Next Article in Journal
Rheological, Thermal and Mechanical Characterization of Toughened Self-Healing Supramolecular Resins, Based on Hydrogen Bonding
Next Article in Special Issue
The Stability of Hybrid Perovskites with UiO-66 Metal–Organic Framework Additives with Heat, Light, and Humidity
Previous Article in Journal
A Comprehensive Study of Synthesis and Analysis of Anisotropic Iron Oxide and Oxyhydroxide Nanoparticles
Previous Article in Special Issue
Recent Advances in Nanostructured Inorganic Hole-Transporting Materials for Perovskite Solar Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Monovalent Copper Cation Doping Enables High-Performance CsPbIBr2-Based All-Inorganic Perovskite Solar Cells

1
State Key Laboratory of Materials-Oriented Chemical Engineering, College of Chemical Engineering, Nanjing Tech University, Nanjing 210009, China
2
WA School of Mines: Minerals, Energy and Chemical Engineering, Curtin University, Perth, WA 6845, Australia
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2022, 12(23), 4317; https://doi.org/10.3390/nano12234317
Submission received: 18 November 2022 / Revised: 2 December 2022 / Accepted: 3 December 2022 / Published: 5 December 2022
(This article belongs to the Special Issue Advanced Nanomaterials for Perovskite Solar Cells)

Abstract

:
Organic–inorganic perovskite solar cells (PSCs) have delivered the highest power conversion efficiency (PCE) of 25.7% currently, but they are unfortunately limited by several key issues, such as inferior humid and thermal stability, significantly retarding their widespread application. To tackle the instability issue, all-inorganic PSCs have attracted increasing interest due to superior structural, humid and high-temperature stability to their organic–inorganic counterparts. Nevertheless, all-inorganic PSCs with typical CsPbIBr2 perovskite as light absorbers suffer from much inferior PCEs to those of organic–inorganic PSCs. Functional doping is regarded as a simple and useful strategy to improve the PCEs of CsPbIBr2-based all-inorganic PSCs. Herein, we report a monovalent copper cation (Cu+)-doping strategy to boost the performance of CsPbIBr2-based PSCs by increasing the grain sizes and improving the CsPbIBr2 film quality, reducing the defect density, inhibiting the carrier recombination and constructing proper energy level alignment. Consequently, the device with optimized Cu+-doping concentration generates a much better PCE of 9.11% than the pristine cell (7.24%). Moreover, the Cu+ doping also remarkably enhances the humid and thermal durability of CsPbIBr2-based PSCs with suppressed hysteresis. The current study provides a simple and useful strategy to enhance the PCE and the durability of CsPbIBr2-based PSCs, which can promote the practical application of perovskite photovoltaics.

1. Introduction

The energy crisis and greenhouse gas emissions are the most crucial worldwide problems nowadays, which are mainly caused by the low efficiency and excessive consumption of nonrenewable fossil fuels [1,2]. Thus, the utilization of renewable energies and the development of relevant energy conversion technologies are highly urgent and essential [3,4,5,6,7]. Among various types of renewable energies/sources, solar energy has gained particular interest due to its inexhaustible and clean nature, which can be efficiently utilized by three main routes: photovoltaic cells, photocatalysis and solar-thermal power generation [1,8,9,10]. In particular, photovoltaic cells (solar cells) have received increasing attention and achieved significant progress in the past decades because of their direct and efficient transformation of sunlight energy into electricity [11,12,13,14,15]. Nowadays, the commercial photovoltaic markets are dominated by silicon-based solar cells, showing excellent stability and high PCE of 27.6% [16]. However, silicon-based solar cells are limited by high-cost and complex fabrication procedures, and toxic by-products in the silicon manufacturing, which are harmful to the environment [17,18]. Consequently, the development of new-type solar cells with simple, cost-effective and environmentally friendly fabrication processes is highly crucial [19,20].
Among various third-generation solar cells, perovskite solar cells (PSCs) with halide perovskites as light absorbers are regarded as more attractive than dye-sensitized and organic solar cells due to the cheap raw materials used, the simple and low-cost fabrication procedures, as well as their rapidly increasing power conversion efficiencies (PCEs) in the last 10 years [21,22]. The theoretical limit of the PCE has been known for PSCs and is approximately 30% [23]. Since the invention of PSCs by Miyasaka et al. in 2009, typical organic–inorganic PSCs have witnessed a rapid increase in their PCEs from only 3.8% to over 25% due to several unique advantages of organic–inorganic halide perovskites, such as proper and adjustable band gaps, large light absorption coefficient, long diffusion length and superior mobility of carriers [24,25,26,27]. The general formula of halide perovskites is ABX3, where A is monovalent cations including CH3NH3+ (MA+), CH3(NH2)2+ (FA+) and cesium (Cs+), B is divalent cations including Pb2+, Sn2+ and Ge2+, while X represents the halogen anions including I, Br and Cl [28,29,30]. It is well accepted that in addition to the PCEs, the long-term durability under humid, high-temperature and sunlight irradiation conditions is another key prerequisite for the widespread application of PSCs [31,32]. The plasmonic effect plays a very effective role in improving the efficiency of perovskite cells. By optimizing the adaptation of the plasmon photovoltaic channel, the relative efficiency was enhanced 40% [33]. Nevertheless, organic–inorganic hybrid PSCs suffer from inferior stability due to the high sensitivity of organic–inorganic perovskites against humidity, heat and light illumination, and they can be decomposed into PbI2 because of the inherent volatility and hygroscopicity of organic A-site cations (e.g., MA+, FA+) in the organic–inorganic perovskites [34,35,36,37].
All-inorganic halide perovskites are considered as attractive alternatives to their organic–inorganic counterparts to tackle the above-mentioned instability issue due to having superior intrinsic structural stability and much lower sensitivity against humidity and heat [38,39]. During the past seven years, CsPbIxBr3−x (0 ≤ x ≤ 3) as an important class of all-inorganic halide perovskites, has received particular attention as an efficient light absorber for PSCs due to the tunable band gaps, high intrinsic stability and superior carrier transport capability, etc. [40,41]. For instance, the PCEs of CsPbI3-based PSCs with a proper band gap of 1.73 eV have reached 21% nowadays. However, CsPbI3 perovskites suffer from the detrimental phase transformation from black CsPbI3 non-perovskite yellow phase with a much wider band gap (3.01 eV) at room temperature, exhibiting negative impacts on the PCEs and the stability of CsPbI3-based PSCs [42,43]. In contrast, CsPbBr3 perovskites have exhibited excellent phase structural stability and high heat/moisture tolerance, which suffered from the extremely large band gap (2.3 eV), remarkably limiting sunlight absorption range and then the cell performance [44,45]. On this basis, mixed halide all-inorganic CsPbIBr2 perovskite has attracted extensive interest due to the well-balanced band gap value (2.05 eV) and phase structural stability [46,47].
For single-junction PSCs, the calculated Shockley–Queisser limit is approximately 33%. Various approaches such as greatly increasing grain size, significantly reducing defect amount and band gaps, significantly improving carrier transport and separation capability, as well as constructing tandem PSCs, may allow approaching or surpassing the Shockley–Queisser limits of PSCs [3,48,49,50]. Nowadays, the PCEs of CsPbIBr2-based PSCs have exceeded 12%, which is much inferior to the organic–inorganic PSCs [51,52]. It has been reported that the limited PCEs of CsPbIBr2-based cells are mainly caused by the inferior quality CsPbIBr2 films fabricated by the conventional solution route, exhibiting numerous pinholes, grain boundaries and defects, which functioned as recombination sites for photoinduced carriers to suppress the cell performance [53,54]. Therefore, a variety of strategies have been devoted to achieving high-quality CsPbIBr2 light-absorbing films, such as interfacial modification, functional/selective cation doping, post-treatment process and additive engineering, etc. [55,56,57]. Among these available strategies, cation doping/substitution is widely employed to tailor the optical/electronic properties and film quality of CsPbIBr2, which plays significant roles in creating homogeneous nucleation sites to control the crystallization and growth of CsPbIBr2 crystals, contributing to remarkably enhanced perovskite film quality [58,59,60]. Nowadays, the cation doping in CsPbIBr2 is mainly focused on the utilization of divalent metal cations to substitute Pb2+, including Sn2+, Cu2+, Zn2+, Ba2+, Eu2+, etc., which effectively enhance the crystallinity and morphology of CsPbIBr2 films [61,62,63,64,65,66]. For instance, Sn2+ plays a crucial role in regulating the band gap and enhancing the film quality of CsPbIBr2, and Zhao et al. have reported that remarkably enhanced PCE (11.33%) and long-term stability were achieved by Sn2+ doping in CsPbIBr2-based PSCs [61,64]. In our previous work, we have found that Cu2+ doping in CsPbIBr2 with optimized doping concentration effectively inhibited the carrier recombination by reducing the defect amount, and it enhanced the perovskite film quality by improving the crystallinity and enlarging the grain sizes, thereby remarkably improving the PCEs and humid/thermal stability of CsPbIBr2-based PSCs [62].
Nevertheless, it should be noted that the monovalent cation doping in the B-site of CsPbIBr2 is much less investigated, which needs further exploration. Herein, we have employed monovalent copper cations (Cu+) as effective B-site dopants for CsPbIBr2 perovskites, which effectively improved the CsPbIBr2 film quality in terms of a suppressed amount of grain boundaries, increased crystallinity and grain sizes, passivated surface defects and inhibited interfacial carrier recombination. After optimizing the Cu+ doping concentration, the CsPbIBr2-0.50%Cu cell produced a superb PCE of 9.11%, 25.8% higher than that of the pristine cell. Furthermore, Cu+ doping also remarkably improved the high-temperature and humid durability of CsPbIBr2-based PSCs with reduced hysteresis effect. Our current work can present some important insights for the design of high-performance all-inorganic PSCs, which may promote the widespread application of perovskite photovoltaics.

