Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Air-stable perovskite photovoltaic cells with low temperature deposited NiOx as an efficient hole-transporting material

Open Access Open Access

Abstract

The electron-beam physical vapor deposition (EBPVD) technique was selected for nickel oxide (NiOx) film deposition at room temperatures. NiOx film (18 nm thick) was deposited as a hole transporting material (HTM) for inverted perovskite solar cells (PSCs) onto a fluorine-doped tin oxide (FTO)-coated glass substrate at a chamber vacuum pressure of 4.6×104 Pa. PSCs were fabricated as a glass/FTO/NiOx(HTM)/CH3NH3PbI3/PC61BM/BCP/Ag structure with as-deposited and annealed (500 °C for 30 min) NiOx films. Under 100 mW cm-2 illumination, as-deposited and annealed NiOx as HTM in PSCs (0.16 cm2) showed a high-power conversion efficiency (PCE) of 13.20% and 13.24%, respectively. The as-deposited and annealed PSCs retained 72.2% and 76.96% of their initial efficiency in ambient conditions, correspondingly. This study highlights the possibility of achieving highly crystalline and finely disseminated NiOx films by EBPVD for fabricating efficient inverted PSCs.

© 2020 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

A hybrid of organic and inorganic halide perovskite has become the center of solar power-related research due to extreme efforts involved in preparing high-quality perovskite thin films [1,2], device structures [3,4], elemental optimization [5,6], and a basic understanding of charge transportation dynamics [7,8]. This hybrid was identified as a promising photovoltaic material in 2009 by Miyasaka et al. [9]. A testified PCE of ∼3.9% was achieved by introducing and integrating perovskite materials as a semiconductor into photovoltaic devices. Kim et al. [10] made substantial development in third-generation solar cells by introducing full solid-state perovskite solar cells (PSCs) in 2012. In the last decade, the PCE of PSCs has grown consistently, and it is now beyond the recently certified record value of 25.2% [11]. In wide-range applications, PSCs with different structures (e.g., mesoporous, planar, and inverted planar) have been fabricated and studied. Planar architecture receives much attention because of its easy fabrication, low processing costs and temperature, and comparatively low hysteresis over mesostructured devices [1214]. Planar inverted structures allow searches for favorable inorganic charge-transporting materials [15] that can provide high PCE and good stability in ambient conditions [16] for robust construction. These inorganic materials cost lower than their counterpart organic materials. Organic lead iodide perovskite is prone to degradation in the presence of humidity and air due to its low formation energy [17]. Nevertheless, all structures can perform competently as long as the charge (electron and hole)-transporting layers of the device successfully separate and collect the charge carrier [1820] from the perovskite absorber layer. Therefore, the charge-transporting layers play a crucial part in the operational principles of PSCs. They influence the device’s performance and provide a shield to the absorber layer to protect it from being exposed to the corresponding unfavorable atmosphere, and, thus, the layers contribute to a remarkably stable inverted PSC [15]. Currently, most high-efficiency PSCs are made from organic HTMs (e.g., poly(3,4-ethylenedioxythiophene) polystyrene sulfonate (PEDOT:PSS) [21], 2,2′,7,7′-Tetrakis[N,N-di(4-methoxyphenyl)amino]-9,9′-spirobifluorene (Spiro-OMeTAD) [22], poly(triaryl amine) (PTAA) [23], and poly(3-hexylthiophene) (P3HT) [24], and these PSCs require additional doping and adhesive materials to increase the conductivity of organic HTMs [22,25]. These p-type HTMs are sensitive to environmental conditions due to their acidic and hygroscopic characteristics, especially when stability becomes a concern. The organic materials previously listed have clear advantages on device fabrication in terms of low processing temperature and easy process route (solution-based) and arbitrary tunable bandgap option. However, their high synthesis and purification costs and degradation must be taken into account for high-volume commercial production [15]. Deploying the perovskite precursor on organic polymers properly is also difficult due to the hydrophobicity of organic polymers. Moreover, designing, synthesizing, and aligning the bandgap of organic HTLs are tedious. These processes result in high production costs for large-scale manufacturing of PSCs [26]. Therefore, inorganic HTMs have become the center of attention for inverted PSCs, especially metal oxides such as NiOx [27], CuI [28], CuSCN [29], CuxO [30], and MO3 [31]. Using metal oxides, including MoO3, WO3, NiOx, and Cr2O3 as HTM has promising outcomes such as improving the carrier transporting properties within solar cell devices. Metal oxides can also serve protection to the photo absorber layers against unfavorable environmental conditions [32,33] though these inorganic oxides still need to go further to catch up organic materials in terms of PCE. Among these metal oxides, NiOx is an ideal choice for holding an adequately large band gap (>3.7 eV) [34], allowing high transparency and suitable and favorable energy levels for effective hole transport. In conjunction with upper valence band maximum (VBM), this HTM can facilitate hole movement to the photoanode, unlike perovskite absorber layers. NiOx also has a higher conduction band minimum (CBM) compared with perovskite absorber layers, which restrict electron movement to the photoanode of perovskite, thus providing good chemical stability [35]. Different approaches have been considered for the successful preparation of NiOx as an HTM in PSC. Seo et al. [36] deposited NiOx as HTM by atomic layer deposition (ALD), which achieved 16.6% efficiency. Park et al. [37] used pulsed laser deposition (PLD), which produced 17.3% device efficiency. Both of these costly processes are known for small-scale deployment, and they may be unsuitable for complete device fabrication in commercial production. Deposition of Li-doped NiOx (LiNiOx) through hot-casting method was introduced by Nie et al. and they achieved a highly efficient, hysteresis-free, and stable PSC with a PCE of approximately 18% [38]. Guo et al. [39] utilized thermolysis, which requires high temperature for nickel hydroxide to form NiOx HTM layers for inverted PSCs, and this method achieved a PCE of 14.55%. Solution-processed NiOx thin films were prepared by spin coating nickel (II) acetylacetonate in ethanol and HCl. The PSC obtained a PCE of 18.8%, retaining 16.92% of the initial PCE upon exposure to environmental conditions (>70% relative humidity) for 720 h [40]. Sol-gel processed NiOx nanocrystals showed a PCE of 9.11% [41]. All these studies indicated the enormous potential of NiOx as a successful HTM for PSC. The characteristic properties of NiOx are primarily influenced by NiOx preparation methods, which allow for versatile fabrication of full device structure. Although the solution process generated expected results, the conversion of crystallize source to polycrystalline NiOx demanded high processing temperatures, resulting in difficult production [16]. By contrast, Jung et al. [42] used a combustion method, a low temperature one, to achieve 17.74% efficiency, but it still required additional Cu-doping. Pre-synthesizing NiOx NPs as HTMs is also complex and tedious, demanding a moderate amount of temperature [43] during synthesis. The material utilization efficiency of the solution process is lower with higher wastage compared with that of other methods [44]. Two important photovoltaic parameters (e.g., Jsc and FF) are affected substantially by using liquid processed NiOx films as HTM, reducing the overall performance of the PSC device [45]. Moreover, any residual solvents such as ethanol or methanol while preparing the precursor solution of HTM can decompose the perovskite layer [15].

