Next Article in Journal
Correction: Yan, X., et al. Study on Utilization of Carboxyl Group Decorated Carbon Nanotubes and Carbonation Reaction for Improving Strengths and Microstructures of Cement Paste. Nanomaterials 2016, 6, 153
Next Article in Special Issue
Non-Enzymatic Glucose Sensor Composed of Carbon-Coated Nano-Zinc Oxide
Previous Article in Journal
Effect of Continuous Multi-Walled Carbon Nanotubes on Thermal and Mechanical Properties of Flexible Composite Film
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Study of Inverted-Type Perovskite Solar Cells with Various Composition Ratios of (FAPbI3)1−x(MAPbBr3)x

Department of Electro-Optical Engineering, National Taipei University of Technology, No. 1, Sec. 3, Chung-Hsiao E. Rd., Taipei 10608, Taiwan
*
Author to whom correspondence should be addressed.
Nanomaterials 2016, 6(10), 183; https://doi.org/10.3390/nano6100183
Submission received: 5 September 2016 / Revised: 24 September 2016 / Accepted: 10 October 2016 / Published: 13 October 2016

Abstract

:
This work presents mixed (FAPbI3)1−x(MAPbBr3)x perovskite films with various composition ratios, x (x = 0–1), which are formed using the spin coating method. The structural, optical, and electronic behaviors of the mixed (FAPbI3)1−x(MAPbBr3)x perovskite films are discussed. A device with structure glass/indium tin oxide (ITO)/poly(3,4-ethylenedioxythiophene) polystyrene sulfonate (PEDOT:PSS)/mixed perovskite/C60/BCP/Ag was fabricated. The mixed perovskite film was an active light-harvesting layer. PEDOT:PSS was a hole transporting layer between the ITO and perovskite. Both C60 and bathocuproine (BCP) were electron transporting layers. MAPbBr3 was added to FAPbI3 with a composition ratio of x = 0.2, stabilizing the perovskite phase, which exhibited a uniform and dense morphology. The optimal device exhibited band matching with C60, resulting in a low series resistance (Rsh) and a high fill factor (FF). Therefore, the device with composition (FAPbI3)1−x(MAPbBr3)x and x = 0.2 exhibited outstanding performance.

1. Introduction

Organometal halide perovskite solar cells have been intensively investigated owing to their high power conversion efficiency and fabrication in solution at low temperatures. Perovskite solar cells with an efficiency of over 20% have been fabricated [1,2,3,4]. The excellent performance of organometal halide perovskite solar cells has two causes: a small bandgap and a large exciton diffusion length. The low absorption bandgap (Eg of CH3NH3PbI3 (MAPbI3) ~1.5 eV) of organometal halide perovskite can harvest most wavelengths of incident sunlight. The long exciton diffusion length increases the thickness of active light-harvesting layers and ensures efficient carrier generation [5,6]. Moreover, the conventional structure of perovskite solar cells use TiO2 for electron transport, but the TiO2 needs to be processed using high temperatures (500–600 °C), and many applications are limited. Inverted-type perovskite solar cells have been developed for low-temperature process and low hysteresis [7].
Interestingly, the compositional engineering of perovskite materials has been extensively utilized to adjust their bandgap and structural properties for use in efficient perovskite solar cells. CH3NH3PbI3−xBrx (x = 0.1–0.15), as an absorbing layer, has been reported to improve the open voltage of photovoltaic devices [8]. CH3NH3PbI3−xClx has been used to increase the exciton diffusion length to improve device performance [9,10,11,12]. HC(NH2)2PbI3 (FAPbI3) [13,14,15,16] can reduce the optical bandgap (Eg ~1.48 eV), with an absorption edge of 840 nm, allowing photons to be absorbed over a broader solar spectrum. Accordingly, FAPbI3 absorbs more light than MAPbI3. Another advantage of FAPbI3 is its thermal stability [17]. It can be processed at a higher temperature than can MAPbI3. The typical annealing process temperature is approximately 130–170 °C. In particular, Jeon et al. explained that (FAPbI3)1−x(MAPbBr3)x can provide a greater balance between electron transport and the hole transport in cells, enabling highly efficient perovskite solar cells to be formed using a regular TiO2 mesoscopic structure (>20%) [4]. However, the origin of their favorable performances and their fabrication process are not yet fully understood.
In this work, solution-processed (FAPbI3)1−x(MAPbBr3)x perovskites were prepared on poly(3,4-ethylenedioxythiophene) polystyrene sulfonate (PEDOT:PSS)-coated indium tin oxide (ITO) substrates for use in inverted perovskite solar cells. The optical, structural, and surface properties of the (FAPbI3)1−x(MAPbBr3)x perovskite films were studied as functions of the composition ratio (x = 0–1). The dependence of cell performance and the properties of the perovskite films is discussed.

