Download PDF
Article  |  Open Access  |  14 Apr 2024

Evolved photovoltaic performance of MAPbI3 and FAPbI3-based perovskite solar cells in low-temperatures

Views: 203 |  Downloads: 27 |  Cited:   0
Energy Mater 2024;4:400034.
10.20517/energymater.2023.86 |  © The Author(s) 2024.
Author Information
Article Notes
Cite This Article

Abstract

Organic-inorganic hybrid perovskites have emerged as an up-and-coming contender for photovoltaic devices owing to their exceptional photovoltaic properties. However, current research predominantly concentrates on their performance under ambient conditions at room temperature. In this work, we delve into the novel territory by investigating MAPbI3-based and FAPbI3-based perovskite solar cells (PSCs) in the temperature range of 300 to 150 K. Remarkable efficiency enhancements of nearly 5% and 20% were obtained at 250 and 210 K, respectively. However, further decreasing the temperature impairs the photovoltaic performance. We propose an underlying mechanism influencing the performance change in perovskite devices at low temperatures by examining the temperature-dependent ultraviolet-visible and photoluminescence spectra results. At the beginning of the cooling process, from 300 to 250 K for MAPbI3 and from 300 to 210 K for FAPbI3, the performance enhancement stems primarily from the enhanced open-circuit voltage by the tuned band gap of the perovskite films. Further lowering the temperature would change the perovskite structure, impairing the performance of PSCs. FAPbI3-based PSCs show a better tolerance in low temperatures owing to the more stable perovskite crystal structure. The present findings offer valuable theoretical guidance for preparing outstanding PSCs for low-temperature applications.

Keywords

Perovskite solar cells, low temperature, phase transition, defective state, carrier trapping

INTRODUCTION

The organic-inorganic perovskite solar cell (PSC), described as ABX3, holds significant promise due to its tunable photoelectric properties and extended carrier lifetime[1], in which A represents monovalent cations such as formamidine cation (FA+) and methylamine cation (MA+), B typically denotes Pb2+ or Sn2+, and X signifies halide anion Cl-, Br- and I-. Since the pioneer work of Dieter Weber on the three-dimensional MAPbX3 PSC in 1978[2], persistent endeavors have been dedicated to refining the formula and enhancing the process, resulting in notable advancements such as achieving a remarkable 26.1% power conversion efficiency (PCE)[3]. At the same time, researchers are continually proposing methods that can further improve the performance of perovskite cells, such as employing small fluorinated organic ammonium molecules to compensate for the pinholes and cracks in perovskites to achieve high-quality perovskite films with larger grain sizes and lower crystal defect densities[4] or using InBr3 modifications to construct homojunctions to expand the electric field and enhance device performance[5].

Perovskites are renowned for their remarkable photoelectric properties at room temperature, prompting interest in their applicability in low-temperature environments such as polar regions and terrestrial space. However, such extreme conditions can lead to decreased efficiency and irreversible damage to photovoltaic cells. For instance, current spacecraft solar cells predominantly employ silicon and GaAs. Measured parameters under the air mass 1.5 global (AM 1.5G) spectrum (1,000 W/m2) at 25 °C reveal the highest PCE of 37.9% and open-circuit voltage (VOC) of 3.06V[6] of GaAs, alongside peak short-circuit current (JSC) for 42.62 mA/cm2 of single crystal Si[7]. In comparison, organic-inorganic perovskites show average parameters in PCE, VOC and JSC. However, they exhibit a high absorption coefficient, impressive mechanical flexibility and radiation tolerance, positioning them as excellent candidates for spacecraft solar systems.

Recently, there has been a growing exploration of perovskite performance in low-temperature conditions, unveiling its potential for development. In a study by Chen et al., conventional n-i-p planar solar cell employing (FA,MA,Cs)Pb(I,Br)3 (FAMACs)-based perovskite materials achieved a remarkable 25.2% efficiency at 220 K, showing an increasing PCE trend from 300 to 220 K, followed by a decline as temperature further decreased to 130 K[8]. Similar trends were observed in the VOC and hysteresis index, but no change was noted in Jsc density across the temperature range. A theory proposing a transition activating the self-elimination of the intrinsic defect was introduced to explain these results, later validated through density functional theory (DFT) calculations.