A facile deposition method for NiOx that offers low cost and high yield with commercial prospect for large-scale production is currently in demand. In the search of substitutional deposition method, physical vapor deposition techniques are also popular where sputtering has been mostly investigated [46] facilitating additional doping option. Magnetron sputtering has been used in PSC to prepare NiOx as HTM by Aydin et al. [47] which exhibited 18.49% efficiency. Electron beam physical vapor deposition (EBPVD) technique is known alongside sputtering to prepare film of metal oxides in optical semiconductor film industries. Due to possessing a few numbers of deposition parameters and simple process, EBPVD is suitable alternative instated of other physical vapor techniques (Fig. 9, Supplementary Information, SI) such as sputtering or thermal evaporator for NiOx deposition. Structural and morphological properties of films deposited by EBPVD can easily be manipulated. Besides, the required raw material volume in EBPVD is lower than that in solution process technique, thereby lowering material utilization. The slow as well as fast, both types of deposition rate of EBPVD allows it to have potential industrial applications, such as in semiconducting optical film and solar cell applications. The mean free path and source-to-substrate distance allow control of films’ properties. A very limited number of investigations has been performed using EBPVD to prepare NiOx film as HTM for perovskite solar cell. Recently T. Abzieher et al. [48] reported competitive efficiency with electron-beam deposited NiOx as HTM with complex inkjet-printed perovskite absorbers layer. Though additional oxygen gas was introduced to enhance NiOx transmission profile which again push this simple system towards some difficulty in controlling oxygen amount. To overcome these issues, this study investigates a simple EBPVD technique to prepare NiOx films as HTM and effective electron blocking layer in PSC without insertion of any heteroatomic dopant given that physical vapor deposition (PVD) techniques are widely used in semiconductor film preparation for efficient and stable perovskite solar cell. Here, we deposited NiOx as HTM for PSCs by EBPDV without providing any intentional heat treatment and achieved very close Voc and FF values compared with the aforementioned solution techniques. This study focused on the simplest perovskite composition and one step perovskite deposition technique in fabrication of PSC devices with glass/FTO/NiOx/CH3NH3PbI3/PCBM/BCP/Ag structure (inverted assembled). The devices with as-deposited and annealed NiOx film exhibited the highest PCE of 13.20% and 13.24%, respectively. This outcome is because of the successful matching to the effective Fermi level (Ef) of NiOx with corresponding valance band of FTO and CH3NH3PbI3. Moreover, NiOx provided such an optical transparency, which produced high Jsc in accordance with light absorption via the active layer of CH3NH3PbI3. The stability of the fabricated devices using EBPVD NiOx film as HTLs under ambient environment is investigated for 4 weeks’ time period. The devices extended their stability by 72.72% and 76.96% of the initial PCE for as-deposited and annealed NiOx HTM, respectively, at aforementioned time. The outcome of this study revealed that a competitive PCE with inorganic HTM alongside organic HTM with stable PSC in environmental conditions was achievable with lower temperature processed EBPVD NiOx films as HTM.

2. Experimental

2.1. Materials

All chemicals were purchased from commercial sources with high purity and used without further purification. Methyl ammonium iodide (CH3NH3I, >98%) was purchased from Tokyo Chemical Industry Co. (Japan), lead iodide (PbI2,99%) was purchased from Sigma–Aldrich, NiOx target (99.99%) was purchased from Kujondo Chemical Laboratories Co., LTD. (Japan), and PC61BM (phenyl-C61-butyric acid methyl ester) (99.5%) was purchased from Lumtec Co. (Taiwan). Other solvents such as gamma butyrolactone (GBL), dimethylsulfoxide (DMSO), chlorobenzene, hydochloric acid, acetone, bathocuproine (BCP) and zinc powder were purchased from Wako Pure Chemical Industries, Ltd. (Japan).

2.2 Deposition of NiOx film on the FTO substrate and fabrication of inverted PSCs

Fluorine doped in oxide (FTO, Pilkington, 14 Ω.cm) glasses (2.5 cm × 2.5 cm) were paternally etched by Zn powder and 2 M HCl solution, consecutively washed with detergent and deionized water, and dried with nitrogen blower. The dried substrates were then sonicated into acetone for 10 min by ultrasonic bath. The substrates were dried again with a nitrogen blower after sonication was completed. Before preparing inverted PSCs, UV/O3 treatment of the dried substrates in UV-O3 stripper was carried out at 115 °C for 15 min. On the patterned FTO glass substrate, NiOx as HTL films were deposited from a 99.99% pure NiOx pallet (1 cm2) by EBPVD technique. The deposition chamber pressure was approximately 4.6 × 104 Pa during deposition. The substrates were deprived from any purposeful heating (room temperature) during deposition, and the deposition rate was adjusted to 0.15 nm/s to get around 20 nm-thick film. The HTM-coated substrates were taken into the N2 glove box directly for further fabrication. Few of the HTM-coated substrates were then heated at 500 °C for 30 min in a muffle furnace (vacuum) for comparative study. The preparation of perovskite precursor solution has been described elsewhere [49]. In brief, 461 mg of PbI2 was measured in a glass container, and 159 mg of CH3NH3I was measured in another different glass container. Subsequently, 78 mg of DMSO and 700 mg of GBL were used to make a clear solution of CH3NH3I. The CH3NH3I solution was added to the PbI2 container and stirred at 55 °C overnight. During device fabrication, 50 µL of CH3NH3PbI3 precursor solution was spread over NiOx substrates. The spin coater was set into two different rotations and time interval with a 2 s slope. The first spin coating step was set at 1000 rpm for 10 s. The second spin step was set at 5000 rpm for 30 s immediately after 2 s slope in between. At the end of 14 s of the second spinning step, 500 µL of anhydrous toluene was dripped onto rotating substrate and annealed. The deposited film’s annealing process was distributed into two parts. First, the films were thermally treated for 10 min at 50 °C onto one hot-plate and then moved to another hot-plate set for annealing for 10 min at 100 °C. Consequently, 50 µL of PC61BM solution (20 mg/mL in chlorobenzene) was spread over the CH3NH3PbI3-deposited substrates, spin-coated for 30 s at 1000 rpm, and treated for 10 min at 80 °C. The BCP solution (150 µL) was dropped on the substrate and rotated for 30 s at 6000 rpm. After spinning was completed, the substrates were annealed at 80 °C (10 min). A 100 nm Ag back contact electrode was deposited on top of the BCP layer by a resistive thermal evaporator under a vacuum chamber pressure of 4.65 × 10−4 Pa. A customized patterned mask was used to attain the expected device structure of final PSCs.