2. Materials and Methods

In this work, a PEDOT:PSS (AI 4083) was spin-coated on a pre-cleaned ITO substrate at 5000 rpm for 30 s. Thereafter, the film was annealed at 120 °C for 10 min. The perovskite layer was deposited as in our previous investigation [18]. HC(NH2)2I (FAI), PbI2, CH3NH3Br (MABr), and PbBr2 were dissolved in 1 mL of cosolvent, which comprised dimethyl sulfoxide (DMSO) and γ-butyrolactone (GBL) (volume ratio = 1:1), to form perovskite precursor solutions. The mole ratios of FAPbI3 to MAPbBr3 in the mixed perovskite varied from 0 to 1. The concentrations of each precursor were 1.2 M. For example, 0.96 mmol of FAPbI3 and 0.24 mmol of MAPbBr3 (i.e., 165 mg of FAI, 27 mg of MABr, 443 mg of PbI2, and 88 mg of PbBr2) were dissolved in the mixing solvent (1 mL) as a (FAPbI3)0.8(MAPbBr3)0.2 precursor solution. The perovskite precursor solutions were then coated onto the PEDOT:PSS/ITO substrate in two consecutive spin-coating steps, at 1000 rpm and 5000 rpm for 10 s and 20 s, respectively, in a glove box that was filled with highly pure nitrogen (>99.999%). The wet spinning film was quenched by dropping 50 μL of anhydrous toluene at 17 s. After spin coating, the film was annealed at 100 °C for 10 min. Subsequently, C60, Bathocuproine (BCP), and a silver (Ag) electrode were deposited with thicknesses of 50, 5, and 100 nm, respectively, using a thermal evaporator. The sample was covered with a shadow mask to define an active area of 0.5 cm × 0.2 cm during C60/BCP/Ag deposition. Figure 1a schematically depicts the complete structure.

Material and Device Measurement

The crystalline microstructures of the films were determined using a PAN analytical X’Pert Pro DY2840 X-ray diffractometer (PANalytical, Naerum, Denmark) with CuKα radiation (λ = 0.1541 nm). A field-emission scanning electron microscope (GeminiSEM, ZEISS, Oberkochen, Germany) was used to observe the surface morphology of the cells. Photoluminescence (PL) and absorption spectra were measured using a fluorescence spectrophotometer (Hitachi F-7000) and a UV/VIS/NIR spectrophotometer (Hitachi U-4100 spectrometers) (Hitachi High-Technologies Co., Tokyo, Japan), respectively. Current density–voltage (J-V) characteristics were measured using a Keithley 2420 programmable source meter (Keithley, Cleveland, OH, USA) under illumination by a 1000 W xenon lamp. The forward scan rate was 0.1 V/s.