Herein, we delve into FA-based and MA-based PSCs across varying operational temperatures (300 to 150 K) and explore FAPbI3 and MAPbI3 perovskite films from 300 to 100 K. Notably, we observed an enhancement in PCE and VOC for both FA and MA solar cells, albeit occurring at different temperature points. To comprehensively understand these variations, temperature-dependent steady-state photoluminescence (PL) measurements and temperature-dependent ultraviolet-visible (UV-vis) transmittance spectra measurements were also conducted to explore the change in bandgap and photovoltaic behavior. Notably, distinct discrepancies emerged between FA- and MA-based perovskites. This investigation uncovers a correlation between perovskite bandgap behavior and device performance.

EXPERIMENTAL

Materials

Herein, all the materials were purchased from Sigma-Aldrich or Alfa-Aesar and used without further purification. Ethanol (anhydrous) and isopropyl alcohol (> 99.7%) were purchased from Greagent. PbI2 (99.99%) was bought from Tokyo Chemical Industry (TCI). Formamidinium iodide (FAI) (> 99.99%), methylammonium iodide (MAI) (> 99.99%), and methylammonium bromide (MABr) (> 99.99%) were purchased from Greatcell Solar. Spiro-OMeTAD (99.7%) was bought from Borun Chemical Co., Ltd. All the materials were used as received without purification.

Materials synthesis

Fabrication of solar cells. Substrates were cleaned in deionized water, ethanol, acetone, and isopropyl alcohol in sequence and treated with O2-plasma for 15 min to remove organic contamination and construct a hydrophilic interface. The SnO2 nanoparticle dispersion obtained was spun on the cleaned indium tin oxide-coated glass substrate at 4,000 rpm for 30 s, followed by annealing at 150 °C for 40 min. The FAPbI3 perovskite was deposited using a two steps program at 1,000 rpm for 10 s in the first step and 5,000 rpm for 30 s in the second, during which diethyl ether was dropped onto the center of the substrate at 22 s after the start of the procedure. The MAPbI3 perovskite was deposited employing a two-step program at 1,000 rpm for 10 s in the first step and 4,000 rpm for 40 s in the second, during which diethyl ether was dropped onto the center of the substrate at 10 s before the end of the procedure. Subsequently, the hole transporting material Spiro-OMeTAD was spun on the perovskite film at 3,000 rpm for 30 s. Finally, 100 nm gold electrodes were deposited on substrates through thermal evaporation.

Fabrication of perovskite films. The FAPbI3 and MAPbI3 perovskite film for characterization was deposited on SnO2 spun indium tin oxide-coated glass substrate with the same procedure above.

Characterization

The current density-voltage (J-V) curves for all devices at room temperature were carried out with a digital source meter (2400, Keithley Instruments, USA) scanning from -0.1 to 1.3 V at a scan rate of 0.05 V·s-1 under AM 1.5G illumination (100 mW·cm-2) with a 0.06 cm2 metal mask over, which was realized using a solar simulator (Oriel, USA, calibrated with a standard crystalline silicon solar cell). The temperature-dependent J-V curves were measured by a solar simulator (Newport, USA) while samples were put in a hermetic probe station connected to a digital source meter (2400, Keithley Instruments, USA).

The morphology of perovskite samples was characterized by a scanning electron microscope (SEM) (Zeiss, LEO1530, Germany). Temperature-dependent steady-state PL measurements were conducted in a step of 2 nm by exciting the sample with a monochromatic xenon lamp (central wavelength λ = 460 nm) under a position of a 45° angle towards the excitation beam. The sample stage was kept in a hermetic and vacuum environment with a temperature control system. Temperature-dependent UV-vis transmittance spectra were performed using a Lambda 950 spectrophotometer (PerkinElmer, USA), sharing the same temperature control system with the PL test. Temperature-dependent Raman spectra measurements were recorded on a laser micro-Raman spectroscope (Horiba Jobin Yvon, Hr800, France) with a laser beam wavelength of 532 nm for the MAPbI3 perovskite film and a 633 nm laser beam for the FAPbI3 film.

