Highly efficient bifacial semitransparent perovskite solar cells based on molecular doping of CuSCN hole transport layer
Hou Shixin1, 2, 3, 4, 5, Shi Biao1, 2, 3, 4, 5, Wang Pengyang1, 2, 3, 4, 5, †, Li Yucheng1, 2, 3, 4, 5, Zhang Jie1, 2, 3, 4, 5, Chen Peirun1, 2, 3, 4, 5, Chen Bingbing1, 2, 3, 4, 5, Hou Fuhua1, 2, 3, 4, 5, Huang Qian1, 2, 3, 4, 5, Ding Yi1, 2, 3, 4, 5, Li Yuelong1, 2, 3, 4, 5, Zhang Dekun1, 2, 3, 4, 5, Xu Shengzhi1, 2, 3, 4, 5, Zhao Ying1, 2, 3, 4, 5, Zhang Xiaodan1, 2, 3, 4, 5, ‡
Institute of Photoelectronic Thin Film Devices and Technology of Nankai University, Tianjin 300350, China
Key Laboratory of Photoelectronic Thin Film Devices and Technology of Tianjin, Tianjin 300350, China
Collaborative Innovation Center of Chemical Science and Engineering (Tianjin), Tianjin 300072, China
Renewable Energy Conversion and Storage Center of Nankai University, Tianjin 300072, China
Engineering Research Center of Thin Film Photoelectronic Technology, Ministry of Education, Tianjin 300350, China

 

† Corresponding author. E-mail: pywang@nankai.edu.cn xdzhang@nankai.edu.cn

Project supported by the National Key Research and Development Program of China (Grant No. 2018YFB1500103), the National Natural Science Foundation of China (Grant No. 61674084), the Overseas Expertise Introduction Project for Discipline Innovation of Higher Education of China (Grant No. B16027), and the Science and Technology Project of Tianjin, China (Grant No. 18ZXJMTG00220).

Abstract

Coper thiocyanate (CuSCN) is generally considered as a very hopeful inorganic hole transport material (HTM) in semitransparent perovskite solar cells (ST-PSCs) because of its low parasitic absorption, high inherent stability, and low cost. However, the poor electrical conductivity and low work function of CuSCN lead to the insufficient hole extraction and large open-circuit voltage loss. Here, 2,3,5,6-tetrafluoro-7,7,8,8-tetracyanoquinodimethane (F4TCNQ) is employed to improve conductivity of CuSCN and band alignment at the CuSCN/perovskite (PVK) interface. As a result, the average power conversion efficiency (PCE) of PSCs is boosted by ≈ 11%. In addition, benefiting from the superior transparency of p-type CuSCN HTMs, the prepared bifacial semitransparent n–i–p planar PSCs demonstrate a maximum efficiency of 14.8% and 12.5% by the illumination from the front side and back side, respectively. We believe that this developed CuSCN-based ST-PSCs will promote practical applications in building integrated photovoltaics and tandem solar cells.

PACS: ;88.40.H-;;88.40.hj;
1. Introduction

Expanding application fields and improving power conversion efficiency (PCE) of solar cells have become inescapable issues to further promote the development of photovoltaic (PV) market. Semitransparent solar cells (ST-SCs) provide an attractive choice to address aforementioned problems, and thus receive overwhelming attention. For example, ST-SCs can be utilized not only as color-tunable building integrated PV (BIPV),[1] wearable electronics,[2] and PV vehicles[3] to excite future applications, but also as sub-cell of multi-junction[4] or bifacial PV devices[5] to convert more solar energy into electricity, and therefore improve PCE. Organic–inorganic hybrid perovskite solar cells (PSCs) have emerged as a promising candidate for ST-SCs, because opaque PSCs have achieved PCE of 25.2%[6] with unprecedented speed attributing to its direct band gap, high absorption coefficient, low exciton binding energy, and a long electron–hole diffusion length. Simple device architecture, facile fabrication process, and higher PCE make the planar heterojunction stand out from all PSCs device structures, especially the planar n–i–p device.[79] Different from opaque solar cells, both charge collection efficiency and light transmittance of HTMs are vital to avoid influencing light absorption in perovskite active layer and realize high performance of ST-PSCs.[10] Generally, 2,2’,7,7’-tetrakis (N,N-di-p-Methoxyphenylamine)-9, 9’-spirobifluorene (Spiro-OMeTAD)[11] or poly (triarylamine) (PTAA)[12] are used as HTMs in most efficient PSCs. Unfortunately, both of them are unstable and expensive, which will greatly affect their wide application in the future. Besides, the serious light absorption in short-wavelength range is also harmful to light harvesting of perovskite absorption layer. Therefore, it is urgent and essential to explore alternative HTMs.[13]