2. Materials and Methods

F-doped tin oxide (FTO) glasses (2.2-mm thick and 7 Ω sq−1) were first cleaned by various solvents, dried by N2 and further treated by oxygen plasma cleaning [56]. Detailed information about the fabrication of CsPbIBr2-based PSCs with a configuration of compact TiO2 (c-TiO2)/perovskite/2,2′,7,7′-tetrakis (N,N-di-p-methoxyphenylamine)-9,9′-spirobifluorene (Spiro-OMeTAD)/Ag can be found in our previous work [62]. As for the Cu+-doped CsPbIBr2-based cells, different amounts of as-prepared 0.2 M cuprous bromide/dimethyl sulfoxide (CuBr/DMSO) solution were added into the 1 M CsPbIBr2 precursor solution to prepare various Cu+-doped CsPbIBr2 films at fixed molar ratios of 0.25, 0.50 and 0.75%. The effective area of the cell was 0.0625 cm2 in this work. For the thermal stability test, the carbon electrode was deposited by doctor blade for the hole-transporting layer (HTL)-free CsPbIBr2-based PSCs and the detailed information can be found in our previous work [62].
X-ray diffraction (XRD) patterns of various samples were acquired by X-ray diffractometer (Rigaku Smartlab, Matsubara-cho, Tokyo, Japan) with Cu radiation. X-ray/ultraviolet photoelectron spectroscopy (XPS/UPS) profiles of various samples were acquired by XPS spectrometer (PHI5600 Versa Probe, Brooklyn, NY, USA). Energy-dispersive X-ray spectroscopy (EDX, Octane Ultra, Lafayette, LA, USA) and scanning electron microscope (SEM, Hitachi S-4800, Chiyoda-ku, Tokyo, Japan) were employed to investigate the microstructures and elemental mapping of various films. The surface roughness values of various perovskite films were acquired by atomic force microscopy (AFM, SmartSPM, Horiba, Minami-ku, Tokyo, Japan). Ultraviolet–visible (UV-vis) and photoluminescence (PL) spectra were obtained by Lambda 750 s spectrometer (PerkinElmer, Waltham, MA, USA) and fluorescence spectrometer (PerkinElmer, FL 6500, Waltham, MA, USA), respectively. Photocurrent density-voltage (J−V) curves of various cells were acquired by Zolix solar simulator under 1 Sun irradiation (100 mW cm−2) after calibration. Electrochemical impedance spectroscopy (EIS) and external quantum efficiency (EQE) and spectra of various cells were acquired by electrochemical workstation (CHI760E) and Zolix Solar Cell Scan 100 instrument, respectively.