RESULTS AND DISCUSSION

We first observed the surface morphology of MAPbI3 and FAPbI3 perovskite thin films prepared on SnO2 substrates using SEM, as shown in Figure 1, where a represents the MAPbI3 sample, and b indicates the FAPbI3 sample. It can be observed that, at room temperature, both types of perovskite thin films exhibit good uniformity, and no pores are present on the surface of the films. The grain size distribution of the FAPbI3 film is broader, with the largest size approaching 1.9 µm; the grain sizes in the MAPbI3 film are more similar, not exceeding 500 nm.

Evolved photovoltaic performance of MAPbI<sub>3</sub> and FAPbI<sub>3</sub>-based perovskite solar cells in low-temperatures

Figure 1. The morphology of perovskite films at room temperature: (A) MAPbI3 film; (B) FAPbI3 film.

We then studied the change of the UV-vis spectra of the polycrystalline perovskite films employing a cooling process from 300 to 110 K [Figure 2]. CH3NH3PbI3 (MAPI) and CH(NH2)2PbI3 (FAPI) were deposited onto SnO2-obtained indium tin oxide (ITO) glass. As the high transparency of the SnO2 film is above the 400 nm range[9], it has negligible impact on spectral pattern. Therefore, any disparities observed in the spectra can be attributed to perovskite behavior. Figure 2A presents the MAPbI3 spectra spanning from 675 to 825 nm. Notably, a distinct peak emerges at approximately 770 nm when measured at 170 K, which is a manifestation of excitonic transition. The peak position also shifts towards shorter wavelengths as the temperature increases, frequently observed in lead composite semiconductors[10]. However, at 300 K, discerning this peak from the conduction band continuum becomes challenging. Another interesting thing is that an excitonic peak at around 715 nm occurs when measured at 160 K. As the temperature further decreased, there was a sudden jump of the peak towards longer wavelengths. This behavior can be attributed to a tetragonal-to-orthorhombic phase transition occurring near 160 K[11].

Evolved photovoltaic performance of MAPbI<sub>3</sub> and FAPbI<sub>3</sub>-based perovskite solar cells in low-temperatures

Figure 2. Temperature-dependent UV-vis spectra for (A and B) MAPbI3 and (C and D) FAPbI3 with their calculated bandgaps in different temperatures. UV-vis: Ultraviolet-visible.

With detailed calculation shown in Supplementary Figure 1, Figure 2B shows the calculated band gap of MAPbI3, where we can observe a positive thermal expansion coefficient of its band gap in both tetragonal and orthorhombic structures besides a sudden jump at the phase-transition point of 170 K. However, it is raised that inorganic semiconductors, such as silicon and germanium, have an increase in their band gap as temperature goes down due to the change of electron-phonon interaction[12]. One theory to explain such abnormal positive thermal expansion coefficient in tetragonal and orthorhombic structure phases is that a photon mode, Pb-I-Pb angular bending vibration, may contribute to the position of the band gap by thermal population as the position of the conduction and valence bands in a semiconductor is decided by phonon populations at specific temperature[11]. First-principles calculation of band gap in other research shows that tetragonal and orthorhombic MAPbI3 perovskite has 1.60 and 1.82 eV bandgaps, respectively, which is reflected well by the calculated bandgaps in Figure 2B[13]. The energy rise in σ-antibonding orbitals of Pb and I may lead to a narrow bandwidth[14].