Recently, copper thiocyanate (CuSCN) is paid much attention due to its excellent material stability, extremely low price, and compatible HOMO (highest occupied molecular orbited) level with perovskite. And it is noteworthy that the hole mobility of CuSCN of 0.01 cm2 · V−1 · s−1–0.1 cm2 · V−1 · s−1 is remarkably higher than that of common organic HTMs, such as 10−3 cm2 · V−1 · s−1 of PTAA and 4 × 10−5 cm2 · V−1 · s−1 of Spiro-OMeTAD.[14] A report of the PSCs based opaque CuSCN with PCE over 20%[15] confirmed that CuSCN can be employed as excellent HTM. In addition, the low parasitic absorption also allows it as potential candidate of HTMs for ST-PSCs. For instance, Jen and co-authors[16] adopted CuSCN as HTM instead of PEDOT:PSS to prepare ST-PSCs to obtain PCE of 10% and average visible transmittance (AVT) of 25%. Fan and co-authors[17] fabricated bifacial semitransparent n–i–p planar PSCs, which achieved a maximum efficiency of 12.47% and 8.74% by illumination from the front side and back side, respectively. Wang and co-authors[18] used the CuSCN and ITO as the HTMs and back electrode in the PSCs, respectively. Compared with the Spiro-OMeTAD based ST-PSCs, the CuSCN-based ST-PSCs exhibited higher PCE of 14.2% and 13.3% by illumination from the front side and back side, respectively. Despite many efforts have been made in employing CuSCN as HTM, few attentions were focused on the poor electrical conductivity and small work function of CuSCN. Those problems will cause weak built-in field within limited film thickness, and at the same time cause energy level offset at the CuSCN/perovskite interface.[19] Consequently, it can be deduced that there will be a large amount of hole accumulation and electron–hole recombination at the interface of CuSCN/perovskite due to the low holes transfer efficiency of intrinsic CuSCN.

In this work, F4TCNQ was doped into CuSCN to be used as the HTM in CsFAMA based n–i–p planar perovskite solar cells, which optimize the energy level alignment at the I/P interface in order to achieve better hole transport. Ultraviolet photoelectron spectroscopy (UPS) and x-ray photoelectron spectroscopy (XPS) indicated that, after F4TCNQ doping, the energy level positions of CuSCN HTMs changed, in which Fermi level (EF) decreased from −4.82 eV to −4.90 eV and valence band maximum (VBM) −EF decreased from 0.71 eV to 0.68 eV. Therefore, the average conversion efficiency of CsFAMA mixed cation PSCs was improved by about 11%, and a PCE of 16.6% for a champion cell. In addition, for bifacial semitransparent n–i–p planar PSC using CuSCN as HTM, maximum PCE of 14.8% and 12.5% were achieved by illuminated from the front side and the rear side, respectively.

2. Experiments
2.1. Materials

Lead bromide (PbBr2, 99.99%), copper thiocyanate (CuSCN, 99%), cesium iodide (CsI, 99.999%), chlorobenzene (CB, 99.8%), dimethylformamide (DMF, 99.9%), and dimethyl sulfoxide (DMSO, 99.9%) were purchased from Sigma-Aldrich. Tin(IV) oxide colloid precursor (15% in H2O colloidal) was purchased from Alfa Aesar. F4TCNQ (99%) were purchased from materwin. Lead iodide (PbI2, 99.99%), formamidinium iodide (FAI), and methylammonium bromide (MABr) wear purchased from Xi’an Polymer Light Technology Corp. Diethyl sulfide (DES, 98%) was purchased from Shanghai TCI.