2.3 Characterization

The crystalline structure of NiOx and CH3NH3PbI3 layers were characterized by X-ray diffraction (XRD) using an x-ray structure analyzer RINT-TTR III/N (Rigaku) with Cu Kα (k=1.5418 Å) radiation to determine 2θ values in the range of 10°–60° and 30°–90° for NiOx and perovskite film, respectively. The surface roughness of the deposited NiOx films were determined by atomic force microscopy (AFM, Nanosurf easyscan 2 AFM) in tapping configuration to a scanning area of 3 µm×3 µm. Surface morphologies of films were analyzed by field emission scanning electron microscope (FESEM, Hitachi SU6600, Japan) in low vacuum condition. Optical transmittance (T%) of NiOx and CH3NH3PbI3 films was measured by an optical spectrophotometer in a UV-near IR (JASCO V-570, Japan) equipped with a double-beam system with single monochromator wavelength ranging from 190 nm to 2500 nm. The ionization potential was measured by AC-3 with 2.0 nW light intensity and 0.05 interval (RIKEN KEIKI, JAPAN). The resistivity and mobility of the deposited NiOx films were measured by HMS ECOPIA 3000. The magnetic field was 0.57 Tesla, and the probe current was set to 4 mA for all the samples. All these measurements were completed under ambient environmental conditions.

The photovoltaic parameters and performances of the fabricated PSCs were measured by a Keithley 2400 source meter with solar simulator (CEP-2000RP, Japan) under simulated AM 1.5 G and evaluated from the characteristic current density–voltage (J–V) arch. The computer-controlled solar illumination was set to 100 mW cm-2 in air. Both the forward scan and reverse scan were set to start from −0.2 V to 1.2 V and vice versa. The external quantum efficiency (EQE) spectra were measured (300–850 nm wavelength) using an EQE system (Bunkoukeiki Co., CEP2000RR, Japan). The photo-illuminated active area of PSCs was 0.16 cm2 (0.4 cm × 0.4 cm) without further protective measures, such as encapsulation. The actual thicknesses of the deposited films were determined by Dektak Veeco Surface profiler.

3. Results and discussion

The surface morphological and crystalline properties of NiOx films were characterized by FESEM and XRD, respectively. The NiOx films were deliberately deposited onto Pilkington glass for XRD intensity. Figure 1(a) shows that the as-deposited NiOx films presented clear and distinct 2θ XRD peaks at 37.36°, 43.53°, 63.20°, 75.66°, and 79.67°, which were indexed to the (1 1 1), (2 0 0), (2 2 0), (1 1 3), and (2 2 2) planes of NiOx [50], respectively agreeing with JCPDS cards no. 731523. The high-temperature annealed NiOx films’ corresponding peak intensity was stronger than that of the as-deposited films, indicating the high crystallinity of the NiOx films once annealed. Moreover, Fig. 1(a) reveals that peaks of the 1 1 1 and 2 0 0 planes of the as-deposited films were almost in the same intensity, but the former dominated while annealed at 500 °C. The average crystallite size (Dp) of NiOx films was measured with the Scherrer formula [51] expressed as follows.

$$\textrm{Dp} = \textrm{K}\mathrm{\lambda} /\mathrm{\beta}\,\textrm{cos}\mathrm{\theta} $$
where λ represents the wavelength of X-ray (1.54 Å), β is the full width at half maximum of diffraction peaks, θ is Bragg’s angle of XRD peak, and K is a constant value of 0.94, namely, “shape factor.” The Dp of the [1 1 1] and [2 0 0] orientations were 14.97 and 13.97 nm for the as-deposited films, respectively. On the other hand, Dp of the 1 1 1 and 2 0 0 orientations were 16.76 and 14.26 nm for the annealed films, respectively. The stoichiometry of NiOx thin films was represented by both [1 1 1] and [2 0 0] XRD peaks [52], as shown in Fig. 1(a). These [1 1 1] and [2 0 0] reflections corresponded to the formation of non-stoichiometric and stoichiometric NiOx, respectively. Defects related to Ni2+ or oxygen vacancy at high annealing temperature might be responsible for the non-stoichiometric formation, enhancing NiOx films’ p-type charge carrier properties. This result was attributed to the manipulation of carrier transportation of films. By contrast, as-deposited NiOx films had good crystallinity with preferred orientation even before heat treatment of these films deposited by EBPVD. The typical energy-dispersive X-ray (EDX) images (Fig. 10, SI) suggest elemental mapping of oxygen, nickel, tin and fluorine with weight% and atomic%).

 figure: Fig. 1.

Fig. 1. XRD data for as-deposited and annealed NiOx film on glass substrate.

Download Full Size | PDF

Therefore, EBPVD technique could facilitate the deposition of highly crystalline NiOx films. Figure 2(a) illustrates the low-magnification SEM image of as-deposited NiOx film on glass substrate. The NiOx film was well distributed, and no clustering was observed. Moreover, to investigate the distribution of NiOx, the FESEM images of bare glass, FTO and NiOx film-coated FTO were obtained and shown in Fig. 2(b-c) and (d). The NiOx was evenly dispersed as a film on FTO and perfectly replicated the morphology of FTO crystals as shown in the Fig. 2(d) inset. The grain boundaries were responsible for the surface roughness of FTO by nature.

 figure: Fig. 2.

Fig. 2. SEM images of (a) as-deposited; FESEM images of (b) as-deposited, (c) annealed NiOx film on bare glass and (d) bare FTO; inset NiOx-coated FTO glass.

Download Full Size | PDF

The resistivity (ρ) and hole concentration (Nb) of the as-deposited and annealed NiOx films were measured by the Hall effect and showed in Table 1.

Tables Icon

Table 1. Summery of electrical properties of NiOx films

The as-deposited films showed ρ and Nb of 4.81×103 Ω.cm and 2.01 × 1015/cm3, respectively. Improvements upon annealing of NiOx films were observed. The resistivity (3.63) Ω.cm decreased by one orders of magnitude, and the carrier concentration (1.16 × 1516/ cm3) increased by one order of magnitude in high-temperature annealed NiOx compared with as-deposited films. But the carrier mobility of annealed NiOx film was increased in a quiet number of folds comparing to as-deposited film. As shown in Fig. 3, the EBPVD processed NiOx considerably leveled the FTO surface, which was demonstrated by reducing the root-mean-square roughness (Rq) of FTO. The Rq value of bare FTO was ∼25 nm, which reduced to ∼18 and ∼17 nm for the as-deposited and annealed FTO/NiOx samples with the EBPVD method as expected, respectively. The uneven surface of HTM was expected to an extent because excess HTM layer smoothness might disturb Perovskite film adhesion on the HTM layer. The actual thicknesses of the deposited films were confirmed by surface profiler measurements, which revealed an average thickness of 18 nm.

 figure: Fig. 3.

Fig. 3. Topographic images by AFM of (a) bare FTO, (b) as-deposited FTO/NiOx film, and (c) annealed FTO/NiOx film; inset shows the 3D images.