Figure 2C illustrates the spectra of the FAPbI3 perovskite film at different temperatures. Compared to those of MAPbI3, it exhibits a similar pattern at 300 K. The featuring peak demonstrates a similar shift towards longer wavelengths when the temperature falls. When the temperature reaches and goes below 160 K, FAPbI3 displays a slight bump, possibly due to a small crystal-structure transient. A previous study indicated that there might be a transition from the tetragonal (I) phase to the tetragonal (II) phase when a more subtle alteration in the chemical and structural environment happens[15]. In this case, FAPbI3 perovskite shows similarity in its tetragonal (I) phase. Still, it undergoes a continuous change and an inverse negative thermal expansion coefficient in the tetragonal (II) phase [Figure 2D]. Such a phenomenon requires further research.

Additionally, we used SEM to observe the microstructure of MAPbI3 thin film samples just after completing the cooling process [Supplementary Figure 2]. Compared to the SEM photos taken before cooling, the previously uniform sample films showed a greater number of pores after the cooling process. We speculate that the phase transition-induced change in lattice size during cooling reduces grain size. Although the temperature is restored during observation under room temperature SEM, the size changes caused by the phase transition, influenced by spatial orientation, cannot fully revert. This results in the observation of numerous pores on the sample films.

The changed energetic structure of MAPbI3 and FAPbI3 films would directly affect the photon-generating carrier dynamics, such as the exciton physics. Herein, we performed temperature-dependent steady-state PL measurements to explore the evolving optoelectronic properties of MAPbI3 and FAPbI3 films. Figure 3A and B summarizes the temperature-dependent PL spectra for MAPbI3 film. Colorful mapping of normalized PL spectra in different temperatures is depicted in Figure 3A to reveal the change of emission peak, with the central positions in Figure 3B. At 300 K, only one peak appears at the wavelength of nearly 780 nm. When the temperature declines, the high-temperature peak (line Peak 2) exhibits a continuous shift towards longer wavelengths in its position before 160 K and shrinks in width. This shifting tendency in position from 290 to 160 K is consistent with the decreasing band gap calculated in Figure 2B.

Evolved photovoltaic performance of MAPbI<sub>3</sub> and FAPbI<sub>3</sub>-based perovskite solar cells in low-temperatures

Figure 3. Temperature-dependent photoluminescence spectra for MAPbI3 and FAPbI3. (A) Color plots of normalized PL spectra for MAPbI3-based perovskites at different temperatures from 110 to 300 K; (B) The corresponding excitonic peak positions of MAPbI3-based perovskites; (C) Color plots of normalized PL spectra for FAPbI3-based perovskites at varying temperatures from 50 to 290 K; (D) The corresponding excitonic peak position of the FAPbI3-based perovskites. PL: Photoluminescence.

However, two PL peaks are observed when the temperature goes down below 150 K [Figure 3B]. Then, the long-wavelength peak showcases a shift tendency towards shorter wavelengths, while the short-wavelength peak (line Peak 1) presents a shift tendency toward long-wavelength direction. As discussed above, there is a transition from the tetragonal to the orthorhombic phase. Thus, we believe that Peak 1 belongs to the orthorhombic phase. A previous study suggested a discontinuity in the crystal structure when MAPbI3 undergoes a phase transition from the high-temperature tetragonal phase to the low-temperature orthorhombic phase, creating new energetic states[16]. Thus, we attributed the PL signal of Peak 2 at the temperature below 160 K to the new states in the orthorhombic phase, where carriers would recombine and emit PL with long wavelengths.

We also conducted temperature-dependent PL measurements for FAPbI3 film and summarized the data [Figure 3]. Colorful mapping of the normalized PL spectra in different temperatures is shown in Figure 3C. The central positions and full width of half maximum of the excitonic peaks are extracted and given in Figure 2D and Supplementary Figure 3. As the temperature decreases, the peak exhibits an apparent shift tendency towards longer wavelengths and a decrease in full width at half maximum (FWHM), displaying the evolution of the free exciton with temperature[17]. Such shift tendency observed in position from 290 to 160 K is consistent with the evolving band gap calculated in Figure 2D. Compared with MAPbI3, FAPbI3 perovskite shows a simpler evolution tendency owing to its more stable crystal structure.