2.2. Fabrication of devices

ITO glass substrates were cleaned with detergent, deionized water, acetone, and isopropanol for 30 min each time. The previously cleaned ITO glass substrates were exposed to ultraviolet light at room temperature for 10 min, spin coated SnO2 layer at 4000 rpm for 30 s then annealed at 150 °C for 30 min in ambient air. The CsFAMA mixed precursor solution was prepared in a glove box. Precursor solution of perovskite was prepared by mixing PbI2 (548.62 mg), PbBr2 (57.06 mg), FAI (178.94 mg), MABr (17.41 mg), CsI (27.04 mg) in DMF/DMSO (V/V 780 μl/220 μl). Subsequently, perovskite films were prepared by spin coating in a two-step programme at 1000 rpm and 5000 rpm for 10 s and 30 s. During the second step, 100 μl of chlorobenzene was dropped on the spinning substrate 10 s prior to the end of the programme and finally annealed on a hot plate at 130 °C in a glove box for 20 min.

The F4TCNQ was dissolved in chlorobenzene at the concentration of 1 mg/ml and added it in CuSCN(30 mg/ml, DES) solution at a volume ratio of 1%, 3%, 5%, and 7%, respectively. Doped-CuSCN and undoped-CuSCN layers were spin-coated on the ITO/SnO2/CsFAMA substrates at the speed of 4000 rpm for 30 s. Drying was performed at room temperature for 10 min, and then it was annealed at 80 °C for 5 min to form the CuSCN films. MoOX with thickness of 9 nm was thermally evaporated in high vacuum of 1 × 10−4 Pa, and grew at a slow rate of 0.1 Å/s, while the substrate temperature was kept at room temperature. Then ITO with thickness of 100 nm was deposited at room temperature by DC magnetron sputter deposition. Finally, gold grid electrode with thickness of 100 nm was deposited by thermal evaporation on the top.

2.3. Characterization

X-ray photoelectron spectroscopy (XPS) was carried out on a Thermo VG ESCALAB 250X with Al Ka radiation as excitation source, and used the C 1s neutral carbon peak at 284.8 eV to calibrate the binding energy. Ultraviolet photoelectron spectroscopy (UPS) with a VG Scienta R4000 analyzer and using a monochromatic He I light source (21.22 eV) was used to characterize the molecular energy levels. Secondary electron cutoff (SEC) was observed with a sample bias of −5 V. The method to determine the work function (ϕ) was based on the difference between the photon energy and the binding energy of the secondary cutoff edge. Scanning electron microscope (SEM, FEI NanoSEM650) was used to measure the surface morphology of undoped and doped CuSCN. In addition, surface roughness of undoped and doped CuSCN was characterized by atomic force microscope (AFM, NanoNavi-SPA400). The transmittance and absorbance of undoped and doped CuSCN were characterized with a UV–vis-NIR spectrophotometer (Cary 5000, VARIAN). A fluorescence spectrophotometer (Edinburgh Instruments, FLS 980) was used to record room temperature steady-state photoluminescence (PL) and time-resolved photoluminescence (TRPL). The photocurrent density–voltage (JV) curves of devices were measured under one sun illumination by using a Keithley 2400 digital source meter. The dark current density (ITO/SnO2/PVK/HTM/Au) and trap state density (ITO/HTM/PVK/HTM/Au) were both examined by Keithley 2400 digital source meter under dark condition. External quantum efficiency (EQE) spectrum was obtained without any voltage bias in the case where the EQE system (QEX10, PV Measurement Co, Ltd.) was used in the DC mode.