3. Results and Discussion

Cu+ cations with three different molar ratios of 0.25, 0.50 and 0.75% were doped into CsPbIBr2, which were labeled as CsPbIBr2, CsPbIBr2-0.25%Cu, CsPbIBr2-0.50%Cu and CsPbIBr2-0.75%Cu, respectively. Figure S1 displays the typical cross-sectional SEM image of as-fabricated CsPbIBr2-0.50%Cu cell and it was found that the 350 nm-thick perovskite film was firmly adhered on the surface of c-TiO2 film. Based on the J−V curves of various CsPbIBr2 cells, as depicted in Figure 1a and relevant photovoltaic parameters in Figure S2, the Cu+ doping in CsPbIBr2 significantly boosted the PCEs of corresponding cells by increasing the fill factor (FF) and short-circuit current density (Jsc). More specifically, the unmodified cell displayed a champion PCE of 7.24% with an open-circuit voltage (Voc) of 1.16 V, a Jsc of 10.4 mA cm−2 and an FF of 0.597. In addition, the cell performance of Cu+-doped CsPbIBr2-based PSCs exhibited a volcano-like trend with increased Cu+ doping contents, suggesting that the introduction of a suitable concentration of Cu+ cations significantly enhanced the PCEs of CsPbIBr2 cells. Particularly, the CsPbIBr2-0.50%Cu cell displayed the highest PCE of 9.11%, with a Voc of 1.19 V, a Jsc of 11.6 mA cm−2 and a FF of 0.658 (Table S1). The CsPbIBr2-0.75%Cu cell exhibited a reduced PCE, which was mainly attributed to decreased FF value induced by the poor quality of CsPbIBr2 film and higher defect concentration, which will be discussed later. Based on the EQE spectra in Figure 1b, the CsPbIBr2-0.50%Cu cell displayed higher EQE values than those of the pristine cell at a wavelength range of 300 to 600 nm. In addition, CsPbIBr2-0.50%Cu cell generated a larger integrated Jsc value of 9.7 mA cm−2 than that of the pristine cell (8.1 mA cm−2), which was consistent with the J−V results (Figure 1a). To confirm the universality of such PCE enhancement, 20 devices for each type of PSC (without and with 0.50%Cu+ substitution) were fabricated and tested with the photovoltaic parameter distributions displayed in Figure 1c. It is clear that the average PCE of the CsPbIBr2-0.50%Cu cell was much larger than that of the unmodified device due to the significantly improved Jsc and FF values (Figure S2). In addition, based on the maximum power point tracking (MPPT) profiles of PSCs as depicted in Figure 1d, the CsPbIBr2-0.50%Cu cell produced much larger stabilized PCE and Jsc values than those of the pristine CsPbIBr2 cell at the maximum power point.
It is well accepted that the PCEs of PSCs are determined by several crucial factors including the transport and recombination behavior of photogenerated carriers, defect density, sunlight absorption capability and energy level alignment between various films [67,68]. In order to elucidate the impacts of Cu+ substitution on the CsPbIBr2 film quality and the performance of CsPbIBr2-based PSCs, we employed XRD technique to investigate the crystallinity and the crystal structures of pristine CsPbIBr2 and various Cu+-doped CsPbIBr2 films. As depicted in Figure 2a, main characteristic peaks assigned to the α-phase CsPbIBr2 structure were observed in all samples, while no obvious impurity phases were found for all investigated films, revealing that Cu+ doping at different concentrations exhibited no obvious influences on the pure-phase cubic structure of CsPbIBr2. Based on the magnified XRD peaks in Figure 2b, the characteristic XRD peaks gradually shifted to higher angles with the increased Cu+ doping amounts, implying that the Cu+ cations with smaller ionic radius (0.60 Å) than that of Pb2+ (1.19 Å) were successfully doped in the lattice of CsPbIBr2 with a lattice contraction [69,70]. UV-vis absorption spectra were employed to evaluate the impacts of Cu+ substitution on the light absorption capability of CsPbIBr2 film. As depicted in Figure 2c, the Cu+ doping effectively increased the light absorption intensity of CsPbIBr2 film at 450–600 nm, especially for the CsPbIBr2-0.50%Cu film. Based on the Tauc plots of various films acquired from the UV-vis spectra in Figure S3, the band gap values of CsPbIBr2, CsPbIBr2-0.25%Cu, CsPbIBr2-0.50%Cu and CsPbIBr2-0.75%Cu were calculated to be 2.08, 2.07, 2.07 and 2.06 eV, respectively. It can be concluded that the Cu+ doping can improve the light absorption capability of CsPbIBr2 film, although the Cu+ doping displayed no obvious influences on the band gap values of CsPbIBr2, which may be beneficial for the PCE improvement of corresponding PSCs. Based on the steady-state PL spectra of various CsPbIBr2 films as depicted in Figure 2d, the Cu+-doped CsPbIBr2 films exhibited much lower PL intensity than the CsPbIBr2 film, demonstrating that the carrier recombination was effectively suppressed after introducing Cu+ cations to reduce the defect amount. Moreover, CsPbIBr2-0.75%Cu film with excessive Cu+ doping amount displayed a higher PL intensity than that of CsPbIBr2-0.50%Cu, implying more defects were formed in the CsPbIBr2-0.75%Cu film, agreeing well with the J−V results. It should be noted that the PL spectra of Cu+-doped CsPbIBr2 films exhibited several split peaks due to the phase segregation of CsPbIBr2 films, which may be beneficial for the PCE enhancement [62,71]. Based on the TRPL spectra as depicted in Figure S4, CsPbIBr2-0.50%Cu film delivered a much higher average carrier lifetime of 3.21 ns than the CsPbIBr2 film (0.88 ns) due to the effectively suppressed defect concentration, which was favorable for the transport and separation of carriers. XPS was further employed to explore the influences of Cu+ substitution on the chemical states of various ions in CsPbIBr2 films, with results displayed in Figure S5. All the XPS peaks of CsPbIBr2-0.50%Cu film shifted to lower binding energies compared with the CsPbIBr2 film due to the shrinkage of the BX6 octahedron caused by Cu+ doping induced by the changed interatomic force [58,72].
The morphology and quality of perovskite film plays a vital role in governing the performance of PSCs [54,73]. More specifically, compact perovskite films with large grain sizes and few grain boundaries are beneficial to suppress the carrier recombination [13,74]. Based on the top-view SEM images of CsPbIBr2 and various Cu+-doped CsPbIBr2 films in Figure 3a–d and Figure S6, the pristine CsPbIBr2 film displayed an inferior morphology with abundant pinholes and small grains, while the quality of CsPbIBr2 film was significantly improved after Cu+ doping in terms of reduced amount of pinholes and larger grain sizes. Particularly, the CsPbIBr2-0.50%Cu film showed a dense and uniform morphology with remarkably enlarged grain sizes and reduced amount of grain boundaries (Figure 3c,d). Nevertheless, CsPbIBr2-0.75%Cu film with excessive Cu+ doping amount exhibited smaller grain sizes than those of CsPbIBr2-0.50%Cu film, as displayed in Figure S6c,d. In addition, it was found that all elements were uniformly distributed on the CsPbIBr2-0.50%Cu film, as depicted in Figure 3e, demonstrating homogeneous distribution of Cu+ dopants. Based on the AFM images as depicted in Figure 3f,g and Figure S7, the CsPbIBr2-0.50%Cu film delivered a lower root-mean-square (RMS) surface roughness of 29.9 nm than the CsPbIBr2 film (31.3 nm), benefiting the interfacial charge transfer at perovskite film/HTL interface [75,76].
The carrier recombination behavior of CsPbIBr2 film after Cu+ doping was investigated by EIS at a fixed voltage of 0.5 V under dark conditions, as depicted in Figure 4a. As can be seen, a low-frequency arc corresponding to the recombination resistance (Rrec) existed in the Nyquist plots, which was inversely proportional to the degree of carrier recombination at the interfaces between the perovskite films and TiO2/Spiro-OMeTAD layer [77,78]. The CsPbIBr2-0.50%Cu cell generated a larger Rrec value than the CsPbIBr2 cell, suggesting that the Cu+ doping remarkably suppressed the interfacial carrier recombination. Based on the dark J−V curves as displayed in Figure 4b, the CsPbIBr2-0.50%Cu cell delivered much smaller current densities than those of the pristine CsPbIBr2 cell, demonstrating that the photo-generated carriers were efficiently transported through charge-transporting layers of the CsPbIBr2-0.50%Cu cell instead of direct shunting, leading to reduced voltage/current loss and enhanced cell performance, which was attributed to the improved morphology and quality of CsPbIBr2-0.50%Cu films.
Hole-only and electron-only PSCs with structures of FTO/PEDOT:PSS/pervoskite/Spiro-OMeTAD/Ag and FTO/TiO2/pervoskite/[6,6]-phenyl-C61-butyric acid methyl ester (PCBM)/Ag were prepared to determine the trap densities of PSCs by measuring the dark J−V curves using the space charge-limiting current model [13,79]. As shown in Figure 4c,d, the traps were gradually filled until the applied voltage reached the trap-filling-limit voltage (VTFL). The trap densities (Ndefects) can be estimated by the equation of Ndefects = 2 ε ε 0 V T F L e L 2 , where ε is the relative dielectric constant of CsPbIBr2 (approximately 8); ε0 is the vacuum dielectric constant; e represents elementary charge; and L is the thickness of CsPbIBr2 film (350 nm, Figure S1). The VTFL values were 0.43 and 0.38 V for electron-only devices with CsPbIBr2 and CsPbIBr2-0.50%Cu films, corresponding to trap densities of 3.10 and 2.75 × 1015 cm−3, respectively. As for hole-only devices, the hole defect densities of CsPbIBr2 and CsPbIBr2-0.50%Cu films were 3.90 and 1.59 × 1015 cm−3 based on VTFL values of 0.54 and 0.22 V, respectively. This suggested that the Cu+ doping effectively reduced the trapping centers for both electrons and holes of CsPbIBr2 film.
UPS technique was used to explore the effects of Cu+ substitution on the energy level alignment of CsPbIBr2-based PSCs. As displayed in Figure S8 and Table S2, the valence band maximum (EVBM) was calculated as −5.58 and −5.52 eV for CsPbIBr2 and CsPbIBr2-0.50%Cu, respectively, while the corresponding conduction band minimum (ECBM) was calculated to be −3.50 and −3.45 eV, respectively, based on the relationship of band gap = ECBMEVBM [80]. In addition, the Fermi level of CsPbIBr2 was slightly increased from −3.74 to −3.68 eV after the introduction of 0.50% Cu+, which was closer to the CBM position (−3.51 eV) due to the n-type nature of Cu+-doped CsPbIBr2 [81]. The upshifted VBM position of CsPbIBr2-0.50%Cu effectively promoted the hole extraction from Spiro-OMeTAD to the perovskite layer. Moreover, the larger energy differences between the CBMs of CsPbIBr2-0.50%Cu and TiO2 may provide a higher driving force for the electron injection from light-absorbing film to the electron-transporting layer [72]. Furthermore, the hysteresis index of the CsPbIBr2 cell was reduced from 0.48 to 0.43 after the introduction of 0.50%Cu based on the J−V curves of corresponding PSCs under reverse and forward scan directions (Figure S9 and Table S3), which was attributed to the improved CsPbIBr2 film quality and promoted charge transfer, benefiting the cell stability.
Besides the device efficiency, the humid and thermal stability is another crucial factor for the development of PSCs. As depicted in Figure 5a, the CsPbIBr2-0.50%Cu cell retained 94% of its primary PCE after storing in humid air with a relative humidity (RH) of 15–30% at 25 °C for 400 h, much superior to the pristine device (57%) under the same conditions. Furthermore, the PCE of the CsPbIBr2-0.50%Cu cell maintained 80% after 600 h storage in ambient condition with a RH of 15–30%. The high-temperature air stability of PSCs was evaluated by preparing HTL-free cells (FTO/c-TiO2/perovskite/carbon). The photovoltaic parameters of HTL-free CsPbIBr2 and CsPbIBr2-0.50%Cu cells are listed in Table S4. It was found that the CsPbIBr2-0.50%Cu cell delivered a superior PCE retention ratio of 97% to that of the CsPbIBr2 device (76%) after storing in ambient condition at 85 °C for 1000 h (Figure 5b).
Cu+ involved in the lattice of CsPbIBr2 partially substituted the Pb2+-occupied B-site, causing a lattice contraction, and resulted in a reduced bandgap for Cu+-doped CsPbIBr2 film. The reduced bandgap is beneficial for broadening light absorption band edge, thus facilitating the increase of photocurrent density. A previous study suggested that lattice contraction caused by doping could lead to improved heat stability for PSCs [82]. The presence of grain boundaries and pinholes might act as defect centers, trapping charge carriers, thereby reducing the PCE of PSC [83] Homogeneous and compact perovskite morphology with large grain sizes, less grain boundaries and pinholes enhanced the light capture of Cu+-doped perovskite film. The chemical bonds were enhanced after Cu+ doping, leading to higher phase stability [84]. Moreover, the defect density and the nonradiative recombination of the films were reduced.

