Elsevier

Optical Materials

Volume 123, January 2022, 111839
Optical Materials

Performance evaluation of an all inorganic CsGeI3 based perovskite solar cell by numerical simulation

https://doi.org/10.1016/j.optmat.2021.111839Get rights and content

Highlights

  • SCAPS-1D simulation of lead-free perovskite CsGeI3-based solar cell was performed.

  • TiO2 and CuI were found to be the best material for ETL and HTL, respectively, to achieve optimal efficiency.

  • The optimized efficiency has been calculated to be 10.8%.

Abstract

The rapid advancement of organo-metal halide perovskite solar cells has led to a certified power conversion efficiency (PCE) of >25%. However, their ability to be implemented and commercialized on a large scale is currently limited due to the presence of toxic elements and the costly hole transport material (HTM) in the highly efficient cells. In the present work, a theoretical study on lead-free, environmental-friendly CsGeI3 based solar cell (SC) is demonstrated. Factors influencing the device parameters, such as thickness, doping, and defect concentration, have been thoroughly investigated to better understand material properties. With optimal material properties, the simulation results demonstrated 10.8% optimized PCE, Jsc of 22.08 mAcm−2, Voc of 0.667 V, and FF of 73.49%. These findings open up a new path for CsGeI3 as light-absorbing material to achieve clean, renewable energy.

Introduction

Perovskite solar cells (PSCs) have established themselves as viable candidates in the solar industry due to their easy preparation process and low cost compared to standard crystalline Si SC [1]. Since the first work on PSC in 2009, PCE is drastically enhanced from 3.8% [2] to more than 25% within a decade [3]. The high PCE is due to several distinguished properties of perovskites like low exciton binding energy, improved absorption properties, and considerable diffusion length of carriers [4,5]. However, instability causes by the volatile organic components in PSC acts as a barrier for its development [6,7]. Recently, inorganic halide perovskites (IHPs) have gained momentum due to their inherent stability and outstanding photovoltaic performance [8,9]. Among these IHPs, cesium lead halide based SCs exhibited high efficiency and stability [10,11]. However, toxicity arises from the heavy toxic element like Pb, which is hazardous to human health and the environment, hinders its commercialization [6,12,13]. Furthermore, the highly efficient CsPbI3 exhibits a non-perovskite yellow δ-phase at room temperature [14,15].

To address the issues of toxicity, lead-free perovskites have been explored by substituting Pb with similar kinds of divalent elements such as Sn [16], Ge [17], Ti [18], Bi [19], etc. In addition, several lead-free double perovskites have been explored [20,21]. Among these alternatives, Sn-based PSCs have demonstrated the highest efficiency, thereby having sparked so much interest in PSC. Furthermore, it was reported that CsSnI3 exhibited large hole mobility of 585 cm2 V−1s−1 compared to its Pb counterpart [22]. Very recently, the best efficiency of 10.1% was achieved from CsSnI3 PSCs [22]. However, the oxidation of Sn element from Sn2+ to Sn4+ under ambient atmosphere induces instability in cesium tin halide perovskites, resulting in degradation of PSCs [23]. In recent years, several theoretical and experimental works have been performed on lead-free Ge-based perovskites for photovoltaic applications. By alloying Sn and Ge, Chen et al. synthesized CsSn0.5Ge0.5I3-based PSCs that showed improved stability and air tolerance at ambient conditions. Because of the highly oxidized characteristics of Ge (II), a homogeneous native oxide surface was developed on CsSn0.5Ge0.5I3, demonstrating improved stability, more than that of CH3NH3PbI3 perovskite [24]. DFT-based results showed that CsGeI3 possesses a direct band gap of 1.63 eV, through strain engineering, its properties such as band gap could be tuned for suitable SC applications [25]. Computational results revealed that CsGeI3 possesses a smaller effective mass for the hole and can be employed as HTM. By the comprehensive theoretical and experimental study of CsGeI3, Krishnamoorthy et al. suggested that Ge is a possible alternative for Pb in halide perovskites.

Furthermore, it was reported that CsGeI3 exhibits a stable perovskite structure up to 350 °C and delivered a PCE of 0.11% [17]. In another work, 3.2% PCE, Jsc of 10.5 mAcm−2 and Voc of 0.57 V, was reported on CsGeI3-based PSCs. Recently, L.J. Chen synthesized CsGeI3 perovskite quantum rods using a solvothermal method and obtained 4.9% PCE by tuning the configuration of quantum rods [26]. In addition, a thermogravimetric study on CsGeI3 perovskite revealed that Ge-based perovskites are more thermally stable under device operating conditions [27]. These findings demonstrate Ge-based inorganic perovskites to be a promising candidate for lead-free PSCs.