Download Full Size | PDF

In an inverted structure of a PSC device, the TCO (FTO) layer is covered by the HTM initially while fabricating the device. Illuminated light first crosses the HTM layer before reaching the perovskite absorber layer. As a result, transparency is a major concern while choosing an HTM for PSC. Figure 4 presents the transmittance spectra of bare FTO, as-deposited NiOx/FTO and annealed NiOx/FTO in the region of 300–900 nm of solar magnetic radiation. A slight difference in bandgap between as-deposited and annealed NiOx film might be recognised by the alterations of Dp, microstrain (ε) and dislocation densities (δ), resulting in lattice inconsistency and defects [53] of films. The absorption coefficient of the films increased at high annealing temperature. This performance could have possibly produced the increased density of states of holes at higher temperature annealing [54]. The results demonstrated that NiOx-coated FTOs showed a negligible difference in light transmission compared with bare FTO. The excellent transparency properties of the as-deposited and annealed NiOx HTLs are advantageous for fabricating inverted PSC because the absorber layer has maximum light for absorption.

 figure: Fig. 4.

Fig. 4. Transmittance of bare FTO and annealed FTO/NiOx

Download Full Size | PDF

Since the device was fabricated in a p-i-n structure so maximum transparency was expected. A careful characterization suggested that upon the incremental thickness of NiOx, the films’ bandgap reduced resulting lower transmission profile (Fig. 11, SI). It has been reported that 4% increment in films absorbance exhibits 1.0 mA cm−2 loss of Jsc [48]. Meanwhile, for the comparison, 500 °C was proposed since at this temperature “interference oscillation” takes place inside the substrate which is the outcome of multiple reflection at NiOx/substrate film’s boundary resulting smooth film formation [55]. High work function (φ), poor crystallinity, and poor dispersion were regarded as substantial drawbacks of low-temperature NiOx films in producing improved PCEs [41]. We measured the ionization potentials (IPs) of the as-deposited (5.34 eV) and annealed (5.29 eV) NiOx films by photoelectron yield spectroscopy (PYS) for further validation as shown in the Fig. 5(a) and Fig. (b) for as-deposited and annealed NiOx films. The estimated Fermi level (Ef) should be 0.2 eV above the VMB of NiOx [15], so the achieved work functions were 5.14 and 5.09 eV for as-deposited and annealed NiOx, respectively.

 figure: Fig. 5.

Fig. 5. Workfunction determination of (a) as-deposited and (b) annealed NiOx film by PYS.

Download Full Size | PDF

As a result, the VBM (−5.4 eV) of CH3NH3PbI3 demonstrated optimum band alignment configuration corresponding to the VBM of NiOx (−5.29 eV for as-deposited or −5.34 eV for annealed) films, thus facilitating efficient charge carrier (hole) transfer from VB of the CH3NH3PbI3 (−5.4 eV) layer to HTM NiOx, as shown in the Fig. 6(b). It is not energetically possible to inject an electron from CB of CH3NH3PbI3 into NiOx because NiOx is a high band-gap semiconductor with CBM ∼−1.8 eV compared with CBM of CH3NH3PbI3 (−3.9 eV). Figure 6(a) shows the full device geometry of fabricated PSC. By contrast, Fig. 7 shows the characteristic JV of the typical devices with as-deposited and annealed NiOx HTLs.

 figure: Fig. 6.

Fig. 6. (a) Fabricated p-i-n PSC and (b) corresponding band diagram.

Download Full Size | PDF

 figure: Fig. 7.

Fig. 7. (a) J-V curves measured at solar simulator under simulated AM 1.5 G in this experiment at day 1 and 28th of fabrication; (b) EQE of the inverted planar PSCs fabricated with NiOx HTLs for both as-deposited and annealed NiOx film.

Download Full Size | PDF

The deposited NiOx HTL films were annealed at 500 °C for 30 min, and PSC devices were fabricated following the same procedure for as-deposited NiOx HTLs to evaluate the effect of high-temperature annealing. The corresponding photovoltaic parameters including Voc, Jsc, FF, and PCE are presented in Table 2. Two other important parameters, namely, series resistance (Rs) and shunt resistance (Rsh), are recorded in Table 2, forming the J-V arcs. However, the thickness tuning of NiOx films may differ in producing PSC devices, showing high performance depending on the deposition technique used. Therefore, the deposition technique of NiOx films while fabricating PSC devices may alter characteristic properties such as crystallite size, plane orientation, band gap, and resistivity [5658].

Tables Icon

Table 2. Summary of fabricated PSC devices’ performance with NiOx HTM at the 1st and 28th days of fabrication

Under 100 mW/cm2 illumination, the devices with as-deposited NiOx HTM films displayed a PCE of 13.20%, with Jsc of 17.90 mA/cm2, FF of 77%, and Voc of 0.95 V. The devices with annealed NiOx HTM films reached a PCE of 13.24% with Jsc of 17.16 mA/cm2, Voc of 0.98 V, and FF of 0.78. The PCEs of the fabricated PSC devices were quite encouraging. The successful integration of the EBPVD NiOx film as HTM were clearly pronounced by the PSC devices’ performance. Moreover, the charge carriers generated in the CH3NH3PbI3 absorber layer were effectively transported to corresponding electrodes by selective contact to show those efficiencies. The data likewise disclose that the main photovoltaic parameters of the device based on as-deposited and annealed NiOx HTL are almost same. Even their EQE (Fig. 7(b), both reaching 70% at 400-800 nm, also revels the identical outcomes which indicate EB deposition at low temperature has successfully managed to form quality NiOx HTL film on top of FTO for better charge carrier transportation. However, Voc of less than 1.0 V was observed, restricting FF and device performance. Rs refers to loss of internal carrier mobility [59], and Rsh refers to loss of photo-induced current due charge carrier recombination at the junction layer formed between layers of the device [60]. Both as-deposited and annealed devices showed a low Rs and high Rsh at the 1st day of fabrication, which were expected for such devices. On the contrary, Rs increased and Rsh decreased, affecting Voc and FF and corresponding PCEs of inverted PSCs but still holding fair photovoltaic values after four weeks of deposition.

Figure 8(a) refers to the perfect XRD peaks with proper crystalline orientation of CH3NH3PbI3 formation, but unchanged PbI2 remained present. Figure 8(b) illustrates the light absorption spectra of CH3NH3PbI3 film on as-deposited and annealed NiOx films, indicating different absorption patterns in strong energy regions. The CH3NH3PbI3 film on high temperature annealed NiOx film showed better absorption in the solar spectrum range of 400–500nm compared with CH3NH3PbI3 film on top of the as-deposited NiOx film. The SEM image of the CH3NH3PbI3 layer on NiOx showed pinhole-free film formation. However, the 500–700 nm grain size of the CH3NH3PbI3 layer possessed grain boundaries (Fig. 8(c)), which might cause electron-hole recombination at the junction interface. Meanwhile, it was expected that annealed NiOx will exhibit high performance than as-deposited films due to possessing better carrier concentration. But consecutive increment carrier mobility in annealed film might be responsible for a minor improvement in PCE between both devices. The highest and average value of PCE, Jsc, Voc and FF of as-deposited and annealed NiOx film as HTM in PSC devices have been shown in the Table 3. Though the overall efficiency of both types PCEs almost identical, but the PSC devices with annealed NiOx as HTM provided very consistent performance while productivity was considered.

 figure: Fig. 8.