4. Conclusions

In summary, we report a monovalent Cu+ cation doping approach to boost the PCEs and humid/thermal durability of CsPbIBr2-based all-inorganic PSCs. It was found that the introduction of a proper amount of Cu+ cations into CsPbIBr2 effectively passivated the defects, improved the perovskite film quality, suppressed the interfacial carrier recombination and enhanced the energy level alignment. Consequently, the optimized Cu+-doped cell exhibited a superb PCE of 9.11% with reduced hysteresis, which was 25.8% higher than that of pristine cell (7.24%). Moreover, the Cu+ doping also remarkably improved the thermal and humid stability of CsPbIBr2-based PSCs. For instance, the CsPbIBr2-0.50%Cu device retained 97% of the primary PCE after 1000 h storage at 85 °C in humid air, while the PCE of the pristine PSC declined to 76% under the same conditions. This work provides some important insights for the fabrication of durable CsPbIBr2-based all-inorganic PSCs with higher PCEs, which may promote the commercialization of this technology.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano12234317/s1, Figure S1: Cross-sectional SEM images of CsPbIBr2-based cell; Figure S2: (a) PCE, (b) Jsc, (c) Voc and (d) FF distributions of CsPbIBr2-based PSCs without and with 0.50% Cu+ doping with the corresponding average values shown inside; Figure S3: Tauc plots of CsPbIBr2 films with different Cu+ doping concentrations; Figure S4: TRPL spectra of CsPbIBr2 and CsPbIBr2-0.50%Cu films; Figure S5: XPS spectra of the pristine CsPbIBr2 and CsPbIBr2-0.50%Cu films: (a) Cs 3d, (b) Pb 4f, (c) I 3d, (d) Br 3d and (e) XPS survey spectra; Figure S6: Top-view SEM images of the (a, b) CsPbIBr2-0.25%Cu and (c, d) CsPbIBr2-0.75%Cu films; Figure S7: 3D-AFM models of (a) pristine CsPbIBr2 and (b) CsPbIBr2-0.50%Cu films; Figure S8: UPS spectra of pristine CsPbIBr2 and CsPbIBr2-0.50%Cu films: (a) the cut-off energies (Ecut-off) and (b) the positions of Fermi edge; Figure S9: JV curves of (a) pristine CsPbIBr2 and (b) CsPbIBr2-0.50%Cu cells under revere and forward scan directions; Table S1: Photovoltaic performance parameters of CsPbIBr2-based PSCs with different Cu+ doping concentrations; Table S2: Calculated energy levels of pristine CsPbIBr2 and CsPbIBr2-0.50%Cu films; Table S3: Photovoltaic parameters of pristine CsPbIBr2 and CsPbIBr2-0.50%Cu cells under revere and forward scan directions; Table S4: Photovoltaic parameters of carbon-based HTL-free PSCs based on CsPbIBr2 and CsPbIBr2-0.50%Cu perovskites.