3. Results and discussion

The device schematic diagram of the undoped and doped CuSCN are illustrated in Fig. 1(a). At present, in order to enhance the characteristics of hole transport layers, lots of strategies were conducted,[2022] such as He and coauthors[23] found that Cu doping NiO results in increased work function and better energy levels alignment with enhanced device performance; Liao and coauthors[24] proved that the incorporation of copper salts achieved a p-type doping with efficient charge transfer complex, which resulted in improved film conductivity and hole mobility in spiro-OMeTAD:copper salts composite films. As a result, the PCE is largely improved from 14.82% to 18.02%. There are many methods to control the carrier concentrations in HTMs, and doping is one of the most effective methods.[25,26] Strong electron acceptor of F4TCNQ can be easily ionized by attracting electrons.[27] With the measurement of UV–vis absorption spectroscopy, the optical bandgap values of undoped CuSCN and doped CuSCN are determined to be 3.66 eV and 3.67 eV, respectively (see Fig. A1 in Appendix A). To explore the mechanism, electronic structures of CuSCN with and without F4TCNQ are probed with UPS.[28] As shown in Fig A2(a), the valance band spectra show that after doping F4TCNQ, the difference between EF and VBM of CuSCN narrows from 0.71 eV to 0.68 eV, indicating the hole concentration is enhanced. From the UPS measurement, the work functions (ϕ) of the undoped and doped CuSCN are calculated to be 4.82 eV and 4.90 eV, respectively (see Fig. A2(b)). The VBM of the doped CuSCN (5.58 eV) went deeper than undoped CuSCN (5.53 eV) as referred to vacuum level. From this result, we can draw the conclusion that the transferring of electrons from CuSCN to F4TCNQ was successful, which is consistent with some reports of F4TCNQ.[29] The negative shifts of EF and VBM effectively reduce the energy level offset between CuSCN and perovskite absorbers and enhance built-in field strength, and hence improve the device performance. From the XPS measurements, the Cu 2p core level peaks have unambiguous shifts about 0.2 eV after doping F4TCNQ (Fig A3), and the core level peaks of S 2p, C 1s, and N 1s also have unambiguous shifts about 0.1 eV–0.4 eV after doping F4TCNQ (Fig. A3) respectively, which further make us confirm that electron transfer from CuSCN to F4TCNQ happens.[30]

Fig. 1. (a) Schematic device architecture. (b) The energy band structure diagram of PSCs.
Fig. A1. The relationship between (αhv)2 and energy of CuSCN and 3% F4TCNQ-doped CuSCN film. The band gap can be determined by tangent of the curve to the base line. It can be found that the band gaps of the CuSCN and 3% F4TCNQ-doped CuSCN are about 3.66 eV and 3.67 eV.
Fig. A2. (a) Valance band spectra of the CuSCN and F4TCNQ-doped CuSCN. (b) UPS spectra of the CuSCN and F4TCNQ-doped CuSCN, measured under −21.22-eV bias. Band bending at CuSCN-perovskite interface of before (c) and after (d) doping F4TCNQ.
Fig. A3. XPS spectra of CuSCN and F4TCNQ-doped CuSCN for the core level of Cu 2p (a), S 2p (b), C 1s (c), and N 1s (d); the core level of Cu 2p1/2 shifted ≈ 0.22 eV and Cu 2p3/2 shifted ≈ 0.27 eV after molecular doped; the core level of S 2p shifted ≈ 0.34 eV after molecular doped; the core level of C 1s shifted ≈ 0.15 eV after molecular doped; the core level of S 2p shifted ≈ 0.26 eV after molecular doped.

Based on the aforementioned problem, the band diagram of the CuSCN/perovskite in the PSCs is shown in Fig. 1(b) and the corresponding band bending at CuSCN-perovskite interface of before and after doping F4TCNQ is shown in Figs. A2(c) and A2(d), where the VBM and EF of CsFAMA perovskite is set at 5.65 eV and at the middle of the band gap, respectively.[23,31] It can be clearly seen that the energy level offset between CuSCN VBM and perovskite VBM for charge transfer is decreased from 0.12 eV to 0.07 eV. Owing to the simultaneous reduction of its work function, the band bending of CuSCN/perovskite interface was enlarged 0.08 eV, doped CuSCN/PVK interface present accelerated charge transfer and reduces holes accumulation, resulting in lower charge recombination and higher charge transport efficiency from perovskite.

Steady-state photoluminescence (PL) and time-resolved photoluminescence (TRPL) decay measurements are used to understand the hole extraction capability and recombination characteristics of the devices with different HTMs. As shown in Fig. 2(a), for the doped CuSCN, a stronger quenching effect is observed, indicating an enhanced hole extraction capability. The measurement of TRPL can detect the time intensity change of radiative processes. The TRPL decay can be ascribed to the perovskite bulk radiative or nonradiative recombination of electrons and holes, charge capture at the surface defects of perovskite surface, or the transfer of hole from perovskite to HTM. The measured TRPL curves are shown in Fig. 2(b). Fitting with exponential, results are shown as equation as follows:

where A1 and A2 are the corresponding decay amplitudes, τ1 is the fast decay time about interface recombination, and τ2 is the slow decay time reflected the bulk recombination, Y0 is a constant for the baseline offset.[32] The obtained parameters for the TRPL data are listed in Table A1 in Appendix A. The better energy level structure makes PL quenching faster, which shows that the ability of carrier extraction is improved. Meanwhile, it is very important to reduce carrier recombination to improve the PCE of device.[33]

Fig. 2. Steady-state (a) and time-resolved (b) photoluminescence of glass/perovskite/CuSCN and glass/perovskite/F4TCNQ-doped CuSCN; SCLC of (c) ITO/NiOx/PVK/F4TCNQ-undoped CuSCN/Au, and (d) ITO/NiOx/PVK/F4TCNQ-doped CuSCN/Au.
Table A1.