Like perovskite materials, organic charge transport materials (CTMs), for example, spiro-OMeTAD, PEDOT:PSS and PCBM become unstable under illumination and ambient conditions [[28], [29], [30], [31]]. Furthermore, several organic CTMs react with perovskite, exhibit hostile behaviour because of their hygroscopic nature, and are expensive to prepare [32]. To overcome these drawbacks, inorganic CTMs, such as NiO [33], TiO2 [34,35], SnO2 [36,37] have been explored, which offer better transparency in UV, visible and infrared spectrums, a wide band gap, high carrier mobility, improved thermodynamic and chemical stability, and a simple synthesis method [38].

Here, we have presented a simulation study on lead-free CsGeI3 PSCs and show its potential application in photovoltaics. We have proposed an all-inorganic PSC by using inorganic CTMs, e.g., TiO2, SnO2, ZnO, CdS, WS2, CuI, Cu2O, CuSCN and NiO. By optimizing various material properties like thickness, doping concentration, defect density, and studying the resistances effect and working temperature, we achieved an optimal efficiency of 10.8%. In addition, we have investigated the impact of various materials as a metal electrode on device performance. Furthermore, the presented simulation findings could aid in the design and fabrication of future lead-free Ge-based PSCs.

The numerical simulation has been conducted by employing 1D-Solar Cell Capacitance Simulator (1D-SCAPS) software [39], under AM 1.5 G. Solving Poisson's equation (1), and continuity equation of electron (2) and hole (3), SCAPS can evaluate several electrical and optical properties like current-voltage characteristics, energy bands and quantum efficiency.x(ε(x)Vx)=q[p(x)n(x)+ND+(x)NA(x)+pt(x)nt(x)]nt=1qJnx+GnRnpt=1qJpx+GpRp

Here q,ε,V,p(x),n(x),ND+(x),NA(x),pt(x)andnt(x) represent electronic charge, dielectric permittivity, electric potential, free hole concentration, free electron concentration, ionized donor concentration, ionized acceptor concentration, trap density of holes and trap density of electrons, respectively. Furthermore, Jn, Jp, Gn, Gp, Rn and Rp denote current density, generation and recombination rate for electrons and holes, respectively.

The structure of the PSC employed in the initial simulation process is illustrated in Fig. 1(a), which comprises of FTO-coated glass, TiO2 for electron transport layer (ETL), CsGeI3 for light-absorbing material, spiro-OMeTAD layer for hole transport layer (HTL), and Ag for the electrode. The basic material properties extracted from various theoretical and experimental results are listed in Table 1. For all the materials, the thermal velocity of electrons and holes are taken as 1 × 107 m/s, while the absorption coefficient α has been calculated using the formula α=Aα(hνEg)2 with the value of Aα as 105 [40]. We have chosen neutral single distribution with 0.1 eV characteristic energy to introduce the defects in each device layer. Moreover, two interface layers, TiO2/CsGeI3 and CsGeI3/spiro-OMeTAD are included in the device structure to investigate the impact of interface recombination, thereby simulating a more realistic situation. The physical parameters of interface layers are described in Table S1. The total defect densities for both the interfaces are taken to be 1 × 1016 cm−3. The work functions of back and front contact are taken as 4.57 and 4.4 eV, respectively.

Using the initial parameters summarized in Table 1 and Table S1, we plotted the energy band diagram, current-voltage characteristics, and QE as depicted in Figs. 1(b) and figure 2 (a) and (b), respectively. The initial simulated device achieved 4.99% PCE, Jsc of 17.87 mAcm−2, Voc of 0.47 V, and FF of 59.68%, almost close to reported experimental findings as illustrated in Table 2. From Fig. 2, a good match between experimental results and our simulation was noticed, demonstrating validation of this work. However, the simulated FF was relatively higher than the experimental value, as we have neglected the series and shunt resistance in the simulation process [41].

Section snippets

Optimization of charge transport layers

The charge transport layers (CTLs), i.e., ETL and HTL of a PSC, influence the overall PCE by blocking the holes and electrons, respectively, to minimize carrier recombination and allow the superior path to reach at the contacts. To achieve improved device performance, the CTLs must meet several criteria, including a suitable band gap, high carrier mobility, optimal thickness, etc. Initially, we simulated the device performance using different inorganic HTLs (CuI, Cu2O, CuSCN and NiO) in the

Conclusion

In the present work, Pb-free, eco-friendly CsGeI3-based all-inorganic PSC was demonstrated through numerical simulation using SCAPS-1D software. By investigating the significance of various inorganic CTLs on photovoltaic performance and by optimizing different properties like thickness, doping concentration, defect density of absorber material and CTLs, a highly efficient n-i-p PSC was proposed. To reduce the device fabrication cost, carbon was suggested as the metal electrode. The proposed

CRediT authorship contribution statement

Dibyajyoti Saikia: Conceptualization, Methodology, Formal analysis, Investigation, Writing – original draft. Jayanta Bera: Writing, Editing. Atanu Betal: Writing, Editing. Satyajit Sahu: Writing the Final Draft, Editing, Investigation.