Author Contributions

Investigation, Z.D. and H.X.; conceptualization, W.W.; formal analysis, Z.D. and H.X.; visualization, Z.D.; writing—original draft, Z.D.; validation, Z.D.; data curation, Z.D., H.X. and A.X.; supervision, W.W. and Z.S.; writing—review and editing, R.R., W.W. and Z.S.; project administration, W.Z.; funding acquisition, W.Z., W.W. and Z.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the National Natural Science Foundation of China (No. 22279057).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lai, C.S.; Jia, Y.; Lai, L.L.; Xu, Z.; McCulloch, M.D.; Wong, K.P. A comprehensive review on large-scale photovoltaic system with applications of electrical energy storage. Renew. Sustain. Energy Rev. 2017, 78, 439–451. [Google Scholar] [CrossRef]
  2. Zou, H.; Du, H.; Ren, J.; Sovacool, B.K.; Zhang, Y.; Mao, G. Market dynamics, innovation, and transition in China’s solar photovoltaic (pv) industry: A critical review. Renew. Sustain. Energy Rev. 2017, 69, 197–206. [Google Scholar] [CrossRef] [Green Version]
  3. Jacak, J.E.; Jacak, W.A. Routes for metallization of perovskite solar cells. Materials 2022, 15, 2254. [Google Scholar] [CrossRef]
  4. Mitrašinović, A.M. Photovoltaics advancements for transition from renewable to clean energy. Energy 2021, 237, 121510. [Google Scholar] [CrossRef]
  5. Carneiro, A.L.; Martins, A.A.; Duarte, V.C.M.; Mata, T.M.; Andrade, L. Energy consumption and carbon footprint of perovskite solar cells. Energy Rep. 2022, 8, 475–481. [Google Scholar] [CrossRef]
  6. Qiao, Y.; Li, S.; Liu, W.; Ran, M.; Lu, H.; Yang, Y. Recent advances of rare-earth ion doped luminescent nanomaterials in perovskite solar cells. Nanomaterials 2018, 8, 43. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Yang, X.; Wang, W.; Ran, R.; Zhou, W.; Shao, Z. Recent advances in Cs2AgBiBr6-based halide double perovskites as lead-free and inorganic light absorbers for perovskite solar cells. Energy Fuels 2020, 34, 10513–10528. [Google Scholar] [CrossRef]
  8. Wang, W.; Xu, M.; Xu, X.; Zhou, W.; Shao, Z. Perovskite oxide based electrodes for high-performance photoelectrochemical water splitting. Angew. Chem. Int. Ed. 2020, 59, 136–152. [Google Scholar] [CrossRef]
  9. He, J.; Liu, P.; Ran, R.; Wang, W.; Zhou, W.; Shao, Z. Single-atom catalysts for high-efficiency photocatalytic and photoelectrochemical water splitting: Distinctive roles, unique fabrication methods and specific design strategies. J. Mater. Chem. A 2022, 10, 6835–6871. [Google Scholar] [CrossRef]
  10. Han, X.; Liu, P.; Ran, R.; Wang, W.; Zhou, W.; Shao, Z. Non-metal fluorine doping in Ruddlesden-Popper perovskite oxide enables high-efficiency photocatalytic water splitting for hydrogen production. Mater. Today Energy 2022, 23, 100896. [Google Scholar] [CrossRef]
  11. Muradov, A.; Frolushkina, D.; Samusenkov, V.; Zhamanbayeva, G.; Kot, S. Methods of stability control of perovskite solar cells for high efficiency. Energies 2021, 14, 2918. [Google Scholar] [CrossRef]
  12. Park, H.H. Inorganic materials by atomic layer deposition for perovskite solar cells. Nanomaterials 2021, 11, 88. [Google Scholar] [CrossRef] [PubMed]
  13. Liu, P.; Xiang, H.; Wang, W.; Ran, R.; Zhou, W.; Shao, Z. A bilateral cyano molecule serving as an effective additive enables high-efficiency and stable perovskite solar cells. J. Energy Chem. 2021, 62, 243–251. [Google Scholar] [CrossRef]
  14. Fang, Z.; Zeng, Q.; Zuo, C.; Zhang, L.; Xiao, H.; Cheng, M.; Hao, F.; Bao, Q.; Zhang, L.; Yuan, Y.; et al. Perovskite-based tandem solar cells. Chin. Sci. Bull. 2021, 66, 621–636. [Google Scholar] [CrossRef]
  15. Xiao, H.; Liu, P.; Wang, W.; Ran, R.; Zhou, W.; Shao, Z. Enhancing the photocatalytic activity of Ruddlesden-Popper Sr2TiO4 for hydrogen evolution through synergistic silver doping and moderate reducing pretreatment. Mater. Today Energy 2022, 23, 100899. [Google Scholar] [CrossRef]
  16. Chen, B.; Wang, P.; Li, R.; Ren, N.; Han, W.; Zhu, Z.; Wang, J.; Wang, S.; Shi, B.; Liu, J.; et al. A two-step solution-processed wide-bandgap perovskite for monolithic silicon-based tandem solar cells with >27% efficiency. ACS Energy Lett. 2022, 7, 2771–2780. [Google Scholar] [CrossRef]
  17. Xiang, H.; Liu, P.; Wang, W.; Ran, R.; Zhou, W.; Shao, Z. Towards highly stable and efficient planar perovskite solar cells: Materials development, defect control and interfacial engineering. Chem. Eng. J. 2021, 420, 127599. [Google Scholar] [CrossRef]
  18. Ansari, M.I.H.; Qurashi, A.; Nazeeruddin, M.K. Frontiers, opportunities, and challenges in perovskite solar cells: A critical review. J. Photochem. Photobiol. C 2018, 35, 1–24. [Google Scholar] [CrossRef]
  19. Huang, D.; Xiang, H.; Ran, R.; Wang, W.; Zhou, W.; Shao, Z. Recent advances in nanostructured inorganic hole-transporting materials for perovskite solar cells. Nanomaterials 2022, 12, 2592. [Google Scholar] [CrossRef]
  20. Liu, P.; Wang, W.; Liu, S.; Yang, H.; Shao, Z. Fundamental understanding of photocurrent hysteresis in perovskite solar cells. Adv. Energy Mater. 2019, 9, 1803017. [Google Scholar] [CrossRef]
  21. Shahinuzzaman, M.; Afroz, S.; Mohafez, H.; Jamal, M.S.; Khandaker, M.U.; Sulieman, A.; Tamam, N.; Islam, M.A. Roles of inorganic oxide based htms towards highly efficient and long-term stable PSC: A review. Nanomaterials 2022, 12, 3003. [Google Scholar] [CrossRef] [PubMed]
  22. Wang, P.; Li, R.; Chen, B.; Hou, F.; Zhang, J.; Zhao, Y.; Zhang, X. Gradient energy alignment engineering for planar perovskite solar cells with efficiency over 23%. Adv. Mater. 2020, 32, 1905766. [Google Scholar] [CrossRef] [PubMed]
  23. Seki, K.; Furube, A.; Yoshida, Y. Theoretical limit of power conversion efficiency for organic and hybrid halide perovskite photovoltaics. Jpn. J. Appl. Phys. 2015, 54, 08KF04. [Google Scholar] [CrossRef] [Green Version]
  24. Kojima, A.; Teshima, K.; Shirai, Y.; Miyasaka, T. Organometal halide perovskites as visible-light sensitizers for photovoltaic cells. J. Am. Chem. Soc. 2009, 131, 6050–6051. [Google Scholar] [CrossRef]
  25. NREL. Best Research-Cell Efficiency Chart. Available online: https://www.nrel.gov/pv/cell-efficiency.html (accessed on 13 November 2022).