We further scrutinized the detailed connection between the PL signal and the perovskite structure. As reported, the integrated PL intensity can be fitted by Arrhenius equations with two thermal activation energies, E1 and E2, as

$$ I(T)=\frac{I_{0}}{1\;+\;C_{1} \exp \left(-\frac{E_{1}}{k_{B} T}\right)+C_{2} \exp \left(-\frac{E_{2}}{k_{B} T}\right)} $$

where I(T) and I0 are the integrated intensities at T and 0 K. C1 and C2 are constants. The reciprocal of integrated PL intensities for FAPbI3 perovskite in different temperatures is fitted using Equation (1) and plotted in Supplementary Figure 4. Then, thermal activation energies E1 and E2 are extracted as 441.7 and 17.6 meV, respectively, showing a high correlation with a R2 of 0.99844. It is reported that E1 and E2, virtually representing exciton binding energies in various phases, can be approximately attributed to the activation energies for cubic and tetragonal perovskite phases. Meanwhile, they are also sensitive to the synthesis, composition, and structure details[18]. As shown in Figure 3D, a temperature-dependent trend of the peak positions indicates a smaller binding energy of excitons when the temperature is higher than 160 K. The featuring peak position shows a temperature-independent tendency due to the easy formation of exciton complexes from photoexcited carriers at extremely low temperatures. Thus, we can witness apparent sharp peaks in the low-temperature range in Figure 2C. The changed crystal phase as tetragonal (II) would be responsible for the higher exciton binding energy.

Undoubtedly, the evolved exciton dynamics in varying temperatures has a central role in determining the photovoltaic performance of PSCs. For example, the change of exciton binding energy might affect the separation of electron and hole and then the IV performance. A lower binding energy would be favorable. According to the above results indicating that excitons at a temperature higher than 150 K have a smaller binding energy, we study the photovoltaic performance of whole PSCs by conducting the J-V measurements from 293 to 153 K in a step of 10 K. The PSCs had an ITO/SnO2/perovskite/Spiro-OMeTAD/Au structure where FAPbI3 and MAPbI3 were used as the light absorber. The obtained PCE, fill factor (FF), VOC, and JSC of the MAPbI3- and FAPbI3-based PSCs are recorded in Figure 4A-D.

Evolved photovoltaic performance of MAPbI<sub>3</sub> and FAPbI<sub>3</sub>-based perovskite solar cells in low-temperatures

Figure 4. The photovoltaic performance of MAPbI3-based PSCs (blue points) and FAPbI3-based PSCs (red points) from 290 to 150 K: (A) PCE; (B) FF; (C) VOC; and (D) JSC. PSCs: Perovskite solar cells; PCE: power conversion efficiency; FF: fill factor; VOC: open-circuit voltage; JSC: short-circuit current.

In MAPbI3-based PSCs, JSC is stabilized at 20 mA in the 290 to 180K range and starts to decrease rapidly below 180 K. VOC rises continuously before 180 K and then falls sharply. FF remains essentially constant at first and shows a downward trend below 250 K. Considering that the photovoltaic parameters are satisfied by

$$ P C E=F F \times V o c \times J s c $$

The evolution of FF primarily determines the performance change of MAPI-based PSCs at low temperatures. At the temperature range of 250 to 280 K, PCE reaches a high level with a champion value of 17.21%, an improvement of nearly 5% compared to that at room temperature (290 K). In FAPbI3-based PSCs, PCEs show a similar trend that they were mainly affected by FF. FF remains almost the same when the temperature exceeds 210 K. JSC is always stabilized in the full range, while VOC rises continuously before 210 K and then falls sharply. As a result, PCE reaches a best efficiency of 20.28% at 210 K, a nearly 20% increase compared to the control efficiency at room temperature, and then decreases fast.