Fig. 8. (a) X-ray diffraction peaks of CH3NH3PbI3; (b) UV-Vis spectra of FTO and PSC on as-deposited and annealed NiOx; (c) SEM topography of CH3NH3PbI3, and (d) cross-sectional FESEM image of the fabricated PSC.

Download Full Size | PDF

Tables Icon

Table 3. Summary of the highest and average values of photovoltaic performances of PSC devices.

To assess air (environmental) stability on NiOx HTM device, we also measured the photovoltaic performance of those devices at the 28th day of fabrication. We measured and compared the PCE and other solar cell parameters of the devices. The devices were kept in 60%-70% relative humidity at 13 °C-25 °C in ambient environment, and no capping as well no intentional illumination was introduced during cell preservation. The measured JV of the fabricated PSCs on the 1st day are shown in Fig. 6(a). Jsc decreased from 17.90 mA/cm2 on the 1st day to 16.39 mA/cm2 on the 28th day of fabrication for the as-deposited NiOx-based PSCs. Meanwhile, PSCs based on annealed-NiOx/CH3NH3PbI3/PCBM/BCP/Ag configuration showed reduced Jsc from 17.16 mA/cm2 to 14.93 mA/cm2 on the 1st and 28th days of measurement, respectively, followed by a decline in Voc and FF. The lower Rsh values of aged devices comparing with initial values of both fresh as-deposited and annealed NiOx PSC suggested the presence of considerable recombination at NiOx interface upsetting those Voc. The Rsh was might be reduced due to presence of pin-hole introducing parasitic loss [61]. However, metal electrode penetration could also be responsible irreversible degradation especially at elevated temperature during measurement [55] which may lead to poor device performance later. As shown in Table 1, the devices showed a great stability retaining 72.72% and 76.96% of the initial PCE for the as-deposited and annealed NiOx HTM, respectively, for the aforementioned time. Thus, EBPVD could successfully facilitate NiOx HTM film formation for efficient and air-stable/ environment-stable PSC.

4. Conclusion

We demonstrated that EBPVD could form NiOx HTL films of superior quality using high-quality NiOx pallet. The NiOx HTL showed decent transparency, allowing a maximum gain of photon flux. The prepared NiOx HTL exhibited rapid charge transfer to reduce carrier recombination and thus can be considered as an effective HTM. In addition, NiOx possessed strong chemical stability, and its robustness toward environmental conditions was encouraging. Consequently, the as-deposited and annealed NiOx HTL-based PSC devices showed PCE of 13.20% and 13.24%, retaining 72.2% and 76.96% of their primary PCE in ambient condition, respectively, after 672 hours of fabrication without any kind of intentional capping. It is clear evident that even without any post deposition heat treatment of electron-beam NiOx film can substantially provide a comparative device efficiency. Moreover, an 18 nm-thick NiOx was sufficient for improved photovoltaic performance to ensure high material utilization efficiency and low-cost production. This study utilized EBPVD as an attainable means to achieve high-quality NiOx inorganic HTM for efficient and stable PSCs. The EBPVD technique could produce well-dispersed, dense, high purity, and unwanted contamination-free NiOx film at low temperature with high material utilization. However, the presence of surface recombination led to poor Voc. This result suggested that the improvement of Voc thorough HTM should be taken into account to minimize current loss due to interface and bulk recombination.

Appendix

 figure: Fig. 9.

Fig. 9. Working principal of EBPVD Technique.

Download Full Size | PDF

 figure: Fig. 10.

Fig. 10. EDX spectrum of FTO/NiOx films. Inset shows the weight% and atomic% of elements of as-deposited NiOx film.

Download Full Size | PDF

 figure: Fig. 11.

Fig. 11. Transmittance of EBPVD NiOx films on glass substrate. A clear decrease of transmission is observed in the higher energy shorter-wavelength when NiOx film thickness is increased and continue to exhibit lower transmission in the visible region of spectrum as well.

Download Full Size | PDF

Funding

Universiti Kebangsaan Malaysia (LRGS/1/2019/UKM-UKM/6/1).

Acknowledgments

The authors would like to acknowledge the Long Term Research Grant Scheme (LRGS) (LRGS/1/2019/UKM-UKM/6/1) grant funded by the University Kebangsaan Malaysia.

Disclosures

“The authors declare no conflicts of interest.”

References

1. J. Burschka, N. Pellet, S.-J. Moon, R. Humphry-Baker, P. Gao, M. K. Nazeeruddin, and M. Grätzel, “Sequential deposition as a route to high-performance perovskite-sensitized solar cells,” Nature 499(7458), 316–319 (2013). [CrossRef]  .

2. N. J. Jeon, J. H. Noh, Y. C. Kim, W. S. Yang, S. Ryu, and S. I. Seok, “Solvent engineering for high-performance inorganic–organic hybrid perovskite solar cells,” Nat. Mater. 13(9), 897–903 (2014). [CrossRef]  .

3. K. Chen, Q. Hu, T. Liu, L. Zhao, D. Luo, J. Wu, Y. Zhang, W. Zhang, F. Liu, T. P. Russell, R. Zhu, and Q. Gong, “Charge-carrier balance for highly efficient inverted planar heterojunction perovskite solar cells,” Adv. Mater. 28(48), 10718–10724 (2016). [CrossRef]  .

4. Z. Yang, A. Rajagopal, C.-C. Chueh, S. B. Jo, B. Liu, T. Zhao, and A. K.-Y. Jen, “Stable Low-Bandgap Pb–Sn Binary Perovskites for Tandem Solar Cells,” Adv. Mater. 28(40), 8990–8997 (2016). [CrossRef]  .

5. N. J. Jeon, J. H. Noh, W. S. Yang, Y. C. Kim, S. Ryu, J. Seo, and S. I. Seok, “Compositional engineering of perovskite materials for high-performance solar cells,” Nature 517(7535), 476–480 (2015). [CrossRef]  .

6. M. Saliba, T. Matsui, K. Domanski, J.-Y. Seo, A. Ummadisingu, S. M. Zakeeruddin, J.-P. Correa-Baena, W. R. Tress, A. Abate, A. Hagfeldt, and M. Grätzel, “"Incorporation of rubidium cations into perovskite solar cells improves photovoltaic performance,” Science 354(6309), 206–209 (2016). [CrossRef]  .

7. S. D. Stranks, G. E. Eperon, G. Grancini, C. Menelaou, M. J. Alcocer, T. Leijtens, L. M. Herz, A. Petrozza, and H. J. Snaith, “Electron-hole diffusion lengths exceeding 1 micrometer in an organometal trihalide perovskite absorber,” Science 342(6156), 341–344 (2013). [CrossRef]  .