Summary of the carrier lifetime from fitting curves of the TRPL measurements.

.

The dark current–voltage (JV) characteristics of hole-only devices based on ITO/NiOx/perovskite/undoped or doped CuSCN/Au were used to evaluate carrier recombination at the interface of CuSCN/Perovskite. The space-charge-limited currents (SCLC) curves is used to measure the trap state density state (ntrap). As shown in Figs. 2(c) and 2(d), the dark JV curve can be divided into three parts. In the case where the applied voltage is higher than the voltage at the first kink point (n > 3), it can be concluded that all the trap states are filled with injected carriers, because the current exhibits a rapid non-linear growth state. The applied voltage at the kink point is the trap-filled limit voltage (VTFL) defined by the trap state density[34]

where d and ε are the thickness and relative dielectric constant of perovskite film, and ε0 is the constant of permittivity in free space. The VTFL is linearly related to the n trap. As shown in Figs. 2(c) and 2(d), the VTFL are 1.36 V and 1.12 V for undoped and doped devices, respectively. Lower VTFL represents the less carrier recombination at the interface of CuSCN/perovskite. This further indicates that the doped CuSCN has a faster charge transfer capability, thereby obtaining faster charge extraction, reducing the accumulation of charge at the interface and charge recombination in the bulk.[35]

The photovoltaic performance of the CsFAMA perovskite PSCs with undoped and doped CuSCN HTMs is analyzed and is shown in Fig. 3. It can be seen that the cell based on doped CuSCN achieves a considerably higher short-circuit current density (Jsc), open-circuit voltage (Voc), and fill factor (FF) and thus higher PCE. The results of the JV curves of the best devices indicate that using F4TCNQ-doped CuSCN as HTMs can significantly improve device performance (Fig. 3(a)). Finally, the champion cell with PCE of 16.6% is obtained, while the undoped device achieves an optimal PCE of 13.5%. Table 1 summarizes all photovoltaic parameters. The EQE spectra are shown in Fig. 3(b). Devices using doped CuSCN as HTMs in the range from 300 nm to 800 nm have higher EQE values compared to reference devices, which indicates that doped CuSCN has stronger charge extraction and transporting capabilities. The photocurrents of the undoped and doped CuSCN based solar cells calculated from the results of EQE are 19.5 mA/cm2 and 20.3 mA/cm2, respectively.

Fig. 3. (a) JV curves from forward-biased (FB) to short-circuit (SC) of the optimal CsFAMA PSCs using CuSCN or F4TCNQ-doped CuSCN as HTMs under 1-sun illumination. The insert is the cross-sectional scanning electron microscope image of PSCs; (b) EQE and integrated Jsc spectra of the corresponding optimal devices. (c) Normalized PCE of PSC (undoped CuSCN and doped CuSCN) as a function of aging time at 25 °C in N2 atmosphere. (d) Statistics of PCE for CuSCN and F4TCNQ-doped CuSCN devices (25 cells are included).
Table 1.

Summary of photovoltaic parameters of n–i–p planar PSCs based on doped and undoped CuSCN HTMs.

.

