This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.
Brought to you by:
Paper

Multilayer evaporation of MAFAPbI3−xClx for the fabrication of efficient and large-scale device perovskite solar cells

, , , and

Published 14 November 2018 © 2018 IOP Publishing Ltd
, , Citation Mohammad Mahdi Tavakoli et al 2019 J. Phys. D: Appl. Phys. 52 034005 DOI 10.1088/1361-6463/aaebf1

0022-3727/52/3/034005

Abstract

FAPbI3 perovskites are excellent candidates for fabrication of perovskite solar cells (PSCs) with high efficiency and stability. However, these perovskites exhibit phase instability problem at room temperature. In this work, to address this challenge we use methylammonium chloride (MACl) as an additive and employed a layer-by-layer thermal evaporation technique to fabricate high-quality perovskite films on a large scale of 25 cm2. The optimized perovskite films show high crystallinity with large grains in the µm-range and reveals phase stability due to the presence of MACl after the annealing process. Finally, we achieved PSCs with 17.7% and 15.9% for active areas of 0.1 cm2 and 0.8 cm2, respectively.

Export citation and abstract BibTeX RIS

Introduction

Hybrid organic–inorganic perovskites have attracted much interest due to their excellent properties such as high absorption, high carrier mobilities and versatile band gap tuneability. Thus, perovskites solar cells (PSCs) emerged as ideal candidates for high performance solar cells with a certified efficiency up to 23.3% [16]. In addition, PSCs can be fabricated with relative ease using evaporation and/or solution-based methods. These methods have the potential for upscalable and cost-effective photovoltaic modules.

Currently, the two dominant fabrication methods are vacuum techniques and processing from solution [712]. Thermal evaporation with one or two step process has the advantage of producing high quality, pinhole-free and uniform perovskite films that are solvent-free, substrate-independent, reproducible, scalable and use more purified precursors during the evaporation process (due to the heating of the precursor before the deposition) [1318]. However, as often reported in literature, thermal evaporation of organic compounds such as the volatile methylammonium iodide (MAI) is challenging and the precise thickness-control is frequently an obstacle [1925]. Previously, we proposed a new evaporation method using a layer-by-layer deposition technique for the fabrication of efficient PSC devices with highly crystalline perovskite films [26].

Formamidinium lead triiodide (FAPbI3) has advantages for achieving highly efficient PSCs because FA is thermally more stable than MA and also has a more red-shifted band gap towards the Shockley–Queisser limit [27, 28]. However, this perovskite suffers from phase instability at room temperature showing a photoinactive 'yellow phase' instead of the desired 'black phase'. Compositional engineering is a key solution to tackle this problem. Adding either A-cations, such as MA and Cs, or X-halides, such as Cl and Br, the black phase of FAPbI3 can be stabilized at room temperature by increasing the entropy of the FAPbI3 system [2931]. In this regard, Isikgor et al [32] employed PbCl2 as a source of chlorine and fabricated mixed MA/FA cation perovskite using a solution method. Based on this composition, they improved the perovskite crystallinity, resulting in a PCE of 18.14%. As shown by Yu et al [33], the presence of extra ${\rm C}{{{\rm H}}_{3}}{\rm NH}_{3}^{+}$ in the composition can slow down the perovskite formation and improve the crystallinity during annealing and thus the role of Cl is to facilitate removing the extra ${\rm C}{{{\rm H}}_{3}}{\rm NH}_{3}^{+}$ at low annealing temperature. In other related works, Jiang et al [34, 35] reported a two-step solution process for fabrication PSCs and they added MACl to the perovskite composition in the second step and improved the crystallinity as well as PCE of device over 20%.

In this work, we demonstrate a layer-by-layer thermal evaporation technique for the deposition of FAPbI3 perovskite. We show a stabilized black phase at room temperature by inserting a thin layer of MACl. We optimized the thickness of MACl and the annealing temperature for the perovskite film which has a grain size of up to 2 µm and very little surface roughness of 14 nm. Based on this modification, PSCs with PCEs of 17.7% and 15.9% were achieved for active areas of 0.1 and 0.8 cm2, respectively.