8. J. Xu, A. Buin, A. H. Ip, W. Li, O. Voznyy, R. Comin, M. Yuan, S. Jeon, Z. Ning, J. J. McDowell, P. Kanjanaboos, J.-P. Sun, X. Lan, L. N. Quan, D. H. Kim, I. G. Hill, P. Maksymovych, and E. H. Sargent, “Perovskite-Fullerene Hybrid Materials Eliminate Hysteresis in Planar Diodes,” Nat. Commun. 6(1), 7081 (2015). [CrossRef]  

9. A. Kojima, K. Teshima, Y. Shirai, and T. Miyasaka, “Organometal halide perovskites as visible-light sensitizers for photovoltaic cells,” J. Am. Chem. Soc. 131(17), 6050–6051 (2009). [CrossRef]  .

10. H.-S. Kim, C.-R. Lee, J.-H. Im, K.-B. Lee, T. Moehl, A. Marchioro, S.-J. Moon, R. Humphry-Baker, J.-H. Yum, J. E. Moser, M. Grätzel, and N.-G. Park, “Lead iodide perovskite sensitized all-solid-state submicron thin film mesoscopic solar cell with efficiency exceeding 9%,” Sci. Rep. 2(1), 591 (2012). [CrossRef]  .

11. NREL Photovoltaic Research, “Best Research-Cell Efficiency Chart” (available on 04/05/2020) https://www.nrel.gov/pv/assets/pdfs/best-research-cell-efficiencies.20200406.pdf

12. D. Bi, C. Yi, J. Luo, J.-D. Décoppet, F. Zhang, S. M. Zakeeruddin, X. Li, A. Hagfeldt, and M. Grätzel, “Polymer-templated nucleation and crystal growth of perovskite films for solar cells with efficiency greater than 21%,” Nat. Energy 1(10), 16142 (2016). [CrossRef]  

13. M. Saliba, T. Matsui, J.-Y. Seo, K. Domanski, J.-P. Correa-Baena, M. K. Nazeeruddin, S. M. Zakeeruddin, W. Tress, A. Abate, A. Hagfeldt, and M. Grätzel, “Cesium-containing triple cation perovskite solar cells: improved stability, reproducibility and high efficiency,” Energy Environ. Sci. 9(6), 1989–1997 (2016). [CrossRef]  .

14. X. Wen, Y. Feng, S. Huang, F. Huang, Y.-B. Cheng, M. Green, and A. Ho-Baillie, “Defect trapping states and charge carrier recombination in organic–inorganic halide perovskites,” J. Mater. Chem. C 4(4), 793–800 (2016). [CrossRef]  .

15. J. You, L. Meng, T.-B. Song, T.-F. Guo, M. Yang, W.-H. Chang, Z. Hong, H. Chen, H. Zhou, Q. Chen, Y. Liu, N. De Marco, and Y. Yang, “Improved air stability of perovskite solar cells via solution-processed metal oxide transport layers,” Nat. Nanotechnol. 11(1), 75–81 (2016). [CrossRef]  .

16. T. H. Chowdhury, M. Ryuji Kaneko, E. Kayesh, M. Akhtaruzzaman, K. B. Sopian, J.-J. Lee, and A. Islam, “Nanostructured NiOx as hole transport material for low temperature processed stable perovskite solar cells,” Mater. Lett. 223, 109–111 (2018). [CrossRef]  .

17. J. H. Noh, S. H. Im, J. H. Heo, T. N. Mandal, and S. I. Seok, “Chemical management for colorful, efficient, and stable inorganic–organic hybrid nanostructured solar cells,” Nano Lett. 13(4), 1764–1769 (2013). [CrossRef]  .

18. W. Kim, S. Kim, S. U. Chai, M. S. Jung, J. K. Nam, J.-H. Kim, and J. H. Park, “Thermodynamically self-organized hole transport layers for high-efficiency inverted-planar perovskite solar cells,” Nanoscale 9(34), 12677–12683 (2017). [CrossRef]  .

19. A. Krishna, D. Sabba, J. Yin, A. Bruno, P. P. Boix, G. Yang, H. A. Dewi, G. G. Gurzadyan, C. Soci, S. G. Mhaisalkar, and A. C. Grimsdale, “Facile Synthesis of a Furan–Arylamine Hole-Transporting Material for High-Efficiency, Mesoscopic Perovskite Solar Cells,” Chem. - Eur. J. 21(43), 15113–15117 (2015). [CrossRef]  

20. J. Tang, D. Jiao, L. Zhang, X. Zhang, X. Xu, C. Yao, J. Wu, and Z. Lan, “High-performance inverted planar perovskite solar cells based on efficient hole-transporting layers from well-crystalline NiO nanocrystals,” Sol. Energy 161, 100–108 (2018). [CrossRef]  .

21. J. You, Y. Yang, Z. Hong, T.-B. Song, L. Meng, Y. Liu, C. Jiang, H. Zhou, W.-H. Chang, G. Li, and Y. Yang, “Moisture assisted perovskite film growth for high performance solar cells,” Appl. Phys. Lett. 105(18), 183902 (2014). [CrossRef]  .

22. J. H. Noh, N. J. Jeon, Y. C. Choi, M. K. Nazeeruddin, M. Grätzel, and S. I. Seok, “Nanostructured TiO2/CH3 NH3PbI3 heterojunction solar cells employing spiro-OMeTAD/Co-complex as hole-transporting material,” J. Mater. Chem. A 1(38), 11842–11847 (2013). [CrossRef]  .

23. J. H. Heo, S. H. Im, J. H. Noh, T. N. Mandal, C.-S. Lim, J. A. Chang, Y. H. Lee, H.-J. Kim, M. Arpita Sarkar, K. Nazeeruddin, M. Grätzel, and S. I. Seok, “Efficient inorganic–organic hybrid heterojunction solar cells containing perovskite compound and polymeric hole conductors,” Nat. Photonics 7(6), 486–491 (2013). [CrossRef]  .

24. B. Conings, L. Baeten, C. De Dobbelaere, J. D’Haen, J. Manca, and H.-G. Boyen, “Perovskite-based hybrid solar cells exceeding 10% efficiency with high reproducibility using a thin film sandwich approach,” Adv. Mater. 26(13), 2041–2046 (2014). [CrossRef]  .

25. Julian Burschka, Florian Kessler, Mohammad K. Nazeeruddin, and M. Grätzel, “Co (III) complexes as p-dopants in solid-state dye-sensitized solar cells,” Chem. Mater. 25(15), 2986–2990 (2013). [CrossRef]  

26. M. L. Petrus, K. Schutt, M. T. Sirtl, E. M. Hutter, A. C. Closs, J. M. Ball, J. C. Bijleveld, A. Petrozza, T. Bein, T. J. Dingemans, T. J. Savenije, H. Snaith, and P. Docampo, “New Generation Hole Transporting Materials for Perovskite Solar Cells: Amide-Based Small-Molecules with Nonconjugated Backbones,” Adv. Energy Mater. 8(32), 1801605 (2018). [CrossRef]  .