  26. Xiang, H.; Liu, P.; Ran, R.; Wang, W.; Zhou, W.; Shao, Z. Two-dimensional Dion-Jacobson halide perovskites as new-generation light absorbers for perovskite solar cells. Renew. Sust. Energy Rev. 2022, 166, 112614. [Google Scholar] [CrossRef]
  27. Yang, X.; Xie, A.; Xiang, H.; Wang, W.; Ran, R.; Zhou, W.; Shao, Z. First investigation of additive engineering for highly efficient Cs2AgBiBr6-based lead-free inorganic perovskite solar cells. Appl. Phys. Rev. 2021, 8, 041402. [Google Scholar] [CrossRef]
  28. Liu, H.; Xiang, L.; Gao, P.; Wang, D.; Yang, J.; Chen, X.; Li, S.; Shi, Y.; Gao, F.; Zhang, Y. Improvement strategies for stability and efficiency of perovskite solar cells. Nanomaterials 2022, 12, 3295. [Google Scholar] [CrossRef]
  29. Mangrulkar, M.; Stevenson, K.J. The progress of additive engineering for CH3NH3PbI3 photo-active layer in the context of perovskite solar cells. Crystals 2021, 11, 814. [Google Scholar] [CrossRef]
  30. Shen, W.; Dong, Y.; Huang, F.; Cheng, Y.; Zhong, J. Interface passivation engineering for hybrid perovskite solar cells. Mater. Rep. Energy 2021, 1, 100060. [Google Scholar] [CrossRef]
  31. Sathiyan, G.; Syed, A.A.; Chen, C.; Wu, C.; Tao, L.; Ding, X.; Miao, Y.; Li, G.; Cheng, M.; Ding, L. Dual effective dopant based hole transport layer for stable and efficient perovskite solar cells. Nano Energy 2020, 72, 104673. [Google Scholar] [CrossRef]
  32. Zhang, J.; Che, B.; Zhao, W.; Fang, Y.; Han, R.; Yang, Y.; Liu, J.; Yang, T.; Chen, T.; Yuan, N.; et al. Polar species for effective dielectric regulation to achieve high-performance CsPbI3 solar cells. Adv. Mater. 2022, 34, 2202735. [Google Scholar] [CrossRef] [PubMed]
  33. Laska, M.; Krzemińska, Z.; Kluczyk-Korch, K.; Schaadt, D.; Popko, E.; Jacak, W.A.; Jacak, J.E. Metallization of solar cells, exciton channel of plasmon photovoltaic effect in perovskite cells. Nano Energy 2020, 75, 104751. [Google Scholar] [CrossRef]
  34. Lee, J.; Park, N. Chemical approaches for stabilizing perovskite solar cells. Adv. Energy Mater. 2020, 10, 1903249. [Google Scholar] [CrossRef]
  35. Maafa, I.M. All-inorganic perovskite solar cells: Recent advancements and challenges. Nanomaterials 2022, 12, 1651. [Google Scholar] [CrossRef]
  36. Eze, V.O.; Carani, L.B.; Majumder, H.; Uddin, M.J.; Okoli, O.I. Inorganic cesium lead mixed halide based perovskite solar materials modified with functional silver iodide. Sci. Rep. 2022, 12, 7794. [Google Scholar] [CrossRef] [PubMed]
  37. Liu, L.; Lu, J.; Wang, H.; Cui, Z.; Giorgi, G.; Bai, Y.; Chen, Q. A-site phase segregation in mixed cation perovskite. Mater. Rep. Energy 2021, 1, 100064. [Google Scholar] [CrossRef]
  38. Zhai, M.; Wang, A.; Chen, C.; Hao, F.; Wang, H.; Ding, L.; Yang, X.; Cheng, M. Construct efficient CsPbI2Br solar cells by minimizing the open-circuit voltage loss through controlling the peripheral substituents of hole-transport materials. Chem. Eng. J. 2021, 425, 131675. [Google Scholar] [CrossRef]
  39. Wang, Y.; Chen, Y.; Zhang, T.; Wang, X.; Zhao, Y. Chemically stable black phase CsPbI3 inorganic perovskites for high-efficiency photovoltaics. Adv. Mater. 2020, 32, 2001025. [Google Scholar] [CrossRef]
  40. Protesescu, L.; Yakunin, S.; Bodnarchuk, M.I.; Krieg, F.; Caputo, R.; Hendon, C.H.; Yang, R.X.; Walsh, A.; Kovalenko, M.V. Nanocrystals of cesium lead halide perovskites (CsPbX3, X = Cl, Br, and I): Novel optoelectronic materials showing bright emission with wide color gamut. Nano Lett. 2015, 15, 3692–3696. [Google Scholar] [CrossRef] [Green Version]
  41. Xiang, W.; Liu, S.; Tress, W. Interfaces and interfacial layers in inorganic perovskite solar cells. Angew. Chem. Int. Ed. 2021, 60, 26440–26453. [Google Scholar] [CrossRef]
  42. Cui, Y.; Shi, J.; Meng, F.; Yu, B.; Tan, S.; He, S.; Tan, C.; Li, Y.; Wu, H.; Luo, Y.; et al. A versatile molten-salt induction strategy to achieve efficient CsPbI3 perovskite solar cells with a high open-circuit voltage >1.2 V. Adv. Mater. 2022, 34, 2205028. [Google Scholar] [CrossRef] [PubMed]
  43. Marronnier, A.; Roma, G.; Boyer-Richard, S.; Pedesseau, L.; Jancu, J.; Bonnassieux, Y.; Katan, C.; Stoumpos, C.C.; Kanatzidis, M.G.; Even, J. Anharmonicity and disorder in the black phases of cesium lead iodide used for stable inorganic perovskite solar cells. ACS Nano 2018, 12, 3477–3486. [Google Scholar] [CrossRef] [PubMed]
  44. Kulbak, M.; Cahen, D.; Hodes, G. How important is the organic part of lead halide perovskite photovoltaic cells? Efficient CsPbBr3 cells. J. Phys. Chem. Lett. 2015, 6, 2452–2456. [Google Scholar] [CrossRef]
  45. Jeon, N.J.; Noh, J.H.; Kim, Y.C.; Yang, W.; Ryu, S.; Seok, S.I. Solvent engineering for high-performance inorganic-organic hybrid perovskite solar cells. Nat. Mater. 2014, 13, 897. [Google Scholar] [CrossRef] [PubMed]
  46. Guo, Y.; Yin, X.; Liu, J.; Wen, S.; Wu, Y.; Que, W. Inorganic CsPbIBr2-based perovskite solar cells: Fabrication technique modification and efficiency improvement. Sol. RRL 2019, 3, 1900135. [Google Scholar] [CrossRef]
  47. Yang, S.; Guo, Z.; Gao, L.; Yu, F.; Zhang, C.; Fan, M.; Wei, G.; Ma, T. Bifunctional dye molecule in all-inorganic CsPbIBr2 perovskite solar cells with efficiency exceeding 10%. Sol. RRL 2019, 3, 1900212. [Google Scholar] [CrossRef]
  48. Wang, K.; Zheng, L.; Hou, Y.; Nozariasbmarz, A.; Poudel, B.; Yoon, J.; Ye, T.; Yang, D.; Pogrebnyakov, A.V.; Gopalan, V.; et al. Overcoming Shockley-Queisser limit using halide perovskite platform? Joule 2022, 6, 756–771. [Google Scholar] [CrossRef]
  49. Wang, Z.; Zhu, X.; Zuo, S.; Chen, M.; Zhang, C.; Wang, C.; Ren, X.; Yang, Z.; Liu, Z.; Xu, X.; et al. 27%-efficiency four-terminal perovskite/silicon tandem solar cells by sandwiched gold nanomesh. Adv. Funct. Mater. 2020, 30, 1908298. [Google Scholar] [CrossRef]
  50. Wang, R.; Huang, T.; Xue, J.; Tong, J.; Zhu, K.; Yang, Y. Prospects for metal halide perovskite-based tandem solar cells. Nat. Photonics 2021, 15, 411–425. [Google Scholar] [CrossRef]
  51. Guo, Q.; Duan, J.; Zhang, J.; Zhang, Q.; Duan, Y.; Yang, X.; He, B.; Zhao, Y.; Tang, Q. Universal dynamic liquid interface for healing perovskite solar cells. Adv. Mater. 2022, 34, 2202301. [Google Scholar] [CrossRef] [PubMed]