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Acknowledgement

This work was supported by the facility provided by the Indian Institute of Technology Jodhpur and the support from Ministry of Human Resource and Development, India.

References (70)

  • P. Zhao

    The crystal anisotropy effect of MAPbI3 perovskite on optoelectronic devices

    Mater. Today Energy

    (2020)
  • J. Wang

    An exploration of all-inorganic perovskite/gallium arsenide tandem solar cells

    Sol. RRL

    (2021)
  • R. Singh et al.

    Review of current progress in inorganic hole-transport materials for perovskite solar cells

    Appl. Mater. Today

    (2019)
  • M. Burgelman et al.

    Advanced electrical simulation of thin film solar cells

    Thin Solid Films

    (May 2013)
  • S. Ahmed et al.

    Numerical development of eco-friendly Cs2TiBr6 based perovskite solar cell with all-inorganic charge transport materials via SCAPS-1D

    Optik

    (2021)
  • Q. Duan

    Design of hole-transport-material free CH3NH3PbI3/CsSnI3 all-perovskite heterojunction efficient solar cells by device simulation

    Sol. Energy

    (2020)
  • M. Azadinia et al.

    Maximizing the performance of single and multijunction MA and lead-free perovskite solar cell

    Mater. Today Energy

    (2021)
  • A.A. Kanoun et al.

    Toward development of high-performance perovskite solar cells based on CH3NH3GeI3 using computational approach

    Sol. Energy

    (2019)
  • A.K. Singh et al.

    Performance optimization of lead free-MASnI3 based solar cell with 27% efficiency by numerical simulation

    Opt. Mater.

    (2021)
  • T. Minemoto et al.

    Theoretical analysis on effect of band offsets in perovskite solar cells

    Sol. Energy Mater. Sol. Cells

    (2015)
  • L. Lin

    Simulated development and optimized performance of CsPbI3 based all-inorganic perovskite solar cells

    Sol. Energy

    (2020)
  • S. Rai et al.

    Modeling of highly efficient and low cost CH3NH3Pb(I1-xClx)3 based perovskite solar cell by numerical simulation

    Opt. Mater.

    (2020)
  • W. Abdelaziz et al.

    Possible efficiency boosting of non-fullerene acceptor solar cell using device simulation

    Opt. Mater.

    (2019)
  • A. Kojima et al.

    Organometal halide perovskites as visible-light sensitizers for photovoltaic cells

    J. Am. Chem. Soc.

    (2009)
  • Best-Research-Cell-Efficiencies.20190923.pdf.”...
  • S.D. Stranks

    Electron-hole diffusion lengths exceeding 1 micrometer in an organometal trihalide perovskite absorber

    Science

    (2013)
  • H.S. Kim et al.

    Organolead halide perovskite: new horizons in solar cell research

    J. Phys. Chem. C

    (2014)
  • T. Leijtens et al.

    Stability of metal halide perovskite solar cells

    Adv. Energy Mater.

    (2015)
  • J. Liang

    All-inorganic perovskite solar cells

    J. Am. Chem. Soc.

    (2016)
  • T. Ma et al.

    The development of all-inorganic CsPbX3 perovskite solar cells

    J. Mater. Sci.

    (2020)
  • A. Babayigit et al.

    Toxicity of organometal halide perovskite solar cells

    Nat. Mater.

    (2016)
  • P. Su

    Pb-based perovskite solar cells and the underlying pollution behind clean energy: dynamic leaching of toxic substances from discarded perovskite solar cells

    J. Phys. Chem. Lett.

    (2020)
  • N. Wang

    Heterojunction-depleted lead-free perovskite solar cells with coarse-grained B-γ-CsSnI3 thin films

    Adv. Energy Mater.

    (2016)
  • T. Krishnamoorthy

    Lead-free germanium iodide perovskite materials for photovoltaic applications

    J. Mater. Chem.

    (2015)
  • M.G. Ju

    Earth-abundant nontoxic titanium(IV)-based vacancy-ordered double perovskite halides with tunable 1.0 to 1.8 eV bandgaps for photovoltaic applications

    ACS Energy Lett.

    (2018)
  • Cited by (76)

    View all citing articles on Scopus
    View full text