27. X. Yin, Z. Yao, Q. Luo, X. Dai, Y. Zhou, Y. Zhang, Y. Zhou, S. Luo, J. Li, N. Wang, and H. Li, “High efficiency inverted planar perovskite solar cells with solution-processed NiO x hole contact,” ACS Appl. Mater. Interfaces 9(3), 2439–2448 (2017). [CrossRef]  .

28. J. A. Christians, R. C. Fung, and V. Prashant, “Kamat “An inorganic hole conductor for organo-lead halide perovskite solar cells. Improved hole conductivity with copper iodide,”,” J. Am. Chem. Soc. 136(2), 758–764 (2014). [CrossRef]  .

29. S. Ye, W. Sun, Y. Li, W. Yan, H. Peng, Z. Bian, Z. Liu, and C. Huang, “CuSCN-based inverted planar perovskite solar cell with an average PCE of 15.6%,” Nano Lett. 15(6), 3723–3728 (2015). [CrossRef]  .

30. W. Sun, Y. Li, S. Ye, H. Rao, W. Yan, H. Peng, Y. Li, Z. Liu, S. Wang, Z. Chen, L. Xiao, Z. Bian, and C. Huanga, “High-performance inverted planar heterojunction perovskite solar cells based on a solution-processed CuOx hole transport layer,” Nanoscale 8(20), 10806–10813 (2016). [CrossRef]  

31. S. Murase and Y. Yang, “Solution processed MoO3 interfacial layer for organic photovoltaics prepared by a facile synthesis method,” Adv. Mater. 24(18), 2459–2462 (2012). [CrossRef]  

32. M. Kaltenbrunner, G. Adam, E. D. Głowacki, M. Drack, R. Schwödiauer, L. Leonat, D. H. Apaydin, H. Groiss, M. C. Scharber, M. S. White, N. S. Sariciftci, and S. Bauer, “Flexible high power-per-weight perovskite solar cells with chromium oxide–metal contacts for improved stability in air,” Nat. Mater. 14(10), 1032–1039 (2015). [CrossRef]  

33. C. Liu, K. Wang, P. Du, T. Meng, X. Yu, S. Z. Cheng, and X. Gong, “High performance planar heterojunction perovskite solar cells with fullerene derivatives as the electron transport layer,” ACS Appl. Mater. Interfaces 7(2), 1153–1159 (2015). [CrossRef]  

34. G. Boschloo and A. Hagfeldt, “Spectroelectrochemistry of nanostructured NiO,” J. Phys. Chem. B 105(15), 3039–3044 (2001). [CrossRef]  

35. X. Yin, P. Chen, M. Que, Y. Xing, W. Que, C. Niu, and J. Shao, “Highly efficient flexible perovskite solar cells using solution-derived NiOx hole contacts,” ACS Nano 10(3), 3630–3636 (2016). [CrossRef]  

36. S. Seo, I. J. Park, M. Kim, S. Lee, C. Bae, H. S. Jung, N.-G. Park, J. Y. Kim, and H. Shin, “An ultra-thin, un-doped NiO hole transporting layer of highly efficient (16.4%) organic–inorganic hybrid perovskite solar cells,” Nanoscale 8(22), 11403–11412 (2016). [CrossRef]  

37. J. H. Park, J. Seo, S. Park, S. S. Shin, Y. C. Kim, N. J. Jeon, H.-W. Shin, T. K. Ahn, J. H. Noh, S. C. Yoon, C. S. Hwang, and S. I. Seok, “Efficient CH3NH3PbI3 perovskite solar cells employing nanostructured p-type NiO electrode formed by a pulsed laser deposition,” Adv. Mater. 27(27), 4013–4019 (2015). [CrossRef]  

38. W. Nie, H. Tsai, J.-C. Blancon, F. Liu, C. C. Stoumpos, B. Traore, M. Kepenekian, O. Durand, C. Katan, S. Tretiak, J. Crochet, P. M. Ajayan, M. G. Kanatzidis, J. Even, and A. D. Mohite, “Critical role of interface and crystallinity on the performance and photostability of perovskite solar cell on nickel oxide,” Adv. Mater. 30(5), 1703879 (2018). [CrossRef]  

39. Y. Guo, X. Yin, J. Liu, Y. Yang, W. Chen, M. Que, W. Que, and B. Gao, “Annealing atmosphere effect on Ni states in the thermal-decomposed NiOx films for perovskite solar cell application,” Electrochim. Acta 282, 81–88 (2018). [CrossRef]  

40. Z. Zhu, Y. Bai, X. Liu, C.-C. Chueh, S. Yang, and A. K.-Y. Jen, “Enhanced efficiency and stability of inverted perovskite solar cells using highly crystalline SnO2 nanocrystals as the robust electron-transporting layer,” Adv. Mater. 28(30), 6478–6484 (2016). [CrossRef]  

41. Z. Zhu, Y. Bai, T. Zhang, Dr. Z. Liu, Dr. X. Long, Z. Wei, Z. Wang, Prof. L. Zhang, J. Wang, Prof. F. Yan, and Prof. S. Yang, “High-performance hole-extraction layer of sol–gel-processed NiO nanocrystals for inverted planar perovskite solar cells,” Angew. Chem. 126(46), 12779–12783 (2014). [CrossRef]  

42. J. W. Jung, C.-C. Chueh, and A. K.-Y. Jen, “A low-temperature, solution-processable, Cu-doped nickel oxide hole-transporting layer via the combustion method for high-performance thin-film perovskite solar cells,” Adv. Mater. 27(47), 7874–7880 (2015). [CrossRef]  

43. F. Jiang, Wallace C. H. Choy, X. Li, D. Zhang, and J. Cheng, “Post-treatment-Free Solution-Processed Non-stoichiometric NiOx Nanoparticles for Efficient Hole-Transport Layers of Organic Optoelectronic Devices,” Adv. Mater. 27(18), 2930–2937 (2015). [CrossRef]  

44. J. Singh, F. Quli, D. E. Wolfe, J. Thomas Schriempf, and J. Singh, “An overview: Electron beam-physical vapor deposition technology-Present and future applications,” Applied Research Laboratory, Pennsylvania State University, USA (1999).