  52. Liu, Q.; Qiu, J.; Yan, X.; Fei, Y.; Qiang, Y.; Chang, Q.; Wei, Y.; Zhang, X.; Tian, W.; Jin, S.; et al. Surface passivation and hole extraction: Bifunctional interfacial engineering toward high-performance all-inorganic CsPbIBr2 perovskite solar cells with efficiency exceeding 12%. J. Energy Chem. 2022, 74, 387–393. [Google Scholar] [CrossRef]
  53. Yang, C.; Wang, H.; Miao, Y.; Chen, C.; Zhai, M.; Bao, Q.; Ding, X.; Yang, X.; Cheng, M. Interfacial molecular doping and energy level alignment regulation for perovskite solar cells with efficiency exceeding 23%. ACS Energy Lett. 2021, 6, 2690–2696. [Google Scholar] [CrossRef]
  54. Pan, J.; Zhang, X.; Zheng, Y.; Xiang, W. Morphology control of perovskite film for efficient CsPbIBr2 based inorganic perovskite solar cells. Sol. Energy Mater. Sol. Cells 2021, 221, 110878. [Google Scholar] [CrossRef]
  55. Chen, Z.; Yang, M.; Li, R.; Zang, Z.; Wang, H. Double-side interface engineering synergistically boosts the efficiency of inorganic CsPbIBr2 perovskite solar cells over 12%. Adv. Opt. Mater. 2022, 10, 2200802. [Google Scholar] [CrossRef]
  56. Xu, Y.; Li, G.; Jing, Y.; Zhang, H.; Wang, X.; Lu, Y.; Wu, J.; Lan, Z. n-type absorber by Cd2+ doping achieves high-performance carbon-based CsPbIBr2 perovskite solar cells. J. Colloid. Interface Sci. 2022, 608, 40–47. [Google Scholar] [CrossRef]
  57. Wang, Q.; Xu, Y.; Zhang, L.; Yang, A.; Bai, T.; Liu, F.; Lyu, M.; Zhu, J. Guanidinium thiocyanate additive engineering for high-performance CsPbIBr2 solar cells with an efficiency of 10.90%. ACS Appl. Energy Mater. 2022, 5, 3110–3118. [Google Scholar] [CrossRef]
  58. Liang, J.; Liu, Z.; Qiu, L.; Hawash, Z.; Meng, L.; Wu, Z.; Jiang, Y.; Ono, L.; Qi, B. Enhancing optical, electronic, crystalline, and morphological properties of cesium lead halide by Mn substitution for high-stability all-inorganic perovskite solar cells with carbon electrodes. Adv. Energy Mater. 2018, 8, 1800504. [Google Scholar] [CrossRef]
  59. Guo, Y.; Zhao, F.; Wang, X.; Tao, J.; Zheng, D.; Jiang, J.; Hu, Z.; Chu, J. Multi-source cation/anion doping towards efficient carbon-based CsPbIBr2 solar cells with superior open voltage up to 1.37 V. Sol. Energy Mater. Sol. Cells 2021, 221, 110918. [Google Scholar] [CrossRef]
  60. Yang, X.; Xiang, H.; Huang, J.; Zhou, C.; Ran, R.; Wang, W.; Zhou, W.; Shao, Z. Thiourea with sulfur-donor as an effective additive for enhanced performance of lead-free double perovskite photovoltaic cells. J. Colloid. Interface Sci. 2022, 628, 476–485. [Google Scholar] [CrossRef]
  61. Liang, J.; Zhao, P.; Wang, C.; Wang, Y.; Hu, Y.; Zhu, G.; Ma, L.; Liu, J.; Jin, Z. CsPb0.9Sn0.1IBr2 based all-inorganic perovskite solar cells with exceptional efficiency and stability. J. Am. Chem. Soc. 2017, 139, 14009–14012. [Google Scholar] [CrossRef]
  62. Liu, P.; Yang, X.; Chen, Y.; Xiang, H.; Wang, W.; Ran, R.; Zhou, W.; Shao, Z. Promoting the efficiency and stability of CsPbIBr2-based all-inorganic perovskite solar cells through a functional Cu2+ doping strategy. ACS Appl. Mater. Interfaces 2020, 12, 23984–23994. [Google Scholar] [CrossRef]
  63. Wang, D.; Li, W.; Zhang, T.; Liu, X.; Jin, X.; Xu, B.; Li, D.; Huang, Z.; Li, Q.; Lan, Z.; et al. Zinc and acetate co-doping for stable carbon-based CsPbIBr2 solar cells with efficiency over 10.6%. ACS Appl. Energy Mater. 2022, 5, 2720–2726. [Google Scholar] [CrossRef]
  64. Zhang, W.; Liu, H.; Qi, X.; Yu, Y.; Zhou, Y.; Xia, Y.; Cui, J.; Shi, Y.; Chen, R.; Wang, H. Oxalate pushes efficiency of CsPb0.7Sn0.3IBr2 based all-inorganic perovskite solar cells to over 14%. Adv. Sci. 2022, 9, 2106054. [Google Scholar] [CrossRef] [PubMed]
  65. Subhani, W.S.; Wang, K.; Du, M.; Liu, S. Goldschmidt-rule-deviated perovskite CsPbIBr2 by barium substitution for efficient solar cells. Nano Energy 2019, 61, 165–172. [Google Scholar] [CrossRef]
  66. Xiang, W.; Wang, Z.; Kubicki, D.; Tress, W.; Luo, J.; Prochowicz, D.; Akin, S.; Emsley, L.; Zhou, J.; Dietler, G.; et al. Europium-doped CsPbI2Br for stable and highly efficient inorganic perovskite solar cells. Joule 2019, 3, 205–214. [Google Scholar] [CrossRef]
  67. Yang, X.; Chen, Y.; Liu, P.; Xiang, H.; Wang, W.; Ran, R.; Zhou, W.; Shao, Z. Simultaneous power conversion efficiency and stability enhancement of Cs2AgBiBr6 lead-free inorganic perovskite solar cell through adopting a multifunctional dye interlayer. Adv. Funct. Mater. 2020, 30, 2001557. [Google Scholar] [CrossRef]
  68. Jena, A.; Ishii, A.; Guo, Z.; Kamarudin, M.; Hayase, S.; Miyasaka, T. Cesium acetate-induced interfacial compositional change and graded band level in MAPbI3 perovskite solar cells. ACS Appl. Mater. Interfaces 2020, 12, 33631–33637. [Google Scholar] [CrossRef]
  69. Hayashi, S.; Hayamizu, K. Nuclear-magnetic-resonance chemical-shifts in alkali iodides, cuprous halides and silver-halides. J. Phys. Chem. Solids 1992, 53, 239–248. [Google Scholar] [CrossRef]
  70. Tang, M.; He, B.; Dou, D.; Liu, Y.; Duan, J.; Zhao, Y.; Chen, H.; Tang, Q. Toward efficient and air-stable carbon-based all-inorganic perovskite solar cells through substituting CsPbBr3 films with transition metal ions. Chem. Eng. J. 2019, 375, 121930. [Google Scholar] [CrossRef]
  71. Guo, Y.; Yin, X.; Liu, J.; Que, W. Highly efficient CsPbIBr2 perovskite solar cells with efficiency over 9.8% fabricated using a preheating-assisted spin-coating method. J. Mater. Chem. A 2019, 7, 19008–19016. [Google Scholar] [CrossRef]
  72. Wang, R.; Zhang, H.; Han, S.; Wu, Y.; Hu, Z.; Zhang, G.; Liu, H.; He, Q.; Zhang, X. Cadmium doping for improving the efficiency and stability of carbon-based CsPbIBr2 all-inorganic perovskite solar cells. New J. Chem. 2021, 45, 9243–9250. [Google Scholar] [CrossRef]
  73. Zhu, W.; Zhang, Q.; Chen, D.; Zhang, Z.; Lin, Z.; Chang, J.; Zhang, J.; Zhang, C.; Hao, Y. Intermolecular exchange boosts efficiency of air-stable, carbon-based all-inorganic planar CsPbIBr2 perovskite solar cells to over 9%. Adv. Energy Mater. 2018, 8, 1802080. [Google Scholar] [CrossRef]
  74. Liu, P.; Han, N.; Wang, W.; Ran, R.; Zhou, W.; Shao, Z. High-quality Ruddlesden-Popper perovskite film formation for high-performance perovskite solar cells. Adv. Mater. 2021, 33, 2002582. [Google Scholar] [CrossRef]