Next, we tried to use a diode model to scrutinize the change of photovoltaic parameters with the decreasing temperature. Previous works pointed out that the J-V characteristics of thin-film PSC can be determined by a single-exponential diode equation[19,20], expressed as

$$ J=J_{0} \exp \left[\frac{q}{A k T}(V-R J)\right]+G V-J_{S C} $$

where R is the series resistance, and G is the parallel conductance, respectively representing losses in the series and parallel circuit model. A stands for the diode ideal factor. The diode reverse current saturation density J0 can be calculated by

$$ J_{0}=J_{00} \exp \left(-\frac{E_{A}}{A k T}\right) $$

where EA is the activating energy (or barrier height) of the main recombination mechanism, and J00 is the pre-exponential factor of J0, which is virtually temperature-independent. Given G << JSC/VOC, the expression of VOC can be obtained, by combining Equations (3) and (4), as

$$ V_{O C}=\frac{E_{A}}{q}-\frac{A k T}{q} \ln \left(\frac{J_{00}}{J_{S C}}\right) $$

where the charge q, ideal factor A, Boltzmann’s constant k, and J00 are constants.

In Figure 4D, JSC of two PSCs is almost unchanged when the temperature is higher than 180 K. Thus, VOC would be a linear function of the temperature T and rise as the temperature decreases. We linearly fitted VOC with temperature for MAPbI3-based and FAPbI3-based PSCs from 290 to 210 K. The results are given in Supplementary Figures 5 and 6. The fitted R2 for MAPbI3 and FAPbI3 are separately 0.99 and 0.95, revealing a high correlation. That means the perovskites are stable in the temperature range with a relatively stable semiconductor property, consistent with the optical and PL measurement results.

In addition, the EA value can also be used to indicate the recombination type of perovskite films. If EA nearly equals Eg, it suggests the recombination in the perovskite film is dominant; if EA < Eg, it indicates that the interface recombination becomes the main mechanism[21]. We extrapolated the linear segment of the two lines to T = 0 K to calculate the value of EA of the two perovskite films. The MAPbI3 device shows a fitted EA of 1.54 ± 0.02 eV and a measured band gap of about 1.57 eV from UV-vis spectra [Figure 2B], indicating a bulk recombination rather than an interfacial recombination. As for the FAPbI3 device, a fitted EA of 1.47 ± 0.03 eV is close to the measured Eg of 1.50 eV [Figure 2D], also showing a dominant mechanism of bulk recombination in the temperature range from 210 to 290 K.

Additionally, the rapid decline in FF at 200 K for MAPbI3 and at 180 K for FAPbI3 is a major factor contributing to the significant drop in the efficiency of the solar cell devices within those corresponding temperature ranges. Given that we observed that cracks appear in the perovskite film during the cooling process along with voids between grain boundaries and considering that these temperatures are close to the phase transition temperatures of the two types of perovskite materials, we speculate that the decrease in FF may originate from the volume changes caused by the low-temperature phase transition, leading to interlayer mismatch.

Thus, it can be concluded that the PSCs comply with a generalized diode model above 210 K when the perovskite crystal structure is stable. Further decreasing the temperature would cause different evolutions. For the MAPbI3 case, as the phase transition from high-temperature tetragonal to low-temperature orthorhombic starts at about 180 K, optoelectronic properties would be changed simultaneously, impairing the photovoltaic performance. However, for the FAPbI3 case, the negative impact of temperature on the photovoltaic performance shows a considerable delay with the decreasing temperature because the phase transition from the high-temperature tetragonal to the low-temperature tetragonal takes place at 140 K. For both types of samples, the interlayer mismatch caused by the phase transition occurring during the cooling process leads to a decrease in FF, resulting in a significant decline in cell efficiency within the corresponding temperature ranges.