Results and discussions

Figure 1 shows the schematic of our fabrication process using the layer-by-layer thermal evaporation technique to produce phase-stable FAPbI3 perovskite by inserting a thin MACl layer (figure 1(a)). We deposited each precursor, i.e. PbI2 and FAI/MACl in ten sequential steps. Note that MACl was deposited in the middle of the multilayers. This technique enables the reaction of the individual precursors to form a high-quality perovskite film and also gives us more freedom to tune the perovskite composition by adjusting the amount and thickness of the interlayers [26]. The photograph of perovskite films on TiO2-coated FTO glass with different scales is shown in figure S1 (stacks.iop.org/JPhysD/52/034005/mmedia). Figure 2 shows the characterization results of the optimized perovskite film deposited by this technique. The top-view scanning electron microscopy (SEM) image shows high quality and a large grain size up to 2 µm (figure 2(a)). Energy dispersive x-ray spectroscopy (EDAX) elemental mapping reveals a uniform distribution of all the chemical elements as well as the presence of trace amounts of Cl after annealing at the optimal 130 °C (figures S2 and S3). In order to further study the role of MACl, the SEM images of perovskite film fabricated by MACl on PbI2 and FAI on PbI2 after annealing are shown in figure S4. As seen, the perovskite with MACl shows larger grain size as compared to its counterpart. Figure 2(b) shows a top-view atomic force microscopy (AFM) image of the optimized perovskite film. The film has a roughness of 14  ±  5 nm which could stem from the presence of MACl during annealing. A 3D AFM image of the corresponding film depicted in figure S5 clearly confirms the surface roughness. Figure 2(c) shows the powder x-ray diffraction (pXRD) pattern with typical peaks at 14.2°, 28.5°, and 43.2° corresponding to (1 1 0), (2 2 0), and (3 3 0) lattice planes, respectively [36]. This indicates that the film has a preferential orientation along with the (1 1 0) direction. The optical UV absorption and photoluminescence (PL) spectra of the perovskite film are shown in figure 2(d) showing a band gap of 1.57 eV. In addition, the UV-visible spectra of the perovskite film without and with MACl were measured as shown in figure S6. The results indicate that there is a slight blue-shift in the spectrum of the perovskite film by adding MACl due to the presence of chlorine atoms in the perovskite composition.

Figure 1.

Figure 1. Schematic of perovskite film fabrication using layer-by layer thermal evaporation. (a) The sequence of each PbI2 (yellow), FAI (purple), and MACl (blue) layer. (b) Resulting perovskite film with large grains after annealing.

Standard image High-resolution image
Figure 2.

Figure 2. Top-view SEM image (a) and AFM image (b) of optimized perovskite film indicating the film crystallinity. (c) Powder x-ray diffraction pattern, (d) UV and PL spectra of the corresponding perovskite film.

Standard image High-resolution image

The main optimization parameter in this study was the thickness of the MACl layer and the annealing temperature. In a first step, the MACl thickness was tuned from 40, 50, 60, 70 to 80 nm. Figure 3(a) shows the device efficiency as a function of the MACl thickness showing optimal device performance for a 60 nm-thick MACl layer. Higher thicknesses than 60 nm reduces the PCE gradually coinciding with the increased blue-shift in the absorption (that accordingly lowers the current). Additionally, we have optimized the amount of chlorine inside the composition by controlling the annealing temperature. As reported in the literature [12, 37], by increasing the annealing temperature beyond the typical 100 °C, the chlorine starts leaving the perovskite film in the form of MACl. Thus, by controlling the annealing temperature, we can tune the amount of the remaining Cl inside the perovskite composition. In this case, we changed the annealing temperature from 100, 110, 120, 130, 140 to 150 °C. Figure 3(b) shows the variation of PCE versus the annealing temperature. Our results show that 130 °C is the optimal annealing condition. Above 130 °C, the PCE is reduced slowly due to less amount of Cl in the lattice and poorer phase stability.

Figure 3.

Figure 3. Variation of PSC device performance as a function of (a) thickness of MACl layer and (b) the annealing temperature (100, 110, 120, 130, 140, 150 °C).