45. W. Chen, Y. Wu, Y. Yue, J. Liu, W. Zhang, X. Yang, H. Chen, E. Bi, I. Ashraful, M. Grätzel, and L. Han, “Efficient and stable large-area perovskite solar cells with inorganic charge extraction layers,” Science 350(6263), 944–948 (2015). [CrossRef]  

46. A. B. Huang, J. T. Zhu, J. Y. Zheng, Y. Yu, Y. Liu, S. W. Yang, S. H. Bao, L. Lei, and P. Jin, “Achieving high-performance planar perovskite solar cells with co-sputtered Co-doping NiOx hole transport layers by efficient extraction and enhanced mobility,” J. Mater. Chem. C 4(46), 10839–10846 (2016). [CrossRef]  

47. E. Aydin, J. Troughton, M. De Bastiani, E. Ugur, M. Sajjad, A. Alzahrani, M. Neophytou, U. Schwingenschlögl, F. Laquai, D. Baran, and S. De Wolf, “Room-Temperature-Sputtered Nanocrystalline Nickel Oxide as Hole Transport Layer for p–i–n Perovskite Solar Cells,” ACS Appl. Energy Mater. 1(11), 6227–6233 (2018). [CrossRef]  

48. T. Abzieher, S. Moghadamzadeh, F. Schackmar, H. Eggers, F. Sutterlüti, A. Farooq, D. Kojda, K. Habicht, R. Schmager, A. Mertens, R. Azmi, L. Klohr, J. A. Schwenzer, M. Hetterich, U. Lemmer, B. S. Richards, M. Powalla, and U. W. Paetzold, “Electron-beam-evaporated nickel oxide hole transport layers for perovskite-based photovoltaics,” Adv. Energy Mater. 9(12), 1802995 (2019). [CrossRef]  

49. I. Raifuku, Y. Ishikawa, S. Ito, and Y. Uraoka, “Characteristics of perovskite solar cells under low-illuminance conditions,” J. Phys. Chem. C 120(34), 18986–18990 (2016). [CrossRef]  

50. M. Tyagi, M. Tomar, and V. Gupta, “Influence of hole mobility on the response characteristics of p-type nickel oxide thin film based glucose biosensor,” Anal. Chim. Acta 726, 93–101 (2012). [CrossRef]  

51. B. D Cullity and S. R. Stock, Elements of X-Ray Diffraction, (Pearson Education, 2014).

52. Y. M. Lu, W.-S. Hwang, and J. S. Yang, “Effects of substrate temperature on the resistivity of non-stoichiometric sputtered NiOx films,” Surf. Coat. Technol. 155(2-3), 231–235 (2002). [CrossRef]  

53. S. Ryu, J. H. Noh, N. J. Jeon, Y. C. Kim, W. S. Yang, J. Seo, and S. I. Seok, “Voltage output of efficient perovskite solar cells with high open-circuit voltage and fill factor,” Energy Environ. Sci. 7(8), 2614–2618 (2014). [CrossRef]  

54. E. J. Juarez-Perez, M. Wußler, F. Fabregat-Santiago, K. Lakus-Wollny, E. Mankel, T. Mayer, W. Jaegermann, and I. Mora-Sero, “Role of the selective contacts in the performance of lead halide perovskite solar cells,” J. Phys. Chem. Lett. 5(4), 680–685 (2014). [CrossRef]  

55. M. Jlassi, I. Sta, M. Hajji, and H. Ezzaouia, “Optical and electrical properties of nickel oxide thin films synthesized by sol–gel spin coating,” Mater. Sci. Semicond. Process. 21, 7–13 (2014). [CrossRef]  

56. J. H. Kim, P.-W. Liang, S. T. Williams, N. Cho, C.-C. Chueh, M. S. Glaz, D. S. Ginger, and A. K.-Y. Jen, “High-performance and environmentally stable planar heterojunction perovskite solar cells based on a solution-processed copper-doped nickel oxide hole-transporting layer,” Adv. Mater. 27(4), 695–701 (2015). [CrossRef]  

57. U. Kwon, B.-G. Kim, D. C. Nguyen, J.-H. Park, N. Y. Ha, S.-J. Kim, S. H. Ko, S. Lee, D. Lee, and H. J. Park, “Solution-processible crystalline NiO nanoparticles for high-performance planar perovskite photovoltaic cells,” Sci. Rep. 6(1), 30759 (2016). [CrossRef]  

58. Y. Yu, Y. Xia, W. Zeng, and R. Liu, “Synthesis of multiple networked NiO nanostructures for enhanced gas sensing performance,” Mater. Lett. 206, 80–83 (2017). [CrossRef]  

59. A. Jain and A. Kapoor, “A new method to determine the diode ideality factor of real solar cell using Lambert W-function,” Sol. Energy Mater. Sol. Cells 85(3), 391–396 (2005). [CrossRef]  

60. K. Bouzidi, M. Chegaar, and A. Bouhemadou, “Solar cells parameters evaluation considering the series and shunt resistance,” Sol. Energy Mater. Sol. Cells 91(18), 1647–1651 (2007). [CrossRef]  

61. D. Liu, M. K. Gangishetty, and T. L. Kelly, “Effect of CH3NH3PbI3 thickness on device efficiency in planar heterojunction perovskite solar cells,” J. Mater. Chem. A 2(46), 19873–19881 (2014). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (11)

Fig. 1.
Fig. 1. XRD data for as-deposited and annealed NiOx film on glass substrate.
Fig. 2.
Fig. 2. SEM images of (a) as-deposited; FESEM images of (b) as-deposited, (c) annealed NiOx film on bare glass and (d) bare FTO; inset NiOx-coated FTO glass.
Fig. 3.
Fig. 3. Topographic images by AFM of (a) bare FTO, (b) as-deposited FTO/NiOx film, and (c) annealed FTO/NiOx film; inset shows the 3D images.
Fig. 4.
Fig. 4. Transmittance of bare FTO and annealed FTO/NiOx
Fig. 5.
Fig. 5. Workfunction determination of (a) as-deposited and (b) annealed NiOx film by PYS.
Fig. 6.
Fig. 6. (a) Fabricated p-i-n PSC and (b) corresponding band diagram.
Fig. 7.
Fig. 7. (a) J-V curves measured at solar simulator under simulated AM 1.5 G in this experiment at day 1 and 28th of fabrication; (b) EQE of the inverted planar PSCs fabricated with NiOx HTLs for both as-deposited and annealed NiOx film.
Fig. 8.
Fig. 8. (a) X-ray diffraction peaks of CH3NH3PbI3; (b) UV-Vis spectra of FTO and PSC on as-deposited and annealed NiOx; (c) SEM topography of CH3NH3PbI3, and (d) cross-sectional FESEM image of the fabricated PSC.
Fig. 9.
Fig. 9. Working principal of EBPVD Technique.
Fig. 10.
Fig. 10. EDX spectrum of FTO/NiOx films. Inset shows the weight% and atomic% of elements of as-deposited NiOx film.
Fig. 11.
Fig. 11. Transmittance of EBPVD NiOx films on glass substrate. A clear decrease of transmission is observed in the higher energy shorter-wavelength when NiOx film thickness is increased and continue to exhibit lower transmission in the visible region of spectrum as well.

Tables (3)

Tables Icon

Table 1. Summery of electrical properties of NiOx films

Tables Icon

Table 2. Summary of fabricated PSC devices’ performance with NiOx HTM at the 1st and 28th days of fabrication

Tables Icon

Table 3. Summary of the highest and average values of photovoltaic performances of PSC devices.

Equations (1)

Equations on this page are rendered with MathJax. Learn more.

Dp = K λ / β cos θ
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.