CONCLUSIONS

In conclusion, we have established a correlation between the optical behavior of MAPbI3- and FAPbI3-based perovskite films and the corresponding photovoltaic performance changes in low temperatures. With the decreasing temperature, the optical and PL properties evolved. Both the perovskite films show phase transitions at specific low temperatures, which further influenced the photovoltaic performance of PSCs. Briefly, the MAPbI3-based and FAPbI3-based PSCs exhibit notable efficiency improvements of 5% (from 16.46% to 17.21%) and 20% (from 16.91% to 20.28%), respectively, at temperatures higher than 210 K for FAPbI3 and 250 K for MAPbI3, primarily attributed to the enhanced VOC. Further decreasing the temperature impaired the photovoltaic performance owing to the decreased FF. By combining the conventional diode model simulation and optical measurement results, one can infer that the diminishing efficiency would be attributed to the change of crystal structure, even phase transition, at extremely low temperatures. Our findings not only hold promise for applying PSCs in more low-temperature scenarios but also show some instructions to improve the photovoltaic performance of PSCs in extremely low temperatures.

DECLARATIONS

Authors’ contributions

Conceived the idea of this study and designed the experiments, fabricated the perovskite films and solar cells, performed the material synthesis, contributed to the SEM, UV-vis and PL characterization, tested the photovoltaic performances: Xu Y, Wu Z

Analyzed the data and wrote the paper: Xu Y, Wu Z, Zhang Z

Supervised this project: Lin H, Li X

Availability of data and materials

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

Financial support and sponsorship

This work was financially supported by the National Natural Science Foundation of China (NSFC52072207, 62175204), Science and Technology Project of Fujian Province (No. 2021H6018).

Conflicts of interest

All authors declared that there are no conflicts of interest.

Ethical approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Copyright

© The Author(s) 2024.

Supplementary Materials

REFERENCES

1. Zhao D, Yu Y, Wang C, et al. Low-bandgap mixed tin-lead iodide perovskite absorbers with long carrier lifetimes for all-perovskite tandem solar cells. Nat Energy 2017;2:17018.

2. Weber D. CH3NH3PbX3, ein Pb(II)-system mit kubischer perowskitstruktur/CH3NH3PbX3, a Pb(II)-system with cubic perovskite structure. Zeitschrift für Naturforschung B 1978;33:1443-5.

3. Min H, Lee DY, Kim J, et al. Perovskite solar cells with atomically coherent interlayers on SnO- electrodes. Nature 2021;598:444-50.

4. Xia Y, Zhu M, Qin L, et al. Organic-inorganic hybrid quasi-2D perovskites incorporated with fluorinated additives for efficient and stable four-terminal tandem solar cells. Energy Mater 2023;3:300004.

5. Lu YN, Zhong JX, Yu Y, et al. Constructing an n/n+ homojunction in a monolithic perovskite film for boosting charge collection in inverted perovskite photovoltaics. Energy Environ Sci 2021;14:4048-58.

6. Sasaki K, Agui T, Nakaido K, Takahashi N, Onitsuka R, Takamoto T. Development of InGaP/GaAs/InGaAs inverted triple junction concentrator solar cells. AIP Conf Proc 2013;1556:22-5.

7. Haase F, Hollemann C, Schäfer S, et al. Laser contact openings for local poly-Si-metal contacts enabling 26.1%-efficient POLO-IBC solar cells. Sol Energy Mater Sol Cells 2018;186:184-93.

8. Chen Y, Tan S, Li N, et al. Self-elimination of intrinsic defects improves the low-temperature performance of perovskite photovoltaics. Joule 2020;4:1961-76.

9. Izydorczyk W, Waczyński K, Izydorczyk J, et al. Electrical and optical properties of spin-coated SnO2 nanofilms. Mater Sci Pol 2014;32:729-36.

10. D’Innocenzo V, Grancini G, Alcocer MJ, et al. Excitons versus free charges in organo-lead tri-halide perovskites. Nat Commun 2014;5:3586.

11. Kim H, Hunger J, Cánovas E, et al. Direct observation of mode-specific phonon-band gap coupling in methylammonium lead halide perovskites. Nat Commun 2017;8:687.

12. Manoogian A, Woolley JC. Temperature dependence of the energy gap in semiconductors. Can J Phys 1984;62:285-7.

13. Geng W, Zhang L, Zhang Y, Lau W, Liu L. First-principles study of lead iodide perovskite tetragonal and orthorhombic phases for photovoltaics. J Phys Chem C 2014;118:19565-71.

14. Umebayashi T, Asai K, Kondo T, Nakao A. Electronic structures of lead iodide based low-dimensional crystals. Phys Rev B 2003;67:155405.

15. Dai J, Zheng H, Zhu C, Lu J, Xu C. Comparative investigation on temperature-dependent photoluminescence of CH3NH3PbBr3 and CH(NH2)2PbBr3 microstructures. J Mater Chem C 2016;4:4408-13.

16. Baikie T, Fang Y, Kadro JM, et al. Synthesis and crystal chemistry of the hybrid perovskite (CH3NH3)PbI3 for solid-state sensitised solar cell applications. J Mater Chem A 2013;1:5628-41.

17. Francisco-lópez A, Charles B, Alonso MI, et al. Phase diagram of methylammonium/formamidinium lead iodide perovskite solid solutions from temperature-dependent photoluminescence and raman spectroscopies. J Phys Chem C 2020;124:3448-58.

18. Wu K, Bera A, Ma C, et al. Temperature-dependent excitonic photoluminescence of hybrid organometal halide perovskite films. Phys Chem Chem Phys 2014;16:22476-81.

19. Phillips JE, Birkmire RW, Mccandless BE, Meyers PV, Shafarman WN. Polycrystalline heterojunction solar cells: a device perspective. Physica Status Solidi 1996;194:31-9.

20. Hegedus SS, Shafarman WN. Thin-film solar cells: device measurements and analysis. Prog Photovolt 2004;12:155-76.

21. Li Y, Zhuang D, Zhao M, et al. Study on the performance of oxygen-rich Zn(O,S) buffers fabricated by sputtering deposition and Zn(O,S)/Cu(In,Ga)(S,Se)2 interfaces. ACS Appl Mater Interfaces 2022;14:24435-46.

Cite This Article

Export citation file: BibTeX | RIS

OAE Style

Xu Y, Wu Z, Zhang Z, Li X, Lin H. Evolved photovoltaic performance of MAPbI3 and FAPbI3-based perovskite solar cells in low-temperatures. Energy Mater 2024;4:400034. http://dx.doi.org/10.20517/energymater.2023.86

AMA Style

Xu Y, Wu Z, Zhang Z, Li X, Lin H. Evolved photovoltaic performance of MAPbI3 and FAPbI3-based perovskite solar cells in low-temperatures. Energy Materials. 2024; 4(3): 400034. http://dx.doi.org/10.20517/energymater.2023.86

Chicago/Turabian Style

Xu, Youcheng, Ziyi Wu, Ziling Zhang, Xin Li, Hong Lin. 2024. "Evolved photovoltaic performance of MAPbI3 and FAPbI3-based perovskite solar cells in low-temperatures" Energy Materials. 4, no.3: 400034. http://dx.doi.org/10.20517/energymater.2023.86

ACS Style

Xu, Y.; Wu Z.; Zhang Z.; Li X.; Lin H. Evolved photovoltaic performance of MAPbI3 and FAPbI3-based perovskite solar cells in low-temperatures. Energy Mater. 2024, 4, 400034. http://dx.doi.org/10.20517/energymater.2023.86

About This Article

© The Author(s) 2024. Open Access This article is licensed under a Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, sharing, adaptation, distribution and reproduction in any medium or format, for any purpose, even commercially, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Data & Comments

Data

Views
203
Downloads
27
Citations
0
Comments
0
2

Comments

Comments must be written in English. Spam, offensive content, impersonation, and private information will not be permitted. If any comment is reported and identified as inappropriate content by OAE staff, the comment will be removed without notice. If you have any queries or need any help, please contact us at support@oaepublish.com.

0
Download PDF
Cite This Article 3 clicks
Like This Article 2 likes
Share This Article
Scan the QR code for reading!
See Updates
Contents
Figures
Related
Energy Materials
ISSN 2770-5900 (Online)
Follow Us

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/