Standard image High-resolution image

In order to further study the quality of the perovskite films, PSCs with a small (0.1 cm2) and a large (0.8 cm2) area were fabricated. Figure 4(a) illustrates the cross-sectional view SEM image of the PSC revealing the FTO-glass, TiO2 electron transporting layer (ETL), perovskite absorber, spiro-OMeTAD as a hole transporting layer (HTL), and the gold electrode. The current density–voltage (JV) curves of the best performing PSCs with small and large active areas are shown in figure 4(b). The figure of merits for these devices under forward and reverse scans are summarized in table 1. The optimized small area PSC resulted in a short circuit current density of (Jsc) of 22.7 mA cm−2, a Voc of 1040 mV, a fill factor (FF) of 75%, and a PCE of 17.7% at reverse bias. The device with large area showed slightly lower photovoltaic parameters as compared to small area one. As seen, the device parameters were: Jsc of 21.94 mA cm−2, a Voc of 1020 mV, a FF of 71% and a PCE of 15.9% under reverse scan. This indicates the potential of the multilayer vacuum technique for large-scale device. In contrast, figure S7 shows the JV curves of a MAPbI3 PSC measured under reverse and forward scans. As can be seen, the PCE is 16.2% which is lower than the MAFAPbI3−xClx perovskite device due to lower Voc and FF. Moreover, the hysteresis indexes of small area and large area devices were calculated by the following formula: hysteresis index  =  (PCEbackward  −  PCEforward)/PCEforward)  ×  100 and obtained to be 2.31% and 2.92%, respectively. These values are smaller than the hysteresis index of the MAPbI3 device with 4.51% (figure S7), maybe due to the larger grain size and better crystallinity. The statistical photovoltaic data for the corresponding devices are shown in figure S8. It can be observed that the average values of the photovoltaic parameters for large scale devices are lower than that found in small area devices. This shortening is not related to the quality of evaporated perovskite film because the performance of the large-scale device in our work has been limited by spin-coating process of TiO2 ETL and spiro HTL.

Table 1. Figure of merits for champion devices under forward and backward scan directions.

Sample Voc (mV) Jsc (mA cm−2) FF (%) PCE (%) Hysteresis index (%)
Small area-forward 1035 22.6 74 17.3 2.31
Small area-backward 1040 22.7 75 17.7
Large area-forward 1010 21.8 70.3 15.9 2.92
Large area-backward 1020 21.94 71 15.45
Figure 4.

Figure 4. (a) Cross-sectional view of PSC device with optimized perovskite composition and annealing temperature. JV curves (b) and their corresponding EQE spectra (c) of the best performing PSC devices. (d) Shelf life stability test of PSC devices under 30% relative humidity for two months.

Standard image High-resolution image

In addition, the external quantum efficiency (EQE) of the corresponding devices has been measured to further confirm the Jsc values as demonstrated in figure 4(c). The EQE response for large area PSC is slightly reduced compared to the small area PSC resulting overall in a lower Jsc, which is still in good agreement with the JV measurements. The integrated current densities calculated from EQE curves were 21.75 mA cm−2 and 20.91 mA cm−2 for small and large area devices, respectively. Since the stability of PSCs is a big challenge for commercialization, we have tested the shelf life stability of our device as well. Figure 4(d) demonstrates the PCE, measured over 60 d, kept in dry box with 30% relative humidity condition. The result shows that the device retains ~97% of its initial PCE value after 60 d.

Conclusions

In summary, FAPbI3 perovskite films have been fabricated through a layer-by-layer thermal evaporation technique and stabilized by inserting an optimum thickness of an MACl layer. After optimization of the annealing temperature, we obtained a perovskite film with high quality, large grain size up to 2 µm, and low roughness. Based on this fabrication technique and optimization processes, devices with PCEs of 17.7% and 15.9% were achieved for 0.1 and 0.8 cm2 active areas, respectively. Our results indicate the potential of our fabrication process and optimized composition as an alternative for commercialization of PSCs.