3. Results and Discussion

Mixed perovskite film can be flexibly modified by changing the concentration ratio of the precursors [4,19]. The lowest unoccupied molecular orbitals (LUMOs) of FAPbI3, MAPbBr3, and C60 are −4.0, −3.6, and −3.9 eV, respectively [20,21]. The bandgap of the mixed (FAPbI3)1−x(MAPbBr3)x film is between that of the FAPbI3 film and that of the MAPbBr3 film, from −4.0 to −3.6. To optimize the band matching with C60, the composition ratio x of FAPbI3 to MAPbBr3 in the mixed (FAPbI3)1−x(MAPbBr3)x perovskite is approximately 0.2 because the LUMO of the (FAPbI3)0.75(MAPbBr3)0.25 was determined by the interpolation to be −3.9, as shown in Figure 1b. Therefore, the device with the mixed (FAPbI3)0.75(MAPbBr3)0.25 perovskite film should have the lowest series resistance (Rsh).
Figure 2 displays a high-resolution image of the cross-section of the obtained perovskite solar cell configuration, which clearly shows the presence of the layers ITO (200 nm), PEDOT:PSS (~50 nm), perovskite (~250 nm), C60 (~60 nm), and BCP (~10 nm). The grain size of the perovskite is approximately 200 nm, as presented in Figure 2. Numerous voids (indicated by red arrows) between grain boundaries were observed. These are characteristic of mixed (FAPbI3)1−x(MAPbBr3)x perovskite films and may be attributed to the supersaturation nucleation and dynamic growth mechanism [22].
Figure 3a shows the X-ray diffraction (XRD) patterns of (FAPbI3)1−x(MAPbBr3)x perovskite films after thermal annealing at various temperatures. The spectrum of the MAPbBr3 film includes three main diffraction peaks at 14.04°, which correspond to the δ-FAPbI3, PbI2, and α-FAPbI3 phases, respectively. As the value of x in the (FAPbI3)1−x(MAPbBr3)x perovskite films increases, the position of the α-FAPbI3 phase peak shifts considerably to a high degree of diffraction, and the δ-FAPbI3 and PbI2 phase peaks disappear. The coexistence of the two FAPbI3 and PbI2 phases can be observed in the (FAPbI3)1−x(MAPbBr3)x perovskite layers with FAPbI3. The spectrum of the MAPbBr3 film includes one diffraction peak at 15.03°, which corresponds to the (100) phase. As presented in Figure 3b, the peak position increases almost linearly with x, revealing that the crystalline FAPbI3 and MAPbBr3 are homogeneous.
Figure 4a presents the room-temperature PL spectra of (FAPbI3)1−x(MAPbBr3)x films with various composition ratios that were deposited on glass substrates. The PL peak shifts nonlinearly from 804 to 533 nm as the composition ratio x is increased from 0 to 1, as displayed in Figure 4a. The bandgaps of MAPbBr3 and FAPbI3 are approximately 2.3 and 1.5 eV, respectively, and correspond to wavelengths of around 540 and 820 nm, respectively. Therefore, bandgap values and PL results match. The bandgap of the mixed (FAPbI3)1−x(MAPbBr3)x perovskite films is calculated from the PL spectra, as shown in Figure 4b. The bandgap over the entire range of the (FAPbI3)1−x(MAPbBr3)x perovskite films can be estimated from the PL spectra. Fitting the PL spectra at room temperature yields the following expression for the bandgap, Eg:
Eg(x) = 1.5 + 0.2x3 + 0.58x6.
The expression is a sixth-order polynomial, rather than the traditional second-order polynomial for compound semiconductors, revealing that the bandgap of the mixed (FAPbI3)1−x(MAPbBr3)x perovskite films is extremely sensitive to the composition of the mixed (FAPbI3)1−x(MAPbBr3)x perovskite films when the concentration of the FAPbI3 is low.
Figure 5 plots the current density as a function of the voltage (J-V) of solar cells that are based on (FAPbI3)1−x(MAPbBr3)x films with various composition ratios. Table 1 presents the power conversion efficiency (Eff), short-circuit current density (Jsc), open-circuit voltage (Voc), and fill factor (FF) of the (FAPbI3)1−x(MAPbBr3)x solar cells. The bandgap of the (FAPbI3)1−x(MAPbBr3)x film is reduced as the proportion of (MAPbBr3) in the (FAPbI3)1−x(MAPbBr3)x films increases. The power conversion efficiency increases with x in the (FAPbI3)1−x(MAPbBr3)x films because Jsc increases with the strength of absorption and the amount of α-FAPbI3 formed. However, the power conversion efficiency decreases as more MAPbBr3 is formed owing to a reduction in the photocurrent and series resistance (Rsh). The optimal device, with a (FAPbI3)0.8(MAPbBr3)0.2 perovskite film, exhibited outstanding performance, where Jsc = 20.6 mA/cm2, Voc = 0.88 V, FF = 65.9%, and Eff = 12.0%. MAPbBr3 was added to FAPbI3 where x = 0.2 to stabilize the perovskite phase with a uniform and dense morphology [4]. Therefore, the device with the (FAPbI3)0.8(MAPbBr3)0.2 perovskite film exhibited the lowest series resistance (Rsh), the best FF, and therefore the best performance.
The bandgap of the (FAPbI3)1−x(MAPbBr3)x film is reduced as the amount of MAPbBr3 in the (FAPbI3)1−x(MAPbBr3)x film increases. Therefore, by increasing the composition ratio x, the (FAPbI3)1−x(MAPbBr3)x film increases the LUMO of the (FAPbI3)1−x(MAPbBr3)x film to match that of the C60 layer and increases the energy barrier to the transportation of electrons, resulting in a low FF. The value of Voc is positively correlated with the difference between the highest occupied molecular orbital (HOMO) of the (FAPbI3)1−x(MAPbBr3)x film and the LUMO of the C60 layer (Figure 1b) [23]. Therefore, Voc is determined by the increase in the HOMO level of the (MAPbBr3)x(FAPbI3)1−x film. The bandgap of the (FAPbI3)1−x(MAPbBr3)x film increases with x, increasing Voc. Additionally, the LUMO level of the (MAPbBr3)x(FAPbI3)1−x film is lower than that of C60, so an energy barrier is formed, lowering FF. Additionally, the photo-generated current density declines as the proportion of MAPbBr3 in the (FAPbI3)1−x(MAPbBr3)x films increases because the absorption range is reduced.