  75. Gao, B.; Meng, J. Highly stable all-inorganic CsPbIBr2 perovskite solar cells with 11.30% efficiency using crystal interface passivation. ACS Appl. Energy Mater. 2020, 3, 8249–8256. [Google Scholar] [CrossRef]
  76. Zhu, W.; Chai, W.; Chen, D.; Xi, H.; Chen, D.; Chang, J.; Zhang, J.; Zhang, C.; Hao, Y. Recycling of FTO/TiO2 substrates: Route toward simultaneously high-performance and cost-efficient carbon-based, all-inorganic CsPbIBr2 solar cells. ACS Appl. Mater. Interfaces 2020, 12, 4549–4557. [Google Scholar] [CrossRef] [PubMed]
  77. Jing, Y.; Liu, X.; Wang, D.; Li, R.; Xu, Y.; Yan, Z.; Sun, W.; Wu, J.; Lan, Z. High-efficiency and ultraviolet stable carbon-based CsPbIBr2 solar cells from single crystal three-dimensional anatase titanium dioxide nanoarrays with ultraviolet light shielding function. J. Colloid. Interface Sci. 2022, 616, 201–209. [Google Scholar] [CrossRef]
  78. Zhang, C.; Wang, K.; Wang, Y.; Subhani, W.S.; Jiang, X.; Wang, S.; Bao, H.; Liu, L.; Wan, L.; Liu, S. Low-temperature crystallization of CsPbIBr2 perovskite for high performance solar cells. Sol. RRL 2020, 4, 2000254. [Google Scholar] [CrossRef]
  79. Chen, Y.; Yang, X.; Liu, P.; Wang, W.; Ran, R.; Zhou, W.; Shao, Z. Improving moisture/thermal stability and efficiency of CH3NH3PbI3-based perovskite solar cells via gentle butyl acrylate additive strategy. Sol. RRL 2020, 5, 2000621. [Google Scholar] [CrossRef]
  80. Xiang, H.; Liu, P.; Wang, W.; Ran, R.; Zhou, W.; Shao, Z. Sodium fluoride sacrificing layer concept enables high-efficiency and stable methylammonium lead iodide perovskite solar cells. J. Mater. Sci. Technol. 2022, 113, 138–146. [Google Scholar] [CrossRef]
  81. Han, Y.; Zhao, H.; Duan, C.; Yang, S.; Yang, Z.; Liu, Z.; Liu, S. Controlled n-doping in air-stable CsPbI2Br perovskite solar cells with a record efficiency of 16.79%. Adv. Funct. Mater. 2020, 30, 1909972. [Google Scholar] [CrossRef]
  82. Zou, S.; Liu, Y.; Li, J.; Liu, C.; Feng, R.; Jiang, F.; Li, Y.; Song, J.; Zeng, H.; Hong, M.; et al. Stabilizing cesium lead halide perovskite lattice through Mn (II) substitution for air-stable light-emitting diodes. J. Am. Chem. Soc. 2017, 139, 11443–11450. [Google Scholar] [CrossRef] [PubMed]
  83. Guo, Y.; Wang, Q.; Saidi, W.A. Structural stabilities and electronic properties of high-angle grain boundaries in perovskite cesium lead halides. J. Phys. Chem. C 2017, 121, 1715–1722. [Google Scholar] [CrossRef]
  84. Tan, X.; Liu, X.; Liu, Z.; Sun, B.; Li, J.; Xi, S.; Shi, T.; Tang, Z.; Liao, G. Enhancing the optical, morphological and electronic properties of the solution-processed CsPbIBr2 films by Li doping for efficient carbon-based perovskite solar cells. Appl. Surf. Sci. 2020, 499, 143990. [Google Scholar] [CrossRef]
Figure 1. (a) J−V curves of CsPbIBr2 and various Cu+-doped CsPbIBr2 cells under a reverse scan direction. (b) EQE profiles, integrated photocurrent densities based on CsPbIBr2 and CsPbIBr2-0.50%Cu devices, (c) statistical PCE distributions of PSCs based on CsPbIBr2 and CsPbIBr2-0.50%Cu, (d) steady PCE and Jsc values of CsPbIBr2 and CsPbIBr2-0.50%Cu cells.
Figure 1. (a) J−V curves of CsPbIBr2 and various Cu+-doped CsPbIBr2 cells under a reverse scan direction. (b) EQE profiles, integrated photocurrent densities based on CsPbIBr2 and CsPbIBr2-0.50%Cu devices, (c) statistical PCE distributions of PSCs based on CsPbIBr2 and CsPbIBr2-0.50%Cu, (d) steady PCE and Jsc values of CsPbIBr2 and CsPbIBr2-0.50%Cu cells.
Nanomaterials 12 04317 g001
Figure 2. (a) XRD patterns, (b) magnified (100) and (200) peaks, (c) UV-vis and (d) PL spectra (deposited on conductive substrate) of CsPbIBr2 and different Cu+-doped CsPbIBr2 films.
Figure 2. (a) XRD patterns, (b) magnified (100) and (200) peaks, (c) UV-vis and (d) PL spectra (deposited on conductive substrate) of CsPbIBr2 and different Cu+-doped CsPbIBr2 films.
Nanomaterials 12 04317 g002
Figure 3. Top-view SEM images of (a,b) CsPbIBr2 and (c,d) CsPbIBr2-0.50%Cu films at different magnifications. (e) EDX elemental mapping images of CsPbIBr2-0.50%Cu film including Cs, Pb, I, Br, Cu elements. Top-view AFM images of (f) CsPbIBr2 and (g) CsPbIBr2-0.50%Cu films.
Figure 3. Top-view SEM images of (a,b) CsPbIBr2 and (c,d) CsPbIBr2-0.50%Cu films at different magnifications. (e) EDX elemental mapping images of CsPbIBr2-0.50%Cu film including Cs, Pb, I, Br, Cu elements. Top-view AFM images of (f) CsPbIBr2 and (g) CsPbIBr2-0.50%Cu films.
Nanomaterials 12 04317 g003
Figure 4. (a) EIS spectra and (b) J−V curves of CsPbIBr2 and CsPbIBr2-0.50%Cu cells in the dark. Dark J−V curves of (c) electron-only and (d) hole-only PSCs based on CsPbIBr2 and CsPbIBr2-0.50%Cu.
Figure 4. (a) EIS spectra and (b) J−V curves of CsPbIBr2 and CsPbIBr2-0.50%Cu cells in the dark. Dark J−V curves of (c) electron-only and (d) hole-only PSCs based on CsPbIBr2 and CsPbIBr2-0.50%Cu.
Nanomaterials 12 04317 g004
Figure 5. Normalized photovoltaic parameters of CsPbIBr2 and CsPbIBr2-0.50%Cu cells stored in humid air at (a) 25 °C and (b) 85 °C without encapsulation.
Figure 5. Normalized photovoltaic parameters of CsPbIBr2 and CsPbIBr2-0.50%Cu cells stored in humid air at (a) 25 °C and (b) 85 °C without encapsulation.
Nanomaterials 12 04317 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Du, Z.; Xiang, H.; Xie, A.; Ran, R.; Zhou, W.; Wang, W.; Shao, Z. Monovalent Copper Cation Doping Enables High-Performance CsPbIBr2-Based All-Inorganic Perovskite Solar Cells. Nanomaterials 2022, 12, 4317. https://doi.org/10.3390/nano12234317

AMA Style

Du Z, Xiang H, Xie A, Ran R, Zhou W, Wang W, Shao Z. Monovalent Copper Cation Doping Enables High-Performance CsPbIBr2-Based All-Inorganic Perovskite Solar Cells. Nanomaterials. 2022; 12(23):4317. https://doi.org/10.3390/nano12234317

Chicago/Turabian Style

Du, Zhaonan, Huimin Xiang, Amin Xie, Ran Ran, Wei Zhou, Wei Wang, and Zongping Shao. 2022. "Monovalent Copper Cation Doping Enables High-Performance CsPbIBr2-Based All-Inorganic Perovskite Solar Cells" Nanomaterials 12, no. 23: 4317. https://doi.org/10.3390/nano12234317

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop