by Hasan et al. demonstrates that FrGeI3 exhibits brittle characteristics, whereas FrGeBr3 resides in the critical state between brittleness and ductility.64 Among other investigated all-inorganic germanium-based perovskite materials, InGeCl3, InGeBr3,52 GaGeF3, InGeF3,61 KGeF3, NaGeF3,51 and Li2GeZ6 (Z = Cl, Br, and I)94 all demonstrate both mechanical stability and ductility. Evidently, the majority of studied all-inorganic germanium-based perovskites exhibit outstanding mechanical properties.
![]() | (6) |
This figure of merit reflects the material's electronic transport characteristics. In practical applications, achieving high-efficiency thermoelectric conversion necessitates materials with high Seebeck coefficient (S), high electrical conductivity (σ), and low thermal conductivity (Ke). Recent studies have demonstrated exceptionally high ZT values in selenides (e.g., SnSe, PbSe, and CuSe2)96–102 and tellurides (e.g., GeTe, SnTe, and PbTe).103–107 However, the scarcity of Se and Te elements renders these materials cost-prohibitive, thereby constraining their large-scale applications to some extent. Emerging research indicates that halide perovskites exhibit ultralow thermal conductivity108–110 and high charge carrier mobility.111,112 Concurrently, the relatively low cost of halide perovskite materials endows them with significant potential for thermoelectric conversion applications. Regarding all-inorganic germanium-based perovskites, they have already demonstrated promising prospects in photovoltaic applications. Building upon this foundation, numerous researchers have redirected their focus to investigating the thermoelectric properties of these materials.
Using the single-mode relaxation time approximation, Jong et al. solved the Boltzmann transport equation to calculate the constant-volume heat capacity, relaxation time, and group velocity, subsequently determining the temperature-dependent lattice thermal conductivity of CsGeX3 (X = I, Br, and Cl) through relevant theoretical formulations. As shown in Fig. 7(a), the thermal conductivity exhibits a gradual decrease with increasing temperature, while at fixed temperatures, it sequentially increases from X = I to Br to Cl. At 300 K, the thermal conductivity values of CsGeI3, CsGeBr3, and CsGeCl3 are 0.10, 0.16, and 0.18 W m−1 K−1, respectively, which are significantly lower than those of conventional thermoelectric materials such as GeTe103 and PbTe.113 Simultaneously, employing deformation potential theory (DPT), Jong et al. estimated charge carrier mobilities of 1,677, 1,401, and 863 cm2 V−1 s−1 for CsGeI3, CsGeBr3, and CsGeCl3, respectively, comparable to state-of-the-art thermoelectric materials. Remarkably, these values significantly surpass those of conventional perovskite counterparts, including FASnI3 (1.6 cm2 V−1 s−1),114 MASnI3 (1.6 cm2 V−1 s−1),115 MAPbBr3 (24 cm2 V−1 s−1), and MAPbI3 (67.2 cm2 V−1 s−1).116 The combined characteristics of ultralow thermal conductivity and high carrier mobility suggest promising potential for CsGeX3 in high-performance pyroelectric applications.117 Bouhmaidi et al. conducted an in-depth investigation on the thermoelectric performance of the CsGeX3 system by calculating the figure of merit ZT using the Quantum Espresso software package based on DFT. As shown in Fig. 7(b), the ZT value of CsGeCl3 remains nearly zero at low temperatures, gradually increases with increasing temperature, and stabilizes at 0.64 above 800 K. In contrast, the ZT value of CsGeBr3 starts at 0.99 under low-temperature conditions (T = 180 K), progressively decreases with increasing temperature, and drops to 0.1 at high temperatures (>900 K), indicating that its superior thermoelectric performance is limited to the low-temperature regime. Notably, the ZT value of CsGeF3 remains consistently above 0.9 throughout the investigated temperature range, demonstrating minimal temperature dependence. Collectively, these three materials exhibit significant potential for pyroelectric applications.57 Alburaih et al. investigated the thermoelectric properties of the double perovskite Li2GeZ6 (Z = Cl, Br, and I) through first-principles calculations based on DFT, combined with the semi-classical Boltzmann equation. Taking Li2GeCl6 as an example, the results shown in Fig. 7(c) and (d) reveal that its power factor (S2σ) increases with temperature, reaching a ZT value of 0.74 at 300 K, with an overall minimal variation in ZT across the temperature range. The same trend is observed for Li2GeBr6 and Li2GeI6, with ZT values approaching unity, suggesting the promising potential of this material series for pyroelectric device applications.94
![]() | ||
Fig. 7 (a) Calculated lattice thermal conductivity of CsGeX3 as functions of temperature (solid lines) and mean free path (dashed lines).117 Reproduced with permission. Copyright 2020, American Physical Society. (b) Temperature evolution of the figure of merit (ZT) for CsGeX3.57 Reproduced with permission. Copyright 2022, Elsevier. (c) Power factor and (d) ZT values of Li2GeCl6 as functions of temperature.94 Reproduced with permission. Copyright 2023, IOP Publishing. |
In all-inorganic systems, germanium-based perovskites have emerged as a key focus for ferroelectricity research due to the off-centering displacement of B-site ions. Using first-principles calculations with the Berry phase approach, Zhang et al. performed interpolation between the undistorted cubic structure and the polar rhombohedral structure to determine the spontaneous polarization (Ps) of CsGeBr3. The computational results are presented in Fig. 8(a). The results reveal a spontaneous polarization value of 19.7 μC cm−2 along the 〈111〉 direction in the distorted phase at room temperature. This value exceeds the estimated Ps (12.2 μC cm−2) derived from multiplying the Born effective charges (BECs) of the displaced ions by their displacements in the polar structure, indicating pronounced charge transfer between cations and anions – a hallmark of ferroelectric polarization. To elucidate the origin of ferroelectricity, the real-space charge density distribution in the polar R3m structure was calculated, revealing non-centrosymmetric charge density along the 〈111〉 direction that correlates with Ge-ion off-centering.122 This theoretical finding provides critical insights into the microscopic mechanism underlying ferroelectric behavior in CsGeBr3. Through DFT calculations, Chen et al. also theoretically investigated the ferroelectricity of CsGeI3. The calculations revealed a substantial ferroelectric polarization of about 15 μC cm−2 and an energy barrier of 0.067 eV (Fig. 8(b)), which are consistent with their experimental measurements. Regarding the piezoelectric response, the calculated piezoelectric modulus d33 of CsGeI3 reached 34.95 pC N−1 (Fig. 8(c)), comparable to conventional ferroelectrics. Notably, under 5% tensile strain, the d33 value was significantly enhanced to 190 pC N−1. Furthermore, as shown in Fig. 8(d), the applied strain notably increased both the polarization magnitude and the switching energy barrier, indicating effective stabilization of ferroelectricity.123 This work demonstrates that strain engineering serves as a promising strategy for tailoring the ferroelectric properties of all-inorganic germanium-based perovskites. Recently, Kashikar et al. developed an effective Hamiltonian parameterization strategy based on hybrid exchange–correlation functionals. This successfully addresses the underestimation of Curie temperature (TC) in halide perovskite ferroelectrics CsGeX3 (X = Cl, Br, and I) by conventional functionals. Theoretically, the experimental TC values of these materials are reproduced with high accuracy. Through molecular dynamics simulations, the finite-temperature properties of these materials are first predicted. These properties are confirmed to rival those of oxide ferroelectrics.124
![]() | ||
Fig. 8 (a) Berry phase calculations of polarization versus the distortion percentage for CsGeBr3, where 0% distortion corresponds to the cubic phase and 100% distortion to the ferroelectric rhombohedral phase.122 Reproduced with permission. Copyright 2022, American Association for the Advancement of Science. For CsGeI3, (b) ferroelectric polarization magnitude and switching energy barrier, (c) piezoelectric coefficients (d33) and lattice parameters under epitaxial strain ranging from −5% to 5%. (d) Spontaneous polarization and energy barrier evolution with epitaxial strain (−5% to 5%).123 Reproduced with permission. Copyright 2023, American Chemical Society. |
(1) Exploration of novel compounds. The AGeX3 perovskite structure exhibits high flexibility in elemental substitution, where A-site elements can be replaced with In, K, Na, Ga, etc., and X-site elements can be varied among different halogens. Furthermore, investigating special-structured perovskites such as 2D perovskites and double perovskites may lead to the discovery of all-inorganic germanium-based perovskites with superior performance.
(2) Composition engineering: Through theoretical studies on perovskite doping, exploring the variation patterns of material properties under different doping ratios will facilitate the screening of high-performance all-inorganic germanium-based perovskites, thereby achieving precise performance modulation.
(3) Defect investigation: Systematic studies on the defect characteristics of all-inorganic germanium-based perovskites should be conducted to identify materials with benign defects. Concurrent exploration of optimal environmental conditions for material synthesis will provide theoretical guidance for fabrication processes.
(4) Expansion of application fields: Continuous exploration of the latent properties of all-inorganic germanium-based perovskites should be pursued, with active investigation into their potential applications in ferroelectrics, thermoelectrics, and other emerging domains. This will broaden application boundaries and enhance the comprehensive value of these materials.
(5) Synergistic integration of multiple theoretical methodologies with experimental investigations. Beyond DFT calculations, advanced theoretical approaches such as software simulations and machine learning should be integrated to enable multifaceted screening and optimization of materials and devices, while GW approximation should be employed for the high-precision analysis of electronic properties. Concurrently, bridging theoretical predictions with experimental validation, particularly in domains like doping and strain engineering where experiments should be guided by theoretical frameworks will significantly enhance research accuracy and efficiency.
Supplementary information available: Table S1 lists bandgap values of all-inorganic Ge-based perovskites in Fig. 4 across computational methods. Table S2 provides elastic constants, critical mechanical parameters, Pugh ratio and Poisson's ratio for these compositions. See DOI: https://doi.org/10.1039/d5tc01917f
This journal is © The Royal Society of Chemistry 2025 |