4. Conclusions

In summary, this work presents mixed (FAPbI3)1-x(MAPbBr3)x perovskite films with various composition ratios (x = 0 to 1), formed using the spin coating method. The bandgap over the entire range of the (FAPbI3)1−x(MAPbBr3)x perovskite films can be estimated from the PL spectra. Fitting the PL spectra yields a sixth-order polynomial for the bandgap of the mixed (FAPbI3)1−x(MAPbBr3)x perovskite films. This result reveals that the bandgap of the mixed (FAPbI3)1−x(MAPbBr3)x perovskite films is extremely sensitive to the composition of the mixed (FAPbI3)1−x(MAPbBr3)x perovskite films when the concentration of the FAPbI3 is low. The optimal device uses (FAPbI3)1−x(MAPbBr3)x with x = 0.2 and exhibited outstanding performance, where short-circuit current density Jsc = 20.6 mA/cm2, open-circuit voltage Voc = 0.88 V, fill factor FF = 65.9%, and power conversion efficiency Eff = 12.0%, perhaps because the addition of MAPbBr3 to FAPbI3 where x = 0.2 stabilized the perovskite phase with a uniform and dense morphology. The optimum device exhibits band matching with C60, resulting in a low series resistance (Rsh) and high FF.

Acknowledgments

Financial support of this paper was provided by the Ministry of Science and Technology of the Republic of China under Contract no. MOST 105-2221-E-027-055.

Author Contributions

Lung-Chien Chen wrote the paper, designed the experiments, and analyzed the data. Zong-Liang Tseng and Jun-Kai Huang prepared the samples and performed all the measurements. All authors read and approved the final manuscript.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Yang, W.S.; Noh, J.H.; Jeon, N.J.; Kim, Y.C.; Ryu, S.; Seo, J.; Seok, S.I. High-performance photovoltaic perovskite layers fabricated through intramolecular exchange. Science 2015, 348, 1234–1237. [Google Scholar] [CrossRef] [PubMed]
  2. Liu, M.; Johnston, M.B.; Snaith, H.J. Efficient planar heterojunction perovskite solar cells by vapour deposition. Nature 2013, 501, 395–398. [Google Scholar] [CrossRef] [PubMed]
  3. Im, J.-H.; Jang, I.-H.; Pellet, N.; Grätzel, M.; Park, N.-G. Growth of CH3NH3PbI3 cuboids with controlled size for high-efficiency perovskite solar cells. Nat. Nano 2014, 9, 927–932. [Google Scholar] [CrossRef] [PubMed]
  4. Jeon, N.J.; Noh, J.H.; Yang, W.S.; Kim, Y.C.; Ryu, S.; Seo, J.; Seok, S.I. Compositional engineering of perovskite materials for high-performance solar cells. Nature 2015, 517, 476–480. [Google Scholar] [CrossRef] [PubMed]
  5. Xing, G.; Mathews, N.; Sun, S.; Lim, S.S.; Lam, Y.M.; Grätzel, M.; Mhaisalkar, S.; Sum, T.C. Long-Range Balanced Electron- and Hole-Transport Lengths in Organic-Inorganic CH3NH3PbI3. Science 2013, 342, 344–347. [Google Scholar] [CrossRef] [PubMed]
  6. Stranks, S.D.; Eperon, G.E.; Grancini, G.; Menelaou, C.; Alcocer, M.J.P.; Leijtens, T.; Herz, L.M.; Petrozza, A.; Snaith, H.J. Electron-Hole Diffusion Lengths Exceeding 1 Micrometer in an Organometal Trihalide Perovskite Absorber. Science 2013, 342, 341–344. [Google Scholar] [CrossRef] [PubMed]
  7. Chen, L.-C.; Wu, J.-R.; Tseng, Z.-L.; Chen, C.-C.; Chang, S.-H.; Huang, J.-K.; Lee, K.-L.; Cheng, H.-M. Annealing Effect on (FAPbI3)1−x(MAPbBr3)x Perovskite Films in Inverted-Type Perovskite Solar Cells. Materials 2016, 9, 747. [Google Scholar] [CrossRef]
  8. Jeon, N.J.; Noh, J.H.; Kim, Y.C.; Yang, W.S.; Ryu, S.; Seok, S.I. Solvent engineering for high-performance inorganic–organic hybrid perovskite solar cells. Nat. Mater. 2014, 13, 897–903. [Google Scholar] [CrossRef] [PubMed]
  9. Shi, Y.; Xing, Y.; Li, Y.; Dong, Q.; Wang, K.; Du, Y.; Bai, X.; Wang, S.; Chen, Z.; Ma, T. CH3NH3PbI3 and CH3NH3PbI3−xClx in Planar or Mesoporous Perovskite Solar Cells: Comprehensive Insight into the Dependence of Performance on Architecture. J. Phys. Chem. C 2015, 119, 15868–15873. [Google Scholar] [CrossRef]
  10. Cao, C.; Zhang, C.; Yang, J.; Sun, J.; Pang, S.; Wu, H.; Wu, R.; Gao, Y.; Liu, C. Iodine and Chlorine Element Evolution in CH3NH3PbI3−xClx Thin Films for Highly Efficient Planar Heterojunction Perovskite Solar Cells. Chem. Mater. 2016, 28, 2742–2749. [Google Scholar] [CrossRef]
  11. Liu, D.; Wu, L.; Li, C.; Ren, S.; Zhang, J.; Li, W.; Feng, L. Controlling CH3NH3PbI3−xClx Film Morphology with Two-Step Annealing Method for Efficient Hybrid Perovskite Solar Cells. ACS Appl. Mater. Inter. 2015, 7, 16330–16337. [Google Scholar] [CrossRef] [PubMed]