Experimental section

Device fabrication

The FTO glasses (Hartford Glass, USA) were etched by zinc powder and diluted HCl solution. The substrates were cleaned using ultra-sonic in the following bathes for 20 min in each step: 3 vol% Triton X-100-containing deionized (DI) water (Millipore, 18 MΩ cm), DI water, acetone, and isopropanol. Then, a 40 nm-thick TiO2 compact layer was spin-coated on the FTO glass at 3000 rpm for 40 s using a precursor solution of titanium diisopropoxide bis(acetylacetonate) in ethanol (0.15 mM), followed by two-step annealing at 150 °C for 20 min and 500 °C for 30 min in air. Afterward, the TiO2-coated FTO glasses were immersed in an aqueous solution of 40 mM TiCl4 (Aldrich) at 70 °C for 30 min and annealed at 500 °C for 30 min.

The perovskite film was deposited on substrates through a layer-by-layer thermal evaporation technique in ten steps. Three quartz crucibles surrounded by two tungsten wire heaters have been loaded by 1 g of lead iodide (PbI2), 0.5 g of formamidinium iodide (FAI), and 0.1 g methylammonium chloride (MACl) powders. The crucibles together with tungsten heaters were located in an evaporator chamber. Then, the substrates were mounted on chamber stage 20 cm far from the sources. The evaporation of each precursor was performed in a vacuum of 4  ×  10−6 mbar. The deposition rates for PbI2, FAI, and MACl were 0.1–0.2 nm s−1, 0.6–1 nm s−1, and 0.6–1 nm s−1, respectively (table S1 show the evaporation parameters of each source). During the evaporation the crystal lifetime was between 80% and 98%. The thickness of each layer was controlled by a quartz sensor and calibrated using Alpha-Step 200 (Tencor). The crystal was located above the shutter. The deposition of all precursors was performed in ten steps, i.e. five steps for PbI2, four steps for FAI and one step for MACl. The thickness of each PbI2, FAI layers were fixed to 24 nm and 68 nm, respectively. In case of MACl, the thickness was changed for purpose of optimization (40, 50, 60, 70 and 80 nm). In this process, after deposition of each PbI2 layer, a layer of FAI was deposited to facilitate the perovskite formation. We inserted MACl in the middle of perovskite film for having a more uniform composition.

The HTL was prepared by dissolving 80 mg spiro-OMeTAD in 1 ml chlorobenzene (CB), followed by adding 17.5 µl of Li-bis(trifluoromethanesulfonyl)imide (Li-TFSI)/acetonitrile (500 mg ml−1), and 28.5 µl of 4-tert-butylpyridine (tBP). The solution was spin-coated at 4000 rpm for 40 s. Finally, a 100 nm-thick gold electrode was deposited on spiro HTL using thermal evaporation (0.08 nm s−1) to complete the device.

Film characterization

Field-emission SEM equipped with EDAX (FESEM, JEOL-7100F) and atomic force microscopy (AFM, a Veeco Dimension 3100) were employed to study the crystallinity and morphology of perovskite films. XRD (Bruker D8 x-ray diffractometer) with Cu Kα radiation was used to study the crystal structure of perovskite. Varian Cary 500 spectrometer and Edinburgh Instruments FLS920P fluorescence spectrometer were used to measure the UV visible and photoluminescence spectra of perovskite film.

Device measurement

All devices were measured using a solar spectrum calibrated under AM 1.5G condition. The light intensity was provided using an Abet Class AAB Sun 2000 simulator and fixed to 100 mW cm−2 after calibration with a KG5-filtered Si reference cell. A 2400 Series SourceMeter (Keithley, USA) instrument was used to record the current density–voltage (JV) curves. The voltage range was  −0.2 V to  +1.2 V. The devices were measured with a step size of 20 mV and a delay time of 150 ms at each point. For EQE measurements, a constant white light bias of nearly 5 mW cm−2 and an Oriel QE-PV-SI (Newport Corporation) were used.

Acknowledgments

M M T wants to thank school of engineering at Hong Kong University of Science and Technology for their support.

Please wait… references are loading.
10.1088/1361-6463/aaebf1