The long-term stability of these PSCs can also be evaluated by the JV characteristics monitoring of the devices stored in glove box (RH < 10%) at room temperature. As shown in Fig. 3(c), after 200 days, 88% of the initial PCE remains for the doped CuSCN, while only 76% remains for the undoped CuSCN. And we also conducted experiments about the steady-state PCEs for the devices with and without doping, which are 16.5% and 13.2%, respectively (Fig. A6). The 3% F4TCNQ is an optimal doping ratio (Fig. A4 and Table A2 in Appendix A), the statistics performances for 25 devices with 3% F4TCNQ doped or undoped devices are presented in Fig. 3(d) and Fig. A5. The averaged PCE for undoped devices is 13.1%, about 8.5% lower than that of the devices with F4TCNQ-doping (14.5%). From these results, the most important improvement is that Voc increases from 1.06 V to 1.10 V (Fig. A7). In addition, after using F4TCNQ-doped CuSCN as HTMs, the series resistance (Rs) of device decreases, and the shunt resistance (Rsh) increases (Table 1). The increase of Voc and FF means that the recombination at the CuSCN/perovskite interface can be reduced by doped CuSCN. The current–voltage curves measured under both reverse and forward scan directions were shown in Fig. A6, the parameters of the devices as shown in Table 1. And we also calculated the JV hysteresis factor, which defined by (PCEReverse-PCEForward)/PCEReverse,[3638] the hysteresis factor of the champion PSCs based on doped and undoped CuSCN are 0.030 and 0.036, respectively. Both devices with negligible hysteresis (Fig. A7 and Table 1).

Fig. A4. Forward scan JV curves of the optimal CsFAMA PSCs with F4TCNQ-doped CuSCN HTM under 1-sun illumination.
Fig. A5. Statistics of device parameters (Jsc, Voc, and FF) for PSCs based on CuSCN or F4TCNQ-doped CuSCN HTM (25 cells are included).
Fig. A6. Steady photocurrent output at fixed bias of Vmpp and steady PCE of the corresponding optimal devices.
Fig. A7. Reverse and forward JV curves of the optimal CsFAMA PSCs using (a) CuSCN and (b) F4TCNQ-doped CuSCN as HTMs under 1-sun illumination.
Table A2.

The device performance with changed F4TCNQ concentration.

.

As depicted in Fig. 4(a), compared with the reference device, the dark current density of the device with doped CuSCN is about one order of magnitude lower. This shows that using doped CuSCN as HTM can effectively inhibit the leakage current density of perovskite. Therefore, photo-generated carriers are dissociated effectively in the photovoltaic process. In addition, according to Eq. (3), Voc can be increased by reducing the leakage current density[39]

where n is the ideal factor, R and F are the ideal gas and Faraday constants, respectively, T is the temperature, Jsc is the short circuit current, and J0 is the dark current. It can be seen that the Voc increases with the decrease of the leakage current, thus devices with the doped hole transport material can obtain higher Jsc and Voc values. Actually, the atomic force microscopy (AFM) images indicate that the root-mean-square roughness (RMS) is reduced from 39 nm to 23 nm after doping (see Figs. 4(b)4(e)), which means that the smoother and denser hole transport layer may prevent direct contact between perovskite and gold electrode. All of these will reduce carrier recombination loss, and then increases the Voc of the device.

Fig. 4. (a) Dark JV curves of PSCs with CuSCN or F4TCNQ-doped CuSCN; SEM [(b) and (c)] and AFM [(d) and (e)] topographic images of CuSCN and F4TCNQ-doped CuSCN on the CsFAMA perovskite films. The scale bar is 1 μm for each SEM image and the scan scale is 10 μm × 10 μm for AFM images. (f) Current–voltage curves of CuSCN and F4TCNQ-doped CuSCN thin films.

The vertical conductivity of the undoped and doped film are measured by JV test system. Figure 4(f) shows that the conductivity of the doped CuSCN film is 2.51 × 10−4 S/cm, which is almost 2 times higher than the undoped CuSCN film (1.48 × 10−4 S/cm). The improvement of the conductivity of the HTM film will lead to the reduction of the Rs, thus improving the charge collection efficiency meanwhile improving the fill factor.[40]