  12. Chen, Q.; Zhou, H.; Fang, Y.; Stieg, A.Z.; Song, T.-B.; Wang, H.-H.; Xu, X.; Liu, Y.; Lu, S.; You, J.; et al. The optoelectronic role of chlorine in CH3NH3PbI3(Cl)-based perovskite solar cells. Nat. Commun. 2015, 6. [Google Scholar] [CrossRef] [PubMed]
  13. Wang, Z.; Zhou, Y.; Pang, S.; Xiao, Z.; Zhang, J.; Chai, W.; Xu, H.; Liu, Z.; Padture, N.P.; Cui, G. Additive-Modulated Evolution of HC(NH2)2PbI3 Black Polymorph for Mesoscopic Perovskite Solar Cells. Chem. Mater. 2015, 27, 7149–7155. [Google Scholar] [CrossRef]
  14. Zhou, Y.; Kwun, J.; Garces, H.F.; Pang, S.; Padture, N.P. Observation of phase-retention behavior of the HC(NH2)2PbI3 black perovskite polymorph upon mesoporous TiO2 scaffolds. Chem. Commun. 2016, 52, 7273–7275. [Google Scholar] [CrossRef] [PubMed]
  15. Song, J.; Hu, W.; Wang, X.-F.; Chen, G.; Tian, W.; Miyasaka, T. HC(NH2)2PbI3 as a thermally stable absorber for efficient ZnO-based perovskite solar cells. J. Mater. Chem. A 2016, 4, 8435–8443. [Google Scholar] [CrossRef]
  16. Lee, J.-W.; Seol, D.-J.; Cho, A.-N.; Park, N.-G. High-Efficiency Perovskite Solar Cells Based on the Black Polymorph of HC(NH2)2PbI3. Adv. Mater. 2014, 26, 4991–4998. [Google Scholar] [CrossRef] [PubMed]
  17. Zhou, Y.; Zhu, K. Perovskite Solar Cells Shine in the “Valley of the Sun”. ACS Energy Lett. 2016, 1, 64–67. [Google Scholar] [CrossRef]
  18. Chen, L.-C.; Chen, C.-C.; Chen, J.-C.; Wu, C.-G. Annealing effects on high-performance CH3NH3PbI3 perovskite solar cells prepared by solution-process. Solar Energy 2015, 122, 1047–1051. [Google Scholar] [CrossRef]
  19. Kulkarni, S.A.; Baikie, T.; Boix, P.P.; Yantara, N.; Mathews, N.; Mhaisalkar, S. Band-gap tuning of lead halide perovskites using a sequential deposition process. J. Mater. Chem. A 2014, 2, 9221–9225. [Google Scholar] [CrossRef]
  20. Aharon, S.; Dymshits, A.; Rotem, A.; Etgar, L. Temperature dependence of hole conductor free formamidinium lead iodide perovskite based solar cells. J. Mater. Chem. A 2015, 3, 9171–9178. [Google Scholar] [CrossRef]
  21. Schueppel, R.; Schmidt, K.; Uhrich, C.; Schulze, K.; Wynands, D.; Brédas, J.L.; Brier, E.; Reinold, E.; Bu, H.B.; Baeuerle, P.; et al. Optimizing organic photovoltaics using tailored heterojunctions: A photoinduced absorption study of oligothiophenes with low band gaps. Phys. Rev. B 2008, 77, 085311. [Google Scholar] [CrossRef]
  22. Zhou, Y.; Game, O.S.; Pang, S.; Padture, N.P. Microstructures of Organometal Trihalide Perovskites for Solar Cells: Their Evolution from Solutions and Characterization. J. Phys. Chem. Lett. 2015, 6, 4827–4839. [Google Scholar] [CrossRef] [PubMed]
  23. Jeng, J.Y.; Chiang, Y.F.; Lee, M.H.; Peng, S.R.; Guo, T.F.; Chen, P.; Wen, T.C. CH3NH3PbI3 perovskite/fullerene planar-heterojunction hybrid solar cells. Adv. Mater. 2013, 25, 3727–3732. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Complete structure and (b) corresponding energy band diagram of the structure.
Figure 1. (a) Complete structure and (b) corresponding energy band diagram of the structure.
Nanomaterials 06 00183 g001
Figure 2. Field emission scanning electron microscope (FESEM) cross-sectional image of device structure.
Figure 2. Field emission scanning electron microscope (FESEM) cross-sectional image of device structure.
Nanomaterials 06 00183 g002
Figure 3. (a) X-ray diffraction (XRD) patterns of (FAPbI3)1−x(MAPbBr3)x perovskite films with various compositions; (b) Relationship between degree of diffraction and composition of (FAPbI3)1−x(MAPbBr3)x.
Figure 3. (a) X-ray diffraction (XRD) patterns of (FAPbI3)1−x(MAPbBr3)x perovskite films with various compositions; (b) Relationship between degree of diffraction and composition of (FAPbI3)1−x(MAPbBr3)x.
Nanomaterials 06 00183 g003
Figure 4. (a) Photoluminescence (PL) spectra of mixed (FAPbI3)1−x(MAPbBr3)x perovskite films with various values of x. (b) Bandgap of mixed (FAPbI3)1−x(MAPbBr3)x perovskite films, estimated from PL spectra.
Figure 4. (a) Photoluminescence (PL) spectra of mixed (FAPbI3)1−x(MAPbBr3)x perovskite films with various values of x. (b) Bandgap of mixed (FAPbI3)1−x(MAPbBr3)x perovskite films, estimated from PL spectra.
Nanomaterials 06 00183 g004
Figure 5. J-V curves of perovskite solar cell (Ag/BCP/C60/(FAPbI3)1−x(MAPbBr3)x/PEDOT:PSS/ITO) obtained under standard 1 sun air mass (AM) 1.5 simulated solar irradiation.
Figure 5. J-V curves of perovskite solar cell (Ag/BCP/C60/(FAPbI3)1−x(MAPbBr3)x/PEDOT:PSS/ITO) obtained under standard 1 sun air mass (AM) 1.5 simulated solar irradiation.
Nanomaterials 06 00183 g005
Table 1. Parameters of solar cells based on perovskite (FAPbI3)1−x(MAPbBr3)x film with various composition ratios.
Table 1. Parameters of solar cells based on perovskite (FAPbI3)1−x(MAPbBr3)x film with various composition ratios.
(FAPbI3)1−x(MAPbBr3)xVoc (V)Jsc (mA/cm2)FF (%)Eff (%)Rsh (Ω)
(FAPbI3)0.6017.344.05.6820.3
(FAPbI3)0.8(MAPbBr3)0.20.8820.665.912.04.6
(FAPbI3)0.6(MAPbBr3)0.40.9017.6352.99.418.5
(FAPbI3)0.4(MAPbBr3)0.60.9011.0151.45.5119.8
(FAPbI3)0.2(MAPbBr3)0.80.958.8649.44.1821.2
(MAPbBr3)11.27.2347.74.1930.6

Share and Cite

MDPI and ACS Style

Chen, L.-C.; Tseng, Z.-L.; Huang, J.-K. A Study of Inverted-Type Perovskite Solar Cells with Various Composition Ratios of (FAPbI3)1−x(MAPbBr3)x. Nanomaterials 2016, 6, 183. https://doi.org/10.3390/nano6100183

AMA Style

Chen L-C, Tseng Z-L, Huang J-K. A Study of Inverted-Type Perovskite Solar Cells with Various Composition Ratios of (FAPbI3)1−x(MAPbBr3)x. Nanomaterials. 2016; 6(10):183. https://doi.org/10.3390/nano6100183

Chicago/Turabian Style

Chen, Lung-Chien, Zong-Liang Tseng, and Jun-Kai Huang. 2016. "A Study of Inverted-Type Perovskite Solar Cells with Various Composition Ratios of (FAPbI3)1−x(MAPbBr3)x" Nanomaterials 6, no. 10: 183. https://doi.org/10.3390/nano6100183

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