As shown in Fig. 5(a), the p-type CuSCN HTM shows excellent transparency throughout the entire UV–vis–NIR range. In addition, the F4TCNQ-doped CuSCN demonstrates almost equivalent transmittance compared with undoped sample. This high transparency ensures that the perovskite absorption layer will have enough light under illumination from HTL side, and thus potentially leads to higher photocurrent in n–i–p ST-PSCs. Accordingly, the wide bandgap CuSCN (Eg = 3.67 eV) may be used as an excellent transparent HTM, applied on the bifacial ST-PSCs. Figure 5(b) shows the structure of a bifacial semitransparent n–i–p planar PSCs: ITO-glass/SnO2/perovskite/CuSCN/MoOx (9 nm)/ITO(100 nm)/Au. The structure can harvest sun-light from both sides.[41] To investigate the effect of F4TCNQ doping on charge collection in ST-PSCs, the JV curves of ST-PSCs with illumination from CuSCN and ITO/glass sides are shown in Figs. 5(c) and 5(d), and Table 2 also lists the corresponding detailed parameters. The parameters of bifacial semitransparent device with doped CuSCN indicate a Voc of 1.06 V, a JSC of 20.4 mA/cm2, a fill factor of 68% and a PCE of 14.8% by illuminating from ITO side, and a Voc of 1.02 V, a JSC of 19.2 mA/cm2, an FF of 64% and a PCE of 12.5% by illuminating from doped CuSCN side. The corresponding EQE of PSCs is shown in Figs. 5(e) and 5(f). Compared with the undoped CuSCN device, the semi-transparent device with doped CuSCN exhibits higher PCE and external quantum efficiency values in the range from 300 nm to 800 nm.

Fig. 5. (a) Transmittance spectra of a bare ITO/glass substrate, ITO-glass/undoped CuSCN substrate, and ITO-glass/doped CuSCN substrate, respectively; (b) ST-PSCs with two illumination directions. (c)–(f) JV measurements and EQE spectra of the ST-PSCs with illumination incident from ITO/glass and transparent electrode side, respectively. The integrated Jsc for both sides are given in the legend.
Table 2.

The JV parameters for ST-PSCs with light incident from ITO/glass and transparent electrode side, respectively.

.
4. Conclusion

In summary, a molecular doping strategy is used to tailor the electronic properties of inorganic CuSCN hole transporting layer. The doped CuSCN has a superior conductivity and well-matched energy level with perovskite, which is beneficial for hole transferring from perovskite to doped CuSCN and reducing the recombination at perovskite/CuSCN interface. As a result, the champion opaque cells with doped CuSCN exhibits PCE of 16.6%, Jsc of 21.5 mA/cm2, Voc of 1.10 V, FF of 70.5%. When employing as HTM in bifacial semitransparent n–i–p planar PSCs, the maximum efficiency of 14.8% and 12.5% are achieved by illuminating from the front side and rear side, respectively. This work can be used for reference to improve the efficiency of ST-PSCs based on CuSCN and will promote the development of BIPV.

Reference
[1] Cannavale A Hörantner M Eperon G E Snaith H J Fiorito F Ayr U Martellotta F 2017 Appl. Energy 194 94
[2] Kim B J Kim D H Lee Y Y Shin H W Han G S Hong J S Mahmood K Ahn T K Joo Y C Hong K S 2015 Energy Environ. Sci. 8 916
[3] Singh P Gupta S J J 2019 Int. J. Sci. Tech. Advancements 5 33
[4] Köhnen E Jošt M Morales-Vilches A B Tockhorn P Al-Ashouri A Macco B Kegelmann L Korte L Rech B Schlatmann R 2019 Sustainable Energy Fuels 3 1995
[5] Hanmandlu C Chen C Y Boopathi K M Lin H W Lai C S Chu C W 2017 ACS. Appl. Mater. Interfaces 9 32635
[6] https://www.nrel.gov/
[7] Wang P Y Li R J Chen B B Hou F H Zhang J Zhao Y Zhang X D 2020 Adv. Mater. 32 1905766
[8] Wang P Y Jiang Q Zhao Y Chen Y Chu Z Zhang X W Zhou Y Q You J B 2018 Sci. Bull. 63 726
[9] Li R J Wang P Y Chen B B Cui X H Ding Y Li Y L Zhang D K Zhao Y Zhang X D 2019 ACS Energy Lett. 5 79
[10] Shi B Duan L Zhao Y Luo J Zhang X 2019 Adv. Mater. 32 1806474
[11] Jeon N J Lee H G Kim Y C Seo J Noh J H Lee J Seok S 2014 J. Am. Chem. Soc. 136 7837
[12] Luo D Yang W Wang Z Sadhanala A Hu Q Su R Shivanna R Trindade G F Watts J F Xu Z 2018 Science 360 1442
[13] Kung P K Li M H Lin P Y Chiang Y H Chan C R Guo T F Chen P 2018 Adv. Mater. Interfaces 5 1800882
[14] Pattanasattayavong P Yaacobi-Gross N Zhao K Ndjawa G O N Li J Yan F O’Regan B C Amassian A Anthopoulos T D 2013 Adv. Mater. 25 1504
[15] Arora N Dar M I Hinderhofer A Pellet N Schreiber F Zakeeruddin S M Grätzel M 2017 Science 358 768
[16] Jung J W Chueh C C Jen A K Y 2015 Adv. Energy Mater. 5 1500486
[17] Fan L Li Y Yao X Ding Y Zhao S Shi B Wei C Zhang D Li B Wang G 2018 ACS Appl. Energy Mater. 1 1575
[18] Wang H Dewi H A Koh T M Bruno A Mhaisalkar S G Mathews N 2020 ACS Appl. Mater. Interfaces 12 484
[19] Jin I S Lee J H Noh Y W Park S H Jung J W 2019 Inorg. Chem. Front. 6 2158
[20] Wang S Huang Z Wang X F Li Y M Gunther M Valenzuela S Parikh P Cabreros A Xiong W Meng Y S 2018 J. Am. Chem. Soc. 140 16720
[21] Pellaroque A Noel N K Habisreutinger S N Zhang Y D Barlow S Marder S Snaith H J 2017 ACS Energy Lett. 2 2044
[22] Wang Q Bi C Huang J 2015 Nano Energy 15 275
[23] Chen W Wu Y H Fan J Djurisic A B Liu F Z Tam H W Ng A Surya C Chan W K Wang D He Z B 2018 Adv. Energy Mater. 8 1703519
[24] Li M Wang Z K Yang Y G Hu Y Feng S L Wang J M Gao X Y Liao L S 2016 Adv. Energy Mater. 6 1601156
[25] Zhang Y Elawad M Yu Z Jiang X Lai J Sun L 2016 RSC Adv. 6 108888
[26] Senthilkumar N Park S Kang H S Park D W Choe Y 2011 J. Ind. Eng. Chem. 17 799
[27] Zhu L Kim E G Yi Y Bredas J L 2011 Chem. Mater. 23 5149
[28] Su P Y Huang L B Liu J M Chen Y F Xiao L M Kuang D B Mayor M Su C Y 2017 J. Mater. Chem. 5 1913
[29] Wu W Q Wang Q Fang Y Shao Y Tang S Deng Y Lu H Liu Y Li T Yang Z 2018 Nat. Commun. 9 1625
[30] Chen W Zhou Y Wang L Wu Y Tu B Yu B Liu F Tam H W Wang G Djurišić A B 2018 Adv. Mater. 30 1800515
[31] Gelmetti I Montcada N F Pérez-Rodríguez A Barrena E Ocal C García-Benito I Molina-Ontoria A Martín N Vidal-Ferran A Palomares E 2019 Energy Environ. Sci. 12 1309
[32] Fan L Ding Y Luo J Shi B Yao X Wei C Zhang D Wang G Sheng Y Chen Y 2017 J. Mater. Chem. 5 7423
[33] Chen P Bai Y Wang S Lyu M Yun J H Wang L 2018 Adv. Funct. Mater. 28 1706923
[34] Muthu C Agarwal S Vijayan A Hazra P Jinesh K B Nair V C 2016 Adv. Mater. Interfaces 3 1600092
[35] Du Y Xin C Huang W Shi B Ding Y Wei C Zhao Y Li Y Zhang X 2018 ACS Sustainable Chem. Eng. 6 16806
[36] Tang Z G Bessho T Awai F Kinoshita T Maitani M M Jono R Murakami T N Wang H Kubo T Uchida Satoshi Segawa H 2017 Sci. Rep. 7 12183
[37] Li Z Tinkham J Schulz P Yang M Kim D H Berry J Sellinger A Zhu K 2017 Adv. Energy Mater. 7 1601451
[38] Liu P Wang W Liu S M Yang H G Shao Z P 2019 Adv. Energy Mater. 9 1803017
[39] Li W Dong H Guo X Li N Li J Niu G Wang L 2014 J. Mater. Chem. 2 20105
[40] Chen W Liu F Z Feng X Y Djurišić A B Chan W K He Z B 2017 Adv. Energy Mater. 7 1700722
[41] Zhu S Yao X Ren Q Zheng C Li S Tong Y Shi B Guo S Fan L Ren H 2018 Nano Energy 45 280